organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment...

145
Organic matter dynamics of hydrologically variable pools in intermittent streams of northwest Australia Andre Robert Siebers BSc (Hons) This thesis is presented for the degree of Doctor of Philosophy The University of Western Australia School of Plant Biology September 2015

Upload: others

Post on 27-Sep-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

Organic matter dynamics of hydrologically variable

pools in intermittent streams of northwest Australia

Andre Robert Siebers

BSc (Hons)

This thesis is presented for the degree of Doctor of Philosophy

The University of Western Australia

School of Plant Biology

September 2015

Page 2: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control
Page 3: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

i

ABSTRACT

Flow regimes of the streams and rivers of many dryland regions are highly variable and

intermittent, often characterised by intense, seasonal floods interspersed by long, dry

periods in which surface water contracts into a series of isolated pools along major

drainage channels. These pools are important refugia for aquatic biota and other

wildlife, but their biogeochemistry, and links to hydrologically driven ecological

processes more generally, remain largely unknown.

This thesis examines the organic matter (OM) and nutrient dynamics of ephemeral and

permanent pools within intermittent streams in the Hamersley Ranges of the semi-arid

Pilbara region of northwest Australia. The relative permanence of these pools through

the dry season is highly dependent on connections to alluvial water (AW). However,

little is known of how ecosystems within these pools function during inter-flood

periods, and particularly how AW connections influence ecosystem structure and

function. This research thus sought to characterise i) how the source and availability of

organic matter changes as pools evaporate and contract, ii) how microbial and metazoan

utilisation of organic matter changes as pools evaporate and contract, and iii) what

influence AW connectivity has on these relationships.

A comparison of four catchments across the Pilbara revealed that the intensity of

fluorescent fractions of dissolved organic matter (DOM) increased with the level of

evaporation. However, the composition of DOM fluorescence was usually characterised

by high contributions of humic-like fluorescence components and low contributions of

protein-like components across pools, suggesting that the production or transformation

of DOM within pools does not change as they evaporate and contract. Closer

investigation of a single catchment indicated that changes in pool volume due to

evaporation and contraction can control concentrations of organic carbon (C), nitrogen

(N) and phosphorus (P) within the water column, but that inorganic N and P

concentrations were less closely linked to changing pool volumes. The stoichiometry of

C, N and P within aquatic plants, terrestrial detritus and the water column also indicated

that AW may provide a source of nitrate and phosphate for aquatic plant growth, while

organic matter may be tightly cycled within pool ecosystems which are strongly limited

by available P.

Page 4: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

ii

Further research suggested tight internal cycling of organic matter within microbial

communities, especially after the cessation of surface flow. Pools were all net

heterotrophic in the late dry season, with primary productivity exceeded by ecosystem

respiration. Ecosystem metabolism varied amongst pools, but was not related to AW

connectivity. However, diel patterns in the stable isotope composition of dissolved

inorganic carbon (δ13C-DIC) were highly consistent across pools, suggesting that DIC

dynamics were more highly affected by temperature-solubility relationships than by

ecosystem metabolism.

Finally, investigation of metazoan food webs showed that macroinvertebrate densities

increased in direct relation to the extent of pool evaporation. Macroinvertebrate food

webs were also smaller and less diverse in the late dry season when compared with

pools post-flooding. However, there was no difference between food webs in AW-

connected and isolated pools. Macroinvertebrate assimilation of C and N also occurred

from a wide range of sources regardless of time or AW connectivity. The trophic base

of metazoan food webs in these intermittent and highly oligotrophic systems may

therefore be determined by general adaptations to intermittent flow regimes rather than

variation in response to changing environmental conditions.

Overall, the effects of AW connections are often masked by the wider ecological effects

of intermittent flow. However, identifying the extent of connectivity to groundwater is

essential in understanding the OM dynamics of intermittent rivers and streams,

particularly where sub-surface flowpaths may contribute to increased mineralization of

recalcitrant OM, chemoautotrophic production, or utilization of ancient OM sources.

Page 5: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

iii

ACKNOWLEDGEMENTS

Funding for this study was provided by an Australian Research Council Linkage Grant

to The University of Western Australia in collaboration with Rio Tinto Pty Ltd

(LP0776626). Many thanks for the collaboration and support of the hydrology team at

Rio Tinto, in particular Glenn Kirkpatrick, Niall Inverarity and Paul Hedley for their

assistance in the field. I am also grateful to Michelle Carey and Phil Davidson at API

for arranging site access and groundwater sample collection at Cane River and Red Hill

Creek.

Foremost, thank you to my supervisors, Pauline Grierson, Neil Pettit, Greg Skrzypek,

Jason Fellman and Shawan Dogramaci, for their encouragement, guidance and help

throughout this PhD. In particular, I will always be grateful to my primary supervisor,

Pauline, for giving me the opportunity to do research in some of the most amazing

places I’ve ever been.

Thank you to my friends and colleagues associated with the Ecosystems Research

Group and West Australian Biogeochemistry Centre at UWA: Doug Ford, Ela

Skrzypek, Kate Bowler, Gerald Page, Alexandra Rouillard, Rachel Argus, Sara Lock,

Romony Coyle, Grace Campbell, Jen Kelley and Ben Anderson, who helped me

considerably in the field and the lab; but mostly by way of their friendliness and

inclusiveness – for which I am also grateful to Alison O’Donnell, Tegan Davies, Kelly

Paterson, Liz McLean, Caroline Mather, Jordan Iles, Libby Trevenen, Matt Fraser,

Belinda Sykes, Judy Dunlop, Sean Tomlinson, Myles Menz, Sonja Jacob, Ché Austin,

Paul Haines, Sam Childs and Janelle Braithwaite (man, I hope that’s everyone!). A

special mention to the Bradshaw family, who have done their very best to make me feel

at home.

To say that this would have been impossible without the patience, understanding, and

support of my family is an understatement. I hope I make you proud.

Most of all, thank you to Emily; for all of the things. This is for you.

Page 6: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

iv

STATEMENT OF CANDIDATE CONTRIBUTION

This thesis is my own work, except where otherwise acknowledged, and has been

completed during the course of my enrolment for the degree of Doctor of Philosophy at

the University of Western Australia. It has not previously been accepted for any other

degree, at this or any other institution.

Credit is due for:

Chapter 1: Pilbara photographs taken by Ryan Mather and Alexandra Rouillard.

Chapters 2-5: Stable isotope and mass spectrometry samples were processed by Doug

Ford, Grzegorz Skrzypek and Ela Skrzypek at the West Australian Biogeochemistry

Centre.

This thesis contains published work: Chapter 2 was published as Siebers, A.R., Pettit,

N.E., Skrzypek, G., Fellman, J.B., Dogramaci, S. & Grierson, P.F. (2015) Alluvial

ground water influences dissolved organic matter biogeochemistry of pools within

intermittent dryland streams. Freshwater Biology doi:10.1111/fwb.12656. I developed

the original premise for the study and was the major contributor to this paper, including

data collection, analysis, and manuscript preparation. My co-authors (also my PhD

supervisors) provided input into the conceptual basis for the study and provided

methodological, technical, and editorial advice. The paper has also benefited from

editorial advice from two anonymous reviewers and the special edition editor, Catherine

Leigh. I have permission from all co-authors to include this work in my thesis.

Date: 23/09/2015

Andre Siebers

Date: 23/09/2015

Pauline Grierson (co-ordinating supervisor)

Page 7: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

v

CONTENTS

CHAPTER ONE: General introduction ....................................................................... 1

HYDROCLIMATIC VARIABILITY AND STREAMFLOW IN SEMI-ARID NORTHWEST

AUSTRALIA .................................................................................................................... 4

Intermittent and ephemeral flow regimes .................................................................. 4

Climatic variability and streamflow in the Pilbara .................................................... 6

Changing hydrology in the Pilbara ............................................................................ 9

LINKING HYDROLOGY TO ECOLOGICAL PROCESSES: ORGANIC MATTER

BIOGEOCHEMISTRY OF DRYLAND STREAMS .................................................................. 10

Organic matter biogeochemistry of freshwater ecosystems - OM sources and

fluxes ........................................................................................................................ 10

Organic matter dynamics in intermittent rivers and streams ................................... 12

Changes in organic matter production and utilisation as pools contract ................. 14

OBJECTIVES OF THIS THESIS .......................................................................................... 16

CHAPTER TWO: Alluvial groundwater influences dissolved organic matter

biogeochemistry of pools within intermittent dryland streams ................................ 18

INTRODUCTION ............................................................................................................. 18

METHODS ..................................................................................................................... 20

Study sites ................................................................................................................ 20

Sampling design ....................................................................................................... 22

Sample collection ..................................................................................................... 24

Stable isotope composition of water samples .......................................................... 24

Pool water chemistry ............................................................................................... 25

DOM fluorescence spectroscopy ............................................................................. 25

Data analyses ........................................................................................................... 27

RESULTS ....................................................................................................................... 29

Hydrologic variation within and among pools ........................................................ 29

Difference in water chemistry among catchments and sampling times ................... 31

Links between DOM fluorescence and pool hydrology .......................................... 32

DISCUSSION .................................................................................................................. 36

CHAPTER THREE: Phosphorus availability and uptake by aquatic plants drive

ecological stoichiometry in pools of an intermittent stream ..................................... 39

Page 8: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

vi

INTRODUCTION ............................................................................................................. 39

METHODS ..................................................................................................................... 41

Study sites and sampling design .............................................................................. 41

Field methods ........................................................................................................... 44

Analytical methods .................................................................................................. 44

Data analysis ............................................................................................................ 45

RESULTS ....................................................................................................................... 47

Hydrology and pool morphology............................................................................. 47

Ecological stoichiometry ......................................................................................... 48

Temporal changes in water quality and water chemistry ........................................ 51

Nutrient storage within aquatic plants and algae ..................................................... 54

DISCUSSION .................................................................................................................. 54

CHAPTER FOUR: Diel cycles of 13C-dissolved inorganic carbon (DIC) and

ecosystem metabolism in hydrologically variable pools of intermittent streams .... 59

INTRODUCTION ............................................................................................................. 59

METHODS ..................................................................................................................... 61

Study sites and sampling design .............................................................................. 61

Field methods ........................................................................................................... 63

Analytical methods .................................................................................................. 64

Data analysis ............................................................................................................ 64

RESULTS ....................................................................................................................... 66

Diel cycles of δ13CDIC .............................................................................................. 66

Diel patterns in ecosystem metabolism ................................................................... 69

Diel cycles in nutrients and DOM ........................................................................... 70

Diel patterns in AW throughflow and evaporation .................................................. 71

DISCUSSION .................................................................................................................. 71

Relationships between ecosystem metabolism and hydrology ................................ 74

Conclusions.............................................................................................................. 76

CHAPTER FIVE: Basal carbon sources maintaining macroinvertebrate food webs

in a dryland stream do not differ among pools of varying hydrology ..................... 77

INTRODUCTION ............................................................................................................. 77

METHODS ..................................................................................................................... 79

Study sites and sampling ......................................................................................... 79

Isotopic analyses ...................................................................................................... 81

Page 9: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

vii

Data analysis ............................................................................................................ 82

RESULTS ....................................................................................................................... 84

Macroinvertebrate community composition ............................................................ 84

Food web size and trophic base ............................................................................... 88

DISCUSSION .................................................................................................................. 93

Evidence for a methanotrophic food source ............................................................ 95

Conclusions .............................................................................................................. 97

CHAPTER SIX: General discussion: organic matter dynamics and ecosystem

functioning of intermittent streams in the semi-arid Pilbara ................................... 98

RESEARCH OUTCOMES: ECOLOGICAL CONSEQUENCES OF INTERMITTENT FLOW ........... 99

Importance of groundwater connectivity in contracting intermittent streams ....... 101

Biogenic methane in intermittent stream food webs ............................................. 103

The source and lability of groundwater carbon in intermittent systems ................ 106

CONCLUSIONS ............................................................................................................ 108

REFERENCES ............................................................................................................ 111

APPENDIX A: Stable isotope analysis of fine particulate organic matter and

dissolved organic matter within Coondiner Creek .................................................. 135

Page 10: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control
Page 11: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

1

CHAPTER ONE: General introduction

The aim of this thesis is to increase understanding of the organic matter dynamics

within surface pools of the intermittent and ephemeral streams of the semi-arid Pilbara

region of northwest Australia. In particular, I sought to characterise the different sources

of organic matter to dryland streams and the role they play in sustaining the productivity

and biodiversity of aquatic ecosystems. I also aimed to increase understanding of how

organic matter is cycled within and among permanent and temporary pools and how

alteration of surface and groundwater flows influence the major ecological processes of

these systems. This thesis thus emphasises the effects of the hydrological regime on the

source and transformations of the energy base of intermittent stream ecosystems.

Rivers and streams in the Australian arid sub-tropics generally only flow for short

periods of time in response to intense but variable seasonal rainfall events associated

with cyclones or monsoonal systems. However, relatively impermeable substrates or

connections to groundwater via the alluvium can result in surface water being

maintained as disconnected pools or waterholes (i.e., as permanent pools) along main

channels throughout long, dry interflood periods (Hamilton et al., 2005; Fellman et al.,

2011a). These pools may thus be considered hotspots of biological and biogeochemical

activity in the wider dry landscape (sensu McClain et al., 2003). Surface and

groundwater reserves provide refuges for aquatic organisms and frequently maintain

distinct communities of riparian vegetation. However, the size and persistence of

individual pools varies greatly both within and among catchments. Highly ephemeral

(temporary) pools that are recharged only by rainfall may evaporate within a matter of

days or weeks, while other perennial (permanent) pools with persistent shallow

groundwater throughflow can last for years between flow events.

Across the 238,592 km2 of major drainage basins in the Pilbara bioregion there are

35,376 km of stream and river channels, not including ephemeral rills and drainage

lines. Only 37 km of this network is classified as perennial (Fig. 1.1). As such, much of

the biodiversity associated with rivers and streams in the Pilbara is confined to reaches

with surface or groundwater reserves during inter-flood periods (Beesley & Prince,

2010). Many organisms are short-range endemics, restricted to particular surface or

ground waters resulting from millions of years of biogeographical separation between

Page 12: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

2

Figure 1.1: Rivers and streams classified as having perennial or intermittent flow

patterns within the Pilbara bioregion of northwest Australia. Locations of field sites

used in this study are marked in white. Major drainage basins are outlined in black.

catchments (Pinder et al., 2010). However, the Pilbara region is currently subject to

increasing demand for water abstraction due to population and industry growth

(Department of Water, 2010). In particular, the Pilbara is home to some of the world's

largest iron ore mines, which often require the re-location of large amounts of

groundwater as a consequence of mining below the water table (Dogramaci et al.,

2015). Changing rainfall patterns in the Pilbara over recent decades (O'Donnell et al.,

2015) will likely also influence the intensity and frequency of the large events required

for regional groundwater recharge (Dogramaci et al., 2012, Rouillard et al., 2015).

