+ All Categories
Home > Documents > The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi...

The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi...

Date post: 18-Jul-2020
Category:
Upload: others
View: 2 times
Download: 0 times
Share this document with a friend
255
The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary, Western Australia: Life history, population structure and effects on sedimentary microbial nitrogen cycling Robert John De Roach, B.Sc. (Hons.) This thesis is presented for the degree of Doctor of Philosophy, The University of Western Australia Zoology School of Animal Biology The University of Western Australia December, 2006
Transcript
Page 1: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

The polychaetes Australonereis ehlersi (Augener) and Simplisetia

aequisetis (Augener) within the eutrophic Swan River Estuary,

Western Australia: Life history, population structure and

effects on sedimentary microbial nitrogen cycling

Robert John De Roach, B.Sc. (Hons.)

This thesis is presented for the degree of Doctor of Philosophy,

The University of Western Australia

Zoology

School of Animal Biology

The University of Western Australia

December, 2006

Page 2: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Copyright © 2007 by Robert De Roach

All rights reserved

ii

Page 3: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

…no one who considers the facts…

will hereafter doubt that worms play an important part in nature.

Darwin – 1881

The Formation of Vegetable Mould through the Action of Worms.

iii

Page 4: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,
Page 5: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Frontispiece: Sediment burrows constructed and irrigated by adult Australonereis ehlersi.

v

Page 6: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,
Page 7: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

ABSTRACT

In my study of Australonereis ehlersi and Simplisetia aequisetis [Polychaeta: Nereididae] from the Swan

River Estuary, Western Australia, I assessed the life history, geographical population structure and production

of both species, then measured their roles in microbial denitrification and nitrogen cycling within the

sediments of the estuary.

Both species exhibit a mean life-span of approximately 1 year, a production:biomass turnover rate of about 3

and potentially are capable of reproducing throughout the year, peaking during winter to spring.

A. ehlersi exhibited a marine euryhaline distribution, occurring only in the main basin and lower estuary,

typically at a very low density of adults; S. aequisetis exhibited a euryhaline distribution, occurring estuary-

wide during both summer and winter.

High density and biomass of A. ehlersi occurred in the middle estuary (at Como), predominantly as winter-

recruiting juveniles. Gravid, atokous adults spawned pelagically, with a 2 to 4 month larval development

period preceding settlement. Intolerance of freshwater by the pelagic larvae possibly is the major reason

excluding specimens from the upper reaches of the Estuary. Adult S. aequisetis brood eggs and embryonic

larvae in tubiculous burrows; the life-cycle presumably progresses entirely in sediments of relatively stable

interstitial salinity (compared to pelagic fluctuations), enabling recruitment by larvae and adults into the upper

reaches of the Estuary.

Adult A. ehlersi inhabited U-shaped burrows; juvenile S. aequisetis were sediment reworkers. The effects of

both polychaete habits on sediment NH4+ (ammonium) and NOx

- (nitrate + nitrite) fluxes, including

denitrification, nitrification and ammonification rates, were assessed in vitro. Nitrogenous fluxes were

estimated in sediment V-cores with an individual polychaete and compared with uninhabited sediment. Each

treatment (A. ehlersi, S. aequisetis, uninhabited) was subject to a range of ambient NO3- concentrations [S] =

0, 20, 50, 100 and 500 μM NO3-, at which NH4

+ and NOx- fluxes (incorporating denitrification rates) were

measured without and then with C2H2 (acetylene). This kinetic-fix adaptation of the C2H2 block technique

facilitated: estimation of potential denitrification rates at saturated NO3- concentration; estimation of

denitrification rates at in situ NO3- concentration; and comparison of the effect of the two polychaete species

on sedimentary nitrification, ammonification and net NOx- and NH4

+ fluxes.

Denitrification was greatest for sediment inhabited by A. ehlersi, moderate in uninhabited sediment and

lowest in sediment inhabited by S. aequisetis. The maximum potential denitrification rate (Vmp) and apparent

Michaelis constant (Kapp) have been reported for each sediment treatment. Sediment nitrification rate was

higher in S. aequisetis- than in A. ehlersi-inhabited cores, and lowest in uninhabited sediment. In the absence

of C2H2, the net result of nitrification minus denitrification resulted in a transition from NOx- efflux from

sediment at lower [S] to sedimentary NOx- uptake at higher [S]. The equilibrium point (zero NOx

- flux)

occurred at lower [S] for A. ehlersi-inhabited sediment cores than for uninhabited cores; S. aequisetis-

vii

Page 8: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

inhabited cores generally effluxed NOx- over the entire range of [S]. The ammonification rate was higher for

A. ehlersi than S. aequisetis-inhabited cores, and lowest in uninhabited cores where polychaete excretion was

absent. In the absence of C2H2, sediments of S. aequisetis inhabited cores indicated a lower net NH4+ influx

than uninhabited cores, whereas A. ehlersi inhabited cores exhibited a slight net efflux of NH4+ from the

sediment.

The difference in magnitude of nitrogenous fluxes imparted by the two polychaete species is hypothesised to

relate to the influence of their respective habits on the composition and activity of their associated

sedimentary microbial community. Juvenile S. aequisetis are hypothesised to homogenise and aerate

sediment continually, enhancing microbial nitrification and retarding anaerobic denitrification. Permanent

A. ehlersi burrows would facilitate vertical and radial oxic/anoxic stratification of sediment which, combined

with enhanced substrate supply through burrow ventilation, resulted in increased rates of microbial

denitrification and nitrification.

I have proposed a preliminary framework by which guilds of benthic fauna, each with similar designated

habits, may be tested for predictable bioturbative influence on nitrogen cycling, i.e. whether particular habits

may be considered ‘functional groups’. In conclusion, the fine-scale effects of A. ehlersi and S. aequisetis on

microbial nitrogen cycling are integrated with details of broader-scale population dynamics to define the role

of polychaetes in estuarine nitrogen cycling, with a view to managing eutrophication.

viii

Page 9: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

TABLE OF CONTENTS Abstract vii

Table of Contents ix

Acknowledgements xi

PART I – THE CONTEXT

Chapter 1 General Introduction 1

Chapter 2

The Swan River Estuary: Hydrodynamics, Eutrophication and

Macrobenthic Ecology 7

Chapter 3

The Nitrogen Cycle: Processes, Methodology and Benthic

Macrofaunal Effects 39

PART II – THE STUDY

Chapter 4

Life History, Population Structure and Production of

Australonereis ehlersi and Simplisetia aequisetis in the Swan

River Estuary, Western Australia 59

Chapter 5 Effects of Australonereis ehlersi and Simplisetia aequisetis on

Sedimentary Nitrogen Cycling, including the Kinetics of

Nitrate Utilisation by Resident Denitrifying Bacteria 89

PART III - CONCLUSIONS

Chapter 6 Polychaete Effects on Nitrogen Cycling at the Sediment-Water

Interface: Testing the Functional Group Concept 119

Chapter 7 General Discussion: Integrating Benthic Faunal Biology with

Sedimentary Nitrogen Cycling 133

References 139

APPENDICES

Appendix A The Salt Wedge: Temporal Variation and Stratification of

Salinity in the Swan River Estuary 169

Appendix B Algal Bloom Succession in the Swan River Estuary 175

Appendix C The Swan River Estuary’s Benthic Macrofaunal Community 181

Appendix D Component Processes of the Nitrogen Cycle 189

Appendix E Denitrification Activity in Sediment Surrounding Polychaete

(Simplisetia aequisetis) Burrows 233

ix

Page 10: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,
Page 11: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

ACKNOWLEDGEMENTS

Primarily, I acknowledge my supervisor Dr. Brenton Knott. Your mentorship, provision of

intellectual stimulation, unfailing support, encouragement, editorial effort and skill have enabled me

to complete this thesis. More than that, you are a good friend. I shall eternally be grateful for all

that you have helped me to achieve: academically, professionally and personally.

A chance reading of a mid 1980s paper by Prof. Erik Kristensen initially sparked my interest in the

capacity of sediment-dwelling animals to regulate supremely the eutrophic status of a water body.

Importantly, he championed the importance of understanding both faunal biology and sediment-

water biogeochemistry. Aside from inspiring my research, at an early stage Prof. Kristensen

provided technical advice and general encouragement for which I am very grateful. My sediment

V-core design is based on his work.

Of no lesser importance, I would like to acknowledge the influence of my high school Biology

teacher, Mr. Rob Gaudet. Your wonderful ability to transfer concepts with chalk drawings catalysed

my decision to become a biologist.

To all of the very helpful and friendly employees of Unisense (Aarhus, Denmark), your technical

assistance in the finer workings of microsensors is very much appreciated.

Polychaete taxonomic advice and clarifications were received from Dr. Pat Hutchings, Dr. Chris

Glasby and Dr. Robin Wilson. I sincerely thank each of you for teaching me the importance and

relevance of taxonomy.

Further, this thesis was markedly improved following editorial review by Prof. Lyn Beazley and Dr.

Jane Prince. I kindly thank you both.

Mr. Andrew Barber, I can not articulate how indebted I am to you for your enormous efforts, over

many years, of field assistance. Your help and unfailing practical and moral support are immensely

appreciated; you are a fantastic friend.

Developing and accomplishing this thesis would have been many-fold harder were it not for the

fruitful discussions and invaluable time spent together with my fellow Ph.D. graduates/candidates,

Dr. Glenn Shiell, Dr. Dennyse Newbound and Ms. Patience Lindhjem.

xi

Page 12: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

The support of my family is unbelievable and unconditional. To my Dad, Dr. John De Roach, you

instilled my sense of wonder and questioning of the natural world. You taught and challenged me to

develop, convey and discuss my own ideas. Our conversation is unique and I cherish it. With all

my heart I thank you for enabling me (in so many ways) to follow in your footsteps. To my mum,

Mrs. Marilyn De Roach, the best teacher in the world, you showed me how to be me. In this past

year, your ability to transform personal tragedy and despair into hope and opportunity is nothing

short of awe-inspiring. Thank you for showing me how to believe in myself and in others. Jacqui, I

couldn’t ask for a better sister. Thank you for always being there for me. To each of my

grandparents, thank you for nurturing me. I love each of you very much.

Ultimately, I am indebted to my fiancé Kirsty. More than anyone else, you have shared in the

experience of completing this thesis. Despite all recent hardship, your love, support,

encouragement, personal sacrifice and eternal commitment are the reason this thesis is finished.

From the bottom of my heart, thank you. Words can’t describe how much I love you, but I can’t

wait to write our next chapter together.

This thesis was financially supported by an Australian Postgraduate Award, an Australian Research

Council Small Grant held by Dr. Brenton Knott, and postgraduate funding by the School of Animal

Biology.

xii

Page 13: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

PART I – THE CONTEXT

xiii

Page 14: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,
Page 15: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

CHAPTER 1

GENERAL INTRODUCTION 1.1 OVERVIEW

I examined the biology of benthic fauna to elucidate their roles in nitrogen cycling within sediments

of the eutrophic Swan River Estuary, Western Australia. Specifically, I studied the life history and

population structure of Simplisetia aequisetis (Augener, 1913) and Australonereis ehlersi (Augener,

1913) (Annelida: Polychaeta: Nereididae). The environmental and socio-economic problems

associated with eutrophication of the estuary contextually drove my interest in determining whether

these polychaetes have a measurable impact on nitrogen cycling at the benthic sediment-water

interface. Given that the benthic fauna play an integral role in biogeochemical nitrogen cycling, an

understanding and quantification of the role of the constituent faunal biology may facilitate the

development of strategies to manage eutrophication effectively within the estuary.

Polychaete worms typically constitute an important component of the benthic fauna of marine and

estuarine ecosystems. In a broad and insightful review of the functional significance of polychaetes

in sedimentary communities, Knox (1977) defined the Polychaeta amongst the most dominant

taxonomic group of macrobenthic animals, with species contributing, on average, greater than 40%

to both community richness and abundance. In many parts of southern Australia, polychaete worms

comprise about half of the species richness of all macrobenthic invertebrates (Wilson et al., 2003).

Knox (1977) also (i) described the typically large contribution of polychaetes to community

biomass, productivity and respiration; and (ii) gave recognition to the often major influence of

benthic polychaetes upon their biophysical surroundings, including synergistic effects on sediment

reworking, particle size distribution, stability and community composition. Furthermore, the role of

polychaetes and other benthic fauna in sedimentary nutrient cycling, particularly of nitrogen, has

gained increasing recognition over the past 30 years. Sediment mixing, irrigation and burrow

construction by benthic fauna with their physiological demands and outputs, influence the chemical,

physical and microbial environment within which sedimentary nitrogen cycling occurs (reviewed

excellently by Petr, 1977; Aller, 1982, 1988; Krantzberg, 1985; Kristensen, 1988, 2000; Andersen

and Kristensen, 1991; Aller and Aller, 1998; Welsh, 2003).

Nitrogen is the chief element limiting primary production in most unpolluted estuaries and coastal

marine systems (Ryther and Dunstan, 1971; Carpenter and Capone, 1983; Seitzinger, 1990), and

also in some freshwater lakes, streams and rivers (Keeney, 1973; Gerhart and Likens, 1975;

Seitzinger, 1990). However, as aquatic systems continue to receive an increasing load of nitrogen

1

Page 16: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

and other nutrients, primarily sourced from fertilisers, eutrophication has become a major problem,

both throughout the world and in south-western Australia (Heathwaite et al., 1996; Colney, 2000;

Meyer-Reil and Köster, 2000; Peters and Donohue, 2001). Eutrophication is the process by which

increasing nutrients (typically nitrogen or phosphorous but also carbon) cause a change in the

nutritional status of a given body of water (Meyer-Reil and Köster, 2000). Increasingly, water

bodies with resident algal populations previously limited in productivity by low nitrogen

availability are becoming ‘unlimited’ in nitrogen supply, often resulting in algal growth of

excessive intensity and frequency (Pedersen and Borum, 1996). Within the Swan River Estuary,

nitrogen is more limiting to algal biomass production than phosphorous, by ≤20 times (Horner

Rosser and Thompson, 2001).

The environmental, health and economic implications of algal blooms can be far-reaching. Excess

accumulations and subsequent biodegradation of phytoplankton, epiphytic and/or ephemeral algae

may be displeasing aesthetically but, more importantly, can cause highly detrimental secondary

effects (Pedersen and Borum, 1996). Blooming phytoplankton species may contain or produce

toxins that directly impact on the health and mortality of trophically-related or surrounding

organisms (including humans) (Atkins et al., 2001; de Jonge et al., 2002). Turbidity and shading

caused by fast-growing algal species may reduce the growth or abundance of benthic microphytes

and macrophytes such as seagrasses and kelp (Pedersen and Borum, 1996). Biodegradation of

excess algae consumes oxygen that can lead to anoxia of both bottom and overlying waters (Meyer-

Reil and Köster, 2000) and to significant mortality of oxygen-dependent organisms including fish

and benthic fauna (Elliott and de Jonge, 2002). Further, loss of benthic fauna and flora may

destabilise sediments, thereby reinforcing the effects of increases in turbidity (Meyer-Reil and

Köster, 2000). Changes in species distribution, abundance, diversity and physiological state have

the potential to transfer cascading effects to higher trophic levels, ranging from slight to major

disruptions to fish, bird and mammal populations, to complete ecosystem collapse in extreme cases

(Meyer-Reil and Köster, 2000; Elliott and de Jonge, 2002). Consequential deleterious impacts on

human values and economies associated with problem waterways include reduced fisheries, direct

threats to public health and lowered potential for recreational activities and tourism.

Consequently, any natural means by which nitrogen can be unloaded from eutrophic waters

deserves attention. The process of sedimentary microbial denitrification, whereby soluble inorganic

forms of nitrogen (nitrate and nitrite) are converted to gaseous forms (nitrous oxide and dinitrogen

gas) that are free to percolate from aquatic systems, is particularly important since it represents an

ultimate sink of nitrogen with the capacity to regulate the eutrophic status of water bodies

(Seitzinger, 1990). For example, in Port Phillip Bay, Victoria, Australia, denitrification removes (as

2

Page 17: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

dinitrogen gas) an amount of nitrogen that is approximately equivalent to the combined total

loading of nitrogen to that system each year from catchment drainage, atmospheric inputs and

outfall from the major wastewater treatment plant servicing the city of Melbourne (Harris et al.,

1996). Whilst the large embayment remains highly productive due to tight internal cycling of

nitrogen, sedimentary denitrification primarily keeps Port Phillip Bay in an oligotrophic to

mesotrophic (low nutrient) state (Harris et al., 1996). Although a critical terminal process,

denitrification needs to be considered in the context of the entire nitrogen cycle and not regarded as

a separate event (see Chapter 3) (Kuenen and Robertson, 1988). Further, a holistic understanding of

the many biological and physico-chemical controls on the flows and transformations of nitrogenous

species within the affected system is crucial to successful eutrophication management strategies.

In this regard, there has been increasing awareness recently of the role benthic fauna plays in

regulating the size, diversity, distribution and activity of populations of denitrifiers and other

nitrogen-transforming bacteria (e.g. see reviews by Petr, 1977; Aller, 1982, 1988; Krantzberg, 1985;

Henriksen and Kemp, 1988; Kristensen, 1988, 2000; Andersen and Kristensen, 1991; Svensson et

al., 2001; Welsh, 2003). For example, the presence of burrows considerably extends the sediment-

water interface area where microbial populations can reside (Hylleberg and Henriksen, 1980;

Kristensen, 1984a; Kristensen et al., 1991), compared to sediment devoid of fauna. Burrow walls

are often bound by organic-rich exudates of the inhabitant fauna which further supports elevated

microbial activity (Aller, 1983, 1988; Aller et al., 1983; Reichardt, 1988). Irrigation of burrows

transports substrates of denitrification (nitrate and nitrite) from the overlying water to depths greater

than expected by diffusion, potentially enhancing the rate of nitrogen removal (Aller, 1988;

Kristensen, 1988; Kristensen et al., 1991). However, the supply of oxygenated water can suppress

anaerobic denitrification but favours other microbial nitrogen transformations such as nitrification

(the conversion of ammonium to nitrate) (Svensson et al., 2001). Intermittent ventilation is typical

of many infaunal species, and burrow waters may quickly turn anoxic during periods of ventilatory

rest (Kristensen, 2000), often imparting alternating redox conditions within the burrow. Such

circumstances may lead to significantly increased rates of inorganic nitrogen removal when

ammonium is first aerobically nitrified to nitrate during active ventilation periods, then to be

anaerobically denitrified during rest periods (coupled nitrification-denitrification) (Jenkins and

Kemp, 1984; Henriksen and Kemp, 1988; Capone, 2000). De Roach et al. (2002) previously

indicated that the burrowing behaviour of adult S. aequisetis may enhance significantly the

denitrification process of sediments in the Swan River Estuary.

The many ways in which benthic fauna, particularly polychaetes, can affect the flow and microbial

transformations of nitrogen within aquatic systems is a major theme of my thesis. A particular aim

3

Page 18: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

of my study was to determine the influence of A. ehlersi and S. aequisetis upon denitrification and

broader microbial nitrogen cycling within their surrounding sediment environment. Highlighting

the importance of faunal impacts on sedimentary nitrogen cycling underscores the potential for

development of benthic fauna management strategies aimed at preventing the deleterious effects of

eutrophication and problem algal blooms.

1.2 THESIS QUESTION, AIMS AND STRUCTURE

In my thesis I address two, over-arching questions:

1. What are the life histories and population structures of A. ehlersi and S. aequisetis

within the Swan River Estuary, Western Australia?

2. What effects do members of both species of polychaete have on microbial nitrogen

cycling within sediments of the estuary?

Three specific aims are addressed, as follows:

I. I present information concerning the life history, geographical population structure and

production of the polychaetes A. ehlersi and S. aequisetis from the Swan River Estuary, in

Chapter 4.

II. Results of my measurements of the effect of A. ehlersi and S. aequisetis individuals on

microbial denitrification and nitrogen cycling within the sediments of the estuary are

presented in Chapter 5.

III. In Chapters 6 and 7 I integrate knowledge on fine-scale effects of benthic fauna on

microbial nitrogen cycling, with a broad-scale understanding of faunal population dynamics

to highlight the functional role of benthic fauna in estuarine nitrogen cycling, with

important implications for eutrophication management.

In Chapter 1, I present a general overview of my thesis content and intent. In Chapter 2, I present a

detailed description of the structure and function of the Swan River Estuary, leading into a review

of the causes of the nitrogenous-based eutrophication and related environmental and socio-

economic problems, and a review of the current state of knowledge regarding A. ehlersi and

S. aequisetis taxonomy, biology, and of the composition of the Swan River Estuary’s macrobenthic

community.

Such an assessment of the species’ basic biology and ecological niche is considered requisite to a

comprehension of their biogeochemical role in the environment. Prior to providing a historical

4

Page 19: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

perspective on the realisation of the effects of benthic fauna on sedimentary microbial nitrogen

cycling, Chapter 3 defines the component processes of the nitrogen cycle, placing emphasis on the

importance of denitrification as an ultimate sink of nitrogen in aquatic environments. Chapter 3

also provides an ideological framework for the methodological consideration of nitrogen cycling,

notably defining denitrification as an enzymatic process. The Appendices provide supporting

information relevant to these contextual issues.

5

Page 20: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

6

Page 21: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

CHAPTER 2

THE SWAN RIVER ESTUARY: HYDRODYNAMICS,

EUTROPHICATION AND MACROBENTHIC ECOLOGY

2.0 INTRODUCTION

The growth of the world’s population from 1.6 billion in 1900 to over 6 billion in 2006 has been

sustained, in large part, by increased crop productivity attributable to development of ammonia-

based fertilisers (Smil, 1999). Unfortunately, the prolonged and intensive use of fertilisers on a

global scale, compounded by the consequences of human population expansion (changing land use

patterns, deforestation, increased discharge of municipal and industrial wastes), has lead to

significantly increased export of nitrogen and other nutrients to near-shore aquatic ecosystems

(Rabouille et al., 2001). Where these increases in nutrients (typically nitrogen or phosphorous but

also carbon) cause a change in the nutritional status of a given body of water, the process can be

defined as eutrophication (Meyer-Reil and Köster, 2000). The world’s rivers, estuaries and coastal

ocean are becoming increasingly eutrophic and subject to the effects of shifts, often deleterious, in

primary productivity. In non-impacted systems, estuarine and near-shore primary productivity is

typically limited by nitrogen availability, whilst phosphorous availability typically limits freshwater

primary production (Nixon, 1995; Peters and Donohue, 2001). Within eutrophic systems, nutrient

limitation of algal productivity is dampened, often resulting in algal growth of excessive intensity

and frequency (Pedersen and Borum, 1996). The environmental and socio-economic implications

of nuisance algal blooms are extensive: problems range from slight aesthetic displeasure to

complete ecosystem collapse (Elliott and de Jonge, 2002; see Section 2.1.3 below). In this section I

outline the pertinent causes and implications of eutrophication and associated algal blooms using

the eutrophication of the Swan River Estuary, Western Australia, as a case study. Such

consideration is intended to highlight not only the broad-scale issues relevant to nitrogenous

eutrophication but also to provide the local context of this thesis.

2.1 THE SWAN RIVER ESTUARY AND EUTROPHICATION

2.1.1 Catchment Geography

The Swan River Estuary, 60 km in length from its mouth to the Indian Ocean at Fremantle to its

confluence with Ellen Brook (Hodgkin, 1987) within a catchment of 121 000 km2 in south-west

Western Australia, covers an area of 53 km2 (Figure 2.1). The riverine source is 375 m above sea

level on the Darling Scarp (Stephens and Imberger, 1996). The estuary has been a major node for

7

Page 22: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

the State’s population growth since European settlement in 1829. The State capital city, Perth, sits

astride the estuary with a present population of approximately 1.4 million (Atkins et al., 2001),

expected to double over the next 30 years (Stephens and Imberger, 1996). Water quality of the

Swan River Estuary, post-settlement, has been related to changes in land-use patterns within the

catchment (Atkins and Klemm, 1987; Kurup et al., 1998).

Drainage into the estuary originates from two main areas:

• The Avon catchment comprising a high escarpment (the Darling Scarp) and dissected

plateau, primarily draining into the Avon River and its tributaries (Figure 2.1; Atkins and

Klemm, 1987; Kurup et al., 1998; Viney and Sivapalan, 2001).

• The lower Swan Coastal Plain sub-catchments extend approximately 30 km inland and are

drained via streams and stormwater drains.

The Avon catchment contributes 85% of total flow and is the larger of the two source areas,

extending 500 km inland and covering 120 000 km2 in area. It is partially forested with remnant

native dry sclerophyll woodland vegetation but has been cleared extensively for agriculture, mainly

between 1940 and 1970 (Atkins and Klemm, 1987; Kurup et al., 1998; Peters and Donohue, 2001;

Viney and Sivapalan, 2001). The Darling Scarp elicits a strong west-east rain-shadow effect,

ranging from mean annual rainfall >800 mm on the scarp, to ~250 mm in the eastern parts of the

catchment (Viney and Sivapalan, 2001). Mean annual rainfall over the entire catchment is

~400 mm.

The lower sub-catchments of the Swan Coastal Plain collectively contain much of the metropolitan

population within 1 000 km2. The area is highly urbanised but also retains much agricultural land

use (Atkins and Klemm, 1987; Kurup et al., 1998). Groundwater discharge exerts primary control

on streamflow on the coastal plain (Peters and Donohue, 2001). The systems responsible for the

discharge are the Gnangara Mound to the north of the estuary, the Cloverdale Groundwater Flow

System in the region between the Swan and Canning Rivers, Jandakot Mound to the south of the

estuary, and the Cottesloe Mound is a small freshwater lens between coast and estuary (Linderfelt

and Turner, 2001). The groundwater systems are recharged solely by rainfall (Donohue et al.,

2001). Soils of the coastal plain are generally coarse and sandy with poor nutrient retention, but

interdigitate with a silt and clay unit at the base of the Darling Scarp (Peters and Donohue, 2001;

Viney and Sivapalan, 2001).

8

Page 23: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

PERTH

5 km

Avon River

Ellen Brook

N

WESTERN AUSTRALIA

FREMANTLE

Helena River

Darling Scarp

Swan

Estuary

River

Canning River

150 km

[A]

[B]

Indian Ocean

300 mm

400 mm

500 mm

600 mm

800 mm

1200 mm

Inset [A]

AVON CATCHMENT BOUNDARY

Figure 2.1: The Swan River Estuary and its sub-catchments: [A] Swan Coastal Plain sub-

catchments; [B] the greater Avon Catchment, illustrated with rainfall contours. (Adapted from Peters and Donohue, 2001).

9

Page 24: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

2.1.2 Hydrology

The hydrodynamics exert primary control over both the ecology, particularly of the spatio-temporal

distribution and abundance of algal and benthic populations, and also of nutrient transport and sinks

within the Swan River Estuary: both aspects constitute a primary focus of this thesis. It is therefore

relevant to review here the hydrology of the estuary, particularly of the main factors governing the

dynamics of this seasonal system. Spencer (1956), Hodgkin (1987), Stephens and Imberger (1996)

and Kurup et al. (1998) have reviewed thoroughly the Swan River Estuary hydrodynamics.

Seasonality and Tides

The hydrological dynamics of the Swan River Estuary are a consequence of its extreme climatic

seasonality, small tidal range and characteristic bathymetry (Hodgkin, 1987; Stephens and

Imberger, 1996; Kurup et al., 1998). The climate is Mediterranean with typically hot, dry summers

and mild, wet winters and greater than 90% of annual rainfall is usually received between May and

October (Peters and Donohue, 2001). As an oversimplification, received freshwater inflows during

winter flush the upper reaches of the estuary thereby reverting the system to a freshwater to

brackish state (Stephens and Imberger, 1996; Linderfelt and Turner, 2001). Decreasing river

discharge in spring allows the upstream intrusion of marine-derived saline water as a salt-wedge,

resulting in marine conditions within the estuary during summer (Stephens and Imberger, 1996;

Linderfelt and Turner, 2001). The Swan River Estuary is therefore classified as a seasonal estuary

(Hodgkin, 1987).

Mean tidal range at Fremantle is 0.6 m, such that the Swan River Estuary is also classified as

microtidal (i.e. estuaries with a tidal range <2 m; Monbet, 1992; Kurup et al., 1998). The low tidal

amplitude allows the persistence of a seasonal cycle of salinity variation, since tidal forcing in

winter is not sufficient to overcome the seaward flow of freshwater (Kurup et al., 1998). This

pattern contrasts with those of meso- and macrotidal estuaries (i.e. tidal range >2 m) where tidal

forcing permits a semi-diurnal, diurnal or fortnightly periodicity in salinity variation, with less

dominance of freshwater discharge (Kurup et al., 1998). Whilst tidal amplitude in the Swan River

Estuary is small, it is nevertheless sufficient to exert an influence up to 60 km upstream during

summer, when freshwater discharge is restricted (Linderfelt and Turner, 2001).

Physiognomy

Underlying the general characterisation of the Swan River Estuary as a seasonal microtidal estuary,

is much variation in estuarine dynamics both on a local geographical scale (i.e. across the spatial

extent of the estuary) and on a temporal scale (i.e. inter-annual variation). The estuary has a free

10

Page 25: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

connection with the open sea and as such approximates Pritchard’s (1967: p. 3) definition of an

estuary as “a semi-enclosed body of water within which sea water is measurably diluted with fresh

water derived from land drainage”. However, Hodgkin (1987) also points out that such a

superficial characterisation of the system as a typical estuary is inadequate for a useful

consideration of its hydrology. Estuarine classification systems may be important for delineating

broad-scale differences between estuaries, but an appreciation of locally-specific features is

necessary to a full understanding of the hydrology of any given estuary. There are several physical

features of the Swan River Estuary that impart implicit effects on its hydrodynamics.

Hodgkin (1987) and Stephens and Imberger (1996) subdivide the Swan River Estuary into five

major regions of hydrological importance: the main basin, two adjacent sills, Fremantle Harbour

and the upper estuary (Figure 2.2). From Preston Point (~5 km upstream of the mouth), a deep (up

to 17 m) and narrow 5 km long trough opens into an approximately 10 km long by 2 km wide

shallower basin (>5 m deep in the centre). The upstream limit of the basin is the constriction

known as the Narrows ~20 km from the mouth. The trough and basin areas are referred to

collectively as the deep or main basin. The Canning River also discharges into the Swan River

Estuary, and for 6 km from its confluence point to the Kent Street Weir is basically an extension of

the main basin. Average depth decreases from 4-5 m at the confluence to less than 1 m upstream of

Salter Point (~4 km upstream of the confluence). The weir acts as a barrier to penetration of marine

water, such that the upstream Canning River remains fresh year round and is not considered a part

of the estuary.

Sills on either side of the main basin regulate the flow of water within the estuary. Downstream of

the deep trough, a shallow (5 m deep) and narrow sill extends for 2 km from Preston Point to the

Fremantle Traffic Bridge (~3 km upstream of the mouth). From the bridge, a 3 km long, 11 m deep,

channel is dredged to the mouth of the estuary to comprise Fremantle Harbour. The downstream

Fremantle Sill controls water exchange between the ocean and the estuary. Upstream of the main

basin, Perth Water comprises a secondary sill, a wide, expanse of shallow water (<2 m deep)

between the Narrows and Heirisson Island. The second sill regulates water flow between the main

basin and the 35 km stretch of upper estuary, i.e. the narrow riverine reaches between Heirisson

Island and the confluence with Ellen Brook (~60 km upstream of the mouth). Average water depth

of the upper estuary is 2 to 3 m, punctuated by pockets of 5 to 6 m, but becomes much shallower

with many sand banks upstream of the confluence with Jane Brook (~50 km from the mouth).

11

Page 26: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Bathymetry

Dep

th (m

)

0

5

10

15

20

0 10 20 30 40Distance from mouth (km)

Upper Estuary

Main

Basin

Fremantle

Harbour

Fremantle Sill

Perth Water

Kent Street

Weir

The Narrows

Heirisson Island

Fremantle Traffic Bridge

Salter Point

Preston Point

Deep Trough

5 km

N

Figure 2.2: Physiognomy and bathymetry of the Swan River Estuary. Adapted from Stephens and

Imberger (1996).

12

Page 27: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Temporal Variation and Stratification

The described physical characters of the Swan River Estuary markedly influence the

hydrodynamics of this seasonal microtidal system. Stephens and Imberger (1996) and Kurup et al.

(1998) consider the hydrodynamics of the system in relation to the annual upstream migration of

tidally forced marine water as a salt wedge, when freshwater discharge decreases following the

cessation of winter rains. Kurup et al. (1998) describe the hydrology of the estuary in relation to

three phases of upstream salt wedge positioning controlled primarily by river discharge:

(1) Salt wedge (marine) dominated in summer and autumn (January to May), when freshwater

discharge is very low. High salinity water extends from the mouth to 40 to 60 km

upstream, i.e. between the confluence with the Helena River and Ellen Brook (Figure 2.3).

(2) (a) Salt wedge waning during early winter (June to July); or, (b) salt wedge emplacement

(propagation) during late spring and early summer (November to December); when

freshwater discharges are low. Fresh to brackish waters extend from the upper estuary to

surficial (<2 m deep) Perth Water (20 to 25 km upstream of the mouth), whilst high salinity

bottom waters (>2 m deep) may extend from Perth Water to 10 – 15 km into the upper

estuary. Thus, the resulting halocline from upstream bottom sediments to downstream

surface waters, i.e. the salt wedge, may extend for >10 km. (Figure 2.3).

(3) Salt wedge absent (freshwater dominated) in late winter and early spring (August to

October), when freshwater discharge is high. Fresh to brackish waters extend from the

upper estuary across the surficial (<5m deep) waters of the main basin to the Fremantle Sill

(5 km upstream of the mouth; Figure 2.3).

These descriptions are valid for the shallow waters (<5 m deep) of the estuary wherein the annual

hydrological cycle depends upon the seasonal variation in freshwater discharge regulating the

upstream migration of a tidally-forced salt wedge. However, the deep waters (>5 m deep) of the

main basin remain at high salinity year-round, with freshwater overlay causing stratification during

phase 3. Detailed consideration of important variations in hydrology occurring within the estuary

due to bathymetry, and also due to inter-annual variation in rainfall and resultant freshwater

discharge, is given in Appendix A. A similar stratification of the deeper pockets of the upper

estuary may also occur during phases 2 and 3, depending upon whether the magnitude of freshwater

discharge has been sufficient (>23-50 m3 s-1) to flush all high salinity water from the upper reaches

(Appendix A). Thus bathymetry, in addition to strong seasonality of river inflow and weak tidal

forcing, is also important in determining the hydrodynamics of the Swan River Estuary. In this

13

Page 28: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

respect the presence of the Fremantle Sill is particularly significant, since its shallow depth

ultimately defines the limits of water exchange between ocean and estuary. Differences in estuarine

hydrology between years are driven by inter-annual variation in rainfall and resultant freshwater

discharge, which governs both the timing and period of the salt-wedge phases and related

stratification events.

Phase 1 – Salt wedge dominated ~ January to May

Phase 2a – Salt wedge waning ~ June to July

Phase 3 – Salt wedge absent ~ August to October

Phase 3 – Salt wedge propagation ~ November to December

0 5 10 15 20 25 30 35 40Distance from mouth (km)

Dep

th (m

)

Figure 2.3: Phases of salinity-depth contours within the Swan River Estuary throughout the

course of a typical year of freshwater discharge. (Adapted from Kurup et al., 1998; Swan River Trust, 2004a).

14

Page 29: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

2.1.3 Eutrophication and Algal Blooms

The Swan River Estuary is a system highly suitable for a case study consideration of the causes and

implications of nitrogenous eutrophication. Nutrients have entered and accumulated within the

estuary over past years, and many authors suggest that levels are now sufficient for it to be

classified as eutrophic (e.g. John, 1994; Stephens and Imberger, 1996; Twomey and John, 2001).

Nitrogen has been identified as the chief nutrient limiting algal productivity within the estuary

(Thompson and Hosja, 1996; Douglas et al., 1997; Thompson et al., 1997), but increased

anthropogenic nitrogen loading has caused a dampening of this limitation on algal growth. The

estuary has a long-standing history of periodic algal blooms which, to some extent, are natural

phenomena (Hodgkin and Vicker, 1987); however, the heightened nutrient load within its more

recent history has increased both the frequency and intensity of problem phytoplankton blooms

(Thompson and Hosja, 1996). Marked anoxia upon algal biodegradation, fish kills (amongst other

fauna), loss of seagrasses and general changes in species’ composition have been associated with an

intensification of both toxic and non-toxic phytoplankton blooms within the Swan River Estuary

(Atkins and Klemm, 1987; Horner Rosser and Thompson, 2001). Further, these environmental

problems have major socio-economic implications including related public health costs, reduced

fisheries, decreased aesthetic and recreational value and loss of tourism (Twomey and John, 2001).

My aim in this section is to detail the main causes that have lead to the increased nutrient loading

and eutrophication of the Swan River Estuary, and to examine the consequences of such

eutrophication upon algal productivity and associated environmental and societal problems.

Nutrient Loading, Transport and Regeneration

Change in catchment land-use patterns has been the driving force of increased nutrient loading to

the Swan River Estuary. As the primary mechanism for nutrient transport, drainage to the estuary is

dominated by winter discharge from the Avon sub-catchment, contributing about 85% of total flow

(Atkins and Klemm, 1987; Kurup et al., 1998). However, during the critical summer period, inflow

is dominated by urbanised coastal plain drainage which typically has nutrient concentrations greater

than that of Avon drainage (Atkins and Klemm, 1987). The combined load is such that over the

period 1986 to 1996 the Swan River Estuary received an input of 550 tonnes of nitrogen per year.

Some of this nitrogen load may be transported to the ocean during freshwater discharge events in

winter and early spring, or during unseasonal rain events; however, ocean exports of nitrogen

during late spring, summer and autumn will typically be limited due to upstream tidal forcing of the

salt wedge (see Figure 2.3). The net effect of land-use changes upon nutrient contributions to the

estuary are separately considered for the two major sub-catchments below. The importance of

sediment nutrient regeneration as an internal supply of nitrogen is then discussed.

15

Page 30: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Agriculture is the main, present day land-use within the Avon sub-catchment. Only 25 to 35% of

the catchment retains its native vegetation of dry sclerophyllous woodland, which was mostly

cleared between 1940 and 1970 to enable sheep grazing and cereal cropping (Peters and Donohue,

2001; Viney and Sivapalan, 2001). Pastures typically are fertilised in autumn (April/May) to

promote good establishment of crops and growth in spring (Donohue et al., 2001). The timing of

fertiliser application heightens the availability of nutrients for surface and leachate transport at the

onset of winter rains (Donohue et al., 2001). Whilst the widespread application of fertilisers within

the Avon catchment has provided the source of nutrients primarily responsible for eutrophication of

the Swan River Estuary (Peters and Donohue, 2001), the potential capacity for nutrient transport

from this area has also increased. During the period 1986 to 1996, it was estimated that the

cumulative effect of land conversion within the Avon catchment facilitated a contribution of 356

tonnes or 65% of the total nitrogen entering the Swan River Estuary each year (Peters and Donohue,

2001).

The remaining 35% of nitrogen loading is generated within the much smaller sub-catchments of the

coastal plain. Over a similar period of monitoring (1987 to 1994), it was estimated that combined

load of nitrogen from the Avon and Ellen Brook sub-catchments was 484 tonnes per year (Viney

and Sivapalan, 2001). The coastal plain sub-catchment of Ellen Brook therefore contributes

approximately 128 tonnes or 23% of total nitrogen entering the Swan River Estuary each year.

Ellen Brook is the largest sub-catchment (664 km2) of the coastal plain and primarily retains an

agricultural land-use, mainly pasture for stock (Donohue et al., 2001). The impact of agriculture

within the Ellen Brook sub-catchment upon nutrient load and transport to the estuary is similar to

that detailed for the Avon catchment; however, the disproportionately higher nitrogen load for its

comparatively small area is a consequence of several other factors. The catchment comprises large

areas of highly permeable siliceous sand and groundwater is a major component of the stream’s

annual ephemeral flow, which is generally maintained from June to November following winter

rains (Donohue et al., 2001). A relatively small amount of superphosphate fertiliser is applied

directly to the discharge zone of the Gnangara groundwater system (Donohue et al., 2001; Peters

and Donohue, 2001). Thus, nutrient transport is significantly elevated via leachate flow and

overland flow when the water table is seasonally high in riparian and near-channel areas (Donohue

et al., 2001). Leguminous pasture has also been suggested to contribute to the large nitrogen pool

of the Ellen Brook sub-catchment (Donohue et al., 2001).

Within the urbanised portion of the coastal plain, nutrient inputs to the Swan River Estuary derive

mainly from industrial discharges, fertilisers (from gardens, parks and golf-courses) and seepage

from septic tanks and landfill sites (Atkins and Klemm, 1987). Soils of the residential and

16

Page 31: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

industrial areas of the catchment are characteristically sandy and highly permeable (Donohue et al.,

2001). Reduced evapotranspiration via loss of native vegetation, combined with importation of

water for irrigation, have increased recharge and nutrient leaching to the unconfined groundwater

aquifers which govern stream flow on the coastal plain (Donohue et al., 2001). As a result, urban

nutrient sources are estimated to contribute 66 tonnes or 12% of total nitrogen to the estuary each

year (Peters and Donohue, 2001; Viney and Sivapalan, 2001). The quantity is small compared with

the agricultural inputs. It is thought that considerable amounts of nitrogen are lost via

denitrification within the groundwater aquifers which are comprised of naturally high levels of

dissolved organic carbon and a pH between 5 and 7 (i.e. ideal conditions for denitrification, see

Appendix D, Section D.7) (Donohue et al., 2001). However, the timing of urban discharge is

highly significant to algal productivity within the estuary (Peters and Donohue, 2001). As opposed

to the mainly ephemeral discharge of the agricultural catchment following winter rains, many of the

urban drains are perennial, with groundwater discharge sustaining flows throughout the dry season

(Donohue et al., 2001). Whilst groundwater discharge on the coastal plain contributes only 6% of

total annual flow to the estuary, the percentage increases to 55% between the summer period of

November to April (Linderfelt and Turner, 2001). Nitrogen-laden groundwater drives urban

drainage and is forced through sediments underlying the estuary (Linderfelt and Turner, 2001), thus

providing the major allochthonous nitrogen source utilised by summer algal blooms. Thompson

and Hosja (1996) suggest that unseasonal summer rainfall may also be important in providing an

external nitrogen supply to surface waters.

More importantly, autochthonous benthic regeneration of dissolved inorganic nitrogen may be

equivalent to or greater than annual allochthonous loads (Pennifold and Davis, 2001). The

sedimentary decomposition of organic nitrogen (including that contained within collapsed and

deposited phytoplankton blooms) by benthic fauna and bacteria, and resupply of ammonium via

excretion and microbial mineralisation may be the most important source of inorganic nitrogen

during the summer period. Transport of the regenerated ammonium from the bottom and pore-

waters of the benthos to the overlying pelagic zone of phytoplankton productivity is driven by

physical and biological processes (Douglas et al., 1997). Physical transport mechanisms include

diffusion, groundwater forcing, wind-induced hydrodynamic forcing and density-driven

displacement of pore-waters by advancement of the salt-wedge (Douglas et al., 1997; Oldham and

Lavery, 1999); bioturbation, resulting from sediment mixing and burrow irrigation of a ubiquitous

benthic fauna, likely is a sediment-water nutrient transfer process of pervasive importance

(Pennifold and Davis, 2001).

17

Page 32: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Algal Blooms

The combined effect of catchment modification and consequential increases in loading and

transport rate of nitrogen and other nutrients, both urban- and agriculturally-derived, has been an

apparent increase in the intensity and frequency of algal blooms within the Swan River Estuary

(Horner Rosser and Thompson, 2001). While this eutrophication has been well documented, the

occurrence of nuisance blooms of algae appears to have been a regular phenomenon of the estuary,

even before the onset of eutrophication. Further, the annual succession of population growth of

various algal species within the estuary has been an on-going topic of relevance to the residents of

Perth since settlement. Hodgkin and Vicker (1987) describe the history of algal pollution in the

Swan River Estuary, which is briefly summarised in Appendix B together with a consideration of

the present day, annual succession of bloom-forming phytoplankton species and the conditions

necessary for bloom-forming species.

Presently, it is the increasingly eutrophic status of the estuary and shift from productive macroalgal

species towards both toxic and non-toxic phytoplankton species, together with the consequences of

an intensification of frequency and severity of blooms, which is driving environmental and public

concern. The spatio-temporal distribution pattern of resident phytoplankton within the Swan River

Estuary is determined primarily by the salinity preference and tolerance of individual species, and

therefore by the factors (e.g. bathymetry, seasonality, rainfall events and freshwater discharge)

governing the hydrology of the system (John, 1987). However, excessive growth (blooms) of

phytoplankton is ultimately regulated by nutrient, mainly nitrogen, availability (John, 1987;

Thompson and Hosja, 1996).

A Recent Example

During February 2000, within the study period of this thesis (see Chapter 3), a massive, toxic

phytoplankton bloom occurred throughout the entire 60 km length of the Swan River Estuary

(Atkins et al., 2001). The rainfall events and subsequent hydrological conditions that facilitated the

bloom within the estuary are detailed in Appendix A. Briefly, heavy and cyclonically-derived rain

fell over the whole catchment during January and the subsequent discharge had reverted the entire

estuary to a winter-like freshwater state by February. The typical catchment of the estuary was

somewhat extended, and areas of the Avon sub-catchment that do not usually receive enough

rainfall to exert considerable flow (even during winter), produced waterways with their first

significant flows in forty years (Atkins et al., 2001). As a result, the discharge deposited over 800

tonnes of nitrogen (i.e. more than the average total annual load) and 35 tonnes of phosphorous into

the subsequently freshwater estuary under summer-time influence (i.e. with long day lengths and

relatively warm, calm and clear water) (Atkins et al., 2001). In the upper and mid-estuary,

18

Page 33: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

concentrations of dissolved inorganic nitrogen were five to twelve times higher than normal

summer conditions (Atkins et al., 2001).

Whilst the high nutrient availability was highly conducive to phytoplankton growth, the low salinity

precluded exploitation by typical summer-time marine diatoms and dinoflagellates. Rather, several

species of freshwater algae proliferated, seeded either from small populations extant in the upper

reaches of the system (Atkins et al., 2001), or possibly from overflow of nearby eutrophic lakes (as

has been suggested by John, 1994, for previous algal blooms). Of these species, the non-nitrogen

fixing cyanobacteria Microcystis aeruginosa out-competed other phytoplankton in utilising the

abundant nutrient source to grow and reproduce. Microcystis aeruginosa has an effective gas

vacuole system which, combined with other carbohydrate metabolic features, enables it to float at

the water’s surface and receive the greatest share of sunlight for photosynthesis (Atkins et al.,

2001). Such a capability during the atypical hydrological circumstance, combined with high

nutrient availability, resulted in an extensive and spectacular bloom of M. aeruginosa. The bloom’s

appearance varied from a slight green tinge in open water to a dense bright green scum in many

sheltered bays and beaches (personal observation). The average cell density reached 15,000 ml-1 in

the general water column and 130 million ml-1 in the accumulated bright green scums at the peak of

the bloom (Atkins et al., 2001).

Figure 2.4: The Microcystis aeruginosa bloom in the Swan River, February, 2000. (Swan River Trust, 2000).

19

Page 34: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

During one period, water concentrations of 8 μg l-1 microcystin toxin produced by this species

exceeded the level recommended by the World Health Organisation by a factor of sixteen (Atkins et

al., 2001). Indeed, 20-25 μg of M. aeruginosa biomass per kilogram of mammal biomass is fatal to

mice (Atkins et al., 2001). The high toxicity posed a tangible hazard to public health and

authorities closed the entire estuary for a period of two weeks, disallowing any contact with the

water until the bloom collapsed with the re-entrance of marine water (Atkins et al., 2001). The

upper salinity tolerance of M. aeruginosa is 10 g kg-1 (Atkins et al., 2001). As freshwater flow

subsided by mid-February, re-establishment of the salt wedge caused the M. aeruginosa bloom to

die-off first in the lower estuary, with subsequent mortality occurring upstream concurrent with

progression of the salt wedge. An estuary-wide collapse of the bloom occurred by the end of

February (Atkins et al., 2001).

Consequences of Eutrophication and Algal Blooms

The environmental, social and economic implications of the February 2000 M. aeruginosa bloom,

and moreover of the recently heightened frequency and intensity of both toxic and non-toxic blooms

within the Swan River Estuary, are far-reaching. Aside from posing a direct threat to human health,

the toxicity of several species of blooming cyanobacteria (e.g. Microcystis spp., Anabaena spp.,

Nodularia spp.) and dinoflagellates (Karlodinium spp., Gymnodinium spp.) may severely impact on

proximate fauna (Blackburn et al., 1997; Kanandjembo, 1998). For example, blooms of the

dinoflagellate Karlodinium micrum, whilst not toxic to humans, clog the gills of fish and cause

fatalities through asphyxiation (Swan River Trust, 2004b). Initial nutrient supply to the Swan River

Estuary provided by the onset of winter rains caused significant blooms of K. micrum and over

100,000 and 30,000 respective fish deaths in 2003 and 2004 (Swan River Trust, 2004b).

Furthermore, microbial oxygen consumption during the biodegradation of blooms and dead fish

may cause or exacerbate water column anoxia, potentially resulting in further fatalities of fish

and/or sessile fauna and flora (e.g. Jørgensen, 1980; Llansó, 1992). For example, the collapse of an

intense dinoflagellate (Gymnodinium spp.) bloom within the upper Swan River Estuary during

March 1994, produced water column anoxia that resulted in numerous fish kills (Kanandjembo,

1998). While it may be hypothesised that the comparatively sessile benthic fauna similarly suffers

mortality during periods of anoxia related to bloom collapse, the severity of stress to both benthic

and fish fauna populations will likely depend on the intensity of the algal bloom. Kanandjembo

(1998) found that an upper Swan River Estuary bloom of dinoflagellates during the summer of 1996

did not effect fish species composition but may have had a detrimental effect on their abundance.

However, there was no evidence that the bloom led to a significant change in either composition or

abundance of macrobenthic fauna. Finally, seagrasses may also be susceptible to mortality

20

Page 35: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

resulting from prolonged periods of anoxia and/or excessive growth of epiphytic algae (Scanes et

al., 1998). Aside from seagrasses providing important habitat, their importance to nutrient cycling

within the Swan River Estuary has been described by Connel and Walker (2001) (also see

Appendix D, Section D.3).

The fish community of the estuary has an inherent ecological and preservation value due to its high

diversity (137 species from 70 families), while the high productivity of commercial species

provides a valuable fishery (John, 1994). More than 300 tonnes of fish are commercially caught on

an annual basis within the estuary, with an equivalent catch rate estimated for the recreational

fishery (John, 1994). Problem algal blooms can have marked impacts on these fisheries. The loss

of fish may also infer cascading effects to higher (and lower) trophic levels. Water bird diversity

(87 species) and abundance (≤10 000 birds at any one time) of the Swan River Estuary is amongst

the highest of any wetland in Western Australia (John, 1994). There is great potential for negative

impacts of algal blooms on bird populations (via either direct toxic effects or indirect effects

through loss of food source), since bird abundance is typically greatest at the time of year

coinciding with most problem blooms, i.e. from mid-spring to autumn. However, it must be

recognised that ecological relationships are complex. A reduction in zooplanktivorous fish may

allow the proliferation of zooplankton that graze on phytoplankton, thereby regulating bloom sizes

(Kanandjembo, 1998; Griffin et al., 2001). Increasing zooplankton abundance may help to

facilitate the re-establishment of planktivorous fish larvae and adults (Kanandjembo, 1998). Within

the Swan River Estuary, examination of these trophic relationships has only recently been instigated

(e.g. see Kanandjembo, 1998; Griffin et al., 2001).

Whilst the administrative and public health costs associated with problem algal blooms may be

high, other social and economic costs become apparent when the Swan River Estuary is considered

as a highly important resource to recreation and tourism (John, 1994). More than 25,000 vessels are

registered for use within the estuary whilst activities of water skiing, rowing, fishing and swimming

are highly popular to both local residents and tourists (John, 1994). The substantial commercial

pleasure cruise industry uses the estuary as a basis for tourism (John, 1994). The economic impact

of closure, or disruptions in use, of the estuary therefore results in loss of income to the many

dependant operators of commercial fisheries and recreational activities. Perhaps more important in

both an economic and cultural sense is the somewhat intangible aesthetic value of the estuary. The

visual landscape of Perth surrounded by the picturesque waterways of the Swan River Estuary is

important to the identity of the city. Moreover, promotion of that image is undefinably valuable for

the attraction of tourists. A sustained decrease in the aesthetic quality of the Swan River Estuary

due to algal blooms could lead to substantial losses in tourism revenue.

21

Page 36: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Consequently, the eutrophication of the Swan River Estuary is an environmental, social and

economic problem in need of scientific investigation, management and remediation. Our

understanding of the many complex interactions between algal production, nutrients, estuarine

ecology, sediment biogeochemistry, hydrology and physical control factors is still in its infancy. It

is not my intent to summarise here the plethora of worthwhile avenues of research regarding

eutrophication management within the Swan River Estuary. However, it is important to emphasise

the central role of nitrogen, rather than phosphorous, as the nutrient with the greatest capacity to

limit phytoplankton blooms. Thompson (1998) notes that phosphate concentrations in the estuary

would have to be reduced by a factor of 15 to 30 times to become more limiting to algal biomass

than nitrogen; and that since such a reduction is unfeasible in the short term, control of summer-

time algal blooms is most likely to be achieved through reduction in nitrogen availability. Horner

Rosser and Thompson (2001: p. 2593) further state that: “If workable strategies for reducing bloom

intensity and frequency, through the manipulation of nitrogen availability, are to be successful for

the Swan River Estuary, then the various flux processes influencing available nitrogen need to be

considered. These processes appear to play an important role throughout the water column in

maintaining nutrient levels and to promoting and sustaining phytoplankton biomass, and thereby

influencing the pattern of phytoplankton bloom succession.”

In this respect, a consideration of the flux of nitrogen from the sediment surface is critical,

especially a recognition of the availability of ammonium near the sediment surface during the

period of highest algal productivity (November to May). While the heightened load and transport

of land-derived nutrients is ultimately responsible for the increased eutrophic status of the estuary

(and catchment nutrient reduction strategies are incredibly important!), it appears that management

of sediment nitrogen fluxes may play a vital role in the short to long term health of the estuary.

Benthic fauna are ubiquitous within the estuary and must be considered an integral constituent of

the sediment composition. An understanding of how benthic fauna influence sedimentary microbial

nitrogen cycling may ultimately allow management of benthic faunal communities to help regulate

the supply of nitrogen to the water column and prevent the deleterious consequences of algal

blooms.

22

Page 37: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

2.2 AUSTRALONEREIS EHLERSI AND SIMPLISETIA AEQUISETIS: POLYCHAETE

CONSTITUENTS OF THE SWAN RIVER ESTUARY’S BENTHIC

MACROFAUNAL COMMUNITY

2.2.1 Benthic Macrofaunal Community Composition of the Swan River Estuary

The overall theme of this thesis is a functional consideration of the two subject species,

Australonereis ehlersi and Simplisetia aequisetis, in the ecology and nitrogen-transforming

processes of the Swan River Estuary. Interpretation of the functional-importance of these two

polychaetes first necessitates an understanding of their positioning within the estuarine benthic

community. In terms of species richness, abundance and productivity, estuarine benthic

macrofaunal communities typically are dominated by polychaetes (nereidids, capitellids and

spionids), molluscs (bivalves and gastropods) and crustaceans (amphipods and decapods) (Day et

al., 1989; Rose, 1994; Kanandjembo et al., 2001). The overall dominance or ecological importance

of any one taxon (e.g. for the Polychaeta see Knox, 1977 and Wilson et al., 2003a) varies both

within and between estuaries, and over time (Day et al., 1989; Rose, 1994; Kanandjembo et al.,

2001; Pennifold and Davis, 2001). Spatio-temporal dynamism in community composition is a

manifestation of the response of constituent species to complex variations and interrelations of

governing environmental factors, both physicochemical (e.g. salinity, temperature and other

chemistry determined by hydrological regime and estuarine physiognomy) and/or biological (e.g.

competition, predation, trophic relations, dispersal mechanisms and evolutionary biogeographical

constraints etc.). As such, every estuary possesses a unique benthic faunal community, and scrutiny

of the distribution and ecological role, dominance or importance of constituent taxa must

necessarily be conducted on an estuary-specific basis. In this Section I consider briefly the broad

expression, and major determinants, of benthic species distribution, population structure and

cumulative community dynamics within the Swan River Estuary; in Appendix C, the position of

A. ehlersi and S. aequisetis within the benthic macrofaunal community of the estuary is discussed.

The Swan River Estuary may be defined by five distinct hydrological regions or zones: (i) the upper

estuary, (ii) the wide sill of Perth Water, (iii) the main basin (i.e. the deep trough and wide basin),

(iv) the shallow Fremantle sill and (v) Fremantle Harbour (see Section 2.1.2; Figure 2.2). To a

large extent, the characteristic hydrological, particularly salinity, regime of these zones governs the

species composition and distribution of the estuarine community (Chalmer et al., 1976; Hodgkin,

1987; Loneragan et al., 1989; Loneragan and Potter, 1990; Rose, 1994). Utilising similar species

composition (mainly of molluscs), Chalmer et al. (1976) separated the estuary into three main

regions or biotopes: the upper, middle and lower estuary (Figure 2.5). The biological and

hydrological definitions of the upper estuary are synonymous. The middle estuary is defined by the

23

Page 38: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

stretch of Perth Water to the wide main basin, excluding the deep trough. The deep trough,

Fremantle Sill and Fremantle Harbour together comprise the lower estuary. A similar classification

based on the community composition and distribution of teleost fish was presented by Loneragan

and Potter (1990).

Hodgkin (1987) argued that salinity regime and the tolerance of constituent taxa primarily

determine this tri-part zonation, which is likely applicable to the broad patterns of distribution and

composition of most biotic groups within the Swan River Estuary, including the Polychaeta. Based

on salinity tolerance, estuarine organisms are classified into the following groups (Chalmer et al.,

1976; Hodgkin, 1987; Rose, 1994):

(i) oligohaline species – freshwater affinity, limited estuarine representation and no marine

occurrence (salinity tolerance <0.5 g L-1)

(ii) ‘true’ euryhaline species – estuarine endemics with neither marine nor freshwater

representation (2 - 60 g L-1);

(iii) marine euryhaline species – marine affinity but with continuous estuarine

representation (5 - 36 g L-1)

(iv) stenohaline species - marine affinity with only temporary or sporadic estuarine

representation (25 - 40 g L-1).

UPPER ESTUARY

Main

Basin

Fremantle

Harbour

Fremantle Sill

Perth Water

Kent Street

Weir

The Narrows

Heirisson Island

Fremantle Traffic Bridge

Salter Point

Preston Point

Deep Trough

MIDDLE ESTUARY

LOWER ESTUARY

N

5 km

Figure 2.5: Biotopic regions of the Swan River Estuary (after Chalmer et al., 1976).

24

Page 39: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

The concept of a biological zonation of the Swan River Estuary into its lower, middle and upper

components implies the definition of boundaries to these zones, and implies a certain level of

stability in community composition within these bounds. While it is conceptually important to

define regions of ecological similarity, the interpretation that regional boundaries are fixed or that

the compositions of constituent taxa are immutable should be rejected. In large part, the valid

broad-scale resolution and recognition of the three biotic zones is determined by the fixed

physiognomic / bathymetric features of the estuary, primarily the sills at the junction between each

zone (Fremantle Sill and Perth Water). These features largely govern the hydrological processes

(including salinity regime) of each region which in turn regulate the broad-scale expression of

community composition. However, while important, physiognomy is only one regulator of

hydrological processes and other factors also impart control and variance on faunal distributions.

Many physicochemical and biological regulatory factors, and their complex relationships, will

ultimately define the expression of species’ population structure and community constituency

(further see Appendix C).

In the following sections I describe current knowledge about the distribution, habit and life history

of the study species, S. aequisetis and A. ehlersi, in the Swan River Estuary and elsewhere. The

environmental factors that may influence polychaete ecology, and particularly the distribution of

each species, will be discussed. However, prior to a consideration of the population biology of

A. ehlersi and S. aequisetis it is pertinent to first briefly discuss recent changes and confusion in

their taxonomy, since each species has previously been referenced under various synonyms and/or

misidentifications.

2.2.2 Taxonomy of A. ehlersi and S. aequisetis

Australonereis ehlersi and S. aequisetis belong to the Nereididae family. Within Australia, about

93 species of nereidids (representing 22 genera) are presently known, with the principal estuarine

genera represented by Australonereis, Olganereis and Simplisetia (formerly Ceratonereis sensu

lato, see below) (Wilson et al., 2003). Nereidids generally are distinguished by having the

following characters: an eversible pharynx with one pair of lateral jaws, a first segment with

tentacular cirri and a prostomium which is bluntly conical to trapezoidal (Wilson, 2000; Wilson et

al., 2003).

Detailed morphological description of A. ehlersi is provided by Hartman (1954), Hutchings and

Turvey (1982), Hutchings and Reid (1990), Wilson et al. (2003b) and Bakken and Wilson (2005).

The species was first described by Augener (1913) as Nereis (Leonnates) ehlersi and by Monro

(1938) as Leptonereis ehlersi. Hartman (1954) first defined the monospecific genus of

25

Page 40: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Australonereis, which differs most notably from other nereidid genera for having, on the ventrum,

paired fleshy transverse ridges/papillae lateral to the ventral cirri of anterior segments (these ridges

are absent in S. aequisetis) (Figures 2.6 and 2.7). Further, compared to S. aequisetis, A. ehlersi has a

flattened body and no simple falcigers (P. Hutchings, pers. comm.). For individuals of similar

length, A. ehlersi is broader and has a greater body volume than S. aequisetis (Hartman, 1954; pers.

obs.; Figure 2.6). Maximum body length may be dependant upon environmental setting. Hutchings

(1982) documents that A. ehlersi specimens may be up to 6 cm long and 0.5 cm wide; however,

Hartman (1954) notes that individuals from Lakes Entrance in Victoria may measure up to 14 cm

long and 1.2 cm wide. In Port Phillip Bay, A. ehlersi attains a length of 10-12 cm (Dorsey, 1981),

while the maximum body length in the Swan River Estuary is ~8 cm (Chapter 4).

Simplisetia aequisetis is described morphologically by Hutchings and Glasby (1985), Wilson et al.

(2003b) and Bakken and Wilson (2005). In comparison to A. ehlersi, the differentiating features of

S. aequisetis are a pharynx with scleroprotein paragnaths and posterior parapodia with simple

falcigers; in contrast, A. ehlersi has a pharynx with fleshy papillae in place of paragnaths

(P. Hutchings and R. Wilson, pers. comm.). Adults from both Port Phillip Bay and the Swan River

Estuary attain a body length of 4-6 cm (Dorsey, 1981; Chapter 4). Augener (1913) first described

the polychaete as Nereis (Ceratonereis) aequisetis1. The species descriptions by Hartmann-

Schröder (1982) and Hutchings and Glasby (1985) elevated the subgenus to full generic status, i.e.

Ceratonereis aequisetis. Hartmann-Schröder (1985) further reviewed species assigned to

Ceratonereis and recognised that the genus (comprising a large number of species ranging from

muddy estuaries to the deep sea), contained several unrelated groups of species. She therefore

formally defined three subgenera, two of which were new (R. Wilson, pers. comm.). Ceratonereis

(Simplisetia) included species with fused heterogomph falcigers in the dorsal neuropodial position,

including aequisetis (Bakken and Wilson, 2005). Finally, Khlebovich (1996) elevated the

Simplisetia subgenera to full generic status. Other synonym attributions have included C. cf.

anchylochaeta2 (Hartmann-Schröder, 1979) and C. pseudoerythraeensis (Hutchings and Turvey,

1982; Hutchings and Murray, 1984). The latter synonym was instigated as a new species to

delineate Australian nereidids which had previously been incorrectly identified as C. erythraeensis 1 The type material of both A. ehlersi and S. aequisetis collected by Augener (1913) originated from Melville

Water in the middle section of the Swan River Estuary. This is within the immediate vicinity of the long-term

sampling site for this study (see Chapter 4). 2 The synonymising of Hartmann-Schröder’s (1979) record of C. cf. anchylochaeta from Port Hedland likely

is incorrect; the species is more similar to the Indonesian C. anchylochaeta. Thus, S. aequisetis probably does

not occur this far north on the Western Australian coastline (Glasby, pers. comm.; Glasby, Wilson and

Bakken, in prep.).

26

Page 41: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

(Hutching and Turvey, 1982; also see below). Arguments for synonymy of S. aequisetis are detailed

in Hutchings and Glasby (1985), Wilson et al. (2003b) and Bakken and Wilson (2005).

(i) A. ehlersi (ii) S. aequisetis

(a) Dorsal

(b) Ventral

3 mm

3 mm

2 mm

2 mm

Figure 2.6: Dorsal (a) and ventral (b) photographs of A. ehlersi (i) and S. aequisetis (ii).

Figure 2.7: Ventral view of A. ehlersi setigers 10-17, illustrating the ventral papillae (‘pap’) that

characterise the genus (Hartman, 1954).

27

Page 42: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Given the previous taxonomic revisions and inaccuracies, when referencing Australian nereidids in

this thesis, I have made every effort to denote species under the present classification system of

Wilson et al. (2003b). Where known, the currently accepted taxonomy of other species groups has

also been utilised; otherwise, species are generally referenced as provided in source documents. It is

highly likely that previous identifications of C. erythraeensis in the Swan River estuary are

erroneous (e.g. Monro, 1938; Chalmer et al., 1976; Pennifold and Davis, 2001) and the species

should be referred to as S. aequisetis. While C. erythraeensis (now Simplisetia erythraeensis)

remains a valid species (described from Madagascar), all published reports of the species in

Australian waters are incorrect (R. Wilson, pers. comm.). Depending on the area, Australian

references to C. erythraeensis more likely indicates the presence of one or more of the following

species: S. aequisetis (southern distribution), S. amphidonta (South Australia, Victoria, Tasmania

and southern New South Wales), S. transversa (South Australia and Western Australia, including

near Bunbury), S. limnetica (Hawkesbury River, New South Wales) or S. turveyi (Merimbula, New

South Wales) (Hutchings and Turvey, 1982; Hutchings and Glasby, 1982, 1985). The only other

Simplisetia species (formerly Ceratonereis sensu lato) known from Australia is S. lizardensis from

a creek on Lizard Island, north-eastern Australia (Wilson et al., 2003b). Australian species of

Ceratonereis sensu stricto are not closely related to the Simplisetia and typically are marine

(Wilson, 2003a). Australian Ceratonereis spp. include C. australis (Hartmann-Schröder, 1985), C.

lapinigensis (Grube, 1878), C. longiceratophora (Hartmann-Schröder, 1985), C. mirabilis (Kinberg,

1866), C. perkinsi (Hartmann-Schröder, 1985), Ceratonereis singularis (Treadwell, 1929) and C. cf.

singularis (Treadwell, 1929).

Regarding populations in the Swan River Estuary, delineation of A. ehlersi and S. aequisetis is

important since they are apparently the only nereidids that inhabit the system. It is likely that past

taxonomic confusion has led not only to erroneous identifications of the two species, but perhaps

also wrongful combinations of the two species into one species group. For example, in a recent

study on the effect of benthic macrofaunal communities on nutrient cycling in the middle Swan

River Estuary, Pennifold and Davis (2001) list C. erythraeensis (i.e. S. aequisetis) amongst the ten

most abundant species in both winter and summer, but do not include the presence of A. ehlersi.

Their study site was adjacent to the long-term study site of this thesis, where both A. ehlersi and

S. aequisetis were found in considerable abundance, at least in one season (see Chapter 4). While it

is possible that A. ehlersi did not rank amongst the ten most abundant species in the study of

Pennifold and Davis (2001), a more likely suggestion is that they failed to recognise that two

species were present in the study and referred to all material as a single species (C. erythraeensis).