Understanding and management of the ecological water requirements of aquatic

ecosystems in the Pilbara clearly requires knowledge of the processes that maintain the

network of pools across the region.

Page 13: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

3

One of the most important ecological processes that may be affected by variation in

groundwater connectivity is organic matter production and utilisation. Organic matter,

generally considered to be composed of organic compounds derived from plant and

animal remains and their waste products in the environment, is the major base of both

bacterial and metazoan (multi-celled animal) food webs. Both quality (chemistry and

decomposability) and quantity of organic matter inputs to freshwater ecosystems are

strong determinants of ecosystem structure and functioning (Findlay et al., 2002; Thorp

& Delong, 2002; Tank et al., 2010). The persistence of surface water during otherwise

dry (climatic) inter-flood periods is likely to strongly influence the production and

cycling of relatively labile (more easily broken down or digested) organic matter due to

the reliance of in-stream processes such as photosynthesis or mineralisation by microbes

on the presence of water (e.g. Amalfitano et al., 2008). Groundwater connections may

also have significant effects on pool organic matter dynamics through the longitudinal

transport of organic matter between pools or via exchange processes in the hyporheic

zone. However, little is known of the major sources of organic matter inputs into the

highly dynamic, intermittent streams and rivers of warm arid environments, especially

as to how these may vary with groundwater connectivity, or the significance of different

components of organic matter to higher trophic levels.

In this introductory chapter, I first describe the distinct hydrological characteristics of

intermittent streams and the ways in which the rainfall patterns and physical

environment of the Pilbara create and maintain flow variation in these streams. I then

summarise the impacts that groundwater abstraction for regional development,

industrial water use, and predicted changes in rainfall patterns can have on flow

regimes. I also provide an overview of the current understanding of organic matter

dynamics within intermittent streams, with particular reference to the effect of the

hydrological regime on external (allochthonous) and internal (autochthonous) sources of

organic matter and their transformations within pools. This background thus provides

context for the experimental rationale that is developed in Chapters 2-5 of this thesis.

Page 14: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

4

Hydroclimatic variability and streamflow in semi-arid northwest Australia

Intermittent and ephemeral flow regimes

Dryland (arid and semi-arid) regions of the world are characterised by high rates of

evaporation that are well in excess of annual rainfall (Middleton & Thomas, 1997).

Rainfall in many dryland regions is also often highly seasonal; in the sub-tropics, for

example, the majority of rain can be delivered in intense but infrequent events such as

cyclones or monsoonal rains (e.g. Van Vreeswyk et al., 2004; Whitford, 2006).

Consequently, surface flow in dryland rivers and streams is frequently intermittent or

ephemeral (see below for further definition of these terms). In the Pilbara, in common

with much of inland Australia, large floods are followed by prolonged dry periods in

which surface water is lost or confined to isolated pools or waterholes along main

drainage channels (Fig. 1.2). However, patterns of surface flow vary widely with the

frequency of floods (e.g. Rouillard et al., 2015), geomorphology of the surrounding

landscape, and the paths that water takes to flow through, in, and out of channels

(Knighton & Nanson, 1997; Kennard et al., 2010).

Given the above, dryland rivers and streams are difficult to group under consistent

hydrological definitions. There have been attempts to group the considerable

hydrological variation exhibited by rivers and streams that regularly lose and regain

surface flow under single terms; recently it has been recommended to describe such

streams consistently as intermittent (Datry, Larned & Tockner, 2014b). However,

describing streams generally as intermittent may not fully capture the often fleeting

nature of surface waters in hot, arid regions (Fig. 1.2; Tooth, 2000). For example, many

inland and northern rivers and streams of Australia flow only infrequently (not every

year) and very briefly (typically, in the order of hours or days), usually in response to

isolated rainfall events. Such streams are more aptly described as ephemeral, while

intermittent rivers and streams have periods where surface flow is maintained for either

a longer time (from weeks to months), or at more regular intervals, by repeated or

sustained hydrological inputs before surface flow is lost. Rivers and streams that

regularly lose and regain surface flow are also sometimes referred to as episodic or

temporary (e.g. Larned et al., 2010), but these are essentially synonyms for intermittent

or ephemeral systems. Many intermittent or ephemeral rivers and streams tend to flow

in response to events that occur almost exclusively during a distinct wet season, for

example, in response to summer cyclone-associated recharge in the arid tropics or

Page 15: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

5

Figure 1.2: Contrast between flooded and dry channels in the Pilbara: Eagle Rock Falls

in a) flood and b) a month after cessation of surface flow; c) floods and d) dry channels

across the Fortescue Valley; and e) inundated and f) dry sections of a creekline.

Page 16: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

6

snowmelt in higher latitude systems (Puckridge et al., 1998). Mediterranean rivers and

streams are a specific subset of these seasonal systems where most rainfall coincides

with the cooler winter months. Identifying how these differences in climate and

hydrology interact with local environmental conditions is essential to understanding

how similar (or dissimilar) intermittent and ephemeral rivers and streams are to each

other in their ecologic functioning.

Climatic variability and streamflow in the Pilbara

The Pilbara is a topographically and geologically complex region encompassing

~500,000 km2 of northwest Australia (for context, about the size of Spain). The

waterways of the Pilbara are endogenic and thus wholly reliant on the rainfall patterns

of the region. Average annual rainfall ranges from ~300 mm in the lowlands of the

north and west to ~500 mm over the inland Hamersley Ranges (Australian Bureau of

Meteorology, www.bom.gov.au). The majority of rainfall originates from seasonal low-

pressure centres and cyclones occurring in the austral summer months from December

to March and can vary anywhere from 60 to 800 mm among years (Australian Bureau of

Meteorology, www.bom.gov.au; Fig. 1.3). Summer thunderstorms may also result in

brief, intense downpours over restricted areas, adding to the high spatial variability in

distribution of rainfall across the region and thus potential recharge to streams. Average

annual pan evaporation is between 2800 and 4000 mm across the region (Australian

Bureau of Meteorology, www.bom.gov.au) and thus many times that of rainfall. Warm

to hot temperatures further contribute to aridity, with mean minimum and maximum

temperatures of 21-39 ⁰C in summer and 6-21 ⁰C in winter (Australian Bureau of

Meteorology, www.bom.gov.au).

Given these climatic condition, most rivers and streams in the Pilbara are truly

ephemeral, flowing briefly (typically, in the order of days) in response to intense rainfall

events before long, dry inter-flood periods in which surface flow is lost. However, for

the purposes of this thesis and for consistency with the wider literature, I will generally

refer to the systems described in this body of research as intermittent.

The geomorphology of catchments and channels varies considerably across the Pilbara.

The inland ranges contain incised, narrow gorges with shallow alluvial sediments,

confining channels to narrow areas typically less than 1 km wide. These systems are

Page 17: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

7

Figure 1.3: Monthly rainfall patterns for Newman, southeastern Pilbara over the period

between 1997 and the end of 2014 (Australian Bureau of Meteorology,

www.bom.gov.au), as an example of the variability in rainfall between years in the

Pilbara.

characterised by high velocity flows, which transport and deposit sediments on broader,

lower-lying washplains where channels can spread out into extensive braided networks

or alluvial fans (Fig. 1.4). These variations in channel geomorphology influence rates

and location of groundwater recharge and connectivity and therefore also the

persistence of surface water. Stable isotope analysis of surface and ground water in the

Pilbara has shown that while the deeper alluvium of lowland washplains generally

separates surface water from groundwater tables, the shallower alluvium and exposure

of bedrock in channels travelling through gorge systems can result in relatively

permanent springs and connections of channel beds to shallow groundwaters

(Dogramaci et al., 2012). In particular, surface flow initiated by flooding can be stored

as shallow alluvial water (AW) through recharge via channel beds (Hornberger et al.,

1998). This AW moves slowly down-gradient through channel sediments and pools

Page 18: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

8

Figure 1.4: Contrast between Pilbara channel systems showing a), b) narrow, gorge

constrained systems and c), d) systems on broader washplains where multiple channels

can spread out across wider areas, including large floodplains.

perched within the alluvium, providing a counterbalance to evaporative water loss and

maintaining surface water in pools during dry inter-flood periods. However, local

differences in channel morphology and the depth of AW create significant variation in

the length of time that pools are connected to AW throughflow. Pools isolated from AW

may still retain surface water due to relatively impermeable substrates and low

transmission losses, but AW connectivity therefore largely controls variation in the

length of time that surface water is retained through drought periods (Fellman et al.,

2011a; Dogramaci et al., 2012).

Page 19: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

9

Changing hydrology in the Pilbara

Over recent decades, the mineral deposits of the Pilbara have been of extremely high

importance to the Australian economy. The production of iron ore has risen consistently

over the past decade, from 233 million tonnes or 25 % of Western Australia's total

commodity exports in 2004, to 73,732 million tonnes or 61 % of total commodities in

2014 (Department of Mines and Petroleum, 2014). Consequently, the number, size, and

intensity of mining operations in the Pilbara have also increased steadily over this

period such that the state of Western Australia is now the world’s third largest exporter

of iron ore (Department of Mines and Petroleum, 2014).

Demand for water as well as altered hydrology due to mining infrastructure (including

roads and rail lines) is an inevitable consequence of these developments. Mining of iron

ore in the Pilbara often focuses on deposits occurring below the water table and within

alluvial deposits, necessitating the abstraction of large volumes of groundwater that may

result in artificially dry reaches where dewatering has occurred (e.g. Zektser, Loáiciga

& Wolf, 2005). While some of this groundwater is utilised in mining operations, the

volumes extracted are usually far in excess of mine site requirements and thus require

disposal elsewhere in the landscape. Excess groundwater released back to the

environment may be discharged along nearby channels, creating persistent surface flow

along usually ephemeral drainage lines (see Dogramaci et al., 2015). Alternatively,

groundwater may be reinjected into nearby aquifers (e.g., Environmental Protection

Authority, 2002). While these groundwater alterations within the Pilbara have been

relatively localised to date, industry usage has also increased from around 99 ML/yr in

2008 to around 157 ML/yr in 2010 (Department of Water, 2010). Consequently,

cumulative impacts across the major catchments being mined may be considerable. In

this context, the relatively unknown resilience and responsiveness of aquatic ecosystems

to altered dynamics of recharge in northwest Australia represents a significant

management concern.

The ecological effects of alteration to flow regimes within the rivers and streams of the

Pilbara are likely to vary greatly among years due to the extremely high hydroclimatic

variability of the region. Over the last twenty years, the Pilbara has experienced a

prolonged period of recurrent cyclones and increased annual rainfall, which represents

an unprecedented wetter period compared to at least the last 200 years (O'Donnell et al.,

2015). Given that regional recharge to groundwater is primarily driven by large cyclone

events (Dogramaci et al., 2012), any change in their frequency will impact on stream

Page 20: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

10

functioning. However, there is high uncertainty in future predictions for both annual

rainfall and the frequency of cyclones (Lavender & Abbs, 2013). A return to a drier

climate and less frequent large rainfall events would amplify the effects of water

abstraction and movement across the landscape. Conversely, an increased frequency of

intense rainfall will increase groundwater tables across the landscape, alleviating the

effects of abstraction but also possibly reducing variation in the permanence of surface

water habitats within rivers and streams as ephemeral habitats become more persistent.

Either eventuality is likely to have significant if unknown effects on surface and

groundwater biodiversity and ecological processes.

Linking hydrology to ecological processes: organic matter biogeochemistry of

dryland streams

Organic matter biogeochemistry of freshwater ecosystems - OM sources and fluxes

Within aquatic ecosystems, organic matter (OM; carbon-containing non-living matter)

is present in a number of forms that are typically differentiated by size. Dissolved

organic matter (DOM) is a complex and variable collection of organic compounds

which can include carbohydrates, proteins, amino acids, lipids, phenols and a wide

range of organic acids (Sinsabaugh & Findlay, 2003). DOM generally composes the

largest fraction of overall OM within aquatic ecosystems (Schlesinger & Melack, 1981).

However, particulate organic matter (POM), defined as any OM which exists out of

solution (in practice, any OM > 0.45 μm; Tank et al., 2010), can predominate in heavily

forested catchments and during events such as seasonal peaks in riparian litterfall

(Benfield, 1997) or transport of POM in storm runoff (e.g. Jacobson et al., 2000;

Larned et al., 2010). POM can be similarly complex in both dimension and composition

to DOM, as it ranges in size from fine particles consisting of dead phytoplankton,

metazoan excrement, or the detritus resulting from mechanical breakdown of leaves to

entire dead trees or logs (Webster et al., 1999; Wetzel, 2001). Consequently, the size,

quantity and chemical composition of OM varies widely amongst freshwater

ecosystems depending on the processes that both create and utilise it.

While both DOM and POM are the basic source of carbon for heterotrophic metabolism

within freshwater environments, they are also important sources of vital nutrients such

as nitrogen and phosphorus. The lability of both DOM (Findlay, 2003) and POM (e.g.

Page 21: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

11

Callisto & Graça, 2013) is therefore highly dependent on its chemical composition, and

is a major control on the structure and function of freshwater ecosystems (Findlay et al.,

2002; Thorp & Delong, 2002). For example, organisms with fast growth rates can

dominate food webs in systems with high-phosphorus content OM that can support their

high metabolic demands (Sardans, Rivas-Ubach & Peñuelas, 2012). In turn, the

chemical composition of both DOM and POM is heavily dependent on their source.

Conventionally, many differences in the lability of OM amongst rivers and streams have

been attributable to the balance between allochthonous (produced outside the aquatic

environment) and autochthonous (produced within the aquatic environment) sources of

OM (Vannote et al., 1980; Junk, Bayley & Sparks, 1989; Thorp, Thoms & Delong,

2006; Battin et al., 2008; Creed et al., 2015). Allochthonous OM is typically

comprised of POM in the form of terrestrial plant detritus such as woody debris and

litterfall (Tank et al., 2010), and DOM largely in the form of compounds leached along

soil flowpaths into the stream and from terrestrial POM within the stream (Aitkenhead-

Peterson, McDowell & Neff, 2003). Conversely, autochthonous POM is largely

composed of excretory products or dead aquatic organisms, particularly phytoplankton

(Hein et al., 2003), and autochthonous DOM is composed of the products of lysis and

metabolism of algae, bacteria and aquatic macrophytes (Bertilsson & Jones, 2003). In

general, aquatic primary producers are smaller and have less allocation to structural

tissues than their terrestrial counterparts, and are therefore more easily broken down by

heterotrophs (Shurin, Gruner & Hillebrand, 2006). Both DOM and POM from

autochthonous sources are therefore generally considered more labile than

allochthonous sources (Thorp et al., 2006; Tank et al., 2010).

However, there is considerable variation in the composition of OM from both terrestrial

and aquatic sources. For example, terrestrial litter from C4 grasses is more labile than

that from Eucalyptus trees, which typically have high tannin and lignin content

(Macauley & Fox, 1980); but both exist in the riparian zones of northern Australian

rivers (e.g. Leigh et al., 2010). Similarly, DOM from groundwater flowpaths may

contain both highly conjugated and recalcitrant humic compounds and the labile

products of methane or sulphate oxidation (Helton et al., 2015) or extensively degraded

soil compounds which can become labile upon exposure to light (McCallister & del

Giorgio, 2012). Allochthonous sources of OM can also more easily enter food webs

through interactions with autochthonous OM, such as the increased mineralisation of

humic OM in the presence of newly produced, labile DOM (Guenet et al., 2010) or the

Page 22: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

12

breakdown and conditioning of litterfall by microbial biofilms (Bärlocher, 1985). The

mass of allochthonous OM is usually much greater within freshwater ecosystems than

autochthonous OM (Webster & Meyer, 1997), and therefore provides a much more

plentiful and predictable carbon and energy source for both microbial and metazoan

metabolism (Jansson et al., 2007). Consequently, the source of OM is important to its

utilisation, but not the only control on its ecological role in freshwater environments.

Organic matter dynamics in intermittent rivers and streams

In the intermittent streams of arid and semi-arid regions, the dynamics of OM and flow

are intimately linked. For example, the frequency of floods and high flows controls the

frequency of floodplain inputs and inundation (e.g. Valett et al., 2005), disturbance of

aquatic plants and consumers (e.g. Fisher et al., 1982; Grimm & Fisher, 1989) and the

distribution of riparian vegetation (e.g. Stromberg et al., 2007). The flow regime also

affects the breakdown of OM in streams by controlling downstream transport and thus

retention of OM and processing length, or the distance and time required for

biogeochemical transformations to take place (Fisher et al., 1998b). As flows recede,

shorter and slower flowpaths, such as internal recycling within biofilms or hyporheic

and parafluvial flow, become more important to the biogeochemistry of rivers and

streams than utilisation of OM produced upstream (Fisher, Sponseller & Heffernan,

2004; Battin et al., 2008). At the lowest flows, local autochthonous and allochthonous

inputs therefore predominate (Humphries, Keckeis & Finlayson, 2014). In this respect,

the OM dynamics of intermittent rivers are much like their perennial counterparts, albeit

with more overall flow variability. It is in the widespread and repeated occurrence of

completely dry phases that the OM dynamics of intermittent rivers may be more clearly

differentiated.

Complete drying of aquatic habitat represents a breakdown in downstream transport and

OM processing. Microbial and macroinvertebrate decomposition of POM accumulating

on dry channels and floodplains is restricted by a lack of water (Langhans & Tockner,

2006; McIntyre et al., 2009). When dry, decomposition of POM instead proceeds

mainly via photolytic processes (Austin & Vivanco, 2006). Instead of being continually

processed down a river network, therefore, dry OM represents a reserve of carbon and

nutrients that “prime” intermittent river ecosystems to respond rapidly to inundation

events (Belnap et al., 2005; Collins et al., 2008). Consequently, much of the

biogeochemical activity related to OM breakdown, transport and transformations in

Page 23: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

13

intermittent rivers may be concentrated within flood pulses, resulting in “punctuated

biogeochemical reactors” (sensu Larned et al., 2010), where the repeated steps of OM

transport and transformation created by downstream transport, utilisation in

biogeochemical processes, and subsequent release back into the stream (Fisher et al.,

1998b; Battin et al., 2008) are interrupted by dry periods in which OM is instead

reserved before rapid processing during the next inundated period (Larned et al., 2010).

However, intermittent streams do not always dry completely between flow pulses or

floods. As discussed earlier, many intermittent rivers and streams are characterised by

the persistence of pools or waterholes along channels during inter-flood periods. In

particular, pools of intermittent rivers in northern and central Australia may persist for

years between floods or flow pulses (Hamilton et al., 2005; Fellman et al., 2011a).

These pools contrast with dry reaches in that the presence of water allows microbial and

metazoan breakdown of both allochthonous and autochthonous OM to continue

unabated, or even at higher levels than during initial inundations. For example, the

highest rates of ecosystem respiration within a Mediterranean intermittent stream were

observed late in the dry season due to extensive accumulations of allochthonous OM

(Acuña et al., 2004). Pools do generally experience much greater fluctuations in

temperature and dissolved oxygen than flowing reaches, especially in warm drylands

(Wood, Fisher & Grimm, 1992; Chan et al., 2005), which could conceivably lead to

lowered rates of heterotrophic metabolism. However, development of periodic anoxia

may also increase the availability and diversity of resources available to pool

ecosystems, such as through the release of inorganic nutrients bound in sediments (von

Schiller et al., 2011) or the production of new carbon sources such as methane (Boon,

Mitchell & Lee, 1997). Rates of autochthonous primary productivity can also be

extremely high within pools, even where limited by turbidity (Bunn, Davies & Winning,

2003; Fellows et al., 2009). Pools therefore represent a substantially different OM

reserve than dry reaches.

A critical question is therefore how OM dynamics in pools might create different

responses to the flood pulse than occur in channels that dry completely. We might

expect that the extremely high quantities of both DOM and POM exported from

terrestrial environments during flood pulses (e.g. Jacobson et al., 2000; Hladyz et al.,

2011) or the high rates of primary productivity supported by inundated floodplains

(Bunn et al., 2006a) would be much in excess of whatever is imported from pools, thus

rendering them of little importance to OM dynamics in flood or flow pulses. Leaf litter

Page 24: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

14

accumulating in pools may also lose nutrients through fermentation under anoxic

conditions, resulting in a poorer substrate for microbial communities once downstream

flow resumes (Dieter et al., 2013). As previously discussed, however, even small inputs

of labile OM can stimulate much higher overall rates of OM decomposition (Guenet et

al., 2010). Photodegredation can also increase the lability of otherwise recalcitrant POM

(Fellman et al., 2013). Consequently, OM reserves in pools may be an important

priming mechanism for the initial breakdown of OM in flood pulses. However, the

extent to which the OM dynamics of pools are consistent across different environmental

gradients is also relatively unknown. Integration of pools into models of intermittent

river biogeochemical cycling will require much more information on the changes in

organic matter production and utilisation that occur within them during long inter-flood

periods.

Changes in organic matter production and utilisation as pools contract

Some of the highest rates of primary productivity (GPP) recorded for streams and rivers

across the continent are of thin bands of littoral algae within the waterholes of Cooper

Creek in arid central Australia (Bunn et al., 2006a). However, primary productivity in

intermittent streams can be heavily light limited even when riparian cover is low (Busch

& Fisher, 1981). For example, GPP within forested Mediterranean intermittent streams

is much lower and highly correlated with canopy cover, even when highly contracted

(Acuña et al., 2004). Autochthonous production of organic matter between pools in

otherwise similar catchments may therefore be extremely heterogeneous due to

differences in the morphology of pools, and thus light availability (Bunn et al., 2006a;

Fellows et al., 2009). Similarly, inputs of terrestrial plant detritus to pools is likely to be

strongly related to the degree of canopy cover, with high inputs in forested catchments

and low inputs in rivers and streams with low canopy cover (Jones et al., 1997; Bunn et

al., 2006a). Large woody debris or dams of terrestrial detritus may also be kept at low

levels due to sparse riparian tree cover and the regular flushing effects of floods (Bunn

et al., 2006b). However, extreme flood events can substantially increase the input of

woody debris to intermittent rivers, as well as creating high variability in the structure

and reestablishment of riparian vegetation (Pettit, Latterell & Naiman, 2006). The

inputs of organic matter to pools during inter-flood periods are therefore likely to be

highly variable over both space and time.

Page 25: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

15

The base of metazoan food webs in dryland rivers and streams is typically algal

production, given that a lack of flow disturbance places little limit on autotrophic

productivity for much of the year (Bunn et al., 2006b). Microbial communities also

generally makes extensive use of autochthonous OM (Findlay et al., 2002; Fellows et

al., 2007), although the lack of downstream transport during periods of little or no

surface flow means that allochthonous organic matter can accumulate for extended

periods of time and provide a substantial resource for microbial metabolism (Acuña et

al., 2004). Dryland rivers and streams are therefore usually net heterotrophic overall,

although primary productivity rates are high in the photic zone; where the surface water

of shallow streams or pools has little turbidity these zones may be net autotrophic

(Fellows et al., 2007). However, the limited food and habitat availability for metazoan

consumers in pools compared with inundated floodplains also leads to short, diffuse,

and highly interconnected food webs in pools of highly seasonal tropical rivers (e.g.

Pusey et al., 2010; Blanchette et al., 2013). Utilisation of different sources of organic

matter in these systems is also highly dynamic over time, particularly during the dry

season (Leigh et al., 2010). Consequently, the utilisation of organic matter within pool

ecosystems is likely to be as heterogeneous as its production.

A critical question arises from the observations described above; which is, how do

groundwater inputs contribute to the heterogeneity of organic matter production and

utilisation within pools of intermittent systems? Groundwater inputs may alleviate or

stop evaporative contraction and rates of aquatic habitat loss, and their associated

effects on plant and animal biomass and community composition (Stanley, Fisher &

Grimm, 1997; Lake, 2003). Groundwater inputs may also have a higher ratio of

inorganic to organic nutrients than surface waters, and stimulate autochthonous

processes in reaches with substantial groundwater connectivity (Dahm et al., 2003).

Groundwater DOM in dryland regions is generally low in concentration, but some older

fractions of groundwater DOM may be highly labile (Fellman et al., 2014). It is likely

that groundwater connectivity therefore has significant effects on organic matter

dynamics, but to date research into the effects of groundwater on pool ecosystem

processes has been restricted. Fellman et al. (2011a) suggest that groundwater

connectivity actually reduces the production and utilisation of organic matter within

pools as they dry, but they also found high levels of spatial heterogeneity among pools.

Overall, however, the effect of groundwater connectivity on ecological processes in

Page 26: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

16

intermittent streams, or even in perennial streams under drought conditions, is still

relatively unknown.

Objectives of this thesis

The general objective of this thesis was to characterise the organic matter fluxes within

pools of differing alluvial water (AW) connectivity in intermittent streams of the

Pilbara, as determined by stable isotope analysis of surface and ground water, and to

assess how these patterns changed both temporally and spatially in response to their

hydrology. As organic matter dynamics in aquatic systems are extremely complex, I

chose to focus mainly on DOM as the largest store of organic matter in most freshwater

ecosystems, although the dynamics of POM and links between POM and DOM are also

discussed. In particular, the principle objectives of this research were to investigate (i)

how the source and availability of organic matter changes as pools evaporate and

contract, (ii) how microbial and metazoan utilisation of organic matter changes as pools

evaporate and contract, and (iii) what influence AW connectivity has on these

relationships.

I examined patterns of organic matter inputs and cycling in response to the hydrological

regime across four catchments of the Pilbara region by characterising the composition

and concentration of DOM across pools with differing AW connections and evaporative

pressure. This information is fundamental to understanding how microbial production

and transformations of OM may change as pools dry (Chapter 2). I also assessed

seasonal changes in the proportion of carbon, nitrogen, and phosphorus stored within

different compartments of pool ecosystems in a creek system of the Hamersley Ranges

in the Pilbara, contrasting pool stoichiometry between relatively dry and wet periods. I

sought to assess the major factors controlling the observed patterns in carbon and

nutrients, such as evaporative concentration of solutes, changes in mineralization or

nitrification and uptake by aquatic vegetation (Chapter 3). Diurnal patterns of 13C-DIC

(dissolved inorganic carbon) and patterns of gross primary productivity (GPP) and

ecosystem respiration (ER) were compared across two creek systems with contrasting

hydrology with the aim of determining whether ecosystem metabolism was affected by

AW connections. Dissolved oxygen and the stable isotope signature of dissolved

inorganic carbon were measured to assess whether AW connections enhanced GPP

relative to ER (Chapter 4). Finally, I examined changes in macroinvertebrate

Page 27: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

17

community composition and feeding strategies across seasons and pools in order to

quantify the movement of carbon and nitrogen from both allochthonous and

autochthonous sources through metazoan food webs. This study sought to determine

whether the trophic base of secondary production changed dynamically in response to

hydrology, a situation that can occur during the flooding and drying cycles of ephemeral

waterways (Chapter 5).

Each experimental chapter of this thesis is presented for the most part as a stand-alone

manuscript, and as such, a small amount of repetition is inevitable across the body of

work presented. However, where sites, methods and concepts have been described in

previous chapters, I have referred to these sections as appropriate in order to improve

readability of the thesis as a body of work. Chapter 6 provides a general discussion of

the outcomes of this research, their relevance to the understanding of organic matter

dynamics in intermittent streams of semi-arid northwest Australia, and possible avenues

for further research, along with the major conclusions of the research.

Page 28: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

18

CHAPTER TWO: Alluvial groundwater influences dissolved organic

matter biogeochemistry of pools within intermittent dryland streams

Introduction

Dissolved organic matter (DOM) is the largest carbon reserve in aquatic ecosystems and

supports bacterial food webs (Sinsabaugh & Findlay, 2003), influences nutrient

availability (Maranger & Pullin, 2003) and limits primary production (Karlsson et al.,

2009). DOM is a variable, heterogeneous mix of numerous compounds and its

composition is the product of many complex and interacting drivers such as catchment

land-use, microbial activity and hydrological variation. Because the composition of

DOM strongly determines its lability and ecological roles (del Giorgio & Davis, 2003;

Steinberg et al., 2006), disentangling the complex processes that determine the

composition of DOM is fundamental to our understanding of how freshwater

ecosystems function.