28

Page 43: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

2.2.3 Distribution and Feeding Habit of A. ehlersi

Swan River Estuary Population

Noting a general paucity of macrobenthic studies of estuaries, Dürr and Semeniuk (2000) comment

that little is known about estuarine polychaete population biology in Western Australia. As such,

limited information is available on the autecology of either A. ehlersi or S. aequisetis within the

Swan River Estuary. The review of Chalmer et al. (1976) noted that A. ehlersi has been recorded

only from the lower and middle estuary, with no records from the upper estuary or Perth Water

portion of the middle estuary (Figure 2.2). Studies by McShane (1977) and Kanandjembo (1998)

further detail an absence of A. ehlersi in the latter regions. Within the shallows of the Melville

Water portion of the middle estuary (Figure 2.2), Rose (1994) found A. ehlersi to be the sixth most

abundant polychaete but comprised only 0.5% of the total polychaete abundance (27 specimens out

of 5516 polychaetes sampled in total) (Appendix C, Table C1[B]). Individuals collected from this

region by McShane (1977) weighed between 50 and 500 mg (presumably wet weight). McShane

(1977) documented that A. ehlersi reside in densely aggregated burrows or tubes to a sediment

depth no greater than 15 cm. Further, he found that A. ehlersi is intolerant of depressed oxygen

conditions, but broadly tolerant of salinities ranging from 2 to 36 g kg-1 (and capable of osmotic

regulation of its coelomic fluid). During the winter study period, individuals in a reproductive state

were observed swimming in overlying waters of lowered salinity (~7 g kg-1), which was interpreted

as indicative of a pelagic spawning event, with larval stages tolerant of brackish salinities

(McShane, 1977). To my knowledge, no further information exists regarding the distribution,

biomass, productivity or life history of the Swan River Estuary population of A. ehlersi.

Australian Distribution

The known Australian distribution of A. ehlersi stretches around the southern half of the continent

from the Swan River Estuary in the southwest to the Calliope River Estuary in Gladstone,

Queensland (Hartman, 1954; Day and Hutchings, 1979; Hutchings, 1982; Hutchings and Turvey,

1982; Hutchings and Murray, 1984; Hutchings and Reid, 1990; Wilson et al., 2003b; Bakken and

Wilson, 2005). The species has also been documented from estuaries in most regions of Tasmania

(Edgar, 1999). This Great Southern biogeographic distribution largely equates to the temperate (i.e.

non-tropical) waters of Australia and is broadly shared with numerous marine fauna, including other

nereidids such as S. aequisetis and Olganereis edmondsi (Figure 2.8; Glasby et al., 2000).

Factors Determining Distribution

Similar to its abundance in the Swan River Estuary, A. ehlersi is not a dominant species in terms of

abundance within the neighbouring Peel-Harvey Estuary, Mandurah (Rose, 1994); nor in the nearby

29

Page 44: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Figure 2.8: The Great Southern Australian biogeographic distribution (dashed line) of

Australonereis ehlersi and Simplisetia aequisetis. Adapted from Glasby et al. (2000), after Wilson and Allen (1987).

Leschenault Estuary, Bunbury (Dürr and Semeniuk, 2000). A comparatively low density of

A. ehlersi may be a general characteristic of the species. In the estuary of the Hawkesbury River

(New South Wales), densities of A. ehlersi are about 6-fold lower than S. aequisetis (MacFarlane

and Booth, 2001), a pattern repeated in Tasmanian estuaries (Edgar, 1999). Nevertheless, dense

beds of A. ehlersi have been reported from some locations, e.g. at Lakes Entrance and in Port

Phillip Bay, Victoria (Hartman, 1954; Dorsey, 1981). Similar to the Swan River Estuary, A. ehlersi

was recorded only within the lower to middle region of the Leschenault Estuary (Dürr and

Semeniuk, 2000), which is suggestive of a marine euryhaline habit. The Lakes Entrance and Port

Phillip Bay populations of A. ehlersi in Victoria are located in waters of marine salinity (Hartman,

1954; Dorsey, 1981). Within the estuaries of New South Wales, A. ehlersi has been found to

underlie waters with salinities ranging from 5 to 37 g kg-1 (Hutchings and Murray, 1984). Given

McShane’s (1977) determination of A. ehlersi’s great capacity for osmoregulation, the general

absence of the species in the upper regions of Australian estuaries is confounding and perhaps

relates less to salinity than to other determining factors.

30

Page 45: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Rose’s (1994) macrobenthic surveys of the Swan River and Peel-Harvey Estuaries sampled only

sediments of shallow waters (depth <1.5 m). Dürr and Semeniuk (2000) found A. ehlersi to be

present in the deeper muddy sediments of Leschenault Estuary; however, they did not observe a

depth-preference as the species was also recorded in the shallows. Australonereis ehlersi is found

within both subtidal and intertidal sediments of Port Phillip Bay (Dorsey, 1981). In seeking to

identify the causal reasons for A. ehlersi distribution patterns in the bay, Dorsey (1981) implicated

factors other than water depth as being important, including a positive association with well-sorted

fine sand. In contrast, Dürr and Semeniuk (2000) did not find A. ehlersi to demonstrate a substrate-

preference, since the species was found within sand, mud and muddy sand. For estuaries of New

South Wales, Hutchings and Murray (1984) similarly note that the species may be found in

sediment substrates ranging from fine to coarse-grained sand to muddy sand. Australia-wide,

Hutchings and Turvey (1982) report that A. ehlersi may be found within sandy mud of estuarine or

sheltered bays, often associated with seagrass beds. An apparent association with stands of

Posidonia, Halophila or Zostera seagrasses, or Rhizophora mangroves, in muddy sand or sand-flats

has been further documented by Hutchings and Reid (1990).

Citing McShane’s (1977) findings of an intolerance towards low dissolved oxygen conditions,

Dorsey (1981) suggested that oxygen depression, associated with organic matter loading, was

implicated in excluding A. ehlersi from near sewage outfalls in Port Phillip Bay. A similar

argument was invoked to help explain the lack of A. ehlersi within the high intertidal zone, wherein

burrow waters are susceptible to hypoxia at low tide (Dorsey, 1981). Additionally, dense shell beds

in some parts of Port Phillip Bay may impart a physical barrier that preclude the construction of

A. ehlersi burrows. Within the bay, A. ehlersi is a prey item of wading birds and a species of

stingarie; predation is probably a significant influence on the size and structure of populations

(Dorsey, 1981; Wilson, 2000). Furthermore, Dorsey (1981) suggested that dense assemblages of

opportunistic species in areas of high organic matter loading may out-compete A. ehlersi for

infaunal space. Thus, many biotic and abiotic factors may potentially control the distribution,

abundance and population structure of A. ehlersi, likely in a synergistic manner (Dorsey, 1981).

Feeding Habit

Australonereis ehlersi is a strictly tubiculous species, inhabiting angular, U-shaped burrows to a

depth of 15 to 20 cm, the ends of which protrude slightly above the sediment surface (Hartman,

1954; McShane, 1977; Dorsey, 1981; Hutchings, 1982; Hutchings and Turvey, 1982; Hutchings and

Reid, 1990). However, documentation regarding the feeding habit of the species is largely

speculative. Rose (1994) considered A. ehlersi within the Swan River Estuary to be a non-specific

deposit feeder. Dorsey (1981) hypothesised that individuals may construct a mucous suspension

31

Page 46: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

feeding net to capture food; however, he too primarily considered A. ehlersi to be a deposit feeder.

In contrast, Rouse (2000) considered A. ehlersi to be raptorial and probably omnivorous, using its

powerful jaws attached to the eversible pharynx to grasp food that comes near the burrow

entrance/s. The common presence of large, curved jaws in the Nereididae is not necessarily

indicative of a predatory or raptorial habit; the jaws may be used for gathering sediments, as well as

for scavenging large food items (Dorsey, 1981; Wilson, 2000). Surface deposit feeding and

herbivory are the most common nereidid feeding habits, but the Family also contains omnivores,

carnivores and suspension-feeders (Wilson, 2000; Wilson et al., 2003a). Feeding behaviour can

vary even within and between populations of a single species (Wilson et al., 2003a).

2.2.4 Distribution and Feeding Habit of Simplisetia aequisetis

Swan River Estuary Population

Simplisetia aequisetis has been recorded from the lower, middle and upper regions of the Swan

River Estuary, typically comprising one of the most abundant macrobenthic species (Chalmer et al.,

1976; Rose, 1994; Kanandjembo, 1998; Kanandjembo et al., 2001; Pennifold and Davis, 2001).

Within shallow waters of the middle estuary (<1.5-2.0 m deep), Rose (1994) found that

S. aequisetis was more numerous than any other benthic macroinvertebrate collected (Appendix C,

Table C1[B]). At Pelican Rocks, a shallow water site in the middle estuary (very close to the long-

term study site of this thesis, see Chapter 4), Pennifold and Davis (2001) found S. aequisetis to be

the most abundant species in winter and the second most abundant species in summer (after the

bivalve mollusc Xenostrobus securis). The density of S. aequisetis retained by Rose (1994) on a 1

mm sieve ranged between 173 individuals m-2 at Dalkeith in the autumn of 1987 and 4074

individuals m-2 at Applecross during the spring of 1986. For the middle estuary, he found that S.

aequisetis density was significantly greater in summer than winter, and at Applecross compared to

other sites. Sampling both shallow and deep waters of the upper estuary using a 500 µm mesh

sieve, Kanandjembo (1998) found S. aequisetis to be the third most abundant macrobenthic species,

with the bivalve mollusc (Arthritica semen) and spionid polychaete (Pseudopolydora sp.) being

more numerous (Appendix C, Table C1[A]). The average density of S. aequisetis over her

sampling period from winter 1995 to autumn 1997 was approximately 250,000 individuals m-2 in

shallow waters and 13,000 individuals m-2 in deep waters. Rose (1994) documented that a sieve size

of 1 mm captured about the same abundance (>95%) as a mesh of 500 µm; thus, it appears that the

density of S. aequisetis may be greater in the upper than middle estuary. No information is

available on the typical abundance or other population statistics of S. aequisetis in the lower Swan

River Estuary.

32

Page 47: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

A wet weight of about 80 mg previously has been determined for large adult S. aequisetis

individuals in the middle estuary (De Roach et al., 2002)3. The only estimates of S. aequisetis

biomass within the Swan River Estuary are provided by Rose (1994), for the shallows of the middle

region. In terms of biomass, S. aequisetis is less dominant than three bivalve mollusc species

(Sanguinolaria biradiata, Xenostrobus securis and Musculista senhousia) but is the most dominant

polychaete species in this region (Appendix C, Table C1[C]; Rose, 1994). At his study sites, wet

weight biomass of S. aequisetis ranged between 5.2 g m-2 at Matilda Bay in winter 1986 and 100.5 g

m-2 at Applecross in spring 1987 (approximately corresponding to a minimum dry weight4 biomass

of 0.68 g m-2 and a maximum of 19.4 g m-2). Similar to patterns of density, S. aequisetis biomass

was significantly greater during spring and summer than autumn and winter, and at Applecross and

Deepwater Bay compared to other sites of the middle estuary. However, there was also a

significant season by site interaction, with the greatest biomass of S. aequisetis at Deepwater Bay

occurring in winter, indicating some site-specificity in the determinants of biomass (Rose, 1994).

Suggested determinants of the spatio-temporal distribution of S. aequisetis in the Swan River

Estuary and elsewhere are discussed below.

Australian Distribution

As noted above, the records of occurrence for S. aequisetis represent a Great Southern, Australian

distribution largely congruent with that of A. ehlersi, stretching around the southern half of the

continent from the Swan River Estuary in the southwest to the Calliope River Estuary in Gladstone,

Queensland (Figure 2.8; Hartman, 1954; Day and Hutchings, 1979; Hutchings, 1982; Hutchings and

Turvey, 1982; Hutchings and Murray, 1984; Hutchings and Glasby, 1985; Glasby et al., 2000;

Wilson et al., 2003b; Bakken and Wilson, 2005). Like A. ehlersi, S. aequisetis has also been

documented from Tasmania, but not from the estuaries of the southwest region of the island state

(Edgar, 1999). Where present, S. aequisetis typically is one of the most abundant constituents of

the polychaete assemblage, if not entire benthic macrofaunal community; including not only within

the Swan River Estuary, but also the Peel-Harvey Estuary, Leschenault Estuary, Hardy Inlet and

Wilson Inlet in Western Australia (Chalmer and Scott, 1984; Rose, 1994; Platell and Potter, 1996;

Dürr and Semeniuk, 2000); Port Phillip Bay in Victoria (Dorsey, 1981, 1982); and the estuary of

Hawkesbury River and many other estuaries of New South Wales (Hutchings and Murray, 1984;

MacFarlane and Booth, 2001) and Tasmania (Edgar, 1999).

3 N.B. – A typographical error exists in the Material and methods section of De Roach et al. (2002): ~ 0.8 g

wet weight should read ~ 0.08 g wet weight. 4 Dry weight calculation based on the assumption that: dry weight = [0.130 to 0.193] x wet weight (Cammen,

1980; Kristensen 1984; Sardá et al., 2000). For ash-free dry weight and/or carbon weight conversion factors, see the methodology section of Chapter 4.

33

Page 48: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Factors Determining Distribution

The inverse relation between S. aequisetis density and overlying water depth reported by

Kanandjembo (1998) within the upper Swan River Estuary has also been documented in other

Western Australian estuaries. Platell and Potter (1996) likewise found S. aequisetis to be far less

abundant in deeper than shallow waters of Wilson Inlet. Noting a reduction in abundance with

increasing sampling depth along transects, Dürr and Semeniuk (2000) described a depth-controlled

concentric distribution pattern of S. aequisetis in the Leschenault Estuary. By contrasting

incidences of S. aequisetis with tidal depth in Tasmanian estuaries, Edgar (1999) denoted the

greatest occurrence of specimens at ~30 cm below the low-water mark, with a trend of decreasing

records for both deeper waters and shallower (including intertidal) depths. In most estuaries, this

concentric distribution pattern is generally concomitant with the distribution of seagrasses and

benthic algae, a purported component of S. aequisetis’ diet (Platell and Potter, 1996). Australia-

wide, it is true that the distribution of S. aequisetis is often associated with estuarine algae,

seagrasses and mangroves (Hutchings and Turvey, 1982; Hutchings and Murray, 1984; Hutchings

and Glasby, 1985). Within the Swan River Estuary, Rose (1994) found S. aequisetis density to be

significantly positively correlated with the biomass of Ulva spp. (green algae), Gracilaria spp. (red

algae) and Halophila ovalis (seagrass). In Wilson Inlet, the relationship was true for S. aequisetis

and the resident seagrass Ruppia spp. (Plattel and Potter, 1996).

Whether or not the plant materials provide a food source that is a direct determinant of S. aequisetis

distribution and abundance, or the relationship is rather more indirect or coincident, requires further

investigation. For example, Platell and Potter (1996) also offered that seagrass habitat may provide

protection from predators. Indeed, Dorsey (1981) suggested that fish and invertebrate predators

consume S. aequisetis at subtidal depths in Port Phillip Bay, thereby helping to explain the species’

absence in this zone. An alternate suggestion is that the shallower photosynthetic zone of estuaries

is largely coincident with other determinant factors such as increased dissolved oxygen levels and

higher temperatures (Platell and Potter, 1996; Dürr and Semeniuk, 2000). Other physicochemical

(e.g. salinity, sediment particle size and organic matter content) and biological parameters (intra-

and inter-specific effects), and their interactions, may also correlate with small and large scale

distribution patterns of S. aequisetis.

In addition to the positive relationships with plant material described above, Rose (1994) also found

Swan River Estuary densities of S. aequisetis to be negatively correlated with the biomass of

Chaetomorpha spp. (green algae), Colpomenea spp. (brown algae) and driftwood. Causal reasons

for these negative associations are unclear, but may include habitat exclusion by these algal species

and/or driftwood, or oxygen stress implicated from their microbial breakdown. In his comparative

34

Page 49: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

study of the Peel-Harvey Estuary, the abundance of S. aequisetis was found to correlate positively

with biomass of H. ovalis but negatively with a species of red algae (Hypnea spp.). Further,

positive associations with coarse sand and redox depth (oxygen availability) were found to explain

most of the variation in S. aequisetis density in sediments of that estuary (Rose, 1994). For the

upper Swan River Estuary, Kanandjembo (1998) likewise argued that a possible reason for higher

S. aequisetis abundance in the shallows was the presence of coarser and less hypoxic sediments,

offering a more conducive environment for inhabitation. However, the association of S. aequisetis

with coarse substrates exhibited in the Peel-Harvey Estuary was not found in the Swan River

Estuary (Rose, 1994), nor is it usually demonstrated elsewhere. Within the Leschenault Estuary, for

example, Dürr and Semeniuk (2000) documented no specific substrate preference of S. aequisetis

since the nereidid was common at sites of sand, sandy mud and muddy sand. Throughout its

Australian distribution, S. aequisetis has been found to inhabit variously sand flats, muddy sand and

mud, both with and without the presence of shell (Dorsey, 1981, Hutchings and Turvey, 1982;

Hutchings and Murray, 1984; Hutchings and Glasby, 1985).

Furthermore, as already suggested by occurrence in the lower, middle and upper reaches of the

Swan River Estuary, S. aequisetis appears to be a true euryhaline species. Edgar (1999) documents

the incidence of S. aequisetis in Tasmanian estuaries with overlying waters ranging in salinity from

<5 g kg-1 to >37 g kg-1, with a tendency for higher salinities. Within Wilson Inlet, Platell and Potter

(1996) demonstrated similarly an inverse correlation of S. aequisetis density with bottom salinity.

In contrast, MacFarlane and Booth (2001) suggested that S. aequisetis is more abundant in areas of

lower salinity within the estuary of Hawkesbury River. It is generally accepted that S. aequisetis

has a distribution encompassing both estuarine and protected inshore marine environments

(Hutchings and Glasby, 1985; Wilson et al., 2003b). The nereidid has been collected from the

inshore marine coast adjacent to the Swan River Estuary at Point Peron (Kott, 1951).

Importantly, throughout his 15 month study of the shallow sediments of the middle Swan River

Estuary, Rose (1994) found that aside from plant materials, S. aequisetis density did not

significantly correlate with any other measured environmental variable, including: salinity,

temperature, dissolved oxygen, turbidity, sediment particle size, redox depth or organic matter

content. Furthermore, his analytical stepwise regression methodology could not define a single

environmental variable to explain spatio-temporal variations in S. aequisetis density. Within their

long-term study of patterns of S. aequisetis (and other polychaetes) abundance within the

Leschenault Estuary, Dürr and Semeniuk (2000) similarly noted a lack of coupling to seasonal

variation in abiotic factors. Furthermore, much variability in abundance occurred between sites,

within a single year and from year to year; with the population dynamics at different sites being

35

Page 50: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

largely asynchronous. They argued that such variability is a common and definitive trait of many

estuarine polychaete populations. Rather than a single dominating causal factor, it is likely that a

synergy of determinant variables governs the distribution and population dynamics of S. aequisetis

(and many other polychaete species) within any given estuary; and the combination of determinant

variables may be quite different in another estuary. Experimental manipulative population studies

are largely required, on an estuary-specific basis, to tease out the most important combinations of

physicochemical and biological determinants of polychaete distributions.

It is important to here reiterate that investigations of biological determinants of polychaete

distributions, e.g. competition, predation, other trophic relations, dispersal/reproductive

mechanisms, intraspecific interactions and biogeographical constraints, are often neglected,

particularly in multivariate studies which usually relate only easily measurable abiotic factors to

population parameters of species. To some extent, insufficient study of potential biological

determinants may contribute to the uncertainties concerning the factors controlling the spatio-

temporal distributions of S. aequisetis and other polychaete species. In a relatively unique study

arising out of the Ph.D. thesis of Dorsey (1981), Kent and Day (1983) found that wading shore birds

(red-necked stints, sharp-tailed sandpipers and curlew sandpipers) and fish (flounders) within Port

Phillip Bay feed selectively on larger individuals of S. aequisetis, which has a significant effect on

the structure of the nereidid population. Further, they document that a decreased density of

S. aequisetis adults increases juvenile recruitment, which apparently counter-balances losses from

predation (in terms of abundance). The Swan River Estuary is also an important habitat for

migratory waders, thus avian predatory effects may similarly be expected within intertidal areas.

Further, Simplisetia aequisetis is a known prey item of four species of fish (Sillaginodes punctatus,

Silago schomburgkii, Pomatomus saltatrix and Argyrosomus hololepidotus) from estuaries of

south-western Australia (Chalmer and Scott, 1984); with at least another two fish species

(Pseudogobius olorum and Amniataba caudavittata) found to consume unidentified nereidids –

S. aequisetis and/or A. ehlersi, within the Swan River Estuary (Kanandjembo, 1998). Largely

unknown are the influences of predation, interspecific competition and other biological interactions

as determinants of nereidid distribution and population structure in Australian estuaries.

Feeding Habit

A degree of herbivory was above insinuated in considering the determinants of S. aequisetis

distribution patterns. Whilst Rose (1994) notes that serrations on the jaws of S. aequisetis are well

adapted to manipulating large pieces of plant material, nonetheless caution should be exercised

when extrapolating jaw morphology to feeding habit. In addition to plant debris, Dorsey (1981)

found the diet of S. aequisetis to include amphipods, arthropod limbs, faecal pellets, foraminiferans,

36

Page 51: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

diatoms, the remains of other polychaetes and also freshwater cladocerans (i.e. Daphnia spp.

discharged with sewage wastewater). Gut contents included a large range of sediment particle

sizes, broadly suggestive of a non-selective deposit feeding habit (although very fine sediments may

not be ingested). The formation of faecal pellets was not observed in the gut. Whether these food

items are consumed with the sediment, or are raptorially seized, remains to be tested. Wilson (2000)

classified S. aequisetis as a scavenging deposit feeder.

Importantly, S. aequisetis may be non-tubiculous throughout most of its life but reproductive adults

have been found to construct burrows or tubes in which eggs are laid for brood protection (Dorsey,

1981, 1982; Hutchings and Turvey, 1982; De Roach et al., 2002). In relation to reproductive traits

(e.g. egg positioning), Dorsey (1981) describes the morphology of these tubes in great detail (also

see Chapter 4). Mucous is used to bind and line the burrows to make a resilient sheath (often even

withstanding the sieving collection process) and that these burrows may be irrigated (Dorsey,

1981). De Roach et al. (2002) also documented burrow construction to a depth of ~6cm in an

experimental situation and vigorous intermittent irrigation by large mature individuals of

S. aequisetis. The possibility that burrow ventilation may equate to some form of suspension

feeding requires further investigation.

37

Page 52: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,
Page 53: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

CHAPTER 3

THE NITROGEN CYCLE: PROCESSES, METHODOLOGY AND

BENTHIC MACROFAUNAL EFFECTS

3.1 PROCESSES DEFINING THE NITROGEN CYCLE

Prior to outlining the impacts of the polychaete and benthic fauna on sedimentary nitrogen

regulation, it is relevant to define the components of the estuarine nitrogen cycle. While an

increased total nitrogen supply is fundamental to the eutrophication of an estuary, and the

dynamics of physical nutrient transport processes are highly important (Section 2.1), it is the

rate of the various microbial nitrogen transforming processes that determine the availability of

nitrogen for algal growth. A summary of the estuarine nitrogen cycle is presented in Figure 3.1,

and the biochemistry, biology and relevance of each nitrogen-transforming process, including:

(i) description of bacterial and enzymatic processes; (ii) consideration of the main factors

regulating ambient rates; (iii) a review of ambient rates in aquatic environments; and (iv) a

review of previous studies examining the enzyme kinetics of denitrification, is provided in

Appendix D.

Denitrification is the ultimate process that liberates gaseous nitrogen from an aquatic

environment (Figure 3.1). The process therefore has potential to act as a nitrogen sink and

alleviate the symptoms of eutrophication. Consequently, denitrification is often the focus of

aquatic research in situations where an abundance of nitrogen has caused deleterious effects.

However, denitrification is only one component of the nitrogen cycle and must be put in context

amongst other components. Each step of the nitrogen cycle is related integrally to all other

processes occurring within it (Figure 3.1).

The potential rate of denitrification within any ecosystem depends upon the reaction rates of all

the other nitrogen transforming processes occurring within that system. Within estuarine

environments, large inorganic nitrogen inputs may be assimilated rapidly into organic material

(definitive of eutrophic estuaries with algal blooms), with a subsequent tight recycling of the

stored nitrogen pool (Capone, 2000). Alternatively, physical processes or typically high

denitrification rates may transport rapidly inorganic nitrogen inputs from the system.

Assimilatory and dissimilatory nitrate reduction to ammonium (DNRA) pathways of nitrate

reduction will be important as competing pathways of nitrate utilisation and will conserve

nitrogen within the system rather than release it. Nitrification may be an important process in

providing the substrate (nitrate) required for denitrification. Both of the latter processes are

typically of quantitative importance in estuarine systems (Capone, 2000). The rate of

nitrification may depend, in turn, on the rates of ammonification and nitrogen fixation, both of

39

Page 54: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

40

which provide ammonia to the system (though the latter process typically is not prevalent).

Alternately, ammonia may be taken up by assimilatory processes and away from the

denitrification pathway. Therefore, if ammonium rather than nitrate is the main source of

nitrogen assimilated into algal blooms, the process of denitrification may be relatively

unimportant in the management of those blooms, unless there is a high rate of coupled

nitrification. It is also becoming increasingly apparent that the novel processes of nitrifier-

denitrification, anammox and heterotrophic nitrification / aerobic denitrification (Appendix D),

add complexity of quantitative importance to ecosystem nitrogen budgets, with the processes

perhaps rivalling conventional denitrification in the production of nitrogenous gases. The

potential for the presence of these processes in estuarine nitrogen cycling systems should, at the

very least, be considered (see Appendix D).

Consequently, an understanding of each nitrogen cycling component is crucial if effective

management of aquatic ecosystems with nitrogen related environmental problems is to be

achieved. Furthermore, the quantification of the rate of turnover of the various nitrogen

compounds, together with the quantitative estimation of the extent to which different biological

and abiotic processes contribute to the cycle in different ecosystems, may allow the successful

management of those ecosystems. For this reason, I have attempted in this study to elucidate

the biological impact of benthic fauna on both denitrification and also on the other important

components of microbial nitrogen cycling within estuarine sediments (Chapter 5).

3.2 NITROGEN CYCLING METHODOLOGY

Quantification of denitrification, in addition to being a focal process of importance, is often a

useful first step in a holistic investigation of the nitrogen cycle. Denitrification activity may be

measured directly by dinitrogen gas production or via a range of indirect approaches based on

(i) measurements of nitrate flux, (ii) application of metabolic inhibitors (e.g. acetylene),

(iii) utilisation of nitrogen isotope tracers (including the isotope pairing technique), (iv) mass

balance and stoichiometric modelling and/or (v) examining kinetic parameters of the

denitrifying community (Seitzinger, 1988, 1993; Nielsen, 1992; Joye et al., 1996). The isotope

pairing technique (Nielsen, 1992) has recently become a common and efficient method to adopt

when access to nitrogen isotope supplies is viable; however, each methodology has its own

assumptions and limitations (Tiedje, 1988), of which Seitzinger (1988, 1993) provides a

comparative review. No one approach has been the obvious method of choice for studies

measuring denitrification, but since 1980 the acetylene inhibition method has often been the

method of first consideration because of its low cost and convenience (Tiedje, 1988; Groffman

et al., 1999). The combination of this method with a kinetic approach, i.e. the kinetic-fix

adaptation of the acetylene block (Joye et al., 1996), provides a powerful technique to measure

not only denitrification but also other nitrogen cycling processes in estuarine sediments. I used

Page 55: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

41

NO3-

NH4+

Organic N

NO2-

N2O N2Organic N

NH4+ NO3

- NO2-

N2O N2

NH4+

NO3-

NO2-

Organic N

Export

NH4+

NO3-

NO2-

Organic N

Import

NH4+, NO3

-, NO2-

Groundwater Import

Ammonification / Mineralisation

Nitrification

Denitrification

Sediment / Water Column / Atmosphere Solute Flux

ATMOSPHERE

WATER COLUMN

SEDIMENT

Nitrogen Fixation

Ammonium Assimilation

Assimilatory Nitrate Reduction

Dissilmilatory Nitrate

Reduction to Ammonium

LEGEND

Figure 3.1: The estuarine nitrogen cycle, showing the main transformations occurring in the sediment and water column. For simplicity many intermediates have been omitted. See also Appendix C.

Page 56: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

a hybridised kinetic-fix acetylene block method in my study. The background theory, advantages

and limitations of both the acetylene block method and general kinetic approach to denitrification

are given below. To give context to results using combined methodology, a comprehensive review

of outcomes is provided from studies based on a kinetic approach to denitrification, from a variety

of environments, in Appendix D.

3.2.1 The Acetylene Block

Nitrous oxide accumulation in soil during exposure to acetylene (C2H2) was first reported in 1973

(Oremland and Capone, 1987). The gas inhibits the nitrous oxide reductase of denitrifying bacteria

(Figure 3.2; Balderston et al., 1976; Yoshinari and Knowles, 1976). The rate of nitrous oxide

evolution is thus a measure of denitrification. The finding was significant because the direct

approach of measuring dinitrogen gas previously had been difficult to achieve (Tiedje, 1988). In

sedimentary environments, most applications of the acetylene inhibition technique for the study of

denitrification have been based on incubation of sediment cores with acetylene and subsequent

extraction of the evolved nitrous oxide after a relatively short (a few hours) incubation period

(Revsbech and Sørensen, 1990).

A potential major problem when using the acetylene inhibition method alone to investigate

denitrification is concurrent blockage of nitrification activity (Figure 3.2; Oremland and Capone,

1987; Kuenen and Robertson, 1988; Tiedje, 1988; Revsbech and Sørensen, 1990; Groffman et al.,

1999). Acetylene blocks the nitrification process by inhibiting the oxidation of ammonium to nitrite

(Hynes and Knowles, 1978, 1982; Walter et al., 1979; Mosier, 1980). The effect of nitrification

inhibition may be small in sediments where there is a high concentration and supply of nitrate from

the overlying water (i.e. where this source of nitrate dominates substrate supply for denitrification);

however, there have been many cases where the acetylene inhibition technique did not satisfactorily

quantify denitrification in situ because of the absence of a natural nitrate supply from nitrification

(i.e. where there is tight coupling of denitrification with nitrification) (Jenkins and Kemp, 1984;

Oremland et al., 1984; Slater and Capone, 1989; Revsbech and Sørensen, 1990). Therefore, nitrate

is often supplied artificially as a substrate at natural or saturating concentrations, and the resultant

nitrous oxide flux is considered the ‘potential’ denitrification rate.

The lack of nitrification in the presence of acetylene is beneficial to studies where the nitrate

substrate supply for denitrification is required to be somewhat precisely controlled or manipulated

(e.g. in approaches based on enzyme kinetics, see below), and/or where it is useful to discount the

nitrification component. In such experimental circumstances (i.e. with C2H2), the absence of

biological nitrate production allows the initial nitrate concentration to be fixed (such that ambient

42

Page 57: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

N2O

N2 NO2-

NH4+ NO3

-

Organic N

Ammonification / Mineralisation

Nitrification

Denitrification

Process blocked by acetylene

Nitrogen Fixation

Ammonium Assimilation

Assimilatory Nitrate Reduction

Dissilmilatory Nitrate

Reduction to Ammonium

LEGEND

C2H2

\

C2H2

\

C2H2

\

Figure 3.2: Acetylene inhibition of nitrous oxide reduction and ammonium oxidation and effects on nitrogen cycling.

concentration cannot increase due to nitrification) and the resultant rate of nitrous oxide production

is a measure of denitrification, expressed relative to the initial nitrate concentration. When nitrate

consumption is concurrently determined, the resultant rate of nitrate loss is due to denitrification,

dissimilatory nitrate reduction to ammonium (DNRA) and assimilatory nitrate reduction (Figure

3.2). Hence, subtraction of the known rate of denitrification (determined via nitrous oxide

production) yields the DNRA plus assimilation component of nitrate loss (Blackburn, 1986;

Revsbech and Sørensen, 1990). Furthermore, dark incubation (and high background ammonium

concentrations) inhibits photosynthetic nitrate assimilation (Kristensen et al., 1985) such that the

component of nitrate consumption by DNRA alone can be determined. This is denoted overleaf.

43

Page 58: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Where ‘Δ [NO3--N]’ and ‘Δ [N2O-N]’ are the measured fluxes of nitrogen (N) respectively

incorporated in nitrate and nitrous oxide; and ‘Nitrification’, ‘Denitrification’, ‘Nitrate

Assimilation’, ‘DNRA’, ‘Ammonification’, ‘Ammonium Assimilation’ and ‘Nitrogen Fixation’

refer to the rates of N turnover via these processes (all set positive), then:

(1) In the absence of acetylene (-C2H2):

Δ [NO3--N] -C2H2 = Nitrification – (Denitrification + DNRA + Nitrate Assimilation)

(2) In the presence of acetylene (+C2H2):

(a) Δ [NO3--N] +C2H2 = – (Denitrification + DNRA + Nitrate Assimilation)

AND

(b) Δ [N2O-N] +C2H2 = Denitrification

THUS

(c) Δ [NO3--N] +C2H2 + Δ [N2O-N] +C2H2 = – (DNRA + Nitrate Assimilation)

(3) In the presence of acetylene and in darkness:

(a) Δ [NO3--N] +C2H2,dark = – (Denitrification + DNRA)

AND

(b) Δ [NO3--N] +C2H2,dark + Δ [N2O-N] +C2H2 = – DNRA

Thus, if DNRA can independently be determined to be negligible (often the case in estuarine

sediment; sensu Capone (1988, 2000), but see Appendix D, Section D.6), nitrate consumption in

dark acetylene incubations provides a good measure of denitrification activity, even when nitrous

oxide production is not measured. That is:

(4) In the presence of acetylene, in darkness and when DNRA ≅ 0 (DNRA):

Δ [NO3--N] +C2H2,dark,DNRA = – Denitrification

Further, when ammonium-bound nitrogenous flux (Δ [NH4+-N]) is considered in the same

progression of experimental manipulation, then:

(5) In the absence of acetylene (-C2H2):

Δ [NH4+-N] -C2H2 = (Ammonification + DNRA + Nitrogen Fixation)

– (Ammonium Assimilation + Nitrification)

44

Page 59: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

(6) In the presence of acetylene (+C2H2):

Δ [NH4+-N] +C2H2 = (Ammonification + DNRA + Nitrogen Fixation)

– (Ammonium Assimilation)

(7) In the presence of acetylene and in darkness:

Δ [NH4+-N] +C2H2,dark = (Ammonification + DNRA + Nitrogen Fixation)

Finally, if both DNRA and nitrogen fixation can be determined independently to be negligible

(which often is the case in estuarine sediment; sensu Postma et al. (1984) and Capone (1988, 2000);

but see Appendix D, Sections D.2 and D.6), then ammonium production provides a good measure

of ammonification (although animal excretion and microbial decomposition components cannot be

separated, unless independently determined). Furthermore, acetylene (an analogue of N2) is

reduced preferentially to ethene (C2H4) by nitrogenase activity of nitrogen fixers, such that

ammonium production by nitrogen fixation is masked (Capone, 2002). That is:

(8) In the presence of acetylene, in darkness and when both DNRA ≅ 0 and nitrogen fixation ≅

0 (Nfix):

Δ [NH4+-N] +C2H2,dark,DNRA,Nfix = Ammonification

Manipulation of the presence/absence of acetylene and darkness in an experimental system,

combined with quantification of nitrate, nitrous oxide and ammonium flux rates therefore can yield

substantial information regarding denitrification and other nitrogen transformations in that system,

particularly when it is possible to verify independently the contribution of DNRA and/or nitrogen

fixation.

Limitations to the acetylene block method need to be considered. One, acetylene inhibition of

nitrous oxide reduction may be incomplete at low (<10 μM) nitrate concentrations, leading to an

underestimation of denitrification where nitrous oxide production is measured (Kaspar, 1982;

Oremland et al., 1984; Slater and Capone, 1989; Rudolf et al., 1991; Seitzinger, 1993; Groffman et

al., 1999). Two, the presence of sulphide hinders acetylene inhibition (Sørensen et al., 1987), and

methods other than the acetylene block (e.g. nitrogen isotope tracing) generally should be applied

where high sulphide concentrations are naturally present (Revsbech and Sørensen, 1990). Revsbech

and Sørensen (1990) have also highlighted the problem of the production of a more oxidised end

45

Page 60: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

product of denitrification after acetylene addition: per nitrate molecule, the reduction to nitrous

oxide requires four electrons whilst reduction to dinitrogen gas requires five electrons. Where the

amount of available organic carbon (as an electron donor) is a limiting factor, the fewer electrons

required in the presence of acetylene may lead to an overestimation of the natural denitrification

rate. It is difficult to quantify this overestimation in the realistic situation where nitrate is supplied

by diffusion to a stratified microbial community (Revsbech and Sørensen, 1990). In the presence of

acetylene, a decreased penetration depth of nitrate may be expected due to the increased demand for

nitrate in surficial and successively deeper layers (Revsbech and Sørensen, 1990). As Revsbech

and Sørensen (1990: p. 263) have stated, “the exact elevation of the total rate of nitrate reduction by

addition of acetylene then depends on the microdistribution of available electron donors”.

The use of acetylene is complicated further when using the inhibitor with sediments inhabited by

fauna. In vitro injection of acetylene saturated water into uninhabited sediment cores through

silicon-rubber injecting ports at appropriate depths has been used successfully by, for example,

Sørensen (1978a), Andersen et al. (1984), Jensen et al. (1988) and Jørgensen and Sørensen (1988)

to quantify large-scale regional and seasonal variation in denitrification activity in sediments.

However, injections into small cores inhabited by infauna are not possible since there is a danger of

damaging or killing the burrow residents (Kristensen et al., 1991; Svensson, 1997). Kristensen et

al. (1991) overcame the problem by introducing inhabited cores to an acetylene-sparged reservoir

for a long period of time (20 h) to ensure an even distribution of acetylene into the denitrifying

sediment volume prior to incubation. Furthermore, Revsbech and Sørensen (1990: p. 262) have

noted that “the denitrification zone within shallow sediments is generally found so close to the

surface or to ventilated infaunal burrows that diffusional supply of acetylene from the overlying

water should suffice to stop nitrous oxide reduction, and it may thus not be necessary to perform the

injection of acetylene in such sediments”. Additionally, “the injection of relatively large volumes

of water does disturb the natural chemical gradients in the sediment core to some extent, and it

would be an advantage to avoid this step” Revsbech and Sørensen (1990: p. 262). An additional

consideration is that acetylene may directly affect the physiology or metabolism of benthic fauna.

Whilst Caffrey and Miller (1995) observed acetylene to have negligible effect on faunal excretion

(of a polychaete and mollusc), elevated rates of ammonium excretion (Kristensen et al., 1991) and

depressed irrigation activity (Chapter 5) by polychaetes are known to occur from exposure to

acetylene. It is important to be aware of such limitations in the application of the acetylene block

method so that appropriate control measures can be taken and/or recognition of potential

implications can be given.

46

Page 61: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

3.2.2 Kinetics of Denitrification

Kinetic analysis is a useful and informative approach to the study of enzyme-driven denitrification.

Kinetic description defines the rate of denitrification as a function of another parameter such as

substrate (nitrate) concentration, enzyme concentration, temperature or pH (Saier Jr., 1987). The

most common method of kinetic analysis is to plot the steady state rate of denitrification (V0)

against substrate concentration (S) (Figure 3.3). The resultant rectangular hyperbola (Figure 3.3) is

defined by a Michaelis-Menten equation:

S KSV V

m

max 0

= Equation 1

where Vmax is the maximal denitrification rate and Km is the nitrate concentration at which

denitrification rate is half maximal, i.e. the Michaelis or half-saturation constant (Firestone, 1982;

Knowles, 1982; Copeland, 1996; Joye et al., 1996). This model can be used to describe systems

more complex than a single-enzyme substrate reaction (Nedwell, 1975). The definition of Km as it

relates to denitrification activity by heterogeneous bacterial populations is not the same as in single

enzyme kinetics, rather it is a useful ecological parameter that describes a population’s hyperbolic

response of denitrification activity to increasing ambient nitrate concentration (Williams, 1973;

Nedwell, 1975). Where denitrification is the net result of enzyme activity from a cohort of

denitrifying organisms in sediment, for example, the kinetic constants should be classed as

“apparent” (Joye et al., 1996). Joye et al. (1996) refer to the apparent Michaelis constant as Kapp

and to the maximum potential denitrification rate as Vmp. Kapp represents a measure of the

physiological adaptation of a denitrifying population to reduce nitrate to nitrogenous gases

(Nedwell, 1975). However, the value of Kapp may be influenced by the physical transport rate of

nitrate to the denitrifiers (i.e. by diffusion or active transport processes), and nitrate availability

must be considered in addition to the general ambient concentration (Nedwell, 1975). Vmp may be

related positively to denitrifier population size but increases in denitrification rate will remain

proportional to nitrate concentration for a steady-sized population (Nedwell, 1975). Therefore, Kapp

should also remain constant for an increased denitrifier population size assuming that no size-

related physiological adaptation occurs (Nedwell, 1975). These kinetic constants are inherently

informative of the denitrification system under study; furthermore, their approximation and

substitution into the Michaelis-Menten equation also allows for an estimate of denitrification rate at

any given ‘natural’ background nitrate concentration (Joye et al., 1996). Natural in situ rates of

denitrification are generally much lower (e.g. 0.1-2.5%; Kaspar, 1982) than maximum potential

rates (Kristensen et al., 1985).

47

Page 62: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Nitrate concentration (S)

Den

itrifi

catio

n ra

te (V

0)

Km

Vmax

1/2 Vmax

Figure 3.3: Michaelis-Menten curve of enzyme-catalysed denitrification. The maximal denitrification rate is defined as Vmax. The Michaelis or half-saturation constant, Km, is the nitrate concentration at ½ Vmax. (Adapted from Saier Jr., 1987, p. 40).

Typically, the Michaelis-Menten (S vs V0) curve is derived empirically and the kinetic constants are

estimated via a linear transformation. For example, by rearrangement of Equation 1 we get:

⎟⎠⎞

⎜⎝⎛+⎟

⎠⎞

⎜⎝⎛=

max

m

max0 VKSx

V1

VS

Equation 2

The resultant linear Hanes-Woolf plot of S/V0 as a function of S (Figure 3.4), allows the kinetic

constants to be estimated from the gradient (1/Vmax) and y intercept (Km/Vmax) (Copeland, 1996;

Joye et al., 1996). Other linear transformations include the Lineweaver-Burk and Eadie-Hofstee

plots (Copeland, 1996). Whilst these rearrangements are each mathematically equivalent, the

measurement of V0 and, to a lesser extent, S is subject to experimental error which affects the

aptitude of the different plots (Dowd and Riggs, 1965). For instance, plotting the reciprocal of a

variable (V0 or S) places undue emphasis on the error associated with the smallest values of that

48

Page 63: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

variable (Dowd and Riggs, 1965). The Lineweaver-Burk method involves plotting against each

other the reciprocal of both the independent variable (S) and the dependent variable (V0) (Copeland,

1996). As such, Dowd and Riggs (1965) strongly argue that the method should be abandoned in

favour of one of the other two aforementioned linear transformations.

The Michaelis-Menten equation describes steady state kinetics. The term steady state applies to

experimental conditions in which the enzyme-substrate complex is at an appreciable steady state

level throughout the course of the denitrification reaction (Copeland, 1996). Copeland (1996: p. 95)

notes that “Experimentally one finds that the time course of product appearance and substrate

depletion is well modelled by a linear function up to the time when about 10% of the initial

substrate concentration has been converted to product.” It is the denitrification rate measured

during the initial linear period of substrate loss that comprises V0 (Copeland, 1996). To satisfy the

assumption of steady state kinetics during the empirical determination of V0 for any given initial

nitrate concentration (S), it is important, therefore, to ensure that background nitrate levels do not

depreciate by more than 10%.

Nitrate concentration (S)

S /

V0

Gradient = 1 / Vmax

Intercept = Km / Vmax

Figure 3.4: Hanes-Woolf linear transformation of a Michaelis-Menten curve. Nitrate

concentration (S) divided by denitrification rate (V0) is plotted against S. The reciprocal of the gradient can be used to estimate the maximal denitrification rate (Vmax). The y intercept multiplied by Vmax then yields an approximation of the Michaelis constant (Km) (Adapted from Copeland, 1996, p. 93).

49

Page 64: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

3.3 THE IMPORTANCE OF BENTHIC MACROFAUNA IN SEDIMENTARY

NITROGEN CYCLING

Darwin championed the idea that infauna are capable of profoundly influencing their habitat, both

biochemically and biomechanically (Johnson, 2002). In his book The formation of mould through

the action of worms, Darwin (1881) demonstrated how the subsurface feeding activity of

earthworms and deposition of faecal castings at the surface actually created the topsoil and stratified

environment in which they live. Johnson (2002: p. 8) notes how Darwin’s ideas “spawned a spate

of biochemical and biomechanical observations in the late 19th and early 20th centuries” across

many fields of study, including marine sedimentology (e.g. Davison, 1891). However, it has been

only in the last 30 years that similar biogeochemical observations have been applied to concepts of

explanatory importance in the paradigms of related disciplines (e.g. archaeology, geomorphology,

pedology). This gap in recognition of faunal influence upon substrate processes results from many

complex factors, chiefly among them: (i) the early practice of limiting world-views and an adoption

of conceptual frameworks in which biomechanical/biochemical processes were largely ignored; and

(ii) lack of a genetic language and supporting theory, including terms such as bioturbation, to

describe and showcase the importance of such processes (Johnson, 2002). The interpretations hold

particularly true when considering the historical development of sedimentary biogeochemistry.

Prior to the 1980s, Aller and Yingst (1978: p. 201) observed that “most studies of near interface

sediments assume that chemically and biologically important properties are stratified vertically in a

deposit”. Their study was one of the first to highlight the importance of the benthic fauna that

bestow localised effects and impart a three-dimensional heterogeneity to the typical sediment

environment: animal burrows, for example, induce a radial plane of diffusion (also see: Gray, 1974;

Rhoads, 1974; Goldhaber et al., 1977; Grundmanis and Murray, 1977; Kikuchi and Kurihara, 1977;

Petr, 1977; Schink and Guinasso, 1977; Aller, 1978, 1980a-c; Billen, 1978; Rhoads et al., 1978; Lee

and Swartz, 1980; McCaffrey et al., 1980; Zeitzschel, 1980). The consolidation of ideas pertaining

to the role of benthic fauna in sediment biogeochemistry (particularly nitrogen cycling) is a major

theme of my thesis and is discussed further in Chapters 4 to 7. Historically, the catalyst driving

early research and insights into benthic faunal effects was eutrophication of aquatic ecosystems and

the realisation that nutrient regeneration in the benthos is an important regulator of primary

productivity in shallow aquatic ecosystems.

In the early 1970s it was realised that nutrient availability generally tended to determine the extent

of primary production in aquatic ecosystems (Rhyther and Dunstan, 1971). Whilst phosphorous

supply was typically implicated in limiting productivity of freshwater environments (Nixon, 1995),

50

Page 65: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

it became understood that nitrogen availability governed the magnitude of primary production in

most estuarine and coastal marine systems (Rhyther and Dunstan, 1971; Goldman et al., 1973;

Nixon et al., 1976; Carpenter and Capone, 1983; Seitzinger, 1990) and in some freshwater lakes,

streams and rivers (Keeney, 1973; Gerhart and Likens, 1975; Seitzinger, 1990). As detailed in

Section 2.1, the discovery and extensive global use of fertilisers during the past century,

compounded by the consequences of human population expansion, led to significantly increased

exports of nitrogen and other nutrients to near-shore aquatic environments (Rabouille et al., 2001).

Many freshwater lakes had become eutrophic in the earlier part of the century, but it was not until

the 1970s that eutrophication of the world’s rivers, estuaries and coastal oceans began to occur on a

broad scale (Cloern, 2001). These water bodies, in which primary productivity had previously been

limited by nutrient availability, were now increasingly subject to algal growth of excessive intensity

and frequency (Pedersen and Borum, 1996). The environmental and socio-economic implications

of eutrophication and problem algal blooms were, and continue to be, far-reaching; ranging from

unsightliness to acute toxic effects and/or complete ecosystem collapse (Section 2.1.3; for an

excellent review see de Jonge et al., 2002). The widespread manifestation of these eutrophic side-

effects in the coastal aquatic environments of the developed world during the 1970s (particularly in

western Europe and in the north-east and south-west of the U.S.A.), is likely the main reason that

spurred researchers to investigate thoroughly the factors regulating primary productivity in aquatic

ecosystems.

Increased nutrient loading arising from accelerated movement of nutrients from the lithosphere to

the hydrosphere had been implicated as the primary causative agent of excessive primary

productivity (Cloern, 2001). However, allochthonous nutrient supplies alone often could not

explain the high rates of algal growth in eutrophic waters. For example, nutrient-laden freshwater

runoff is typically diminished during the dry summer of temperate regions and cannot account for

the high rates of near-shore productivity frequently observed during this period (Rowe et al., 1975).

For many water bodies, it became obvious that tight internal mechanisms of organic algal matter

breakdown and nutrient re-supply (i.e. autochthonous nutrient regeneration), within both the pelagic

and benthic zones, were highly important factors regulating primary productivity (at least during

certain times of the year). Indeed, Nixon et al. (1976: p. 271) predicted that benthic nutrient

regeneration likely played only a minor role in the surface productivity of deep oceans (wherein

pelagic regenerative processes were expected to dominate5); however, they argued that: “the

5 Eppley and Peterson (1979) later highlighted the fact that the majority (82-94%) of nitrogen necessary to support primary production in deep oceanic waters was regenerated from organic matter in the pelagic zone; and Smith et al. (1978) and Zeitschel (1980) documented that nutrient effluxes of deep ocean sediments are two to four orders of magnitude less than those of shallow coastal sediments.

51

Page 66: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

importance of benthic communities may be expected to increase dramatically in coastal waters, and

it is surprising that so little is known about their role in nutrient cycling or in maintaining the

characteristically high levels of productivity found in shallower marine and estuarine regions”.

The significance of benthic nitrogen regeneration in supporting primary production of aquatic

ecosystems had previously been highlighted by the classic limnological study of Mortimer (1941,

1942). A similar role of the marine benthos was conceptualised by Rittenberg et al. (1955);

however, their findings that the deep-sea basin sediments adjacent to southern California provided

only 0.4% of the overlying phytoplankton’s nitrogen requirements perhaps retarded recognition of

the process within the marine environment. Nevertheless, Barnes (1957) hypothesised that near-

shore nutrient regeneration occurred in bottom sediments and Pomeroy et al. (1965) showed benthic

phosphate regeneration to be an important nutrient source for estuarine plants and phytoplankton.

However, it was not until the widespread coastal eutrophication crisis of thirty years ago that

nitrogen recycling by the benthos was fully recognised to be a major determinant of primary

productivity in shallow near-shore aquatic ecosystems (Hale, 1974; Hartwig, 1974; Davies, 1975;

Rowe et al., 1975).

Between 20-100% of the nitrogen requirements of phytoplankton in shallow estuarine and marine

environments are provided by benthic regeneration (Hale, 1974; Davies, 1975; Billen, 1978;

Boynton et al., 1980; Aller and Benninger, 1981; Nixon, 1981; Callender and Hammond, 1982).

However, in certain areas or at certain times of the year, sediment inorganic nitrogenous effluxes

either exceed algal demand by 200-800% (Hale, 1974; Rowe et al., 1975; Boynton et al., 1980) or,

conversely, provide only a negligible (0-16%) contribution of nitrogen towards primary

productivity (Hartwig, 1976; Boynton et al., 1980; Hopkinson and Wetzel, 1982). Furthermore,

some sediments may constitute a sink whereby inorganic nitrogen is influxed, likely via

denitrification (Smith et al., 1978). The general importance of denitrification in near-shore

sediments was later demonstrated by Seitzinger et al. (1980). Although benthic nutrient

regeneration could support ecosystem productivity, it was also realised that sediments may limit

overall productivity by acting as a sink for combined nitrogen (Christensen et al., 1987). Whilst the

importance of the benthic-pelagic link in the nitrogen cycle of shallow aquatic ecosystems had been

established, such large variation in strength and direction of the link implied the participation of

other factors regulating the supply of benthic-regenerated nutrients and/or usage of those nutrients

by primary producers.

The rate of organic material (algal, plant, bacterial and animal matter) deposition to the sediments,

as a substrate for biogenic diagenetic reactions, is possibly the primary regulator of benthic nutrient

52

Page 67: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

cycling that controls nutrient regeneration and denitrification (Hargrave, 1973; Berner, 1976; Suess,

1980). Biological, physical and chemical processes (e.g. trophic relationships, water movements

and abiogenic reactions, respectively) occurring in the overlying water column profoundly influence

the supply of organic matter to the sediment surface and also the resupply of nutrients to the pelagic

environment (Berner, 1976). These pelagic processes could partially explain geographic and

temporal variation in use of sediment-derived nutrients by phytoplankton and other primary

producers. However, for a given organic matter load, it is the rate of diagenetic reactions (including

biogenic organic matter decomposition, nutrient regeneration and denitrification) and exchange of

pertinent substrates and products at the sediment-water interface that markedly influence water

column concentration of nutrients (Berner, 1976; McCaffrey et al., 1980; Hammond et al., 1985).

Factors affecting reaction rates within, and transfer of substances across, the benthic boundary layer

probably account for spatio-temporal variability in sedimentary nutrient fluxes.

Biogenic diagenetic reaction rates are governed by well-defined kinetic parameters (e.g. substrate

concentration, temperature, salinity, redox, pH and pressure) (Hartwig, 1976; Zeizschel, 1980;

Hammond et al., 1985). In addition, the process of assimilation by benthic primary producers

(microphytobenthos, macroalgae and seagrasses) intercepts nutrients at the sediment-water interface

(e.g. Jansson, 1980). Processes important in the transport of solutes and solids across the sediment-

water interface include diffusion (resulting from concentration differences) and advection (the mass

transfer of constituent substances due to the flow of water and/or sediments) (Berner, 1980).

Advective processes define mechanisms of non-diffusive exchange and include: (i) turbulent

resuspension (due to tidal and wave action); (ii) depositional burial of sediments; (iii) groundwater

forcing; (iv) density-displacement of pore-water (e.g. flushing by bottom water of higher salinity);

(v) transport via bubble tubes; and importantly, (vi) bioturbation, the transportation of particles

(sediment reworking) and fluid (irrigation) by the activity of benthic organisms (Rhoads, 1974;

Berner, 1976, 1980; Hartwig, 1976; Aller, 1980a; McCaffrey et al., 1980; Zeitzschel, 1980,

Callender and Hammond, 1982; Hammond et al., 1985). The flux of nutrients across the sediment-

water interface was considered a net result of their consumption or production within the benthos,

as governed by pertinent diagenetic reactions and transport processes.

Factors controlling sedimentary reaction rates and transport processes were acknowledged to play a

role in bentho-pelagic nutrient cycling and were therefore duly investigated. However, it was the

process of bioturbation (Rhoads, 1974) that began to receive a large proportion of research attention

and soon challenged conventional thought regarding benthic biogeochemistry. The classical

sediment geochemistry paradigm characterised the average sedimentary deposit as dominated by

one-dimensional, vertical transport processes with stratified zones of diagenetic reactions (Berner,

53

Page 68: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

1976, 1980; Aller and Yingst, 1978). Given the vertical pore-water concentration profile and the

diffusion coefficient of a constituent solute (e.g. nutrients), Fick’s First Law could be used to

determine the diffusive flux of that solute across the sediment-water interface (Berner, 1976).

Discrepancies between diffusive nutrient fluxes predicted via pore-water profiling and directly

measured fluxes were demonstrated and the enhanced flux rates explained by advective processes

(Rhoads, 1974; Goldhaber et al., 1977; Aller, 1978; McCaffrey et al., 1980; Callender and

Hammond, 1982). Whilst the importance of bioturbation was acknowledged, the one-dimensional

formalism was retained by integrating the process (and other advective processes) into an ‘apparent’

or ‘effective’ diffusion coefficient that was ascertained empirically (Goldberg and Koide, 1962;

Berner, 1976, 1980; Aller, 1982). However, the diffusion analogy “is based on the assumption that

the activity of sediment-dwelling and irrigating organisms causes random movement of particles or

solutes, resulting in a transport of particles or pore-water along a concentration gradient in a similar

manner as molecular diffusion” (Timmermann et al., 2002: p. 152). The treatment of a sediment

deposit as an averaged or spatially-integrated entity only works well “where the length scale for

biological transport of material is less than the scale of the solute profile or where the mixing is

approximately stochastic” (Timmermann et al., 2002: p. 152).

It was the insightful research of Robert Aller and his co-workers during the late 1970s and early

1980s, from which understanding of the integral controls that benthic animals impart on diagenetic

processes and nutrient transport at the sediment-water interface, progressed beyond the view of the

sedimentary environment as one dominated by one-dimensional processes. Large- and fine-scale

patchiness in the magnitude and direction of sediment-water nutrient fluxes had previously been

attributed to differences in benthic faunal composition, abundance and biomass (e.g. Smith et al.,

1978), but the underlying mechanisms were still poorly known. In many instances, the assumptions

of utilising the diffusion analogy in considering bioturbative transport were breached. For example,

in the presence of irrigated burrows, exchange of solutes and solids between nonadjacent points in

the sediment cannot be modelled as a diffusive transport because transport distances are too great

and directional (i.e. non-random) (Timmerman et al., 2002). The radial diffusion model was

developed (Aller and Yingst, 1978; Aller, 1978, 1980a-c, 1982) to simulate these situations

whereby water and constituent solutes are transported over large vertical distances into burrows,

during and after which solutes are locally and radially exchanged with the surrounding pore-water

via diffusion6 (Timmerman et al., 2002). The benefit of this model was not only an improved

practical consideration of the presence of irrigated burrows on sediment-water nutrient fluxes; but

6 The related and practically simpler nonlocal irrigation model (Emerson et al. 1984; Bordreau, 1984) was developed later for practical use in similar situations (accounting for the presence of irrigated burrows), but reverted to a one-dimensional plane of reference.

54

Page 69: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

much more importantly by invoking a radial plane of diffusion and an allowance for lateral

heterogeneity, conceptualisation of the typical sediment environment had been transformed from a

one-dimensional diagenetic stratification into a dynamic, three-dimensional mosaic of diagentic

stratifications and biogenic microenvironments (Aller, 1982, 1988; Kristensen, 2000).

The consideration of irrigated burrows introduced a major advance in elucidating the faunal effects

on sedimentary biogeochemistry. Subsequently, other faunal affects on diagenetic processes (e.g.

the influence of epibenthic filter-feeders upon algal biodeposition and nutrient regeneration; Dame

et al., 1980; Cloern, 1982; Officer et al., 1982) were considered in a more holistic framework.

Aller’s radial diffusion model marked the beginning of a realisation that the pervasive presence,

diverse habits and complex interactions of benthic fauna were integral components of aquatic

ecosystems with much explanatory power in the consideration of sediment-water nutrient fluxes and

regulation of primary productivity. The increasing frequency of global eutrophication has ensured

the continuing relevance and scientific interest in benthic faunal effects on bentho-pelagic nutrient

cycling. Indeed, in the early 1980s, Aller (1982: p. 94) stated, prophetically: “The interactions

between different groups of sediment-dwelling organisms as determined by, or reflected in,

sediment chemistry remains one of the most intriguing and understudied aspects of marine

biogeochemistry”. Consequently, my study of the role of two polychaete species (as a case study)

in the Swan River Estuary, Western Australia contributes to understanding of the effects and

interactions of sediment dwelling organisms on sedimentary nitrogen biogeochemistry.

55

Page 70: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,
Page 71: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

PART II– THE STUDY

57

Page 72: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,
Page 73: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

CHAPTER 4

LIFE HISTORY, POPULATION STRUCTURE AND PRODUCTION

OF AUSTRALONEREIS EHLERSI AND SIMPLISETIA AEQUISETIS IN

THE SWAN RIVER ESTUARY, WESTERN AUSTRALIA

4.1 INTRODUCTION

Polychaete worms typically constitute an important component of the benthic fauna of marine

and estuarine ecosystems. In reviewing their functional significance in sedimentary

communities, Knox (1977) noted that polychaetes are one of the dominant elements of the

macrobenthic fauna, with species contributing, on average, >40% to both community richness

and abundance. In southern Australia, polychaetes often comprise about half of the species

richness of all macrobenthic invertebrates (Wilson et al., 2003). Further, polychaetes typically

contribute substantially to overall community biomass, productivity and respiration (Knox,

1977). Secondary production of polychaetes often is of pivotal ecological importance in the

food-web dynamics and trophic inter-relationships including other macroinvertebrates, fish,

birds and humans of estuaries and nearshore marine environments (Dorsey, 1981; Kent and

Day, 1983; Petti et al., 1996; Omena and Amaral, 2000; Morris and Keough, 2003). Species

such as Nereis virens and Perinereis aibuhitensis are cultured or harvested in many places

around the world and represent considerable commercial value as fish bait (Dorsey, 1981; Choi

and Lee, 1997; Wilson, 2000; Kristensen, pers. comm.). Knox (1977) highlighted the fact that

benthic polychaetes may exert substantial influence on their biophysical surroundings, including

having synergistic effects on sediment reworking, particle size distribution, stability and

community composition. The role of polychaetes and other benthic fauna in sedimentary

biogeochemical processes (including nutrient cycling) has gained increasing recognition over

the past thirty years (Section 3.3). To ascertain fully the role of benthic polychaetes in these

ecosystem functions/processes, a detailed knowledge of the basic biology of constituent species

is required – yet is often lacking.

Polychaetes represent a significant faunal component of the macrobenthos of the Swan River

Estuary, Western Australia. Within those habitats studied most intensively, namely in shallow

water <2 m deep, members of the families Nereididae (with two species recorded,

Australonereis ehlersi and Simplisetia aequisetis), Spionidiae, Capitellidae, Sabellidae and

Orbiniidae together comprise about half of the total macrobenthic abundance of the main basin

and upper estuary (Rose, 1994; Kanandjembo et al., 2001). The distribution of A. ehlersi is

restricted to the middle and lower estuary (Chalmer et al., 1976; Rose, 1994). Simplisetia

aequisetis occurring in the lower, middle and upper regions of the Estuary, typically is one of

59

Page 74: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

the most abundant elements of the macrobenthic fauna (Chalmer et al., 1976; Rose, 1994;

Kanandjembo, 1998; Kanandjembo et al., 2001; Pennifold and Davis, 2001). Burrowing adult

specimens may enhance significantly denitrification within sediments of the Estuary (De Roach

et al., 2002). I included both species of nereidid in this study; their distribution and abundance

in the Estuary and elsewhere are described in Chapter 2 (Section 2.2).

The most detailed account of the life history and reproductive biology of A. ehlersi and

S. aequisetis relate to populations near a major treated-sewage outfall in Port Phillip Bay,

Victoria (Dorsey, 1981). The only other life history study of an Australian nereidid, noted by

Wilson (2000), is by Glasby (1984, 1986) of the freshwater species S. limnetica from the

Hawkesbury River (New South Wales). Both A. ehlersi and S. aequisetis display rapid growth

and complete their life cycles in about 1-1.5 years, reproducing from spring to autumn (Dorsey,

1981). Australonereis ehlersi is tubiculous but likely produces a free-swimming larvae;

however, details of post-larval development are unknown (Dorsey, 1981). Simplisetia

aequisetis is dioecious, and the adult male broods embryos in specially constructed burrows or

tubes (Chapter 2, Section 2.2).

In this Chapter I describe the life history, geographical population structure and production of

the polychaetes A. ehlersi and S. aequisetis in the Swan River Estuary. The dynamics and

reproductive biology of both species from the estuary are compared in the Discussion with

Dorsey’s (1981) results. My investigation into the reproductive biology of both nereidid species

was intended to resolve not only their inherent ecological importance, but also to facilitate

insight and quantification of their respective roles in the biogeochemical cycling of nitrogen

within the estuary that constitute the substance of Chapters 5, 6 and 7.

4.2 MATERIALS AND METHODS

4.2.1 Study Area

The spatial distribution of A. ehlersi and S. aequisetis during summer (February) and winter

(August) conditions within the estuary was investigated by sampling at 12 shallow water

locations during 2000 (Figure 4.1; Table 4.1). Recruitment and progression of nereidid cohorts

were ascertained following monthly sampling from December 1999 to March 2001 at the Como

sampling site in the middle estuary. Temperature and salinity were determined for bottom water

(~2 m depth) on all sampling occasions via a conventional laboratory mercury thermometer and

AO temperature-compensated salinity refractometer.

60

Page 75: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Applecross

East Fremantle

South Perth

Heirisson Island

Bassendean

Ascot

Attadale

The Spit

Point Walter

Pelican Point

Matilda Bay

Como

N

5 km

Figure 4.1: Study sites within the Swan River Estuary.

61

Page 76: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Table 4.1: Swan River Estuary sampling sites and bottom water salinities and temperatures, 2000. Salinity throughout the estuary in February, 2000, was atypical, depressed due to freshwater runoff from a large, unseasonal rain event (further see Section 2.1.3). Salinity concentrations in February 2001 (Water and Rivers Commission, 2001), representing typical summer-time conditions, are given in parentheses.

Salinity

(g kg-1)

Temperature

(°C) Sampling Site

Location Substratum

Distance

from mouth

of estuary

(km) Feb

2000*

Aug

2000

Feb

2000

Aug

2000

East Fremantle Clayey/sand bank 4 32

(36)

36 26 17

Point Walter Sand/shell bed with lime-stone

rocks, macroalgal cover

8 11

(36)

28 28.5 19

‘The Spit’ Sand bank

9 10

(36)

28 28 16.5

Attadale Sand with sparse seagrass

(Halophila sp.) cover

12 9.5

(36)

16 28.5 19

Applecross Sand with sparse Halophila

cover

14 8.5

(35)

17 28 16

Como Sand/silt/shell bed

15 6

(35)

8 28 15

Pelican Point Sand/shell with patchy

Halophila cover

15 8

(35)

5 28 14.5

Matilda Bay Sand/shell bed

16 6

(35)

6 29 17

South Perth Mud/sand/shell bed

20 5

(34)

8 28.5 17

Heirisson

Island

Muddy sand 22 5

(31)

4 29 15

Ascot Sand bank

32 6

(27)

5 28.5 14.5

Bassendean Muddy sand

37 5.5

(23)

2 28 14

4.2.2 Sampling Procedure

Sediment was sampled by snorkellers using PVC pipe to extract a 25 cm long core, cross-

sectional area of 60 cm2. On every sampling occasion at each site, cores were taken within a

20 m wide transect extending from the shoreline (at low tide) to the 2 m deep water mark.

Within this transect, 12 cores were taken in total; with 4 cores taken randomly in water of three

depths: (i) <0.7 m deep, (ii) between 0.7 and 1.3 m deep, and (iii) between 1.3 m and 2 m deep

62

Page 77: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

(Figure 4.2). Each core was considered the primary sampling unit, such that between site

comparisons of parameters were based on the mean of 12 cores. The contents of cores were

transferred carefully to plastic jars and returned to the laboratory. Sediment from each jar was

passed through a series of 4 mm, 2 mm then 1 mm sieves and the polychaetes were preserved in

4% formalin in ambient estuary water. For benthic communities underlying waters <1.5 m deep

in the middle Swan River Estuary, Rose (1994) recorded that a 1 mm sieve captured 95.5% of

abundance and 99% of all species in sediment cores. Further, a total sampling area of 120 cm2

(5 cores) captured most species at each site, with 98% of both abundance and biomass in the top

9 cm of sediment (Rose, 1994). Consequently, the present study’s total sampling area of 240

cm2 (4 cores) to a depth of 25 cm, in each of the three sampling regions of each transect (Figure

3.2), was considered adequate to determine the abundance and biomass of nereidids at each site;

however, statistical comparison of abundance or biomass between each site was not undertaken

due to a lack of statistical power. The sampling effort required to generate significant statistical

power was not viable due to constraints imposed by the breadth of the sampling regime (12 sites

across the estuary) and the time taken to process samples (by one researcher). Regardless, the

sampling effort was considered sufficient to provide meaningful qualitative descriptions of

spatio-temporal variation in A. ehlersi and S. aequisetis distributions across the estuary. In the

laboratory, the two nereidid species were separated but no distinction was made between the

sexes.