The DOM dynamics of intermittent streams and rivers in dryland regions are generally

dominated by biogeochemical inputs and transport of organic matter during episodic

rainfall and spates, which deliver and transport large amounts of terrestrial organic

matter accrued in channels and on the floodplain during long, dry inter-flood periods

(Jacobson et al., 2000; Romaní, Vázquez & Butturini, 2006; Brooks, Haas & Huth,

2007; Catalán et al., 2013). DOM inputs are much lower during periods of

disconnection and drying, although little is known about the DOM dynamics in pools of

intermittent streams after the surface flow ceases. Concentrations of DOM and

inorganic nutrients have been strongly associated with contraction of wetted area in

drying pools, suggesting that evaporative concentration of solutes is a major driver of

water chemistry during these times (Townsend, 2002; Vazquez et al., 2011).

Evaporative contraction may also be associated with shifts in the fluorescence

characteristics of DOM through an increase in the autochthonous production of carbon

after surface flow ceases (Fellman et al., 2011a; Vazquez et al., 2011). However, these

carbon inputs can be highly variable, which at least in part may be attributable to

sustained connectivity to groundwater of some pools but not others (Fellman et al.,

2011a).

Page 29: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

19

DOM concentrations of shallow groundwater in the alluvium can be very low in dryland

regions (Valett et al., 1996; Dahm et al., 2003; Vazquez et al., 2011), which has been

attributed to dominance of recharge to streams by overland flow rather than soil and

sub-surface flowpaths that are more likely to accumulate soil-derived DOM (Marti et

al., 2000). Groundwater DOM may also be mineralized at the sediment-water interface

(Jones, 2002), further reducing the concentration of DOM that reaches surface water.

Pools and stream reaches with substantial groundwater inputs may thus experience

substantial dilution of DOM and nutrients while also receiving inputs of water that may

compensate for evaporative loss. However, how DOM composition varies within and

among intermittent streams after surface flow ceases is still mostly unknown. A recent

study, based on limited sampling of a single intermittent stream in the arid northwest of

Australia, suggested that the DOM of pools isolated from shallow alluvial groundwater

(AW) have higher proportions of some fluorescence components (e.g. protein-like) than

those that remain connected to AW (Fellman et al., 2011a). The extent to which these

observations can be generalised remains unknown; further studies that encompass

multiple catchments of differing geology, hydrology and land use are required to

determine whether AW connectivity substantially affects DOM fluorescence

characteristics of intermittent rivers.

The stable isotopic composition (δ18O and δ2H) of stream pool water reflects local

evaporation rates, volume of rainfall events and mixing of water sources and thus is a

useful measure of hydrological change (Gat, 1996). In many dryland regions, surface

water flows of intermittent rivers are derived primarily from infrequent large rainfall

events that have very distinct negative δ18O and δ2H signatures (e.g. Yuan & Miyamoto,

2008). Smaller rainfall events do not usually affect water budgets so that pools in losing

reaches progressively dry down during droughts, exhibiting predictable and linear

increases in δ18O and δ2H values based on local evaporative conditions (e.g. Dogramaci

et al., 2012). In contrast, the isotopic composition of groundwater that is rapidly

recharged through channel beds reflects the strongly 18O-depleted signatures of flood

events (e.g. Baillie et al., 2007; Dogramaci et al., 2012). Thus, disconnected and drying

pools can be differentiated from those pools with higher contributions of groundwater

flow by changes in their isotopic signatures post-flood, particularly in δ18O (Fellman et

al., 2011a). Intermittent rivers in dryland regions are thus ideal systems for investigating

the hydrologic dependency of key ecological processes, including the supply of DOM

that ultimately supports much of the trophic structure of these systems.

Page 30: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

20

In this study, I investigated the origins and composition of DOM in isolated pools of

intermittent streams from four catchments in the semi-arid Pilbara region of northwest

Australia. I used δ18O and δ2H values of surface water and AW together with DOM

fluorescence excitation-emission spectroscopy to identify changes in i) the hydrological

status (water origin and extent of evaporation) and ii) changes in DOM fluorescence of

pools across a period of recharge from seasonal floods followed by extended surface

flow cessation. My objectives were to i) determine if evaporative concentration of

solutes is the main driver of DOM fluorescence intensity as pools are isolated from AW,

and ii) if the composition of DOM is consistently affected by AW throughflow across

catchments.

Methods

Study sites

In 2011 and 2012, I sampled 26 pools along the main channels of Coondiner Creek,

Caves Creek, Cane River/ Red Hill Creek and Weeli Wolli Creek in the Hamersley

Ranges of the Pilbara region of northwest Australia (Fig. 2.1). The climate of the region

is described in Chapter 1.

The four study creeks flow briefly (typically, on the order of days) in response to

intense summer rainfall, so surface flows are almost always associated with large, high

energy and rapidly rising overbank floods. The creeks subsequently dry into a series of

disconnected pools along main channels during the extended periods between surface

flows. Surface water associated with flooding also recharges through creek beds and can

be stored as shallow alluvial groundwater (AW), which moves slowly down gradient

through alluvium. Pools are commonly perched within the alluvium and subject to

throughflow of this shallow AW, with little influence from deeper groundwater.

However, local differences in alluvium depth, water tables and channel morphology

create variation in the length of time individual pools are connected to AW, and isolate

some pools from alluvial throughflow altogether (Fellman et al., 2011a).

The four catchments are dominated by sedimentary geology, especially banded-iron

formations in association with mudstone and siltstone. Carbonaceous rocks and calcrete

platforms are relatively common at Weeli Wolli and Caves Creek, while felsic igneous

rocks predominate in the upper reaches of Coondiner Creek. Nitrogen and phosphorus

Page 31: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

21

Figure 2.1: Locations of stream pools (white circles) sampled across four catchments

within the Pilbara region of northwest Australia. Thick black lines indicate main

channels in each catchment. Dashed lines indicate the borders of insets shown for the a)

Cane River/Red Hill Creek, b) Caves Creek, c) Weeli Wolli Creek and d) Coondiner

Creek catchments. Altitude key refers to topography shown in inserts.

concentrations are near or below detection limits in soils, sediments and water (Ford et

al., 2007; McIntyre et al., 2009; Fellman et al., 2011a). All of the catchments are

subject to low levels of grazing on native vegetation, but otherwise modified land use is

minimal. Vegetation of all catchments is characterised by trees and shrubs (Eucalyptus

and Acacia spp.) scattered over Triodia hummock grassland along hills and ridges.

Riparian and floodplain vegetation is a denser assemblage of red river gum (E.

camaldulensis) and coolibah (E. victrix) woodlands, in association with Acacia and

Melaleuca spp., the latter being associated with relatively permanent alluvial

groundwater resources. Aquatic vegetation commonly consists of benthic charophytes

together with aquatic macrophytes, predominantly Potamogeton tricarinatus, with the

emergent macrophytes Typha orientalis and Schoenoplectus subulatis sometimes

Page 32: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

22

forming dense stands along banks. Filamentous algae proliferate on aquatic

macrophytes in the mid-late dry season.

Sampling design

The Pilbara region is remote (c. 1200 km from Perth, Western Australia’s largest

population centre), vast (~500,000 km2; c. the size of France) and sparsely populated,

making access to many of the region’s waterways difficult even for the few systems that

are accessible by unsealed road. As surface flow is almost exclusively associated with

intense rainfall events and large floods, some of my study sites were inaccessible for up

to three months following flooding. Consequently, much of my sampling was

necessarily opportunistic, although I have attempted to maintain consistency in the

number and hydrology of pools sampled across all catchments.

Within each catchment I aimed to select pools that i) would be sustained by AW

throughflow between rainfall and flow events to be compared with ii) pools that I

expected to become isolated from AW with increasing time since flooding. I sampled

seven pools in Coondiner Creek: four remain connected to AW except in multi-year

droughts and three quickly become isolated from AW after floods (Fellman et al.,

2011a). At Caves Creek, I sampled six pools, three that were expected to be connected

to AW because of their underlying carbonaceous geology (shallower alluvium) and

three associated with more siliciclastic geology (deeper alluvium), which were expected

to become isolated pools. At Cane River/Red Hill Creek, I sampled eight pools: four

deeper and larger pools suspected to retain connections to AW throughflow owing to

their observed persistence in dry years, and four shallower and smaller pools that I

expected would be isolated from AW.

Weeli Wolli Creek receives discharge of groundwater from nearby iron ore mining

operations and thus experiences year-round surface flow for c. 24 km of its length

(Dogramaci et al., 2015). The discharged groundwater is of high quality and generally

low in DOM and nutrients. This stream thus provides us with an opportunity to assess

how large volumes of surface water flow influence pool δ18O and δ2H values and DOM

fluorescence. At Weeli Wolli Creek, I sampled two isolated pools upstream from the

discharge point, which long-term observations (> 10 years) suggest are sustained by

AW throughflow throughout much of the dry season. I also sampled three pools that are

now connected by continual surface water flow but with unknown AW connectivity at

0, 5 and 10 km downstream of the discharge point.

Page 33: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

23

Figure 2.2: Monthly rainfall across the study catchments. Data for Weeli Wolli Creek,

Caves Creek and Cane River/Red Hill Creek were obtained from the Australian Bureau

of Meteorology (www.bom.gov.au) for sites closest to these catchments, whereas data

for Coondiner Creek were obtained from a HOBO weather station (Onset Computer

Corporation, USA) installed at the catchment; • represents months when samples were

taken.

Coondiner Creek pools were sampled in both May and September 2011, c. 3 and 7

months respectively after the cessation of surface flows arising from large rainfall

events in February 2011 (Fig. 2.2). Pools were again sampled following floods arising

from Cyclone Heidi (10-13 Jan 2012) in April, June and October 2012, corresponding

to 2, 4 and 8 months after cessation of surface flow (Fig. 2.2). Caves Creek pools were

sampled in February 2012, less than a month after cessation of surface flow initiated by

floods, while Cane River/Red Hill Creek pools were sampled in March 2012, also c. 1

month following cessation of surface flow initiated by floods (Fig. 2.2). Weeli Wolli

Creek pools were sampled only once in July 2012, c.5 months following flooding from

Cyclone Heidi and the cessation of surface flow above the discharge point (Fig. 2.2).

Page 34: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

24

Sample collection

Water samples (40 mL) for δ18O and δ2H analysis were collected from each pool 5 cm

below the water surface, filtered immediately through 0.2μm sterile PES syringe filters

and stored without headspace in airtight, sterile glass vials at 4oC in a dark refrigerator

until analysis. Three to six replicate water samples (100 mL) for DOM and water

chemistry (dissolved organic carbon and nitrogen, dissolved inorganic nitrogen and

phosphorus) analysis were collected at most pools at the same locations as δ18O and δ2H

samples, filtered through 0.7μm glass fibre or quartz filters in the field and stored in

acid-washed PET bottles at 4oC in a dark refrigerator until analysis. For the Cane River

sampling in March 2012 and the Coondiner Creek sampling in April 2012, equipment

failure prevented collection of separate samples for DOM and water chemistry. On this

occasion, water samples collected for δ18O and δ2H analysis were also used for DOM

analysis.

Regional groundwater has been extensively characterised for stable isotope composition

(Fellman et al., 2011a; Dogramaci et al., 2012; Skrzypek, Dogramaci & Grierson,

2013; Dogramaci et al., 2015), with a regional δ18O average of between -8 and -9 ‰.

Given the consistency of this regional data set, I considered it sufficient to collect a few

AW samples opportunistically to establish that my study systems conformed to this

range of values. At Coondiner Creek, AW samples were collected in May 2011 and

April 2012 from a c. 30 m deep alluvium bore located c. 5 km downstream from Pool 1.

At Caves Creek, I sampled three c. 10-15 m deep bores between 500 to 1300 m adjacent

to the channel. At Cane River/Red Hill Creek, five c. 15 m deep bores were sampled

between 1 and 2 km adjacent to the channel. At Weeli Wolli Creek, groundwater

samples were collected directly from the discharge point. Discharge water originated c.7

km upstream and 3 to 5 km adjacent to the channel at a depth of c. 20 m. In addition, a

water sample for δ18O and δ2H analysis of rainfall was collected during Cyclone Heidi

in February 2012 by a bulk rainfall collector located c. 28 km northeast of Weeli Wolli

Creek.

Stable isotope composition of water samples

The stable isotope composition (δ18O and δ2H) of all water samples was analyzed using

a Picarro L1102-i Isotopic Liquid Water Analyser (Picarro, Santa Clara, CA, USA).

Page 35: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

25

δ18O and δ2H values are reported in per mil (‰) after normalization to VSMOW scale

(Vienna Standard Mean Ocean Water), following a three-point normalization based on

three laboratory standards each replicated twice (Coplen, 1994). All laboratory

standards were calibrated against international reference materials for the

VSMOW/SLAP (Standard Light Antarctic Precipitation, relative to VSMOW) scale

provided by the International Atomic Energy Agency (Coplen, 1995). Analytical

precision was 1.0 ‰ for δ2H and 0.1 ‰ for δ18O.

Pool water chemistry

Dissolved organic carbon (DOC; determined by non-purgeable organic carbon analysis)

and total dissolved nitrogen (TDN) were measured by high-temperature catalytic

oxidation on a Shimadzu TOC-V total carbon and nitrogen analyzer (Shimadzu, Kyoto,

Japan). Dissolved inorganic nitrogen (DIN; NH4-N and NO3-N) was determined by

spectrophotometric colorimetric detection method on a Technicon AutoAnalyzer

(Technicon, Tarrytown, NY, USA) with a lower detection limit of 8 µg N L -1.

Dissolved organic nitrogen (DON) was calculated as the difference between TDN and

DIN. Dissolved inorganic phosphorus (as PO4, or soluble reactive phosphorus; SRP)

was measured using the ascorbic acid method (Murphy & Riley, 1962) with a lower

detection limit of 2.5 µg P L-1.