0.7 m @ low tide1.3 m

2.0 m

Shore

20 m

25 cm

4mm mesh

2mm mesh

1mm mesh

SievesCollect

polychaetes

Determine length

Figure 4.2: Experimental design and polychaete sampling procedure at each study site.

63

Page 78: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

4.2.3 Size determinations

Following Kristensen (1984b), total body length of nereidids fixed in formalin (formalin-length,

Lf) was determined under a dissecting microscope to the nearest 0.5 mm, and used as an

estimate of size. The relationships between the preserved formalin-length and the extant length

of MgCl2-narcotised subsamples (live-length, Ll) of each species were as follows:

A. ehlersi: Ll = 1.0502 x Lf + 0.3421 (n = 10; r2 = 0.99)

S. aequisetis: Ll = 1.0435 x Lf - 0.0876 (n = 14; r2 = 0.99)

The original length of nereidids that were fragmented during the sampling/sieving process was

estimated by the following relationships between formalin-length and width of the third

setigerous segment (W) in mm (Figure 4.3):

A. ehlersi: Lf = 19.8897 x W - 0.7571 (n = 105; r2 = 0.91)

S. aequisetis: Lf = 27.6285 x W - 1.7654 (n = 157; r2 = 0.86)

0

10

20

30

40

50

60

70

80

0.00 0.50 1.00 1.50 2.00 2.50 3.00Width of 3rd segment (mm)

Poly

chet

e le

ngth

(mm

)

C. aequisetis A. ehlersi Linear (C. aequisetis) Linear (A. ehlersi)

S. aequisetis A. ehlersi

Figure 4.3: Relationship between width of the 3rd setigerous segment and nereidid length for preserved (4% formalin) specimens of Australonereis ehlersi and Simplisetia aequisetis. (See text for linear equations, sample sizes and r2 values.)

Size cohorts (age classes) were identified by fitting normal (Gaussian) components to monthly

length-frequency-distribution histograms using the modal progression analysis routine of the

computer program FiSAT II (FAO-ICLARM, 2005). Preliminary assessment indicated that

length modes were reasonably apparent for the A. ehlersi population at Como, but size cohorts

64

Page 79: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

of S. aequisetis were not clearly discernible by eye (Figure 4.5). Within FiSAT II, where a fit-

by-eye was insufficient to obtain initial estimates for mode means, Bhattacharya’s (1967)

graphical method was applied initially to the data. First guess mode means were then refined

using the NORMSEP optimisation procedure (after Hasselblad, 1966). The latter method

applies the maximum likelihood concept to SEParation of the NORMally distributed

components, and provides measures of standard deviation for the estimated modes. All mode

estimates were accepted since, in every case, the separation index between mean modes was

above the critical value of 2 (Gayanilo et al., 2002). The desirability of using an optimisation

refinement procedure has been highlighted by Grant (1989), who recognised that: (i) graphical

methods alone are often unreliable unless modes are well separated and/or sample sizes are very

large; and (ii) the methods used should place confidence limits around estimates of modes (e.g.

standard errors or deviations) to warn when estimates are not reliable. In this study, measures

of standard deviation are given on the growth curve of each cohort identified (Figure 4.6[A]).

The estimated density of nereidids within each cohort every month is also given (Figure 4.6[B]).

Finally, incidental observations of the presence of gravid or brooding adults, in addition to the

dimensions of eggs or embryonic juveniles, were also noted.

4.2.4 Biomass and Production

Several narcotised individuals of various sizes from both species were measured for live-length,

then dried for 48 hours at 60°C and weighed. The relationships between dry weight (Mdw in

mg) and length (in mm) were determined as:

A. ehlersi: Mdw = 0.0085 x Ll 1.96 (n = 12; r2 = 0.92)

S. aequisetis: Mdw = 0.0100 x Ll 1.65 (n = 13; r2 = 0.94)

Utilising these regressions for each identified cohort, monthly mean lengths were converted to

monthly mean biomass. For rough comparison of Mdw with studies stating biomass in terms of

wet weight (Mww), ash-free dry weight (Mafdw) or carbon weight (Mc), calculations may be based

on the assumptions that:

Mdw = [0.130 to 0.193] x Mww

Mdw = [1.125 to 1.330] x Mafdw

Mdw = [2.466] x Mc (Cammen, 1980; Kristensen, 1984b; Sardá et al., 2000).

Monthly production of both nereidid species was estimated by using the monthly mean density

and biomass increments of each cohort as described by Crisp (1971). Where months were

sampled in more than one year (December to March), a monthly average was obtained. Annual

production (P) was determined by summing the average monthly production over 12 months. In

the same manner, mean annual biomass (B) was calculated by summing average monthly

biomass of all cohorts, and used to determine the P/B ratio.

65

Page 80: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

4.3 RESULTS

4.3.1 Salinity, Temperature and an Unexpected Algal Bloom

The temporal variation in bottom water temperature and salinity over the 16 month

sampling period at Como is illustrated in Figure 4.4. Temperature ranged from 15 °C in

winter (August 2000) to 32 °C in summer (January 2001). Excluding the datum from

February 2000 (see below), salinity ranged from 8 g kg-1 in winter (August/September

2000) to 38 g kg-1 in summer (February 2001).

Variation in bottom water temperature and salinity throughout the 12 sampling sites in the

estuary, for both summer (February 2000) and winter (August 2000), is presented in Table

4.1. Bottom water temperature was relatively stable throughout the estuary in both summer

(26-29 °C) and winter (14-19 °C). For both seasons, there was generally an upstream to

downstream cline of increasing salinity. In winter, salinity ranged from 2 g kg-1 at

Bassendean to 36 g kg-1 at East Fremantle. In summer, salinity of the surface 2 m of water

throughout most of the estuary (Bassendean to Point Walter) was a brackish 5 to 11 g kg-1,

with salinity increasing to 32 g kg-1 nearer the mouth (East Fremantle). During the

following summer (February 2001), bottom water salinity ranged from 23 g kg-1 at

Bassendean to 36 g kg-1 at East Fremantle (Table 4.1).

The dip in salinity at Como during summer (Figure 4.4) and generally throughout the

estuary in February 2000 (Table 4.1), was due to freshwater runoff from unseasonal heavy

rain associated with the aftermath of a tropical cyclone in January (see Appendix A). The

estuary was flushed with enough freshwater to fill it five times over and the rainfall event

and resultant hydrological conditions were considered highly atypical (Atkins et al., 2001).

Additionally, the summer freshwater conditions concomitant with a large, runoff-associated

nutrient input triggered a record estuary-wide bloom of the toxic blue-green alga

Microcystis aeruginosa (see Section 2.1.3). Subsequent easing of freshwater discharge

enabled the tidal re-entrance of marine water, upstream propagation of the salt wedge and

collapse of the blue-green algal bloom, such that the system had reverted to its more typical

marine dominated state by March/April 2000 (Figure 4.4; Atkins et al., 2001). The direct

and indirect implications of both the unusual salinity regime and faunal mortality effects of

the widespread toxic algal bloom during the early stages of the study period are reviewed in

the following discussion of A. ehlersi and S. aequisetis population dynamics.

66

Page 81: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

0

5

10

15

20

25

30

35

40

1999

D

2000

J F M A M J J A S O N D

2001

J F M

Salin

ity (g

kg

-1)

0

5

10

15

20

25

30

35

Tem

pera

ture

(°C

)

SalinityTemperature

Month

Year

*

Figure 4.4: Salinity and temperature of bottom water (depth ~ 2 m) at Como during the study period. Salinity atypically was depressed in February/March 2000 due to freshwater runoff from a large, unseasonal rain event (see text).

4.3.2 Life History – Age Distribution and Growth at Como

Australonereis ehlersi

The monthly size-frequency distributions of A. ehlersi at Como are given in Figure 4.5(i). The

mean body length and density of the cohorts identified are respectively given in Figure 4.6 [A(i)

and B(i)]. Six cohorts were identified over the 16-month study period, including: Cohort 1

(represented at the beginning of the study by medium to large-sized mature individuals; Cohorts

2 and 3 (two cohorts initially recruited as small-sized juveniles); and three cohorts M1, M2 and

M3 that immigrated as medium-sized mature individuals.

The mean body length of individuals from Cohort 1 increased from 45.8 mm (in December

1999) to 65.0 mm (in February 2000). The density of the cohort declined from 84 to 14

nereidids m-2 during the same period. The cohort was not present during March 2000.

Cohort M1 migrated to Como during autumn (March and April 2000) at a density of 84-98

nereidids m-2, with a medium-sized mean body length of 40.0-49.3 mm. The cohort was not

captured during May 2000 (likely due to lowered sampling effectiveness at lower nereidid

densities) and was last collected at the beginning of winter (June 2000) at the low density of 28

nereidids m-2 and a mean body length of 72.5 mm.

67

Page 82: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Cohort M2 migrated to Como during late autumn/early winter (May and June 2000) also as

medium-sized individuals (mean body length 45.0-47.0 mm) at a density of 28-70 nereidids m-2.

The density of the cohort declined over winter to be 14 nereidids m-2 in September 2000, by

which time mean body length had increased to 75 mm. During this month one large, bloated

A. ehlersi individual (73.1 mm live length) from Cohort M2, ruptured between the dorsal and

ventral parapodia when probed with forceps; the coelomic fluid contained numerous eggs, with

a mean diameter of about 310 µm.

A third migration event of medium-sized A. ehlersi in mid- to late-winter (July and August

2000) of Cohort M3 yielded a density of 42-154 nereidids m-2, with a mean body length of 43.3-

48.7 mm. The cohort was last present during September 2000 at a density of 28 nereidids m-2

and mean length of 60 mm.

The appearance of Cohort 2 in late-autumn (May 2000) marked the first recruitment event of

small-sized juvenile A. ehlersi individuals of mean body length of 17.2 mm and cohort density

of 378 nereidids m-2. A continual increase in density occurred over the winter months, peaking

at 1836 m-2 in August 2000 by which time mean body length of Cohort 2 had increased to

24.6 mm. This winter peak in density of Cohort 2 was an order of magnitude higher than the

maximum density attained by any other cohort. The density of the cohort then decreased

steadily throughout the rest of the year to 28 nereidids m-2 by December, 2000, and was absent

during January 2001. The cause of the reduction in density, whether due to mortality or

emigration, was not determined. The mean body length of Cohort 2 increased to a medium-

sized 40.0 mm over the period to December 2000.

A second recruitment event of small-sized A. ehlersi individuals occurred during the summer of

2000/01, with Cohort 3 first appearing during December at a mean body length of 17.5 mm and

a density of 224 nereidids m-2. The density of the cohort (238 nereidids m-2) remained relatively

stable through to March 2001. Over this time, the mean body length of Cohort 3 increased to

33.5 mm.

Simplisetia aequisetis

The monthly size-frequency distributions of S. aequisetis at Como are shown in Figure 4.5(ii).

The mean body length and density of the cohorts identified are given in Figure 4.6 [A(ii) and

B(ii)]. Ten cohorts were identified over the 16-month study period.

Cohorts 1, 2 and 3 were present initially in December 1999. Cohort 1 comprised larger

individuals of mean body length 38.4 mm, at a density of 322 nereidids m-2. The density of

Cohort 1 decreased to 14 nereidids m-2 by the end of summer (March 2000), whilst mean body

68

Page 83: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

length increased to 50.0 mm. Although not detected in April/May and from July to September,

a few individuals from Cohort 1 likely remained present in low density (~14 nereidids m-2) until

June if not October. Large individuals with mean body lengths of 60.0 mm and 65.0 mm,

respectively, were captured in those latter months.

Cohort 2, present in a relatively high density of 687 nereidids m-2 during December 1999,

comprised individuals with a mean body length of 22.9 mm. The density of this cohort declined

to 14–154 nereidids m-2 during autumn/early-winter. Cohort 2 was last detected during June

2000, with a mean body length of 41.0 mm.

Initially detected in December 1999 at a low density of 56 nereidids m-2 and mean body length

of 12.7 mm, Cohort 3 exhibited a continual increase in density throughout summer to reach a

peak of 617 nereidids m-2 in March, 2000. By this time mean body length of the cohort had

increased to 24.9 mm. The density of Cohort 3 then decreased over autumn to 196 nereidids m-2

in May 2000. This was the last month in which the cohort was detected, having grown to a

mean body length of 30.1 mm.

Cohort 4 was first detected in mid-autumn (April 2000) at a density of 308 nereidids m-2, with a

mean body length of 20.1 mm. Further recruitment to the cohort was exhibited in late-autumn

(May 2000), with density doubling to 617 nereidids m-2. Over winter, cohort density declined

only slightly to 547 nereidids m-2 in September 2000. At this time mean body length of Cohort

4 had increased to 37.7 mm. The density of the cohort then decreased considerably during

spring to 14 nereidids m-2 in November 2000. At a mean body length of 55.0 mm, November

2000 was the last month in which Cohort 4 was noted.

Cohort 5 was initially detected in low densities (56-98 nereidids m-2) and at a small mean size

(11.8-15 mm) during late-autumn to early-winter (April - June, 2000). A large increase in

cohort density to 575 nereidids m-2 was then exhibited during mid-winter (July 2000), with

individuals having a mean body length of 19.6 mm. The density of Cohort 5 decreased

gradually over winter/early-spring to 168–364 nereidids m-2, but then considerably to only 14

nereidids m-2 in the last month of spring. The cohort was last recorded during November 2000

when the mean body length was 50.0 mm.

Cohorts 6, 7 and 8 were first detected in September, October and November, respectively. At

first detection, each cohort demonstrated similar average body length of ~20 mm and density

within the range 750–890 nereidids m-2. These cohorts had a longevity of only 3 - 4 months,

during which time the size of Cohorts 6, 7 and 8 attained an ultimate mean length of 43.4, 34.8

69

Page 84: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

and 26.5 mm, respectively; whilst the density of each cohort declined to 56, 168 and 308

nereidids m-2, respectively.

Whilst Cohort 9 was first detected at a density of 70 nereidids m-2 and mean body length of

6.7 mm during late-spring (November 2000), it then exhibited a similar progression as the three

preceding cohorts. Mean body length of Cohort 9 in December 2000 was 18.7 mm, with a

cohort density of 757 nereidids m-2. During the next 4 months over summer, mean body length

of the cohort increased to 30.4 mm, whilst the density decreased to 98 nereidids m-2.

Finally, Cohort 10 was detected in early-summer (December 2000) at a density of 266 nereidids

m-2 and mean body length of 10.1 mm. Cohort density and mean length increased over summer

to be 589 nereidids m-2 and 16.4 mm during February 2001. By the last month of the study

period (March 2001), mean body length of Cohort 5 had increased to 20.5 mm, whilst density

decreased to 224 nereidids m-2.

Brooding individuals of the species S. aequisetis, ranging in live length between 31.2-53.1 mm,

were observed in cohorts 2 (May 2000), 4 (November 2000), and 7 (November and December

2000, January 2001. The individuals occasionally were still encapsulated in mucous-bound

tubes despite the sieving process. Tubes contained hundreds of 5 to 8 setiger embryonic

juveniles with body lengths of 500 to 1000 µm; the mean diameter of the eggs was about

225 µm (Figure 4.7).

70

Page 85: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

5 10 15 20 25 30 35 40 45 50 55 60 65 70 75

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

0100200300400

5 10 15 20 25 30 35 40 45 50 55 60 65 70

(i) Australonereis ehlersi (ii) Simplisetia aequisetis

Body Length (mm)

December 1999

January 2000

February

March

April

May

5

6

7

8

9

10

4

3 2

1

2 M2

M1

1

July

August

September

October

November

December 2000

January 2001

February

March

Den

sity

(ner

eidi

ds m

-2)

631 631

M3

3

June

71

Figure 4.5: Monthly size-frequency distributions of (i) A. ehlersi and (ii) S. aequisetis at Como from December 1999 to March 2001. Shading delineates cohorts and bold numbers indicate cohort appearance. Two-tone shading of S. aequisetis body length intervals indicates a mixed contribution by sequential cohorts. Dashed line indicates division between sub-adult and potentially reproductive S. aequisetis (see text for explanation).

Page 86: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Figure 4.6: [A] Mean body length (± s.d.) and [B] density of (i) A. ehlersi and (ii) S. aequisetis cohorts at Como from December 1999 to March 2001. [C] Total monthly biomass and [D] production of combined cohorts are also depicted. Dashed line at ~ 31 mm on [A(ii)] indicates division between sub-adult and potentially reproductive S. aequisetis (see text for explanation).

11394

4992

1289 1019426

987

1734 15011055

414 1128

307 564577

2658

728

851

2579 5064

8310

6148

5935

2478

328

521 931 1040

1979

426

200

428541

701

305

65

244119

322

345

796

498

956

864

1040

226

354457

251

546

202

641

114

78

118181 166 151

393234

352

37

518

89

24

242188

396

240

499

809

455

1264

348

471

35

424

80 239

26

45

158

182

2152 10 10 11 4 4 4 6128 17 6

80

78

1

M1

M2

M3

2

3

1327

498236 110 120 125 129 133 275

1201

747

474317

706

240778

1244 1136 541

436 1276 1635

1375

1975 2178

473

778

902 475

1357

1009

979

282

1244

2138

1891

587

1129

1052

688

953

492

596

172211 274

150

328

120

1

2

3

4

5

6

7

8

9

10

1 2 3 4 5 6 7 8 9 10

Cohort No

[A] [B] [C]

[D]

(i) Australonereis ehlersi (ii) Simplisetia aequisetis M

ean

Bod

y L

engt

h (m

m)

Den

sity

(ner

eidi

ds m

-2)

Bio

mas

s (m

g dw

m-2

) Pr

oduc

tion

(mg

dw m

-2 m

onth

-1)

80

70

60

50

40

30

20

10

0

2000

1800

1600

1400

1200

1000

800

600

400

200

0

8000

7000

6000

5000

4000

3000

2000

1000

0

1400

1200

1000

800

600

400

200

0

Cohort No

Cohort No

Month D J F M A M J J A S O N D J F M D J F M A M J J A S O N D J F M

Year 1999 2000 2001 1999 2000 2001

Cohort No

1 M1 M2 M3 2 3

3874

72

Page 87: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

[A]

2 mm

[B]

2 mm

Figure 4.7: Brooding tube of S. aequisetis [A] with eggs [B].

73

Page 88: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

4.3.3 Biomass and Production at Como

Australonereis ehlersi

The pattern of progression in the monthly biomass (standing stock) of A. ehlersi at Como

generally reflected the changes in combined densities of the cohorts (Figure 4.6[Bi & Ci]). A

maximum monthly biomass of 11.4 g dw m-2 in winter (August 2000) resulted predominantly

from the very high density of small-sized recruits in Cohort 2, supplemented by the lower

density of large, immigrant Cohort M3 adults. Cohort 2 comprised the dominant component of

total monthly biomass throughout winter and spring (June to November, 2000), and was largely

responsible for the gradual increase and then decline in total biomass, either side of the temporal

peak. A minimum monthly biomass of 0.43 g dw m-2 in February 2000, comprised solely a low

density of large, Cohort 1 individuals. A minimum of 0.85 g dw m-2 was recorded during the

subsequent summer (December 2000) representing medium-sized, Cohort 2 adults and newly

recruited small, Cohort 3 individuals. The annual mean biomass (B) for A. ehlersi at Como was

3.89 g dw m-2.

Total monthly production of combined cohorts of A. ehlersi at Como is depicted in Figure

4.6[Di]. The temporal pattern of change in monthly production largely followed that of

monthly biomass. Maximum monthly production peaked at 4.99 g dw m-2 month-1 in

September 2000, principally due to growth by the high density of Cohort 2 recruits in the

preceding month but also supplemented by the growth of migrant Cohort M3 adults. Production

of A. ehlersi was absent in March 2000, due to the disappearance of Cohort 1. During the

following summer, production was lowest in the months of November 2000 (0.23 g dw m-2

month-1) and February 2001 (0.25 g dw m-2 month-1), respectively, indicating only low monthly

increases in biomass of Cohorts 2 and 3. Total annual production (P) for A. ehlersi at Como

was estimated at 12.5 g m-2 year-1, yielding an annual turnover (P/B) ratio of 3.2.

Simplisetia aequisetis

Total monthly biomass of combined cohorts of S. aequisetis at Como in 2000 is summarised in

Figure 4.6[Cii]. Highest monthly biomass was recorded during spring (September to

December), ranging between 3.76 - 4.32 g dw m-2. During this time, total monthly biomass

comprised various components from Cohorts 4 to 9. A minimum monthly biomass of

approximately 1.49 g dw m-2 was recorded during the summer months (January to March 2000),

with components progressing from Cohort 1 to 3. Thereafter, monthly biomass gradually

increased towards the spring peak. Monthly biomass then again declined over the 2000/01

summer to be at a minimum of 0.60 g dw m-2 during March 2001, comprising similar biomass

components from Cohorts 9 and 10. Annual mean biomass of S. aequisetis at Como was

estimated as 2.58 g dw m-2.

74

Page 89: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

For S. aequisetis at Como, a marked peak in monthly productivity of 2.22 g dw m-2 month-1

occurred during late-spring (November) 2000 (Figure 4.6[Dii]), due largely to growth

contributions from Cohort 7 recruits during the preceding month. Productivity was also high in

October 2000 = 1.57 g dw m-2 month-1 and December 2000 = 1.24 g dw m-2 month-1; with major

growth contributions from recruits in Cohorts 6 and 8, respectively. Aside from a relatively

high productivity of 0.92 g dw m-2 month-1 in January 2000 (and a large growth contribution

from Cohort 2 during the first month in the sampling period for which productivity could be

calculated), productivity was low (0.14 – 0.54 g dw m-2 month-1) during the remaining months

of the study period. Minimal values near the lower end of this range were observed during

February to April in 2000 (combined low productivity mainly of Cohorts 2 and 3) and in

January and March, 2001 (combined low productivity mainly of Cohorts 9 and 10). Total

annual production for S. aequisetis at Como was estimated to be 8.74 g dw m-2 year-1, providing

a P/B ratio of 3.4.

4.3.4 Population Structure within the Swan River Estuary

For both A. ehlersi and S. aequisetis populations of the Swan River Estuary from waters <2 m

deep, the total density and biomass at each of the 12 study sites sampled during both summer

(February 2000) and winter (August 2000) is presented in Figure 4.8. Where a species was

present at a site, the size-frequency distributions of individuals are given in Figure 4.9. Cohort

analysis was not applied to these size-frequency distributions since it is impossible to determine

congruency of cohorts both between different sites and also at the same site with 6 months

between samplings.

Australonereis ehlersi

During summer, A. ehlersi was present at only four sites in the middle estuary at low densities

(14 to 56 nereidids m-2; Figure 4.8[Ai]). The density at Attadale was at the upper end of this

range and comprised large individuals (40-70 mm live length; Figure 4.9[Ai]), which

represented a substantial biomass (1.90 g dw m-2; Figure 4.8[Bi]). Low biomass (0.25–0.45 g

dw m-2) of A. ehlersi during February 2000 occurred at Como, Pelican Point and Applecross; at

these latter two sites, however, the mean length of individuals (~50 mm and 25 mm,

respectively) were considerably smaller than at Como (~65 mm).

In contrast, during winter (August 2000) A. ehlersi was present at 7 sites, including Matilda Bay

in the middle estuary plus Point Walter and East Fremantle in the lower estuary (Figures 4.8[Aii

& Bii] and 4.9[Aii]). Excluding Como, the density of A. ehlersi at each site was low (14-168

nereidids m-2), representing a biomass at each site in the range 0.21-1.12 g dw m-2. The high

density (1991 nereidids m-2) and high biomass (11.0 g dw m-2) of A. ehlersi at Como during

August 2000 was anomalous. No other site yielded a comparable high density standing stock of

75

Page 90: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

small-sized recruits, however a low density cohort of small individuals 10-20 mm in length was

observed at East Fremantle. In summary, the biomass of A. ehlersi at sites other than Como

during winter resulted from a low density of individuals >35 mm in length.

Simplisetia aequisetis

Simplisetia aequisetis was present at all 12 sampling sites during both the summer and winter

(Figures 4.8 and 4.9B). During February, both density (1710 nereidids m-2) and biomass (4.47 g

dw m-2) were greatest at The Spit (Figure 4.8[Ai & Bi]), located at the most downstream end of

the middle estuary. At this site, the body lengths of S. aequisetis individuals were in the range

of 10-60 mm, with a polymodal size frequency distribution which likely represented several

cohorts (age-classes) (Figure 4.9[Bi]). The 20 mm body length category was the most common

size-interval of S. aequisetis individuals. Within the adjacent lower estuary, the density and

biomass of S. aequisetis decreased sharply and were lowest at East Fremantle, i.e. nearer the

mouth of the estuary (70 nereidids m-2, 0.12 g dw m-2). At these lower densities within the

lower estuary, the size-frequency distribution remained polymodal and the 20 mm body length

category was most common; however, S. aequisetis individuals >40 mm in length were not

present. Regarding the remaining sites in the middle and upper estuary during summer, there

was a general upstream trend of decreasing density and biomass of S. aequisetis although the

species was present in only very low density and biomass at both Ascot and South Perth during

this time. Hence, sites within the main basin of the middle estuary (Attadale, Applecross,

Como, Pelican Point and Matilda Bay) supported a higher density (435-1262 nereidids m-2) and

biomass (0.85-3.65 g dw m-2) of S. aequisetis than sites within Perth Water (South Perth and

Heirisson Island) and the upper estuary (Ascot and Bassendean); wherein density and biomass

ranged, respectively, from 14–252 nereidids m-2 and 0.04–0.58 g dw m-2. The size-

constituencies of S. aequisetis individuals at sites within the middle estuary were generally

polymodal and within the body length range of 10-50 mm. Simplisetia aequisetis >50 mm in

length during summer were observed only at the Spit and Attadale. The low density of S.

aequisetis in the upper estuary during summer comprised medium-sized individuals 25-50 mm

in length.

Although the magnitude of S. aequisetis density and biomass at sites within the lower reaches of

the estuary (from Como downstream to East Fremantle) during winter were generally similar to

that observed during summer (Figure 4.8), the species’ distribution pattern upstream of Como

was markedly different from that exhibited during summer. The density and biomass of S.

aequisetis at Pelican Point and Matilda Bay (within the upper main basin of the middle estuary)

was 2-5 times higher than that during summer. Further, the density and biomass of S. aequisetis

at sites within Perth Water and the upper estuary were 1-2 orders of magnitude higher than

summer-time values. Thus, whilst S. aequisetis remained at high density (1318 – 1514

76

Page 91: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

77

nereidids m-2) and biomass (4.73 – 5.44 g dw m-2) at the Spit and Attadale during winter (and at

low density and biomass within the lower estuary), the absolute peak in density and biomass

shifted upstream to Heirisson Island (2930 nereidids m-2) and Ascot (5.87 g dw m-2). In

summary, the estuary-wide population of S. aequisetis was greatly supplemented in winter by

the increases in density and biomass at locations upstream of Como (i.e. within the upper

estuary, Perth Water and at Matilda Bay and Pelican Point). The size-constituencies of

S. aequisetis individuals at these upstream sites were polymodal, with the mean body length of

likely cohorts (age-classes) encompassing the full range of sizes from 5-65 mm (Figure

4.9[Bii]). However, the most common/dense size-interval was always represented by the 15 or

20 mm body length category, suggesting that small-sized recruits of S. aequisetis contributed

largely to the heightened upstream density/biomass during winter.

Page 92: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Figure 4.8: Mean density [A] and biomass [B] of A ehlersi and S. aequisetis during (i) February and (ii) August 2000 at 12 sampling locations throughout the Swan River Estuary.

East

Fre

man

tle

Poin

t Wal

ter

‘The

Spi

t’

Atta

dale

App

lecr

oss

[A] Density [B] Biomass

Com

o

Pelic

an P

oint

Mat

ilda

Bay

Sout

h Pe

rth

Hei

rris

son

Is.

Asc

ot

Bas

send

ean

6000

5000

4000

3000

2000

1000

6000

5000

4000

3000

2000

1000

0

0

78

(i) February 2000 (ii) August 2000

Den

sity

(n

erei

dids

m-2

)

Bio

mas

s (m

g dw

m-2

)

2500

2000

1500

1000

500

0

Bas

send

ean

Asc

ot

Hei

rris

son

Is.

Sout

h Pe

rth

Mat

ilda

Bay

Pelic

an P

oint

Com

o

App

lecr

oss

Atta

dale

‘The

Spi

t’

Poin

t Wal

ter

East

Fre

man

tle

UPSTREAM DOWNSTREAM

Bio

mas

s (m

g dw

m-2

)

Australonereis ehlersi

Den

sity

(n

erei

dids

m-2

)

2500

2000

1500

1000

500

0

Simplisetia aequisetis

Page 93: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

(B) Simplisetia aequisetis

(i) February 2000 (ii) August 2000 UPSTREAM

Bassendean

(A) Australonereis ehlersi

(i) February 2000 (ii) August 2000

Ascot Heirisson Island South Perth

Matilda Bay Pelican Point Como Apple- cross Attadale

65

‘The Spit’ Point Walter East Fremantle 65

DOWNSTREAM

400

300

200

100

400

300

200

100

400

300

200

100

400

300

200

100

400

300

200

100

400

300

200

100

400

300

200

100

400

300

200

100

400

300

200

100

400

300

200

100

400

300

200

100

400

300

200

100

5 10

15

20

25

30

35

40

45

50

55

60

65

70

5

10

15

20

25

30

35

40

45

50

55

60

65

70

5 10

15

20

25

30

35

40

45

50

55

60

65

70

5

10

15

20

25

30

35

40

45

50

55

60

70

5

10

15

20

25

30

35

40

45

50

55

60

70

Den

sity

(ner

eidi

ds m

-2)

Body Length (mm)

Figure 4.9: Size-frequency distributions of A. ehlersi [A] and S. aequisetis [B] during (i) February

and (ii) August 2000 at 12 sampling locations throughout the Swan River Estuary.

645

449

631 631

547 505

785 855

79

Page 94: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

4.4 DISCUSSION

4.4.1 Secondary Production of A. ehlersi and S. aequisetis

Sardá et al. (2000) defined secondary production in terms of the formation of heterotrophic

biomass with time, and primarily is a function of the growth of individuals, recruitment patterns

and mortality. Variation in these parameters, affected by a multitude of physicochemical and

biological factors, for example food resource quality and quantity, extent of larval dispersion,

magnitude of predation and competition, typically will lead to differences in secondary

production between separate populations. By including reference to biomass or standing stock,

the ratio of production to biomass (P/B) expresses the turnover of a population and is more

informative than comparing the productivity of populations (Kristensen, 1984b).

As far as the author is aware, this is the first study to quantify the mean annual biomass and

production for any population of A. ehlersi. Whilst the S. aequisetis population co-inhabiting

the site at Como had both lower mean annual biomass and productivity, the resultant P/B ratio

of 3.4 was slightly higher than that of A. ehlersi (P/B = 3.2). These turnover rates in the Swan

River Estuary are slightly higher than that observed by Dorsey (1981) for S. aequisetis in Port

Phillip Bay (P/B = 2.9). However, the latter population had both a high mean annual biomass

and production, based on ash-free dry weight to dry weight conversions (Section 4.2.4). These

production estimates and P/B ratios suggest that Australian nereidids are likely an important

component of total productivity in the ecosystem; offering a potential food resource for

constituent crustaceans, fish and bird species.

Turnover rates of the nereidids Simplisetia keiskama (P/B = 1.8) and S. erythraeensis7 (P/B =

1.9) in the Berg River Estuary, South Africa (Kalejta, 1992), were lower than the rates reported

for Australian nereidids; whereas derivative mean annual production and biomass estimates

were generally intermediate to those exhibited by nereidids in the Swan River Estuary and Port

Philip Bay. Similar to South African species, the Brazilian nereidid Laeonereis acuta had a P/B

ratio of 2.0 and comparable annual productivity (Omena and Amaral, 2000). The range of

turnover rates for those few southern hemisphere nereidids P/B = 1.8-3.4 is within the range

0.8-4.8 of northern hemisphere nereidids including Nereis diversicolor, Neanthes

arenaceodentata, Neanthes succinea, Neanthes virens, Lumbrinereis acicularum and

Lumbrinereis fragilis (Omena and Amaral, 2000; Medernach et al., 2000).

7 The genus of Simplisetia is presumed to replace the genus of Ceratonereis for these two South African species (see p. 27).

80

Page 95: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

4.4.2 Australonereis ehlersi: Life History and Distribution

Life History

The first major arrival of juvenile A. ehlersi recruits (Cohort 2) occurred in late-autumn and

peaked during late-winter, a similar pattern of intensive winter-time recruitment of juveniles

recorded for Port Phillip Bay (Dorsey, 1981). Additional cohorts of A. ehlersi at Como first

appeared as medium-sized adults in low density during autumn through winter. It is not known

whether the medium-sized A. ehlersi individuals of Cohort 1 first migrated at a medium-size in

the summer of 1999/2000 or first appeared as small-sized juveniles earlier in 1999. Following

initial migration of the medium-sized A. ehlersi cohorts during autumn and winter, growth

ensued over the subsequent 2 to 4 months while cohort density concomitantly decreased. The

sudden appearance of these medium-sized groups at Como implies immigration from elsewhere,

indicative of movement within the estuary. Reproductive stages of A. ehlersi have been

observed swimming in the winter in the Swan River Estuary (McShane, 1977) and in spring and

summer in Port Phillip Bay, Victoria (Dorsey, 1981), and on both occasions were interpreted as

indicative of pelagic spawning events. The large gravid female collected from Como sediment

in September did not show morphological changes indicative of epitoky. However, large

atokous individuals about to spawn probably leave the sediment to release gametes in the

overlying water. The absence of gravid A. ehlersi specimens throughout the whole study period

supports the idea that reproductive individuals do not remain in the sediment. Thus, the decline

and eventual disappearance of cohorts M1-3 likely was due predominantly to pelagic spawning

by atokous individuals.

Pelagic spawning by epitokal stages is not uncommon in the Nereididae (Dorsey, 1981), but

atokous gravid females of A. ehlersi are more comparable to those of Nereis japonica (Dorsey,

1981). Whilst atoky may not be unusual for estuarine or brackish/freshwater nereidids, it is

usually characteristic of species which produce benthic larvae and/or brood, where limited

dispersal or confinement of larvae to an osmotically/thermally stable sediment environment,

may confer some form of reproductive advantage (see also p. 87). Hence, atokous pelagic

spawning may represent a modified reproductive strategy of estuarine nereidids facilitating

larval dispersion (Dorsey, 1981).

Prior to the likely pelagic spawning events of large individuals, Cohorts M1, M2 and M3

presumably migrated into the sediments at Como as medium-sized adults. Thus, swimming

behaviour may also facilitate migration of A. ehlersi. Indeed, the decline in this cohort’s density

over the preceding 2 months, and its ultimate disappearance, likely was due to emigration of

medium-sized adults in addition to mortality. Such emigration was not apparent in Dorsey’s

(1981) study of A. ehlersi in Port Phillip Bay, possibly because he pooled size-frequency data

between different sites.

81

Page 96: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

The reproductive strategy and pelagic spawning potential of A. ehlersi is unknown, however the

species may be iteroparous (polytelous). Glasby (1986) suggests that the presence of

coelomoducts in both males and females of A. ehlersi (Hartman, 1954; Hutchings and Reid,

1990), which enable external transport of gametes, indicates potential for iteroparity. Nereidids

generally lack coelomoducts and are semelparous (monotelic), often spawning via rupture of the

body wall preceding death of the individual (Glasby, 1984; Wilson, 2000). Further

investigation into the reproductive behaviour of A. ehlersi, combined with histological

examination of the gametogenic cycle, is required to elucidate the frequency of spawning events

in its life history.

The mean body length of the winter recruits increased by 23 mm over the ensuing 6 month

period until the end of spring. Assuming that these medium-sized adults then emigrate and

settle elsewhere in the estuary, the life-span of A. ehlersi is estimated at 8-10 months, plus the

undetermined duration of the larval period. The length of the larval period is estimated to last

for 2-4 months, by assuming an initial embryonic length of ~2 mm and extrapolating the growth

curves of Cohorts 2 and 3. Hence, the total life-span of A. ehlersi in the Swan River Estuary is

estimated at about 10-14 months, comparable with the estimated 12-16 month total life-span for

the Port Phillip Bay population (Dorsey, 1981).

Dorsey (1981) concluded that, in Port Philip Bay, A. ehlersi spawn mainly from summer to

early-winter. Given a 2-4 month pelagic larval period, the main recruitment period of A. ehlersi

juveniles settling at Como from late-autumn to mid-winter (Cohort 2), indicates a similarly

timed major spawning period in the Swan River Estuary. However, the timing of major

spawning events may be specific to a particular location; in addition, secondary spawning

events may occur year-round. For example, Dorsey (1981) noted the beginning of a low density

juvenile recruitment event during late-summer (at the end of his sampling period), indicative of

spring-time spawning. Similarly, the low-density recruitment of juveniles to Como during the

2000/01 summer of this study (Cohort 3) indicates a spawning event during late-winter to early

spring. Uncertainty regarding the extent of larval dispersion prior to settling makes

determination of the parent cohort problematic. If these summer-time recruits were locally-

derived (from the spawn of adults located at or near Como), then candidate parents include

large-sized adults from Cohort M3 and/or medium-sized adults from Cohort 2, potentially

lending support to the hypothesis of iteroparity, with spawning at medium- and then again at

large-size. Candidate local parents of the more abundant winter-time recruits at Como (Cohort

2) include those large-sized A. ehlersi present from summer to early-winter (2000), i.e.

individuals in Cohorts 1, M1 and M2. Alternately, settling juveniles may be derived from

adults residing elsewhere in the estuary.

82

Page 97: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Regardless, it is apparent that A. ehlersi populations are capable of spawning throughout the

year. Furthermore, spawning and recruitment patterns may demonstrate considerable inter-

annual variability. The juvenile recruits present in the summer of 2000/01 at Como were not

observed at the same time during the preceding year. Similarly, Dorsey (1981) reported the

presence of juvenile recruits in relatively high density in the March/April of 1977 but only in

very low density during the March/April of 1978. In relation to A. ehlersi’s spatio-temporal

distribution pattern within the Swan Rive Estuary, factors potentially governing the timing and

success of spawning and settling events are discussed below.

Distribution

Evidence from this study and others (Chalmer et al., 1976; McShane, 1977; Rose, 1994;

Kanandjembo, 1998) indicates that A. ehlersi resides exclusively within the lower, and main

basin of the middle, estuary. Furthermore, of all sites sampled during summer and winter, high

density and biomass of the species were recorded at one site only (Como). The high abundance

at Como was observed only during winter and comprised mainly juveniles. Dorsey (1981)

similarly noted generally low densities of A. ehlersi at his sites throughout Port Philip Bay

except for two locations with high densities. This patchy distribution pattern possibly reflects

areas with suitable conditions for the settling and development of larval A. ehlersi. Whether a

more intensive sampling effort throughout the lower- to mid-reaches of the estuary would reveal

other locations with higher densities of A. ehlersi remains to be determined, as does the

distribution of the species into sediments underlying water depths >2 m.

Adult A. ehlersi inhabit sediment underlying water of salinity ranging between 5-37 g kg-1 (both

this study and estuaries of New South Wales (Hutchings and Murray, 1984)). Further, adult

stages of A. ehlersi regulate osmotically their body fluids (McShane, 1977). Hence, given that

adult stages theoretically could tolerate the salinity regime within the upper reaches of

Australian estuaries, it is surprising, a priori, that the species exhibits a marine euryhaline

distribution (Hartman, 1954; Dorsey, 1981; Dürr and Semeniuk, 2000). The reason they are

absent from the upper reaches likely is due to the intolerance of the pelagic larval stage of even

brief exposures to freshwater (<8 g kg-1). Juveniles settled at Como at salinities ~ 8 g kg-1, but

this is likely the lower limit of their tolerance.

Since freshwater conditions in summer will occur, albeit infrequently and unpredictably

depending on the passage of rain-bearing depressions emanating from tropical cyclones across

the catchment, it is relevant to consider, therefore, the implications of the atypical summer-time

freshwater inundation of the Swan River Estuary that occurred during this study in February

2000. Summer salinities may be expected to be within the range 23-36 g kg-1 down-stream

from Bassendean, substantially higher than the 5-11 g kg-1 recorded in 2000, during which time

83

Page 98: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

only adult stages were collected, and at 4 locations only. These lowered salinities may have

caused considerable mortality of juvenile recruits, particularly since juveniles were present at

Como during the subsequent summer (February 2001); although recruitment of juveniles was

not occurring at Como during February 2000 prior to the freshwater flushing. Estuaries

typically experience salinity fluctuations of the water column that are substantially damped in

the sediment: larvae may find refuge from low water salinity by resting in the sediment. The

extent to which benthic populations, possibly comprising larval and adult stages, were affected

by the summer freshwater flushing in 2000 was not determined, but the impact of summer

flushing on distribution and recruitment of A. ehlersi are biologically significant issues to

elucidate in future research.

Determining the undoubtedly numerous, physical, chemical and biological factors controlling

the spatio-temporal distribution of A. ehlersi in the estuary was not an objective of this study,

but the results presented here point towards patchiness in distribution that results initially each

year from the balance between larval dispersal and post-settlement success in what is a very

dynamic environment (Chapter 2, Section 2.1) subject to considerable inter-annual variability.

Local populations of high abundance and biomass reflect successful recruitment and post-

settlement survival of settling larvae controlled by estuarine hydrodynamics and larval salinity

tolerance. Retention of such populations into adult stages may also rely on the availability of

sufficient allochthonous nutrients, possibly supplied via the nearby stormwater drain at Como,

and in the Port Philip Bay examples of local high-density populations, via outfalls from a

nearby sewage treatment plant (Dorsey, 1981).

4.4.3 Simplisetia aequisetis: Life History and Distribution

Life History

Mature and potentially reproductive S. aequisetis were present within the middle Swan River

Estuary during every month of the year and juveniles were recruited almost continuously. The

10 cohorts identified were very closely-packed in time. During the 1999/2000 summer, Cohort

1 comprised entirely potentially sexually mature individuals; their progeny probably were first

captured as Cohort 4 in mid- to late-autumn, based again on a forward-extrapolation on the

growth curve of Cohort 4. This extrapolation indicates an approximate 4-month lag period of

embryonic development during which larval S. aequisetis were generally too small to be

retained on a 1 mm sieve. Although the abundance of Cohort 1 decreased rapidly and

continually in numbers over summer (likely due to mortality), a very low density of individuals

that had grown to a large size remained at Como until mid-spring. Consequently, progeny from

long-lived individuals in Cohort 1 may have contributed to the recruitment of Cohorts 4, 5, 6, 7,

8, 9 and 10 throughout autumn 2000 to summer 2000/01.

84

Page 99: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Cohort 2 was present during the initial month of sampling mainly as immature individuals, with

growth and potential sexual maturity occurring from mid-summer through to early-winter.

Progeny from Cohort 2, together with progeny from Cohort 1, likely first appeared as Cohort 5

but perhaps also supplemented recruitment to Cohorts 4 and 6. Small juveniles from Cohort 3

were detected in early summer, 1999, but greatest abundance was observed during late-summer/

early autumn, 2000. Individuals in Cohort 3 likely were derived mainly from a cohort extant in

the spring prior to the start of sampling, but Cohort 1 probably contributed some late recruits. A

low proportion of individuals from Cohort 3 reached sexual maturity during late-summer, 2000,

to early-winter, 2000, with progeny perhaps contributing to the composition of Cohorts 5 and 6.

A proportion of individuals in Cohort 4 first reached sexual maturity in early-winter, 2000.

Further, Cohort 4 was the only potentially sexually mature cohort definitely present throughout

the remainder of winter and also in early spring of 2000. Its progeny likely largely contributed

towards juvenile recruits in Cohorts 7, 8, 9 and 10. Thus, winter-time reproduction, largely by

Cohort 4 (but with some contribution by Cohort 3 and perhaps Cohort 1) and resultant progeny

(Cohorts 5, 6, 7, 8 and 9) was responsible predominantly for the spring to early-summer peak in

biomass and productivity of S. aequisetis at Como. Finally, Cohorts 5, 6 and 7 potentially

reached sexual maturity in mid- to late-spring, 2000, with their progeny contributing to Cohort

10 during the subsequent 2000/01 summer. Simplisetia aequisetis biomass and productivity

were generally lowest during late-summer to mid-autumn, indicative of a decrease in the

abundance of: (i), Cohorts 1 and 2 during 2000; and of (ii), Cohort 9 plus the loss of Cohorts 7

and 8 in 2001 (likely due to mortality). It is interesting, therefore, that greatest reproductive

output likely occurred during winter when the overlying salinity was lowest; conversely, the

greatest mortality occurred during late-summer/early-autumn when overlying salinity was

highest (excluding the dip in salinity in February 2000 due to heavy unseasonal rain). The

influence of salinity and other factors on the reproductive strategy and population structure of

S. aequisetis in the Swan River Estuary are further discussed in the next Section.

Given this interpretation of cohort progression through the sampling period, the minimum and

maximum estimates of S. aequisetis life-span, assuming a 4-month larval development period,

are 7 (derived from Cohort 8) and 15 months (derived from Cohort 1 of adults reproducing in

spring, 1999, and potentially sexually active for up to 11 months), respectively. Thus, variation

in life-span may depend on the time of conception of each cohort. Regardless of finer-scale

variation in cohort life-spans, the broad estimate of 7-15 months compares with the

approximation of “about one year” life-span of S. aequisetis in Port Phillip Bay (Dorsey, 1981).

Dorsey (1981) also reported almost year-round reproduction of S. aequisetis; however, he

argued that reproduction did not occur during the winter months because of failure to collect

breeding tubes in June, July and August despite the presence of individuals of reproductive size.

85

Page 100: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Distribution

Previous studies have documented S. aequisetis in the upper and middle estuary (Chalmer et al.,

1976; Hodgkin, 1987; Rose, 1994), but this study is the first to confirm that the species’

distribution extends into the lower Swan River Estuary. Simplisetia aequisetis is thus a

euryhaline species, with neither dominant marine nor oligohaline affinities within the Swan

River Estuary. The density and biomass of S. aequisetis at sites within the downstream reaches

of the middle and lower estuary generally were similar during summer and winter in 2000,

suggesting some stability in the factors regulating biomass/density within these lower reaches

regardless of the season. However, whether this is a result of the freshwater inundation during

the summer needs to be evaluated. Within the lower estuary, S. aequisetis density and biomass

were greatest at The Spit, indicating that the extensive sand flats of this locality comprise a

highly conducive habitat. Nevertheless, the factors prescribing an optimal habitat for the

species, still to be defined likely are subject to spatio-temporal variation. For example, in the

late 1980s very high densities (~4000 m-2) of S. aequisetis were recorded at Attadale,

significantly greater than elsewhere in the middle estuary (Rose, 1994). Within the lower

estuary, S. aequisetis density and biomass decrease substantially towards the mouth of the

estuary; an observation consistent with a euryhaline as opposed to a marine status of the species.

The distribution pattern of S. aequisetis upstream from Como differed markedly in winter

compared to summer. During winter, S. aequisetis density and biomass peaking at locations

upstream of Como, were much higher than in summer by factors ranging approximately from 2-

100x, due predominantly to juvenile recruits. Kanandjembo (1998), using a 500 μm sieve that

presumably retained a larger juvenile component than the present study, reported extremely

high mean annual densities ~250,000 m-2 of S. aequisetis in shallow waters of the upper Swan

River Estuary during the mid-1990s. MacFarlane and Booth (2001) also found S. aequisetis to

be more abundant in areas of lower salinity within the estuary of the Hawkesbury River, New

South Wales. Given the above information regarding life history at Como, the greatest

reproductive output of S. aequisetis in the mid-estuary probably occurs during winter when

overlying salinities are lowest. However, the higher rate of reproduction in winter is likely to be

due to increased allochthonous organic matter inputs and food availability rather than to the

lowered salinity per se. Indeed, there may be a general benefit in recruiting as far upstream as

the salinity tolerance of juveniles allows, i.e. where there may be greater resource availability or

lowered competition/predation.

Since extreme salinity fluctuations are less likely to occur within the burrow environment than

in the overlying water (Smith, 1956; McShane, 1977), brooding behaviour may have evolved

predominantly as a protective mechanism enabling larvae to avoid stresses associated with

pelagic salinity fluctuations (Dorsey, 1981; Glasby, 1984, 1986). Thus, where there is a

86

Page 101: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

pronounced gradient or stratification of salinity between benthic and pelagic environments,

sedimentary brooding may facilitate larval retention within upstream benthic habitats defined by

physiologically tolerable interstitial salinities that are intermittently overlain by inhospitable

brackish or fresh waters that would exclude pelagic larvae. Times of low salinities in the upper

reaches of the Swan River Estuary, probably exclude infiltration by pelagic A. ehlersi larvae,

but the interstitial sediments remain available for benthic S. aequisetis larvae and adults. Thus,

the atypical, estuary-wide depression in salinity in February, 2000, likely had much less of an

impact on reproduction in S. aequisetis than in A. ehlersi.

Adult stages of S. aequisetis inhabit tubiculous burrows in sediments of the Swan River Estuary

(De Roach et al., 2002; Appendix E). In the current study, specimens ≥31 mm of unknown sex

were recorded brooding embryos in the burrows. Brooding was described for S. aequisetis from

Port Phillip Bay (Dorsey, 1981). Only larger, mature reproductive stages of S. aequisetis in Port

Phillip Bay are tubiculous, wherein males brood the larvae (Dorsey, 1982). Females of the

closely-related freshwater species, Simplisetia limnetica, from the Hawkesbury River, New

South Wales, also brood larval stages (Glasby, 1984, 1986). While the details of fertilisation

are unknown for both S. aequisetis and S. limnetica, Glasby (1984, 1986) suggests that it

follows deposition of the eggs on the wall of the male or female’s burrow.

The mean diameter of S. aequisetis eggs are similar from the Swan River Estuary (225 µm) and

Port Phillip Bay (210 µm; Dorsey, 1981). The diameter of S. limnetica eggs (245-270 µm) is

slightly larger. Although larger still in diameter, eggs of A. ehlersi (310 µm) are intermediate to

those of other dioecious, benthic spawning nereidids elsewhere that range from 135-162

(Laeonereis culveri) to 420-600 µm (Neanthes caudata) (Glasby, 1984, 1986). Embryonic

development of S. aequisetis was not examined in this study, but in Port Phillip Bay juveniles

are wholly or partially nourished by egg-yolk reserves and remain in the parental tube until

about the 13-setiger stage (Dorsey, 1981).

In contrast to the spawning strategy of the estuarine A. ehlersi, for example, the brooding

behaviour of S. aequisetis will result in developing larvae being retained within a sedimentary

environment that is likely more osmotically and thermally stable than the overlying water

(Smith, 1956; McShane, 1977). Dorsey (1981) and Glasby (1984, 1986) have hypothesised that

brooding protects the young from osmotic and thermal stresses and also predation, and that

limited dispersal is an opportunistic trait facilitating a rapid increase in abundance when

conditions are favourable. Consequently, brooding species could be predicted a priori to

exploit the upper regions of estuaries subject to wide fluctuations in salinity. The trade-off is a

more limited ability to disperse and colonise new habitat in the short-term (Dorsey, 1981;

Glasby, 1984, 1986).

87

Page 102: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

4.5 SUMMARY

Despite clear differences in life histories and population structure of A. ehlersi and S. aequisetis

in the Swan River Estuary, some similarities exist. For example, both species exhibit a mean

life-span of approximately 1 year, a P/B turnover rate of about 3 and potentially are capable of

reproducing throughout the year, peaking during winter to spring. However, A. ehlersi

illustrates a marine euryhaline distribution and is present only in the main basin and lower

estuary, typically at a very low density of adults; while S. aequisetis exhibits a true euryhaline

distribution, with an estuary-wide presence during both summer and winter.

High density and biomass of A. ehlersi was observed only in the middle estuary, mainly as

juveniles that were recruited throughout winter. It is hypothesised that gravid atokous adults

leave the sediment to spawn pelagically, with larval development occurring over a 2-4 month

period prior to settlement. It is further hypothesised that a physiological intolerance of

freshwater by the pelagic larval stage is the major reason limiting the expansion of the species

into the upper reaches of the Swan River Estuary. The extent of larval dispersal is unknown and

therefore it is unknown whether recruited juveniles are derived from local or external adult

populations. Regardless of derivation, settling larvae at Como grew to a body length of ~40 mm

over a period of 6 months and then, it is hypothesised, migrated elsewhere. The reproductive

status of medium-sized (migrating) A. ehlersi is also unknown, but the presence of

coelomoducts suggests a potential for iteroparity. Migrant cohorts of A. ehlersi arrived at Como

during autumn through winter, and grew to a body length of 60-75 mm over a period of 2-4

months. A life-span of 10-14 months is estimated for A. ehlersi in the Swan River Estuary.

In contrast, adult S. aequisetis brood eggs and embryonic larvae in tubiculous burrows. Sexual

maturity is reached at a minimum body length of ~30 mm. Brooding of young indicates a life-

cycle that is based entirely in sediments within a relatively stable interstitial salinity. Brooding

within burrows, it is suggested, enables exploitation by S. aequisetis larvae and adults of the

upper reaches of the Swan River Estuary. Simplisetia aequisetis is a true euryhaline species and

ubiquitous in distribution within the estuary. At sampling locations downstream from Como, S.

aequisetis density and biomass were generally similar during the summer and winter in 2000

when the summer-time salinity of overlying water was atypically low due freshwater runoff

from heavy unseasonal rain. Within the lower reaches, S. aequisetis abundance peaked at The

Spit, presumably indicative of optimal benthic habitat. The estuary-wide population of

S. aequisetis was supplemented substantially in winter by increases in density and biomass at

locations upstream of Como predominantly by juvenile recruits; thus imposing a winter-time

upstream shift in maximal density and biomass to the upper estuary. Larval development was

estimated to take 4 months. In the middle estuary, S. aequisetis cohorts were extant for a period

ranging from 3-11 months from which it was concluded the life-span ranged between 7-15

months.

88

Page 103: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

CHAPTER 5

EFFECTS OF AUSTRALONEREIS EHLERSI AND SIMPLISETIA AEQUISETIS

ON SEDIMENTARY NITROGEN CYCLING, INCLUDING THE

KINETICS OF NITRATE UTILISATION BY RESIDENT

DENITRIFYING BACTERIA

5.1 INTRODUCTION

The need for management strategies to combat the problem of increasing eutrophication within

near-shore waters including within the Swan River Estuary, Western Australia (Sections 2.1.3 and

3.3) has spurred research into the processes governing nutrient availability in aquatic ecosystems.

For a given loading of organic matter, the rate of diagenetic reactions (including biogenic organic

matter decomposition, nutrient regeneration and denitrification) and exchange of pertinent

substrates and products at the sediment-water interface can influence markedly the water column

concentration of nutrients available for algal production (Berner, 1976; McCaffrey et al., 1980;

Hammond et al., 1985). Bioturbation, sediment reworking and/or fluid irrigation by benthic

organisms, has a substantial role in determining bentho-pelagic nutrient fluxes (Rhoads, 1974;

Berner, 1976, 1980; Hartwig, 1976; Petr, 1977; Aller, 1980a, 1982, 1988; McCaffrey et al., 1980;

Zeitzschel, 1980; Callender and Hammond, 1982; Hammond et al., 1985; Krantzberg, 1985;

Kristensen, 1988, 2000; Andersen and Kristensen, 1991; Aller and Aller, 1998; Welsh, 2003).

Benthic communities rarely comprise single-species assemblages, and elucidating the holistic

impact of nutrient recycling by benthic macrofauna will involve quantification of the dynamics of

mixed, complex macrofaunal communities (Welsh, 2003). However, a reductionist approach has

been followed here to measure the contribution to nitrogen cycling, particularly in terms of

denitrification, nitrification, ammonification and net NOx- and NH4

+ flux rates, by the two nereidid

polychaetes of the Swan River Estuary, Australonereis ehlersi and Simplisetia aequisetis. This

approach is in accordance with the suggestion of Kristensen (2000) that in studying the effects of

bioturbation a range of animal species or functional groups of infauna should be investigated.

5.2 METHODS

5.2.1 Sampling site

Nereidids (A. ehlersi and S. aequisetis) and sediment were collected on 17 October, 2001, from

Como, Swan River Estuary, Western Australia (31°59.6’S, 115°51.1’E, Figure 4.1). Physico-

chemical properties of this location are described in Table 4.1 and in De Roach et al. (2002;

89

Page 104: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Appendix E). At the time of sampling, water temperature was 19 °C and salinity 21 g kg-1.

Ammonium concentration [NH4+] and nitrate (+ nitrite) concentration [NOx

-] in the overlying water

were both ~2 μM. Sediment with nereidids was collected manually to a depth of 15 cm while

snorkelling. Water (60 L) used in the initial set-up of the experiment was collected in situ. For

continued replenishment of water quality (see below), additional water (60 L) was collected every

two days from the more accessible Pelican Point (Figure 4.1), until termination of the experiment at

the end of December. During this period, salinity of the estuary at Point Pelican increased slowly

from 24 to 36 g kg-1. The mean (± s.e.) values of [NOx-] and [NH4

+] of estuary water over the

period were 1.99 ± 0.05 μM and 12.7 ± 0.90 μM, respectively.

5.2.2 Sediment and nereidid preparation

In the laboratory, sediment and nereidids were transferred into V-cores (Figures 5.1 and 5.2;

Kristensen, 1984a) constructed according to De Roach et al. (2002; Appendix E). V-core

dimensions were 47 mm (length) by 16.5 mm inner diameter (i.d.) for uninhabited and S. aequisetis

inhabited cores, and 55 by 25 mm i.d. for A. ehlersi inhabited cores, reflecting spring population

densities at the study site (~1000 A. ehlersi m-2 and ~ 2300 S. aequisetis m-2), and natural burrowing

depth (De Roach, personal observation). To maintain homogeneity between V-cores, all sediment

utilised was sieved simultaneously through a 2 mm mesh and thoroughly mixed. This also

facilitated collection of the nereidids. Sediment re-establishment after sieving was chosen over

other methods of sediment collection and defaunation for reasons outlined in De Roach et al. (2002,

Appendix E).

Fifty-five V-cores were filled with sieved sediment to within 1-2 mm from the top of the cores

(Figure 5.1). A worm was introduced to 40 of these (20 with A. ehlersi ~ 200 mg wet weight, 20

with S. aequisetis ~ 10 mg wet weight) and 15 were left uninhabited. The polychaetes rapidly

buried into the sediment and each core was randomly placed into one of two aerated 20 L holding

aquaria containing water collected in situ, at room temperature (~ 22 °C). Within the aquaria, both

ends of each V-core were capped with a plastic lid containing a central aperture of 2 mm diameter

(Figure 5.1). The hole allowed solute flux between sediment and overlying water and, in the case of

A. ehlersi inhabited cores, provided a site of attachment for a silicone tube (12 x 1.6 mm i.d.,

considered an extension of the burrow) in which burrow ventilation could be measured. To avoid

the build-up of excretory products, ~ 80% of the water was replaced every two days (the remaining

20% maintained saturation of the cores), until the end of the experiment. The period of acclimation

of V-cores within the holding tanks was no less than 31 days, enabling sediment microbial

communities to re-establish after the disturbance of sediment sieving.

90

Page 105: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Each core was visually checked daily; any core with a dead nereidid (easily distinguished by a rapid

reduction to black sediment) was removed. For cores with A. ehlersi, only those exhibiting a

permanent burrow, continuous from one end of the core to the other, were used. Initial inspections

of cores with S. aequisetis revealed that burrow formation did not occur. Sacrifice of a few cores

revealed that the specimens were alive but living interstitially. This habit was unexpected since De

Roach et al. (2002; Appendix E) had demonstrated that larger specimens of this species construct

well-defined, U-shaped burrows. Mortality and loss of core-suitability over the period were such

that eight each of A. ehlersi and S. aequisetis inhabited cores, and nine uninhabited cores, were

utilised.

Figure 5.1: Experimental set-up depicting a sediment-filled V-core inhabited by A. ehlersi

(adapted from Kristensen et al., 1991). Other species treatments of V-cores (S. aequisetis inhabited, uninhabited) are interchangeable. Acetylene supply and flow sensor can be omitted where appropriate (for further details see text).

5.2.3 Experimental Procedure

The kinetic-fix adaptation of the acetylene block method outlined here was utilised to determine not

only denitrification rate in vitro, but also nitrification, ammonification, total NOx- flux and total

NH4+ flux rates within individual inhabited or uninhabited sediment cores. The influence of

acetylene on various microbial nitrogen cycling processes is described in Section 3.2.1 and the

theory of denitrification kinetics in Section 3.2.2. For each V-core (uninhabited, A. ehlersi and

2 cm

Acetylene Air

Water Sampler Flow sensor

Worm

91

Page 106: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

S. aequisetis inhabited), NOx- and NH4

+ fluxes within the overlying water were measured both in the

presence (+) and absence (-) of C2H2 over a range of idealised ambient NO3- concentrations [S] ≅ 0,

20, 50, 100, 500 μM. KNO3 addition to a sample of estuary water of near 0 μM NO3- (but not

exactly known), resulted in actual values of [S] slightly deviating from idealised concentrations

(due to slight variation in initial estuary water NO3- concentration and small error in KNO3

amendments). Individual V-cores were subject to one ambient NO3- treatment only, with NH4

+ and

NOx- flux monitored first in the absence and then in the presence of acetylene. Flux rates of NOx

-

and NH4+ within water overlying V-cores were determined by the difference in initial and final

solute concentrations over a three hour period.

Initially, a randomly chosen V-core was placed into an 8 L incubation aquarium containing estuary

water amended with NO3- to an appropriate [S] (the sequence of exposure to ambient NO3

-

concentration was also randomised for subsequent cores). The aquarium was aerated and held at

20 °C by a constant temperature water bath in darkness, for a NO3- incubation period of 48 hours.

The same ambient conditions, including [S], were maintained in all experimental containers and

aquaria as described below. The volume of water in each aquarium was calculated such that the

estimated reduction in NO3- concentration of the overlying water throughout the period of V-core

exposure did not exceed 10%, even at very high denitrification rates. This limited proportion of

NO3- depreciation was necessary to maintain enzyme-driven denitrification at a linear rate, an

important requirement for Michaelis-Menten modelling of denitrification kinetics (Copeland, 1996;

Section 3.2.2).

Following NO3- incubation, the V-core was transferred to an aerated experimental set-up (Figure

5.1) in which solute fluxes could be measured in the absence of C2H2 and, in the case of A. ehlersi

inhabited cores, burrow ventilation monitored. The core was placed into a pre-moulded groove

within a plaster-of-Paris base and an initial (t = 0 min) 3 mL sample of the overlying water was

taken by syringe via a sampling port (Figure 5.1). It is assumed that the distribution of solutes

within the overlying water was sufficiently mixed by the aeration process. The water sample was

injected into a 3 mL ‘Venoject’ evacuated test-tube (Terumo Corporation, Belgium) and

immediately placed in a freezer to be stored at -4 ˚C until analysis. A final water sample was taken

after three hours (t = 180 min) using the same method. The experimental container held 500 mL

(A. ehlersi treatment) or 400 mL (uninhabited and S. aequisetis treatments) of estuary water

amended to the appropriate [NO3-]. These volumes were estimated to be minimal enough to allow

sediment N-transforming processes to impart a measurable effect upon overlying solute

concentrations over the 3 hour period, whilst being large enough not to allow a 10% depreciation in

ambient NO3- concentration.

92

Page 107: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Australonereis ehlersi burrow ventilation was monitored continuously by placing the tip of a FS20

micro-flow sensor (Unisense, Denmark) within the centre of the opening of the silicone tube burrow

extension (Figure 5.1). The flow sensor was positioned and held by a micromanipulator and

attached to a PA2000 Picoammeter (Unisense). An analog-digital processor (DI-705, DATAQ

Instruments, U.S.A) and data acquisition system (DI-700, DATAQ) facilitated PC recording of

ventilation waveform data using WinDaq (DATAQ) software. To minimise the need for

calibrations of the flow sensor, it was kept polarized between measurements. Calibration of the

flow sensor was conducted every two days in accordance with the manufacturer’s instructions8.

Subsequent to the three hour C2H2-free monitoring period, the V-core was transferred to an 8 L

C2H2 incubation aquarium containing ambient water but continuously purged by a mixture of 10%

C2H2, 90% air, for 20 hours (Kristensen et al., 1991). As described in Section 3.2.1, acetylene

blocks the nitrification process by inhibiting the oxidation of NH4+ to NO2

- (Hynes and Knowles,

1978; Walter et al., 1979). The lack of NO3- production and measurement of NO3

- loss in the

presence of C2H2, permits an estimate of denitrification rate (provided that assimilatory NO3- uptake

is prevented by dark incubation, and that the rate of dissimilatory NO3- reduction to NH4

+ is

negligible – both here assumed true in vitro; further see Section 3.2.1 for methodological theory and

the footnote of Section 5.3.4 for justification of the latter assumption). Also, determination of

denitrification rates over a range of [S] facilitates investigation of the apparent enzyme kinetics of

the denitrifier community within sediments of each treatment (i.e. either inhabited by A. ehlersi or

S. aequisetis, or uninhabited) (further see Section 3.2.2). If Michaelis-Menten (steady-state) type

kinetics are exhibited, then the apparent Michaelis (half-saturation) constant (Kapp) and maximum

potential denitrification rate (Vmp) can be determined (Section 3.2.2).

The V-core was subsequently returned to the experimental set-up, with a slow stream of C2H2

incorporated (Figure 5.1), and the water sampling process repeated. The experimental and C2H2

incubation set-ups were housed within a darkened fume-cupboard which collected escaping C2H2

and also flow sensor tracer (H2) gas. After the final period of +C2H2 water sampling, a visual record

of sediment redox status was taken by photographing the towel-dried V-core. The wet sediment

weight and height within both halves of the V-core were measured to ascertain sediment bulk

density.

8 Briefly, the flow sensor was placed in a calibration set-up whereby the tip was immersed in a known gravity-fed flow of estuary water exiting a from silicon tube. The velocity of the flow was controlled by clamping and calculated by measuring the time it took to fill a standard volume. The flow sensor was exposed progressively to a series of set flow rates within the range of 0.5 to 20 mL min-1 and the signal (pA) was logged when it had stabilized at each velocity. These readings were used to produce a calibration curve, through which a measured signal was converted into a velocity value.

93

Page 108: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

For each core, the duration of the entire process totalled 3 days 2 hours; initial NO3- incubation (48

h), -C2H2 monitoring (3 h), C2H2 incubation (20 h) and +C2H2 monitoring (3 h). To enable

monitoring of two cores within this period, the NO3- and C2H2 incubation aquaria were duplicated

and the two cores were subject to the same process at a staggered interval of 5 hours (i.e. the first

core incubation was initiated in the morning and the second in the afternoon). Four days were

required, therefore, to monitor both + and - C2H2 nitrogenous fluxes in the water overlying two V-

cores, and experimental completion of all 25 cores took 52 days.

5.2.4 NH4+ and NOx

- analyses

Water samples were thawed to room temperature (~ 22˚C) and analysed for [NH4+] and [NO3

- +

NO2-] using a Skalar Sanplus Segmented Flow Autoanalyser System. The automated colourimetric

procedures for the determination of [NH4+] and [NO3

- + NO2-] are given by the American Public

Health Association (1995).

5.3 RESULTS

5.3.1 Sediment Colour and Bulk Density

The presence of A. ehlersi and S. aequisetis in sediment cores had a marked effect on sediment

colouration, visible from the second day of core acclimation until experiment termination (Figure

5.2). All cores were vertically colour-stratified into three layers of approximately the same

proportions: 0 to 3.0 cm (surface), 3.0 to 3.1-3.3 cm (mid) and 3.3 to 4.0 cm (deep). The deep

sediment colour of all cores was light grey, whilst the mid-depth interval was a band of orange

sediment that was generally 1-2 mm thicker in uninhabited cores. Uninhabited and A. ehlersi

inhabited cores contained dark brown surface sediment, whilst the same depth interval was light

brown for C. aequisetis inhabited cores. Average (± s.e.) wet bulk densities for uninhabited, A.

ehlersi inhabited and S. aequisetis inhabited cores were 1.61 ± 0.02, 1.68 ± 0.03 and 1.54 ± 0.04 g

cm-3, respectively.

Uninhabited A. ehlersi inhabited S. aequisetis inhabited

Figure 5.2: Photographs of representative V-cores, taken at termination of the experimental period, illustrating sediment colouration profiles indicative of redox status.

5 cm

94

Page 109: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

5.3.2 Ventilation of A. ehlersi Burrows

The patterns of burrow ventilation exhibited by individual A. ehlersi specimens within sediment

cores over the experimental range of ambient NO3- concentrations, both pre- and post C2H2

incubation, are presented in Figure 5.3. Average ventilation was determined as the cumulative area

under each ventilation curve (absolute flow), standardised for time (Table 5.1 and Figure 5.3). The

degree and pattern of burrow ventilation in cores inhabited by A. ehlersi varied. Ventilation was

absent in the two cores subject to +500 μM NO3- (idealised) both prior to and during acetylene

exposure, except for two periods of low ventilation in one of the cores whilst exposed to acetylene.

Excluding the cores at +500 μM NO3-, average ventilation decreased 5-27-fold in burrows after

acetylene exposure, except for one core subject to +0 μM NO3- (idealised) that showed a 4-fold

increase in average ventilation. Average ventilation through burrows ranged from 21-578 mL h-1

prior to, and 3-113 mL h-1 following, acetylene exposure. The percentage of time spent ventilating

the burrow (i.e. flow >0 mL min-1) also decreased following acetylene exposure, and with the

exception of the core subject to +0 μM NO3-, the percentage prior to acetylene exposure was always

≥ that following exposure (Table 5.1). In the absence of acetylene, typical ventilation patterns were

assessed to be represented by Figures 5.3 [ai], [di], [ei] and [fi], and normal average ventilation was

considered the mean of the average ventilation rates exhibited by these inhabited cores, 415 mL h-1.

Table 5.1: Average burrow ventilation rate and the percentage of time for which burrow ventilation >0 ml min-1 for A. ehlersi individuals exposed to increasing initial nitrate treatments.

Average burrow ventilation

(mL h-1)

% time burrow ventilation

>0 mL min-1Initial Nitrate

Treatment (μM) - C2H2 + C2H2 - C2H2 + C2H2

0 578 25 100 53

0 27 113 59 80

20 21 3 72 16

50 323 69 100 100

50 373 66 98 43

100 385 14 99 22

500 0 0 0 0

500 0 11 0 24

95

Page 110: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

0 20 40 60 80 100 120 140 160 180

500 µM

0

2

4

6

8

10

500 µM

0

2

4

6

8

10

50 µM

0

2

4

6

8

10

- C2H2[Nitrate]

0 µM

0

2

4

6

8

10

0 20 40 60 80 100 120 140 160 180

+ C2H2

0 µM

0

2

4

6

8

10

20 µM

0

2

4

6

8

10

50 µM

0

2

4

6

8

10

100 µM

0

2

4

6

8

10

Time (min)

Bur

row

ven

tilat

ion

rate

(ml m

in-1

)

Figure 5.3: Time-specific burrow ventilation pattern of A. ehlersi inhabited V-cores at increasing

initial nitrate treatments [Nitrate], in the presence (right) and absence (left) of acetylene. To reduce ‘noise’, maximal ventilation rate has been smoothed for each treatment core.

96

Page 111: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

5.3.3 Denitrification

Potential denitrification rates (V0) in acetylene-inhibited, NO3--amended sediment cores increased

with respect to ambient NO3- concentration (S) for all three treatments (A. ehlersi and S. aequisetis

present, uninhabited), in a general Michaelis-Menten type manner (Figure 5.4[ai-iii]). The linear

Hanes-Woolf transformations used to determine the apparent kinetic constants are presented in

Figure 5.4[bi-iii]. Substrate (NO3-) loss never amounted to a >10-12.5% reduction of initial levels

in water overlying uninhabited and S. aequisetis inhabited cores; however, for NO3- amendments

less than +500 μM (idealised), NO3- loss in water overlying A. ehlersi inhabited cores typically was

11.5%-17.7%, but 25% and 31% in the +0 μM amendments.

The microbial sediment community of uninhabited cores demonstrated a higher Vmp (4.33 mmol N

m-2 h-1) and lower Kapp (133 μM NO3-) than that of S. aequisetis inhabited cores (Vmp = 3.83 mmol

N m-2 h-1; Kapp = 202 μM NO3-) (Fig’s 5.4[ai-ii], 5.4[bi-ii]). The estimated denitrification rate at

‘natural’ overlying NO3- concentration (at the sampling site during the study period - V2μM) was

higher for uninhabited sediment (64 μmol N m-2 h-1) than for S. aequisetis inhabited sediment (38

μmol N m-2 h-1).

Potential denitrification rates of A. ehlersi sediment cores measured over the experimental range of

S, as well as those values subjectively standardised to account for depressed burrow ventilation in

the presence of C2H2, are presented in Figure 5.4[aiii]. Measured denitrification rates in A. ehlersi

inhabited cores (Figure 5.4[aiii], black dots) were generally similar to or slightly greater than the

denitrification curve established for uninhabited cores (Figure 5.4[aii]), albeit in burrows where

acetylene had, for the most part, greatly retarded ventilation (see above, Ventilation results).

Further, it was generally observed that the measured V0 values of A. ehlersi inhabited cores most

similar (closest) to the uninhabited denitrification curve were those with zero to very low burrow

ventilation; and conversely that A. ehlersi cores demonstrating higher ventilation rates exhibited V0

values proportionately greater than the uninhabited denitrification curve. To estimate potential

denitrification of A. ehlersi cores at normal (typical non-acetylene affected) burrow ventilation, the

measured V0 values were therefore standardised utilising the uninhabited denitrification curve as a

baseline (representing A. ehlersi potential denitrification at zero burrow ventilation), and using the

realised average ventilation flow of each core as a scaling factor with respect to normal average

ventilation (415 ml h-1) (Figure 5.4[aiii]). Algebraically:

97

Page 112: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Amden = Uden + (p * (Aden – Uden)) Equation 3

where, Amden is the measured denitrification rate of the A. ehlersi inhabited core in the presence of

C2H2; Uden is the expected denitrification rate of uninhabited cores at relevant S (equivalent to

expected denitrification rate of A. ehlersi cores at zero burrow ventilation or the baseline estimate);

p is the proportion of normal, average ventilation exhibited by the A. ehlersi specimen in the

presence of C2H2; and Aden is the expected denitrification rate of the A. ehlersi sediment core at

normal ventilation in the absence of acetylene (ventilation-standardised estimate). Aden values were

estimated by rearrangement of Equation 3:

Aden = Uden + ((Amden – Uden) / p) Equation 4

The Hanes-Woolf linear transformation regression fit (r2 = 0.98) of ventilation-standardised

denitrification estimates is presented in Figure 5.4[biii]. The ventilation-standardised Vmp estimated

for A. ehlersi inhabited sediment cores (4.75 mmol N m-2 h-1) was greater than that of either

uninhabited or S. aequisetis inhabited sediment cores, whilst the associated Kapp estimate for A.

ehlersi inhabited cores (50 μM NO3-) was lower than either of the other two core treatments. At

natural overlying nitrate concentration (2 μM), the denitrification rate of A. ehlersi inhabited

sediment (standardised to normal ventilation) was estimated at 182 μmol N m-2 h-1, which is greater

than that determined for either uninhabited or S. aequisetis inhabited cores.

5.3.4 Nitrification and Net NOx- Flux

In the absence of acetylene, total NOx- flux in uninhabited, S. aequisetis and A. ehlersi inhabited

sediment cores responded to increasing initial NO3- concentration in the same generalised manner;

net efflux from sediment to overlying water at lower initial NO3- concentration, tending towards net

influx from overlying water to sediment at higher initial NO3- concentration (Figure 5.4[ci-iii],

black dots). NOx- flux ranged from 2.77 mmol N m-2 h-1 (at [S] = 24 μM) to -0.56 mmol N m-2 h-1

(at [S] = 510 μM) within S. aequisetis sediment cores (Figure 5.4[ci]); from 1.87 mmol N m-2 h-1 (at

S = 0.5 μM) to -2.21 mmol N m-2 h-1 (at [S] = 475 μM) within uninhabited cores (Figure 5.4[cii]);

and, from 2.65 mmol N m-2 h-1 (at [S] = 1.0 μM) to -1.65 mmol N m-2 h-1 (at [S] = 480 μM) within

A. ehlersi inhabited cores (Fig. 5.4[ciii]).

98

Page 113: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

NO

x- flu

x(m

mol

m-2

h-1

)

-2

-1

0

1

2

3

(i) C. aequisetis

NO

x- flu

x [v

](m

mol

m-2

h-1

)

-5

-4

-3

-2

-1

0

(ii) Uninhabited (iii) A. ehlersi

NO

x- flux

/ In

itial

[NO

3- ](S

/v)

-200

-150

-100

-50

0S/v = -0.261S - 52.77

R2 = 0.925

S/v = -0.230S - 30.59R2 = 0.957

S/v = -0.211S - 10.56R2 = 0.983

+ C2H2

Ventilationstandardised

estimate

NH

4+ flux

(mm

ol m

-2 h

-1)

0

1

2

3

0 100 200 300 400 500

NH

4+ flux

(mm

ol m

-2 h

-1)

-2

-1

0

1

Initial NOx- concentration [S]

(µM)

0 100 200 300 400 500 0 100 200 300 400 500

(a)

(b)

(c)

(d)

(e)

+ C2H2

+ C2H2

- C2H2

- C2H2 Nitrification estimate

(i) S. aequisetis

Figure 5.4: Dissolved inorganic nitrogen (DIN) fluxes measured in the water overlying (i) S. aequisetis inhabited, (ii) uninhabited and (iii) A. ehlersi inhabited V-cores, at increasing initial concentrations of nitrate [S].

(a) Nitrate flux measured in the presence of acetylene, indicating potential denitrification rates [v]. (b) Linearisation [S/v] of (a) illustrating the regression equation used to generate the fitted curves in (a). (c) Nitrate flux measured in the absence of acetylene and estimates of nitrate flux due to nitrification (illustrating

constant average). The idealised curve illustrated represents the net product of estimated nitrification (= constant), minus the denitrification curve in (a).

(d) Ammonium flux measured in the presence of acetylene, indicating the average production from mineralisation plus, in the case of inhabited V-cores, excretion.

(e) Ammonium flux measured in the absence of acetylene, indicating the net product of ammonium produced via mineralisation/excretion minus that lost via nitrification.

99

N.B. Estimated curves of nitrate flux in water overlying A. ehlersi inhabited V-cores have been subjectively standardised to represent ‘normal’ burrow ventilation activity. The rate of dissimilatory nitrate reduction to ammonia is assumed to be negligible. See text for further details.

Page 114: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Total NOx- flux was assumed to be the net result of NOx

- produced via nitrification minus that

consumed by denitrification. For uninhabited and S. aequisetis inhabited sediment cores,

nitrification estimates were therefore determined by adding the expected denitrification rate

(calculated via substitution of the relevant [S] into the appropriate Michaelis-Menten equation) to

the measured total NOx- flux (Figures 5.4[ci-ii], cross-hatches). Sediment nitrification rate was

generally constant over the range of [S] and average nitrification rate was highest within

S. aequisetis inhabited cores (2.76 ± 0.17 [s.e.] mmol N m-2 h-1) and lowest in uninhabited cores

(1.86 ± 0.13 mmol N m-2 h-1) (Figure 5.4[ci & ii]).

To account for (non-C2H2 related) variance in burrow ventilation, determination of nitrification

estimates for A. ehlersi inhabited cores incorporated a subjective scaling process (similar to the

ventilation-standardisation of denitrification estimates). The baseline of nitrification (i.e. that

presumed at zero-flow ventilation) generated by A. ehlersi cores was set to equal the average

nitrification rate of uninhabited cores, with increases above the baseline directly-proportional to

average ventilation of individual burrows, again using 415 mL min-1 to represent ‘normal’

ventilation. Measured total NOx- flux was inferred to be resultant of the ventilation-related

proportion of normal nitrification (as described) minus the same ventilation-related proportion of

normal denitrification (calculated at relevant [S], as described for ventilation standardisation of

denitrification estimates). Algebraically:

ANOx = [Unit + (p * (Anit – Unit))] – [Uden + (p * (Aden – Uden))] Equation 5

where, ANOx is the measured total NOx- flux of the A. ehlersi inhabited sediment core (non-

standardised for variance in burrow ventilation); Unit is the estimated nitrification rate of

uninhabited cores (equivalent to expected nitrification rate of A. ehlersi cores at zero burrow

ventilation or the baseline estimate); Anit is the expected nitrification rate of the A. ehlersi sediment

core at ‘normal’ ventilation (ventilation-standardised estimate); and other variables are as described

above. Anit values were estimated by rearrangement of Equation 5:

Anit = Unit + {[[Uden + (p * (Aden – Uden))] + ANOx – Unit] / p} Equation 6

The ventilation-standardised nitrification estimates of A. ehlersi inhabited sediment cores are

presented in Figure 5.4[ciii]. The average nitrification rate (2.43 ± 0.07 mmol N m-2 h-1) was lower

than that estimated for S. aequisetis inhabited cores but higher than that of uninhabited sediment

(Figure 5.4[ci-iii]). Figure 5.4[ci-iii] also exhibits, for each treatment, the curve of expected total

100

Page 115: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

NOx- flux estimated by subtraction of the defined Michaelis-Menten denitrification curve from the

curve (constant) of average nitrification rate. Expected total NOx- flux determined by this method

displays a very good fit to measured total NOx- fluxes for each species treatment of sediment cores9

(Figures 5.4[ci-iii]). The values of [S] at which total NOx- flux shifts from efflux to influx (zero net

NOx- flux, predicted from expected total NOx- flux curves) were estimated at 520 μM for S.

aequisetis inhabited cores, 100 μM for uninhabited cores and 52 μM for A. ehlersi inhabited cores.

5.3.5 Ammonification and Net NH4+ Flux

NH4+ production in the presence of acetylene provided an estimate of ammonification - here defined

as microbial mineralisation plus, in the case of inhabited sediment, nereidid excretion. Within each

of the species treatments of sediment cores, ammonification in the presence of acetylene was

generally constant in relation to [S] (Figure 5.4[di-iii]). Average ammonification was greatest in

A. ehlersi inhabited cores (2.20 ± 0.05 mmol N m-2 h-1, Figure 5.4[diii]), intermediate in S.

aequisetis inhabited cores (1.84 ± 0.16 mmol N m-2 h-1, Figure 5.4[di]), and lowest in uninhabited

cores (0.91 ± 0.05 mmol N m-2 h-1, Figure 5.4[dii]). Due to the uncertainty of the relationship

between A. ehlersi ventilation activity and magnitude of excretion, and also the inability to separate

the excretion and mineralisation components of ammonification, NH4+ fluxes could not be

standardised with respect to burrow ventilation within either ± C2H2 treatments of A. ehlersi cores.

Within sediment cores inhabited by A. ehlersi, net NH4+ flux in the absence of acetylene ranged

from an efflux of 0.54 mmol N m-2 h-1 to an influx of -0.43 mmol N m-2 h-1 across the range of S,

representing a slight average efflux of 0.07 ± 0.13 mmol N m-2 h-1 (Figure 5.4[eiii]). In the absence

of acetylene there was a net uptake of NH4+ from the overlying water at all ambient NO3

-

concentrations of, on average, -0.62 ± 0.08 mmol N m-2 h-1 from S. aequisetis inhabited cores and,

on average, -0.85 ± 0.12 mmol N m-2 h-1 from uninhabited cores (Figure 5.4[ei-ii]).

9 For each core species treatment, the good fit of the observed NOx

- flux curve to that expected by subtracting the denitrification curve from the observed nitrification constant (both illustrated in Figure 5.4[c]), supports the assumption that denitrification was the major source of NOx

- loss and that the rate of dissimilatory nitrate reduction to ammonium was negligible (the observed and expected curves would otherwise show substantial deviation, if the rate of DNRA was significant). The likely absence of requisite highly anoxic micro-environments is probably implicated in the observed lack of DNRA within V-cores (further see discussion of the DNRA process in Appendix D, Section D.6).

101

Page 116: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

5.4 DISCUSSION

5.4.1 Overview

I will describe initially the occurrence and magnitude of the main nitrogen transformation processes

(denitrification, nitrification and ammonification) and net flux rates of NOx- and NH4

+ occurring

within the sediments of uninhabited V-cores (Figure 5.5[a]). Although a comparison to the relative

magnitude of nitrogenous processes and flux rates occurring within sediments of A. ehlersi and S.

aequisetis inhabited V-cores (respectively, Figures 5.5[b] and 5.5[c]) is necessary, my focus firstly

is to describe the base-line case within uninhabited sediment. Consideration of the likely reasons

for the observed effects of each nereidid species on these nitrogenous processes/fluxes is delayed

until the habit of each species is discussed. Also, in all cases the environmental conditions in vitro

entailed dark incubation of V-cores and, within the overlying water, a constant initial NH4+

concentration of about 13 µM.

Under the conditions of the measurements, the average rate of both nitrification and ammonification

were lowest in the absence of macrofauna, whilst denitrification was moderate in comparison to the

two nereidid species sediment treatments (Figure 5.5). The balance of these nitrogen cycling

processes in uninhabited V-cores resulted in a net high average rate of NH4+ influx to sediment

(Figure 5.5[a]). The flux direction of NOx- in uninhabited (and inhabited) sediments was dependent

on the NO3- concentration of the overlying water [S]: NOx

- flux shifted from an efflux at low [S] to

an influx to sediment at high [S]. The shift is due to the denitrification rate increasing in a

Michaelis-Menten type fashion in response to increasing [S] (Figure 5.4[a]); i.e. the efflux of NOx-

at low [S] switches to an influx at higher [S] because of the higher rate of sedimentary NOx-

consumption by denitrifying bacteria. Within uninhabited sediment at low [S], the average efflux of

NOx- was low in comparison to the two nereidid species sediment treatments; whilst at high [S] the

average influx of NOx- was moderate compared to the inhabited treatments (Figure 5.5). Also,

within uninhabited sediment cores, zero NOx- flux occurred when [S] was about 100 µM, which is

intermediate to the comparative zero NOx- flux [S] value of V-cores inhabited by the two nereidid

species (Figures 5.4[c] and 5.5).

My interest lies in determining the impact of A. ehlersi and S. aequisetis on these processes

(denitrification, nitrification, ammonification) and resultant net fluxes of NOx- and NH4

+. In order

to interpret the role of these two resident nereidid species on nitrogen cycling processes in

sediments of the Swan River Estuary, it is imperative to consider the potential effects of their

infaunal habit.

102

Page 117: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Low [S]

NH4+ NOx

-

N2

NH4+ NOx

-

N2

NH4+ NOx

-

N2

(a) Uninhabited

(b) A. ehlersi

(c) S. aequisetis

High [S]

-0 NOx flux @

Water den Sediment

nit

min

org-N

Low [S] High [S] -0 NOx flux

@ Water

den Sediment

nit

min + excr

org-N

Low [S] High [S]

-0 NOx flux @

Water den Sediment

nit

min + excr

org-N

Figure 5.5: Conceptualisation of the major nitrogen transformation and flux processes within V-core sediment in vitro (incubated in darkness) that is either: (a) uninhabited by macrofauna; (b) inhabited by individual adult A. ehlersi or (c) inhabited by individual juvenile S. aequisetis. Black arrows indicate the relative rate of denitrification (den), nitrification (nit) and ammonification (i.e. mineralisation (min) plus, in the case of inhabited sediment, nereidid excretion (excr)). Grey arrows indicate the net rate and direction of NH4

+ flux. The net rate and direction of NOx- flux (open arrows) is dependent on the

initial ambient concentration of NO3- in the overlying water [S]. Thus illustrated are NOx

- flux rate and direction at both low and high [S], in addition to the approximate concentration of [S] when NO - +

103 flux = 0. Initial ambient NHx 4 concentration in the overlying water was held constant at ~ 13 µM. The size of arrows indicates the relative magnitude of each nitrogen transformation or flux process (NB. not to scale, see Sections 5.3.3 to 5.3.5 for actual rates).

Page 118: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

5.4.2 Nereidid Habit

Each comparatively large A. ehlersi individual (200 mg wet weight) constructed, remained within

and sporadically irrigated permanent, U-shaped burrows within the experimental V-cores. The

sediment in these cores had the highest bulk density and superficially was similar in colour (dark

brown) to that within uninhabited cores. Burrow ventilation presumably enhanced the transport of

overlying oxygenated water to deeper sediment, however either oxygen consuming processes within

the burrow were too prolific or radial diffusion of oxygen was not sufficient to produce a noticeable

discolouration of sediment at the outer edge of the V-cores (radius 12.5 mm). In cross-section,

however, a 1-2 mm ring of light brown sediment was always observed around the burrow lining

with generally wider annuli surrounding the more heavily ventilated burrows. The presence of a

burrow space likely compacted the outer sediment and increased the bulk density of A. ehlersi

inhabited V-cores. Although a decrease in interstitial spacing may restrict solute (including

oxygen) transport, it is unclear whether this compaction effect is an artefact of the presence of the

perspex core lining acting as barrier to lateral sediment movement. The habit of A. ehlersi observed

here is similar to that of many northern hemisphere Nereis spp. (e.g. N. diversicolor, N. virens, N.

succinea) which have been well studied for their effects on sedimentary nitrogen cycling; and thus

comprise a point of reference for inter-species comparisons (discussed below).

Simplisetia aequisetis, on the other hand, reworked the sediment in the present study and did not

generate permanent burrows. The V-core sediments had a lower bulk density and lighter brown

surficial colouration than both A. ehlersi inhabited and uninhabited sediments. The lighter colour is

suggestive of a more aerobic environment and it is hypothesised that the sediment reworking by

these polychaetes likely increased interstitial spacing, facilitating greater oxygenation of the entire

3 cm surface layer of sediments. The absence of permanent burrow construction by S. aequisetis in

this study contrasts to the previous study of De Roach et al. (2002), wherein permanent U-shaped

tubes were constructed and ventilated by individuals residing within similar V-cores. Dorsey

(1982) has noted previously that polychaetes of the species S. aequisetis do not inhabit permanent

burrows during most of their life-span, although reproductive individuals construct burrows wherein

eggs are deposited and larvae brooded. Within the Swan River Estuary, sexual maturity of S.

aequisetis specimens occurs at a body size within the range of ~17-39 mg wet weight

(corresponding to 30–50 mm total length; Chapter 4). Juveniles of about 10 mg wet weight were

utilised in the present study in contrast to the large (~80 mg wet weight), mature tube-building

individuals studied by De Roach et al. (2002)10. Thus, construction of permanent burrows by

specimens of S. aequisetis apparently is age-dependent: mobile immature juveniles rework 10 In the Materials and methods section of De Roach et al. (2002), the biomass ~0.8 g wet weight should read ~ 0.08 g wet weight.

104

Page 119: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

sediment, while reproductively mature individuals inhabit permanent brooding burrows. In contrast

to the well-examined role of burrows in sedimentary nitrogen cycling, fewer studies have

investigated the effects of polychaete species that do not form burrows. Thus, for purposes of

comparison with juvenile S. aequisetis, where possible, I have considered below the impact of

sediment reworking by polychaetes and also various other species upon nitrogenous processes and

flux rates. However, the extent to which the habits of different species may represent functional

groups, i.e. sets of species or habits that impart similar effects on nitrogen-cycling processes (sensu

Chapin et al., 1992), is beyond the scope of this discussion, but is the subject of Chapter 6.

5.4.3 Denitrification

The presence of A. ehlersi considerably enhanced denitrification in comparison with uninhabited

sediment (Figure 5.5). At in situ ambient nitrate concentration, the denitrification rate (V2μM) of V-

core sediment in the presence of A. ehlersi was estimated to be 284% of uninhabited sediment. This

compares with a range of 138-500% enhancement of denitrification over defaunated sediment,

previously documented for nereidid species inhabiting permanent U-shaped burrows (Henriksen et

al., 1980; Kristensen et al., 1985; Pelegri and Blackburn, 1995a). Alternately, these U-burrowing

nereidids are responsible for between 28-80% of denitrification in the sediments they inhabit (when

considered as a single macrofaunal species addition). This effect of denitrification enhancement has

been observed for: Nereis virens (Hylleberg and Henriksen, 1980; Henriksen et al., 1980, 1983;

Kristensen, 1985; Kristensen et al., 1985, 1991; Kristensen and Blackburn, 1987; Christensen et al.,

2000), N. diversicolor (Law et al., 1991; Gilbert et al., 1995, 1997; Hansen and Kristensen, 1997,

1998; Mortimer et al., 1999; Christensen et al., 2000; Heilskov and Holmer, 2001), N. succinea

(Bartoli et al., 2000) and unidentified Nereis spp. (Binnerup et al., 1992; Pelegri and Blackburn,

1995a; Meyer et al., 2001). De Roach et al. (2002; Appendix E) similarly observed enhanced

denitrification activity surrounding the burrows of adult S. aequisetis in the Swan River Estuary.

While there may be several factors accounting for the enhanced rate of denitrification within

sediments inhabited by A. ehlersi and other U-burrowing nereidids, De Roach et al. (2002;

Appendix E) summarised the likely main contributors as: (i) the presence of a burrow provides an

increased sediment-water interface, extending the surface area (and oxic-anoxic boundary) available

for microbial inhabitation and solute exchange (Hylleberg and Henriksen, 1980; Kristensen, 1984a;

Kristensen et al., 1991); (ii) the mucus binding the burrow walls may elevate the organic matter

content and resultant microbial activity of the sediment (Aller, 1983, 1988; Aller et al., 1983;

Reichardt, 1988), and/or (iii) active ventilation of a burrow may increase nitrate supply (as a

substrate for denitrification) to the sediment; furthermore, nitrification may also be enhanced by

105

Page 120: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

active ventilation (see Section 5.4.4), which provides another NOx- source for denitrification (Aller,

1988; Kristensen, 1988; Kristensen et al., 1991).

Each of these factors may have contributed to the enhancement of denitrification rate observed

within the A. ehlersi inhabited sediments of this study, but it was not possible to determine their

relative importance. The intermittent burrow ventilation pattern demonstrated by

A. ehlersi (Figure 5.3) likely facilitated enhanced coupling of nitrification-denitrification; since

enhanced nitrification heightened NOx- production (see Figure 5.5; Section 5.4.4) when the burrow

is being ventilated by oxygenated water (i.e. aerobic conditions within the burrow wall), which may

be consumed in surrounding sediment, probably at a heightened rate of denitrification during

periods of non-ventilation (i.e. when oxygen consuming processes allow more anaerobic conditions

to prevail) (Kristensen, 2000). Studies employing nitrogen isotope tracers within A. ehlersi

inhabited sediment are required to determine the partitioning of denitrification resultant either from

NOx- supplied by nitrification or by direct drawdown from the overlying water (e.g. see Pelegri and

Blackburn, 1995a; Bartoli et al., 2000).

The maximum potential denitrification rate of A. ehlersi inhabited sediments at non-limiting

ambient nitrate concentration (Vmp) was only 110% of the maximal denitrification rate of

uninhabited V-core sediments. Also, the denitrifying community associated with A. ehlersi

inhabited sediment demonstrated the lowest Michaelis-Menten constant (Kapp), indicating that

maximal denitrification was achieved at lower ambient nitrate concentration than in uninhabited, or

S. aequisetis inhabited, sediment. Thus, to some extent, the magnitude of denitrification

enhancement will depend on ambient nitrate concentration. This may help to explain why the

degree of enhancement has not previously appeared related to species. For example, N. virens has

been observed to stimulate denitrification by either the lower (138%) or upper (500%) end of the

above-stated range (see respectively, Henriksen et al., 1980 and Kristensen et al., 1985), whilst the

increases resultant from N. diversicolor addition are mid-range (177-300% stimulation; Hansen and

Kristensen, 1998). However, it is equally likely that the variation in the magnitude of

denitrification enhancement is due to differences and interrelations of the many other factors

controlling the rate of denitrification; for example, the specific increase in sediment-water interface

area provided by individual burrows, or the organic matter content and quality of the mucus-bound

sediment wall, and/or the actual pattern and magnitude of burrow irrigation that may markedly

influence the redox conditions which control absolute denitrification rate, amongst other factors. It

is probable that a combination of these and other regulatory factors determine the overall influence

of nereidids upon sedimentary denitrification, but elucidation of their relative magnitude of

importance largely waits for explicit testing.

106

Page 121: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

In contrast, the presence of juvenile S. aequisetis was found to retard denitrification in comparison

with uninhabited sediment. At in situ ambient NO3- concentration, the denitrification rate within

S. aequisetis inhabited sediment was 59% of that within uninhabited sediment and only 21% of the

rate within A. ehlersi inhabited sediment. In response to increasing ambient NO3- concentration, the

denitrifying community within S. aequisetis inhabited sediment reached a maximal denitrification

rate that was 88% of uninhabited sediment and 81% of A. ehlersi inhabited sediment, but the

maximum occurred at a much higher ambient NO3- concentration (i.e. as defined by the highest

Kapp). Although comparative studies are scarce, this net inhibitory effect of inhabitation by a

sediment reworker upon denitrification activity has been observed in marine mesocosms containing

the heart urchin Brissopsis lyrifera (Widdicombe and Austen, 1998). This species ‘bulldozes’ and

aerates the top few centimetres of sediment which, the authors argue, has the effect of suppressing

anaerobic denitrification. It is likely that the degree of oxygenation caused by the sediment

reworking habit of S. aequisetis similarly inhibited denitrification within the V-cores of this study.

Conversely, freshwater oligochaetes such as Tubifex tubifex display a habit comparable to juvenile

S. aequisetis, but have been demonstrated to enhance the denitrification rate of aquatic sediments

due to tight coupling with likewise enhanced nitrification (Chartapaul et al., 1979, 1980; Pelegri

and Blackburn, 1995b; Svensson et al., 2000). In this study, juvenile S. aequisetis did enhance

sedimentary nitrification (see Section 5.4.4 below) but the implicated increase in NOx- production

was not coupled to an increase in denitrification activity. However, it is important to consider that

the observed net reduction in denitrification activity may not occur in situ, if a deeper active zone of

denitrification naturally develops (i.e. deeper than the V-cores allowed in vitro). While not

resulting in a net decrease, suppression of denitrification activity within the surface layer of

sediments and burrow walls due to oxygenation by resident fauna has often been documented

(Sayama and Kurihara, 1983; Kristensen et al., 1985; De Roach et al., 2002). Heilskov and Holmer

(2001) have suggested that the net affect of macrofauna stimulating or inhibiting anaerobic

processes [such as denitrification] probably results from the balance between counteracting effects,

namely stimulation due to an increased supply of substrates and inhibition resulting from more

oxidised conditions.

5.4.4 Nitrification

The estimated average nitrification rate of sediment within A. ehlersi inhabited V-cores was 130%

of uninhabited sediment. The magnitude of this enhancement effect is at the lower end of the range

documented previously for N. virens (138 to 228%: Henriksen et al., 1980; Kristensen and

Blackburn, 1987; Kristensen et al., 1991) and N. diversicolor (133 to 240%: Hansen and

Kristensen, 1998). Mayer et al. (1995) similarly found that potential nitrification rates within the

107

Page 122: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

burrow walls of N. virens were 170 to 410% of that in surface sediment. Furthermore, their study

demonstrated that potential nitrification rates were stimulated considerably in the burrows walls of

11 out of 12 other infaunal species that construct definite burrow or tube structures. Factors

explaining the heightened nitrification rates of sediments containing infaunal burrows are

considered below.

The oxygen requirement of chemoautotrophic nitrifying bacteria largely limits their distribution to

aerobic layers of sediment (Henriksen et al., 1981). The presence of an infaunal burrow and

concomitant irrigation with oxygenated water considerably extends the surface sediment-water oxic

layer to depth, thus increasing the volume of sediment available for inhabitation by nitrifying

bacteria (Mayer et al., 1995). As discussed above, the ventilation rate and pattern may control both

the redox conditions and substrate supply (of NH4+ in the case of nitrification) within the burrow,

which in turn govern the process rates of aerobic/anaerobic microbial nitrogen transformations.

Significantly, Mayer et al. (1995: p. 168) suggest that “average rate of water turnover in tubes or

burrows might provide the best ‘single’ trait for predicting the availability of oxygen in tubes and

burrows and consequently for predicting species differences in the enhancement of nitrification in

these macrofaunal structures”. They further observed that those infaunal species that were strong

irrigators, i.e. with high rates of burrow ventilation (>150 mL h-1) or water turnover (>1mL min-1),

only mildly enhanced nitrification; and suggested that because higher irrigation exports NH4+ from

the sediment, enhancement of nitrification may be greatest at intermediate irrigation rates. The

results of this study lend support to the idea, since A. ehlersi exhibited a relatively high average

burrow ventilation rate (415 mL h-1) and the observed percentage enhancement of nitrification was

only at the lower end of the reported range of enhancement (see above). Nevertheless, Mayer et al.

(1995) concede that, while aspects of burrow ventilation are undoubtedly important determinants of

net nitrification rate, they are not solely responsible.

Excretion by resident fauna is a source of NH4+ that is undoubtedly of importance when considering

the enhancement of nitrification due to a heightened provision of substrate (Henriksen et al., 1983;

Welsh and Castadelli, 2004). In the present study, the rate of ammonification within A. ehlersi

inhabited V-cores was much greater than that within uninhabited sediment (Figure 5.5). While the

contribution of NH4+, either from excretion or microbial mineralisation could not be separated, it is

very likely that A. ehlersi excretion stimulated nitrification in adjacent sediment to some extent

(further see Section 5.4.6). Furthermore, intra- and inter-species variation in excretion rates may

have profound implications on the potential for nitrification within macrofaunal inhabited

sediments. Indeed, Mayer et al. (1995) found the availability of NH4+ to be correlated with

nitrification potential of sediment inhabited by the polychaete Loimia medusa. A worthwhile

108

Page 123: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

avenue of study would be to examine whether a Michaelis-Menten type relationship exists between

ambient NH4+ concentration and nitrification rate of inhabited sediments, similar to the relationship

between ambient NO3- concentration and denitrification rate documented above. As Mayer et al.

(1995) have pointed out, the existence of substrate-dependant relationships determining the rates of

both nitrification and denitrification is not surprising, but the influence of ambient substrate

concentrations have not been recognised widely when comparing studies.

Kristensen (2000) has also pointed out that other governing factors must necessarily be considered.

For example, faecal pellet deposition may enhance nitrification, as has been shown for Macoma

balthica (Henriksen et al., 1983). Nitrification rates within burrow walls have also been correlated

with the content of mucus and fine particles (Kristensen, 2000). Despite the chemoautrophic nature

of nitrifying bacteria, Kristensen (2000) suggested that their association with the often high organic

content and fine particles of burrow walls may be due to an optimal pH environment that such

particles provide. Finally, Welsh and Castadelli (2004) recently documented nitrification due to the

activity of nitrifying bacteria occurring directly on or within the bodies of various macrofaunal

species, including an estimated 15-63% increase in nitrification rate of the benthos due to nitrifying

bacteria resident upon N. succinea. This novel aspect of nitrification enhancement and the causal

mechanisms are only just beginning to be investigated. Again, elucidation of the relative

contribution of the various factors governing the enhancement of nitrification within inhabited

sediment, largely still requires explicitly directed manipulative experiments.

The enhancement of nitrification within sediment inhabited by fauna that do not construct burrows

has been less documented. In this study, the presence of juvenile S. aequisetis enhanced the

nitrification rate of sediment in V-cores to a greater extent than adult A. ehlersi; i.e. the estimated

average nitrification rate of S. aequisetis inhabited cores was 148% of that within uninhabited cores

and 113% of that within A. ehlersi inhabited cores. A similar enhancement effect has been

demonstrated for freshwater sediment reworking tubificid oligochaetes (such as T. tubifex and

Limnodrilus spp.), wherein nitrification rates of inhabited sediment have been within the range of

123-180% of defaunated sediment (Chartapaul et al., 1979, 1980; Pelegri and Blackburn, 1995b;

Svensson et al., 2001). Obviously, the disturbance caused by this type of macrofaunal habit (at

lower worm densities, see below) does not concurrently result in a disruption to the activity of the

resident nitrifier population; on the contrary, it stimulates nitrification rate. Some of the reasons for

nitrification enhancement within macrofaunal burrows may extend to reworked sediment, e.g.

increased NH4+ supply from macrofaunal excretion and perhaps enhanced nitrification directly

associated with nitrifiers on the body of macrofauna.

109

Page 124: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Furthermore, it is hypothesised that the reworking habit has a similar effect on nitrification as

infaunal burrow irrigation. The constant turnover of sediment and reduction in sediment bulk

density caused by juvenile S. aequisetis likely facilitates transport of oxygenated water at a rate and

to a depth much greater than expected by diffusion, thereby considerably enhancing the volume of

sediment available for inhabitation by nitrifiers. Svensson et al. (2000) consider both sediment

reworking and burrow irrigation as forms of biological advection, but suggest that the random

nature of solute/particle transport of the former habit and directed flow of the latter may result in

very different effects on nitrification and denitrification. Certainly, it appears that the enhancement

of aerobic nitrification but suppression of anaerobic denitrification within V-cores inhabited by

juvenile S. aequisetis may be due to the homogenisation and aeration of the entire surface 3 cm

layer of sediment. In contrast, the directed flow of water in A. ehlersi inhabited V-cores and

persistence of a lower oxygen environment in the surface sediment surrounding burrows may be the

major difference facilitating the enhancement of not only nitrification, but also denitrification.

It is important to note that the studies of freshwater tubificid oligochaetes by Pelegri and Blackburn

(1995b) and Svensson et al. (2001) found that nitrification was suppressed at very high worm

densities. The authors of both studies implicated heightened tubificid net respiratory oxygen

consumption to explain the retardation of aerobic nitrification. However, it may also be that a

higher sediment turnover rate at high worm density flushes NH4+ from the sediment (similar to the

effect of high irrigation rates; see above) or causes appreciable physical disturbance to nitrifier

populations. Indeed, Svensson et al. (2001) and other authors have found high levels of sediment

disturbance (e.g. by sieving) to impact aerobic and slow-growing nitrifier populations to a greater

extent than anaerobic organisms, such as denitrifiers. Further investigation is required into the

causal mechanisms generating variability in the effect of sediment reworkers and other types of

infaunal habit upon nitrification, denitrification and other nitrogen cycling processes.

5.4.5 Net NOx- Flux

The balance of NOx- producing (e.g. nitrification) and consuming (e.g. denitrification) processes

will determine the net NOx- flux rate of a sediment. At low ambient NO3

- concentration, the net

efflux rate of NOx- from V-core sediment in vitro was enhanced by the presence of both A. ehlersi

and juvenile S. aequisetis, due to their stimulatory effect on nitrification (Figure 5.5). Within

A. ehlersi inhabited sediment this enhancement of NOx- efflux occurred even though denitrification

was also stimulated; however, the enhancement of NOx- efflux was greater within S. aequisetis

inhabited cores because not only was nitrification stimulated to a higher extent, but the consumption

of NOx- by denitrification was retarded.

110

Page 125: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

In response to increasing ambient NO3- concentration in the overlying water, denitrification activity

increased in a Michaelis-Menten type fashion for all V-core sediment treatments, although the rate

of increase was highest in A. ehlersi inhabited sediments due to this species’ stimulatory effect on

denitrification. The heightened stimulation of denitrification caused the apparent switch from

sedimentary NOx- efflux to influx to occur at a lower ambient NO3

- concentration in A. ehlersi

inhabited sediment (i.e. at ~ 50 µM), than in uninhabited sediment (~100 µM) or S. aequisetis

inhabited sediment (~ 520 µM) (Figures 5.4[c] and 5.5). The stimulatory effect of juvenile

S. aequisetis on nitrification and concurrent inhibitory effect on denitrification entailed that NOx-

was effluxed until a very high concentration of NO3- in the overlying water (i.e. when the rate of

denitrification eventually outweighed nitrification).

Thus, although at low ambient NO3- concentration in the overlying water both species stimulated

nitrate efflux, at higher ambient NO3- concentrations the presence of A. ehlersi increased the influx

of NOx-, whilst juvenile S. aequisetis suppressed the influx of NOx

-. These observations have very

important implications when considering the effect of each species in oligotrophic (low ambient N

concentration) versus eutrophic (high ambient N) environments. That is, for the conditions in vitro,

both species would increase the NOx- available for algal uptake in the overlying water when at

lower ambient NO3- concentration; whereas, at higher ambient NO3

- concentration, the presence of

A. ehlersi in sediment would remove NOx- from the overlying water at a greater rate than in waters

overlying uninhabited sediment. Conversely, the presence of juvenile S. aequisetis at these higher

ambient NO3- concentrations would reduce the ability of the sediment to remove NOx

- from the

overlying water.

5.4.6 Ammonification and Net NH4

+ Flux

As illustrated in Figure 5.5, the ammonification rate was lowest in uninhabited sediment, higher in

S. aequisetis inhabited sediment (202% of the rate in uninhabited sediment) and highest in

A. ehlersi inhabited sediment (242% of the uninhabited rate). Whilst the excretion rates of each

species were not measured, thus disallowing segregation from the contribution made by microbial

mineralisation, uninhabited cores were obviously not subject to the former source of NH4+ supply

and it is suggested that macrofaunal excretion is a major cause of increased ammonification rates

within inhabited sediments. However, the effects of various infauna on sediment redox conditions

and resultant implications for microbial mineralisation, whether enhanced or inhibited, cannot be

discounted (for an excellent review see Kristensen, 2000). Resolution is still required of these

effects in A. ehlersi and S. aequisetis inhabited sediment.

111

Page 126: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Notwithstanding the inability to designate enhanced ammonification rates of inhabited cores either

to increased microbial mineralisation or excretion (or a combination of both), the net effect of

nitrogen cycling processes in sediment V-cores in vitro, resulted in: (i) a slight average efflux of

NH4+ from sediment in A. ehlersi inhabited cores and (ii) an influx of NH4

+ to sediment of cores

inhabited by juvenile S. aequisetis that was lower than the influx observed in uninhabited cores

(Figure 5.5). It is hypothesised that even though the nitrification rates within uninhabited sediments

were lower than in inhabited sediments, the high net NH4+ consumption of uninhabited sediment

was due the lack of infaunal excretion to provide a supplementary substrate source and thus a

concomitant dependence on diffusional NH4+ from the overlying water to supply nitrification. The

stimulation of ammonification within inhabited burrows (due either to excretion or enhanced

microbial mineralisation), and resultant in situ availability of NH4+ for nitrification, likely caused

the lower net influx of NH4+ to S. aequisetis inhabited sediment and slight efflux of NH4

+ from A.

ehlersi inhabited sediment (because ammonification was greatest in these sediments).

Most studies have similarly documented that the presence of macrofauna either enhances NH4+

efflux or suppresses NH4+ influx. The magnitude of the effect is often correlated with macrofaunal

density (e.g. Asmus, 1986; Law et al., 1991; Pelegri and Blackburn, 1995b; Lohrer et al., 2004) and

has been observed previously for macrofaunal communities in the Swan River Estuary (Pennifold

and Davis, 2001). The enhancement of sedimentary NH4+ efflux by a variety of species, both in the

dark and under light, is well illustrated by Andersen and Kristensen (1988), who examined the

effect of single species additions of: (i) the amphipod Corophium volutator (a 3-5 cm deep U-

burrower); (ii) the gastropod Hydrobia spp. (a sediment reworker); (iii) the nereidid polychaete

N. virens (a 10-40 cm deep U-burrower); and (iv) a combined community of all three species.

Compared to defaunated sediment, they found that NH4+ efflux was always increased by between

110 and 1000%. This range of NH4+ efflux enhancement by macrofauna is common to that often

reported by other researchers, who usually relate the effect either to the increased sediment-water

surface area provided by burrows, the presence of biological advection (either by irrigation or

sediment reworking) and/or increased ammonification due either to excretion or microbial

mineralisation. Whilst the extra provision of NH4+ via macrofaunal excretion is obvious, it is

worthwhile considering the study of Henriksen et al. (1983) to exemplify that other explanatory

factors may be very important. For both organic rich and poor sediments inhabited by similar

species to those examined by Andersen and Kristensen (1988; see above), NH4+ efflux rates were

equivalent to 60-270% of excretion rates. Efflux rates greater than 100% of excretion were thus

due to supplementary enhancement factors other than excretion.

112

Page 127: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Although more rarely reported, examples exist whereby the presence of macrofauna causes a

reduction in NH4+ efflux due to significant enhancement of nitrification (i.e. great enough to

outweigh any increased production of NH4+). In Long Island Sound sediments, Aller (1978)

documented this effect for the protobranch bivalve Yoldia limatula, a mobile subsurface deposit

feeder that reworks sediment to depths of about 4 cm. Within a marine sediment community off

Sweden, Enoksson and Samuelsson (1987) suggested that the presence of a brittle star (Amphiura

filiformis) can decrease the NH4+ efflux of sediment via enhanced nitrification caused both by

oxygenation of the upper layer of sediment and the ability of its calcium carbonate body structure to

buffer sediment pH to an optimal level for nitrifiers. As a final example, Caffrey and Miller (1995)

found in South San Francisco Bay that the nitrification rate within a sediment community

dominated by the tube-dwelling polychaete Asychis elongata was positively correlated with

macrofaunal biomass; and further, that net NH4+ efflux was negatively correlated with macrofaunal

biomass.

The net effect of infaunal habits and excretion rates on net sedimentary NH4+ fluxes has

implications for the regulation of the amount of NH4+ available for algal assimilation in overlying

waters. Even though the general trend documented in the literature is that the presence of fauna

generally stimulates NH4+ release from sediments, different macrofaunal species, habits and

excretion rates will undoubtedly have variant magnitudes of effect on NH4+ availability in the

overlying water. Whilst this study demonstrates the differing net effects of a mobile sediment

reworker and stationary U-burrower upon net NH4+ flux in vitro, the studies of Watson et al. (1993)

and Davey and Watson (1995) provide similar examples for in situ sediment communities of the

Tamar Estuary, England. At locations throughout the estuary, they found that the presence of

infauna resulted in effluxes of NH4+ that were 3 to 63 times greater than expected by diffusive flux.

Importantly, they observed that the degree of enhancement depended on the density and population

structure of resident macrofauna and, moreover, whether irrigating or sediment reworking fauna

predominated (with irrigators enhancing NH4+ efflux to a greater extent). The effect of different

faunal habits upon sedimentary nitrogen cycling and the availability of total dissolved inorganic

nitrogen (NOx- + NH4

+) to bloom forming phytoplankton species is further considered in Chapters 6

and 7.

5.4.7 Efficacy of the Kinetic-fix Adaptation of the Acetylene Block Method

As far as the author is aware, this is the first study in which the novel ‘kinetic-fix’ adaptation of the

C2H2 block technique (Joye et al., 1996) has been employed to examine nitrogen cycling within

sediments inhabited by macrofauna. The method is briefly re-described here to highlight its

benefits. Estimates of nitrogenous process and flux rates were determined in sediment V-cores

113

Page 128: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

separately inhabited by individual polychaetes of the two species, and compared with uninhabited

sediment. Briefly, each V-core species treatment (A. ehlersi, S. aequisetis, uninhabited) was subject

to a range of ambient NO3- concentrations, S = 0, 20, 50, 100 and 500 μM NO3

-, at which NH4+ and

NOx- fluxes (incorporating denitrification rates) were measured first without and then with C2H2

(acetylene). This kinetic-fix adaptation of the C2H2 block technique allowed for estimates of

potential denitrification rates at saturated NO3- concentration as well as for estimates of

denitrification rates at in situ NO3- concentration. The technique also allowed resolution of

sediment nitrification and ammonification rates and determination of net NOx- and NH4

+ fluxes

within the sediment of each V-core species treatment. Although appropriate corrections and

considerations were required to account for the effect of acetylene on A. ehlersi ventilation, the

technique is cheap and technically easy, yet has been demonstrated to be powerful in resolving the

effects of macrofauna upon sedimentary nitrogen cycling.

5.5 SUMMARY

The effects of polychaete inhabitation on sediment NH4+ (ammonium) and NOx

- (nitrate + nitrite)

fluxes, including the contributing denitrification, nitrification and ammonification rates, were

assessed for two species of polychaete (A. ehlersi and S. aequisetis) within the Swan River Estuary,

Western Australia. Australonereis ehlersi reside in well-defined U-shaped burrows, whereas

juvenile S. aequisetis are sediment reworkers.

For the denitrifying communities associated with uninhabited cores and those inhabited by

A. ehlersi or S. aequisetis, the apparent Michaelis constant (Kapp) and the maximum potential rate

(Vmp) are reported. As defined by resultant Michaelis-Menten curves, the magnitude of

denitrification at in situ NO3- concentration (V2μM), and across the range of experimental S, was

greatest for sediment inhabited by A. ehlersi (V2μM = 182 μmol N m-2 h-1; Vmp = 4.75 mmol N m-2

h-1; Kapp = 50 μM NO3-), moderate in uninhabited sediment (V2μM = 64 μmol N m-2 h-1; Vmp = 4.33

mmol N m-2 h-1; Kapp = 133 μM NO3-) and lowest in sediment inhabited by S. aequisetis (V2μM = 38

μmol N m-2 h-1; Vmp = 3.83 mmol N m-2 h-1; Kapp = 202 μM NO3-). Sediment nitrification rate was

generally constant over the range of S and estimated to be higher in S. aequisetis (2.76 ± 0.17 [s.e.]

mmol N m-2 h-1) than in A. ehlersi (2.43 ± 0.07 mmol N m-2 h-1) inhabited cores, and lowest in

uninhabited sediment (1.86 ± 0.13 mmol N m-2 h-1). In the absence of C2H2, the net result of

nitrification minus denitrification was such that a transition from NOx- efflux from sediment at

lower S to sedimentary NOx- uptake at higher S occurred. The point of equilibrium (zero flux)

occurred at a lower S for A. ehlersi inhabited sediment cores (52 μM) than for uninhabited sediment

cores (100 μM), with S. aequisetis inhabited cores generally effluxing NOx- over the entire range of

114

Page 129: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

S, up to 520 μM. For each V-core species treatment, ammonification in the presence of C2H2 was

generally constant over the range of S, and indicated the net rate of NH4+ production from sediment

mineralisation plus, in the case of inhabited cores, polychaete excretion (though these components

could not be separated). Estimated net ammonification was higher for A. ehlersi inhabited (2.20 ±

0.05 mmol N m-2 h-1) than S. aequisetis inhabited cores (1.84 ± 0.16 mmol N m-2 h-1), and lowest in

uninhabited cores (0.91 ± 0.05 mmol N m-2 h-1), within which polychaete excretion was absent. In

the absence of C2H2, and associated unblocking of nitrification, sediments of S. aequisetis inhabited

and uninhabited cores indicated a net NH4+ influx of -0.62 ± 0.08 mmol N m-2 h-1 and -0.85 ± 0.12

mmol N m-2 h-1, respectively, whereas A. ehlersi inhabited cores exhibited a slight average efflux of

NH4+ (0.07 ± 0.13 mmol N m-2 h-1) from the sediment.

The high resultant rates of nitrogenous flux determined in vitro may not reflect rates at natural, in

situ conditions. Further, the estimates of nitrification and kinetic denitrification parameters for V-

cores inhabited by A. ehlersi are more subjective than for the other two V-core species treatments

because of the requirement to extrapolate to ventilation-standardised NOx- flux estimates (due to

polychaete burrow ventilation being suppressed by C2H2). It is suggested that the difference in

magnitude of nitrogenous fluxes imparted by the two polychaete species is related to the influence

that their respective habits exert on the composition and activity of their associated sedimentary

microbial community. The interstitial habit of juvenile S. aequisetis is hypothesised to homogenise

and aerate sediment continually, thereby enhancing microbial nitrification and retarding anaerobic

denitrification. The permanence of A. ehlersi burrows is suggested to allow oxic/anoxic

stratification of sediment (both vertically and radially) which, combined with enhanced substrate

supply through burrow ventilation, provides for increased rates of microbial denitrification and

nitrification. Thus, due to the different habit of each nereidid species, the resultant net fluxes of

both NOx- and NH4

+ are variant (see above). The habits of constituent benthic faunal species may

therefore play a profound role in regulating the type and concentration of dissolved inorganic

nitrogen available for assimilation by bloom-forming algal species in the overlying water.

115

Page 130: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,
Page 131: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

PART III– CONCLUSIONS

117

Page 132: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

118

Page 133: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

CHAPTER 6

POLYCHAETE EFFECTS ON NITROGEN CYCLING AT THE

SEDIMENT-WATER INTERFACE: TESTING THE

FUNCTIONAL GROUP CONCEPT

6.1 INTRODUCTION To predict, reliably, the effect of each element of the benthos exerting some biogenic

modification of sediments requires basic knowledge of the species, particularly of the

processes involved in forming, maintaining and degrading micro-environments surrounding

individual organisms (Aller and Yingst, 1985). In quantifying the impact of a benthic

species on sedimentary processes associated with nitrogen cycling, for example, the habit or

life mode, life history and population biology of members of the species need to be

considered. Within a population, intraspecific variation in habit and associated differences

in sedimentary microenvironments may have marked implications for biogeochemical

processes. Such variation may result from either: (i) ontogenetic differences in habit over a

life-cycle; (ii) short-term behavioural changes of an individual’s habit; and/or (iii) variant

predisposed habits within or between populations.

Further, spatio-temporal variation and interactions of controlling environmental parameters

(organic matter supply, inorganic nitrogen availability, redox conditions, photoperiod,

hydrodynamic setting, temperature, salinity, population density/burrow spacing, for

example) may be either influential in determination of an individual’s habit, or, for a given

habit, be a source of variation in its biogeochemical effects. Thus, as part of my aim in this

study to elucidate the effect of polychaete species upon nitrogen-cycling at the sediment

water interface of the Swan River Estuary, I consider in this Chapter the influence of both

intra- and inter-specific variation in polychaete and other macrofaunal habits in the context

of their regulatory capacity of sedimentary nitrogen biogeochemistry.

6.2 CLASSIFICATION OF BENTHIC MACROFAUNAL HABITS: THE

FUNCTIONAL GROUP CONCEPT

The challenge in determining the effects of benthic fauna upon diagenetic reactions and

biogeochemical cycles, in this case of nitrogen at the sediment-water interface, is to define

appropriate working groups of animals to facilitate useful discussion instead of attempting to

reduce the analysis to consideration of individual species or even individuals. In identifying

the complex issues involved, Mermillod-Blondin et al. (2001, 2002) adopted the concept of

functional groups, a set of species that have similar effects on ecosystem processes, a

119

Page 134: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

concept developed previously by Chapin et al., 1992; Naeem et al., 1994; Tilman et al,

1997) to identify succinctly the relationships between species diversity and ecosystem

functioning. In a study of a New Zealand soft bottom community, Lohrer et al. (2004)

suggested that the loss of a functionally important spatangoid urchin resulted in decreased

sedimentary ammonium effluxes and reduced system primary productivity; thereby

instigating biogeochemical changes that broadly impacted on ecosystem functioning. Given

the present threats to benthic biodiversity due to eutrophication, habitat destruction, over-

harvesting and climate change (Tilman, 1999; Lohrer et al., 2004), identification of

functionally important groups may be critical to our understanding of consequential shifts in

aquatic ecosystem function.

In Section 6.2 I outline a preliminary framework within which guilds of benthic fauna

defined by habit, each with bioturbative influence on sedimentary nitrogen cycling, may be

tested as functional groups. I discuss the benefits and limitations of the approach in Section

6.3 and an unavoidable lack of functional specificity within each guild is emphasised (i.e. an

inability to explicitly categorise a functional group with immutable effects). Although my

primary focus throughout my thesis concerns the effect of polychaete habits upon

sedimentary nitrogen cycling, my focus is broader in this Chapter to include a wide range of

benthic fauna taxa to facilitate both testing in future and expansion of the concept to include

these groups. This kind of holistic approach is desirable given that, within a benthic faunal

community, polychaetes represent only a sub-set of all elements that impact on sediment-

based nitrogen cycling.

Traditionally, benthic faunal species have been classified into ecologically important feeding

groups based on their role in organic matter processing (Mermillod-Blondin et al., 2002).

Five categories of benthic feeding strategy have been delineated (Rhoads, 1974; Kennish,

1990):

1) suspension feeders that consume seston suspended in the water column;

2) deposit feeders that consume deposited detritus;

3) herbivores that ingest plant matter;

4) carnivore-scavengers that eat live animals or recently dead animal tissue; and

5) parasites that obtain nutrition via the fluids of living organisms.

However, an organism’s size, microhabitat and sediment preference (Day et al., 1989) are

often used as a basis of classification. Based on their size, benthic fauna may be grouped

120

Page 135: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

into macro-, meio- and microfauna: macrofauna and meiofauna are those animals that are

retained on sieves with a mesh gauge of 0.5-2 mm (e.g. many molluscs, polychaetes and

decapods) and 0.04-0.1 mm (e.g. most nematodes and harpacticoid copepods, many

turbellaria and juvenile macrofauna), respectively; microfauna pass through sieves of the

latter mesh dimensions (generally protozoans, for example foraminiferans and ciliates) (Day

et al., 1989; Kennish, 1990). Benthic animals are also categorised in two broad microhabitat

zonations: the epifauna and the infauna (Thorson, 1957). Infauna live below the sediment

surface and within the sediment fabric, either interstitially or within burrows and tubes;

epifauna live on the surface of sediments or hard substrates (Day et al., 1989; Kennish,

1990). Finally, the sediment substrate preference of benthic organisms may be used to

define groups such as soft bottom forms (preferring silty to clayey muds), sandy bottom

forms (preferring sand to sandy silt) and hard bottom forms (epifauna living on rocky

substrates) (Day et al., 1989). These secondary descriptors (size, microhabitat and substrate

preference) are useful at refining pertinent subdivisions within the primary (feeding) groups.

For example, delineation of soft bottom deposit-feeding meio-infauna or hard bottom

suspension-feeding macro-epifauna etc. is descriptive and informative of the size relations

and spatial positioning of benthic fauna within defined feeding groups. Variations in the

definition of functional feeding groups exist (Day et al. 1989), but the traditional emphasis

of group separation is on the trophic status of constituent fauna.

When considering faunal impacts on diagenetic and biogeochemical processes, Mermillod-

Blondin et al. (2002) have argued that trophic-based classification systems are more

restrictive than functional group definitions that include non-trophic activities such as

bioturbation (for example, see definitions by Lee and Swartz, 1980; Chapin et al., 1992).

The feeding mode of a benthic animal may be indicative of its impact through bioturbation

on a sediment (Welsh, 2003); nevertheless, as Aller (1978) noted previously, factors other

than feeding relations and including the taxonomic group, size, mobility, life habit, particle

size selectivity, depth and rate of feeding, and population densities in which the species

normally occurs are important factors in defining functional groups relevant to chemical

diagenesis at the sediment-water interface. Further, in recognising that the macrobenthic

faunal mode of feeding, locomotion and relation to the substrate affect sediment structure

and chemistry, Lee and Swartz (1980) developed a classification based on feeding-mobility

to group organisms in relation to bioturbation processes affecting the distribution of

pollutants in marine sediments. In considering the role of burrowing animals in organic

matter diagenesis in coastal marine sediments, Kristensen (2000) similarly noted that

functional groups may be defined by feeding type, life habit and mobility.

121

Page 136: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

In Table 6.1, I propose a classification of guilds of benthic animals relevant to their role in

biogeochemical nitrogen transformations at the sediment-water interface. It is readily

acknowledged that the validity is contentious of considering each of the initially defined

guilds of benthic animals as functional groups, with members of each guild exerting similar

effects on the nitrogenous biogeochemical processes. Hence, the rationale in defining the

subdivisions firstly is discussed, followed by consideration of the functional redundancy of

each guild.

Table 6.1: Bioturbative guilds of benthic faunal habits: A preliminary classification. (See text for details on group segregation rationale).

Habit Guild Description of Habit

Epifaunal Suspension feeding

Epifaunal habits which enhance biodeposition of organic matter/nitrogen via suspension feeding. (e.g. mussels, oysters)

Non-suspension feeding

Non-suspension feeding epifaunal habits that do not directly enhance biodeposition, e.g. epibenthic grazing, carnivory and omnivory. (e.g. mud snails, crabs, prawns)

Surface Infaunal Stationary burrow Reside in the surface oxidised layer of sediment (0-3cm depth) within

permanent or relatively stationary burrows or tubes. (e.g. amphipods, spionid polychaetes, some juvenile nereidid polychaetes)

Mobile interstitial reworker

Live interstitially or in highly transitory burrows within the surface oxidised layer of sediment (0-3cm depth). (e.g. some juvenile nereidid polychaetes, many bivalves and gastropods, sea urchins, brittle stars)

Subsurface Infaunal

Stationary burrow Capable of residing in the deeper anoxic layer of sediment (0 to >10 cm depth) within permanent or relatively stationary burrows or tubes. (e.g. nereidid polychaete U-burrowers, I- and L-shaped burrowers such as maldanid polychaetes and Arenicola marina, i.e. ‘conveyor-belt’ species).

Mobile interstitial reworker

Capable of living interstitially or in highly transitory burrows within the deeper anoxic layer of sediment (0 to >10 cm depth). (e.g. many polychaetes and bivalves, freshwater oligochaetes and isopods)

An initial major division recognises the spatial distribution: epifaunal vs infaunal habits

(Table 6.1). Regnault et al. (1988) have observed, as a gross simplification, that epibenthic

macrofauna are usually associated with ammonia regeneration but have lesser effects on

other nitrogen transforming processes, but the benthic infauna stimulate either mineralisation

or nitrification/denitrification through activity of the associated microbiota. Epifaunal habits

may be divided into those which enhance biodeposition of organic matter/nitrogen via

suspension feeding, namely epibenthic suspension-feeding; and those that do not directly

122

Page 137: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

enhance biodepostion such as epibenthic grazing, carnivory and omnivory, epibenthic non-

suspension feeding (Table 6.1).

The depth of inhabitation can be utilised further to classify the infauna into surface vs.

subsurface habits (Table 6.1). Surface fauna typically inhabit the surficial oxidised layer of

sediment (~0-3 cm deep), whereas subsurface fauna are capable of penetrating and living in

the deeper anoxic sediment (0->10 cm) by maintaining a connection with the overlying

oxygenated water and/or sediment via transitory movements or the construction of irrigated

burrows. Henriksen et al. (1983), in investigating the effect of four species of infauna on

sediment-water dissolved inorganic nitrogen fluxes within a Danish coastal sediment, found

that two shallow/surface dwellers, the amphipod Corophium volutator and the bivalve

Macoma balthica, each increased the efflux of nitrate from the sediment (likely due to a

stimulation of nitrification over denitrification). Conversely, the polychaetes Nereis virens

and Arenicola marina, both deeper sub-surface infauna, each caused an influx of nitrate into

the sediment (likely due to an enhancement of denitrification over nitrification).

A final category recognises the mobility of infaunal habits with regard to the capacity for

transport of sediment particles (sediment reworking) vs. water irrigation (Table 6.1). The

two habits are not mutually exclusive (Section 6.3), but sediment re-workers generally tend

to be more mobile and live interstitially or in highly transitory burrows; whereas, irrigating

fauna tend to reside in relatively stationary and more permanent burrows or tubes within

which water is actively or passively injected to facilitate respiration and/or suspension

feeding. The functional importance of each habit has been identified by Watson et al.

(1993) in their study on the Tamar Estuary, Cornwall or Devon, England, when they found

that enhanced fluxes (of nitrate and ammonium) resulted from the activities of benthic

macrofauna, but differences between sites depended on which habit was predominant,

irrigation or sediment reworking.

Given the paucity of information concerning the biology of polychaete species likely to

impact on their biogeochemical roles, I have assigned the polychaetes to guilds in Table 6.1

based on minimal information, primarily regarding the mobility and depth of sediment

inhabitation of each polychaete species. Thus, the classification may be undertaken with

some confidence because detailed information regarding a species habit is not required.

Most polychaete species for which information concerning their effects on nitrogen cycling

is available, are assigned to one of two broad bioturbative guilds, either subsurface

stationary burrow dwellers or subsurface mobile interstitial dwellers. Investigations

conducted have been limited to a few species only, with most belonging predominantly to

123

Page 138: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

the former guild. Of these, nereidids that construct U-shaped burrows, predominantly Nereis

virens and N. diversicolor, constitute the most commonly studied forms. A functional

distinction may be made between N. diversicolor which may irrigate burrows vigorously to

suspension-feed (given appropriate overlying phytoplankton concentration), and N. virens

which does not (Christensen et al., 2000; Section 6.3). A second sub-category of subsurface

stationary burrowers comprises polychaetes that reside within L- or I-shaped burrows, for

example Arenicola spp., the maldanids Clymenella spp. and Asychis spp., and the capitellid

Heteromastus. spp. These polychaetes and similar taxa have been described as conveyor-

belt species by Rhoads (1974) since typically they ingest sediments at the deeper end of their

burrows and deposit ingested particles as faeces at the sediment-water interface (Knox,

1977). The subsurface mobile interstitial habit for which information is available is almost

solely represented by polychaetes belonging to the Nephtys genus. Polychaetes belonging to

the surficial guilds of either stationary burrowers (some spionids, for example) or mobile

interstitial dwellers (for example Capitella spp.) no doubt constitute an important component

of the benthic fauna, however there are few studies elucidating their effect on nitrogen

cycling. Juvenile polychaetes may be important functional representatives of these surficial

groups. For example, Watson et al. (1993) suggested that a large efflux of nitrate during

winter at their study site was associated with a high abundance of juvenile N. diversicolor

that enhanced oxygenation and nitrification of surface sediments.

6.3 TESTING THE FUNCTIONAL GROUP CONCEPT

Mermillod-Blondin et al. (2001) argued that if two taxa belong to the same functional group,

then according to the redundancy model (sensu Walker, 1992), they should generate similar

effects on ecosystem processes. The six guilds of faunal habits identified above (Table 6.1)

represent distinct groups of bioturbative influence in terms of sedimentary nitrogen

transformations but they are not necessarily functional groups. The classification system of

Lee and Swartz (1980) recognises three groups of benthic organisms, epifaunal/infaunal,

deposit/suspension feeding and mobile/stationary, but based on their criteria, 12 guilds are

recognised. Further refinement (subdivision) of either system possibly could increase the

functional redundancy of each guild identified and also the number of groups. Lee and

Swartz (1980) tolerated a trade-off between maintaining functional redundancy and

minimising the number of groups for practical purposes; for example, tubiculous and non-

tubiculous groups were not separated to minimise the number of guilds. Similarly, the

classification presented here could include, for example, recognition of suspension feeding

and non-suspension feeding burrowers/irrigators; to do so may have some functional

relevance, however only the six guilds defined above are initially retained to facilitate a

more simplified consideration.

124

Page 139: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Moreover, although the functional group concept is useful for defining relevant bioturbative

guilds of benthic animals, the task of delineating ideal functional groups, i.e. those with

members of each group exerting immutable effects on biogeochemical processes at the

sediment-water interface (i.e. with absolute functional specificity), may be theoretically, but

not practically, possible. The inherent intraspecific variation in life histories, habits and

feeding modes of benthic fauna makes it difficult (if not impossible) to segregate exclusive

groups of species or individuals on such bases.

Juvenile and adult animals may exert varying effects upon nitrogen cycling. Lewin et al.

(1979) found that, although the excretion rate of ammonium is positively correlated to the

length of razor clams (Siliqua patula), the weight-specific rate of excretion (i.e. per gram of

animal) is negatively related to clam size. Therefore, a juvenile population of razor clams

will excrete a larger amount of ammonium than an adult population of the same biomass.