DOM fluorescence spectroscopy

DOM of all pool water samples was characterized by excitation-emission fluorescence

spectroscopy on a Varian Cary Eclipse fluorescence spectrometer (Varian Inc.,

Mulgrave, VIC, Australia). Pool water samples were allowed to warm to room

temperature and then filtered through 0.2 μm PES syringe filters in the laboratory

immediately prior to analysis. Fluorescence intensity of these water samples was

measured at excitation (Ex) wavelengths from 240- 500 nm (5 nm increments) and

emission (Em) wavelengths from 300- 600 nm (2 nm increments). Ex and Em slit

widths were 5 nm and photomultiplier tube voltage was set to 725 V. Samples were

corrected for inner filter effects, by dilution with Milli-Q water, when optical density

was > 0.05 at 254 nm within a 1 cm quartz cuvette (Green & Blough, 1994). All

fluorescence excitation-emission matrices (EEMs) were corrected for instrument bias

Page 36: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

26

using correction files provided by Varian Cary, and normalized to Raman Units (R.U.)

by the area under the Raman water peak at Ex350nm.

Replicate EEMs were averaged, resulting in 53 EEMs in the final data set that

represented a single pool at a single sampling time. These EEMs were analysed with

parallel factor analysis (PARAFAC), a multivariate decomposition of the DOM

fluorescence signal to individual components (Stedmon, Markager & Bro, 2003), using

the DOMFLuor Toolbox v1.7 (Stedmon & Bro, 2008) in MATLAB v7.14 (MathWorks,

Natick, MA, USA). To reduce the influence of highly concentrated samples, each EEM

was standardized pre-PARAFAC by dividing the entire EEM by total sample

fluorescence. A triangle of zeroes was inserted into each EEM in the area of no

fluorescence to aid in component discrimination (Thygesen et al., 2004).

Four fluorescence components were identified by the PARAFAC model (Fig. 2.3) and

validated by split half analysis. Components 1 and 4 (C1, C4) were humic-like, with

emission maxima at wavelengths greater than 420 nm. C1 (peaks at Ex 240/330 nm, Em

338 nm) closely resembled a combination of the ubiquitous humic and fulvic-like peaks

A and C of Coble’s (1996) characterization of aquatic environments, which represent

aromatic, conjugated and high molecular weight compounds (Hudson, Baker &

Reynolds, 2007; Fellman, Hood & Spencer, 2011b). C4 (peaks at Ex 270/395/445 nm,

Em 512/572 nm) was also similar to common humic-like fluorescence components,

with its greater emission wavelengths possibly representing more highly conjugated

compounds than C1 (Fellman et al., 2011b). C2 (peaks at Ex 300/<240 nm, Em 340 nm)

and C3 (peaks at Ex 273/<240 nm, Em 302 nm) had spectra closely resembling those of

tryptophan and tyrosine respectively, and often correlate with the presence of amino

acids, intact proteins and peptide material (Hudson et al., 2007; Fellman et al., 2011b).

However, a large proportion of protein-like fluorescence may also arise from non-

proteinaceous compounds, such as N-containing aromatics or carboxylic-rich alicyclic

molecules (Stubbins et al., 2014; Kellerman et al., 2015). PARAFAC results are given

as maximum fluorescence intensity (Fmax) of each component, from each averaged

EEM, at the same scale as the original EEMs (R.U.).

Page 37: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

27

Figure 2.3: The four PARAFAC components identified in this study: a) humic-like C1,

b) tryptophan-like C2, c) tyrosine-like C3 and d) humic-like C4. White in the bottom-

right corners indicates areas of no fluorescence that were excised from the EEM data.

Data analyses

The δ18O and δ2H values of pool water, AW and cyclonic rainfall were plotted against

each other and compared with the local meteoric water line (LMWL) and local

evaporation line (LEL) of the Hamersley Basin (Dogramaci et al., 2012) to determine

the original source and evaporation of pool water. The LMWL is derived from the

relationship between δ18O and δ2H in local precipitation events, with a linear increase in

both occurring as the size of rainfall events decreases. LELs are calculated from the

linear relationship between pool water δ18O and δ2H values as evaporation occurs and

generally have a slope significantly less than that of the LMWL. The position of pool

water values that fall upon the LMWL can therefore be inferred as being mostly derived

from rainfall events of the corresponding size. Conversely, pool water values that

deviate from the LMWL can be traced back via an LEL to the point along the LMWL at

Page 38: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

28

which they began evaporating, and thus their original water source. The same rationale

can be applied to AW δ18O and δ2H values.

In addition, differences in the slopes of LELs between different systems can indicate

local differences in humidity which affect the degree to which oxygen and hydrogen

isotopes are discriminated against during evaporation, and therefore the applicability of

using single isotope values (i.e., δ18O) as a universal indicator of AW connectivity and

evaporation across these systems. LELs were therefore calculated for each individual

sampling occasion as well as for the entire pool water data set by least-squares linear

regression of pool water δ18O and δ2H values. Differences in the slopes of LELs were

examined using Student’s t-tests at a confidence level of 95%.

I assessed changes in water chemistry along δ18O gradients by applying linear least-

squares regression at a confidence level of 95% to the relationships between DOC,

DON, DIN and PO4 concentrations and δ18O values in pool water. I also assessed how

much DOM fluorescence was related to bulk DOC measurements by applying linear

least-squares regression at a confidence level of 95% to the relationships between the

Fmax values of each fluorescence component and DOC concentrations.

To assess whether the gradient between AW throughflow and evaporation affected

DOM fluorescence intensity, I applied linear least-squares regression at a confidence

level of 95% to the relationships between the Fmax values of each fluorescence

component and δ18O values in pool water. I also calculated the proportional contribution

of each fluorescence component within each pool water sample by dividing Fmax values

by the sample’s total EEM fluorescence. The residuals for linear regressions between

these proportional Fmax values and δ18O were not normally distributed. I therefore

assessed whether the composition of DOM fluorescence also changed along the gradient

from AW-connected to highly evaporated pools by applying linear least-squares

regression at a confidence level of 95% to the relationships between the proportional

Fmax values of each fluorescence component and the log10 of δ18O values (specifically

δ18O +10, to avoid any logarithmic transformations of values below 1) in pool water.

All r2 values given are adjusted for sample size.

Page 39: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

29

Results

Hydrologic variation within and among pools

All δ18O and δ2H values for AW samples fell within the known range of the Hamersley

Basin (-10 to -6 ‰ for δ18O; -70 to -40 ‰ for δ2H; Dogramaci et al., 2012) and varied

little across catchments (Fig. 2.4). Based on their isotopic compositions, AW samples

across all four creeks were derived from floods caused by large rainfall events, such as

those associated with Cyclone Heidi (Fig. 2.4; Dogramaci et al., 2012). As the LELs for

most catchments also intersected the LMWL around values associated with large events,

the original source of all pool water is likely to have been either cyclonic rainfall or AW

derived from cyclonic rainfall. Pool water δ18O values of less than -6 ‰ were therefore

characterised as likely to have had AW throughflow at the time of sampling.

Pool surface water δ18O values across all creeks and sample times ranged from -8.21

(similar to rainfall inputs from large-volume events) to 13.1 ‰ (highly evaporated).

Pool surface water δ18O values were not correlated with time since the most recent flood

(P = 0.99). At Caves Creek, three of the six pools sampled (two pools from the reaches

with carbonaceous geology and one from a reach with siliciclastic geology) had

substantially more positive δ18O values (-3.7 to -0.28 ‰) compared to AW (average -

8.1 ‰), reflecting greater evaporative loss as a result from disconnection from alluvial

throughflow (Fig. 2.4). At Cane River/Red Hill Creek, five of the eight pools sampled

(including the four suspected to be isolated from AW) had higher δ18O values (-3.4 to

10.2 ‰) than AW (average -8.03 ‰) and thus probably were also isolated from

throughflow (Fig. 2.4). At Weeli Wolli Creek, δ18O values were lowest immediately

below the groundwater discharge point (-8 ‰) and increased (to a maximum of -7.87

‰) with distance downstream, reflecting slight progressive evaporation of water. At

Weeli Wolli Creek, pools upstream from the discharge point had less depleted δ18O

values (>-6.5‰), which was unexpected as these pools were the deepest across all

catchments as well as heavily shaded. One pool had a consistently evaporated signature

(-3.42 ‰) indicating it was likely isolated from AW (Fig. 2.4). At Coondiner Creek,

four of the seven pools (those originally classified as being AW-connected) consistently

had δ18O values (-7.6 to -5.8 ‰) similar to AW and thus likely remained connected to

AW throughflow across the sampling period (Fig. 2.4). The other three pools had values

similar to or slightly higher than AW (-5.6 to -4.1 ‰) for samples collected within

Page 40: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

30

Figure 2.4: Relationship between δ18O and δ2H values for alluvial groundwater (AW),

rainfall and pool water samples collected between May 2011 and October 2012 from a)

Caves Creek, Cane River/Red Hill Creek and Weeli Wolli Creek, and b) Coondiner

Creek. The local meteoric water line (LMWL) and local evaporation line (LEL) for the

Hamersley Basin are taken from Dogramaci et al. (2012). LELs significantly different

from the entire data set are shown for a) Caves Creek and b) Coondiner Creek.

Page 41: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

31

relatively short time periods since surface flow ceased (May 2011, April and June

2012), when pools were likely still connected or only just beginning to be isolated from

AW throughflow. Samples from these same pools had substantially more positive δ18O

and δ2H values (-1.9 to 13.1 ‰) in September 2011 and October 2012 when pools were

likely completely isolated from AW.

The LEL calculated from all pool water samples (δ2H = 4.60 × δ18O – 15.2; r2 = 0.95, P

< 0.001) was not significantly different from the LELs for the Cane River/Red Hill

Creek (δ2H = 4.70 × δ18O – 14.7; r2 = 0.98, P < 0.001) and Weeli Wolli Creek (δ2H =

5.31 × δ18O – 15.5; r2 = 0.94, P = 0.005) catchments (P = 0.22 to 0.39). The Caves

Creek LEL (δ2H = 6.78 × δ18O – 4.3; r2 = 0.98, P < 0.001) had a significantly higher

slope than the combined LEL (P < 0.001). However, this may be due to mixing of

recent rainfall in the most highly evaporated pool (Fig. 2.4).

For Coondiner Creek, the slopes of local evaporation lines were not statistically

different from each other except for samples collected in April 2012 (P = 0.30 to 0.96).

However, the May 2011 LEL was not statistically different from that in April 2012 (P =

0.64). An LEL for Coondiner Creek was therefore calculated for pools across all

sampling occasions (δ2H = 4.06 × δ18O – 17.7; r2 = 0.97, P < 0.001), which had a slope

significantly lower than that for the combined LEL (P = 0.04), possibly due to lower

humidity during October 2012 than other sampling times.

Difference in water chemistry among catchments and sampling times

The highest DOC values were at Caves Creek (5.5 to 13.4 mg L-1) and during October

2012 in Coondiner Creek (0.9 to 14.0 mg L-1); however, DOC concentrations were

highly variable among pools (Table 3.1). DOC concentrations were lowest at Weeli

Wolli Creek (0.4 to 1.4 mg L-1) and Coondiner Creek in June 2012 (0.4 to 1.4 mg L-1).

DOC concentrations were generally higher in pools with more enriched δ18O values (r2

= 0.41, P < 0.001) and were also strongly correlated with DON (r2 = 0.82, P < 0.001).

All sampled pools had low concentrations of DIN (< 0.2 mg L-1) and SRP (< 2.5 μg L-1)

except for Coondiner Creek in June 2012 (mean 7.5 μg L-1 SRP) and Weeli Wolli Creek

(mean 16.4 μg L-1 SRP), regardless of sampling time (Table 3.1). DIN and SRP were

not correlated with δ18O.

Page 42: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

32

Table 3.1: Average water chemistry values for each sampling occasion, with standard

errors in parentheses. Sampling occasions are separated by year and listed in

chronological order. The Cane River/Roy Hill Creek sampling in March 2012 and

Coondiner Creek sampling in April 2012 were not assessed for water chemistry. *

indicates average for all samples was below detection limits (2.5 μg L-1 SRP).

Catchment Date DOC

(mg L-1)

DON

(mg L-1)

DIN

(NO3 +NH4)

(mg L-1)

SRP (PO4)

(µg L-1)

Coondiner Creek May 2011 3.27 (±0.54) 0.15 (±0.01) 0.11 (±0.03) <2.5*

Coondiner Creek Sept 2011 5.90 (±1.01) 0.17 (±0.03) 0.01 (±0.0001) <2.5*

Caves Creek Feb 2012 7.67 (±1.53) 0.51 (±0.10) 0.19 (±0.1) <2.5*

Coondiner Creek June 2012 1.05 (±0.17) 0.09 (±0.01) 0.02 (±0.006) 7.5 (±1.2)

Weeli Wolli Creek July 2012 0.58 (±0.18) 0.08 (±0.01) 0.04 (±0.02) 16.4 (±6.6)

Coondiner Creek Oct 2012 4.43 (±1.78) 0.31 (±0.05) 0.07(±0.01) <2.5*

Fmax values were highly positively correlated with DOC concentrations for C1 (r2 = 0.8,

P < 0.001) and C4 (r2 = 0.73, P < 0.001). C2 Fmax values were also higher in pools with

high DOC concentrations (r2 = 0.61, P < 0.001). While C3 Fmax values were higher in

pools with high DOC concentrations, there was much more variability in this

relationship than with the other fluorescence components (r2 = 0.15, P = 0.01).

Links between DOM fluorescence and pool hydrology

Across all catchments, the fluorescence intensities of DOM components C1, C2 and C4

were generally greater in pools with enriched δ18O values indicating greater evaporative

losses and lowest in pools with δ18O values indicating throughflow of AW (Fig. 2.5a, b,

d). In contrast, C3 fluorescence was still significantly and positively correlated with

δ18O values, but showed considerable variation which was not attributable to the δ18O

gradient (Fig. 2.5c).

Page 43: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

33

Figure 2.5: Relationships between δ18O and maximum fluorescence (Fmax) values of

PARAFAC components a) C1, b) C2, c) C3 and d) C4 across all pool water samples.

Dotted lines show 95% confidence bounds for the regressions. Note different scales for

Fmax across different components.