Similarly, juvenile cockles (Cardium edule) will excrete 93-96% of absorbed organic

nitrogen as ammonium and retain 4-7% for somatic growth while adult cockles excrete, as

ammonium, only 60-64% of absorbed organic nitrogen, retaining 36-40% for somatic

growth and the production of gametes (Loo and Rosenberg, 1989). Appositely, it is

documented in Chapter 5 that juveniles of the nereidid polychaete Simplisetia aequisetis

suppressed in vitro anaerobic sedimentary denitrification activity, likely due to the

homogenisation and aeration of sediments generated by the juvenile’s mobile interstitial

burrowing habit. On the other hand, De Roach et al. (2002) established that adult S.

aequisetis are capable of constructing permanent U-shaped burrows. Converse to the effect

of juveniles, these authors reported that actively irrigating adult specimens significantly

enhanced the denitrification activity of proximal burrow sediment in vitro. It is

hypothesised that, whilst similar to the juvenile habit, the adult creation of a permanent

burrow allows the provision of oxygen (and nitrate) at depth, it contrastingly maintains and

extends a stratified anoxic layer within the sediment in which denitrification is enhanced.

Information on age structure of a population may therefore be integral to any predictive

attempt at ascertaining a species net effect on sedimentary nitrogen cycling.

Variation in habit and resultant nitrogen cycling effects may also occur within age-specific

cohorts. In an in vitro study of the effects of bioturbation on freshwater midge larvae

(Chironomus riparius) on benthic nitrogen cycling, Stief and de Beer (2002) recognised two

depth-specific life modes: approximately 90% of larvae were surface sediment reworkers but

10% of larvae made permanent and irrigated subsurface U-shaped burrows. Sediment

ingestion was supposed to be predominant in the former habit and was hypothesised to

reduce the abundance and activity of nitrogen-transforming microbes; conversely, the

125

Page 140: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

additional sediment-water interface area for microbial colonisation provided by the latter

burrow habit, combined with an enhanced substrate (nitrate, oxygen, organic nitrogen)

supply via irrigation, was suggested to increase the abundance and activity of nitrogen

cycling micro-organisms (Stief and de Beer, 2002). As an adult, the New Zealand

spantangoid urchin (Echinocardium) may graze on epibenthic microphytobenthos,

potentially reducing oxygen production and thereby favouring anaerobic processes such as

denitrification; alternatively, it may deposit-feed, possibly reducing particulate organic

substrates and the abundance or activity of denitrifying bacteria (Lohrer et al., 2004).

Variation in behaviour associated with alternative habits by an individual may occur in

accordance with spatial or temporal differences in local environmental parameters such as

organic matter supply, inorganic nitrogen availability, redox conditions, photoperiod,

hydrodynamic setting, temperature, salinity, population density/burrow spacing, for

example. Consequently, the behavioural habit of benthic residents may change under

different environmental settings. This potential for change is well exemplified by the

polychaete N. diversicolor that inhabits and irrigates sub-surface U-shaped burrows. Under

conditions of high phytoplankton concentration in the overlying water column (~ 1-3 μg chl

a l-1), the polychaete tends towards a suspension feeding habit of vigorous irrigation, whilst

at lower algal concentrations a habit of deposit-feeding (on burrow and surficial sediment)

predominates, with less intensive respiratory burrow ventilation (Riisgård et al., 1996;

Vedel, 1998; Nithart et al., 1999; Christensen et al., 2000). For deposit feeding

N. diversicolor habituated to a low phytoplankton load, Christensen et al. (2000)

documented a three-fold increase in ammonium efflux and a halving of nitrate efflux

compared to defaunated sediment. At comparatively high phytoplankton concentration, the

polychaete’s shift to suspension feeding resulted in an almost twelve-fold stimulation of

ammonium efflux and a two-thirds increase in nitrate efflux (compared to defaunated

sediment under high phytoplankton load). The increased stimulation of ammonium efflux

and shift from a reduction to increase in nitrate efflux when suspension feeding was due

mainly to enhanced biodeposition of organic matter/nitrogen, i.e. to a rate 30-times faster

than passive sedimentation (Christensen et al., 2000). Therefore, variation in the intensity of

irrigation and related microbial provisioning of pertinent substrates markedly influences the

magnitude, direction and composition of nitrogenous fluxes at the sediment-water interface.

Another example is given by the intertidal soldier crab (Mictyris longicarpus) which, over

the period of a tidal cycle, will alternate its mode of habit between epibenthic grazing (at low

tide) and subsurface burrowing/deposit feeding (at high tide); each of the habits have explicit

and differing effects on sedimentary nitrogen cycling (Webb and Eyre, 2004a). Thus, the

representation of an ideal functional group (or indeed of a bioturbative guild with lower

126

Page 141: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

functional redundancy) is not necessarily composed of a group of species but rather of a

group of habits. A single, benthic species may display more than one habit, either

ontogenetically, due to intra-population variation in life habits, and/or via short-term

behavioural shifts, and therefore belong to more than one functional group or bioturbative

guild. Consequently, it is the habit of a benthic species, rather than the species per se, that

determines membership to a defined functional group or guild. For any given species,

variance in expression of alternate habits (life modes) of individuals and/or populations may

result in net nitrogen cycling effects that are highly dissimilar.

Further complexity is introduced from the effect of a defined functional group changing

paradoxically with variation in environmental settings (i.e. which comprise an array of

natural habitats within a constituent individual’s or population’s residing range). The habit

of benthic fauna may not necessarily change in response to local environmental variation,

but differing impacts upon sedimentary nitrogen cycling may result. Christensen et al.

(2000) provide an elegant example describing the effects of both N. diversicolor and N.

virens on sediment-water nitrogen fluxes under two ambient phytoplankton loads. In

contrast to the change in habit of N. diversicolor from deposit feeding to suspension feeding

and greater stimulation of ammonium efflux under higher phytoplankton concentration, N.

virens maintained a deposit feeding habit irrespective of phytoplankton load. However, N.

virens also stimulated ammonium efflux (compared to that from defaunated sediment) four-

fold under high, compared to two-fold stimulation at low, ambient phytoplankton

concentrations (Christensen et al., 2000). Thus, environmental factors, in this case

phytoplankton (organic matter) loading or degree of eutrophication, can affect the magnitude

of stimulation of nitrogenous effluxes by benthic fauna.

Although some allowance for differences in effect intensity may be possible in the definition

of a functional group (Mermillod-Blondin et al., 2001), variations in the composition and/or

direction of sediment-water nitrogenous fluxes, as influenced by a given benthic faunal

habit, are not permissible. Yet such deviations in the effect of habit do occur, depending on

variation in local environmental parameters. For example, following identical additions of

chironomid larvae (Chironomus plumosus) to sediments of three freshwater Swedish lakes,

Granéli (1979) found the initial effect upon sediment-water inorganic nitrogen flux to be

different for each lake (although the manipulation of chironomid numbers had little long-

term influence on the exchange): ammonium efflux increased initially in sediments of Lake

Årungen and Lake Vombsjön, but ammonium was consumed in sediments of Lake

Trummen; nitrate was initially consumed in sediments of Lake Årungen but nitrate efflux

was increased from sediments of Lake Trummen and Lake Vombsjön. Despite similarities

127

Page 142: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

in the burrowing habit of the chironomids in each lake, it was hypothesised that the variation

in direction and composition of resultant nitrogenous fluxes among the lakes resulted from

lake-specific differences in ambient conditions, particularly of redox status and alkalinity

concentration, that affected microbially mediated reactions, especially those controlling the

balance between aerobic nitrification and anaerobic denitrification (Granéli, 1979).

Similarly, diurnal shifts in redox status and inorganic nitrogen availability caused by

photosynthetic oxygen production and nitrogenous assimilation of benthic primary

producers greatly influences the net result of faunal habit effects upon nitrogen cycling. In

the dark, both ammonium and nitrate are often effluxed from sediment inhabited by benthic

fauna; whereas light-associated assimilation of each nitrogen species due to benthic

phototrophic growth typically reduces the efflux of ammonium and may cause a shift to

nitrate influx within inhabited sediments (for example by amphipods and polychaetes

(Andersen and Kristensen, 1988) and decapods (Webb and Eyre, 2004a,b)). For a given

benthic faunal habit, variation in light intensity is thus capable of altering the magnitude,

composition and/or direction of nitrogenous fluxes, both temporally (over the period of a

day) and also spatially: light intensity and periodicity will vary with depth and latitude. The

same is true for variation in redox status which will also change diurnally, as will many

other regulatory environmental factors such as population density, burrow spacing,

temperature and water movements, for example.

A change in habit effect, aside from slight changes in effect intensity, defeats the very

definition of a functional group, the constituents of which must exert similar influence on

ecosystem processes. The possibility that a defined benthic habit may exert a variable effect

on ecosystem functioning, albeit in accordance with natural variation in regulatory

environmental parameters, challenges the usefulness and even validity of the functional

group concept. The qualification must be made that a functional group of similar habits is

only relevant to a particular environmental setting, wherein the habit effect is constant.

Thus, the functional group definition must be relaxed to: A group of habits which, within a

defined environmental setting, impart similar effects on ecosystem functioning – in this case,

on nitrogen cycling. The usefulness of this refined concept in delineating practical

functional groups depends on the robustness of the habit effect in comparison to the effect

and natural variation of other regulatory environmental parameters. If the effect of the

benthic habit is a major determinant of nitrogen cycling processes, overriding or even

determining the influence of other environmental parameters, then recognition of that habit

as a functional group will have broad application in many environmental settings. However,

if variation in other environmental parameters (such as organic and inorganic nitrogenous

128

Page 143: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

substrate availability, redox status, photoperiod, for example) is independently more

important in causing change in nitrogen cycling processes, then the effect of habit may only

be constant within a restricted and defined range of those parameters within which the

functional group concept would apply.

For a given habit, the multiplicity of functional groups then depends on the extent to which

environmental variation results in habit effects that are appreciably different. As a

hypothetical example, if magnitude and quality of organic matter supply to sediments largely

determines nitrogen cycling effects, and this parameter is highly variable across a range of

habitats, then many functional groups may need to be recognised for a given habit, because

the habit effect is likely to change in accordance with variation in organic matter supply.

However, if supply of organic matter is invariable across habitats, then the effect of a

specific habit is likely to remain constant (provided there is limited influence from other

factors), and represents only one functional group. Many other environmental parameters

control the magnitude, composition and direction of nitrogenous fluxes at the sediment-

water interface, which may or may not dominate habit effects.

Resolution of the potential number of functional groups for any given habit (which

determines practicality of the concept), thus depends on a concurrent consideration of the

influence and natural variation of regulatory environmental parameters. Since these

parameters interact with each other and with the effect of habit in complex relationships, the

task is complicated. Expanding on the previous example, organic matter supply may control

the potential rate of production of remineralised ammonium. Even if supply is invariant, the

intensity and direction of ammonium flux may be altered by interactions with benthic

primary producer activity (ammonium assimilation, oxygen production) which, in turn, may

be dependent on benthic faunal habit (for example grazing vs deposit feeding) or variation in

light intensity/periodicity, amongst other factors (e.g. see Lohrer et al., 2004). The

culmination of such interactions comprises a distinct benthic environmental setting, which is

represented by a particular balance of nitrogenous fluxes.

Thus, to ascertain the potential for marked deviations in the effect of a given faunal habit

upon the balance of sedimentary nitrogen cycling (i.e. identifying the potential number of

functional groups) requires, firstly, knowledge of the degree of heterogeneity in regulatory

environmental parameters within and between benthic environmental conditions. To some

extent this depends on the scale of environmental setting being considered. When

encompassing marine, estuarine and freshwater sedimentary environments, i.e. adopting a

global view of the benthos, a defined benthic habit is expected to be subject to a large range

129

Page 144: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

of variation in parameters and their interactions that influence nitrogen cycling. Variation in

regulatory parameters is also expected within each of these broadly defined environments,

not necessarily on a smaller scale: for example (i) between near-shore and deep sea marine

sediments (subject to differences in terrestrially-derived organic matter supply, light

availability, temperature and pressure, for example); (ii) along an estuarine gradient (subject

to variation in salinity and a multitude of related and influential environmental parameters);

and (iii) between lakes and rivers subject to differing hydrological regimes and associated

effects. Indeed, considerable environmental variation is expected at even smaller scales, due

to the multitude of possible feedbacks coincident with slight variation of just one regulating

parameter. Thus, from a biogeochemical viewpoint, patchiness or heterogeneity within

benthic environments is the norm rather than the exception.

6.4 CONCLUSION

Discerning the effects of polychaete habits on sedimentary nitrogen cycling may be a very

useful test of the functional group concept in its broad application to the general effect of

habit upon ecosystem processes. Regarding high versus low usability of the functional

group concept, whether only a few or a plethora of groups need to be recognised for any

given faunal habit, the pertinent question becomes:

Is the effect of habit robust enough to remain (relatively) constant, despite variation in

environmental setting; or does such variation result in too much multiplicity of habit

effects?

In an attempt to explore the above question, either by future research or reinterpretation of

previous studies, I am suggesting that:

(i) polychaete habits initially be assigned to one of the six bioturbative guilds

defined in Table 6.1; then,

(ii) assess the variation in polychaete habit effects within each guild; and,

(iii) relate such variation to either:

(a) finer-scale differences in habit that may comprise functional groups; and/or,

(b) differences in regulatory parameters (and their interactions) which comprise

particular environmental settings, perhaps making it harder to define

functional groups (or making them specific to defined environments).

130

Page 145: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

It is not suggested a priori that the six preliminary guilds defined in Table 6.1 represent

practical functional groups, rather a lack of functional specificity is acknowledged. For

example, the subsurface stationary burrow habit as represented by N. diversicolor has been

demonstrated already to exert a multiplicity of effects depending on organic matter supply

and irrigation intensity). The relevant question of whether it is feasible to subdivide further

each bioturbative guild, either in recognition of alternate effects of a particular habit for

different environmental settings, and/or in relation to finer-scale definition of that habit to

represent functional groups, or such division would lead to too many groups for the concept

to be practicable, awaits exploration. Importantly, the bioturbative guilds defined above

offer a hierarchical starting point to examine the potential for structuring benthic faunal

habits with respect to their effects on ecosystem functions.

131

Page 146: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,
Page 147: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

CHAPTER 7

GENERAL DISCUSSION: INTEGRATING BENTHIC FAUNAL

BIOLOGY WITH SEDIMENTARY NITROGEN CYCLING

In this thesis I have addressed one aspect of the regulation of nitrogen availability to primary

producers in aquatic environments: namely, the role of the benthos and its constituent

macrofauna. I have presented information concerning the life history, geographical population

structure and production of the polychaetes A. ehlersi and S. aequisetis within the Swan River

Estuary before describing the marked and varying effects of individuals from each species on

microbial denitrification and nitrogen cycling within sediments of the estuary. Further, I have

suggested an exploratory framework by which guilds of benthic fauna, each with similar

designated habits, may be tested for predictable bioturbative influence on nitrogen cycling, i.e.

whether particular habits may be considered ‘functional groups’. In this concluding discussion

chapter I integrate my results towards defining the fine-scale effects of A. ehlersi and

S. aequisetis on microbial nitrogen cycling within a broader-scale understanding of their

population dynamics. Within this context I emphasise the functional role of benthic fauna in

estuarine nitrogen cycling and the important implications for eutrophication management.

Whilst benthic fauna have long been recognised to influence the biogeochemistry of sediments

(e.g. see reviews by Petr, 1977; Aller, 1982; Krantzberg, 1985; Kristensen, 1988), acceptance of

the pervasive importance of the benthos as an actual component of sedimentary biogeochemical

cycling has been more recent and not yet fully realised. The biogeochemical effects of well-

studied individual species are becoming known (e.g. Nereis virens and Corophium volutator;

Henriksen et al., 1983; Kristensen et al., 1985, 1991; Andersen and Kristensen, 1988,

Christensen et al., 2000) and the net effect of some benthic communities on certain processes

have been determined empirically (e.g. via the use of bulk cores or in situ benthic chambers;

Hammond et al., 1985; Enoksson and Samuelsson, 1987; Watson et al., 1993; Caffrey and

Miller, 1995; Gilbert et al., 1997); however, the integral mechanisms of variation in species

effects and ecological interactions that drive observed community biogeochemistry are less well

understood. As such, prediction of spatial and temporal differences in biogeochemical cycling

due to variation in benthic species composition has been difficult to achieve. There is a great

need to investigate both inter- and intra-specific effects of benthic fauna on sedimentary

nitrogen cycling, but this task is often problematic due to a lack of detailed information

regarding the biology of individual species and/or general community ecology. In this regard,

use of the functional group concept to classify habit effects may offer a starting framework to

explore these issues (Chapter 6), but an assessment of the basic biology and ecological niche of

a species is considered requisite to developing understanding of their biogeochemical role in the

133

Page 148: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

environment. Given the importance of sediment biogeochemistry to local and global nutrient

budgets (Barbanti et al., 1992), it is worth striving for a better understanding of the population

biology of benthic fauna and interactions with nutrient cycling.

In Table 7.1, my results describing the distribution of A. ehlersi and S. aequisetis in the Swan

River Estuary (Chapter 4) have been combined with their effects on nitrogen cycling process

and flux rates determined in vitro (Chapter 5). My intention is to illustrate the potential

magnitude of effect of the presence of each nereidid species when extrapolated and considered

at an estuary-wide scale. The scaled-up annual nitrogenous flux rates are indicative of A.

ehlersi and S. aequisetis occurring throughout their potential habitat at the experimental size and

density cited in Chapter 5 (i.e. adult A. ehlersi, 60-70 mm in length at 1,000 m-2, juvenile

S. aequisetis, 15-20 mm in total length at a density of 2,300 m-2). The total area of intertidal

flats and tidal sand banks within the Swan River Estuary is 2.46 km2 (Geoscience Australia,

2005). The estimated flux rates provided in Table 7.1 assume that this whole area comprises

potential habitat for the true euryhaline S. aequisetis, while only a portion of this (approximately

50%) potentially will be inhabited by the marine euryhaline A. ehlersi, which is restricted in its

distribution to the lower and middle estuary. Since A. ehlersi was found at 7 out of the 12 sites

examined within the estuary (Chapter 4), this is an adequate estimate of its percentage cover of

the intertidal area.

Table 7.1: In vitro nitrogen cycling processes and flux rates (tonnes N yr-1) extrapolated to an estuary-wide annual basis (see text for further explanation and assumptions).

Sediment species

treatment

Den

itrifi

catio

n at

am

bien

t* N

O3-

Max

imum

Po

tent

ial

Den

itrifi

catio

n

Nitr

ifica

tion

Net

NO

x- Flu

x*

Net

NH

4+ Flu

x

Net

Dis

solv

ed

Inor

gani

c N

itrog

en

(NO

x- + N

H4+)

Fl

ux*

Uninhabited

19 1307 561 542 -257 286

A. ehlersi

27 717 367 339 11 350

S. aequisetis

11 1156 833 822 -187 634

*At an ambient NO3- concentration of 2 µM.

It is acknowledged that the resultant flux rate estimates should not be interpreted as indicative of

conditions in situ, since experimental manipulation necessarily involves constraining

environmental factors (e.g. temperature, salinity, light availability, water quality, nereidid size

and density) such that the effect of only one to a few factors may be elucidated (in this case the

134

Page 149: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

effect of acetylene addition and manipulation of ambient NO3- concentration). Since in situ

benthic conditions of an estuary typically are subject to much spatio-temporal variation in those

artificially constrained environmental factors, it is difficult (if not impossible) to extrapolate

results of nitrogenous flux rates obtained in vitro (at a small scale and over a short-time frame)

to be reflective of the variant natural condition. Nevertheless, in the absence of requisite in situ

studies in the Swan River Estuary, my first-order attempt at gauging the potential magnitude of

the estuary-wide influence of each nereidid on nitrogenous nutrient supply (Table 7.1) is

relevant to my consideration in this thesis of the importance of benthic species in regulating

primary productivity and affecting estuarine eutrophication.

The annual allochthonous load of nitrogen entering the estuary from all tributaries has been

estimated to be within the range of 550-770 tonnes (Donohue et al., 1994). The scaled-up, net

annual efflux rate of 286 t N y-1 dissolved inorganic nitrogen (DIN = NOx- + NH4

+) from

uninhabited sediment is approximately 40-50% of the allochthonous load. Furthermore, this

autochthonous sediment regeneration of DIN estimated across the intertidal area represents only

about 10% of the estuary’s total area (Geoscience Australia, 2005). Thus, the approximate

order of magnitude of DIN efflux estimated here is consistent with the observation by Pennifold

and Davis (2001) that autochthonous benthic regeneration of nitrogen within the Swan River

Estuary may be equivalent to or greater than annual allochthonous loads. Davis (1997) has

previously suggested that the sediments of most Australian estuaries provide a larger and more

accessible source of nutrients than does runoff from catchments and, furthermore, that this

means it may be more important to control internal nutrient sources than to control catchment

sources. This is particularly true for estuarine catchments where accepted land-use practices

such as intensive agriculture and urbanisation are in direct conflict with effective terrestrial

nutrient retention. Within the Swan River Estuary, it certainly appears that management of

sedimentary nutrient fluxes and sinks may be just as, if not more, beneficial than attempting to

regulate agricultural and urban fertiliser nutrient inputs, at least in the short-term.

In this regard, an understanding of the capacity of benthic fauna to govern not only the

magnitude of total DIN flux rates, but also the type (NOx- vs NH4

+) and direction (efflux vs

influx) of sedimentary nitrogenous fluxes, may be crucial to the development of strategies

aimed at controlling pelagic nutrient concentrations and algal productivity. As already noted in

Chapter 5, the differences in the U-burrowing habit of A. ehlersi and sediment reworking habit

of juvenile S. aequisetis lead to very different net effects on sedimentary nitrogen cycling.

Under in vitro conditions at an estuary-wide scale, estimates summarised in Table 7.1

demonstrate that whilst both species increase the efflux of DIN, the stimulation by S. aequisetis

(221%) is greater than by A. ehlersi (122%). Furthermore, due to the marked enhancement of

sedimentary nitrification by S. aequisetis, NOx- is the only form of DIN effluxed and NH4

+ is

135

Page 150: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

influxed (although to a lesser extent than in uninhabited sediment). Conversely, the presence of

A. ehlersi decreases the efflux rate of NOx- from sediment (Table 7.1), mainly due to stimulation

of denitrification. Both species enhance ammonification (either due to excretion or stimulation

of microbial mineralisation; Chapter 5) and this results in a lowered influx of NH4+ within

S. aequisetis inhabited sediment and a net efflux of NH4+ from A. ehlersi inhabited sediment

(Table 7.1).

These examples of the potentially disparate effects of different benthic species habits on NOx-

and NH4+ flux rates have large implications for the assimilatory capacity of overlying

phytoplankton and other estuarine primary producers. Ammonium assimilation is less energetic

than assimilatory NO3- uptake (Appendix D; Section D.3) and NH4

+ is therefore often deemed

the preferred substrate for incorporation into algal biomass (Brezonik, 1972; Postma et al.,

1984). This generally has been confirmed within estuaries and near-shore waters whereby

comparative studies typically indicate NH4+ to be the preferred substrate (Paasche, 1988;

Middelburg and Nieuwenhuize, 2000a). However, at high ambient NO3- concentrations

significant amounts of NO3- may be assimilated into algal biomass, despite the lower preference

for this substrate (Heathwaite, 1993; Middelburg and Nieuwenhuize, 2000b). The latter point is

particularly relevant for eutrophic estuaries that are increasing in NO3- concentration due to

anthropogenic inputs (Middelburg and Nieuwenhuize, 2000b), which potentially may be further

enhanced by the stimulatory effect on nitrification and net NOx- efflux by benthic organisms

such as juvenile S. aequisetis. Whether, and to what extent, NH4+ and/or NO3

- will be utilised as

an assimilatory nitrogen source will depend not only on the enzymatic capacity of the primary

producer but also upon the spatial and temporal availability of substrate (whether NO3- or

NH4+). In this respect, the influence of benthic fauna may be of paramount importance, as

demonstrated above. Additionally, identification of the major biosynthetic nitrogen source/s for

a particular primary producer is of equal importance when attempting to regulate and manage

the occurrence and intensity of nitrogen limited algal blooms. For phytoplankton blooms within

the Swan River Estuary, Horner Rosser and Thompson (2001) found no physiological

preference for NH4+ or nitrate at any time of year, except within bottom waters during summer

whereby NH4+ was preferred, apparently in relation to the high rate of NH4

+ flux from sediments

at that time.

Net primary production within the Swan River Estuary in situ typically is highest during

summer when water temperatures, growing period and water clarity are high and flows are low

(Thompson, 1998). It is thus interesting to note that the summer phytoplankton blooms occur

during the period when runoff (allochthonous) nutrient inputs are very low and combined

available DIN concentrations (NOx- + NH4

+) are at an annual minimum due to depletion by the

preceding spring-time algal bloom (Thompson, 1998, 2001). Ammonium is the dominant

136

Page 151: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

inorganic nitrogen source from November to May in situ, and occurs in highest concentrations

near the sediment surface (Thompson, 1998). Thompson (1998, 2001) therefore suggests that

summer phytoplankton blooms rely on, and are regulated by, the supply of autochthonous

nitrogen recycled within sediments or in the water column near the sediment surface, which, as

this thesis has demonstrated, may be influenced strongly by bioturbative processes. The annual

succession of pelagic algal species in the Swan River Estuary and potential influence of

autochthonous and allochthonous NOx- and NH4

+ supplies are further discussed in Appendix B.

I have demonstrated explicitly that the role of polychaetes (and by implication other benthic

fauna) in regulating the supply, type and availability of pelagic nitrogen is a very important

factor when considering the issue of estuarine eutrophication and problem algal blooms.

Crucially, it must also be recognised that benthic regulation is only one component of a

multitude of factors governing nutrient supply and availability to water bodies. Although this

thesis necessarily adopted a reductionist approach to demonstrate the marked effect of benthic

fauna on sedimentary nitrogen cycling, much benefit is to be gained, by integrating within a

holistic context, appreciation of other component mechanisms of explanatory importance in

eutrophication (nutrient sources and transport mechanisms, pelagic biophysical controls on algal

productivity for example). An intelligent systematic consideration of each of these component

mechanisms is required to give due weight to each functionally important process.

Furthermore, only by contextualising and integrating holistically finer scale regulatory

mechanisms can we hope to understand, model reliably and/or manage the broader issue of

eutrophication and associated nuisance algal blooms.

In summation, this thesis has contributed knowledge of both (i) the basic biology and population

structure of A. ehlersi and S. aequisetis and (ii) the potential impact of each species on nitrogen

cycling within the Swan River Estuary. I have also defined a framework by which the effect of

benthic faunal habits upon sedimentary nitrogen cycling may be further explored. Future

identification of important or ‘priority’ benthic faunal species and communities, combined with

their conservation and/or ecological manipulation, may facilitate better management of

eutrophic ecosystems. However, dedicated manipulative experiments and multivariate analyses

are still required to elucidate, predict and/or model the synergistic physico-chemical and

biological factors that govern the distribution of benthic faunal species. Moreover, further study

of the biology of benthic species and ecology of benthic communities, and the governing role

they play in regulating nutrient supply and availability to pelagic algae, when incorporated with

knowledge of other important processes (e.g. hydrodynamics and algal biology), will constitute

a significant advance in the development of algal bloom management strategies.

137

Page 152: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,
Page 153: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

REFERENCES

Abdelmagid, H.M. and Tabatabai, M.A. (1987) Nitrate reductase activity of soils. Soil Biology

and Biochemistry, 19(4): 421-427.

Aller, J.Y. and Aller, R.C. (1986) Evidence for localised enhancement of biological activity

associated with tube and burrow structures in deep-sea sediments at the Hebble site,

Western North Atlantic. Deep-Sea Research, 33: 755-790.

Aller, R.C. (1978) Experimental studies of changes produced by deposit feeders on pore water,

sediment, and overlying water chemistry. American Journal of Science, 278: 1185-1234.

Aller, R.C. (1980a) Diagenetic processes near the sediment-water interface of Long Island

Sound. I. Decomposition and nutrient element geochemistry (S, N, P). Advances in

Geophysics, 22: 237-350.

Aller, R.C. (1980b) Relationships of tube-dwelling benthos with sediment and overlying water

chemistry. In: K.R. Tenore and B.C. Coull eds, Marine Benthic Dynamics. University of

South Carolina Press, Columbia, pp. 285-308.

Aller, R.C. (1980c) Quantifying solute distributions in the bioturbated zone of marine sediments

by defining an average microenvironment. Geochimica et Cosmochimica Acta, 44(12):

1955-1965.

Aller, R.C. (1982) The effects of macrobenthos on chemical properties of marine sediment and

overlying water. In: P.L. McCall and M.J.S. Tevesz eds, Animal-Sediment Relations: The

Biogenic Alteration of Sediments. Plenum Press, New York, pp. 53-102.

Aller, R.C. (1983) The importance of the diffusive permeability of animal burrow linings in

determining marine sediment chemistry. Journal of Marine Research, 41: 299-322.

Aller, R.C. (1988) Benthic fauna and biogeochemical processes in marine sediments: The role

of burrow structures. In: T.H. Blackburn and J. Sørensen eds, Nitrogen Cycling in Coastal

Marine Environments. SCOPE 33. Wiley, Chichester, pp. 301-339.

Aller, R.C., Yingst, J.Y. and Ullman, W.J. (1983) Comparative biogeochemistry of water in

intertidal Onuphis (Polychaeta) and Upogebia (Crustacea) burrows: Temporal patterns

and causes. Journal of Marine Research, 41: 571-604.

Aller, R.C. and Aller, J.Y. (1998) The effect of biogenic irrigation intensity and solute exchange

on diagenetic reaction rates in marine sediments. Journal of Marine Research, 56: 905-

936.

Aller, R.C. and Benninger, L.K. (1981) Spatial and temporal patterns of dissolved ammonium,

manganese, and silica fluxes from bottom sediments of Long Island Sound, U.S.A.

Journal of Marine Research, 39(2): 295-314.

Aller, R.C. and Yingst, J.Y. (1978) Biogeochemistry of tube-dwellings: A study of the

sedentary polychaete Amphitrite ornata (Leidy). Journal of Marine Research, 36(2): 201-

254.

139

Page 154: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Ambus, P. (1993) Control of denitrification enzyme activity in a streamside soil. FEMS

Microbiology Ecology, 102: 225-234.

American Public Health Association (1995) Standard methods for the examination of water and

wastewater, 19th Edition. APHA/AWWA/WEF.

An, S. and Gardner, W.S. (2002) Dissimilatory nitrate reduction to ammonium (DNRA) as a

nitrogen link, versus denitrification as a sink in a shallow estuary (Laguna Madre/Baffin

Bay, Texas). Marine Ecology Progress Series, 237: 41-50.

An, S. and Joye, S.B. (2001) Enhancement of coupled nitrification-denitrification by benthic

photosynthesis in shallow estuarine sediments. Limnology and Oceanography, 46(1): 62-

74.

Andersen, F.Ø. and Kristensen, E. (1988) The influence of macrofauna on estuarine benthic

community metabolism: A microcosm study. Marine Biology, 99: 591-603.

Andersen, F.Ø. and Kristensen, E. (1991) Effects of burrowing macrofauna on organic matter

decomposition in coastal marine sediments. In: P.S. Meadows and A. Meadows eds, The

Environmental Impact of Burrowing Animals and Animal Burrows. Symposia of the

Zoological Society of London, Number 63. Clarendon Press, pp. 69-88.

Andersen, J.M. (1977) Rates of denitrification of undisturbed sediment from six lakes as a

function of nitrate concentration, oxygen and temperature. Archives of Hydrobiology, 80:

147-159.

Andersen, T.K., Jensen, M.H. and Sørensen, J. (1984) Diurnal variation of nitrogen cycling in

coastal, marine sediments I. Denitrification. Marine Biology, 83: 171-176.

Asmus, R. (1986) Nutrient flux in short-term enclosures of intertidal sand communities.

Ophelia, 26: 1-18.

Atkins, R.P. and Klemm, V.V. (1987) The effects of discharges, effluent and urbanisation on

the Swan River. In: J. John ed., Swan River Estuary, Ecology and Management. Curtin

University Environmental Studies Group Report, 1: 296-313.

Atkins, R.P., Rose, T., Brown, R.S. and Robb, M. (2001) The Microcystis cyanobacteria bloom

in the Swan River – February 2000. Water Science and Technology, 43(9): 107-114.

Augener, H. (1913) Polychaeta I: Errantia. In: W. Michaelsen and R. Hartmeyer eds, Die Fauna

Südwest-Australiens: Ergebnisse der Hamburger südwest-australischen Forschungsreise

1905. Verlag von Gustav Fischer in Jena, pp. 63-304 [In German].

Bakken, T. and Wilson, R.S. (2005) Phylogeny of nereidids (Polychaeta, Nereididae) with

paragnaths. Zoologica Scripta, 34(5): 507-547.

Balderston, W.L., Sherr, B. and Payne, W.J. (1976) Blockage by acetylene of nitrous oxide

reduction in Pseudomonas perfectomarinus. Applied and Environmental Microbiology,

31(4): 504-508.

140

Page 155: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Barbanti, A., Ceccherelli, V.U., Frascari, F., Rosso, G. and Reggiani, G. (1992) Nutrient release

from sediments and the role of bioturbation in the Goro Lagoon (Italy). In: R.A.

Vollenweider, R. Marchetti and R. Viviani eds, Marine Coastal Eutrophication. Elsevier,

Amsterdam, pp. 475-487.

Barnes, H. (1957) Nutrient elements. In: J.W. Hedgpeth ed., Treatise on Marine Ecology and

Paleoecology. Volume 1. Ecology. The Geological Society of America, Memoir 67.

Waverly Press, Baltimore, pp. 297-344.

Bartoli, M., Nizzoli, D., Welsh, D.T. and Viaroli, P. (2000) Short-term influence of

recolonisation by the polychaete worm Nereis succinea on oxygen and nitrogen fluxes

and denitrification: A microcosm study. Hydrobiologia, 431: 165-174.

Berner, R.A. (1976) The benthic boundary layer from the viewpoint of a geochemist. In: I.N.

McCave ed., The Benthic Boundary Layer. Plenum Press, New York, pp. 33-55.

Berner, R.A. (1980) Early Diagenesis: A Theoretical Approach. Princeton University Press,

New Jersey, 241pp.

Betlach, M.R. and Tiedje, J.M. (1981) Kinetic explanation for accumulation of nitrite, nitric

oxide, and nitrous oxide during bacterial denitrification. Applied and Environmental

Microbiology, 42(6): 1074-1084.

Bhattacharya, C.G. (1967) A simple method of resolution of a distribution into Gaussian

components. Biometrics, 23: 115-135.

Billen, G. (1978) A budget of nitrogen recycling in North Sea sediments off the Belgian coast.

Estuarine, Coastal and Shelf Science, 7: 127-146.

Binnerup, S.J., Jensen, K., Revsbech, N.P., Jensen, M.H. and Sørensen, J. (1992)

Denitrification, dissimilatory reduction of nitrate to ammonium, and nitrification in a

bioturbated estuarine sediment as measured with 15N and microsensor techniques. Applied

and Environmental Microbiology, 58(1): 303-313.

Blackburn, S., Bolch, C., Jones, G., Negri, A. and Orr, P. (1997) Cyanobacterial blooms: Why

are they toxic? (The regulation of toxin production and persistence). In: J.R. Davis ed.,

Managing Algal Blooms: Outcomes from CSIRO’s Multi-divisional Blue-green Algal

Program. CSIRO Land and Water, Canberra, pp. 15-30.

Blackburn, T.H. (1986) Nitrogen cycle in marine sediments. Ophelia, 26: 65-76.

Blackburn, T.H. and Sørensen, J. (1988) Nitrogen Cycling in Coastal Marine Environments.

SCOPE 33. Wiley, Chichester, 451 pp.

Bonin, P. and Raymond, N. (1990) Effects of oxygen on denitrification in marine sediments.

Hydrobiologia, 207: 115-122.

Bonin, P., Gilewicz, M. and Bertrand, J.C. (1989) Effects of oxygen on each step of

denitrification in Pseudomonas nautica. Canadian Journal of Microbiology, 35: 1061-

1064.

141

Page 156: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Bonin, P., Tamburini, C. and Michotey, V. (2002) Determination of the bacterial processes

which are sources of nitrous oxide production in marine samples. Water Research, 36:

722-732.

Bordreau, B.P. (1984) On the equivalence of nonlocal and radial-diffusion models for porewater

irrigation. Journal of Marine Research, 42: 731-735.

Bothe, H., Jost, G., Schloter, M., Ward, B.B. and Witzel, K.-P. (2000) Molecular analysis of

ammonia oxidation and denitrification in natural environments. FEMS Microbiology

Reviews, 24: 673-690.

Bowman, R.A. and Focht, D.D. (1974) The influence of glucose and nitrate concentrations upon

denitrification rates in sandy soils. Soil Biology and Biochemistry, 6: 297-301.

Boynton, W.R., Kemp, W.M. and Osborne, C.G. (1980) Nutrient fluxes across the sediment-

water interface in the turbid zone of a coastal plain estuary. In: V.S. Kennedy ed.,

Estuarine Perspectives. Academic Press, New York, pp. 93-109.

Brezonik, P.L. (1972) Nitrogen: Sources and transformations in natural waters. In: H.E. Allen

and J.R. Kramer eds, Nutrients in Natural Waters. Wiley, New York, pp. 1-50.

Brown, C.M., Macdonald-Brown, D.S. and Stanley, S.O. (1975) Inorganic nitrogen metabolism

in marine bacteria: Nitrate uptake and reduction in a marine pseudomonad. Marine

Biology, 31: 7-13.

Buresh, R.J. and Patrick Jr., W.H. (1981) Nitrate reduction to ammonium and organic nitrogen

in an estuarine sediment. Soil Biology and Biochemistry, 13(4): 279-283.

Cabrita, M.T. and Brotas, V. (2000) Seasonal variation in denitrification and dissolved nitrogen

fluxes in intertidal sediments of the Tagus estuary, Portugal. Marine Ecology Progress

Series, 202: 51-65.

Caffrey, J.M. and Miller, L.G. (1995) A comparison of two nitrification inhibitors used to

measure nitrification rates in estuarine sediments. FEMS Microbiology Ecology, 17: 213-

220.

Callender, E. and Hammond, D.E. (1982) Nutrient exchange across the sediment-water interface

in the Potomac River Estuary. Estuarine, Coastal and Shelf Science, 15: 395-413.

Cammen, L.M. (1980) Ingestion rate: An empirical model for aquatic deposit feeders and

detritivores. Oecologia, 44: 303-310.

Capone, D.G. (1988) Benthic nitrogen fixation. In: T.H. Blackburn and J. Sørensen eds,

Nitrogen Cycling in Coastal Marine Environments. SCOPE 33. Wiley, Chichester, pp. 85-

123.

Capone, D.G. (2000) The marine microbial nitrogen cycle. In: D.L. Kirchman ed., Microbial

Ecology of the Oceans. Wiley, New York, pp. 455-493.

Capone, D.G. (2002) Microbial nitrogen cycling. In: C.J. Hurst ed., Manual of Environmental

Microbiology (2nd Edition), Section IV: Aquatic Environments. American Society for

Microbiology Press, Washington D.C., pp. 439-449.

142

Page 157: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Capone, D.G. and Taylor, B.F. (1980) Microbial nitrogen cycling in a seagrass community. In:

V.S. Kennedy ed., Estuarine Perspectives. Academic Press, New York, pp. 153-161.

Carpenter, E.J. and Capone, D.G. (1983) Nitrogen in the Marine Environment. Academic Press,

New York, 900 pp.

Chalmer, P.N. and Scott, J.K. (1984) Fish and benthic faunal surveys of the Leschenault and

Peel-Harvey estuarine systems of south-western Australia in December 1974. Department

of Conservation and Environment Bulletin (Western Australia), 149: 40pp.

Chalmer, P.N., Hodgkin, E.P. and Kendrick, G.W. (1976) Benthic faunal changes in a seasonal

estuary of south-western Australia. Records of the Western Australian Museum, 4(4): 383-

410.

Chapin III, F.S., Schulze, E.-D. and Mooney, H.A. (1992) Biodiversity and ecosystem

processes. Trends in Ecology and Evolution, 7(4): 107-108.

Chartapaul, L., Robinson, J.B. and Kaushik, N.K. (1979) Observations on the role of tubificid

worms in nitrogen transformations in stream sediment. Journal of the Fisheries Research

Board of Canada, 36: 673-678.

Chartapaul, L., Robinson, J.B. and Kaushik, N.K. (1980) Effects of tubificid worms on

denitrification and nitrification in stream sediment. Canadian Journal of Fisheries and

Aquatic Science, 37: 656-663.

Choi, J. and Lee, J. (1997) Secondary production of a nereid species, Perenereis aibuhitensis in

the intertidal mudflat of the west coast of Korea. Bulletin of Marine Science, 60: 517-528.

Christensen, B., Vedel, A. and Kristensen, E. (2000) Carbon and nitrogen fluxes in sediment

inhabited by suspension-feeding (Nereis diversicolor) and non-suspension feeding (N.

virens) polychaetes. Marine Ecology Progress Series, 192: 203-217.

Christensen, J.P., Smethie Jr., W.M. and Devol, A.H. (1987) Benthic nutrient regeneration and

denitrification on the Washington continental shelf. Deep-Sea Research, 34(5/6): 1027-

1047.

Christensen, P.B., Nielsen L.P., Revsbech, N.P. and Sørensen, J. (1989) Microzonation of

denitrification activity in stream sediments as studied with a combined oxygen and nitrous

oxide microsensor. Applied and Environmental Microbiology, 55(5): 1234-1241.

Christensen, S. and Tiedje, J.M. (1988) Sub-parts-per-billion nitrate method: Use of an N2O-

producing denitrifier to convert NO3- or 15NO3

- to N2O. Applied and Environmental

Microbiology, 54(6): 1409-1413.

Cloern, J.E. (1982) Does the benthos control phytoplankton biomass in South San Francisco

Bay? Marine Ecology Progress Series, 9: 191-202.

Cloern, J.E. (2001) Our evolving conceptual model of the coastal eutrophication problem.

Marine Ecology Progress Series, 210: 223-253.

143

Page 158: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Cole, J.A. (1988) Assimilatory and dissimilatory reduction of nitrate to ammonia. In: J.A. Cole

and S.J. Ferguson eds, The Nitrogen and Sulphur Cycles, 42nd Symposium of the Society

for General Microbiology. Cambridge University Press, Cambridge, pp. 282-329.

Cole, J.A. (1990) Physiology, biochemistry and genetics of nitrate dissimilation to ammonia. In:

N.P. Revsbech and J. Sørensen eds, Denitrification in Soil and Sediment. Plenum Press,

New York, pp. 57-76.

Colliver, B.B. and Stephenson, T. (2000) Production of nitrogen oxide and dinitrogen oxide by

autotrophic nitrifiers. Biotechnology Advances, 18: 219-232.

Colney, D.J. (2000) Biogeochemical nutrient cycles and nutrient management strategies.

Hydrobiologia, 410: 87-96.

Connel, E.L. and Walker, D. I. (2001) Nutrient cycling associated with the seagrass Halophila

ovalis in the Swan-Canning Estuary based on seasonal variations in biomass and tissue

nutrients. Hydrological Processes, 15: 2401-2409.

Copeland, R.A. (1996) Enzymes: A Practical Introduction to Structure, Mechanism, and Data

Analysis. Wiley-VCH, New York.

Crisp, D.J. (1971) Energy flow measurements. In: N.A. Holme and A.D. McIntyre eds, Methods

for the Study of the Marine Benthos. Blackwell Scientific, Oxford, pp. 197-279.

Dalsgaard, T. and Thamdrup, B. (2002) Factors controlling anaerobic ammonium oxidation

with nitrite in marine sediments. Applied and Environmental Microbiology, 68(8): 3802-

3808.

Dalsgaard, T., Canfield, D.E., Peteresen, J., Thamdrup, B. and Acuña-González, J. (2003) N2

production by the anammox reaction in the anoxic water column of the Golfo Dulce,

Costa Rica. Nature, 422: 606-608.

Dame, R., Zingmark, R., Stevenson, H. and Nelson, D. (1980) Filter feeder coupling between

the estuarine water column and benthic subsystems. In: V.S. Kennedy ed., Estuarine

Perspectives. Academic Press, New York, pp. 93-109.

Darwin, C. (1881) The Formation of Vegetable Mould Through the Action of Worms with

Observations on Their Habits. Murray, London. (Reprinted in 1948 as Darwin on Humus

and the Earthworm. Faber and Faber Ltd, London, 153pp.)

Davey, J.T. and Watson, P.G. (1995) The activity of Nereis diversicolor (Polychaeta) and its

impact on nutrient fluxes in estuarine waters. Ophelia, 41: 57-70.

Davies, J.M. (1975) Energy flow through the benthos in a Scottish sea loch. Marine Biology,

31: 353-362.

Davis, J.R. (1997) The CSIRO multi-divisional blue-green algal program: Outcomes and

successes [Introduction]. In: J.R. Davis ed., Managing Algal Blooms: Outcomes from

CSIRO’s Multi-divisional Blue-green Algal Program. CSIRO Land and Water, Canberra,

pp. vii-x.

144

Page 159: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Davison, C. (1891) On the amount of sand brought up by lobworms to the surface. Geological

Magazine, 8: 489-493.

Day, Jr., J.W., Hall, C.A.S., Kemp, W.M. and Yáñez-Arancibia, A. (1989) Estuarine Ecology.

Wiley, New York, 558 pp.

de Bie, M.J.M., Middelburg, J.J., Starink, M. and Laanbroek, H.J. (2002) Factors controlling

nitrous oxide at the microbial community and estuarine scale. Marine Ecology Progress

Series, 240: 1-9.

de Jonge, V.N., Elliott, M. and Orive, E. (2002) Causes, historical development, effects and

future challenges of a common environmental problem: Eutrophication. Hydrobiologia,

475/476: 1-19.

De Roach, R.J., Rate, A.W., Knott, B. and Davies, P.M. (2002) Denitrification activity in

sediment surrounding polychaete (Ceratonereis aequisetis) burrows. Marine and

Freshwater Research, 53: 35-41.

de Wilde, H.P.J. and de Bie, M.J.M. (2000) Nitrous oxide in the Schelde estuary: Production by

nitrification and emission to the atmosphere. Marine Chemistry, 69: 203-216.

Dettmann, E.H. (2001) Effect of water residence time on annual export and denitrification of

nitrogen in estuaries: A model analysis. Estuaries, 24(4): 481-490.

Devol, A.H. (2003) Solution to a marine mystery. Nature, 422: 575-576.

Dong, L.F., Thornton, D.C.O., Nedwell, D.B. and Underwood, G.J.C. (2000) Denitrification in

sediments of the River Colne estuary, England. Marine Ecology Progress Series, 203:

109-122.

Donohue, R., Deeley, D.M., Robb, M.R. and Young, L. (1994) Nutrient loads for the Swan-

Canning estuarine catchment 1987-1992. Swan River Trust Report [Perth], 20: 61 pp.

Donohue, R., Davidson, W.A., Peters, N.E., Nelson, S. and Jakowyna, B. (2001) Trends in total

phosphorous and total nitrogen concentrations of tributaries to the Swan-Canning Estuary,

1987 to 1998. Hydrological Processes, 15: 2411-2434.

Dorsey, J.H. (1981) The ecology of Australonereis ehlersi (Augener, 1913) and Ceratonereis

erythraeensis (Fauvel, 1919) (Polychaeta, Nereidae) living offshore from the Werribee

sewage-treatment farm, Port Phillip Bay, Victoria, Australia. Ph.D. Thesis, University of

Melbourne, Victoria.

Dorsey, J.H. (1982) Intertidal community offshore from the Werribee sewage-treatment farm:

An opportunistic infaunal assemblage. Australian Journal of Marine and Freshwater

Research, 33: 45-54.

Douglas, G., Hamilton, D., Gerritse, R., Adeney, J. and Coad, D. (1997) Sediment

geochemistry, nutrient fluxes and water quality in the Swan River Estuary, Western

Australia. In: J.R. Davis ed., Managing Algal Blooms: Outcomes from CSIRO’s Multi-

divisional Blue-green Algal Program. CSIRO Land and Water, Canberra, pp. 15-30.

145

Page 160: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Dowd, J.E. and Riggs, D.S. (1965) A comparison of estimates of Michaelis-Menten kinetic

constants from various linear transformations. The Journal of Biological Chemistry,

240(2): 863-869.

Dürr, V. and Semeniuk, T.A. (2000) Long-term spatial dynamics of polychaetes in Leschenault

Inlet estuary. Journal of the Royal Society of Western Australia, 83: 463-474.

Edgar, G. (1999) A Classification of Tasmanian Estuaries and Assessment of their Conservation

Significance using Ecological and Physical Attributes, Population and Land Use. Marine

Research Laboratories, Tasmanian Aquaculture and Fisheries Institute, University of

Tasmania, Australia.

Edwards, R.M. and Tiedje, J.M. (1981) Determination of Km for dissimilatory nitrate reduction

by cells of Pseudomonas fluorescens. Abstracts of the Annual Meeting of the American

Society for Microbiology, 1981(N48): 181.

Elliott, M. and de Jonge, V.N. (2002) The management of nutrients and potential eutrophication

in estuaries and other restricted water bodies. Hydrobiologia, 475/476: 513-524.

Emerson, S., Jahnke, R. and Heggie, D. (1984) Sediment-water exchange in shallow water

estuarine sediments. Journal of Marine Research, 42: 709-730.

Enoksson, V. and Samuelsson, M.-O. (1987) Nitrification and dissimilatory ammonium

production and their effects on nitrogen flux over the sediment-water interface in

bioturbated coastal sediments. Marine Ecology Progress Series, 36: 181-189.

Eppley, R.W. and Peterson, B.J. (1979) Particulate organic matter flux and planktonic new

production in the deep ocean. Nature, 282: 677-680.

Esteves, J.L., Mille, G., Blanc, F. and Bertrand, J.C. (1986) Nitrate reduction activity in a

continuous flow-through system in marine sediments. Microbial Ecology, 12: 283-290.

Eyre, B.D. and Ferguson, A.J.P. (2002) Comparison of carbon production and decomposition,

benthic nutrient fluxes and denitrification in seagrass, phytoplankton, benthic microalgae-

and macroalgae-dominated warm-temperate Australian lagoons. Marine Ecology

Progress Series, 229: 43-59.

Eyre, B.D. and McKee, L.J. (2002) Carbon, nitrogen, and phosphorous budgets for a shallow

subtropical coastal embayment (Moreton Bay, Australia). Limnology and Oceanography,

47(4): 1043-1055.

Falkowski, P.G. (1983) Enzymology of nitrogen assimilation. In: E.J. Carpenter and D.G.

Capone eds, Nitrogen in the Marine Environment. Academic Press, New York, pp. 839-

868.

FAO-ICLARM (2005) FiSAT II: Stock Assessment Tools [software version 1.2.0]. Food and

Agriculture Organisation of the United Nations and the International Center for Living

Aquatic Resources, Rome, Italy.

Firestone, M.K. (1982) Biological denitrification. In: F.J. Stevenson ed., Nitrogen in

Agricultural Soils. Agronomy Monographs, 22: 289-363.

146

Page 161: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

García-Ruiz, R., Pattinson, S.N. and Whitton, B.A. (1998) Kinetic parameters of denitrification

in a river continuum. Applied and Environmental Microbiology, 64(7): 2533-2538.

Gardner, W.S., Yang, L., Cotner, J.B., Johengen, T.H. and Lavrentyev, P.J. (2001) Nitrogen

dynamics in sandy freshwater sediments (Saginaw Bay, Lake Huron). Journal of Great

Lakes Research, 27(1): 84-97.

Gayanilo Jr., F.C., Sparre, P. and Pauly, D. (2002) FiSAT II Online User’s Guide. <http://

www.fao.org/fi/statist/fisoft/fisat/WebHelp/FiSAT2.htm>, Accessed May 2005.

Geoscience Australia (2005) OzEstuaries – Information about Australia’s estuaries and coasts.

Estuary Database Query Module. Accessed from <http://www.ozestuaries.org/>, October

2005. Geoscience Australia, Australian Government, Canberra.

Gerhart, D.Z. and Likens, G.E. (1975) Enrichment experiments for determining nutrient

limitation: Four methods compared. Limnology and Oceanography, 20: 649-653.

Gilbert, F., Bonin, P. and Stora, G. (1995) Effect of bioturbation on denitrification in a marine

sediment from the West Mediterranean littoral. Hydrobiologia, 304: 49-58.

Gilbert, F., Stora, G., Bonin, P., LeDreau, Y., Mille, G. and Bertrand, J.-C. (1997) Hydrocarbon

influence on denitrification in bioturbated Mediterranean coastal sediments.

Hydrobiologia, 345: 67-77.

Glasby, C.J. (1984) The Population Structure and Reproductive Biology of Ceratonereis

limnetica (Polychaeta: Nereidida) at Lower Portland, Hawkesbury River. M.Sc. Thesis,

University of Sydney, New South Wales, Australia.

Glasby, C.J. (1986) Population structure and reproductive biology of Ceratonereis limnetica

(Polychaeta: Nereididae) at Lower Portland, Hawkesbury River, Australia. Marine

Biology, 90: 589-595.

Glasby, C.J., Hutchings, P.A. and Wilson, R.S. (2000) Class Polychaeta: Biogeography. In: P.L.

Beesley, G.J.B. Ross and C.J. Glasby, eds, Polychaetes and Allies: The Southern

Synthesis. Fauna of Australia, Volume 4a: Polychaeta, Myzostomida, Pogonophora,

Echiura, Sipuncula. CSIRO Publishing, Melbourne, pp. 39-43.

Goeyens, L., De Vries, R.T.P., Bakker, J.F. and Helder, W. (1987) An experiment on the

relative importance of denitrification, nitrate reduction and ammonification in coastal

marine sediment. Netherlands Journal of Sea Research, 21(3): 171-175.

Goldberg, F.D., and Koide, M. (1962) Geochronological studies of deep sea sediments by

ionium/thorium methods. Geochimica et Cosmochimica Acta, 26: 417-450.

Goldhaber, M.B., Aller, R.C., Cochran, J.K., Rosenfeld, J.K., Martens, C.S. and Berner, R.A.

(1977) Sulfate reduction, diffusion, and bioturbation in Long Island Sound sediments:

Report of the FOAM Group. American Journal of Science, 277: 193-237.

Goldman, J.C., Tenore, K.R. and Stanley, H.I. (1973) Inorganic nitrogen removal from

wastewater: Effect on phytoplankton growth in coastal marine waters. Science, 180: 955-

956.

147

Page 162: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Granéli, W. (1979) The influence of Chironomus plumosus larvae on the exchange of dissolved

substances between sediment and water. Hydrobiologia, 66(2): 149-159.

Grant, A. (1989) The use of graphical methods to estimate demographic parameters. Journal of

the Marine Biological Association of the United Kingdom, 69: 367-371.

Gray, J.S. (1974) Animal-sediment relationships. Oceanography and Marine Biology Annual

Review, 12: 223-261.

Griffin, S.L., Herzfeld, M., Hamilton, D.P. (2001) Modelling the impact of zooplankton grazing

on phytoplankton biomass during a dinoflagellate bloom in the Swan River Estuary,

Western Australia. Ecological Engineering, 16: 373-394.

Groffman, P.M., Holland, E.A., Myrold, D.D., Robertson, G.P. and Zou, X. (1999)

Denitrification. In: G.P. Robertson ed., Standard Soil Methods for Long-Term Ecological

Research. Oxford University Press, New York, pp. 272-288.

Grube, A.-E. (1878) Annulata Semperiana. Beiträge zur kenntniss der annelidenfauna der

Phillipinen nach den von Herrn Prof. Semper mitgebrachten sammlungen. Mémoires

l’Académie Impériale des Sciences de St.-Petersbourg, serie 7, 25: ix, 1-300, pls. 1-15.

Grundmanis, V. and Murray, J.W. (1977) Nitrification and denitrification in marine sediments

from Puget Sound. Limnology and Oceanography, 22(5): 804-813.

Hale, S.S. (1974) The role of benthic communities in the nutrient cycles of Narragansett Bay.

M.Sc. Thesis, University of Rhode Island, Kingston, Rhode Island, 123 pp.

Hamilton, D.P. and Turner, J.V. (2001) Integrating research and management for an urban

estuarine system: The Swan-Canning Estuary, Western Australia. Hydrological

Processes, 15: 2383-2385.

Hamilton, R.D. (1964) Photochemical processes in the inorganic nitrogen cycle of the sea.

Limnology and Oceanography, 9: 107-111.

Hammond, D.E., Fuller, C., Harmon, D., Hartman, B., Korosec, M., Miller, L.G., Rea, R.,

Warren, S., Berelson, W. and Hager, S.W. (1985) Benthic fluxes in San Francisco Bay.

Hydrobiologia, 129: 69-90.

Hansen, K. and Kristensen, E. (1997) Impact of macrofaunal recolonization on benthic

metabolism and nutrient fluxes in a shallow marine sediment previously overgrown with

macroalgal mats. Estuarine, Coastal and Shelf Science, 45: 613-628.

Hansen, K. and Kristensen, E. (1998) The impact of the polychaete Nereis diversicolor and

enrichment with macroalgal (Chaetomorpha linum) detritus on benthic metabolism and

nutrient dynamics in organic-poor and organic-rich sediment. Journal of Experimental

Marine Biology and Ecology, 231: 201-223.

Hargrave, B.T. (1973) Coupling carbon flow through some pelagic and benthic communities.

Journal of the Fisheries Research Board of Canada, 30: 1317-1326.

148

Page 163: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Harris, G.P. (2001) Biogeochemistry of nitrogen and phosphorous in Australian catchments,

rivers and estuaries: Effects of land use and flow regulation and comparisons with global

patterns. Marine and Freshwater Research, 52: 139-49.

Harris, G.P., Batley, G., Fox, D., Hall, D., Jernakoff, P., Molloy, R., Murray, A., Newell, B.,

Parslow, J., Skyring, G. and Walker, S. (1996) Port Phillip Bay Environmental Study

Final Report. CSIRO Publishing, Canberra, 248 pp .

Hartman, O. (1954) Australian Nereidae including descriptions of three new species and one

genus, together with summaries of previous records and keys to species. Transactions of

the Royal Society of South Australia, 77: 1-41.

Hartmann-Schröder, G. (1979) Teil 2. Die Polychaeten der tropischen Nordwestküste

Australiens (zwischen Derby im Norden und Port Hedland im Süden). In: G. Hartmann-

Schröder and O. Hartmann eds, Zur Kenntnis des Eulitorals der australischen Küsten

unter besonderer Berücksichtigung der Polychaeten und Ostracoden (Teil 2 und Teil 3).

Mitteilungen aus dem Hamburgischen Zoologischen Museum und Institut 76: 75-218.

Hartmann-Schröder, G. (1982) Teil 8. Die Polychaeten der subtropisch-antiborealean Westküste

Australiens (zwischen Cervantes im Norden und Cape Naturaliste im Süden). In: G.

Hartmann-Schröder and O. Hartmann eds, Zur Kenntnis des Eulitorals der australischen

Küsten unter besonderer Berücksichtigung der Polychaeten und Ostracoden. Mitteilungen

aus dem Hamburgischen Zoologischen Museum und Institut 79: 51-118.

Hartmann-Schröder, G. (1985) Teil 11. Die Polychaeten der antiborealean Sudküste Australiens

(zwischen Port Lincoln im Westen und Port Augusta im Osten). In: G. Hartmann-

Schröder and O. Hartmann eds, Zur Kenntnis des Eulitorals der australischen Küsten

unter besonderer Berücksichtigung der Polychaeten und Ostracoden. Mitteilungen aus

dem Hamburgischen Zoologischen Museum und Institut 82: 61-99.

Hartwig, E.O. (1974) Physical, chemical and biological aspects of nutrient exchange between

the marine benthos and the overlying water. Ph.D. Thesis, University of California, San

Diego and Scripps Institute of Oceanography, La Jolla, California, 174 pp.

Hartwig, E.O. (1976) The impact of nitrogen and phosphorous release from a siliceous sediment

on the overlying water. In: M. Wiley ed., Estuarine Processes. Volume 1: Uses, Stresses

and Adaptation to the Estuary. Academic Press, New York, pp. 103-117.

Hasselblad, V. (1966) Estimation of parameters for a mixture of normal distributions.

Technometrics, 8: 421-444.

Hattori, A. (1983) Denitrification and dissimilatory nitrate reduction. In: E.J. Carpenter and

D.G. Capone eds, Nitrogen in the Marine Environment. Academic Press, New York, pp.

191-232.

Heathwaite, A.L. (1993) Nitrogen cycling in surface waters and lakes. In: T.P. Burt, A.L.

Heathwaite and S.T. Trudgill, eds, Nitrate: Processes, Patterns and Management. Wiley,

Chichester, pp. 99-140.

149

Page 164: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Heathwaite, A.L., Johnes, P.J., Peters, N.E. (1996) Trends in nutrients. Hydrological Processes,

10: 263-293.

Heggie, D.T., Skyring, G.W., Orchardo, J., Longmore, A.R., Nicholson, G.J. and Berelson,

W.M. (1999) Denitrification and denitrifying efficiencies in sediments of Port Phillip

Bay: Direct determination of biogenic N2 and N-metabolite fluxes with implications for

water quality. Marine and Freshwater Research, 50: 589-596.

Heilskov, A.C. and Holmer, M. (2001) Effects of benthic fauna on organic mater mineralization

in fish-farm sediments: importance of size and abundance. ICES Journal of Marine

Science, 58: 427-434.

Helder, W. and De Vries, R.T.P. (1983) Estuarine nitrite maxima and nitrifying bacteria (Ems-

Dollard Estuary). Netherlands Journal of Sea Research, 17(1): 1-18.

Henriksen, K. and Kemp, W.M. (1988) Nitrification in estuarine and coastal marine sediments.

In: T.H. Blackburn and J. Sørensen eds, Nitrogen Cycling in Coastal Marine

Environments. SCOPE 33. Wiley, Chichester, pp. 207-249.

Henriksen, K., Hansen, J.I. and Blackburn, T.H. (1980) The influence of benthic infauna on

exchange rates of inorganic nitrogen between sediment and water. Ophelia Supplement, 1:

249-256.

Henriksen, K., Rasmussen, M.B. and Jensen, A. (1983) Effect of bioturbation on microbial

nitrogen transformations in the sediment and fluxes of ammonium and nitrate to the

overlaying water. In: R. Hallberg ed., Environmental Biogeochemistry. Ecological

Bulletin (Stockholm) 35: 193-205.

Hodgkin, E.P. (1987) The hydrology of the Swan River Estuary: Salinity the ecological master

factor. In: J. John ed., Swan River Estuary, Ecology and Management. Curtin University

Environmental Studies Group Report, 1: 34-44.

Hodgkin, E.P. and Lenanton, R.C. (1981) Estuaries and coastal lagoons of south Western

Australia. In: B.J. Neilson and L.E. Cronin eds, Estuaries and Nutrients. Humana Press,

New Jersey, pp. 309-321.

Hodgkin, E.P. and Vicker, E. (1987) A history of algal pollution in the estuary of the Swan

River. In: J. John ed., Swan River Estuary, Ecology and Management. Curtin University

Environmental Studies Group Report, 1: 65-70.

Hopkinson, C.S. and Wetzel, R.L. (1982) In situ measurements of nutrient and oxygen fluxes in

a coastal marine benthic community. Marine Ecology Progress Series, 10: 29-35.

Hordijk, C.A., Snieder, M., van Engelen, J.J.M. and Cappenberg, T.E. (1987) Estimation of

bacterial nitrate reduction at in situ concentrations in freshwater sediments. Applied and

Environmental Microbiology, 53(2): 217-223.

Horner Rosser, S.M.J. and Thompson, P.A. (2001) Phytoplankton of the Swan-Canning

Estuary: A comparison of nitrogen uptake by different bloom assemblages. Hydrological

Processes, 15: 2579-2594.

150

Page 165: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Hutchings, P.A. (1982) Bristleworms (Phylum Annelida). In: S.A. Shepherd and I.M. Thomas

eds, Marine Invertebrates of Southern Australia, Part 1. Government Printer, South

Australia, pp. 228-298.

Hutchings, P.A. and Glasby, C.J. (1982) Two new species of Ceratonereis (Polychaeta:

Nereididae) from estuarine areas of New South Wales, Australia. Proceedings of

Biological Society of Washington, 95: 515-521.

Hutchings, P.A. and Glasby, C.J. (1985) Aditional nereidids (Polychaeta) from eastern

Australia, together with a redescription of Namanereis quadraticeps (Gay) and the

synonomising of Ceratonereis pseudoerythraeensis (Hutchngs and Turvey) with C.

aequisetis (Augener). Records of the Australian Museum, 37: 101-110.

Hutchings, P.A. and Murray, A. (1984) Taxonomy of polychaetes from the Hawkesbury River

and the southern estuaries of New South Wales, Australia. Records of the Australian

Museum Supplement, 3: 1-119.

Hutchings, P.A. and Reid, A. (1990) The Nereididae (Polychaeta) from Australia –

Gymnonereidinae sensu Fitzhugh, 1987: Australonereis, Ceratocephale, Dendronereides,

Gymnonereis, Nicon, Olganereis and Websterinereis. Records of the Australian Museum,

42: 241-274.

Hutchings, P.A. and Turvey, S.P. (1982) The Nereididae of South Australia. Transactions of the

Royal Society of South Australia, 106: 93-144.

Hylleberg, J. and Henriksen, K. (1980) The central role of bioturbation in sediment

mineralisation and element recycling. Ophelia Supplement, 1: 1-16.

Hynes, R. K. and Knowles, R. (1978) Inhibition by acetylene of ammonia oxidation in

Nitrosomonas europaea. FEMS Microbiology Letters, 4: 319-321.

Hynes, R.K. and Knowles, R. (1982) Effect of acetylene on autotrophic and heterotrophic

nitrification. Canadian Journal of Microbiology, 28: 334-340.

Iizumi, H., Hattori, A. and McRoy, C.P. (1980) Nitrate and nitrite in interstitial waters of

eelgrass beds in relation to the rhizosphere. Journal of Experimental Marine Biology and

Ecology, 47: 191-201.

Jansson, M. (1980) Role of benthic algae in transport of nitrogen from sediments to lake water

in a shallow clearwater lake. Archiv für Hydrobiologie, 89: 101-109.

Jenkins, M.C. and Kemp, W.M. (1984) The coupling of nitrification and denitrification in two

estuarine sediments. Limnology and Oceanography, 29(3): 609-619.

Jensen, M.H., Andersen, T.K. and Sørensen, J. (1988) Denitrification in coastal bay sediment:

Regional and seasonal variation in Aarhus Bight, Denmark. Marine Ecology Progress

Series, 48: 155-162.

151

Page 166: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Jeter, R.M. and Ingraham, J.L. (1981) The denitrifying prokaryotes. In: M.P. Starr, H. Stolp,

H.G. Trüper, A. Balows and H.G. Schlegel eds, The Prokaryotes: A Handbook of

Habitats, Isolation, and Identification of Bacteria, Volume I. Springer-Verlag, Berlin, pp.

913-925.

Jetten, M.S.M. (2001) New pathways for ammonia conversion in soil and aquatic systems.

Plant and Soil, 230: 9-19.

Jetten, M.S.M., Logemann, S., Muyzer, G., Robertson, L.A., de Vries, S., van Loosdrecht,

M.C.M. and Kuenen, J.G. (1997) Novel principles in the microbial conversion of nitrogen

compounds. Antonie van Leeuwenhoek, 71: 75-93.

Jetten, M.S.M., Sliekers, O., Kuypers, M., Dalsgaard, T., van Niftrik, L., Cirpus, I., van de Pas-

Schoonen, K., Lavik, G., Thamdrup, B., Le Paslier, D., Op den Camp, H.J.M., Hulth, S.,

Nielsen, L.P., Abma, W., Third, K., Engström, P., Kuenen, J.G., Jørgensen, B.B.,

Canfield, D.E., Sinninghe Damsté, J.S., Revsbech, N.P., Fuerst, J., Weissenbach, J.,

Wagner, M., Schmidt, I., Schmid, M. and Strous, M. (2003) Anaerobic ammonium

oxidation by marine and freshwater planctomycete-like bacteria. Applied Microbiology

and Biotechnology, 63: 107-114.

Johannes, R.E. (1968) Nutrient regeneration in lakes and oceans. In: M. Droop and E.J.F. Wood

eds, Advances in Marine Microbiology. Academic Press, New York, pp. 203-213.

John, J. (1987) Phytoplankton of the Swan River Estuary. In: J. John ed., Swan River Estuary,

Ecology and Management. Curtin University Environmental Studies Group Report, 1: 71-

89.

John, J. (1994) Eutrophication of the Swan River Estuary, Western Australia and the

management strategy. In: W.J. Mitsch ed., Global Wetlands: Old World and New.

Elsevier, Amsterdam, pp. 749-759.

Johnson, D.L. (2002) Darwin would be proud: Bioturbation, dynamic denudation, and the

power of theory in science. Geoarchaeology: An International Journal, 17(1): 7-40.

Jones, J.G., Simon, B.M. and Horsley, R.W. (1982) Microbiological sources of ammonia in

freshwater lake sediments. Journal of General Microbiology, 128: 2823-2831.

Jørgensen, B.B. (1980) Seasonal oxygen depletion in the bottom waters of a Danish fjord and its

effects on the benthic community. Oikos, 34: 68-76.

Jørgensen, B.B. and Revsbech, N.P. (1985) Diffusive boundary layers and the oxygen uptake of

sediments and detritus. Limnology and Oceanography, 30(1): 111-122.

Jørgensen, K.S. (1989) Annual pattern of denitrification and nitrate ammonification in estuarine

sediment. Applied and Environmental Microbiology, 55(7): 1841-1847.

Jørgensen, K.S. and Sørensen, J. (1988) Two annual maxima of nitrate reduction and

denitrification in estuarine sediment, (Norsminde Fjord, Denmark). Marine Ecology

Progress Series, 48: 147-154.

152

Page 167: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Joye, S.B., Smith, S.V., Hollibaugh, J.T. and Paerl, H.W. (1996) Estimating denitrification rates

in estuarine sediments: A comparison of stoichiometric and acetylene based methods.

Biogeochemistry, 33: 197-215.

Kalejta, (1992) Distribution, biomass and production of Ceratonereis erythraeensis (Fauvel)

and Ceratonereis keiskama (Day) at the Berg River Estuary, South Africa. South African

Journal of Zoology, 27(3): 121-129.

Kanandjembo, A.N. (1998) Fish and macrobenthic faunas of the upper Swan Estuary:

Interrelationships and possible influence of a dinoflagellate bloom. M.Sc. thesis, Murdoch

University, Western Australia.

Kanandjembo, A.N., Platell, M.E. and Potter, I.C. (2001) The benthic macroinvertebrate

community of the upper reaches of an Australian estuary that undergoes marked seasonal

changes in hydrology. Hydrological Processes, 15: 2481-2501.

Kaplan, W.A. (1983) Nitrification. In: E.J. Carpenter and D.G. Capone eds, Nitrogen in the

Marine Environment. Academic Press, New York, pp. 139-190.

Kaspar, H.F. (1982) Denitrification in marine sediment: Measurement of capacity and estimate

of in situ rate. Applied and Environmental Microbiology, 43(3): 522-527.

Kaspar, H.F. and Tiedje, J.M. (1981) Dissimilatory reduction of nitrate and nitrite in the bovine

rumen: Nitrous oxide production and effect of acetylene. Applied and Environmental

Microbiology, 41(3): 705-709.

Kaspar, H.F., Asher, R.A., and Boyer. I.C. (1985a) Microbial nitrogen transformation in

sediments and inorganic nitrogen fluxes across the sediment/water interface on the South

Island West Coast, New Zealand. Estuarine, Coastal and Shelf Science, 21: 245-255.

Kaspar, H.F., Gillespie, P.A., Boyer, I.C., and MacKenzie, A.L. (1985b) Effects of mussel

aquaculture on the nitrogen cycle and benthic communities in Kenepuru Sound,

Marlborough Sounds, New Zealand. Marine Biology, 85: 127-136.

Keeney, D.R. (1973) The nitrogen cycle in sediment-water systems. Journal of Environmental

Quality, 2(1): 15-29.

Kemp, W.M., Sampou, P., Caffrey, J., Mayer, M., Henriksen, K. and Boynton, W.R. (1990)

Ammonium recycling versus denitrification in Chesapeake Bay sediments. Limnology

and Oceanography, 35(7): 1545-1563.

Kennish, M.J. (1990) Ecology of Estuaries. Volume II – Biological Aspects. CRC Press, Boca

Raton, Florida, 391 pp.

Kent, A.C. and Day, R.W. (1983) Population dynamics of an infaunal polychaete: The effect of

predators and an adult-recruit interaction. Journal of Experimental Marine Biology and

Ecology, 73: 185-203.

Khlebovich, V.V. (1996) Fauna of Russia and neighbouring countries. Polychaetous Annelids

Volume III. Polychaetes of the family Nereididae of the Russian Seas and the adjacent

waters. NAUKA Publishing House, St Petersberg [In Russian].

153

Page 168: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Kikuchi, E. and Kurihara, Y. (1977) In vitro studies on the effects of tubificids on the

biological, chemical and physical characteristics of submerged ricefield soil and overlying

water. Oikos, 29: 348-356.

Kinberg, J.G.H. (1866) Annulata Nova. Ofversigt af Kongliga Vetenskaps-Akademiens

Forhandlingar, Stockholm, 23: 239-258.

Klemedtsson, L., Svensson, B.H., Lindberg, T. and Rosswall, T. (1977) The use of acetylene

inhibition of nitrous oxide reductase in quantifying denitrification in soils. Swedish

Journal of Agricultural Research, 7: 179-185.

Klump, J.V. and Martens, C.S. (1983) Benthic nitrogen regeneration. In: E.J. Carpenter and

D.G. Capone eds, Nitrogen in the Marine Environment. Academic Press, New York, pp.

411-457.

Knowles, R. (1982) Denitrification. Microbiological Reviews, 46(1): 43-70.

Knox, G.A. (1977) The role of polychaetes in benthic soft-bottom communities. In: D.J. Reish

and K. Fauchald eds, Essays on Polychaetous Annelids: In Memory of Dr. Olga Hartman.

Allan Hancock Foundation, University of Southern California, Los Angeles, pp. 547-604.

Kohl, D.H., Vithayathil, F., Whitlow, P., Shearer, G. and Chien, S.H. (1976) Denitrification

kinetics in soil systems: The significance of good fits of data to mathematical forms. Soil

Science Society of America Journal, 40: 249-253.

Koike, I. and Hattori, A. (1979) Estimates of denitrification in sediments of the Bering Sea

shelf. Deep-Sea Research, 26A: 409-415.

Koike, I. and Sørensen, J. (1988) Nitrate reduction and denitrification in marine sediments. In:

T.H. Blackburn and J. Sørensen eds, Nitrogen Cycling in Coastal Marine Environments.

SCOPE 33. Wiley, Chichester, pp. 251-273.

Koike, I., Hattori, A. and Goering, J.J. (1978) Controlled ecosystem pollution experiment:

Effect of mercury on enclosed water columns. VI. Denitrification by marine bacteria.

Marine Science Communications, 4(1): 1-12.

Kott, P. (1951) Nereidae and Eunicidae of south Western Australia; also notes on the ecology of

Western Australian limestone reefs. Journal of the Royal Society of Western Australia,

35: 85-130.

Krantzberg, G. (1985) The influence of bioturbation on physical, chemical and biological

parameters in aquatic environments: A review. Environmental Pollution (Series A), 39:

99-122.

Kristensen, E. (1984a) Effect of natural concentrations on nutrient exchange between a

polychaete burrow in estuarine sediment and the overlying water. Journal of

Experimental Marine Biology and Ecology, 75: 171-190.

Kristensen, E. (1984b) Life cycle, growth and production in estuarine populations of the

polychaetes Nereis virens and N. diversicolor. Holarctic Ecology, 7: 249-256.

154

Page 169: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Kristensen, E. (1985) Oxygen and inorganic nitrogen exchange in a Nereis virens (Polychaeta)

bioturbated sediment-water system. Journal of Coastal Research, 1: 109-116.

Kristensen, E. (1988) Benthic fauna and biogeochemical processes in marine sediments:

Microbial activities and fluxes. In: T.H. Blackburn and J. Sørensen eds, Nitrogen Cycling

in Coastal Marine Environments. SCOPE 33. Wiley, Chichester, pp. 275-299.

Kristensen, E. (2000) Organic matter diagenesis at the oxic/anoxic interface in coastal marine

sediments, with emphasis on the role of burrowing animals. Hydrobiologia, 426: 1-24.

Kristensen, E., Jensen, M.H. and Aller, R.C. (1991) Direct measurement of dissolved inorganic

nitrogen exchange and denitrification in individual polychaete (Nereis virens) burrows.

Journal of Marine Research, 49: 355-377.

Kristensen, E., Jensen, M.H. and Andersen, T.K. (1985) The impact of polychaete (Nereis

virens Sars) burrows on nitrification and nitrate reduction in estuarine sediments. Journal

of Experimental Marine Biology and Ecology, 85: 75-91.

Kuai, L. and Verstraete, W. (1998) Ammonium removal by the oxygen-limited autotrophic

nitrification-denitrification system. Applied and Environmental Microbiology, 64(11):

4500-4506.

Kuenen, J.G. and Robertson, L.A. (1988) Ecology of nitrification and denitrification. In: J.A.

Cole and S.J. Ferguson eds, The Nitrogen and Sulphur Cycles, 42nd Symposium of the

Society for General Microbiology. Cambridge University Press, Cambridge, pp. 161-218.

Kurup, G.R., Hamilton, D.P. and Patterson, J.C. (1998) Modelling the effect of seasonal flow

variations on the position of salt wedge in a microtidal estuary. Estuarine, Coastal and

Shelf Science, 47: 191-208.

Kuwae, T., Hosokawa, Y. and Eguchi, N. (1998) Dissolved inorganic nitrogen cycling in Banzu

intertidal sand-flat, Japan. Mangroves and Salt Marshes, 2: 167-175.

Kuypers, M.M.M., Sliekers, A.O., Lavik, G., Schmid, M., Jørgensen, B.B., Kuenen, J.G.,

Sinninghe Damsté, J.S., Strous, M. and Jetten, M.S.M. (2003) Anaerobic ammonium

oxidation by anammox bacteria in the Black Sea. Nature, 422: 608-611.

LaMontagne, M.G., Astorga, V., Giblin, A.E. and Valiela, I. (2002) Denitrification and the

stoichiometry of nutrient regeneration in Waquoit Bay, Massachusetts. Estuaries, 25(2):

272-281.

Law, C.S., Rees, A.P. and Owens, N.J.P. (1991) Temporal variability of denitrification in

estuarine sediments. Estuarine, Coastal and Shelf Science, 33: 37-56.

Lean, D.R.S., Liao, C.F.H., Murphy, T.P. and Painter, D.S. (1978) The importance of nitrogen

fixation in lakes. Ecological Bulletin (Stockholm), 26: 41-56.

Lee II, H. and Swartz, R.C. (1980) Biological processes affecting the distribution of pollutants

in marine sediments. Part II. Biodeposition and bioturbation. In: R.A. Baker ed.,

Contaminants and Sediments, Volume 2. Science Publishers, Ann Arbor, Michigan, pp.

555-605.

155

Page 170: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Lewin, J., Eckman, J.E. and Ware, G.N. (1979) Blooms of surf-zone diatoms along the coast of

the Olympic Peninsula, Washington. XI. Regeneration of ammonium in the surf

environment by the Pacific razor clam Siliqua patula. Marine Biology, 52: 1-9.

Linderfelt, W.R. and Turner, J.V. (2001) Interaction between shallow groundwater, saline

surface water and nutrient discharge in a seasonal estuary: The Swan-Canning system.

Hydrological Processes, 15: 2631-2653.

Llansó, R.J. (1992) Effects of hypoxia on estuarine benthos: The lower Rappahannock River

(Chesapeake Bay), a case study. Estuarine, Coastal and Shelf Science, 35: 491-515.

Lohrer, A.M., Thrush, S.F. and Gibbs, M.M. (2004) Bioturbators enhance ecosystem function

through complex biogeochemical interactions. Nature, 431: 1092-1095.

Loneragan, N.R. and Potter, I.C. (1990) Factors influencing community structure and

distribution of different life-cycle categories of fishes in shallow waters of a large

Australian estuary. Marine Biology, 106: 25-37.

Loneragan, N.R., Potter, I.C. and Lenanton, R.C.J. (1989) Influence of site, season and year on

contributions made by marine, estuarine, diadromous and freshwater species to the fish

fauna of a temperate Australian estuary. Marine Biology, 103: 461-479.

Loo, L.-O. and Rosenberg, R. (1989) Bivalve suspension-feeding dynamics and benthic-pelagic

coupling in an eutrophicated marine bay. Journal of Experimental Marine Biology and

Ecology, 130: 253-276.

Maag, M., Malinovsky, M. and Nielsen, S.M. (1997) Kinetics and temperature dependence of

potential denitrification in riparian soils. Journal of Environmental Quality, 26: 215-223.

MacFarlane, G.R. and Booth, D.J. (2001) Estuarine macrobenthic community structure in the

Hawkesbury River, Australia: Relationships with sediment physicochemical and

anthropogenic parameters. Environmental Monitoring and Assessment, 72: 51-78.

Mayer, M.S., Schaffner, L. and Kemp, W.M. (1995) Nitrification potentials of benthic

macrofaunal tubes and burrow walls: Effects of sediment NH4+ and animal irrigation

behaviour. Marine Ecology Progress Series, 121: 157-169.

McCaffrey, R.J., Myres, A.C., Davey, E., Morrison, G., Bender, M., Luedtke, N., Cullen, D.,

Froelich, P. and Klinkhammer, G. (1980) The relation between pore water chemistry and

benthic fluxes of nutrients and manganese in Narragansett Bay, Rhode Island. Limnology

and Oceanography, 25(1): 31-44.

McShane, P. (1977) Studies on the Environmental Physiology of some Errant Polychaetes from

the Swan River Estuary. Honours Thesis, The University of Western Australia, Perth.

Medernach, L., Jordana, E., Grémare, A., Nozais, C., Charles, F. and Amouroux, J.M. (2000)

Population dynamics, secondary production and calcification in a Meditteranean

population of Ditrupa arietina (Annelida: Polychaeta). Marine Ecology Progress Series,

199: 171-184.

156

Page 171: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Mermillod-Blondin, F., Gérino, M., Degrange, V., Lensi, R., Chassé, J.-L., Rard, M. and des

Châtelliers, M.C. (2001) Testing the functional redundancy of Limnodrilus and Tubifex

(Oligochaeta, Tubificidae) in hyporheic sediments: An experimental study in microcosms.

Canadian Journal of Fisheries and Aquatic Science, 58: 1747-1759.

Mermillod-Blondin, F., Gérino, M., des Châtelliers, M.C. and Degrange, V. (2002) Functional

diversity among 3 detritivorous hyporheic invertebrates: An experimental study in

microcosms. Journal of the North American Benthological Society, 21(1): 132-149.

Messer, J.J. and Brezonik, P.L. (1984) Laboratory evaluation of kinetic parameters for lake

sediment denitrification models. Ecological Modelling, 21: 277-286.

Meyer, R.L., Kjær, T. and Revsbech, N.P. (2001) Use of NOx- microsensors to estimate the

activity of sediment nitrification and NOx- consumption along an estuarine salinity,

nitrate, and light gradient. Aquatic Microbial Ecology, 26: 181-193.

Meyer-Reil, L.-A. and Köster, M. (2000) Eutrophication of marine waters: Effects on benthic

microbial communities. Marine Pollution Bulletin, 41(1-6): 255-263.

Middelburg, J.J. and Nieuwenhuize (2000a) Uptake of dissolved inorganic nitrogen in turbid,

tidal estuaries. Marine Ecology Progress Series, 192: 79-88.

Middelburg, J.J. and Nieuwenhuize (2000b) Nitrogen uptake by heterotrophic bacteria and

phytoplankton in the nitrate-rich Thames estuary. Marine Ecology Progress Series, 203:

13-21.

Monbet, Y. (1992) Control of phytoplankton biomass in estuaries: A comparative analysis of

microtidal and macrotidal estuaries. Estuaries, 15: 563-571.

Monro, C.C.A. (1938) On a small collection of Polychaeta from Swan River, Western Australia.

The Annals and Magazine of Natural History (Series II), 2: 614-624.

Morris, L. and Keough, M.J. (2003) Testing the effects of nutrient additions on mudflat

macroinfaunal assemblages in the presence and absence of shorebird predators. Marine

and Freshwater Research, 54: 859-874.

Mortimer, C.H. (1941) The exchange of dissolved substances between mud and water in lakes.

Journal of Ecology, 29(2): 280-329.

Mortimer, C.H. (1942) The exchange of dissolved substances between mud and water in lakes.

Journal of Ecology, 30(1): 147-201.

Mortimer, R.J.G., Davey, J.T., Krom, M.D., Watson, P.G., Frickers, P.E. and Clifton, R.J.

(1999) The effect of macrofauna on porewater profiles and nutrient fluxes in the intertidal

zone of the Humber Estuary. Estuarine, Coastal and Shelf Science, 48: 683-699.

Mosier, A.R. (1980) Acetylene inhibition of ammonium oxidation in soil. Soil Biology and

Biochemistry, 12: 443-444.

Moura, I. and Moura, J.J.G. (2001) Structural aspects of denitrifying enzymes. Current Opinion

in Chemical Biology, 5: 168-175.

157

Page 172: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Mulder, A., van de Graaf, A.A., Robertson, L.A. and Kuenen, J.G. (1995) Anaerobic

ammonium oxidation discovered in a denitrifying fluidised bed reactor. FEMS

Microbiology Ecology, 16: 177-183.

Murray, R.E., Parsons, L.L. and Smith, M.S. (1989) Kinetics of nitrate utilization by mixed

populations of denitrifying bacteria. Applied and Environmental Microbiology, 55(3):

717-721.

Myrold, D.D. and Tiedje, J.M. (1985) Diffusional constraints on denitrification in soil. Soil

Science Society of America Journal, 49: 651-657.

Naeem, S., Thompson, L.J., Lawler, S.P., Lawton, J.H. and Woodfin, R.M. (1994) Declining

biodiversity can alter the performance of ecosystems. Nature, 368:734-737.

Nedwell, D.B. (1975) Inorganic nitrogen metabolism in a eutrophicated tropical mangrove

estuary. Water Research, 9: 221-231.

Nedwell, D.B. (1982) Exchange of nitrate, and the products of bacterial nitrate reduction,

between seawater and sediment from a U.K. saltmarsh. Estuarine, Coastal and Shelf

Science, 14: 557-566.

Nicholson, G.J. and Longmore, A.R. (1999) Causes of observed temporal variability of nutrient

fluxes from a southern Australian marine embayment. Marine and Freshwater Research,

50: 581-588.

Nishio, T., Koike, I. And Hattori, A. (1983) Estimates of denitrification and nitrification in

coastal and estuarine sediments. Applied and Environmental Microbiology, 45(2): 444-

450.

Nithart, M., Alliot, E. and Salen-Picard, C. (1999) Production, respiration and ammonia

excretion of two polychaete species in a north Norfolk saltmarsh. Journal of the Marine

Biological Association of the United Kingdom, 79: 1029-1037.

Nixon, S.W. (1981) Remineralisation and nutrient cycling in coastal marine ecosystems. In: B.J.

Neilson and L.E. Cronin eds, Estuaries and Nutrients. Humana Press, Clifton, New Jersey,

pp. 111-138.

Nixon, S.W. (1995) Coastal marine eutrophication: A definition, social causes and future

concerns. Ophelia, 41: 199-220.

Nixon, S.W., Oviatt, C.A. and Hale, S.S. (1976) Nitrogen regeneration and the metabolism of

coastal marine bottom communities. In: J.M. Anderson and A. Macfadyen eds, The Role

of Terrestrial and Aquatic Organisms in Decomposition Processes. Blackwell Scientific,

Oxford, pp. 269-283.

Nowicki, B.L., Requintina, E., Van Keuren, D.V. and Portnoy, J. (1999) The role of sediment

denitrification in reducing groundwater-derived nitrate inputs to Nauset Marsh Estuary,

Cape Cod, Massachusetts. Estuaries, 22(2A): 245-259.

Officer, C.B., Smayda, T.J. and Mann, R. (1982) Benthic filter feeding: A natural

eutrophication control. Marine Ecology Progress Series, 9: 203-210.

158

Page 173: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Ogilvie, B., Nedwell, D.B., Harrison, R.M., Robinson, A. and Sage, A. (1997) High nitrate,

muddy estuaries as nitrogen sinks: The nitrogen budget of the River Colne estuary

(United Kingdom). Marine Ecology Progress Series, 150: 217-228.

Oldham, C.E. and Lavery, P.S. (1999) Porewater nutrient fluxes in a shallow fetch-limited

estuary. Marine Ecology Progress Series, 183: 39-47.

Omena, E.P. and Amaral, A.C.Z. (2000) Population dynamics and secondary production of

Laeonereis acuta (Treadwell, 1923) (Nereididae: Polychaeta). Bulletin of Marine Science,

67(1): 421-431.

Oremland, R.S. and Capone, D.G. (1987) Use of specific inhibitors in microbial ecology and

biogeochemical studies. In: K.C. Marshall ed., Advances in Microbial Ecology. Plenum

Press, New York, pp. 285-383.

Oremland, R.S., Umberger, C., Culbertson, C.W. and Smith, R.L. (1984) Denitrification in San

Francisco Bay intertidal sediments. Applied and Environmental Microbiology, 47: 1106-

1112.

Oren, A. and Blackburn, T.H. (1979) Estimation of sediment denitrification rates at in situ

nitrate concentrations. Applied and Environmental Microbiology, 37(1): 174-176.

Owens, N.J.P. (1993) Nitrate cycling in marine waters. In: T.P. Burt, A.L. Heathwaite and S.T.

Trudgill, eds, Nitrate: Processes, Patterns and Management. Wiley, Chichester, pp. 169-

209.

Paasche, E. (1988) Pelagic primary production in nearshore waters. In: T.H. Blackburn and J.

Sørensen eds, Nitrogen Cycling in Coastal Marine Environments. SCOPE 33. Wiley,

Chichester, pp. 33-57.

Paerl, H.W. and Zehr, J.P. (2000) Marine nitrogen fixation. In: D.L. Kirchman ed., Microbial

Ecology of the Oceans. Wiley, New York, pp. 387-426.

Parsonage, D., Greenfield, A.J. and Ferguson, S.J. (1985) The high affinity of Paracoccus

denitrificans cells for nitrate as an electron acceptor. Analysis of possible mechanisms of

nitrate and nitrite movement across the plasma membrane and the basis for inhibition by

added nitrite of oxidase activity in permeabilised cells. Biochimica et Biophysica Acta,

807: 81-95.

Payne, W.J. (1973) Reduction of nitrogenous oxides by microorganisms. Bacterial Reviews, 37:

409-452.

Payne, W.J. (1981) Denitrification. Wiley, New York, 214 pp.

Pedersen, M.F. and Borum, J. (1996) Nutrient control of algal growth in estuarine waters.

Nutrient limitation and the importance of nitrogen requirements and nitrogen storage

among phytoplankton and species of macroalgae. Marine Ecology Progress Series, 142:

261-272.

159

Page 174: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Pelegri, S. P., and Blackburn, T. H. (1994) Bioturbation effects of the amphipod Corophium

volutator on microbial nitrogen transformations in marine sediments. Marine Biology,

121: 253-258.

Pelegri, S. P., and Blackburn, T. H. (1995a) Effect of bioturbation by Nereis sp., Mya arenaria

and Cerastoderma sp. on nitrification and denitrification in estuarine sediments. Ophelia,

42: 289-299.

Pelegri, S. P., and Blackburn, T. H. (1995b) Effects of Tubifex tubifex (Oligochaeta:

Tubificidae) on N-mineralization in freshwater sediments, measured with 15N isotopes.

Aquatic Microbial Ecology, 9: 289-294.

Pennifold, M. and Davis, J. (2001) Macrofauna and nutrient cycling in the Swan River Estuary:

experimental results. Hydrological Processes, 15: 2537-2553.

Peters, N.E. and Donohue, R. (2001) Nutrient transport to the Swan-Canning Estuary, Western

Australia. Hydrological Processes, 15: 2555-2577.

Petr, T. (1977) Bioturbation and exchange of chemicals in the mud-water interface. In: H.L.

Golterman ed., Interactions Between Sediments and Freshwater: Proceedings of an

International Symposium Held at Amsterdam, The Netherlands, September 6-10, 1976.

Dr. W. Junk B.V. Publishers, The Hague, pp. 216-226.

Petti, M.A., Nonato, V.E.F. and Paiva, P.C. (1996) Trophic relationships between polychaetes

and brachyuran crabs on the southeastern Brazilian coast. Rev. Bras. Oceanogr., 44: 9-21.

Philips, S., Wyffels, S., Sprengers, R. and Verstraete, W. (2002) Oxygen-limited autotrophic

nitrification/denitrification by ammonia oxidisers enables upward motion towards more

favourable conditions. Applied Microbiology Biotechnology, 59: 557-566.

Plattel, M.E. and Potter, I.C. (1996) Influence of water depth, season, habitat and estuary

location on the macrobenthic fauna of a seasonally closed estuary. Journal of the Marine

Biological Association of the United Kingdom, 76:1-21.

Pomeroy, L.R., Smith, E.E. and Grant, C. (1965) The exchange of phosphate between estuarine

water and sediments. Limnology and Oceanography, 10(2): 167-172.

Postma, H., Kemp, W.M., Colebrook, J.M., Horwood, J., Joint, I.R., Lampitt, R., Nixon, S.W.,

Pilson, M.E.Q. and Wulff, F. (1984) Nutrient cycling in estuarine and coastal marine

ecosystems. In: M.J.R. Fasham ed., Flows of Energy and Materials in Marine Ecosystems

– Theory and Practice. Plenum Press, New York, pp. 651-661.

Pritchard, D.W. (1967) What is an estuary, physical viewpoint. In G. Lauff ed., Estuaries.

American Association for the Advancement of Science Publication, 83: 3-5.

Rabouille, C., Mackenzie, F.T. and Ver, L.M. (2001) Influence of the human pertubation on

carbon, nitrogen and oxygen biogeochemical cycles in the global coastal ocean.

Geochimica et Cosmochimica Acta, 65(21): 3615-3641.

160

Page 175: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Raymond, N., Bonin, P. and Bertrand, J.-C. (1992) Comparison of methods for measuring

denitrifying activity in marine sediments from the Western Mediterranean coast.

Oceanologica Acta, 15(2): 137-143.

Regnault, M., Boucher-Rodoni, R., Boucher, G. and Lasserre, P. (1988) Effects of macrofauna

excretion and turbulence on inorganic nitrogenous exchanges at the water-sediment

interface. Experimental approach in microcosms. Cahiers de Biologie Marine, 29: 427-

444.

Reichardt, W. (1988) Impact of bioturbation by Arenicola marina on microbiological

parameters in intertidal sediments. Marine Ecology Progress Series, 44: 149-158.

Revsbech, N.P. and Sørensen, J. (1990) Combined use of the acetylene inhibition technique and

microsensors for quantification of denitrification in sediments and biofilms. In: N.P.

Revsbech and J. Sørensen eds, Denitrification in soil and sediment. Plenum Press, New

York, pp. 259-275.

Revsbech, N.P., Sørensen, J., Blackburn, T.H. and Lomholt, J.P. (1980) Distribution of oxygen

in marine sediments measured with microelectrodes. Limnology and Oceanography,

25(3): 403-411.

Rhoads, D.C. (1974) Organism-sediment relations on the muddy sea floor. Oceanography and

Marine Biology Annual Review, 12: 263-300.

Rhoads, D.C., Aller, R.C. and Goldhaber, M.B. (1978) The influence of colonising benthos on

physical properties and chemical diagenesis of the estuarine seafloor. In: B.C. Coull ed.,

Ecology of Marine Benthos. University of South Carolina Press, Columbia, pp. 113-138.

Rhyther, J.H. and Dunstan, W.M. (1971) Nitrogen, phosphorous and eutrophication in the

coastal marine environment. Science, 171: 1008-1013.

Riisgård, H.U., Poulsen, L. and Larsen, P.S. (1996) Phytoplankton reduction in near-bottom

water caused by filter-feeding Nereis diversicolor – implications for worm growth and

population grazing impact. Marine Ecology Progress Series, 141: 47-54.

Risgaard-Petersen, N. and Ottosen, L.D.M. (2000) Nitrogen cycling in two temperate Zostera

marina beds: Seasonal variation. Marine Ecology Progress Series, 198: 93-107.

Risgaard-Petersen, N., Nielsen, L.P., Rysgaard, S., Dalsgaard, T. and Meyer, R.L. (2003)

Application of the isotope pairing technique in sediments where annamox and

denitrification coexist. Limnology and Oceanography: Methods, 1: 63-73.

Rittenberg, S.C., Emery, K.O. and Orr, W.L. (1955) Regeneration of nutrients in sediments of

marine basins. Deep-Sea Research, 3: 23-45.

Rose, T.H. (1994) Comparisons of the benthic and zooplankton communities in the eutrophic

Peel-Harvey and nearby Swan estuaries in south-western Australia. Ph.D. Thesis,

Murdoch University, Western Australia.

161

Page 176: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Rouse, G.W. (2000) Class Polychaeta: Morphology and physiology. In: P.L. Beesley, G.J.B.

Ross and C.J. Glasby, eds, Polychaetes and Allies: The Southern Synthesis. Fauna of

Australia, Volume 4a: Polychaeta, Myzostomida, Pogonophora, Echiura, Sipuncula.

CSIRO Publishing, Melbourne, pp. 9-32.

Rowe, G.T., Clifford, C.H., Smith Jr., K.L. and Hamilton, P.L. (1975) Benthic nutrient

regeneration and its coupling to primary productivity in coastal waters. Nature, 255: 215-

217.

Rudolph, J., Frenzel, P. and Pfenning, N. (1991) Acetylene inhibition technique underestimates

in situ denitrification rates in intact cores of freshwater sediment. FEMS Microbiology

Ecology, 85: 101-106.

Ryther, J.H. and Dunstan, W.M. (1971) Nitrogen, phosphorous and eutrophication in the coastal

marine environment. Science, 171: 1008-1013.

Ryzhova, I.M. (1979) Effect of nitrate concentration on the rate of soil denitrification.

Pochvovedeniye, 3: 54-58.

Saier Jr., M.H. (1987) Enzymes in Metabolic Pathways: A Comparative Study of Mechanism,

Structure, Evolution, and Control. Harper and Row, New York, 246 pp.

Sardá R., Pinedo, S. and Dueso, A. (2000) Estimating secondary production in natural

populations of polychaetes: Some general constraints. Bulletin of Marine Science, 67(1):

433-447.

Sayama, M. (2001) Presence of nitrate-accumulating sulfur bacteria and their influence on

nitrogen cycling in a shallow coastal marine sediment. Applied and Environmental

Microbiology, 67(8): 3481-3487.

Sayama, M. and Kurihara, Y. (1983) Relationship between burrowing activity of the

polychaetous annelid, Neanthes japonica (Izuka) and nitrification-denitrification

processes in the sediments. Journal of Experimental Marine Biology and Ecology, 72:

233-241.

Scanes, P., Coade, G., Large, D. and Roach, T. (1998) Developing criteria for acceptable loads

of nutrients from catchments. In: Australian Water and Wastewater Association ed.,

Proceedings of the Coastal Nutrients Workshop, Sydney (October 1997). Artarmon,

Sydney, pp. 89-99.

Schink, D.R. and Guinasso Jr., J.L. (1977) Effects of bioturbation on sediment-seawater

interaction. Marine Geology, 23: 133-154.

Schipper, L.A., Cooper, A.B., Harfoot, C.G. and Dyck, W.J. (1993) Regulators of

denitrification in an organic riparian soil. Soil Biology and Biochemistry, 25(7): 925-933.

Schmidt, I. and Bock, E. (1997) Anaerobic ammonia oxidation with nitrogen dioxide by

Nitrosomonas eutropha. Archives of Microbiology, 167: 106-111.

162

Page 177: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Schmidt, I., Sliekers, O., Schmid, M., Cirpus, I., Strous, M., Bock, E., Kuenen, J.G. and Jetten,

M.S.M. (2002) Aerobic and anaerobic ammonia oxidizing bacteria - competitors or

natural partners? FEMS Microbiology Ecology, 39: 175-181.

Seitzinger, S.P. (1988) Denitrification in freshwater and coastal marine ecosystems: Ecological

and geochemical significance. Limnology and Oceanography, 33(4 part 2): 702-724.

Seitzinger, S.P. (1990) Denitrification in aquatic sediments. In: N.P. Revsbech and J. Sørensen

eds, Denitrification in Soil and Sediment. Plenum Press, New York, pp. 301-322.

Seitzinger, S.P. (1993) Denitrification measurements in aquatic sediments: A comparison of

three methods. Biogeochemistry, 23: 147-167.

Seitzinger, S.P., Nixon, S., Pilson, M.E.Q. and Burke, S. (1980) Denitrification and N2O

production in near-shore marine sediments. Geochimica et Cosmochimica Acta, 44: 1853-

1860.

Serventy, D.L. (1955) The fauna of the Swan River Estuary. In: Swan River Reference

Committee ed., Report on Pollution of the Swan River. Government Printer, Perth, pp. 70-

77.

Šimek, M. (2000) Nitrification in soil – terminology and methodology (review). Rostlinná

Výroba, 46(9): 385-395.

Simpson, C.J., Burt, J.S., Cary, J.L., D’Adamo, N.D., Masini, R.J. and Mills, D.A. (1996)

Southern Metropolitan Coastal Waters Study (1991-1994), Final Report. Department of

Environmental Protection, Perth, Western Australia. Report 17, 288 pp.

Slater, J.M. and Capone, D.G. (1989) Nitrate requirement for acetylene inhibition of nitrous

oxide reduction in marine sediments. Microbial Ecology, 17: 143-157.

Smil, V. (1999) Detonator of the population explosion. Nature, 400: 415.

Smith Jr., K.L., White, G.A., Laver, M.B. and Haugsness, J.A. (1978) Nutrient exchange and

oxygen consumption by deep-sea benthic communities: Preliminary in situ measurements.

Limnology and Oceanography, 23(5): 997-1005.

Smith, R.I. (1956) The ecology of the Tamar Estuary VII: Observations on the interstitial

salinity of intertidal muds in the estuarine habitat of Nereis diversicolor. Journal of the

Marine Biological Association of the United Kingdom, 35: 81-104.

Sørensen, J. (1978a) Denitrification rate in a marine sediment as measured by the acetylene

inhibition technique. Applied and Environmental Microbiology, 36: 139-143.

Sørensen, J. (1978b) Capacity for denitrification and reduction of nitrate to ammonia in a

coastal marine sediment. Applied and Environmental Microbiology, 35: 301-305.

Sørensen, J. (1980) Inhibition by sulfide of nitric and nitrous oxide reduction by denitrifying

Pseudomonas fluorescens. Applied and Environmental Microbiology, 39(1): 105-108.

Sørensen, J., Rasmussen, L.K. and Koike, I. (1987) Micromolar sulfide concentrations alleviate

acetylene blockage of nitrous oxide reduction by denitrifying Pseudomonas fluorescens.

Canadian Journal of Microbiology, 33: 1001-1005.

163

Page 178: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Spencer, R.S. (1956) Studies in Australian estuarine hydrology: II. The Swan River. Australian

Journal of Marine and Freshwater Research, 7: 193-253.

Sprent, J.I. (1987) The Ecology of the Nitrogen Cycle. Cambridge University Press, Cambridge,

151 pp.

Stephens, R. and Imberger, J. (1996) Dynamics of the Swan River Estuary: The seasonal

variability. Marine and Freshwater Research, 47: 517-529.

Stief, P. and de Beer, D. (2002) Bioturbation effects of Chironomus riparius on the benthic N-

cycle as measured using microsensors and microbiological assays. Aquatic Microbial

Ecology, 27: 175-185.

Strong, D.T. and Fillery, I.R.P. (2002) Denitrification response to nitrate concentrations in

sandy soils. Soil Biology and Biochemistry, 34: 945-954.

Strous, M, Fuerst, J.A., Kramer, E.H.M., Logemann, S., Muyzer, G., van de Pas-Schoonen,

K.T., Webb, R., Kuenen, J.G. and Jetten, M.S.M. (1999) Missing lithotroph identified as

new planctomycete. Nature, 400: 446-449.

Suess, E. (1980) Particulate organic carbon flux in the oceans – Surface productivity and

oxygen utilization. Nature, 288: 260-263.

Svensson, J.M. (1997) Influence of Chironomus plumosus larvae on ammonium flux and

denitrification (measured by the acetylene blockage- and the isotope pairing-technique) in

eutrophic lake sediment. Hydrobiologia, 346: 157-168.

Svensson, J.M., Carrer, G.M. and Bocci, M. (2000) Nitrogen cycling in sediments of the

Lagoon of Venice, Italy. Marine Ecology Progress Series, 199: 1-11.

Svensson, J.M., Enrich-Prast, A. and Leonardson, L. (2001) Nitrification and denitrification in a

eutrophic lake sediment bioturbated by oligochaetes. Aquatic Microbial Ecology, 23: 177-

186.

Swan River Trust (2000) The Febtuary 2000 Microcystis bloom. Accessed January 2001,

<http://www.wrc.wa.gov.au/srt/alert/moreinfo.html>.

Swan River Trust (2004a) Swan-Canning Estuary vertical contour plots. Accessed June 2004,

<http://www.wrc.wa.gov.au/srt/riverscience/profiles.html>.

Swan River Trust (2004b) Swan River Trust Media Statements. Accessed July 2004,

<http://www.wrc.wa.gov.au/srt/algalalert/media.html>.

Thamdrup, B. and Dalsgaard, T. (2000) The fate of ammonium in anoxic manganese oxide-rich

marine sediment. Geochimica et Cosmochimica Acta, 64(24): 4157-4164.

Thamdrup, B. and Dalsgaard, T. (2002) Production of N2 through anaerobic ammonium

oxidation coupled to nitrate reduction in marine sediments. Applied and Environmental

Microbiology, 68(3): 1312-1318.

Thompson, P.A. (1998) Spatial and temporal patterns of factors influencing phytoplankton in a

salt wedge estuary, the Swan River, Western Australia. Estuaries, 21(4B): 801-817.

164

Page 179: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Thompson, P.A. (2001) Temporal variability of phytoplankton in a salt wedge estuary, the

Swan-Canning Estuary, Western Australia. Hydrological Processes, 15: 2617-2630.

Thompson, P.A. and Hosja, W. (1996) Nutrient limitation of phytoplankton in the Upper Swan

River Estuary, Western Australia. Marine and Freshwater Research, 47: 659-667.

Thompson, P.A., Adeney, J. and Gerritse, R. (1997) Phytoplankton in the Swan River: Research

results 1993 to 1996 and management implications. In: J.R. Davis ed., Managing Algal

Blooms: Outcomes from CSIRO’s Multi-divisional Blue-green Algal Program. CSIRO

Land and Water, Canberra, pp. 15-30.

Thorson, G. (1957) Bottom communities (sublittoral or shallow shelf). In: J.W. Hedgpeth ed.,

Treatise on Marine Ecology and Paleoecology. Volume 1. Ecology. The Geological

Society of America, Memoir 67. Waverly Press, Baltimore, pp. 461-534.

Tiedje, J.M. (1988) Ecology of denirification and dissimilatory nitrate reduction to ammonium.

In: A.J.B. Zehnder ed., Biology of Anaerobic Microorganisms. Wiley, New York, pp.

179-244.

Tilman, D. (1999) The ecological consequences of changes in biodiversity: A search for general

principles. Ecology, 80: 1455-1474.

Tilman, D., Knops, J., Wedin, D., Reich, P., Ritchie, M. and Siemann, E. (1997) The influence

of functional diversity and composition on ecosystem processes. Science, 277: 1800-1802.

Timmermann, K., Christensen, J.H. and Banta, G.T. (2002) Modeling of advective solute

transport in sandy sediments inhabited by the lugworm Arenicola marina. Journal of

Marine Research, 60: 151-169.

Tobias, C.R., Anderson, I.C., Canuel, E.A. and Macko, S.A. (2001) Nitrogen cycling through a

fringing marsh-aquifer ecotone. Marine Ecology Progress Series, 210: 25-39.

Treadwell, A.L. (1929) New species of polychaetous annelids in the collections of the American

Museum of Natural History, from Porto Rico, Florida, Lower California, and British

Somaliland. American Museum Novit., 392: 1-13.

Trimmer, M., Nedwell, D.B., Sivyer, D.B. and Malcolm, S.J. (2000a) Seasonal benthic organic

matter mineralisation measured by oxygen uptake and denitrification along a transect of

the inner and outer River Thames estuary, UK. Marine Ecology Progress Series, 197:

103-119.

Trimmer, M., Nedwell, D.B., Sivyer, D.B. and Malcolm, S.J. (2000b) Seasonal organic

mineralisation and denitrification in intertidal sediments and their relationship to the

abundance of Enteromorpha sp. and Ulva sp. Marine Ecology Progress Series, 203: 67-

80.

Trimmer, M., Nicholls, J.C. and Deflandre, B. (2003) Anaerobic ammonium oxidation

measured in sediments along the Thames Estuary, United Kingdom. Applied and

Environmental Microbiology, 69(11): 6447-54.

165

Page 180: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Twilley, R.R. and Kemp, W.M. (1986) The relation of denitrification potentials to selected

physical and chemical factors in sediments of Chesapeake Bay. In: D.A. Wolfe ed.,

Estuarine Variability. Academic Press, Orlando, pp. 277-293.

Twomey, L. and John, J. (2001) Effects of rainfall and salt-wedge movement on phytoplankton

succession in the Swan-Canning Estuary, Western Australia. Hydrological Processes, 15:

2655-2669.

Vanderborght, J.-P. and Billen, G. (1975) Vertical distribution of nitrate concentration in

interstitial water of marine sediments with nitrification and denitrification. Limnology and

Oceanography, 20(6): 953-961.

Vanderborght, J.-P., Wollast, R. and Billen, G. (1977) Kinetic models of diagenesis in disturbed

sediments. Part 2. Nitrogen diagenesis. Limnology and Oceanography, 22(5): 794-803.

Vedel, A. (1998) Phytoplankton depletion in the benthic boundary layer caused by suspension-

feeding Nereis diversicolor (Polychaeta): Grazing impact and effect of temperature.

Marine Ecology Progress Series, 163: 125-132.

Viney, N.R. and Sivapalan, M. (2001) Modelling catchment processes in the Swan-Avon river

basin. Hydrological Processes, 15: 2671-2685.

Walker, B.H. (1992) Biodiversity and ecological redundancy. Conservation Biology, 6(1): 18-

23.

Wallace, J. (1977) The macrobenthic invertebrate fauna of Pelican Rocks, March-April 1977.

Unpublished Report. Department of Conservation and Environment, Perth, 10pp.

Walter, H.M., Keeney, D.R. and Fillery, I.R. (1979) Inhibition of nitrification by acetylene. Soil

Science Society of America Journal, 43: 195-196.

Ward, B.B. (1996) Nitrification and denitrification: Probing the nitrogen cycle in aquatic

environments. Microbial Ecology, 32: 247-261.

Ward, B.B. (2000) Nitrification and the marine nitrogen cycle. In: D.L. Kirchman ed., Microbial

Ecology of the Oceans. Wiley, New York, pp. 427-453.

Water and Rivers Commission (2001) Swan-Canning Cleanup Program Environmental

Performance and Progress Reporting (27049): Swan River Estuary – Weekly Report for

18th June 2001. Aquatic Sciences Branch, Waters and Rivers Commission, Perth, Western

Australia.

Watson, P.G., Frickers, P.E. and Howland, R.J.M. (1993) Benthic fluxes of nutrients and some

trace metals in the Tamar Estuary, SW England. Netherlands Journal of Aquatic Ecology,

27(2-4): 135-146.

Webb, A.P. and Eyre, B.D. (2004a) The effect of natural populations of the burrowing and

grazing soldier crab (Mictyris longicarpus) on sediment irrigation, benthic metabolism

and nitrogen fluxes. Journal of Experimental Marine Biology and Ecology, 309: 1-19.

166

Page 181: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Webb, A.P. and Eyre, B.D. (2004b) Effect of natural populations of burrowing thalassinidean

shrimp on sediment irrigation, benthic metabolism, nutrient fluxes and denitrification.

Marine Ecology Progress Series, 268: 205-220.

Webb, K.L. (1981) Conceptual models and processes of nutrient cycling in estuaries. In: B.J.

Neilson and L.E. Cronin eds, Estuaries and Nutrients. Humana Press, Clifton, New Jersey,

pp. 25-46.

Weiss, R.F. and Price, B.A. (1980) Nitrous oxide solubility in water and seawater. Marine

Chemistry, 8: 347-359.

Welsh, D.T. (2003) It’s a dirty job but someone has to do it: The role of marine benthic

macrofauna in organic matter turnover and nutrient recycling to the water column.

Chemistry and Ecology, 19(5): 321-342.

Welsh, D.T. and Castaldelli, G. (2004) Bacterial nitrification activity directly associated with

isolated benthic marine animals. Marine Biology, 144: 1029-1037.

Welsh, D.T., Bartoli, M., Nizzoli, D.,Castaldelli, G., Riou, S.A. and Viaroli, P. (2000)

Denitrification, nitrogen fixation, community primary productivity and inorganic-N and

oxygen fluxes in an intertidal Zostera noltii meadow. Marine Ecology Progress Series,

208: 65-77.

Widdicombe , S. and Austen, M.C. (1998) Experimental evidence for the role of Brissopsis

lyrifera (Forbes, 1841) as a critical species in the maintenance of benthic diversity and the

modification of sediment chemistry. Journal of Experimental Marine Biology and

Ecology, 228: 241-255.

Williams, P.J. le B. (1973) The validity of the application of simple kinetic analysis to

heterogeneous microbial populations. Limnology and Oceanography, 18: 159-165.

Wilson, B.R. and Allen, G.R. (1987) Major components and distribution of marine fauna. In:

G.R. Dyne and D.W. Walton eds, Fauna of Australia, Volume 1: General Articles.

Australian Government Publishing Service, Canberra, pp. 43-68.

Wilson, R.S. (2000) Class Polychaeta: Nereididae. In: P.L. Beesley, G.J.B. Ross and C.J.

Glasby, eds, Polychaetes and Allies: The Southern Synthesis. Fauna of Australia, Volume

4a: Polychaeta, Myzostomida, Pogonophora, Echiura, Sipuncula. CSIRO Publishing,

Melbourne, pp. 138-141.

Wilson, R.S., Bakken, T. and Glasby, C.J. (2003b) Nereididae (Polychaeta) – A DELTA

database of genera and Australian species. In: R.S. Wilson, P.A. Hutchings and C.J.

Glasby eds, Polychaetes: An Interactive Identification Guide. CD-ROM and manual.

CSIRO Publishing, Melbourne.

Wilson, R.S., Hutchings P.A. and Glasby C.J. (2003a) Polychaetes: An Interactive Identification

Guide. CD-ROM and manual. CSIRO Publishing, Melbourne.

Winer, B. J. (1971). Statistical Principles in Experimental Design. McGraw-Hill, New York.

167

Page 182: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Winston, J. E., and Anderson, F. E. (1971). Bioturbation of sediments in a northern temperate

estuary. Marine Geology, 10: 39-49.

Wrage, N., Velthof, G.L., van Beusichem, M.L. and Oenema, O. (2001) Role of nitrifier

denitrification in the production of nitrous oxide. Soil Biology and Biochemistry, 33:

1723-1732.

Yin, S.X., Chen, D., Chen, L.M. and Edis, R. (2002) Dissimilatory nitrate reduction to

ammonium and responsible microorganisms in two Chinese and Australian paddy soils.

Soil Biology and Biochemistry, 34: 1131-1137.

Yoshinari, T., and Knowles, R. (1976). Acetylene inhibition of nitrous oxide reduction by

denitrifying bacteria. Biochemical and Biophysical Research Communications, 69: 705-

710.

Yoshinari, T. and Knowles, R. (1978) Inhibition by acetylene of ammonia oxidation in

Nitrosomonas europaea. FEMS Microbiology Letters, 4: 319-321.

Yoshinari, T., Hynes, R. and Knowles, R. (1977) Acetylene inhibition of nitrous oxide

reduction and measurement of denitrification and nitrogen fixation in soil. Soil Biology

and Biochemistry, 9: 177-183.

Zehr, J.P. and Ward, B.B. (2002) Nitrogen cycling in the ocean: New perspectives on processes

and paradigms. Applied and Environmental Microbiology, 68(3): 1015-1024.

Zeitzschel, B. (1980) Sediment-water interactions in nutrient dynamics. In: K.R. Tenore and

B.C. Coull eds, Marine Benthic Dynamics. University of South Carolina Press, Columbia,

pp. 195-218.

Zopfi, J., Kjær, T., Nielsen, L.P. and Jørgensen, B.B. (2001) Ecology of Thioplaca spp.: Nitrate

and sulfur storage in relation to chemical microgradients and influence of Thioplaca spp.

on the sedimentary nitrogen cycle. Applied and Environmental Microbiology, 67(12):

5530-5537.

168

Page 183: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

APPENDIX A

THE SALT WEDGE: TEMPORAL VARIATION AND

STRATIFICATION OF SALINITY IN THE SWAN RIVER ESTUARY

Stephens and Imberger (1996) and Kurup et al. (1998) consider the hydrodynamics of the Swan

River Estuary system in relation to the annual upstream migration of tidally forced marine water

as a salt wedge, when freshwater discharge decreases at the cessation of winter rains. Kurup et

al. (1998) describe the hydrology of the estuary in relation to three phases of upstream salt

wedge positioning controlled primarily by river discharge. Slightly modified, the typical

hydrological phases are:

(1) Salt wedge (marine) dominated in summer and autumn (January to May), when

freshwater discharge is low. High salinity water extends from the mouth to between the

confluences of the estuary with Helena River and Ellen Brook, 40 to 60 km upstream.

(Figure A1).

(2) (a) Salt wedge waning during early winter (June to July); or, (b) salt wedge

emplacement (propagation) during late spring and early summer (November to

December); when freshwater discharges are low. Fresh to brackish waters extend from

the upper estuary to surficial (<2 m deep) Perth Water (20 to 25 km upstream of the

mouth), whilst high salinity bottom waters (>2 m deep) may extend from Perth Water to

10-15 km into the upper estuary. Thus, the resulting halocline from upstream bottom

sediments to downstream surface waters, i.e. the salt wedge, may extend for >10 km.

(Figure A1).

(3) Salt wedge absent (freshwater dominated) in late winter and early spring (August to

October), when freshwater discharge is high. Fresh to brackish waters extend from the

upper estuary across the surficial (<5m deep) waters of the main basin to the Fremantle

Sill (5 km upstream of the mouth). (Figure A1).

This general conceptualisation allows effective consideration of important variations in

hydrology occurring within the estuary at a more localised geographic scale, and also due to

inter-annual variation in rainfall and resultant freshwater discharge.

The phase-descriptions are mainly relevant to the shallow (<5 m deep) areas of the estuary, i.e.

most of the upper estuary and outer (near-shore) edges of the main basin. Significant

stratification can occur in the deeper areas of both the upper estuary and the main basin, and

169

Page 184: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

deeper waters can remain marine year-round. Tidally forced high salinity water enters the upper

estuary in spring to summer when freshwater discharge in the upper-stream regions decreases

below 6 m3 s-1 (Stephens and Imberger, 1996), i.e. during phase 2b - salt wedge emplacement.

Waning discharge below this flow rate allows the deeper waters of the upper estuary to move

towards phase 1, marine dominated, although some freshwater stratification may remain in the

upper reaches with persistent freshwater flow. Increasing freshwater discharge (>6 m3 s-1)

during winter (phase 3), flushes the surficial salt water from the upper estuary (Stephens and

Imberger, 1996). However, high salinity water and thus stratification will remain in the deeper

areas unless discharge exceeds 23-50 m3 s-1, which is sufficient to flush all saline water from the

upper estuary (Stephens and Imberger, 1996; Kurup et al., 1998).

Phase 1 – Salt wedge dominated ~ January to May

Phase 2a – Salt wedge waning ~ June to July

Phase 3 – Salt wedge absent ~ August to October

Phase 3 – Salt wedge propagation ~ November to December

0 5 10 15 20 25 30 35 40Distance from mouth (km)

Dep

th (m

)

Figure A1: Phases of salinity-depth contours within the Swan River Estuary throughout the

course of a typical year of freshwater discharge. (Adapted from Kurup et al., 1998; Swan River Trust, 2004a).

170

Page 185: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Deep waters (>5 m) of the main basin typically remain marine year-round, but are overlain by

seaward flowing fresh to brackish water in late winter to early spring, i.e. during phase 3

(Stephens and Imberger, 1996). When the depth of freshwater discharge exceeds the depth of

the Fremantle Sill (~5 m), upstream tidal movement of marine water into the main basin is

limited or nonexistent (Figure A2; Stephens and Imberger, 1996). The presence of the halocline

at a depth below that of the sill entombs the deeper high salinity water of the main basin, and

disallows marine exchange. Persistence of such stratification and lack of mixing, combined

with sedimentary biochemical oxygen demand, often causes significant deoxygenation of

bottom waters in the main basin during late winter to early spring (Figure A2; Hodgkin, 1987;

Stephens and Imberger, 1996). Similarly, locally stratified deep saline pockets of the upper

estuary may experience deoxygenation during the same period of freshwater overlay. As

freshwater discharge recedes and the depth of out-flowing water over Fremantle Sill is reduced,

richly oxygenated marine water is tidally forced into the deeper part of the main basin, and

eventually upstream as a wedge, thereby relieving anoxia (Figure A2; Stephens and Imberger,

1996). Additionally, lack of mixing within the highly stratified leading edge of the saline

wedge may also cause transitory yet significant deoxygenation of bottom waters during phase

2a (early winter) and phase 2b (late spring to early summer), as the salt wedge respectively

propagates and wanes (Figure A3).

Dep

th (m

)

Deoxygenated basin water

Surface discharge layer

Marine water

2 km

(a) (b)

(c) (d)

Figure A2: Seasonal entombment, deoxygenation and recovery of deep water within the main

basin of the Swan River Estuary. (a) Freshwater overflow in late winter to early spring at a depth greater than that of the Fremantle Sill; (b) subsequent decrease in freshwater discharge and depth of overflow allows tidal intrusion of oxygenated marine water; (c) further subsidence in freshwater discharge and resultant tidal flushing of the deep basin with oxygenated water; (d) summer-time absence of freshwater influence and upstream migration of marine water.

171

Page 186: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Dep

th (m

)

0 5 10 15 20 25 30 35 40

Distance from mouth (km)

(a)

(b)

Figure A3: Salinity (a) and dissolved oxygen (b) depth profiles in the Swan River Estuary during July 1997. Lack of circulation in the bottom waters of the waning salt wedge, including the upstream deeper pockets, causes marked deoxygenation. (Adapted from Swan River Trust, 2004a).

The seasonality of rainfall and resultant freshwater discharge governs both the salt wedge and

other stratification dynamics of the Swan River Estuary. Kurup et al. (1998) consider that

freshwater inflow is the most important factor affecting salt wedge position in the estuary.

Surficial freshwater flow is also the causal agent of winter/spring stratification and associated

anoxia in both the main basin and, when present, in the upper estuary. It is therefore apparent

that inter-annual variation in rainfall and freshwater discharge has profound potential to alter the

hydrodynamics of the estuary. To highlight the dependence of the hydrology of the estuary on

freshwater discharge, Kurup et al. (1998) considered salt wedge dynamics within the Swan

River Estuary during very wet (1974) and very dry (1993) years.

Subsequent to April 1974, the estuary received greater than average river inflow which persisted

until the end of the year (Kurup et al., 1998). As a result, the salt wedge had been flushed from

the upper estuary by July (i.e. early initiation of phase 3) and the estuary remained freshwater

dominated until the following summer (i.e. salt wedge emplacement, phase 2b, was inhibited

and the estuary remained in phase 3 throughout the remainder of the year). Additionally, it was

noted that the magnitude of inflow by July (>50 m3 s-1) was sufficient to flush high salinity

water from the deeper upstream pockets, thereby diminishing the potential for salinity

stratification and deoxygenation in these areas. Since Kurup et al.’s (1998) study area did not

172

Page 187: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

include the main basin, the effect of the extended period of heightened river inflow upon

stratification in this area is uncertain. Presumably, the period of salinity stratification and

perhaps anoxia within the main basin would likewise have been extended. In a similar year of

above average rainfall (1992), Stephens and Imberger (1996) demonstrated that freshwater

discharge maintained stratification and deoxygenation within the main basin at least until early

October.

In contrast, river inflow by July 1st of the dry year 1993 was insufficient (<20 m3 s-1) to flush the

salt wedge from the upper estuary (Kurup et al., 1998). A subsequent burst in river discharge

(>100 m3 s-1) flushed the salt wedge to downstream of the Narrows by mid-September (i.e. late

initiation of phase 3, freshwater dominance); however, a resumption of low inflow resulted in

salt wedge emplacement 10 km upstream of Heirisson Island by mid-spring (i.e. earlier than

usual initiation of phase 2b, salt wedge propagation in the upper estuary). By early summer

(October) the salt wedge had migrated beyond the upstream boundary of their study, i.e. the

estuary’s confluence with Jane Brook (~50 km upstream of the mouth). As such, the Swan

River Estuary was dominated by marine conditions (phase 1) for most of this very dry year.

The effect of the low freshwater inflow upon upper estuary and main basin stratification is

uncertain. While the late winter burst of freshwater discharge was enough to flush the upper

reaches of saline water, the preceding period of low inflow may have stratified the deeper

pockets of this area and caused substantial deoxygenation. Presumably, the lack of freshwater

influence over the main basin throughout most of the year would have allowed greater than

usual tidal exchange, thereby alleviating winter stratification and deep water anoxia.

Aside from deviations in winter rainfall, large-scale unseasonal summer rainfall events can also

alter the typical hydrology of the Swan River Estuary. For example, heavy rain in early January

2000 saturated much of the Avon catchment (Atkins et al., 2001). One week later this was

followed by further atypical heavy rainfall associated with the aftermath of a northern tropical

cyclone (Atkins et al., 2001). As a result, the Swan River Estuary was flushed with enough

freshwater to fill it five times over (Atkins et al., 2001). Whilst the estuary had settled

previously into its usual marine dominated summer-time state (phase 1), by the end of January

the entire estuary had reverted to a winter-like state of freshwater dominance (phase 3), with

marked stratification of the main basin (Atkins et al., 2001). These freshwater conditions,

combined with a large runoff-associated nutrient input, triggered a record bloom of toxic blue-

green freshwater algae (see Chapter 2, Section 2.1.3). Subsequent easing of freshwater

discharge allowed the tidal re-entrance of marine water and upstream propagation of the salt

wedge, such that the system had reverted back to its typical marine dominated state by the

beginning of March.

173

Page 188: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

The influence of seasonality upon the hydrology of the Swan River Estuary also extends to

water temperature. Within the Mediterranean climate, average temperature of surface water

(<5 m deep) overlying the main basin typically ranges from 12-14 °C in winter to 22-24 °C in

summer (Figure A4; Hodgkin, 1987). The annual range in average water temperature of the

shallow upper estuary is greater, from 10-13 °C in winter to 28-30 °C in summer (Figure A4;

Hodgkin, 1987). Very shallow areas of the estuary (<2 m deep; typically where benthic fauna is

most abundant), such as nearshore and at several upstream locations, may experience even

greater seasonal variation in water temperature (Hodgkin, 1987). The upper estuary and

shallower areas are also subject to diurnal fluctuations in temperature of 4-5 °C (Kurup et al.,

1998). Deeper water of the main basin is less variant, seldom falling below 15 °C in winter or

rising above 22 °C in summer (Figure A4; Hodgkin, 1987). Haloclines generated by the

trapping of high salinity water under cold freshwater winter flow (i.e. in the deep basin, deep

pockets of the upper estuary and at the leading/waning edge of the salt wedge), are also subject

to reversed thermoclines during the period of freshwater influence. The difference in salinity,

and thus density, maintains the warmer marine water at depth (Figure A4). Salinity, not

temperature, overwhelmingly drives the density variations within the Swan River Estuary

(Kurup et al., 1998).

Summer

Winter

Temperature(ºC)

Figure A4: Typical temperature-depth profiles of the Swan River Estuary in winter (July) and summer (December). Adapted from Swan River Trust (2004a).

174

Page 189: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

APPENDIX B

ALGAL BLOOM SUCCESSION IN THE SWAN RIVER ESTUARY

The annual succession of population growth of various algal species within the Swan River

Estuary has been an on-going topic of relevance to the residents of Perth since settlement.

Hodgkin and Vicker (1987) detail the history of algal blooms in the Swan River Estuary, which

is briefly summarised below prior to a consideration of the present day annual succession of

bloom-forming phytoplankton species.

Pre-dating the problems associated with blooming phytoplankton species in recent history, the

Swan River Estuary was subject to recurrent, large populations of macroalgae. The following is

recounted from Hodgkin and Vicker (1987). From at least the 1870s, shore-line accumulation

and decay of dense mats of macroalgae were a major source of complaint by Perth residents.

During spring and summer, green algae (e.g. Cladophora, Chaetomorpha and Rhizoclonium

spp.) and filamentous diatoms (e.g. Melosira spp.) grew rapidly in the warm, calm and shallow

waters of the main basin and Perth Water. Subsequent detachment and drift allowed

accumulation and then degradation upon the estuary’s beaches and sand banks. Newspaper

articles of the 1870s record of swans avoiding “large masses of algae floating near the

causeway” and a “large collection of seaweed in the river near Mill Street Jetty”. At the turn of

the century, “rotting weed was being collected and carted away from beaches as far away as

Claremont” (some distance from the main settlement for the period). In 1913 it was recorded

that “some half mile of foreshore is covered with filthy green vegetable growth”. Not only did

these accumulations provide a blanket cover of the estuary’s beaches and a barrier to

recreational enjoyment, their ensuing decay produced hydrogen sulphide and other malodorous

gases that were the major cause of complaint. The stench together with the associated flies,

maggots and other detritivores were also perceived to be a public health threat. Up to 400

tonnes of seaweed was manually removed from the estuary per year, until the 1960s.

Reclamation of sheltered areas and an increase in river flow (resultant from channel dredging

and catchment changes) are most likely the causes of a subsequent and significant decrease in

annual macroalgal accumulations. Some collection of algal weed still occurs, but in contrast to

the magnitude of the problem prior to the 1970s, the accumulations are small and only of minor

public nuisance. In summation, large algal accumulations within the Swan River Estuary pre-

date its current eutrophic status and perhaps even European settlement.

It is the increasingly eutrophic status of the estuary and shift from productive macroalgal

species towards both toxic and non-toxic phytoplankton species, together with the consequences

of an intensification of frequency and severity of blooms, which is presently driving

environmental and public concern. The modern annual succession pattern of phytoplankton

within the estuary is described below, in addition to the conditions necessary for bloom-forming

species.

175

Page 190: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Twomey and John (2001: p. 2656) state that “inter-annual variation of phytoplankton succession

can be large, but in general the patterns are broadly repeated from one year to the next”. The

‘normal’ succession and blooming of phytoplankton species within the Swan River Estuary is

defined for those years of typical rainfall, freshwater discharge and salt-wedge migration (see

Chapter 2, Section 2.1.2), and is as follows. When freshwater discharge is high during winter,

the upper estuary is dominated by dislodged epiphytic and periphytic diatoms (e.g. Synedra

ulna, Synedra pulchella, Surirella ovalis, Eunotia pectinalis, Melosira variens and Synedra

fasciculata), with occasional mild blooms of euplanktonic diatoms (e.g. Cyclotella atomus,

Cyclotella meneghiniana, Thalassiosira weissflogii, Nitzschia closterium and Cyclotella striata)

(John, 1987; Twomey and John, 2001). The phytoplankton community of the main basin is also

comprised of these species but is mixed with marine diatom species (e.g. Coscinodiscus

asteromphalus, Coscinodiscus janischii and Rhizosolenia setigera) (John, 1987; Twomey and

John, 2001). Dinoflagellates are typically absent from the upper estuary and within the main

basin both dinoflagellates and cryptophytes are present but in low biomass compared to diatoms

(Twomey and John, 2001). Although normal winter freshwater discharge delivers high

concentrations of dissolved inorganic nitrogen (and it is the only time that nitrate rather than

ammonium is the dominant species), it is concurrent with a high flow rate, particulate load,

discolouration and general increase in turbidity of the estuary (Thompson, 1998, 2001).

Combined with low temperatures, decreased day length, greater cloud cover and lower surface

irradiance, the normal, winter freshwater conditions are comprised of low phytoplankton

primary production and biomass (Thompson, 2001).

The spring phytoplankton community is dominated by a chlorophyte (Chlamydomonas

globulosa) in the upper estuary and a marine diatom (Skeletonema costatum) in the main basin,

and both species frequently bloom under typical hydrological conditions (John, 1987;

Thompson, 1998; Twomey and John, 2001). Dinoflagellate species (Prorocentrum spp.) may

bloom and precede or accompany diatom blooms in the main basin (John, 1987; Twomey and

John, 2001). As the salt wedge migrates upstream towards the end of spring to early summer so

too does the dominant diatom species Skeletonema costatum, accompanied by other marine

species of centric diatoms (e.g. Coscinodiscus centralis, Rhizosolenia setigera, Coscinodiscus

janischii, Coscinodiscus oculus-iridius and Lithodesmium undulatum) (John, 1987; Twomey

and John, 2001). Within the main basin Skeletonema costatum may be replaced by

Rhizosolenia setigera as the dominant diatom species by the end of spring, along with several

Chaetoceros spp., other marine diatoms and less frequently Prorocentrum spp. dinoflagellates

(John, 1987; Twomey and John, 2001). Spring phytoplankton blooms are often the most prolific

(chlorophyll-a biomass in the upper estuary may reach 250-300 μg l-1), occurring at the time of

year when day length period and water temperatures are increasing, whilst freshwater discharge

is waning but still providing an allochthonous nutrient supply (mainly derived from agricultural

drainage) (John, 1994; Douglas et al., 1997).

176

Page 191: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Marine influence of the salt wedge during summer allows the estuary-wide persistence and

dominance of marine diatom species (e.g. Skeletonema costatum, Coscinodiscus spp.,

Thalassiosira spp., Chaetoceros spp., Lithodesmium spp., Rhizosolenia spp., Asterionella spp.

and Actinocyclus spp.) (John, 1987; Twomey and John, 2001). Phytoplankton community

composition is similar in the upper estuary and main basin; however, net primary production

and the susceptibility to blooms is far greater in the upper estuary (Twomey and John, 2001).

Periodic blooms of dinoflagellates (e.g. Scrippsiella sp.) may be observed throughout the

estuary, although their magnitude and frequency is typically greater in the main basin

(Thompson, 1998; Twomey and John, 2001). Towards the end of summer (February),

cryptophytes may be observed in the upper estuary (Twomey and John, 2001). Net primary

production within the Swan River Estuary is typically highest during summer when water

temperatures, growing period and water clarity are high and flows are low (Thompson, 1998).