The percent contribution of DOM fluorescence components varied little across δ18O

gradients (Fig. 2.6). In particular, the contribution of C4 to total fluorescence was very

consistent (c. 0.015 to 0.027 %) across both evaporating pools and those connected to

AW (Fig. 2.6d). The contribution of C2 to total fluorescence was also highly consistent

among the 26 pools, with the exception of four pools from Coondiner Creek with

relatively low δ18O values and the pool below the Weeli Wolli Creek discharge, which

had noticeably higher values (>0.025 %; Fig. 2.6b). While the contribution of C3 to

total fluorescence was not significantly correlated with δ18O values, C3 variation within

pools connected to AW throughflow was greater than the proportional contributions of

Page 44: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

34

Figure 2.6: Relationships between δ18O and the proportional contribution of maximum

fluorescence (Fmax) values to total EEM fluorescence for PARAFAC components a) C1,

b) C2, c) C3 and d) C4 across all pool water samples. The pool directly below the Weeli

Wolli groundwater discharge is filled in black. Dotted lines show 95% confidence

bounds for the regressions. Note different scales for proportional contribution of Fmax

across different components.

C2 and C4 (Fig. 2.6c). Conversely, the contribution of C1 to total fluorescence was

positively correlated with δ18O values and was lowest in pools with δ18O values

indicating AW throughflow, especially in the pool below the Weeli Wolli discharge

(Fig. 2.6a).

Pools with low contributions from C1 were mainly those connected to AW throughflow

in Weeli Wolli Creek and Cane River and in Coondiner Creek during September 2011

(Fig. 2.7a). Pools with the highest contributions from C3 were also largely AW-

Page 45: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

35

Figure 2.7: Median, upper and lower quartiles and maximum and minimum values for

the proportional contribution of maximum fluorescence (Fmax) values to total EEM

fluorescence for PARAFAC components a) C1, b) C2, c) C3 and d) C4 across the

different sampling dates and catchments, separated by alluvial groundwater (AW)

connectivity. Catchment names are abbreviated to Cnd (Coondiner Creek), Caves

(Caves Creek), Cane (Cane River and Red Hill Creek) and W.W. (Weeli Wolli Creek).

Samples are presented in chronological order. Note different scales for proportional

contribution of Fmax across different components. * represents the single AW-isolated

pool sampled in Weeli Wolli Creek.

Page 46: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

36

connected, and were all from either Weeli Wolli Creek or the June 2012 sampling of

Coondiner Creek. However, pools isolated from AW in these samplings also had higher

contributions from C3 than either AW-connected or isolated pools at other times (Fig.

2.7c). High contributions from C2 were much less specific to catchment, sampling time

or hydrology (Fig. 2.7b), while those from C4 were extremely consistent across all

catchments (Fig. 2.7d).

Discussion

My study confirms that the amount and composition of DOM in pools within

intermittent streams of the Pilbara are strongly determined by the extent to which pools

remain connected to alluvial groundwater. In general, more evaporated pools had

greater DOM fluorescence and DOC concentrations than pools receiving continuous

AW inputs. However, my study also revealed that the intensity of tyrosine-like

fluorescence (C3) was highly variable across both AW-connected and isolated pools.

While some AW-connected pools had high percentage contributions from tyrosine-like

fluorescence and low contributions from humic-like fluorescence (C1 and C4), most

had negligible contributions from tyrosine-like fluorescence. Evaporated pools had

consistently low percentage contributions of tyrosine-like fluorescence. These findings

contrast directly with earlier observations at Coondiner Creek (Fellman et al., 2011a).

Tyrosine-like fluorescence is generally attributed to biodegradation of DOM in aquatic

systems as a result of greater bacterial or algal metabolism (Elliott, Lead & Baker, 2006;

Fellman et al., 2011b; Cory & Kaplan, 2012). My results therefore suggest that little

modification of DOM occurred as pools evaporated. The variable contribution of

tyrosine-like fluorescence in pools maintained by AW is more difficult to explain. Small

(< 10 kDa) and extensively degraded molecules that accumulate in groundwaters

(Romaní et al., 2006) can contribute to tyrosine-like fluorescence (Yamashita &

Tanoue, 2004). Thus, higher tyrosine-like fluorescence in my study may reflect highly

degraded DOM inputs from alluvial water. The high proportions of tyrosine-like

fluorescence in the discharge-dominated pools at Weeli Wolli Creek also suggest this

component may strongly reflect degradation status, particularly if the molecules

associated with protein-like fluorescence are actually nitrogen-depleted (Stubbins et al.,

2014). However, further characterisation of DOM in groundwater is required to confirm

this hypothesis.

Page 47: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

37

The proportion of tryptophan-like fluorescence (C2) varied considerably over time and

across catchments. Overall, proportional contributions of tryptophan-like fluorescence

did not consistently differ between AW-connected and evaporated pools. Fmax values of

tryptophan-like fluorescence (i.e. indicating greater concentrations of amino acids and

proteins; Hudson et al., 2007; Fellman et al., 2011b) were often three to four-fold

greater than observed previously at Coondiner Creek (Fellman et al., 2011a).

Tryptophan-like Fmax values were also positively correlated with DOC and δ18O values,

which suggests that evaporation increases the concentration of tryptophan-like

components but does not increase its production. This result is unusual, as tryptophan-

like fluorescence is considered to reflect the production of DOM from microbial or algal

metabolism (Cammack et al., 2004; Elliott et al., 2006; Jaffé et al., 2008), which I

would expect to increase after surface flow ceases. Light, nutrient supply and grazing

may become limiting at different rates across pools, thus increasing variability.

Conversely, tryptophan-like fluorescence at my sites is also likely to arise from

polyphenols and tannins (Maie et al., 2007) leached from riparian litterfall into pools

after surface flow ceases, as well as downstream transport (Reid et al., 2008).

Tryptophan-like fluorescence may also be produced from the photochemical

degradation of humic-like DOM (Howitt et al., 2008). There is growing evidence that

large proportions of the compounds contributing to tryptophan-like fluorescence in

some freshwater environments may be N-free and similar to degradation products of

lignin (Stubbins et al., 2014; Kellerman et al., 2015). Given that tryptophan-like

fluorescence displays a similar relationship with δ18O values here as humic-like

fluorescence does, it is likely that tryptophan-like fluorescence in my sites is also not

strongly associated with proteins or amino acids.

The intensity of humic-like DOM fluorescence components was higher in more

evaporated pools, consistent with the evaporative concentration of solutes previously

observed in other intermittent rivers (Townsend, 2002; Vazquez et al., 2011), including

Coondiner Creek (Fellman et al., 2011a). Conversely, pools with AW throughflow had

low fluorescence of humic-like components. These results demonstrate that at my sites

in the Pilbara, much as in other low productivity dryland regions, AW has low

contributions from soil flowpaths and the humic-rich organic matter contained in them

(Dahm et al., 2003; Vazquez et al., 2011). The low intensity of humic-like fluorescence

in the discharge-affected pools at Weeli Wolli Creek suggest this is mostly due to a

dilution effect of AW with low DOM, but humic-like fluorescence was also low in sites

Page 48: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

38

upstream of the discharge. AW throughflow may therefore also enhance the microbial

decomposition of humic-like DOM, perhaps through enhanced movement of water

column DOM into the hyporheic zone (Jones, 1995).

Overall, my results suggest that while disconnection from groundwater and subsequent

evaporation in stream pools are important in determining the characteristics of DOM,

consistent and direct effects of AW on DOM composition are less clear. Environmental

isotopes coupled with further characterisation of DOM can be used to better understand

changes in groundwater-surface water mixing and have recently proven effective for

identifying links between non-contemporary carbon fixation and bacterial metabolism

(Fellman et al., 2014). For example, the rapid biotic uptake of old, bioavailable DOC

originating in groundwater springs may be gradually replaced downstream by the

accumulation of modern, terrestrially derived DOC. Increased knowledge of the age,

source and composition of DOM will further elucidate the role that groundwater

connectivity plays in maintaining and possibly subsidising the ecological processes of

intermittent streams, particularly in oligotrophic systems.

Page 49: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

39

CHAPTER THREE: Phosphorus availability and uptake by aquatic

plants drive ecological stoichiometry in pools of an intermittent stream

Introduction

Ecosystem trophic structure, nutrient cycling and productivity are highly dependent on

the elemental composition of organisms and their environment, or ecological

stoichiometry (Sterner et al., 1998; Elser & Urabe, 1999; Vanni, 2002). Metabolic

processes place intrinsic constraints on the elemental composition of all living

organisms, with higher metabolic rates requiring higher acquisition of N and P due to

the need to produce more ATP and RNA (Ågren, 2004; Cross et al., 2005). Higher

metabolic rates are promoted by lower C:N:P ratios in resources such as particulate

organic matter (POM) for decomposers (Enríquez, Duarte & Sand-Jensen, 1993) or

dissolved inorganic nutrients for plants and algae (Koerselman & Meuleman, 1996).

The C:N:P ratios of dissolved compounds within the water column can therefore be a

strong indicator of certain biogeochemical processes. For example, the biological

oxidation of NH4 to NO2 and then NO3 (i.e., nitrification) by chemoautotrophic bacteria

decreases as C:N ratios increase, reflecting reduced production of NH4 from the

mineralization of organic matter by heterotrophic bacteria (Jones, Fisher & Grimm,

1995a). Similarly, low N:P ratios can indicate that the energetically expensive process

of nitrogen fixation can be promoted in N-limited ecosystems (Howarth, Marino &

Cole, 1988). These shifts in ecological stoichiometry remain virtually undescribed for

most intermittent streams.

In Chapter 2, I found that the evaporative contraction of pools in intermittent rivers is

likely to cause a strong concentration of solutes within the water column (Townsend,

2002; Vazquez et al., 2011). However, whether this contraction affects the ecological

stoichiometry of pools is unclear. Previously, von Schiller et al. (2011) observed

relatively consistent N:P ratios in the water column of contracting and fragmenting

pools within a Mediterranean forest stream, despite high spatial heterogeneity in N and

P concentrations. This apparent contradiction, where ratios of C, N and P to each other

remain consistent during periods when the spatial heterogeneity of organic matter,

nutrients and water quality are at their highest (Sheldon & Fellows, 2010; von Schiller

et al., 2011; von Schiller et al., 2015), suggests that evaporative contraction

concentrates organic matter and nutrients in the water column but does not alter the

Page 50: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

40

basic processes underlying their source or production (Chapter 2).

Evaporative contraction (i.e., reduced volume) of pools within intermittent streams may

be alleviated or delayed by inputs of groundwater via the alluvium (e.g. Fellman et al.,

2011a; Dogramaci et al., 2012; Chapter 2). Groundwater connectivity may also

influence the ecological stoichiometry of pools via mechanisms other than alleviation of

evaporative concentration. For example, a lack of sub-surface flowpaths, which

accumulate soil-derived organic matter (Marti et al., 2000), leads to generally low

concentrations of dissolved organic matter (DOM) and inorganic nutrients in the

groundwater of dryland regions (Dahm et al., 2003). In contrast, groundwater

upwellings can provide inputs of organic matter and nutrients that pools without

groundwater connectivity lack and which may be crucial resources for autotrophic

production within drying intermittent streams (Valett et al., 1994). High ratios of

inorganic nutrients such as NO3 and PO43-, to DOM in groundwater inputs may also

favour autotrophic over heterotrophic processes (Dahm et al., 2003), for example the

production of algal biomass and exudates over coupled mineralisation of imported DON

and subsequent nitrification (Jones, 2002). Increased autotrophic production may in turn

lower C:N ratios of the trophic base of food webs, favouring metazoan consumers with

high growth rates and metabolic demands (Cross et al., 2005) or encouraging bacterial

over fungal growth (Findlay et al., 2002). Alternatively, groundwater inputs of DOM

may be low but also highly labile (Fellman et al., 2014) and therefore rapidly

mineralized at the sediment-water interface, resulting in the release of inorganic

nutrients to the water column (Jones, 2002). While we might therefore expect that pools

with differing groundwater connectivity will show substantially different patterns of

ecological stoichiometry in relation to expansion and contraction of isolated pools, the

nature of these differences remains unclear.

The quantity and nature of organic matter and nutrients in the water column can also be

strongly influenced by the presence of aquatic vegetation, especially in low nutrient

systems (Scheffer & Nes, 2007). Aquatic macrophytes, such as charophytes and

Potamogeton spp., can act as a sink for inorganic nutrients both through their own

demand for N and P and through enhanced oxygenation of sediments, which in turn

enhances microbial removal of N from the water column and reduces anoxic release of

P (Kufel & Kufel, 2002). Expansion and contraction of wetted area are also likely to

provide strong controls on the biomass of autotrophic communities within intermittent

streams (Stanley, Fisher & Grimm, 1997). However, groundwater inputs of inorganic

Page 51: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

41

nutrients may stimulate aquatic vegetation growth even when pools are contracting, i.e.,

when the rate of evaporation and transmission losses exceed the rate of replenishment

by groundwater. This poses the question as to how the abundance of aquatic vegetation

affects the relationships between evaporative contraction and the concentration and

stoichiometry of water column nutrients.

The results of Chapter 2 revealed that changes in DOM biogeochemistry across several

catchments of the Pilbara region of northwest Australia were related to the connectivity

of pools to shallow alluvial groundwater (AW). This study focuses more closely on

pools of a single catchment to further unravel the biogeochemical role of AW

connections within these intermittent streams, with particular focus on changes in water

chemistry and ecological stoichiometry over periods of recharge from floods and

subsequent evaporative contraction. I sought to determine if: i) C:N:P stoichiometry

changed within the water column and aquatic plants after pools expanded or contracted;

ii) whether these stoichiometric changes were able to explain the differences in DOM

and nutrient concentrations between sampling periods (e.g., lower C:N ratios promoting

increased nitrification and thus NO3 concentrations), and; iii) if aquatic plants affect

resource stoichiometry by taking up and immobilising inorganic fractions of N and P,

thereby reducing their availability.

Methods

Study sites and sampling design

All of the pools sampled are situated along Coondiner Creek, an intermittent stream that

originates in the Hamersley Range in the south-east of the Pilbara region of northwest

Western Australia and drains approximately northeast through incised gorge country out

onto the low-lying Fortescue Valley (Fig. 3.1). The climate of the region is described in

Chapter 1, while the study site at Coondiner Creek is as described in Chapter 2.

The seven pools identified in Chapter 2 as having differences in alluvial water (AW)

connectivity (pools 1 - 7; sensu Fellman et al., 2011a) were chosen for sampling.