Both algal biomass and production are particularly high in the upper estuary whereby the

relatively warm shallow waters promote the greatest growth (Thompson, 1998). It is interesting

that the summer phytoplankton blooms occur during the period when runoff (allochthonous)

nutrient inputs are very low and combined available dissolved inorganic nitrogen concentrations

(nitrate + nitrite + ammonium) are at an annual minimum (due to depletion by the preceding

spring-time algal bloom) (Thompson, 1998, 2001). Ammonium is the dominant inorganic

nitrogen source from November to May, and occurs in highest concentrations near the sediment

surface (Thompson, 1998). During this period phosphate concentrations are relatively high and

there is strong potential for nitrogen limitation (Thompson, 1998, 2001). Thompson (1998,

2001) suggests that summer phytoplankton blooms rely on, and are regulated by, the supply of

autochthonous nitrogen (ammonium) recycled within sediments or in the water column near the

sediment surface (perhaps including that supplied by groundwater discharge). There is evidence

that motile species actively migrate over a diurnal period to maximise night-time near-bottom

nitrogen uptake and daytime near-surface photosynthesis (Thompson, 1998; Horner Rosser and

Thompson, 2001).

With the increasing marine domination of the estuary during autumn there is a continued

presence and high abundance of marine diatoms. Within the main basin, diatom blooms (e.g. of

Skeletonema costatum, Coscinodiscus spp.,) may be accompanied by cryptophytes (Twomey

and John, 2001). The phytoplankton composition of the upper estuary is also characteristic of

marine waters and in addition to high densities of diatom species (e.g. Thalassiosira spp.,

Skeletonema costatum, Rhizosolenia spp., Chaetoceros lorenzianus, Lithodesmium undulatum

and Coscinodiscus spp.), dinoflagellate species (e.g. Scripsiella spp., Gymnodinium spp.,

Protoperidinium spp. and Prorocentrum spp.) are also responsible for blooms (Twomey and

John, 2001). The euplanktonic diatoms more characteristic of the winter period (e.g. Amphora

spp., Cyclotella spp. and Nitzschia spp.) may also be noticeable in the upper estuary during the

autumn period (Twomey and John, 2001), perhaps associated with early rainfall events of

limited discharge. With the onset of winter rains and subsequent freshwater discharge the

177

Page 192: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

biomass of marine species declines, periphytic diatoms are dislodged and the phytoplankton

community composition reverts to the winter-time condition under freshwater influence (John,

1987; Twomey and John, 2001). During autumn and prior to the winter freshwater discharge of

land-derived nutrients, phytoplankton are similarly reliant on autochthonous nitrogen sources as

detailed for the summer period above. While ammonium concentration of bottom waters

typically increases to a peak in late autumn (Thompson, 1998), phytoplankton biomass

generally wanes from early to late autumn (Twomey and John, 2001). This suggests that factors

such as decreasing day lengths, water column irradiance and temperatures more than counteract

the increased nutrient conditions that are potentially conducive to heightened phytoplankton

growth (Thompson, 1998).

It is inter-annual variation in hydrology and associated nutrient loads that may cause

phytoplankton assemblages to deviate from normal succession and bloom patterns, and

potentially trigger atypical bloom events in the Swan River Estuary (Twomey and John, 2001).

While it is salinity preference and tolerance that principally determines the distribution of

phytoplankton species in the estuary (John, 1987), it is rainfall and freshwater discharge that

governs both the position of the salt wedge and allochthonous nutrient supply (Thompson,

2001). As such, inter-annual variation in rainfall drives deviations from the typical occurrence,

timing and persistence of algal blooms (Thompson, 2001). Thompson (2001) has tabulated a

predicted set of phytoplankton responses to variations in the magnitude and timing of annual

rainfall, to be compared to the normal years of rainfall and phytoplankton succession as

described above. While acknowledging that the data set is limited, consideration provides

insight into the factors governing the dynamics of problematic and nuisance phytoplankton

species, providing some predictive power and a platform to help facilitate better the

management of algal blooms. The reader is referred to the source (Thompson, 2001) for a more

detailed account, but given below are some examples that highlight the consequences of inter-

annual variation in rainfall on phytoplankton succession and atypical bloom events.

Primary production in the Swan River Estuary is typically lowest during winter. Significant

winter blooms are restricted to marine diatom species (e.g. Skeletonema costatum) and occur

only when rainfall and freshwater discharge is less than normal (Thompson, 2001). During the

winter of dry years, the lack of freshwater flow and upstream persistence of the salt wedge

allows the continuance of autumn-like conditions for phytoplankton, as described above.

During years of normal winter flow, Thompson (1998) suggests that part of the reason for low

productivity in the estuary is that much of the allochthonous nutrients associated with

freshwater discharge are transported out of the system. He notes that Simpson et al. (1996)

have detected a winter plume of low-salinity, high-nutrient water extending for some tens of

kilometres into the Indian Ocean from the estuary’s mouth. If correct, the interpretation implies

that the Swan River Estuary is normally spared from some of the more serious consequences of

high nutrient loading due to this flow-through effect and the winter timing of the high nutrient

178

Page 193: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

pulse (and associated factors inhibiting algal growth, i.e. high turbidity and decreased day

length, water temperatures and irradiance; see above) (Thompson, 1998). Drier-than-normal

winters and reduced flow may cause more of the allochthonous nutrient load to be retained

within the estuary, thereby promoting diatom blooms. Conversely, algal blooms may not be

expected during wetter-than-usual winters due to sustained flow and non-ideal physical

conditions for algal growth, in addition to transport of most of the land-derived nutrients from

the estuary (Thompson, 2001).

During normal years of freshwater discharge, spring blooms of diatoms, chlorophytes and

dinoflagellates seem able to capture and utilise much of the allochthonous nutrient supply

(Thompson, 1998). This may not be the case if a wetter than usual winter and/or spring flushes

the land-derived nutrients from the estuary (Thompson, 2001). Further, Thompson (2001)

hypothesises that lack of normal spring blooms deprives the summer period of readily

recyclable nutrients (as labile phytoplankton detritus), and precludes typical summer blooms.

Such a scenario occurred in the summer of both 1995/96 and 1996/97 (Thompson, 2001).

Alternatively, a drier-than-normal winter and spring followed by unseasonable summer and/or

autumn rains seem likely to facilitate bloom events in the latter periods (Thompson, 2001). It is

hypothesised that decreased flow during winter/spring allows storage of nutrients either on land

or in the upper estuary, whilst subsequent large bursts of rain would allow discharge of high

concentration nutrients at an appropriate rate for algal uptake, and at a time of year

(summer/autumn) when the physical conditions required for primary production are maximal.

An example of such an event is described in Chapter 2, Section 2.1.3. It details the recent case

of the largest and most significant phytoplankton bloom within the recorded history of the

Estuary, exemplified by a two-week total closure of the estuary to the public. While the rainfall

factors causing the event were atypical and extreme, its occurrence illustrates the increasingly

eutrophic status of the estuary, heightened potential for (and intensification of) such blooms, as

well as the highly detrimental environmental and socio-economic consequences of

eutrophication.

179

Page 194: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,
Page 195: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

APPENDIX C

THE SWAN RIVER ESTUARY’S BENTHIC MACROFAUNAL

COMMUNITY

Characterisation of the benthic macrobenthos of the Swan River Estuary has been the subject of

relatively few published studies (e.g. Chalmer et al., 1976; Kanandjembo et al., 2001; Pennifold

and Davis, 2001), but some further information may be found in unpublished theses and

government reports (e.g. Serventy, 1955; McShane, 1977; Wallace, 1977; Rose, 1994;

Kanandjembo, 1998). In this Appendix I present a brief review of these studies and the current

status of knowledge pertaining to benthic macrofaunal community composition and trophic

relations in the Swan River Estuary.

A review conducted by Rose (1994) identified that 134 benthic macrofaunal species have been

noted in the Swan River Estuary, including both continuous and temporary inhabitants. At a

high taxonomic level and similar to other estuaries, the Polychaeta, Mollusca and Crustacea

together comprise the vast majority of species and individuals. T he best-studied regions are the

shallows (overlying water <2 m deep) of the main basin and upper estuary wherein polychaetes

of the families Nereididae, Spionidiae, Capitellidae, Sabellidae and Orbiniidae contribute 46-

47% of the total abundance (Rose, 1994; Kanandjembo et al., 2001). Molluscan bivalves and

gastropods together contribute 14-43% towards abundance, while amphipod, decapod and

isopod crustaceans contribute 7-39%, depending on the region (Rose, 1994; Kanandjembo et

al., 2001).

Regarding mollusc species, Chalmer et al. (1976) document that the community of the lower

estuary is dominated by species of marine affinity, both temporary (64 stenohaline species) and

continuous (25 marine euryhaline species). They argue that the high molluscan diversity (89

species) demonstrates a distinctiveness of the lower estuary and gives weight to the three-part

biotopic subdivision of the Swan River Estuary (see Section 2.2.1). The group is characterised

by great attenuation above the lower estuary, with only 6 of the marine euryhaline species

capable of living continuously in the middle estuary and only two species having made

‘temporary incursions’ into the upper estuary (during sustained periods of low river discharge)

(Chalmer et al., 1976; Hodgkin, 1987). The molluscan fauna of the middle and upper estuary is

generally characterised by 7 true euryhaline species (4 of which may extend into the lower

estuary during years of strong winter floods). At any one time, abundance and diversity of these

true euryhaline species typically decreases in the upper reaches (Chalmer et al., 1976; Hodgkin,

1987). The only true oligohaline mollusc is a freshwater gastropod, which is restricted to the

upper estuary. Stenohaline and marine euryhaline species from other taxonomic groups

181

Page 196: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

(including the Polychaeta) have not been so well studied, but similar broad patterns of

distribution are expected for the lower and middle estuary. For example, Chalmer et al. (1976)

further documented that two polychaete species (A. ehlersi and Diopatra sp.) and two

crustaceans (the crab Halicarcinus bedfordi and barnacle Balanus sp.) of marine affinity, each

resided in the lower estuary but at few or no locations in the middle and upper estuary.

The distinction between the middle and upper estuary may be further elucidated by comparing

the macrobenthic community studies of Rose (1994) and Kanandjembo (1998) (Table C1).

Succeeding thorough sampling of shallow water sites (<2 m deep) over a comparable period, the

former found 53 macrobenthic species within the middle estuary, while the latter revealed only

37 species from the upper estuary. During a brief sampling period in autumn, Wallace (1977)

similarly found a greater species richness in the middle estuary than the upper estuary, both for

shallow and deep waters (depth >2 m). Kanandjembo (1998) suggested that the lower species

richness of the upper estuary was due to a more variable, extreme and sometimes very low

salinity environment, which fewer euryhaline species were able to tolerate. For example, two

euryhaline crustaceans (the amphipod Grandidierella sp. and the isopod Tanais dulongii) were

abundant in the middle estuary but not present in the upper estuary (Table C1). In general, the

abundance of small crustaceans was much higher in the middle than upper estuary.

Nevertheless, some species were characterised as upper estuary specialists (e.g. the spionid

polychaetes Pseudopolydora sp. and Desdemona ornata, as well as oligochaete and chironomid

species), that were somewhat abundant in the upper estuary but not present in the middle

estuary (Table C1). Further, the abundance of the euryhaline bivalves Arthritica semen and

Xenostrobis securis was much greater in the upper than middle estuary. In summary, while

many euryhaline species are capable of inhabiting both the middle and upper estuary, other

species are less tolerant of the lower and more variable salinities of the upper estuary, leading to

a lower species richness. Oligohaline species and some euryhaline species have a lower salinity

preference, residing exclusively or in greater abundance in the upper estuary, which results in an

appreciable contrast in community composition between the upper and middle estuary.

While many species may be capable of inhabiting a certain region of the estuary, only a few

species will typically dominate the benthic community, both in terms of total abundance and

biomass (Table C1). For example, in rank order the four most abundant species within the

shallows of both the middle estuary (S. aequisetis, Grandidierella sp., Capitella ‘capitata’ and

T. dulongi) and upper estuary (A. semen, Pseudopolydora sp., S. aequisetis and X. securis),

together comprise approximately 70% of total abundance in each locale (Rose, 1994;

Kanandjembo et al., 2001). Further, the ten most abundant species together encompass greater

than 90% of total abundance in each area (Table C1). It is pertinent that abundance is only one

of many population descriptors that provide information about the dominance or role of

182

Page 197: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

community constituents. For instance, within shallows of the middle estuary the constituency of

the top four ranked species in terms of biomass dominance is much different to those species

dominating abundance. That is, while polychaetes and crustaceans dominate abundance (see

above), 70% of total biomass is dominated by three bivalve molluscs (Sanguinolaria biradiata,

X. securis and Musculista senhousia) (Table C1; Rose, 1994). Thus, dominance is a subjective

term unless the context of consideration is clearly defined, i.e. dominance of taxa in terms of

abundance, biomass, productivity or other functional role in ecology (e.g. trophic effects and

community structuring) or biogeochemical effects (e.g. upon sediment stability, reworking,

particle size distribution or nutrient cycling).

Table C1: The rank dominance of the most common macrobenthic inhabitants of the upper and middle Swan River Estuary, either in terms of percentage contribution to total abundance [A & B] or biomass [C]. Adapted from Rose (1994) and Kanandjembo et al. (2001).

[A] Upper Estuary [B] Middle Estuary [C] Middle Estuary

Ran

k

Spp Group*

% o

f tot

al

abun

danc

e

Spp Group*

% o

f tot

al

abun

danc

e

Spp Group*

% o

f tot

al

biom

ass

1 Arthritica semen

M, B 25.6 Simplisetia aequisetis

P, N 19.4 Sanguinolaria biradiata ##

M, B 32.4

2 Pseudopolydora sp. #

P, S 16.8 Grandidierella sp. ##

C, A 18.8 Xenostrobus securis

M, B 24.7

3 Simplisetia aequisetis

P, N 15.2 Capitella ‘capitata’

P, C 17.8 Musculista senhousia

M, B 13.2

4 Xenostrobus securis

M, B 13.7 Tanais dulongi ##

C, I 11.0 Simplisetia aequisetis

P, N 11.2

5 Capitella ‘capitata’

P, C 4.1 Arthritica semen

M, B 7.7 Tellina deltoidalis

M, B 3.0

6 Desdemona ornata #

P, S 3.9 Corophium minor ##

C, A 5.0 Leitoscoloplos normalis

P, O 3.0

7 Leitoscoloplos normalis

P, O 3.5 Boccardiella limnicola

P, S 4.7 Marphysa sanguinea

P, E 2.9

8 Paracorophium excavatum

C, A 3.3 Leitoscoloplos normalis

P, O 3.9 Irus crenata ##

M, B 1.9

9 Erichthonius sp.

C, A 3.0 Sanguinolaria biradiata ##

M, B 2.9 Grandidierella sp. ##

C, A 1.3

10 Fluviolanatus subtorta #

M, B 2.9 Musculista senhousia

M, B 1.8 Arthritica semen

M, B 0.8

11 Oligochaete sp. 1 #

O 1.9 Paracorophium excavatum

C, A 1.4

12 Boccardiella limnicola

P, S 1.6 Xenostrobus securis

M, B 1.3

13 Prionospio cirrifera

P, S 0.9 Syncassidina aestuaria

C, I 1.3

14 Oligochaete sp. 2 #

O 0.7 Melita matilda

C, A 0.6

15 Syncassidina aestuaria

C, I 0.5 Melita zeylanica ##

C, A 0.4

16 Musculista senhousia

M, B 0.5 Marphysa sanguinea

P, E 0.3

17 Tatea preissi M, G 0.4 Australonereis ehlersi ##

P, N 0.2

18 Chironomid sp. 1 #

I 0.4 Heteronemertean sp. ##

N 0.2

183

Page 198: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Table C1: (continued)

[A] Upper Estuary [B] Middle Estuary

Ran

k

Spp Group*

% o

f tot

al

abun

danc

e

Spp Group*

% o

f tot

al

abun

danc

e

19 Arachnid sp. #

A 0.2 Irus crenata

M, B 0.2

20 Amarinus laevis #

C, D 0.2 Ficopmatus enigmaticus ##

P, P 0.2

21 Nassarius burchardi

M, G 0.2 Ericthonius sp.

C, A 0.1

22 Melita matilda

C, A 0.1 Halicarcinus bedfordi

C, D 0.1

23 Velacumanthus australia

M, G 0.1 Tellina deltoidalis

M, B 0.1

24 Chironomid sp. 2 #

I 0.1 Munna brevicornis ##

C, I 0.1

25 Marphysa sanguinea

P, E 0.1 Tatea preissi

M, G 0.1

26 Tellina deltoidalis

M, B <0.1 Nodilittorina unifasciata ##

M, G 0.1

# Not found within the middle estuary. ## Not found within the upper estuary. * 1st letter: P = Polychaeta, O = Oligochaeta, M = Mollusca, C = Crustacea, I = Insecta, A = Arachnida, N = Nemertean.

2nd letter: for polychaetes, S = Spionidae, P = Serpulidae, N = Neredidae, O = Orbiniidae, E = Eunicidae; for molluscs, B = Bivalvia, G = Gastropoda; for crustaceans, A = Amphipoda, D = Decapoda I = Isopoda.

Geographic variation in salinity regime has above been argued to be the main delineator of the

major biotic zones within the estuary. Similarly and perhaps most importantly, it is seasonal and

inter-annual variations in the dynamics of the characteristic salt-wedge of the estuary (driven by

variation in rain and freshwater discharge, Section 2.1.2 and Appendix A), that largely govern

spatio-temporal variability in community structure. Firstly on a seasonal basis, the boundaries of

the ranges of the species, as defined by salinity tolerance, within the lower, middle and upper

estuary may be expected to shift upstream during summer and autumn, concomitant with upstream

emplacement of higher salinity water, i.e. the salt-wedge (Phase 1). Conversely, the distribution of

species may be expected to shift downstream during late winter to early spring, as freshwater

discharge increases and the salt-wedge wanes (Phase 3). Thus, the theoretical upper and lower

boundary of each species’ range as determined by salinity tolerance, will seasonally fluctuate

concomitant with progression of the salt-wedge. From year to year, the absolute magnitude of these

upstream and downstream shifts in biotic zones will be dependant upon inter-annual variation in

rainfall and freshwater discharge which ultimately governs the progression of the salt wedge.

These boundaries are based hypothetically only on the influence of salinity. Within these bounds,

many other regulatory factors such as faunal dispersal capability and habitat/substrate suitability

and availability will regulate the true expression and dynamics of a macrobenthic species

distribution. Salinity may primarily govern the broad distribution and composition of the biota

184

Page 199: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

within the Swan River Estuary (Hodgkin, 1987); however, at a finer scale, a multitude of

physicochemical and biological factors, and their interactions, will regulate where particular species

live and how abundant or productive they are. Some of the major determinants in fine-scale

patterns of species distribution and resultant community composition are briefly discussed below.

Fauna of sediments that underlie shallow waters have been more comprehensively studied than

deeper benthic communities within the Swan River Estuary, likely due to differences in the ease

and accessibility of sampling. However, water depth and correlating parameters (such as light

availability, primary productivity and temperature, for example) have been documented to

largely influence community composition and abundance of constituent fauna at a local scale.

Those studies that have sampled deeper sediments have found that both species richness and

abundance are generally greater in shallow than deeper waters. For the upper estuary,

Kanandjembo et al. (2001) established that the difference in composition was mainly due to the

greater presence and abundance of the nereidid S. aequisetis and bivalve X. securis in shallow

waters. The brief sampling regime of Wallace (1977) also found that species densities were

generally greater in the shallows than in deeper water of both the upper and middle estuary.

Kanandjembo et al. (2001) offer the explanation that the aforementioned shallow-water species

either: (i) prefer and are proliferating on an increased abundance of photosynthetic food material

in shallower water, or (ii) coarser sediments in the shallows buffer against hypoxia which is a

more conducive environment for burrowing invertebrates, or (iii) a combination of both these

reasons results in the observed discrepancy between shallow and deep water environments.

Contrastingly, the spionid polychaete Prionospio cirrifera and sabellid polychaete Desdemona

ornata were more abundant in deeper than shallow water; it was argued that these species more

exclusively feed on detrital material that settles out in deep waters (Kanandjembo et al., 2001).

Manipulative studies are required to test these hypotheses and determine which factor, or

combination of factors, are responsible for the differences in community composition with

depth.

The observation that regulatory factors likely do not act unilaterally, but rather synergistically, is

pertinent to an understanding of population and community dynamics. For example, the seasonal

ebbing and waning of the salt-wedge will be accompanied by changes not only in salinity but also

in temperature, dissolved oxygen, nutrients, organic matter supply, water flow rates and

stratification events (Section 2.1.2 and Appendix A). To reiterate, while salinity may determine the

theoretical distribution of a species, it will be variations in these other factors that result in the

tangible expression of population structure and community composition. An eloquent example to

note is again provided by the study of Kanandjembo et al. (2001), wherein community composition

at any given time did not differ significantly between sites of the upper estuary; however,

community composition did vary significantly between seasons of a year and also between seasons

185

Page 200: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

of successive years. An argument may be invoked that similar salinities in the upper reaches at any

one time resulted in comparable community composition and that seasonal shifts in salinity

likewise imparted change on the distribution of constituent species. Yet, this does not explain the

variance in community composition from year to year, since salinities between seasons of

successive years were reasonably similar. Kananjembo et al. (2001) argue that while salinity

regime was largely responsible for the seasonal, and perhaps some of the inter-annual, variation in

community composition, water flow rates associated with discharge events also affected species

richness and densities. It was proposed that extremely high rates of discharge during one of the

winters caused much scouring of substrates, particularly in the shallows, which resulted in

dislodgement of fauna and both depressed species richness and overall abundance. Somewhat

lower flow rates in other winters, while resulting in similarly low salinities in the upper estuary, did

not appear to cause this scouring effect to shallow substrates; wherein species richness and

abundance remained comparatively high. The recovery phase from a disturbance scouring event

led to community composition of subsequent seasons being appreciably different to seasons of the

previous, non-impacted year.

Finally, although previous examples have placed emphasis on physicochemical regulators of

species distribution and community composition, biological determinants may be just as

important (Day et al., 1989). To some extent, biological influences on community dynamics –

e.g. competition, predation, other trophic relations, dispersal/reproductive mechanisms,

intraspecific interactions and biogeographical constraints etc. – are often overlooked, since they

are usually not as easily quantified as physicochemical determinants that only require a quick

measurement or sample (e.g. salinity, dissolved oxygen, substrate type etc.). As such, empirical

evidence of biological controls on the distribution of species within the Swan River Estuary are

largely lacking. Nevertheless, certain hypotheses have been proposed. As alluded to above,

food resource availability such as plant material in shallow waters may be an important

determinant of species presence and abundance (Kanandjembo et al., 2001). Hodgkin and

Lenanton (1981) suggested that the most important item in the food chain of the Swan River

Estuary is detritus. Many benthic invertebrates are detrital feeders and the distribution of

detritus, in terms of both abundance and quality, may influence species distributions. Further

many resident fish species are either detritivores or feed on benthic macroinvertebrates

(Hodgkin and Lenanton, 1981). As such, fish predation may exert considerable influence on the

abundance and even presence of sedimentary fauna. Kanandjembo (1998) studied the diet of

four fish species (Acanthopagrus butcheri, Amniataba caudavittata, Pseudogobius olorum and

Leptatherina wallacei) and found that benthic macrofauna were a component of each, but that

the composition of macroinvertebrates was different for each fish species. Thus, the

composition of the fish community may in turn regulate benthic community composition.

186

Page 201: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

As a final example of biological and synergistic effects on the distribution and abundance of

macrobenthic fauna, it is salient to reconsider the potential influence of algal blooms. In

Section 2.1.3 I observed that the decomposition of collapsed algal blooms may result in

localised anoxia or toxic effects, which have profound implications for the mortality and/or

abundance of both relatively sessile macrofauna and mobile fish species. Of course, nutrient

availability may ultimately regulate the extent and severity of algal blooms and the resultant

potential impacts on the macrobenthos. This illustrative, if not simplistic, example of the effects

of eutrophication and the interaction between nutrients, algal blooms, dissolved oxygen

concentration and toxic effects on both benthic fauna and fish (a potential source of predation),

clearly highlights the type of synergistic effects that may govern macrobenthic faunal

community composition. The contention of this thesis that macrobenthic fauna may themselves

regulate nutrient availability, conceptually completes a circular linkage whereby the effects of

macrofauna may govern factors which influence their own distribution.

187

Page 202: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,
Page 203: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

The reaction is catalysed by nitrogenase, a complex molybdo-iron enzyme (Capone, 1988).

Nitrogen fixation is energetically expensive requiring about 15 adenosin triphosphate molecules

(ATP) per N2 reduced, with the metabolic cost being met by ATP production from

photosynthesis or chemosynthesis (Capone, 1988; Heathwaite, 1993).

189

APPENDIX D

COMPONENT PROCESSES OF THE NITROGEN CYCLE

D.1 OVERVIEW

It is not the intent of this Appendix to comprehensively review nitrogen cycling in aquatic

environments, and the reader is referred to Brezonik (1972), Keeney (1973), Carpenter and

Capone (1983), Blackburn (1986), Sprent (1987), Blackburn and Sørensen (1988), Heathwaite

(1993) and Capone (2000, 2002). A consideration of the biochemistry, biology and relevance of

each process will suffice here to enhance the unfamiliar reader’s understanding of the dynamic

positioning of denitrification within the estuarine nitrogen cycle. Necessarily, denitrification is

reviewed in greater detail and supplemented with information regarding a methodological

approach to the process using acetylene and enzyme kinetics, of relevance to this thesis’

practical consideration of denitrification within estuarine sediments (Chapter 5). Finally, some

novel microbial nitrogen transforming pathways are detailed to highlight potential implications

to nitrogen cycling and facilitate awareness of the processes.

D.2 NITROGEN FIXATION

Nitrogen fixation is a bacterially mediated, exergonic process whereby nitrogen gas (dinitrogen,

N2) is reduced to ammonium (NH4+) which is then incorporated into cellular material (Figure

D1; Capone, 1988; Heathwaite, 1993). The chemical reaction is illustrated below:

½ N2 + 1½ H2O + H+ NH4+ + ¾ O2 ΔG = + 317 kJ mol N-1

(amino acids)

Bacteria containing nitrogenase and capable of fixing atmospheric nitrogen are disparately

found amongst all major prokaryotic groups including: phototrophic cyanobacteria (e.g.

Anabaena spp., Aphanizomenon spp., Calothrix spp., Gloeocapsa spp., Lyngbya spp., Nodularia

spp., Nostoc spp., Oscillatoria spp. and Phormidium spp.), anoxyphotobacteria (e.g. some

sulphur and purple non-sulphur bacteria) and Prochlorales (e.g. Prochloron sp., an ascidian

symbiont); and, heterotrophic aerobes (e.g. Azotobacter spp.), microaerophiles (e.g.

Azospirillum spp., Campylobacter sp. and Beggiatoa spp.), facultative anaerobes (e.g.

Enterobacter spp., Klebsiella spp. and Vibrio spp.) and strict anaerobes (e.g. Desulfovibrio spp.

and Clostridium spp.) (Capone, 1988).

Page 204: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

190

NO3-

NH4+

Organic N

NO2-

N2O N2Organic N

NH4+ NO3

- NO2-

N2O N2

NH4+

NO3-

NO2-

Organic N

Export

NH4+

NO3-

NO2-

Organic N

Import

NH4+, NO3

-, NO2-

Groundwater Import

Ammonification / Mineralisation

Nitrification

Denitrification

Sediment / Water Column / Atmosphere Solute Flux

ATMOSPHERE

WATER COLUMN

SEDIMENT

Nitrogen Fixation

Ammonium Assimilation

Assimilatory Nitrate Reduction

Dissilmilatory Nitrate

Reduction to Ammonium

LEGEND

Figure D1: The estuarine nitrogen cycle. A conceptualisation of the main transformations occurring in the sediment and water column. For simplicity many intermediates have been omitted. See text for further details.

Page 205: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Perhaps of most importance to estuarine systems are the bloom-forming nitrogen fixing

cyanobacteria. In the absence of available (dissolved inorganic) nitrogen, many species are

capable of exploiting a phosphorous nutrient source (together with an ambient dinitrogen gas

supply) to form dense microbial or algal mats either in the plankton (some species have a gas

vacuole buoyancy system) or on the sediment bed (Capone, 1988; Heathwaite, 1993). Some of

the most notorious (i.e. toxin-producing, anoxia-generating) nuisance algal bloom species are

nitrogen fixing cyanobacteria (e.g. Anabaena spp., Aphanizomenon spp., Lyngbya spp.,

Nodularia spp. and Oscillatoria spp.). Toxic species with a freshwater physiology (e.g.

Anabaena spp.) are of considerable consequence in freshwater lakes and rivers (Lean et al.,

1978), but may also be of importance during the freshwater phase or within the upper reaches of

estuaries (Paerl and Zehr, 2000). The potential for toxic salinity-tolerant species such as

Nodularia spp. and Aphanizomenon spp. to bloom in marine and estuarine environments is also

great (Paerl and Zehr, 2000). Where seagrasses are present within estuaries, symbiotic

relationships with heterotrophic nitrogen fixers may be particularly important in nitrogenous

input to the system (Capone, 1988). Within unvegetated sediments some heterotrophic nitrogen

fixation activity is usually detected; however, it is usually only a minor component of the

estuarine nitrogen cycle (Capone, 1988, 2000). In general, because external loads and recycled

nitrogen are in relatively high supply, rates of nitrogen fixation in unvegetated estuarine

environments and their sediments are generally of little significance (Postma et al., 1984,

Capone, 2000). Nevertheless, the potential for nitrogen fixation, especially in the presence of

seagrass beds or cyanobacterial blooms, should be considered in any estuarine nitrogen cycle

budget.

D.3 AMMONIUM ASSIMILATION AND ASSIMILATORY NITRATE

REDUCTION

Assimilation is a process whereby inorganic nitrogenous compounds are incorporated into the

organic biomass of primary producers and some heterotrophic bacteria and fungi (Keeney,

1973; Hattori, 1983; Kuenen and Robertson, 1988) (Figure D1). In a biological sense, it is the

most important component of the nitrogen cycle because it makes inorganic nitrogen available

for synthesis into cellular materials (i.e. for growth of phytoplankton, macroalgae,

microphytobenthos, seagrasses and bacteria). Influx of inorganic nitrogen into organisms

generally results either from ammonium (NH4+) or nitrate (NO3

-) assimilation (Brezonik, 1972).

Ammonium assimilation is the process whereby inorganic ammonium (NH4+) is converted into

the organic nitrogen biomass of an organism as biosynthetic nitrogen (“R-NH2”) (Sprent, 1987).

The process may be illustrated as follows:

NH4+ “R-NH2”

191

Page 206: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Ammonium is in the same oxidation state (-3) as organic nitrogen and is therefore an ideal

substrate for assimilation (Brezonik, 1972). The general toxicity of ammonium to cells,

however, makes rapid assimilation very desirable (Sprent, 1987). For this reason, the major

route of assimilation is energetically expensive. Two ATP molecules are required to convert

one ammonium (NH4+) molecule into a glutamine molecule utilising the enzyme glutamine

synthetase (GS) (Sprent, 1987; Capone, 2000). Glutamine is then enzymatically converted to

glutamate using glutamate synthase (GOGAT) (Capone, 2000). Finally, glutamate is used to

synthesize other nitrogen containing products (e.g. amino acids, proteins and polynucleotides =

“R-NH2”) within the cell (Capone, 2000). This GS/GOGAT pathway is the main route of

ammonium assimilation in marine organisms utilising inorganic nitrogen as a source, including

nitrogen fixers (Capone, 2000). Alternative pathways exist for the assimilation of ammonium.

For example, ammonium may be biosynthesised via glutamic acid in a reversible, less energetic

and much slower reaction, the glutamate dehydrogenase pathway (Sprent, 1987; Capone, 2000).

This second pathway of ammonia assimilation is less prevalent than the first due to the

increased potential for toxic ammonium build-up and it is thought to be more significant in the

degradative direction (ammonification, see Section 1.5.3) (Sprent, 1987). The reader is referred

to Falkowski (1983) and Capone (2000) for a more comprehensive review of biochemical

ammonium assimilation pathways. Generally, any nitrogenous compound that is to be

incorporated into an organism as biosynthetic nitrogen must first be reduced to the same

oxidation state as ammonium (Brezonik, 1972).

Assimilatory nitrate reduction or nitrate assimilation is defined as the reduction of nitrate to

ammonia via nitrite (NO2-) for incorporation into biomass (Kuenen and Robertson, 1988). The

process is illustrated below:

NO3- NO2

- NH4+ “R-NH2”

Nitrate is utilised as a source of biosynthetic nitrogen as the ammonium produced is assimilated

in the synthesis of cellular materials including amino acids, proteins and polynucleotides

(Webb, 1981; Hattori, 1983; Kuenen and Robertson, 1988). The process is differentiated from

ammonium assimilation because the ammonium utilised in biosynthesis is produced

intracellularly rather than being incorporated from the external environment and nitrate is the

ultimate biosynthetic nitrogen source. In the first two steps of the process, specific assimilatory

nitrate and nitrite reductase enzymes are respectively required to reduce the oxidised nitrogen

species to the required ammoniacal oxidation level for biosynthetic incorporation (Brezonik,

1972; Capone, 2000). Organisms capable of ammonium assimilation may also be able to

assimilate nitrate if they have the appropriate enzymatic apparatus. In essence, assimilatory

192

Page 207: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

nitrate reduction is simply an extension of the ammonium assimilation process since the final

step of converting ammoniacal nitrogen to biosynthetic nitrogen is the same as that described

above. Detection of the presence of specific nitrate and nitrite reductase enzymes in organisms

therefore offers a tool for identification of the capacity for those organisms to utilise nitrate as a

biosynthetic nitrogen source (Brezonik, 1972). Assimilatory nitrate reduction can be mediated

by plants, phytoplankton, aerobic bacteria and fungi and proceeds at the expense of light energy

in photoautotrophs or energy reserved in organic compounds in heterotrophs (Hattori, 1983).

Assimilatory nitrate reduction generally requires aerobic conditions so that photosynthesis can

provide the energy required to reduce the oxidised nitrogen species (Tiedje, 1988). The process

is also repressed by large concentrations of ammonia and organic nitrogen typical to anaerobic

environments (Tiedje, 1988).

Of the two processes (ammonium assimilation and assimilatory nitrate reduction), ammonium

assimilation is the less energetic (the prior reduction of oxidised nitrogen species obviously not

being required) and ammonium is therefore often deemed the preferred substrate for

incorporation into biomass (Brezonik, 1972; Postma et al., 1984). This has been generally

confirmed within estuaries and near-shore waters whereby comparative studies typically

indicate ammonium to be the preferred substrate (Paasche, 1988; Middelburg and

Nieuwenhuize, 2000a). However, at high ambient nitrate concentrations significant amounts of

nitrate may be assimilated, despite the lower preference for this substrate (Heathwaite, 1993;

Middelburg and Nieuwenhuize, 2000b). The latter point is particularly relevant for eutrophic

estuaries that are increasing in nitrate concentration due to anthropogenic inputs (Middelburg

and Nieuwenhuize, 2000b). Whether, and to what, extent ammonia and/or nitrate will be

utilised as an assimilatory nitrogen source will depend primarily upon the spatial and temporal

availability of substrate and the enzymatic capacity of that organism to reduce nitrate.

Identification of the major biosynthetic nitrogen source/s for a particular primary producer is of

fundamental importance in attempting to regulate and manage the occurrence and intensity of

nitrogen limited algal blooms. For phytoplankton blooms within the Swan River Estuary,

Horner Rosser and Thompson (2001) found no physiological preference for ammonium or

nitrate at any time of year, except within bottom waters during summer whereby ammonium

was preferred, apparently in relation to the high rate of ammonium flux from sediments at that

time.

It is important to note that nitrogen assimilation by heterotrophic bacteria, benthic microalgae

(microphytobenthos) and seagrasses may present significant sources of nitrogen uptake, that

compete with phytoplankton (algal bloom) assimilatory pathways. Heterotrophic bacteria

assimilate inorganic nitrogen because organic substrates often do not contain enough nitrogen to

fully support bacterial growth (Middelburg and Nieuwenhuize, 2000b). In estuarine and marine

193

Page 208: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

systems, bacterial nitrogen assimilation has been reported to account for a significant but highly

variable (<5% to >90% ) proportion of total nitrogen uptake (Middelburg and Nieuwenhuize,

2000b). During maximal growth of estuarine seagrasses and microphytobenthos in spring and

summer, assimilatory nitrogen uptake may both reduce the flux of nitrogen from the sediment to

the water column (where phytoplankton reside) and enhance nitrogen flux from the water

column towards the bottom (Risgaard-Petersen and Ottosen, 2000). While seagrass and

microphytobenthos nitrogen assimilation most likely do not alter estuarine retention of total

nitrogen on an annual scale (since subsequent decomposition allows nitrogen to re-enter the

environment), the seasonal timing of assimilation may influence phytoplankton nitrogen

availability (Risgaard-Petersen and Ottosen, 2000). Benthic microalgae actually may enhance

water column primary productivity because nutrients are transferred to the sediment via algal

uptake during periods of high nutrient concentrations (winter and spring) (Risgaard-Petersen

and Ottosen, 2000). When nutrients are scarce during summer, decomposition of the algal

biomass causes the release of ammonium, which may be used to fuel planktonic primary

production (Risgaard-Petersen and Ottosen, 2000). On the other hand, rooted seagrasses have

been found to be a net sink of dissolved inorganic nitrogen over the entire spring-summer

growing period, whereby nitrogen uptake is mainly and more permanently allocated to plant

tissue (Risgaard-Petersen and Ottosen, 2000). Release of nitrogen from seagrasses is mainly via

loss of leaves which are known to decompose relatively slowly (Risgaard-Petersen and Ottosen,

2000). Net nitrogen sequestration may occur over the seagrass growing period with nutrient

release mainly occurring outside this period, when phytoplankton growth is light-limited and

nutrients are more likely to be flushed from the estuary (Risgaard-Petersen and Ottosen, 2000).

A scenario like this has been documented for Halophila ovalis meadows that spatially occupy

approximately 20% of the Swan River Estuary (Connell and Walker, 2001). During summer,

H. ovalis biomass increases, pore-water ammonium is assimilated and meadows act as a nutrient

sink; while during winter, there are large losses of plant biomass and meadows become a source

of nutrients (Connell and Walker, 2001).

D.4 AMMONIFICATION / MINERALISATION

Ammonification is essentially the opposite of assimilation since it is the process whereby

ammonium is released back to the environment during the catabolism of nitrogen containing

biological materials (“R-NH2”) (Brezonik, 1972; Webb, 1981) (Figure D1). It is also often

referred to as mineralisation or remineralisation (particularly in the marine sciences) (Webb,

1981). The process can be illustrated as follows:

“R-NH2” NH3 ↔ NH4+

194

Page 209: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

There are two major pathways of ammonium release. Firstly, organic nitrogen breakdown and

excretion of dissolved inorganic nitrogen by higher organisms via the urea cycle mainly has an

end-product of ammonium in estuarine environments (although urea, amino acids and other

nitrogen compounds may account for ~20% of nitrogen release) (Webb, 1981; Capone, 2000).

Secondly, the degradative catabolism (decomposition) of organic nitrogen by heterotrophic

micro-organisms (including many bacteria and fungi) also results in ammonium release.

Microbial decomposition is metabolically expensive and requires respiratory pathways based on

oxygen or anoxic equivalents (Klump and Martens, 1983). Aerobic bacteria are capable of

completely degrading complex nitrogenous polymers (e.g. amino acids) to ammonium, whilst

many individual species of anaerobic bacteria can not (Klump and Martens, 1983). In anoxic

environments, decomposition therefore usually occurs in a step-wise manner, with the

metabolite of one species becoming the growth substrate of another (Klump and Martens,

1983). It is generally thought that the most important decomposers in marine environments are

aerobic bacteria in combination with anaerobic fermenters and sulphate reducers (Klump and

Martens, 1983). Due to their differing respiratory requirements these processes are typically

segregated in space (e.g. biogeochemical zonation and stratification of sediments) and/or time

(e.g. due to fluctuating oxic/anoxic periods) (Klump and Martens, 1983). Additionally,

temperature is a factor of paramount importance in determining the rate of microbial ammonium

regeneration (Klump and Martens, 1983). Rates increase exponentially with temperature: a rise

of 10 °C is generally accompanied by a two to four fold increase in ammonium release. In

temperate estuaries and nearshore waters that exhibit up to a 20 °C temperature range, this

imparts significant seasonality to microbially mediated ammonium regeneration (Klump and

Martens, 1983). Seasonal asymmetry and summer maxima of sediment ammonium effluxes are

frequently observed (Boynton et al., 1980; Klump and Martens, 1983; Kemp et al., 1990),

including within the Swan River Estuary (Thompson, 1998, 2001).

Bacterial decomposition of organic nitrogen was traditionally assumed to be the major agent of

ammonium release in aquatic ecosystems; however, many studies have since shown that animal

excretion is frequently the dominant aquatic ammonium regenerating process (Brezonik, 1972;

Keeney, 1973). An often cited example is the early research of Johannes (1968), who pointed

out that zooplankton feeding on phytoplankton and detritus may excrete amounts of dissolved

nitrogen many times greater than their body content in a period of days to weeks (Brezonik,

1972; Keeney, 1973). In systems with high zooplankton productivity, excretion is suggested as

the dominant mechanism of ammonification (Brezonik, 1972). Similarly, excretion may be a

major component of ammonium regeneration in estuarine systems with dense benthic fauna

assemblages. For example, Gardner et al. (2001) found that zebra mussel excretion dominated

total ammonium regeneration in areas where the bivalve was abundant. The influence of

excretion by benthic fauna on estuarine nitrogen cycling is more fully discussed in Chapter 5.

195

Page 210: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Organic matter not consumed by animals is available for bacterial and fungal ammonification.

When living tissues die, nitrogenous polymers become available for decomposition (Klump and

Martens, 1983). Due to the collection, deposition and degradation of organic matter on and

within the sediment bed, ammonium regeneration in benthic environments is often dominated

by the catabolic activity of microrganisms rather than animal excretion (Klump and Martens,

1983). For example, in the heavily studied Narragansett Bay, Rhode Island, nitrogen inputs are

largely (70%) recycled within the system (Klump and Martens, 1983). Of the recycled amount,

approximately 70% is due to benthic ammonium regeneration and the remaining 30% due to

pelagic regeneration, primarily by zooplankton excretion (Klump and Martens, 1983). There is

considerable ongoing debate regarding whether animals or bacteria and fungi are more

important mineralisers of organic nitrogen (Webb, 1981). Most likely, the dominant

mineralisation source of ammonium within estuaries will be situation-specific (both

geographically within and between estuaries and over time within an estuary), depending on

factors such as faunal abundance and appropriate sediment conditions for bacterial/fungal

decomposition of organic nitrogen.

Whatever the source process, the rate and timing of ammonium release from organic matter in

estuarine environments may be a critical regulator of nitrogen assimilation by phytoplankton

and other primary producers. Where ammonium regeneration is the major provider of a

nitrogen source, the rate of its release may be an important factor in determining nutrient

limitation of phytoplankton (Heathwaite, 1993). If ammonium release is relatively slow, then

assimilation by phytoplankton potentially can become nitrogen limited (Heathwaite, 1993).

Alternatively, where ammonium release rates are high, phytoplankton and other primary

producers may become unlimited in growth by nitrogen. Further, it has often been found in

estuaries and continental shelf sediments that much of the nitrate substrate that is denitrified

(see Section D7) is derived from ammonium produced from organic matter mineralisation that

has been subsequently nitrified (see Section D5 and Figure D1) (Seitzinger, 1988; Kemp et al.,

1990; Gardner et al., 2001). This ammonification-nitrification-denitrification process may be a

major shunt of nitrogen away from the assimilatory pathways of phytoplankton. While some

studies have confirmed that heightened ammonium regeneration in summer (due to temperature

increases) causes increased rates of nitrification and denitrification, others have found that

enhanced decomposition causes sediment anoxia, thereby suppressing aerobic nitrification

(Kemp et al., 1990).

196

Page 211: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

D.5 NITRIFICATION

Nitrification typically occurs in a two-step biological process whereby ammonium is oxidised to

nitrite followed by the further oxidation of nitrite to nitrate (Figure D1) (Ward, 2000). It links

the most oxidised and most reduced forms of nitrogen as such:

NH4+ NO2

- NO3-

Each stage (NH4+ NO2

- and NO2- NO3

-) is respectively mediated by a particular group of

bacteria (Table D1), the ammonium oxidisers (nitrosobacteria) and the nitrite oxidisers

(nitrobacteria) (Colliver and Stephenson, 2000; Šimek, 2000; Capone, 2002). Most nitrifiers are

aerobic chemolithoautotrophic proteobacteria that obtain energy for growth through either

oxidation process (see below) (Kaplan, 1983; Kuenen and Robertson, 1988; Ward, 2000;

Capone, 2002). Genera of nitrifying bacteria (Table D1) are not necessarily closely related, but

all appear to have arisen from a common photosynthetic ancestor (Ward, 2000). No known

species is capable of undertaking both ammonium oxidation and nitrite oxidation (Bothe et al.,

2000; Ward, 2000). Nitrification and chemolithoautotrophy are unique metabolic traits that

allow nitrifying bacteria to exploit a unique environmental niche; however, growth is usually

slow or inefficient due to their inflexible nutritional requirements and energy needs for carbon

dioxide reduction (Henriksen and Kemp, 1988; Ward, 2000). The implication is that

populations of low abundance can impart high rates of nitrification (Henriksen and Kemp,

1988).

Table D1: Genera of autotrophic nitrifying bacteria (Adapted from Šimek, 2000).

Ammonium oxidisers (NH4+ NO2

-)

nitrosobacteria

Nitrite oxidisers (NO2- NO3

-)

nitrobacteria

Nitrosomonas

Nitrosococcus

Nitrosospira

Nitrosolobus Nitrosospira*

Nitrosovibrio

Nitrobacter

Nitrococcus

Nitrospira

Nitrospina } * Recently proposed on the basis of 16S rRNA sequence homology (Bothe et al., 2000).

Kaplan (1983) denotes the reactions for the first step (ammonium oxidation) of nitrification as:

2 e-

NH4+ + ½ O2 NH2OH + H+ ΔG = + 16.7 kJ mol N-1

4 e-

NH2OH + O2 NO2- + H2O + H+ ΔG = - 289 kJ mol N-1

ΔG = - 272 kJ mol N-1

197

Page 212: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

The first reaction of ammonium to hydroxylamine (NH2OH) conversion is catalysed by the

enzyme ammonia mono-oxygenase, which requires molecular oxygen and a source of reducing

power (Bothe et al., 2000; Colliver and Stephenson, 2000). Hydroxylamine (NH2OH) is an

intermediate that is unstable in aqueous solution and therefore rarely accumulates (Kaplan,

1983). The second reaction is carried out by the enzyme hydroxylamine oxidoreductase and

yields two pairs of electrons, one pair being used to drive the first reaction and the other pair is

used for energy production (Bothe et al., 2000; Colliver and Stephenson, 2000). As such, the

energy requirements of ammonium oxidisers (such as Nitrosomonas spp.) are generally thought

to only be derived from the oxidation of hydroxylamine, whilst the first reaction is not believed

to contribute to the organism’s energy needs (Colliver and Stephenson, 2000).

Kaplan (1983) notes that second step of nitrification (nitrite reduction) is a simple, two electron

transfer, with molecular oxygen as the terminal electron acceptor:

2 e-

NO2- + ½ O2 NO3

- ΔG = - 76.1 kJ mol N-1

Nitrite oxidation by bacteria such as Nitrobacter spp. is catalysed by the enzyme nitrite

oxidoreductase, with oxygen being supplied by water (Colliver and Stephenson, 2000). Nitrite

oxidation yields two electrons that are utilised for energy production and the reduction of O2 to

water (Colliver and Stephenson, 2000).

Many environmental variables control the rate of nitrification but the main regulatory factors are

ammonium as an initial substrate, and oxygen for respiration and substrate oxidation (Šimek,

2000; Ward, 2000). The importance of substrate concentration is highlighted by the general

relationship between ambient ammonium concentration and ammonium oxidation rate (Ward,

2000). Such a relation also holds between nitrite oxidation rate and nitrite substrate

concentration (Henriksen and Kemp, 1988; Ward, 2000). Michaelis-Menten type kinetics that

link a hyperbolic dependence of rate with substrate concentration, generally are exhibited in

pure cultures (for a definition of Michaelis-Menten kinetics see Section D.7.1) (Henriksen and

Kemp, 1988; Ward, 2000). Reported Km values (i.e. the half-saturation or Michaelis constant:

the substrate concentration at which the reaction rate is half maximal) range from 70 to 700 μM

NH4+ for ammonium oxidisers and 350 to 600 μM NO2

- for nitrite oxidisers in pure culture

(Henriksen and Kemp, 1988). However, in natural environments (e.g. in the sediment)

nitrifying bacteria are capable of adaptation to very low substrate levels (Km values as low as

0.1 μM have been reported for both ammonium and nitrite oxidation), and it has been suggested

that substrate affinity depends largely on ambient substrate availability (Henriksen and Kemp,

198

Page 213: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

1988). In general, the activity of nitrifying bacteria is usually highest at the sediment water

interface where the rate of ammonification is high (Heathwaite, 1993).

Oxygen is needed for both ammonia and nitrite oxidation and an aerobic environment is an

expected requirement for nitrification. Whilst molecular oxygen is definitely required, it has

been found that nitrifying bacteria actually thrive best under relatively low oxygen

concentrations (Ward, 2000). Such microaerophily has probably evolved in response to the

supplies of ammonium at oxic-anoxic interfaces (Ward, 2000). For example, at the sediment-

water interface, ammonium may be supplied from above by diffusion from the aerobic water

column and from below by anaerobic ammonification (Ward, 2000). Ammonium supply at the

interface is optimal. The aerobic zone appropriate for nitrification may be less than a millimetre

to a few centimetres thick, depending on organic loading (as an ammonification source) and the

oxygenation status of the environment (Ward, 2000). The high rate of oxygen demand for both

oxidative processes and aerobic respiration by nitrifying bacteria can create anoxic conditions in

some environments (Heathwaite, 1993).

Nitrifying bacteria are also temperature, pH, sulphide and light susceptible. Optimum

temperature for nitrifying bacteria in pure culture is typically in the range of 25 to 35 °C, with

growth occurring between 3 to 45 °C (Henriksen and Kemp, 1988). Ward (2000) suggests that

temperature is not generally an important environmental variable because nitrifying populations

are adapted to the temperature of their environment. That is, nitrifiers adapted to low

temperatures, when warmed, can nitrify at similar rates as nitrifiers adapted to high

temperatures (Ward, 2000). While a dependence on temperature can be demonstrated for

nitrification in any particular environment, it is less sensitive to variation by temperature than

other processes and is usually regulated by some other variable (Ward, 2000). Optimal

nitrification occurs in a narrow pH range (pH 7 to 8.5), with a slightly wider limiting range (pH

6 to 9.5) (Henriksen and Kemp, 1988). Shallow-water sediments with benthic microalgae can

exhibit marked diurnal pH variation (e.g. pH 7 to 10) as day-time algal carbon dioxide

consumption increases the pH, potentially limiting nitrification (Henriksen and Kemp, 1988).

Nitrification inhibition by organosulphur compounds has been demonstrated (Ward, 2000).

Sulphide (HS-), as the product of anaerobic sulphate reduction, is quantitatively the most

important compound and has been found to suppress or inhibit nitrification at concentrations

from 0.4 to 40 μM HS-. Sulphide-rich environments therefore may be expected to have

depressed levels of nitrification. Light, too, has an inhibitory effect on pure culture nitrification,

with light intensity having a clear negative relationship on rate measurements (Ward, 2000).

However, nitrifying bacteria isolated from estuarine environments are apparently less

susceptible and the light limitation of nitrification is only considered relevant in the upper layer

of surface waters (Ward, 2000). In estuarine systems generally, the amount of nitrate generated

199

Page 214: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

internally through nitrification, particularly in eutrophic waters, is often of a lower magnitude in

comparison with the nitrate load received from the drainage basin (Heathwaite, 1993).

However, nitrifying bacteria are found to be more active in temperate estuarine environments

than in most other ecosystems (Capone, 2000), and the provision of an in situ nitrate source is

highly relevant to phytoplankton production.

Ambient rates of sediment nitrification reported for estuaries are within the range of 0 to 350

μmol N m-2 h-1 (Owens, 1993). The importance of nitrification in providing a nitrate source for

denitrification has already been alluded to in Section D.5 above. In estuaries and shallow

marine systems in which anaerobic denitrification typically is limited by nitrate availability (see

Section D.7), nitrate may be produced very rapidly via nitrification of regenerated (mineralised)

ammonium (Henriksen and Kemp, 1988). A high rate of nitrification therefore may be essential

in the efficient provision of nitrate substrate for denitrification, the rate of which may be

enhanced by anoxic conditions created by high oxygen consumption of nitrifying bacteria

(Henriksen and Kemp, 1988; Heathwaite, 1993). In fact, the two processes often are tightly

coupled (Jenkins and Kemp, 1984; Henriksen and Kemp, 1988; Capone, 2000). Nitrification-

denitrification coupling may be temporal (i.e. due to fluctuating redox conditions) or spatial (i.e.

within zones of sharp oxic/anoxic gradients, for example at the sediment-water interface). As

has already been stated in Section D.5, the conversion of ammonium to nitrate by nitrification

followed by reduction of nitrate to nitrogenous gases by denitrification provides a shunt of

dissolved inorganic nitrogen away from the assimilative pathways of phytoplankton (Henriksen

and Kemp, 1988). The degree to which coupled nitrification-denitrification leads to overall

nitrogen reduction (denitrification efficiency) in estuaries is discussed in Section D.7. Less

prevalently, this coupling may be broken as nitrate produced by nitrification may be

anaerobically reduced back to ammonium by alternate dissimilatory processes (see Section D.6

below), to be re-nitrified or assimilated into algal biomass. When denitrification activity is low,

for example in highly aerobic sediments, another scenario may exist. During summer,

catchment inputs are often at a minimum whilst algal utilisation of nitrogen is at a maximum

(Heathwaite, 1993). During this period, nitrate production in situ may be considered an

extended regenerative component of the aquatic nitrogen cycle. While ammonification is

usually identified as the major process regenerating nitrogen for algal assimilation, nitrate

production via nitrification could potentially become important in providing an assimilatory

nitrogen source.

The process of nitrification has been extensively reviewed in the literature (e.g. Helder and De

Vries, 1983; Kaplan, 1983; Henriksen and Kemp, 1988; Kuenen and Robertson, 1988; Ward,

1996; Bothe et al., 2000; Šimek, 2000; Ward, 2000). These sources provide for a detailed

account of nitrifier physiology, ecology and biogeochemical significance.

200

Page 215: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

D.6 DISSIMILATORY NITRATE REDUCTION TO AMMONIUM

Two types of biological nitrate reduction can be distinguished: assimilatory nitrate reduction

and dissimilatory nitrate reduction (Hattori, 1983). The first type has been discussed previously

(Section D.3) and involves the reduction of nitrate into ammonium for incorporation into

organic nitrogen (Figure D1), and occurs predominantly in aerobic environments. Dissimilatory

processes are distinguished from the assimilatory process by the fact that nitrogen reduced is not

used in biosynthesis (Tiedje, 1988). Rather, nitrate serves as an electron acceptor (as an

alternative to oxygen) for the cell’s respiratory metabolism (Tiedje, 1988). Since dissimilatory

processes are inhibited by oxygen, they only occur in anaerobic environments (Tiedje, 1988).

Assimilatory nitrate reduction could possibly also occur under anaerobic conditions, but most

anoxic environments also have large concentrations of ammonium and organic nitrogen, which

repress this process or make it quantitatively insignificant (Tiedje, 1988). Nitrate reduction in

anaerobic environments is therefore dominated by two dissimilatory processes: dissimilatory

nitrate reduction to ammonium (DNRA) and respiratory denitrification (Figure D1; Tiedje,

1988). The former is discussed below and the latter in Section D.7.

In the process of dissimilatory nitrate reduction to ammonium, nitrate is converted to

ammonium via nitrite (Kuenen and Robertson, 1988) (Figure D1):

NO3- NO2

- NH4+

In the beginning step (nitrate to nitrite), nitrate reduction is coupled to energy metabolism

(anaerobic respiration): nitrate is utilised as an alternative terminal electron acceptor to oxygen

as it is first reduced to nitrite (Koike and Sørensen, 1988). This first step may be common to the

denitrification process (Section 1.5.6), but where organisms can undertake only the first step,

the more restricted term of nitrate respiration may be used (e.g. anaerobic nitrate respiration by

some fermentative bacteria) (Webb, 1981; Koike and Sørensen, 1988). However, nitrite must

be further reduced to ammonium for the overall process to be deemed dissimilatory nitrate

reduction to ammonium, or sometimes nitrate ammonification (Koike and Sørensen, 1988). In

the second step (nitrite to ammonium), nitrite may likewise be used as a respiratory alternative

as it is reduced to ammonium; or alternatively, nitrite may serve as an electron sink as it is

reduced to ammonium in a fermentative reaction (Koike and Sørensen, 1988). In the latter case,

nitrite reoxidises reduced NADH without a direct coupling to ATP production thereby allowing

substrate (e.g. glucose) to be fermented (Koike and Sørensen, 1988; Kuenen and Robertson,

1988). When fermentation is not possible, DNRA is entirely an anaerobic respiratory process

(Kuenen and Robertson, 1988). Dissimilatory nitrate and nitrite reductase enzymes are required

for each step of DNRA, and are different to the enzymes catalysing individual steps of

assimilatory and denitrification processes (Kuenen and Robertson, 1988). The capacity of

201

Page 216: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

DNRA is found in strictly anaerobic bacteria (e.g. Clostridium spp.), facultatively anaerobic

bacteria (e.g. Enterobacter spp.) and in some aerobic bacteria (e.g. Bacillus spp.) (Hattori,

1983; Kuenen and Robertson, 1988; Yin et al., 2002).

Unlike assimilatory reduction, in the process of DNRA nitrate is not utilised as a biosynthetic

source of nitrogen, and the ammonium produced is excreted from the cells to the environment

(Hattori, 1983). Thus, DNRA conserves or recycles inorganic (biologically available) nitrogen

within an aquatic environment, unlike the gas-liberating process of denitrification (Kuenen and

Robertson, 1988). For unclear reasons, ammonium conservation by DNRA has historically

been thought quantitatively unimportant (or even worse, sometimes ignored) in the nitrogen

budget of natural systems, and has therefore traditionally received little scientific attention

(Cole, 1990). Whilst the presumption may be justified in many environments, there is evidence

that DNRA plays a significant role in the nitrogen cycling of numerous systems. DNRA is

generally an anaerobic process that competes with denitrification for nitrate/nitrite as a substrate

(Cole, 1988). Since many denitrifying bacteria are facultative anaerobes that use nitrate only

when oxygen is not available (see Section D.7), it appears that denitrification is the dominant

process in environments that favour a respiratory mode of life (rich in nitrate but relatively poor

in organic carbon), whilst DNRA may predominate in more highly anaerobic carbon-rich

systems favouring colonisation by fermentative bacteria (Cole, 1988, 1990; Capone, 2002). In

support of this notion, Cole (1990) cites an early study by Sørensen (1978b) (amongst others)

who demonstrated that the denitrification capacity of a coastal marine sediment decreased

rapidly with depth, but the potential for DNRA in deeper anaerobic sediment was better

maintained. Further, ammonium was the dominant product of nitrate reduction in a highly

anaerobic digested sludge (Kaspar and Tiedje, 1981; Cole, 1988). Favourable conditions for

DNRA are generally not found in natural soils (typically 1 to 5% of total nitrate reduction is due

to dissimilation to ammonium); however, exceptions exist and DNRA was found recently to

account for 15% of total nitrate reduction in an organically-rich anaerobic Australian paddy soil

(Cole, 1990; Yin et al., 2002). Whilst understudied, DNRA in freshwater systems has been

found important when nitrate is available and may account for a moderate proportion of total

nitrate reduction (approximately 10% in one system) (Jones et al., 1982; Cole, 1990). Tobias et

al. (2001) similarly found that in a fringing marsh, DNRA was quantitatively more important

when nitrate supply was high during freshwater discharge in spring.

Within marine and estuarine sediments, the proportion of total nitrate reduced to ammonium by

dissimilatory pathways has often been assumed negligible, but under appropriate circumstances

it may in fact be considerable. For example, proportions of 35 to 42% (Buresh and Patrick Jr.,

1981), 32% (Goeyens et al., 1987) and 4% to 21% (Jørgensen, 1989) have been cited. In a

recent study of a shallow estuary in Texas, An and Gardner (2002) found that DNRA accounted

202

Page 217: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

for between 15 and 75% of all radio-labelled nitrate reduced in the sediment, while the

equivalent range for denitrification was only between 5% and 29%. In this case, the authors

argued that high sulphide concentrations inhibited nitrification and denitrification but enhanced

DNRA by providing an electron donor. Consequentially, ammonium was recycled efficiently

within the system and was inferred to allow the persistence of a long-lasting algal bloom.

However, as Buresh and Patrick Jr. (1981) have pointed out, care should be taken when

extrapolating results from laboratory-based results, especially where nitrate has been added

and/or sediments have been (perhaps artificially) anaerobically incubated under an inert gas

atmosphere. Whilst intensely-anaerobic sediment conditions do exist in nature, it is unclear if

much nitrate is ever exposed to such environments (Buresh and Patrick Jr., 1981). When more

natural conditions were considered in the study of Buresh and Patrick Jr. (1981), whereby

nitrate moved downward from the overlying water into the sediment, the estimate of total nitrate

dissimilatorily reduced to ammonium decreased from 35-42 to 15%. Nitrate supplied either

from nitrification or the overlying water and diffusing into anaerobic zones is quickly reduced

(Buresh and Patrick Jr., 1981). The sequential diffusion of nitrate is such that anaerobic zones

of denitrification may often be more closely related (spatially and/or temporally) to areas of

nitrate supply than the more highly anaerobic (e.g. deeper) fermentative DNRA zones. It is very

likely that this type of reasoning, sometimes justifiably, has lead to ignorance of the DNRA

process.

Nonetheless, scenarios exist whereby highly anaerobic organic-rich sediments are supplied with

substantial quantities of nitrate such that DNRA may be prolific (for example where nitrate-

laden groundwater forces a substrate supply from below or in situations where denitrification is

inhibited, by sulphide for instance). The recent identification of nitrate accumulating sulphur

bacteria that are capable of DNRA in microaerobic coastal marine and continental shelf

sediments provides an interesting example of the potential importance of the DNRA process

(Sayama, 2001; Zopfi, 2001). Three closely related genera (Thioplaca, Beggiatoa and

Thiomargarita) have been isolated from coastal upwelling zones off Peru, Chile, Namibia and

in Tokyo Bay (Sayama, 2001; Zopfi, 2001). Each species stores nitrate intracellularly in large

vacuoles at high concentrations (40 to 800 mM; Sayama, 2001). Therefore, whilst a large

nitrate pool exists in these sediments it is neither a pore-water constituent nor available for

denitrification (Sayama, 2001). Rather, these sulphur bacteria dissimilatorily reduce the stored

nitrate to ammonium which is released to the environment, potentially exacerbating

eutrophication (Sayama, 2001). The microbiology, ecology and general quantitative importance

of DNRA bacteria are still not well understood (Yin et al., 2002). However, the consequences of

DNRA, including the potential for nitrogen conservation within aquatic environments, are such

that the process warrants detailed study and at the very least deserves consideration in any

nitrogen cycling budget.

203

Page 218: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

D.7 DENITRIFICATION

The potential importance of denitrification in removing inorganic nitrogen from eutrophic

systems is of major consequence to this thesis. An explanation of the bacteria and inherent

enzymes that mediate and define denitrification, together with regulatory factors controlling this

process and typical ambient denitrification rates will therefore be discussed in some detail.

Thereafter, studies examining the kinetics of denitrification will be reviewed to place my thesis’

kinetic consideration of denitrification (Chapter 5) within the context of prior research.

D.7.1 The Bacterial and Enzymatic Process of Denitrification

Denitrification may be defined as the reduction of soluble nitrogen oxides, nitrate (NO3- ) and

nitrite (NO2-), to the gases nitric oxide (NO), nitrous oxide (N2O) and dinitrogen (N2), by

bacteria using the pathway as a respiratory alternative to oxygen (Kuenen and Robertson, 1988;

Groffman et al., 1999) (Figure D1):

NO3- NO2

- NO N2O N2

The major implication of denitrification in aquatic systems is that dissolved inorganic nitrogen

(DIN) atoms (within nitrate and nitrite) are removed from the ecosystem as they are returned to

the atmosphere as nitrogenous gases (nitric oxide, nitrous oxide and dinitrogen). Denitrification

therefore has great potential to reduce the availability of DIN to primary producers. Both

facultative anaerobes and strictly anaerobic bacteria from a wide range of taxa, including

Pseudomonas, Bacillus, Thiobacillus, Propionibacterium amongst others, are capable of

denitrification (Brezonik, 1972; Hattori, 1983; Wrage et al., 2001). Denitrifiers are

predominantly heterotrophic, facultative anaerobes respiring nitrate (instead of oxygen) as an

electron acceptor in anoxic or low-oxygen conditions, and the terminal product is usually

dinitrogen gas (Tiedje, 1988; Wrage et al., 2001). However, chemoautotrophic and

photosynthetic denitrifying bacteria exist, and whilst in all groups the end product of

denitrification must by definition be gaseous, not all reductive steps may occur (Knowles,

1982). The maximum possible reactions of denitrification by certain bacteria are given in Table

D2. For a more comprehensive consideration of denitrifying microflora the reader is referred to

Payne (1973, 1981), Jeter and Ingraham (1981), Knowles (1982) and Kuenen and Robertson

(1988). Recently, some autotrophic and heterotrophic nitrifying bacteria have been

demonstrated to be capable of denitrification in oxygen-limited or even aerobic environments

(Jetten et al., 1997). The consequences of potential denitrification by these organisms may be

great; however, the process of denitrification by nitrifying bacteria will not be considered

further here, and the reader is referred to Section D.8. This section pertains only to conventional

anaerobic denitrification. Further, although nitrate and nitrite may be reduced photochemically,

in water these processes are restricted to a very thin layer at the surface (Hamilton, 1964). In

204

Page 219: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

aquatic ecosystems only biological processes are of quantitative importance in the

transformation of nitrate (Hattori, 1983; Tiedje, 1988).

Table D2: Examples of bacterial genera which exhibit various reactions of nitrate respiration and denitrification. (Adapted from: Knowles, 1982; Kuenen and Robertson, 1988)

Maximum possible

reaction Genera exhibiting the reaction

NO3- NO2-

(nitrate respirers)

Lysobacter, Thiobacillus

NO3- N2O

(nitrous oxide producers)

Achromobacter, Aquaspirillum, Bacillus, Propionibacterium,

Pseudomonas

NO3- N2

(‘typical’ denitrifiers)

Alcaligenes, Bacillus, Halobacterium, Hyphomicrobium,

Kingella, Neisseria, Pseudomonas, Paracoccus,

Propionibacterium, Rhodopseudomonas, Thiobacillus

NO2- N2

(nitrite dependent

denitrifiers)

Achromobacter, Alcaligenes, Bacillus, Flavobacterium,

Pseudomonas, Neisseria,

Bacillus, Vibrio N2O N2

(nitrous oxide reducers)

Like all other biological components of the nitrogen cycle, denitrification is an enzyme driven

process and each of the four inherent steps is catalysed by a specific reductase: NO3- to NO2

- by

nitrate reductase (NAR); NO2- to NO by nitrite reductase (NIR); NO to N2O by nitric oxide

reductase (NOR); and, N2O to N2 by nitrous oxide reductase (NOS) (Jetten et al., 1997; Bothe et

al., 2000; Moura and Moura, 2001). The reactions involved in each respective step are detailed

below (Bothe et al., 2000):

NO3- + 2 H+ + 2 e- NO2

- + H2O

NO2- + 2 H+ + e- NO + H2O

NO + 2 H+ + 2 e- N2O + H2O

N2O + 2 H+ + 2 e- N2 + H2O

As was noted in Section D.6 above, the first step of nitrate reduction to nitrite may be common

to the DNRA process and may generally be termed nitrate respiration (Table D2), but further

reduction to a gaseous end-product must occur for the process to be deemed denitrification.

Bacterial genes encoding at least four different classes of NAR have been described (Bothe et

al., 2000; Moura and Moura, 2001). Denitrifying bacteria lacking NAR but capable of reducing

nitrite are defined as nitrite-dependant (Table D2; Knowles, 1982). Two types of NIR have

205

Page 220: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

been found in denitrifiers and for unclear reasons their presence in individual bacteria is

mutually exclusive (approximately ⅔ contain a heme-cd1 NIR whilst ⅓ contain a copper NIR)

(Jetten et al., 1997; Bothe et al., 2000; Moura and Moura, 2001). The production of nitric oxide

as a free intermediate during denitrification has historically been questioned. However, there is

now strong evidence that NO is a kinetically competent intermediate, with its steady-state

concentration ranging between 1 to 65 nM, depending on bacterial species (Jetten et al., 1997).

There is evidence of two genes encoding nitric oxide reductase (Bothe et al., 2000) but the

structure of NOR has yet to be confirmed (Moura and Moura, 2001). One type of nitrous oxide

reductase is currently defined and its structure has recently been characterised (Bothe et al.,

2000; Moura and Moura, 2001). Denitrifiers lacking NOS have nitrous oxide as their terminal

product and have been dubbed nitrous oxide producers (Table D2; Knowles, 1982). On the

other hand, certain bacteria (nitrous oxide reducers) can not produce N2O from nitrate or nitrite

(and technically are not denitrifiers), but contain NOS and therefore are capable of producing N2

(Table D2; Knowles, 1982).

An increasing build-up of both nitric oxide and nitrous oxide within the atmosphere, potentially

related to anthropogenic impacts on the global nitrogen cycle, has implications for ozone

depletion and greenhouse warming (see Section D.8; Colliver and Stephenson, 2000; Bonin et

al., 2002). Dinitrogen gas may not be the ultimate product of denitrification, and release of the

gaseous intermediates of NO and N2O may occur for a variety of reasons. Firstly, the genetic

lack of nitric oxide reductase (NOR) or nitrous oxide reductase (NOS) in certain denitrifying

bacteria physically prevents conversion to dinitrogen gas and the intermediates are in fact end-

products of the denitrification process (see Table D2). Within the majority of bacteria that

possess the full enzymatic apparatus to produce dinitrogen gas, sudden changes in the

environment may result in termination of the denitrification process at an intermediate stage.

Jetten (2001) notes that in response to environmental perturbations, bacterial regulation of the

various enzymes involved in denitrification can be immediate, and that the transition between

oxic and anoxic conditions often results in the release of intermediate gases. A limitation of

electron donors (e.g. organic carbon) or the presence of toxic and/or inhibitory compounds (e.g.

sulphide; Sørensen, 1980) may also stimulate the emission of gaseous intermediates (Jetten,

2001).

D.7.2 Factors Regulating Denitrification

Denitrification is dependent primarily upon availability of nitrate, anoxic conditions and the

presence of an energy (electron) source (Kristensen, 1988). Indeed, the supply of nitrate, the

rate of oxygen supply and consumption, and the presence of organic matter (carbon) as an

electron source are the chief regulating factors of denitrification at a particular habitat microsite

(Tiedje, 1988). These major regulatory factors are discussed below. For simplicity,

206

Page 221: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

denitrification is preposed to occur across a flat two-dimensional sediment-water interface,

described as ambient sediment, without the presence of bioturbating macrofauna. The influence

of bioturbation and the creation of burrow micro-environments considerably alters the potential

for denitrification in comparison with ambient sediment (e.g. Aller and Yingst, 1978; Henriksen

et al., 1983; Kristensen et al., 1985; Aller and Aller, 1986; Aller, 1988; Reichardt, 1988;

Kristensen et al., 1991; Svensson, 1997; De Roach et al., 2002). The impact of bioturbation,

creation of burrows and ventilatory activity by benthic infauna in regulating denitrification rates

is the subject of Chapter 5.

To a great extent, rates of denitrification appear to be limited by nitrate availability in estuarine

and coastal marine sediments. Diffusional nitrate supply from the overlying water and from

sediment nitrification, may largely regulate denitrification in these systems (Kristensen, 1984;

Henriksen and Kemp, 1988). Diffusional processes are of great importance in exchange across

the sediment-water interface and the amount of nitrate in overlying water is essential in

determining both the rate and direction of nitrate exchange (Kristensen, 1984). Kristensen

(1984) demonstrated that an equilibrial (no net flux across the sediment-water interface)

concentration of nitrate existed in overlying waters with 10-15 μM of nitrate. Nitrate diffused

into the sediment (for subsequent denitrification) at overlying concentrations above equilibrium,

whereas nitrate was released from the sediment (due to nitrification) at overlying concentrations

below equilibrium (Kristensen, 1984). The equilibrial concentration will vary for differing

sediment environments, but nitrate concentrations in overlying water will obviously affect the

diffusional rate into sediment and considerably regulate the degree (and even presence) of

denitrification activity.

In situ nitrification in sediments may also considerably regulate denitrification activity (Jenkins

and Kemp, 1984; Henriksen and Kemp, 1988) and the process has been discussed previously in

Section D.5. Generally, nitrification provides the denitrification process with a substrate

(nitrate) and may also provide denitrifiers with an anoxic environment as the aerobic nitrifiers

consume oxygen (Heathwaite, 1993). The relationship between supply of nitrate from

nitrification and consumption of those nitrates by denitrification is by no means simple. This is

primarily because the two processes require opposite redox conditions; nitrification and

denitrification are characteristically oxic and anoxic processes, respectively (Henriksen and

Kemp, 1988). Several models and quantitative studies have demonstrated that in ambient

sediment the ‘available nitrate zone’ typically is 1 to 5 cm deep (depending upon oxygen

regulatory factors, i.e. supply and consumption); and the maximum nitrate concentration is in

the uppermost 0.5 to 1 cm due to nitrification in the oxic zone, with decreasing concentration

below 1cm due to denitrification in the anoxic zone (Vanderborght and Billen, 1975;

Vanderborght et al., 1977; Kristensen, 1988; Henriksen and Kemp, 1988). The nitrate produced

207

Page 222: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

by nitrification in the oxic layers will pass either to the overlying water or enter the anoxic

sediment where it is reduced to nitrogenous gas by denitrification (Kristensen, 1988). Evidence

suggests that not all reduced nitrate resulting from nitrification is accounted for by

denitrification; the remainder is reduced to ammonium by dissimilatory pathways (Section D.6).

It is also found that temporal rather than spatial separation of nitrification and denitrification is

an effective means of relating the two processes (Henriksen and Kemp, 1988). For example,

nitrification may be a more important process in the daytime when primary producers (e.g.

microphytobenthos) are releasing oxygen for the aerobic respiration of nitrifiers, whilst

denitrification may occur at a greater rate at night under lower oxygen concentrations.

Generally, denitrification takes place at aerobic-anaerobic interfaces (Bonin and Raymond,

1990). However, there is on-going debate in the literature concerning the lowest oxygen

concentration that permits denitrification (Bonin and Raymond, 1990). Much early confusion

probably related to ignorance of aerobic nitrifier-denitrification processes (see Section D.8).

Conceptual separation of conventional anaerobic denitrification from these novel processes

should help to settle the debate in the future. Bonin and Raymond (1990) have demonstrated

that in ambient sediment denitrification was found to occur when the oxygen concentration was

increased to 5 mg l-1, but the activity was only 10% of that obtained under anaerobic conditions.

In addition, the enzymes associated with each step of denitrification were affected differently

with respect to oxygen concentration (Bonin et al., 1989). Nitrate reductase is the least

sensitive toward oxygen and its activity was completely blocked at oxygen concentrations

above 4.05 mg l-1, whereas nitrite and nitrous oxide reductases were more sensitive, the activity

of each was blocked at oxygen concentrations above 2.15 and 0.25 mg l-1 respectively (Bonin et

al., 1989). The reactions that these enzymes catalyse will therefore exhibit corresponding

sensitivity and the loss of the products dinitrogen gas, nitrous oxide and nitrite will be lost in

that order following increasing oxygen levels. Additionally, when ambient sediments are

restored to anoxic conditions following a period of oxygen enrichment, denitrification rates are

never fully restored to original anoxic rates (Bonin et al., 1989; Bonin and Raymond, 1990).

The oxygen status of a habitat microsite is controlled by the rate of oxygen supply to that site

and the rate of oxygen consumption (respiration) (Tiedje, 1988). Sediments or detritus exposed

to aerated water can be almost anoxic at the surface, providing they have sufficiently high rates

of oxygen uptake (Jørgensen and Revsbech, 1985). This can explain the occurrence of

anaerobic bacteria on exposed sediments where fully oxic conditions would intuitively be

expected (Jørgensen and Revsbech, 1985). Oxygen generally penetrates no more than a few (3

to 10) millimetres in ambient sediment (Revsbech et al., 1980; Jørgensen and Revsbech, 1985).

The thinness of this layer is due in part to the slow diffusive rate of oxygen into sediment and

also to the high sedimentary respiration rate (Jørgensen and Revsbech, 1985; Tiedje, 1988).

The extent and concentration gradient of the oxic zone in ambient sediment will regulate the

208

Page 223: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

occurrence and reaction rates of denitrifying bacteria. In response to decreasing oxygen levels,

an elevated activity and abundance of denitrifiers is generally expected with depth.

The role of carbon in the denitrification process is to provide the electron donor for nitrate

reduction (Tiedje, 1988). Nitrate will accept five electrons when reduced all the way to

dinitrogen gas, and four when reduced to nitrous oxide (see equations above; Revsbech and

Sørensen, 1990). In ambient sediment these electrons are primarily provided by the carbon

present in the detrital organic component of the sediment profile (Tiedje, 1988; Seitzinger,

1990). As detrital matter generally accumulates and cycles on the top surface layer of ambient

sediments, carbon content is expected to decrease with depth. If the available carbon used to

drive the respiration of aerobic nitrifiers (and other aerobic electron consuming processes) in the

surface layer of sediments is diminished (either in time or space), before becoming available to

denitrifying populations, then net denitrification rates may become carbon limited (Tiedje,

1988). Sediments containing low amounts of organic matter are likely to exhibit lower

denitrification rates than highly organic sediments.

In summary, the gradient profiles of oxygen, nitrate and available carbon usually define the

denitrification zone (Tiedje, 1988), as illustrated in Figure D2. Generally, denitrification rates

increase as a function of nitrate concentration but are inhibited by high oxygen concentrations.

Additionally, when there is a low availability of organic carbon to the denitrifying population,

the denitrification process may become limited by the supply of electrons.

Figure D2: Conceptualised gradient profiles of oxygen, nitrate and available carbon, and relative zone of denitrification, in sediments. (Adapted from Tiedje, 1988).

Aerobic Fermentation and sulphate reduction Denitrification respiration

O2

Con

cent

ratio

n

Sediment Depth

NO3-

Available carbon

209

Page 224: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

D.7.3 Ambient Rates and Efficiencies of Sedimentary Denitrification

An extensive review of the literature on denitrification rates measured in sediments from lakes,

streams, rivers, estuaries, coastal marine systems, continental shelf and deep sea environments

was given by Seitzinger (1990). Since that time, continued problems of nitrogenous

eutrophication and concern over increasing emissions of the ozone-depleting / greenhouse gas

intermediates of denitrification, has lead to an explosion in the number of studies addressing

denitrification in aquatic environments. It is beyond the scope of this thesis to summarise here

all of this work. A consideration of Seitzinger’s (1990) review as well as more recently

reported denitrification rates pertaining to estuarine sediments, in addition to the current state of

knowledge regarding denitrification in Australian sediments, will suffice.

Seitzinger (1990) gives the following comparisons of denitrification rates between differing

aquatic environments. Lowest denitrification rates (0.03 to 2.4 μmol N m-2 h-1) generally occur

in deep sea sediments. Denitrification rates in continental shelf sediments are in the range of 0

to 54 μmol N m-2 h-1 and are generally lower than those in estuarine and coastal environments.

Denitrification rates in oligotrophic to moderately eutrophic lakes range from 0.3 to

56 μmol N m-2 h-1, whilst generally higher rates exist in eutrophic lakes (20 to 292 μmol N m-2

h-1). Most denitrification measurements from rivers and streams are from temperate systems

that receive considerable inputs of anthropogenic nutrients, and as such, estimates of

denitrification in these environments are generally very high (40 to 2121 μmol N m-2 h-1). More

information is still required on natural rates of denitrification in unpolluted aquatic

environments. The differences in denitrification rates within and between each specified

environment are predominantly due to variation in broad-scale regulatory factors (nitrate/carbon

availability and redox conditions), and the reader is referred to Seitzinger (1990) for specific

details.

Seitzinger (1990) noted that highly polluted estuarine sediments exhibit some of the highest

sedimentary denitrification rates of all aquatic systems (e.g. >500 μmol N m-2 h-1 within the

Tama Estuary in Japan and the Tejo Estuary in Portugal); however, rates in most estuaries and

coastal marine systems range from 5 to 250 μmol N m-2 h-1. Although several more estuarine

denitrification studies have been conducted in the fifteen years since Seitzinger’s important

review, her summary is still remarkably valid. An inclusion to the list of polluted estuaries that

exhibit very high ambient denitrification rates is the River Thames Estuary in England; an

extraordinary upstream ambient denitrification rate of up to 19,916 μmol N m-2 h-1 may occur in

sediments adjacent to where London’s major sewage works discharge, but this tapers off to

between 0 and 70 μmol N m-2 h-1 at the mouth of the estuary (Trimmer et al., 2000a). The upper

rate of denitrification is by far the highest ambient activity ever reported for an aquatic

sediment, and is probably related to the fact that the Thames’ catchment has the highest rate of

210

Page 225: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

nitrate export per unit area (10 tonne N m-2 yr-1) known in the world (Trimmer et al., 2000a).

The Tamar Estuary, also in England, has been characterised as having a relatively high ambient

rate of denitrification, and may be within the range of 423 to 541μmol N m-2 h-1 during late

spring to summer (Law et al., 1991). It has also been reported that the Westerschelde Estuary in

the Netherlands had a very high ambient denitrification rate in the 1970s (640 μmol N m-2 h-1),

although following eutrophication intervention this rate declined to 335 μmol N m-2 h-1 in the

1980s (Dettman, 2001). In his review of 11 North American and European estuaries, Dettman

(2001) noted a range of average denitrification rates between 18 and 94 μmol N m-2 h-1

(excluding the Westerschelde Estuary). In fact, most estuarine denitrification rates reported

since Seitzinger’s (1990) review are within this lower range; however, notable and recently

reported exceptions of 154 μmol N m-2 h-1 (for maximal ambient day-time rates in Galveston

Bay, Texas; An and Joye, 2001) and 160 μmol N m-2 h-1 (within Waquoit Bay, Massachusetts;

LaMontagne et al., 2002) may deem that the original and more conservative range of 5 to

250 μmol N m-2 h-1 is still appropriate for typical estuarine denitrification rates. Ambient

denitrification data for estuaries outside of Europe, North America and Japan are largely

missing.

An increasing effort to quantify the proportionate contribution of denitrification to estuarine

nitrogen budgets has occurred over the last decade. Estuarine denitrification efficiencies, i.e. the

percentage of an estuary’s total nitrogen load that is denitrified, as low as 0.1 to 1.3% (Welsh et

al., 2000), 0.3 to 1.7% (Trimmer et al., 2000b) and 3 to 9% (Cabrita and Brotas, 2000), have

respectively been reported for the Bassin d’Arcachon in France, Langstone and Chichester

Harbours in England, and the Tagus Estuary in Portugal. These systems are characterised by

very high primary production by seagrasses, macroalgae and/or benthic microalgae, and

assimilatory pathways outweigh denitrification by one or two orders of magnitude.

LaMontagne et al. (2002) similarly demonstrated decreased denitrification rates in macroalgal

beds compared to bare sediments of Waquoit Bay, and suggested that increasing algal cover

could create a positive feedback as assimilation continues to proportionately out-compete

denitrification in dissolved inorganic nitrogen consumption. Nitrogen fixation may also be

prevalent in systems of proportionately lower denitrification (particularly in seagrass meadows),

and nitrification may be negligible and/or not coupled to denitrification (Welsh et al., 2000).

Meyer et al. (2001) found that denitrification was not coupled to denitrification in sediments of

Randers Fjord, Denmark. Further, Andersen et al. (1984) demonstrated that day-time oxygen

production by benthic microalgae was inhibitory to anaerobic denitrification within sediments

of Lendrup Vig, a shallow estuary in Denmark. Lowered denitrification efficiencies may also

be partly caused by short water residence times in estuaries, i.e. increased flushing and reduced

physical contact between water and sediment (Svensson et al., 2000).

211

Page 226: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

The depressed influence of denitrification in the above-mentioned systems is in contrast to the

69 to 75% denitrification efficiency range reported by Dettmann (2001) for 11 estuaries in

Europe and North America. Within these estuaries, higher denitrification efficiencies were

mainly related to longer freshwater residence times, i.e. decreased flushing (Dettmann, 2001).

While important, water residence time is not the only factor governing denitrification

efficiencies. As noted above, the presence of primary producers is of pervasive importance, and

their effect may often be to stimulate rather than depress denitrification. In their study of

Galveston Bay, Texas, An and Joye (2001) reported that denitrification efficiency (>50%) was

greater than that expected by water residence time alone. They found that oxygen production by

benthic microalgae actually enhanced rates of, and coupling between, nitrification and

denitrification (and warned that denitrification estimates of shallow estuaries may be

underestimated if the influence of benthic primary production is discounted, e.g. in dark

incubations of sediment). Oxygen production by the microphytobenthos of the River Colne

estuary, England, enhanced rates of denitrification coupled to nitrification but inhibited

denitrification of nitrate sourced from the overlying water column (Dong et al., 2000).

Denitrification efficiency in this estuary is moderate (18 to 27%; Ogilvie et al., 1997; Dong et

al., 2000). Where ammonium is the major source of dissolved inorganic nitrogen, stimulation

of nitrification may be very important in providing a substrate for denitrification. Jenkins and

Kemp (1984) demonstrated almost complete coupling of nitrification with denitrification in

sediments of the Patuxent Estuary, Chesapeake Bay, as more than 99% of nitrogen lost as

dinitrogen gas was sourced from ammonium. Tight coupling of nitrification and denitrification

has been reported for many estuaries, including the River Colne estuary (see above), Nauset

Marsh Estuary, Massachusetts (Nowicki et al., 1999), Tokyo Bay (Kuwae et al., 1998) and the

Tama Estuary, Japan (Nishio et al, 1983), amongst many others. While the relationship usually

results in at least moderate denitrification efficiencies (e.g. in the River Colne estuary, see

above), competing pathways of nitrogen consumption and other regulatory factors may lessen

the proportional importance of denitrification. For example, the denitrification efficiency of

Waquoit Bay is estimated to be between 32 and 37%, and the high importance of macroalgal

nitrogen assimilation in this system has already been described (LaMontagne et al., 2002).

Within this system, denitrification efficiency was lowered in summer and was inversely

correlated with oxygen availability and ammonium flux (LaMontagne et al., 2002). Further,

while organic matter addition did not affect denitrification efficiency in Waquoit Bay, it did

increase net rates of denitrification (LaMontagne et al., 2002). A similar positive relationship

between organic carbon and denitrification rate was also demonstrated for the nearby Nauset

Marsh Estuary (Nowicki et al., 1999).

There is obviously a great range (<1 to 75%) in the efficacy of estuaries to unload nitrogenous

inputs via denitrification. The interactions between the competing or intrinsically dependant

212

Page 227: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

nitrogen-transforming processes (e.g. ammonification, nitrification, denitrification), and

between the parameters governing these processes (e.g. nitrogenous substrate, oxygen and

carbon availability), are often complex and result in denitrification efficiencies largely specific

to individual estuarine systems. Resolution of the factors determining estuarine denitrification

efficiency requires careful consideration.

There are scarce data on the importance and rates of denitrification in Australian estuaries.

Harris (2001) recently reviewed both the scientific and grey literature for a limited number of

Australian systems. He found strong evidence that sediment denitrification becomes of greater

quantitative importance as the water residence time of estuaries increases. When residence

times are short, the ammonification and nitrification products of the bottom waters of Australian

estuaries are frequently discharged downstream (Harris, 2001). However, the long water

residence time of other estuaries and coastal lagoons can lead to denitrification efficiencies at

the high end of the global scale (Harris, 2001). For a shallow sub-tropical coastal embayment

(Moreton Bay), Eyre and McKee (2002) determined an average ambient denitrification rate of

36 μmol N m-2 h-1 and found that about 56% of the total nitrogen load was lost through

denitrification and 41% was physically exported to the ocean. Of the nitrogen not exported, this

equates to a 95% loss of nitrogen via denitrification whilst the remainder (5%) was recycled

within the system. For the heavily studied Port Phillip Bay (a temperate marine embayment),

ambient denitrification rates are about 54 μmol N m-2 h-1, with denitrification efficiencies

around 75% to 85% (Heggie et al., 1999; Nicholson and Longmore, 1999; Harris, 2001). It was

found that most of the nitrate utilised in denitrification was sourced from in situ nitrification,

and the importance of ammonification and subsequent coupled nitrification-denitrification was

highlighted (Heggie et al., 1999). Further, for several warm-temperate Australian lagoons, Eyre

and Ferguson (2002) found that ambient rates of denitrification were in the range of 8 to 69

μmol N m-2 h-1 and that denitrification efficiency was mainly regulated by carbon loading. This

latter point was also true of the cooler Port Phillip Bay sediments (Heggie et al., 1999). As

carbon load rates increased, denitrification efficiency decreased in response to either: (i) a lack

of oxygen and therefore coupled nitrification-denitrification; (ii) elevated sulphide levels and

inhibition of nitrification and denitrification; and/or, (iii) an increased proportion of

dissimilatory nitrate reduction to ammonium (Eyre and Ferguson, 2002). Furthermore the

diurnal activities of dominant primary producers in each lagoon, characteristic of different

stages of eutrophication, (i.e. seagrasses, phytoplankton, benthic microalgae and macroalgae),

were highly important in determining overall nutrient fluxes (Eyre and Ferguson, 2002). A

characteristic pattern of sedimentary N-efflux during dark conditions and N-influx during light

conditions was observed (Eyre and Ferguson, 2002). As estuarine primary producer

assemblages change in response to nutrient loading, so too does their impact on the net ratio of

productivity to respiration (p/r) and therefore on N-fluxes (Eyre and Ferguson, 2002). In fact,

213

Page 228: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

change in p/r was considered one of the key indicators of eutrophication in these estuaries. Eyre

and Ferguson’s (2002) study highlights the managerial importance of limiting carbon loading,

and maintaining a balance of autotrophy and heterotrophy, so as to facilitate better

denitrification and nitrogen removal within eutrophic systems.

D.7.4 Studies of Denitrification Kinetics

To place this thesis’ kinetic consideration of denitrification (Chapter 5) into the context of prior

research, an extensive review of the literature examining the kinetics of denitrification, nitrate

and/or nitrite reduction in various systems (estuarine/marine, freshwater, soil, pure cultures and

cell-free reductase extracts) has been undertaken. The maximal potential denitrification rates

(Vmax or Vmp) and half-saturation constants (Km or Kapp ) of reported studies, together with

methods utilised, relevant location and sediment/soil depth (where appropriate) have been

summarised in Table D3. The reader is referred to this table for reference in the following

discussion. Pertinent results of this thesis are included in the table but their implications and

relevance are considered in the discussion of Chapter 5.

Depending on the methods utilised, kinetic constants (Vmp and Kapp) reported in the literature

can pertain to either denitrification, nitrate reduction or nitrite reduction. Where nitrous oxide

or nitrogen gas production has been measured the resultant kinetic constants are specific to

denitrification, whilst measurement of the rate of nitrate or nitrite consumption refers to the

more general case of nitrate or nitrite reduction (either to nitrogenous gases or to ammonium,

perhaps incorporated into organic biomass). In many cases the rates of dissimilatory nitrate

reduction to ammonium (DNRA) and assimilatory nitrate uptake are assumed or demonstrated

to be negligible, and nitrate/nitrite consumption is approximated to be the rate of denitrification.

However, it must be noted that even under these special circumstances, the actual rate of nitrate

reduction (presumed denitrification) may be underestimated if nitrate production via

nitrification is not accounted for. Generally, the distinction must be made between reported

kinetic constants for nitrate/nitrite reduction whereby end products may not necessarily be

gaseous, and kinetic constants specifically pertaining to denitrification. In kinetic studies

relating to bacterial populations in sediment or soil environments it is also essential to consider

reported sediment/soil depth increments or total bulk. Where Vmp is reported on a per area

basis, the depth to which nitrate reduction/denitrification is obviously important. A deeper

sediment/soil column equates to a larger volume where bacterial activity can occur, and

therefore a potentially greater population size and value of Vmp. However, it is often found that

denitrification and nitrate/nitrite reduction activity tapers off very quickly in deeper sediment

where substrate supply is limited (Table D3). It is assumed that most studies incorporate the

relevant zone where the majority of denitrification or nitrate/nitrite reduction occurs, but this

may not always be the case. The affinity of bacteria for nitrate/nitrite (as measured by Kapp) in

redox-stratified sediment/soil is also expected to vary for different depth increments.

214

Page 229: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

The maximum potential rate of denitrification (Vmp) in estuarine or marine sediments reported

by studies utilising a kinetic approach ranges between 0.04 and 3.33 mmol N m-2 h-1

respectively within the Potomac and Choptank Tributaries of Chesapeake Bay (Table D3;

Twilley and Kemp, 1986). Variation in salinity, nitrate and organic carbon availability was

proposed to account for the wide range in maximal denitrification rates (Twilley and Kemp,

1986). Kinetic studies of sediments off the coast of California (Oremland et al., 1984; Joye et

al., 1996), France (Raymond et al., 1992), Denmark (Oren and Blackburn, 1979) and England

(Dong et al., 2000) have demonstrated maximal denitrification rates intermediate to the reported

range (Table D3), thus highlighting the capacity for denitrification potentials to vary greatly

within a limited geographical range (i.e. within the vicinity of Chesapeake Bay). The reported

range of maximum potential rates of general nitrate reduction is broader (Table D3). Absence

of nitrate reduction was reported for sediments of the North Sea off the Belgian coast (Billen,

1978) whereas a maximal nitrate reduction rate of 12.9 mmol N m-2 h-1 was demonstrated for an

intertidal salt marsh in the United Kingdom (Nedwell, 1982). Within the North Sea sediment,

Billen (1978) contends that nitrate reduction is reflective of denitrification rate and that activity

is absent in the oxidised layer (up to 8 cm deep) of certain sediments. The high rate of maximal

nitrate reduction in the deep (30 cm) salt marsh sediment is argued mainly to result in

denitrification, although Nedwell (1982) estimates an 8% contribution by DNRA and a

potentially small assimilatory consumption of nitrate.

In estuarine or marine environments, the affinity of denitrifying populations for nitrate as

judged by the concentration required for half-maximal activity (Kapp), ranges from less than

2 μM (high affinity) measured during September in the top 3 cm of sediment of Tomales Bay,

California (Joye et al., 1996), to 1818 μM (low affinity) determined in the 2 to 4 cm depth

interval of sediment from Carteau Cove, France (Table D3; Raymond et al., 1992). Reported

Kapp values for both denitrification and more general nitrate reduction for other estuarine and

marine areas, including off the coasts of Fiji (Nedwell, 1975), Japan (Koike et al., 1978; Koike

and Hattori, 1979), France (Esteves et al., 1986), Denmark (Oren and Blackburn, 1979),

Belgium (Billen, 1978), England (Nedwell, 1982; Dong et al., 2000) and North America (Koike

et al., 1978; Capone and Taylor, 1980; Iizumi et al., 1980; Oremland et al., 1984; Twilley and

Kemp, 1986), are intermediate to these extremes (Table D3). The variation in Kapp values has

often been correlated with natural nitrate availability, whereby communities exhibiting a high

affinity (low Kapp) are adapted to an environment of low nitrate concentration or supply, whilst a

low affinity (high Kapp) is more reflective of a community adapted to an ample or abundant

nitrate supply (Nedwell, 1975; Joye et al., 1996; García-Ruiz et al., 1998). Stated conversely,

nitrate reduction would be inefficient by high affinity denitrifying bacteria in an environment of

high nitrate concentration, and low affinity bacteria would likewise be inefficient in a low

nitrate concentration environment (García-Ruiz et al., 1998). Elevated nitrate availability may

215

Page 230: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

alter either cell-specific enzyme activity (i.e. increased activity of individual denitrifying

bacteria) or enhance growth rates (i.e. an increased number of bacteria), both of which may lead

to a greater Kapp; however, separation of denitrifier biomass and enzyme expression is

technically challenging (Joye et al., 1996). Variations in Kapp have also been observed for

similar ambient nitrate concentrations, and related to differing composition or physiology of the

denitrifying community, variation in organic carbon availability (as an electron donor),

dissimilar redox conditions and/or differences in transport rate of nitrate (via diffusion or active

processes) to active sites of denitrification (Nedwell, 1975; Joye et al., 1996; Strong and Fillery,

2002). A combination of these factors (natural ambient nitrate, oxygen and organic carbon

concentration; abundance, composition and physiology of the adapted denitrifying community;

rate of transport to active denitrifying zones) including a consideration of relevant temporal

changes in these parameters, explains the large variation of reported Kapp values both within and

between geographical regions (large and fine scale) and over time.

Furthermore, Myrold and Tiedje (1985) suggest that dissimilatory nitrate reducers may possess,

and be able to switch between, both high and low affinity nitrate uptake mechanisms. This

explanation would account for the apparent discrepancy between reportedly high Km / low

affinity values (250 – 1300 μM) for enzyme extracts of nitrate reductases (Myrold and Tiedje,

1985), and the generally much lower Km / high affinity values (1.7 – 15 μM) reported for pure

cultures of nitrate reducers (Table D3; Betlach and Tiedje, 1981; Edwards and Tiedje, 1981;

Parsonage et al., 1985; Christensen and Tiedje, 1988). The exception of a low affinity (Km =

260 μM) Pseudomonas pure culture (Brown et al., 1975) is important, and Myrold and Tiedje

(1985) argue that pure culture studies have generally not been conducted at high enough nitrate

concentrations to detect a low affinity system.

The number of kinetic studies conducted within estuarine, marine and other environments is

somewhat limited and it is insightful to consider kinetic constants reported for freshwater

sediments and soil. Within lake and river sediments of Denmark (Anderson, 1977), England

(García-Ruiz et al., 1998) and the U.S.A. (Messer and Brezonik, 1984), maximum potential

rates of denitrification or nitrate/nitrite reduction (Vmp) are generally within the range

determined for estuarine/marine sediments (Table D3); although higher nitrate reduction Vmp’s

were reported for farm pond sediments in Kentucky, U.S.A (12.4 to 26.6 mmol N m-2 h-1;

Murray et al., 1989), and were even higher in sediments of Lake Vechten, Denmark (24.0 to

120 mmol N m-2 h-1; Hordijk et al., 1987). Likewise, reported Kapp values of freshwater

sediments are within the range of marine/estuarine sediments, although very high (low affinity)

half saturation constants have been reported for nitrate reduction in lakes Kvind sø (Kapp = 2856

μM) and Kalgaard sø (Kapp = 6024 μM) of Denmark (Anderson, 1977).

216

Page 231: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Whilst reported Kapp values of soils are similarly variable and generally within the range

determined for sediments (estuarine, marine or freshwater), even higher (lower affinity) half-

saturation constants may exist in soil environments (Table D3). For example, a fine sandy soil

from Coachella, California had a nitrate reduction Kapp of 12,137 μM (Bowman and Focht,

1974). The lower affinities of bacterial communities for nitrate within agricultural soil

presumably largely relates to higher nitrate supply and ambient concentration due to

fertilisation. Higher fertiliser-nitrate availability may also partially explain the generally higher

maximal rates of denitrification, nitrate and/or nitrite reduction (Vmp) in soils of New Zealand

(Schipper et al., 1993), Denmark (Ambus, 1993; Maag et al., 1997), Sweden (Klemdtsson et al.,

1977), Russia (Ryzhova, 1979), U.S.A (Bowman and Focht, 1974; Kohl et al., 1976; Myrold

and Tiedje, 1985; Abdelmagid and Tabatabai, 1987; Murray et al., 1989) and Canada

(Yoshinari et al., 1977) compared to estuarine/marine sediments (Table D3). The Vmp range of

soils is typically between 1.38 mmol N m-2 h-1 (for nitrite reduction in a riparian agricultural soil

in Copenhagen, Denmark; Ambus, 1993) to 512 mmol N m-2 h-1 (for nitrate reduction in soil

under a forb-grass meadowland within Moscow Oblast, Russia; Ryzhova, 1979). The only

reported exceptions, with much lower denitrification Vmp’s (0.003 to 0.89 mmol N m-2 h-1), have

been determined within south-western Australia for poorly nutrient-retentive, sandy soils under

arable crop or pasture (Strong and Fillery, 2002). The implication is important within the

regional context of this thesis, since the comparatively low maximal rate of terrestrial

(catchment) denitrification suggests that a potentially higher load of nitrate is available for

export to the aquatic systems of south-western Australia. These low terrestrial denitrification

Vmp’s therefore may be a contributing factor to the recent eutrophication of many of these

systems. To the best of the author’s knowledge, apart from this thesis, no studies examining the

kinetics of nitrate reduction or denitrification have been undertaken for Australian estuarine,

marine or freshwater sediments.

217

Page 232: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Table D3: Half saturation constants (Km or Kapp) and maximal rates (Vmax or Vmp) of denitrification, nitrate and nitrite reduction of various systems.

Location Process Sediment or

soil depth range (cm)

Km or Kapp (μM NO3

-)

Vmax or Vmp (converted# to

mmol N m-2 h-1) Vmax or Vmp

[reported units] Method Reference

Marine/Estuarine Sediment Swan River Estuary, Western Australia - A. ehlersi inhabited sediment - uninhabited sediment - C. aequisetis inhabited sediment

Nitrate Reduction 0 – 5.5 0 – 4.7 0 – 4.7

50

133 202

4.75 4.33 3.83

[mmol N m-2 h-1] 4.75 4.33 3.83

C2H2 block, NO3-

consumption above in vitro sediment cores.

This study

Vatuwaga River Estuary, Fiji (polluted mangrove sediment) - upstream, adjacent to effluent outfall - mouth, 2.5 km from effluent outfall

Nitrate Reduction

NR NR

180 600

3.65 1.09

[mmol N m-2 h-1]

3.65 1.09

In situ NO3-

consumption of overlying water.

Nedwell 1975

Tokyo Bay, Japan Mangoku-Ura, Japan

Denitrification Denitrification

NR

NR

24

27 – 42

[μmol N g-1 h-1] 0.11

0.017 – 0.019

15NO3- 15N2.

Sediment slurries. Koike et al. 1978

Bering Sea, Japan

Denitrification 0 – 2

~ 4

NR

15NO3- 15N2.

Sediment slurries. Koike and Hattori 1979

Port Cros Island, France Nitrate Reduction 0 – 1.2a

78

1.47

[μmol N cm-3 d-1] 2.94

NO3- consumption

of in vitro sediment cores.

Esteves et al. 1986

West Mediterranean, France

- Lavera - Berre - Cassis - Carry le Rouet - Port Cros - Carteau

Denitrification

0 – 2 0 – 2 0 – 2 0 – 2 0 – 2

0 – 2 0 – 4 0 – 6 0 – 8

0 – 10

228 256 499 539 560

1428 1818b

740b

1666b

1428b

0.39 0.32 0.25 0.25 0.25

0.50 0.66 0.68 0.71 0.81

[μmol l-1 (sediment) d-1]

464 386 302 298 300

600 192b

20b

37b

120b

C2H2 block, N2O production. Sediment slurries.

Raymond et al. 1992

218

Page 233: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Table D3: (continued)

Location Process Sediment or

soil depth range (cm)

Km or Kapp (μM NO3

-)

Vmax or Vmp(converted# to

mmol N m-2 h-1)

Vmax or Vmp[reported units] Method Reference

Kysing Fjord, Denmark

Denitrification 0 – 1

344

0.18

[nmol N cm-3 d-1] 422

15NO3- 15N2.

Sediment slurries. Oren and Blackburn 1979

North Sea, Belgium (coastal ‘zone’) - Station M1230 (April 1975) - Station M1283 (March 1976) - Station M01 (August 1975) - Station M01 (March 1976) - Station M11 (April 1975)

Nitrate Reduction 10 – 15

0 – 5 0 – 3

20 – 25 6 – 11 0 – 5 0 – 8

50 1.17 0.58 0.49 0.81 0.14

0 0

[μmol N cm-3 s-1] 6.5 x 10-6

3.2 x 10-6

4.5 x 10-6

4.5 x 10-6

0.75 x 10-6

0 x 10-6

0 x 10-6

NO3- consumption

Sediment slurries. Billen 1978

Colne Point, Colchester, U.K. (intertidal salt marsh sediment)

Nitrate Reduction 0 – 30

319 ± 114

12.9 ± 2.11

[mmol N m-2 h-1] 12.89 ± 2.11

NO3- consumption

above in vitro sediment cores.

Nedwell 1982

River Colne Estuary, Colchester, U.K. (sewage affected intertidal muddy sediment)

Denitrification 0 – 30

188

0.22

[μmol N m-2 h-1] 222

15NO3- 15N2.

Sediment slurries. Dong et al. 2000

[nmol N cm-3 d-1] Tomales Bay, California, U.S.A.

Denitrification

0 – 3 0 – 6 0 – 9

0 – 12 0 – 15 0 – 18 0 – 21 0 – 24 0 – 27 0 – 30

Marc

53 148 19 34

103 68 28

172 63 NR

Sepc

<2 2 5 4 5 5 5 7

10 12

Mar 0.20 0.40 0.43 0.52 0.63 0.70 0.77 0.86 0.92 0.95

Sepd

NR 0.32 0.52 0.69 0.92 1.09 1.32 1.42 1.51 1.58

Marc

163 161 20 73 87 60 52 72 46 29

Sepc

NR 252 160 137 187 137 182 85 67

C2H2 block, N2O production.

60

Sediment slurries. Joye et al. 1996

South San Francisco Bay, U.S.A. (polluted intertidal mudflat)

Denitrification 0 – 7

50

0.36

[nmol 50cm-3 h-1] 256

C2H2 block, N2O production. Sediment slurries.

Oremland et al. 1984

219

Page 234: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Table D3: (continued)

Location Process Sediment or

soil depth range (cm)

Vmax or VmpKm or Kapp Vmax or Vmp(converted# to (μM NO3-) [reported units] mmol N m-2 h-1)

Method Reference

Soldier Key, Florida, U.S.A. (seagrass rhizosphere sediment)

Denitrification NR

~ 20

NR

C2H2 block, N2O production. Sediment slurries.

Capone and Taylor 1980

Izembek Lagoon, Alaska, U.S.A. (seagrass bed sediment)

Denitrification 0 – 3

53

NR

15NO3- 15N2.

Sediment slurries Iizumi et al. 1980

Maryland, U.S.A. - Chesapeake Bay (Centre Basin Sth) - Chesapeake Bay (Centre Basin Nth) - Potomac Tributary (Downstream) - Potomac Tributary (Upstream) - Patuxent Tributary (Downstream) - Patuxent Tributary (Upstream) - Choptank Tributary (Downstream) - Choptank Tributary (Upstream)

Denitrification 0 – 1.5 0 – 1.5 0 – 1.5 0 – 1.5 0 – 1.5 0 – 1.5 0 – 1.5 0 – 1.5

8.4 8.5 2.0 67 22 78 11 92

0.10 0.10 0.04 1.86 0.86 0.37 1.16 3.33

[nmol N cm-3 h-1] 6.99 6.67 2.65 124 57.1 24.4 77.3 222

C2H2 block, N2O production. Sediment slurries.

Twilley and Kemp 1986

Saanich Inlet, British Columbia, Canada - Non-polluted - Mercury polluted

Denitrification NR 31

189

– –

[μmol N g-1 h-1] 1.9 3.0

15NO3- 15N2.

Sediment slurries. Koike et al. 1978

Freshwater Sediment [mg N m-2 d-1] Lakes of Denmark

- Frederiksborg Slotsø - Almind sø - Esrom sø - Kalgaard sø - Bryrup Langsø - Kvind sø

Nitrate Reduction

0 – 25 0 – 25 0 – 25 0 – 25 0 – 25 0 – 25

-O2 7.5 170 177 363 612 892

+O2599 548 376 6024 451 2856

-O2 0.04 0.54 1.10 0.54 4.25 1.49

+O20.48 0.53 1.14 3.10 1.06 4.25

-O2 11.9 182 370 182 1429 500

+O2161 179 385 1042 357 1429

NO3- consumption

above in vitro sediment cores.

Andersen 1977

Lake Vechten, Denmark Nitrate Reduction 0 – 8

17 – 100e

24.0 – 120f

[μmol N cm-3 d-1] 7.2 – 36e

NO3- consumption.

Sediment slurries. Hordijk et al. 1987

Lowland streams Arhus Å, Gudenå and Døde Å, Jutland, Denmark

Denitrification 0 – 8 <10 – NR C2H2 block, N2O microsensor, NO3

- profile modelling in sediment cores.

Christensen et al. 1989

220

Page 235: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Table D3: (continued)

Location Process Sediment or

soil depth range (cm)

Vmax or VmpKm or Kapp Vmax or Vmp(converted# to (μM NO3-) [reported units] mmol N m-2 h-1)

Method Reference

Rivers of north-east England, U.K. - Swale-Ouse River System - Wiske Tributary (polluted)

Denitrification 0 – 25 0 – 25

13 – 90

351 – 640

0.04 – 0.32

1.19

[μmol N m-2 h-1] 36 – 324

1194

C2H2 block, N2O production within in vitro sediment cores.

Garcia-Ruiz et al. 1998

Lake Okeechobee, Florida, U.S.A.

Denitrification 0 – 10

218

7.78 – 8.94g

[μg N g-1 h-1] 1.09

C2H2 block, N2O production. Sediment slurries.

Messer and Brezonik 1984

Spindletop Farm, Kentucky, U.S.A. (pond sediment) - May 1987 - September 1987 - June 1988

Nitrate Reduction

0 – 10h

0 – 2 0 – 4 0 – 6

0 – 2 0 – 4 0 – 6

18.3

NR 1.8b

10.1b

2.4

10.1b

4.8b

12.4i

2.20ij

8.77i

12.6i

4.68i

18.2i

26.6i

[nmol N g-1 min-1]

1.80

NR 4.76b

2.75b

3.39 9.80b

6.06b

C2H2 block, NO3-

consumption. Sediment slurries.

Murray et al. 1989

Soil Esperance, Katanning and Qualeup, south-western Australia (sandy soils under arable crop or pasture)

Denitrification 0 – 20

71 - 7139

0.003 – 0.89

[g N ha-1 d-1] 10 – 3000

C2H2 block, N2O production above in vitro soil cores.

Strong and Fillery 2002

Tairua Forest, Coromandel Peninsula, North Island, New Zealand (riparian soil receiving sewage irrigation)

Denitrification 0 – 30

150

7.07

[μg N g-1 h-1] 1.1

C2H2 block, N2O production. Soil slurries.

Schipper et al. 1993

Copenhagen, Denmark (riparian soil) k

- wet meadow near rural stream

Denitrification

0 – 5 0 – 10 0 – 25 0 – 35

0 – 425

26 21l

NR 6m

1.3n

3.78 – 5.40 5.58 – 8.40 9.52 – 14.7p

11.1 – 17.1 31.2 – 44.1

[μmol N l-1 (water) min-1] 1.8 1.0l

NR 0.4m

0.02n

C2H2 block, N2O production. Soil slurries.

Maag et al. 1997

221

Page 236: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Table D3: (continued)

Location Process Sediment or

soil depth range (cm)

Vmax or VmpKm or Kapp Vmax or Vmp(converted# to (μM NO3-) [reported units] mmol N m-2 h-1)

Method Reference

- reedswamp adjacent to lake: low N load - reedswamp adjacent to lake: high N load

0 – 5 0 – 10 0 – 20 0 – 30 0 – 80 0 – 5

0 – 10 0 – 20 0 – 30 0 – 80

43 15l

4.7m

NR 43o

90 13l

2.1m

NR 2.2o

0.92 – 1.08 2.45 – 2.88 4.51 – 5.46 6.09 – 7.50p

10.6 – 14.7 4.59 – 5.40 4.87 – 5.73 5.40 – 6.39 5.96 – 7.11p

9.11 – 11.3

0.36 0.6l

0.43m

NR 0.24o

1.8 0.11l

0.11m

NR 0.14o

Copenhagen, Denmark (agricultural riparian soil)

Nitrate Reduction Nitrite Reduction

0 – 10

0 – 10

4.2 ± 5.0

6.3 ± 1.0*

1.69 – 2.32q

(± 0.15 – 0.21)

1.38 – 1.90q

(± 0.84 – 1.16)

[μmol l-1 (slurry) min-1]

2.2 ± 0.2

1.8 ± 1.1

C2H2 block, NO3-

or NO2-

consumption. Soil slurries.

Ambus 1993

Funbo-Lövsta, Uppland, central Sweden (mineral soil)

Denitrification NR

228

[μg N g-1 h-1] 2.3

C2H2 block, N2O production. Soil slurries.

Klemdtsson et al. 1977

Moscow Oblast, Russia (soil under forb-grass meadowland)

Nitrate Reduction 0 – 10

3141

347 – 512r

[mg N l-1 d-1] 222 NO3

- consumption. Soil slurries. Ryzhova 1979

Spindletop Farm, Kentucky, U.S.A. - Maury fine-silty soil under sod - Lanton fine-silty fertilised soil with corn - Lanton fine-silty soil under sod - Lanton fine-silty soil under sod

Nitrate Reduction Nitrite Reduction

0 – 10 0 – 10 0 – 10 0 – 10

3.9 – 19 14 – 17 5.8 – 34

2.7*

3.59 – 4.14s

3.52 – 4.42s

22.0 – 50.2s

43.6s

[nmol N g-1 min-1] 0.52 – 0.60 0.51 – 0.64 3.19 – 7.27

6.31

C2H2 block, NO3-

consumption. Soil slurries.

Murray et al. 1989

Illinois, U.S.A. - fertilised silt loam with corn / soybean - non-fertilised silt loam with bromegrass

Nitrate Reduction 0 – 30 0 – 30

290 3477

16.4t

133t

[mg N l-1 d-1] 9.6

77.7

NO3- consumption

Soil slurries. Kohl et al. 1976

222

Page 237: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Table D3: (continued)

Location Process Sediment or

soil depth range (cm)

Vmax or VmpKm or Kapp Vmax or Vmp(converted# to (μM NO3-) [reported units] mmol N m-2 h-1)

Method Reference

Iowa, U.S.A. - Ames soil - Okoboji soil

Nitrate Reduction 0 – 15 0 – 15

3700 2900

59.9 – 76.2u

61.8 – 78.7u

[μg N g-1 d-1] 122 126

NO2- production.

Soil slurries. Abdelmagid and Tabatabai 1987

East Lansing, Michigan, U.S.A. (agricultural clay loam soil)

Denitrification NR

450

NR

C2H2 block, N2O production. Soil slurries.

Myrold and Tiedje 1985

Coachella, California, U.S.A. (fine sandy soil)

Nitrate Reduction NR

12137

29.0v

[μg N ml-1 d-1] 150 NO3

- consumption. Soil slurries.

Bowman and Focht 1974

St Bernard, Quebec, Canada - sandy loam + glucose - sandy loam – glucose

Denitrification NR 1214 129

– NR C2H2 block, N2O production. Soil slurries.

Yoshinari et al. 1977

Pure Culture - Paracoccus denitrificans Nitrate Reduction – <5 – NR NO3

- consumption. Parsonage et al. 1985 - Alcaligenes sp.

- Flavobacterium sp. - Pseudomonas fluorescens

Nitrate Reduction Nitrite Reduction Nitrate Reduction Nitrite Reduction Nitrate Reduction Nitrite Reduction

– <15

12.9*

<15

5.6*

<15

5.5*

– NR

NO3- or NO2

- consumption.

Betlach and Tiedje 1981

- Pseudomonas fluorescens Nitrate Reduction – 5 – 10 – NR NO3- consumption. Edwards and Tiedje

1981

223

Page 238: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Table D3: (continued)

Location Process Sediment or

soil depth range (cm)

Vmax or VmpKm or Kapp Vmax or Vmp(converted# to (μM NO3-) [reported units] mmol N m-2 h-1)

Method Reference

- Pseudomonas chlororaphis - Pseudomonas aureofaciens

Denitrification – 1.7 1.8

– 47.6 ± 0.8 51.3 ± 2.9

[μg N g-1 cells min-1]

N2O production. Christensen and Tiedje 1988

- Pseudomonas sp. (strain PL1 - isolate from Trondheims Fjord, Norway)

Nitrate Reduction – 260 – NR NO3- consumption. Brown et al. 1975

Cell-free Reductase Extract Dissimilatory nitrate reductase from: - Pseudomonas aeruginosa - Micrococcus denitrificans - Micrococcus halodenitrificans

Nitrate Reduction – 300 – 500

250 1300

– NR Myrold and Tiedje 1985 (review) NO2

- production.

Notes

NR - Not Reported. * - Nitrite concentration. # - For comparative purposes and where possible, reported Vmax or Vmp units were converted to mmol N m-2 h-1 by utilising available information on soil/sediment depth and/or bulk density

and/or water content (see References for further data):

a - Sediment height estimated from Figure 1a of Esteves et al. (1986).

b - Km and Vmax values are for the deepest 2 cm increment within the depth range.

c - Estimates of Kapp and Vmp for each 3 cm depth interval were respectively taken from Figures 3 and 4 of Joye et al. (1996). Values are for the deepest 3 cm increment within the depth range.

d - September values are underestimates since they do not include the Vmp data for the top 3 cm of sediment (not reported).

e - Denotes range of Km or Vmax values within depth interval (incomplete information regarding values at specific depths within interval).

f - Lower estimate assumes lowest Vmax throughout entire depth interval; vice versa for higher estimate.

g - Assuming a wet sediment bulk density of 1 to 1.15 g cm-3 (which must be true for the reported sediment water content of 87%, whereby the collective sediment is more dense than water).

h - Sediment assumed to be sampled to a depth of 10 cm.

i - Assuming sediments have a ‘fine-silty’ bulk density of 1.15 g cm-3, i.e. similar to the surrounding (often flooded) soil (see ‘s’).

j - Assuming September 1987 (0 – 2 cm) estimate of Vmax ≅ ½ June 1988 (0 – 2 cm) value (as are the values for the 0 – 4 cm and 0 – 6 cm depth increments in September 1987 compared to June 1988).

k - Km and Vmax values are for the deepest 5 cm (l), 10 cm (m), 25 cm (n) or 50 cm (o) increments within the depth range. Converted Vmax values are based on estimates of soil water content (1 – dry matter) taken from Figure 1 of Maag

et al. (1997); and assume that the collective soil is more dense than water. For the depth increments where Vmax values were not reported (p), the converted Vmax estimate assumes that the missing increment equals the average of the

Vmax values immediately above and below that increment.

q - Assuming 12.8 to 17.6% of slurry is original wet soil, by volume. (Slurry = 15g original wet soil + 70 ml buffer; reported original wet soil water content = 68.6 to 81.6%; and, whereby the collective soil is more dense than water).

224

Page 239: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

225

Notes (continued)

r - Given the slurry mix of [20 g dry soil : 100 ml solution], and where soil type is undefined the lower estimate assumes a ‘clayey’ bulk density (1.05 g cm-3), whilst the upper estimate assumes a ‘sandy’ bulk density (1.55 g cm-3).

s - Assuming all soils are sampled to a depth of 10 cm and have a ‘fine-silty’ bulk density of 1.15 g cm-3.

t - Given the slurry mix of [75 g dry soil : 125 ml solution], and assuming a typical in situ silt loam bulk density of 1.15 g cm-3.

u - Range assumes a bulk density between 1.1 and 1.4 g cm-3.

v - Given the slurry mix of [50 g soil : 50 ml solution], assuming soil is taken from the surface 5 cm of the profile and dried, and has an in situ ‘fine-sandy’ bulk density of 1.3 g cm-3.

Page 240: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

D.8 NOVEL PROCESSES OF NITROGENOUS TRANSFORMATION AND

GAS PRODUCTION

The biological processes of nitrogen transformation described above conventionally are

considered to be the major pathways of nitrogen conversion in natural environments.

Traditionally, denitrification has been assumed to be the major source of gaseous nitrogen (NO,

N2O and N2) from inorganic derivatives, and thereby an all-important nitrogen sink in aquatic

environments. Nitrous oxide is a 200-fold more effective greenhouse gas than carbon dioxide,

whilst nitric oxide is implicated in a set of key reactions that cause depletion of the ozone layer

(Colliver and Stephenson, 2000; Bonin et al., 2002). Concern over increasing concentrations of

N2O and NO in the atmosphere (an annual rate of 0.2 to 0.3% for N2O,) potentially linked to

anthropogenic activities (such as mass global fertilisation), in addition to discrepancies in

nitrogen budgets of waste-water treatment and natural systems, has triggered more thorough

research into the sources of gaseous nitrogen production (Colliver and Stephenson, 2000; Bonin

et al., 2002). Recent identification of a variety of novel microbial nitrogen conversion

pathways, including nitrifier-denitrification, anaerobic ammonia oxidation (anammox) and

heterotrophic nitrification linked with aerobic denitrification, has challenged conventional

thought and forced a re-assessment of possible nitrogen transfers and sinks in aquatic

environments. When rationalising ecosystem nitrogen budgets, a consideration of the potential

for each of these processes to occur is now necessary. The characterisation and general nitrogen

cycling implications of these novel microbial pathways have been extensively reviewed by

Jetten et al. (1997), Jetten (2001) and Zehr and Ward (2002), and are summarised below.

D.8.1 Nitrifier-denitrification and Chemodenitrification

Autotrophic ammonium oxidisers (nitrifiers) have been implicated in the production of the

gases dinitrogen (N2), nitrous oxide (N2O) and nitric oxide (NO) in many marine environments

(Capone, 2000; Wrage et al., 2001). While the mechanisms of gaseous nitrogen production by

ammonium oxidisers are still unclear, they are thought to produce mainly the nitrogenous gases

in a process termed nitrifier-denitrification (Bonin et al., 2002). In nitrifier-denitrification,

ammonium oxidisers are capable of reversing what is considered to be their natural reaction

process by reducing nitrite in the presence of low levels of oxygen (Colliver and Stephenson,

2000). The oxidation of ammonium to nitrite (NH4+ NO2

-) is followed by the reduction of

nitrite to nitric oxide, nitrous oxide and/or dinitrogen gas (NO2- NO N2O N2) (Wrage et

al., 2001). The end product is dependant on the species involved and the presence of a suitable

electron acceptor, but importantly the nitrification and denitrification path both occur in the

same organism (Colliver and Stephenson, 2000; Wrage et al., 2001). Ammonium oxidisers that

are capable of undertaking this kind of denitrification (so far only isolates from the genus

Nitrosomonas) are thought to possess and use essentially the same enzymes that are involved in

denitrification by conventional denitrifying bacteria (Colliver and Stephenson, 2000; Wrage et

226

Page 241: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

al., 2001). Since the denitrification step usually only occurs under conditions of

microaerophily, nitrifier-denitrification is sometimes also termed oxygen-limited autotrophic

nitrifier-denitrification or OLAND (Kuai and Verstraete, 1998; Philips et al., 2002). A

particular point should be made that nitrifier-denitrification by autotrophs is distinct from the

two-or-more organism process of coupled nitrification-denitrification, whereby conventional

denitrifying bacteria reduce nitrate and nitrite (produced by nitrifying bacteria) to N2O and N2

(Jetten, 2001; Wrage et al., 2001). Autotrophic nitrifier-denitrification is also different to

heterotrophic nitrification linked with denitrification in the presence of oxygen (see below)

(Wrage et al., 2001).

Alternatively, N2O may be formed from decomposition of the intermediates of ammonium

oxidation (e.g. hydroxylamine - NH2OH, see Section D.5) or nitrite itself (Kaplan, 1983; Wrage

et al., 2001). These processes are usually regarded as a special case of chemodenitrification and

probably only constitute a small fraction of the nitrogenous gases produced by ammonium

oxidisers (Colliver and Stephenson, 2000; Wrage et al., 2001). Additionally, the anaerobic

oxidation of ammonia and amino-level nitrogen in organic compounds to dinitrogen gas via

inorganic (non-biological) catalysis by manganese oxide has been demonstrated in some

sediments (Ward, 2000); however, the quantitative significance of this process is questionable

(Thamdrup and Dalsgaard, 2000).

The contribution of nitrifier-denitrification and other processes within sediments, surface waters

and the marine environment to atmospheric flux of and NO, N2O and N2, is only just beginning

to be quantified. Together with incomplete conventional denitrification by anaerobic

denitrifying bacteria (see Section D.7), nitrification by ammonium oxidisers is thought to be a

major biological nitrous oxide producing process (Colliver and Stephenson, 2000; Bonin et al.,

2002). Bonin et al. (2002) found for marine surface layers of the Mediterranean Sea that

nitrification is the dominant process that produces nitrous oxide, whilst incomplete

denitrification accounted for more production in surface waters near the mouth of a river and

near the bottom. Estuaries are thought to be an important source of atmospheric N2O, whereby

production is associated with high rates of both nitrification and denitrification (de Bie et al.,

2002). Net fluxes of N2O to the atmosphere will depend on the factors controlling these

processes, e.g. ammonium and nitrate supply, respectively (de Wilde and de Bie, 2000; de Bie

et al., 2002).

D.8.2 Anaerobic Ammonium Oxidation (Anammox)

It was discovered recently that ammonium oxidation can also occur in the absence of oxygen

(Mulder et al., 1995). Under anaerobic conditions, ammonium is oxidised with nitrite serving

227

Page 242: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

as the electron acceptor with an end-product of dinitrogen gas (Kuypers et al., 2003). The

process has been coined ‘anammox’ (anaerobic ammonium oxidation) and can be denoted as:

NH4+ + NO2- N2 + 2 H2O

Bacteria that are capable of this reaction belong to the order Planctomycetales. Isolates from

wastewater treatment plants and natural environments have been provisionally named and

include two freshwater species (Brocadia anammoxidans and Kuenenia stuttgartiensis) and one

marine species (Scalindula sorokinii) (Strous et al., 1999; Jetten et al., 2003). All three species

contain a distinctive organelle, the anammoxsome, which houses a unique set of enzymes to

catalyse the characteristic reaction (Jetten et al., 2003). Like all planctomycetes, anammox

bacteria grow very slowly, dividing only once every 11 to 14 days (Strous et al., 1999; Jetten et

al., 2003). The rate of anaerobic ammonium oxidation may be substrate dependent. Ammonium

typically is plentiful in anoxic ecosystems, whilst nitrite may be provided by nitrate reducing

bacteria (nitrate respirators including denitrifiers and bacteria capable of dissimilatory nitrate

reduction to ammonium) or microaerophilic ammonium-oxidising bacteria (nitrifiers) (Jetten et

al., 2003). In systems with limited oxygen supply it is thought that these bacteria may occur in

close association with anammox bacteria (Schmidt et al., 2002; Jetten et al., 2003).

Additionally, anaerobic ammonium oxidation also has been recently exhibited in the

conventionally aerobic (or microaerophilic) nitrifier Nitrosomonas eutropha (Schmidt and

Bock, 1997; Schmidt et al., 2002). The process is similar to that described for aerobic nitrifier-

denitrification; however, nitrogen tetroxide (N2O4, the dimeric form of nitrogen dioxide, NO2) is

used as an electron acceptor (instead of O2) in the conversion of ammonia to nitrite, with NO

being released as an additional product (Schmidt et al., 2002):

NH3 + N2O4 NO2- + 2 NO + 3 H+ + 2 e-

The nitrite produced is used partly as an electron acceptor leading to the formation of dinitrogen

gas (Schmidt et al., 2002):

NO2- + 4 H+ + 3 e- ½ N2 + H2O

Nitrogen di/tetroxide (NO2/N2O4) is not readily available in anaerobic environments and the

process is thus limited by transport of the substrate/s from oxic layers (Schmidt et al., 2002).

The term anammox is restricted to the reaction defined by the above-mentioned planctomycete-

like bacteria; anaerobic ammonium oxidation by N. eutropha may perhaps be termed anaerobic

nitrifier-denitrification. Nitrite production by this latter process may also provide a substrate for

228

Page 243: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

annamox, and the two types of bacteria may be closely related in anaerobic environments

(Schmidt et al., 2002; Jetten et al., 2003). To complicate matters further, under anaerobic

conditions some nitrite-oxidisers (strains of the genus Nitrobacter) have been reported to have a

similar nitrifier-denitrification process as that described for autotrophic ammonium oxidisers

above, producing N2O via reduction of nitrate but using pyruvate as an electron acceptor

(Wrage et al., 2001). Very little information is available regarding this pathway (Wrage et al.,

2001). As Schmidt et al. (2002) have acknowledged, we are far from understanding the

prevalence, variations and complexities of all these novel nitrogen transforming processes.

The most striking implication of the anammox process is its capacity to facilitate removal of

both ammonium and nitrite from aquatic systems as they are converted to dinitrogen gas,

thereby providing an alternative inorganic nitrogen sink to denitrification. Justifiably, the

application of anammox to the treatment of nitrogenous waste has received much attention and

the process has been patented (European Patent 0327184A1, US Patent 427849[5078884];

Mulder, 1995; Jetten et al., 1997; Trimmer, 2003). Indeed it was in a denitrifying fluidised bed

reactor treating effluent from a methanogenic reactor that anammox was first discovered

(Mulder, 1995; Jetten et al., 1997). It is also becoming obvious that anammox may play a

marked role in nitrogen turnover in natural sediments, estuaries and marine environments.

Risgaard-Petersen et al. (2003) suggest that annamox accounts for less than 6% of N2

production in estuarine sediments. In sediments of the Thames Estuary, anammox accounted for

between 1% and 8% of total N2 production (the remaining percentage attributable to

denitrification), respectively at the mouth of the estuary and near the river head (Trimmer et al.,

2003). Anammox was found to be insignificant in the sediment of a coastal eutrophic bay

(Aarhus Bay), whilst it accounted for between 24% and 67% of total N2 production in Baltic-

North Sea continental shelf sediments (Dalsgaard and Thamdrup, 2002; Thamdrup and

Dalsgaard, 2002). In reviewing studies of the annamox process in anoxic waters of the Black

Sea (Kuypers et al., 2003) and off the coast of Costa Rica (Dalsgaard et al., 2003), Devol

(2003) suggests that anammox could account for between 30% to 50% of all biological N2

production in the world’s oceans, and that a major revision of the relative importance of

denitrification in the global marine nitrogen budget may be necessary. However, research is

required before the quantitative importance of anammox and similar processes in estuarine,

marine and other environments can be accurately ascertained.

D.8.3 Heterotrophic Nitrification and Aerobic Denitrification

In addition to autotrophic nitrifying bacteria (see Section D.5 and Table D1), there is a

heterogeneous range of prokaryotes (bacteria) and eukaryotes (fungi and some unicellular algae)

that are capable of oxidising both ammonium and organic nitrogen compounds in the process

called heterotrophic nitrification (Sprent, 1987; Kuenen and Robertson, 1988; Šimek, 2000).

229

Page 244: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

For an expansive list of these species, the reader is referred to Kuenen and Robertson (1988).

The enzymes utilised in heterotrophic nitrification are argued to be quite different to

comparative autotrophic nitrifier enzymes (Ward, 2000; Wrage et al., 2001). In contrast to

autotrophic nitrification, oxidation of reduced nitrogen compounds by heterotrophic nitrifiers

requires energy, generally resulting in net energy loss (Sprent, 1987; Ward, 2000; Jetten, 2001).

In fact, the process can actually serve as an electron sink (Jetten, 2001). Many aspects of the

physiology of heterotrophic nitrification are still unclear (Šimek, 2000), although it has been

suggested that nitrogenous substrates are oxidised to inactivate toxins (including nitrite) and to

metabolise substrates which may be required for the growth of other competing organisms,

amongst other reasons (Sprent, 1987).

The existence of heterotrophic nitrifiers has long been known; however, their rates of

nitrification traditionally have been considered insignificant or very low in comparison to

autotrophic nitrifiers (Keeney, 1973; Sprent, 1987; Jetten et al., 1997; Jetten, 2001).

Heterotrophic nitrification has been studied mainly in those terrestrial systems where it has been

hard to document autotrophic nitrification (e.g. in acid forest soils) (Ward, 2000; Wrage et al.,

2001). In these environments, heterotrophic fungi have been considered more prolific in

nitrification than heterotrophic bacteria, although the net rates of heterotrophic nitrification have

usually been low (Ward, 2000; Wrage et al., 2001). Quite recently, Ward (2000; p. 427) stated

that “no information is available on the occurrence or significance of heterotrophic nitrification

in marine systems”. It has become realised, recently, that purportedly negligible or low rates of

heterotrophic nitrification may partially be due in part to an artefact of the experimental

procedure. Detection methods typically involve the determination of product (nitrite/nitrate)

appearance but heterotrophic denitrifiers do not accumulate large amounts of these compounds

(Jetten et al., 1997). This fact deems that autotrophic nitrification is by far the more important

process of in situ nitrate/nitrite supply to the external environment, but the possibility for

internal use of nitrification products by heterotrophic organisms has forced a re-think of

potential rates of heterotrophic nitrification.

The recent discovery that some heterotrophic nitrifiers are capable of simultaneous aerobic

denitrification has lead to the notion that heterotrophic nitrification rates by some species may

actually be higher than first estimated (Jetten et al., 1997; Jetten, 2001). Furthermore, the

implications of this heterotrophic nitrification / aerobic denitrification process, including the

generation of gaseous end-products, may be significant to ecosystem nitrogen budgets. The

aerobic process is in stark contrast, and offers an alternative to conventional anaerobic

denitrification and also to oxygen-limited autotrophic nitrifier-denitrification (Wrage et al.,

2001). Although the enzymes are different, the substrates (NH4+, NO2

-), intermediates

(hydroxylamine NH2OH) and products (NO2-, NO3

-) of heterotrophic and autotrophic

230

Page 245: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

nitrification are the same (Wrage et al., 2001). Definitively, the further reduction of nitrite

(NO2-) to nitrous oxide (N2O) and dinitrogen gas (N2) occurs in the same heterotrophic nitrifier

under aerobic conditions (Wrage et al., 2001). The observation of aerobic denitrification in a

heterotrophic nitrifier was first made in the bacterium Thiosphaera pantotropha and has since

been verified in representatives from other genera including Microvirgula and Alcaligenes

(Jetten et al., 1997; Jetten, 2001). The simultaneous use of oxygen and nitrate as electron

acceptors by heterotrophic nitrifiers may confer an ecological advantage by facilitating an

increased growth rate in these species (Jetten, 2001).

Currently, nitrous oxide production by heterotrophic nitrifiers is generally considered to be of

relatively minor quantitative importance in comparison to global production by autotrophic

nitrifiers (Wrage et al., 2001). However, under certain (e.g. oxic) natural conditions,

heterotrophic nitrifiers may produce more nitrous oxide (Wrage et al., 2001). As Šimek (2000)

acknowledges, the significance of both autotrophic and heterotrophic sources of NO and N2O

has not yet been sufficiently ascertained.

231

Page 246: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,
Page 247: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

APPENDIX E

DENITRIFICATION ACTIVITY IN SEDIMENT SURROUNDING

POLYCHAETE (CERATONEREIS AEQUISETIS) BURROWS1

1 A reprint of the article published in Marine and Freshwater Research (2002), 53: 35-41. Note: Ceratonereis should read Simplisetia.

233

Page 248: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

234

Page 249: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

235

Page 250: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

236

Page 251: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

237

Page 252: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

238

Page 253: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

239

Page 254: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,
Page 255: The polychaetes Australonereis ehlersi (Augener) and ......The polychaetes Australonereis ehlersi (Augener) and Simplisetia aequisetis (Augener) within the eutrophic Swan River Estuary,

Recommended