Typically, there is high temporal variation in wetted area both within and amongst pools

(Fig 3.2). Sampling occurred during the 8th – 9th September 2011 (excepting pool 7, due

to issues with site access), 26th – 28th June 2012, and 19th – 21st October 2012. In both

2011 and 2012, floods brought on by heavy rainfall early in the year (approx. 190 mm

during February 2011, and 130 mm over two days in January 2012) were followed by

Page 52: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

42

Figure 3.1: Map showing the locations of pools sampled along Coondiner Creek,

northwest Australia. Black square (inset) indicates area shown in the map relative to the

extent of the Pilbara bioregion. Pools with persistent alluvial water (AW) connections

are shaded black.

extended periods without surface flow in which rainfall events were generally low (<30

mm per day) and isolated (Chapter 2). Samplings thus corresponded to the late dry

season (Sept 2011), early-mid dry season (June 2012) and late dry season again (Oct

2012).

Aquatic and riparian vegetation within and around the pools was largely as described in

Chapter 2. However, pools within Coondiner Creek also contained the submerged

macrophytes Najas tenuifolia, N. marina and Utricularia australis, although these

species were much less common than the typical aquatic assemblages of benthic

charophytes (Chara spp.) and the floating macrophyte Potamogeton tricarinatus.

Page 53: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

43

Figure 3.2: Pools of differing persistence due to varying alluvial water (AW)

connections within Coondiner Creek, northwest Australia. Shown are a) “Pool 4” in

May 2011, b), September 2011, c) June 2012, and d) October 2012, as an example of a

pool which loses AW throughflow a few months after cessation of surface flow; and e)

“Pool 2” in May 2011, f), September 2011, g) June 2012, and h), October 2012, as an

example of a pool with AW throughflow which may persist for years between flood

events.

Page 54: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

44

Field methods

Pools were mapped for length, width, maximum depth, aspect, distance to and height of

adjoining gorge walls, and cover of terrestrial and aquatic vegetation. GPS coordinates

were taken along the edges of pools for reference. Conductivity (EC) and dissolved

oxygen (DO) concentrations were measured using a YSI 85 handheld probe (Xylem

Inc., USA), while pH was measured using a handheld ecoTestr pH 2 probe (Oakton

Instruments, USA).

Water samples for stable isotope and water chemistry analysis were collected as

outlined in Chapter 2. Water samples (approx. 0.5 – 2.5 L) for seston fine particulate

organic matter (FPOM) analysis were collected from 5 cm below the pool surface and

filtered through 0.7 µm pre-combusted quartz or glass fiber filters until the filters

became clogged. Filter papers were kept and stored at 4 oC until analysis. Between three

and six replicate FPOM samples were taken per pool.

Samples of aquatic plant and algal material were collected from pools in September

2011 and June 2012, composed of material taken from multiple individual plants or

colonies and consisting of two to three replicate samples per pool depending on pool

size. Samples of aquatic macrophytes and filamentous algae were removed from living

plants. In addition, whole conditioned (i.e. abscised, wet and partly decomposed) leaves

of Eucalyptus spp. and Melaleuca spp. (where present) were collected from the beds of

pools as the main component of detrital POM. All samples were stored at 4 oC until

analysis.

Analytical methods

Water samples were analysed for stable isotope composition and water chemistry as

outlined in Chapter 2, as well as for total dissolved phosphorus (TDP) and dissolved

organic phosphorus (DOP). Total dissolved phosphorus (TDP) was measured after

acidic oxidation to orthophosphate using the method of Solorzano and Sharp (1980).

Dissolved organic phosphorus (DOP) was calculated as the difference between TDP and

soluble reactive phosphorus (SRP).

Filters for FPOM analysis were air dried at 50 oC for 24-48 hours, after which

particulate matter was removed from the filters by scraping. Total C and N content were

determined by analysis using a continuous flow system consisting of an Elemental

Page 55: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

45

Analyser Thermo Flush 1112 (Thermo-Finnigan, Germany). However, both the C

(approx. 1 %) and N (approx. 0.1 %) content of all FPOM samples was too low to

obtain readings via this method, despite the large volumes of water sampled. I therefore

concluded that seston FPOM was an insignificant fraction of pool organic matter stores,

and excluded it from further analyses.

Detrital POM, plant and algal samples were air dried at 50 ⁰C for 24-48 hours, then

finely ground. Total C and N content were determined as above for FPOM. Phosphorus

content was determined using acid digestion (H2SO4/H2O2) and subsequent analysis on

a Technicon AutoAnalyzer (Technicon, USA).

Data analysis

The alluvial water connectivity of pools was determined as outlined in Chapter 2. Pool

area maps were created by cross-checking field measurements of pool and channel

morphology across all sampling periods with GPS coordinates taken on site, satellite

photos of the catchment from February 2012, and 1:250,000 topographical maps. The

morphology of pools was represented well by modelled pools with asymmetrically bell-

shaped distributions of depth. Volume of pool water was therefore estimated using

measurements of the area and maximum depth of each pool, where

volume (m3) = area (m2) ⁄ c × maximum depth (m)

and c is a constant of value 2.58, which I derived from the linear increase in volume of a

modelled pool with increases in area given constant maximum depth.

Ratios of C, N, and P to each other in water column solutes, detrital POM, aquatic

plants and algae were calculated using molar quantities. Differences between sampling

periods and pool hydrology (AW-connected versus isolated pools) in the physical

characteristics of the water column (DO, conductivity, and pH), the C:N:P

stoichiometry of solutes and the C:N:P stoichiometry of aquatic plants and algae were

identified using single-factor ANOVA at a confidence level of 95 %. C:N:P ratios were

compared against other and common indicators of shifts in ecological stoichiometry of

organisms in the Redfield ratio (106 C:16 N:1 P; the theoretically optimal ratio in

aquatic algae and bacteria) and the theoretical limit between C and N limitation of

microbial processes (20 C: 1 N) to broadly evaluate which nutrients may be limiting

(Sardans, Rivas-Ubach & Peñuelas, 2012).

Page 56: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

46

Chapter 2 indicated that evaporative concentration is a major control on solute

concentrations in the water column. I therefore identified changes in water chemistry

due to changes in volume alone. Theoretically, if the overall mass of a solute is

constant, then every percent decrease in volume (Δv; v1-v2/v1 × 100) leads to an

equivalent percent increase in concentration (Δc; c2-c1/c2 × 100). I thus calculated Δv

and Δc for each nutrient between sampling periods within each pool, and then compared

the observed changes in concentration between samplings with an idealised model

where Δc = Δv (i.e., increases in concentration were only due to decreases in volume)

using t-tests at α = 0.05. Significant differences between expected (E) and observed (O)

values of Δc therefore indicated that changes in concentration were occurring which

were due to production or utilisation of solutes, rather than changes in volume alone.

The area of pool beds covered by aquatic plants (aquatic macrophytes together with

filamentous algae) was calculated as the difference between total pool area and the

benthic area not covered by any plant biomass. Wetted area lost or gained was presented

as a ratio between each pool’s total area at subsequent samplings (Areat2:Areat1). Given

the non-linear nature of the scale for change in wetted area (i.e., contraction could only

result in values between 0 and 1 but expansion could effectively result in limitless

values above 1), I applied a log10 transformation to Areat2:Areat1 values before using

least-squares regression to identify the correlation between change in wetted area and

aquatic plant cover.

Biomass of aquatic plants within each pool was estimated by multiplying average

density of each major group (charophytes, P. tricarinatus, filamentous algae, T.

orientalis and S. subulatis) by that groups’ area of pool bed covered. Other aquatic

macrophytes were not considered in the biomass analysis due to their extremely low

density across all samplings. The amount of C, N, and P stored within each pool for

each major aquatic plant group was estimated by multiplying biomass by their average

total C, N, and P content. In order to determine whether C, N and P stores within plants

and algae were consistent across pools of different sizes, I transformed the total C, N

and P stores into densities (g m-2) by dividing by pool area, and then compared across

sampling periods and pool hydrology using single-factor ANOVA at a confidence level

of 95 %.

Page 57: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

47

Results

Hydrology and pool morphology

The alluvial water connectivity of pools is as described in Chapter 2. Pools were larger

in area by between 10 % and 260 % in the early-mid dry season (June 2012) compared

to nine months earlier in the late dry season (Sept 2011; Fig. 3.3). However, the pools

with the highest relative wetted area between the late and early-mid dry seasons had

Figure 3.3: Wetted area maps of pools within Coondiner Creek showing comparisons

between September 2011 (late dry season), June 2012 (early-mid dry season), and

October 2012 (late dry season). Note different scales for each pool.

Page 58: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

48

Table 3.1: Average (± standard error) molar C:N:P ratios in aquatic plants, terrestrial

detritus and algae within Coondiner Creek.

C:N N:P

Charophytes (Chara spp.) 22.6 (±4.2) 28.8 (±5.1)

Potamogeton tricarinatus 20.4 (±2.4) 25.5 (±2.7)

Typha orientalis 23.4 (±1.9) 29.0 (±2.2)

Schoenoplectus subulatis 32.4 (±7.8) 20.8 (±3.0)

Filamentous algae 22.4 (±3.9) 21.1 (±4.5)

Detrital POM 68.6 (±3.2) 58.8 (±2.8)

δ18O values indicative of both alluvial water connections (Pool 1: 260% expansion; Pool

5: 110% expansion) and isolation from AW (Pool 4: 120% expansion). Pools lost

between 43 and 97% of their wetted area between the early mid dry season and the

second late dry season sampling (Oct 2012; Fig. 3.3). The highest change in wetted area

was again observed in pools with δ18O values indicative of both AW throughflow (Pool

1: 97% contraction; Pool 6: 89% contraction) and isolation from AW (Pool 4: 98%

contraction; Pool 3: 88% contraction). Changes in the δ18O values of AW-isolated pools

were highly correlated with wetted area change (r2 = 0.92, P = 0.006), but changes in

the δ18O values of AW-connected pools were not correlated with wetted area change (r2

= 0, P = 0.51).

Ecological stoichiometry

The C:N ratios of aquatic plants and algae ranged from 13:1 to 55:1 (Table 3.1) but

were not significantly different between species (P = 0.35), sampling periods (P = 0.80)

or pools with differing AW connectivity (P = 0.41). Aquatic plant and algal N:P ratios

ranged from 12:1 to 38:1 (Table 3.1) and were also not significantly different between

species (P = 0.44), sampling periods (P = 0.11) or pools with differing AW connectivity

(P = 0.39). By comparison, detrital POM had much higher C:N:P ratios (Table 3.1),

although its stoichiometry also did not differ between AW-connected and isolated pools

(P = 0.5).

Page 59: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

49

Figure 3.4: Temporal variation in the ratios of average molar total dissolved a) carbon

(DOC) to nitrogen (TDN), b) organic nitrogen (DON) to phosphorus (DOP) and c)

inorganic nitrogen (DIN) to phosphorus (SRP) in pool water (± standard error) with

respect to pool alluvial water (AW) connectivity. Shown for comparison are a) the

general C:N ratio for microbial carbon limitation (dashed line; 20:1) and Redfield Ratio

for C:N (dotted line; 6:1), b) the Redfield ratio for N:P (dashed line; 16:1) and average

N:P of detrital POM (dotted line; 59:1), and c) the average (dashed line; 24.6:1) ±

standard deviation (dotted line; ±7.7) of N:P ratios for aquatic plants and algae across

all sampling periods. Note logarithmic scales in b) and c).

Page 60: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

50

DOC:DON ratios were not significantly different to those for DOC:TDN across any

sampling period (ANOVA P = 0.65 to 0.87), and so for succinctness only DOC:TDN

values are presented here. DOC:TDN ratios in the water column were highly variable

across both sampling periods and pool hydrology (P = 0.003). AW-connected pools had

very high water column DOC:TDN (52:1 ± 12) in the first late dry season sampling

(Sept 2011) but had low values (c. 10:1) in both of the following periods (Fig. 3.4a).

Conversely, pools isolated from AW had DOC:TDN values close to the theoretical

point at which microbial metabolism switches from C to N limitation (when C:N ratios

increase above c. 20:1) in both late dry season samplings, but lower values (14:1 ± 1) in

the early-mid dry season (Fig. 3.4). Overall, there were therefore few consistent

temporal patterns in C:N ratios or across pools with differing hydrology.

DON:DOP ratios did not differ across pool hydrology (P = 0.8) but were extremely high

in both late dry season samplings (>100:1) and relatively lower in the early-mid dry

season (c. 35:1 to 57:1). However, even when lower in the early-mid dry season,

DON:DOP ratios were still much closer to the N:P ratios of detrital POM than the

Redfield ratio (Fig. 3.4b). DIN:SRP ratios were also similar across AW-connected and

isolated pools (P = 0.25), and showed a similar pattern to DON:DOP ratios in that the

late dry season samplings had much higher average values (c. 70:1 to 130:1) than in the

early-mid dry season (c. 4:1 to 10:1). Thus, DIN:SRP values decreased well below

those of aquatic plants and algae in the early-mid dry season (Fig. 3.4c).

Table 3.2: Temporal variation in water quality represented by dissolved oxygen

concentrations (DO), specific conductivity (EC) and pH. Values in bold indicate a

significant difference (at α = 0.05) between AW-connected pools and pools isolated

from AW in the same sampling period.

Sept 2011 (late dry) June 2012 (early-mid dry) Oct 2012 (late dry)

AW-connected

Isolated AW-connected

Isolated AW-connected

Isolated

DO (mg L-1) 8.9 7.6 9.6 9.6 7.6 8.9

EC (μS m-1) 611.9 461.4 1015.4 1061.4 764.3 961.3

pH 8.1 8.3 7.9 7.6 7.8 8.0

Page 61: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

51

Table 3.3: Temporal variation of average dissolved C, N and P concentrations (in mg L-

1) in pools with respect to alluvial water (AW) input. Values in bold indicate a

significant difference (at α = 0.05) between AW-connected pools and pools isolated

from AW in the same sampling period. Values in italics indicate that average values

were below detection limits for N (0.008 mg L-1) or P (0.003 mg L-1).

Sept (late dry) June (early-mid dry) Oct (late dry)

AW-connected Isolated AW-connected Isolated AW-connected Isolated

DOC 6.48 4.74 0.84 1.34 1.88 7.84

DON 0.15 0.31 0.08 0.10 0.13 0.30

DOP 0.06 0.06 0.03 0.03 0.02 0.04

DIN 0.012 0.009 0.025 0.012 0.074 0.059

NO3 < 0.008 < 0.008 0.016 < 0.008 0.046 < 0.008

NH4 0.012 0.009 0.009 < 0.008 0.028 0.053

SRP 0.009 < 0.003 0.007 0.008 < 0.003 < 0.003

Temporal changes in water quality and water chemistry

Water quality (physical parameters of the water column) was generally similar between

all sampling periods and also between AW-connected and isolated pools (Table 3.2).

Dissolved oxygen (DO) did not differ significantly between sampling periods or

hydrological groups (P = 0.49). Conductivity was lower in the late dry season sampling

in September when compared with the other two sampling periods (P = 0.002), and was

also lower in pools isolated from AW at that time (Table 3.2). pH was remarkably

consistent across pools and sampling periods (mean pH = 7.8; P = 0.83).

Dissolved organic C and N concentrations were lower in the early-mid dry season than

in the late dry season samplings, but DOP concentrations were similar across all

sampling periods (Table 3.3). Difference between AW-connected and isolated pools

were also not consistent over time, with the only significant differences occurring in

DOC concentrations during the early-mid dry season, DON concentrations during the

first late dry season sampling and both DOC and DON in the second late dry season

sampling (Table 3.3). Most changes in DOC, DON and DOP concentrations between

sampling periods were not statistically different from what could be expected from

Page 62: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

52

Table 3.4: Average changes in C, N and P concentrations (in mg L-1) in pools between

sampling periods with respect to alluvial water (AW) connectivity. Bold values indicate

significant differences (at α = 0.05) between observed values (O) and values expected if

changes in concentration were only due to changes in volume (E). Note that values for

DIN (including NO3 and NH4) and SRP included some samples below detection limits

for N (0.008 mg L-1) and P (0.003 mg L-1).

Expansion: Sept (late dry) to June (early-mid dry)

Contraction: June (early-mid dry) to Oct (late dry)

AW-connected Isolated from AW AW-connected Isolated from AW

DOC E -3.39 -2.76 1.63 7.22

O -5.64 -3.17 1.04 6.50

DON E -0.09 -0.20 0.11 0.27

O -0.07 -0.20 0.05 0.20

DOP E -0.03 -0.03 0.02 0.03

O -0.02 -0.03 -0.01 0.01

DIN E -0.006 -0.005 0.067 0.045

O 0.013 0.006 0.049 0.047

NO3 E 0.000 -0.001 0.042 0.005

O 0.015 0.007 0.030 -0.001

NH4 E -0.006 -0.005 0.024 0.040

O -0.003 -0.002 0.019 0.047

SRP E -0.006 -0.001 0.003 0.002

O -0.002 0.005 -0.004 -0.006

changes in volume alone (Table 3.4), and were therefore consistent with the evaporative

concentration effect of DOM indicated by the results of Chapter 2. The exceptions were

seen in AW-connected pools after contraction, with a lower DON increase than

expected and a decrease in DOP concentrations when an increase was expected (Table

3.4).

DIN concentrations increased over all samplings, which in AW-connected pools was

mainly due to consistent increase in NO3 concentrations (Table 3.3). Conversely, NO3

Page 63: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

53

Figure 3.5: Changes in the proportion of pool beds covered in aquatic vegetation

(change in vegetative cover) between consecutive sampling periods in relation to

relative changes in pool wetted area (Areat2:Areat1) between the same samplings.

Dashed lines represent the boundaries between expansion and contraction (horizontal)

and loss or gain in vegetative cover (vertical).

concentrations remained mostly below detection limits in pools isolated from AW, with

greater differences in NH4 concentrations between sampling periods (Table 3.3). SRP

concentrations were low across all samplings (Table 3.3). Consequently, while NH4

concentrations were largely predictable by changes in volume, NO3 and therefore DIN

concentrations increased between the first late dry season sampling and the early-mid

dry season when they were expected to decrease (Table 3.4). SRP concentrations were

also largely different from what was predicted by changes in pool volume (Table 3.4).

Page 64: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

54

Nutrient storage within aquatic plants and algae

The proportion of pool beds covered in aquatic vegetation decreased as wetted area

contracted in pools isolated from AW, while AW-connected pools maintained relatively

consistent proportions of aquatic vegetation cover despite large changes in wetted area

(Fig. 3.5). Consequently, the amount of C, N and P stored within aquatic plant and algal

biomass was consistent across sampling periods in AW-connected pools, with means of

13.8 to 18.9 mg C m-2, 0.7 to 1.0 mg N m-2 and 0.06 to 0.08 mg P m-2 (Fig. 3.6).

However, nutrients held in aquatic plants and algae were more variable across pools

isolated from AW, averaging 21.8 mg C m-2, 1.2 mg N m-2 and 0.10 mg P m-2 in the

early-mid dry season but with a sharp decrease from the early-mid to late dry season in

October 2012 (ranging from 0 to 9.6 mg C m-2, 0 to 0.5 mg N m-2 and 0 to 0.04 mg P m-

2; Fig. 3.6).

Discussion

In this study, the N:P ratios of aquatic plants and algae were above the DIN:SRP ratios

of the water column within pools during the early-mid dry season but below those of

inorganic nutrients in the water column in both of the late dry season samplings. These

results indicate that a clear shift from N limitation to P limitation in aquatic vegetation

may occur as pools dry (Sterner & Elser, 2002). However, aquatic plants and algae

maintained consistent C:N and N:P ratios across sampling periods and pools of differing

AW connectivity. Aquatic plants and algae therefore either did not modify their growth

rates to respond to changing environmental conditions (Ågren, 2004) or, given their

relatively high N:P ratios here (c. >20:1), were generally limited by P availability

(Koerselman & Meuleman, 1996; Sardans et al., 2012). Furthermore, pools isolated

from AW lost aquatic plant cover upon contraction but AW-connected pools did not.

The AW-connected pools also maintained consistent stores of C, N and P within aquatic

vegetation across all sampling periods. Aquatic plants, and charophytes in particular,

can rapidly take up and immobilize inorganic nutrients in shallow, oligotrophic lotic

environments, as well as binding and oxygenating sediments to prevent the

resuspension of nutrients under anoxic conditions (Kufel & Kufel, 2002; Scheffer &

Nes, 2007). N and P concentrations within the water column were low across all

samplings, especially in comparison with other Australian waterholes (e.g., Sheldon &

Fellows, 2010). Soils and rocks of the ancient and arid environments of the Pilbara are

Page 65: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

55

Figure 3.6: Changes in the average density of a) C, b) N and c) P stored within aquatic

plants and algae (± standard error) in pools between sampling periods with respect to

alluvial water connectivity (AW).

Page 66: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

56

also generally low in P (Bentley et al., 1999; Ford et al., 2007; McIntyre et al., 2009).

AW throughflow may therefore provide a source of inorganic nutrients, particularly of

PO4, which is quickly removed from the water column by aquatic plants and algae

under relatively constant P limitation. In contracting pools, these nutrient inputs support

new growth and therefore help counteract the effect of loss of aquatic vegetation along

drying fronts.

The high water column DON:DOP ratios across all samplings suggest that microbial

communities in Coondiner Creek should also be highly limited by P. DOP availability

should increase with accumulations of terrestrial litterfall late into the dry season (von

Schiller et al., 2011; see also Fellman et al., 2013). However, given low concentrations

of inorganic P, P-limited bacteria are likely to devote much of their resources towards

the production of extracellular enzymes to break down organic P (Mooshammer et al.,

2011). The extremely high DON:DOP ratios in the late dry season samplings may

therefore be reflective of high mineralization and immobilization of P from detrital

POM, especially given the low PO4 availability at those times. Organic P availability

may also have been related to the composition of DOM, with high proportions of what

appeared to be highly-degraded DOM from AW throughflow (Chapter 2). The low C:N

ratios of DOM would appear to indicate that it was more labile in the early-mid dry

season, and possibly had lower N:P ratios as well. Alternatively, high concentrations of

humic DOM during the late dry season (Chapter 2) may have resulted in PO4 adsorption

(Maranger & Pullin, 2003); preventing its use by microbes and therefore necessitating

increased mineralisation of detrital POM.

Stoichiometric ratios in the water column also indicated that N availability was highly

variable over time. N:P ratios for both organic and inorganic forms were higher in the

late dry season compared with the early-mid dry season sampling, suggesting that DIN

(particularly NH4) concentrations may have increased due to increasing N fixation

(Sardans et al., 2012). However, NH4 concentrations were highest in the second late dry

season sampling when N:P ratios were high, and did not change more than expected

from the expansion and contraction of pools. Coondiner Creek also appears relatively

free of the dense cyanobacterial mats that form at downwelling sites and dominate N

fixation in other intermittent dryland streams (Grimm & Petrone, 1997; Fisher et al.,

2001). Given the high C:N ratios of OM in the late dry season samplings, NH4 may

therefore be linked to the mineralization of OM by labile C-limited microbial

Page 67: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

57

communities (Acuña et al., 2005; Fellman et al., 2011a; von Schiller et al., 2011).

Conversely, NO3 concentrations increased between the first late dry season sampling

and the early-mid dry season when increases in volume would suggest NO3 should

decrease. C:N ratios were low in the early-mid dry season and could therefore have

promoted nitrification and the accumulation of NO3 (Jones et al., 1995a; Jones, 2002).

DOM at this time may have been more labile than in the late dry season and also

predominantly derived from AW flowpaths (Chapter 2), thus promoting nitrifier activity

if the competing heterotrophic bacteria are more limited by P (Bengtsson, Bengtson &

Månsson, 2003). Low C:N could alternatively point to a reduction in DOC and

increased NO3 inputs in AW throughflow (Dahm et al., 2003). C:N ratios also remained

low in AW-connected pools in the second late dry season sampling, and although NO3

concentrations did not increase more than expected they were still much higher than

NH4. As NO3 concentrations were relatively low in pools isolated from AW over all

sampling periods, it is therefore likely that AW throughflow acts as a source of NO3 as

well as SRP.

In light of the above analysis, this study suggests that the ecological stoichiometry of

pools within Coondiner Creek is intrinsically linked to the low availability of P. The

ecological consequences of low P availability are likely to be seen in higher internal

recycling of nutrients within biofilms or sediments (Mulholland et al., 1995; Kufel &

Kufel, 2002; Scott et al., 2008). In Coondiner Creek, this tendency towards P

conservancy is also expressed through storage within aquatic plant biomass. Low P

availability is also likely to limit new pathways of C and N into pool ecosystems, such

as through algal production of labile DOM or nitrogen fixation (Howarth et al., 1988;

Wyatt et al., 2014). The strong evaporative concentration effect on solutes seen both

here and previously (Chapter 2) may therefore be due to the production and utilisation

of dissolved OM and nutrients not being expressed within the water column, rather than

there being no change in ecosystem processes as pools expand and contract. Similarly,

the C:N:P stoichiometry of the water column may also not be a consistent indicator of

the biogeochemical processes operating within pools. In addition, AW throughflow may

further obscure stoichiometric indicators of DOM and nutrient cycling due to consistent

dilution or purging of the water column. However, the effects of AW connectivity do

appear to be expressed in the ecological stoichiometry of pools through NO3 and PO4

inputs and their effect on the persistence of aquatic vegetation. These results therefore

emphasize that the effects of flow intermittency on ecological processes within dryland

Page 68: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

58

rivers and streams will be highly dependent on connections between surface and sub-

surface water.

Page 69: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

59

CHAPTER FOUR: Diel cycles of 13C-dissolved inorganic carbon (DIC)

and ecosystem metabolism in hydrologically variable pools of

intermittent streams

Introduction

In previous chapters, I have shown that dissolved organic matter (DOM) concentrations

are highly affected by cycles of expansion and contraction in pools of the Pilbara

(Chapters 2 and 3). However, dissolved inorganic carbon (DIC; including CO2, HCO3-

and HCO32-) is also an important component of the biogeochemistry and functioning of

aquatic ecosystems. Production and removal of CO2 from the water column strongly

affects pH and therefore chemical speciation, mineral precipitation and dissolution,

nitrification, rates of primary productivity, and microbial growth (Nimick, Gammons &

Parker, 2011). In freshwater systems, DIC dynamics are the result of a complex

combination of external (allochthonous) and internal (autochthonous) processes

including microbial respiration, photosynthetic activity, atmospheric exchange,

carbonate weathering and groundwater inputs (Finlay, 2003). The relative importance of

each of these factors can vary over hourly, daily and seasonal cycles as a function of

temperature, light and hydrology (Tobias & Böhlke, 2011) and also at different spatial

scales, e.g. among different reaches or with distance from groundwater seepage or if a

system is high in carbonate materials (Waldron, Scott & Soulsby, 2007; Dogramaci et

al., 2015; Perkins et al., 2015). In particular, diel variation in DIC is often linked to

ecosystem metabolism (broadly defined as the variation in gross primary production and

total ecosystem respiration; Smith & Hollibaugh, 1997) owing to utilization of aqueous

CO2 or HCO3- by photosynthetic organisms during the day and respiration at night

(Nimick et al., 2011). However, most studies of riverine ecosystem metabolism, and

carbon fluxes more generally, have focussed on catchments of temperate and boreal

regions as well as the wet tropics. Relatively little is known of the dynamics of

ecosystem metabolism of temporary and intermittent rivers more common in dryland

regions; especially with regard to short-term DIC dynamics and changes in the

transitions between hydrological phases, for example as pools contract and dry during

inter-flood periods.

While the dynamics of flood or flow pulse cycles are recognised as vital to the

Page 70: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

60

biogeochemical cycles and productivity of dryland rivers and streams (Bunn et al.,

2006b), the effects of drought on aquatic ecosystems are less well understood (Lake,

2003), in large part due to the complexity of interacting processes. Between floods,

dryland rivers and streams in warm regions have infrequent flow disturbance and high

light availability (Fellows et al., 2009). The long-term accrual of algal biomass is

typically limited more by disturbance (e.g. Grimm & Fisher, 1989) and overall

productivity by light (Hill, 1996). Consequently, primary productivity in dryland

streams tends to be relatively high, even when nutrients are considered limiting (Grimm

& Fisher, 1986) or there is high turbidity (Bunn et al., 2003). Higher rates of

photosynthesis by macrophytes and algae during these low or no flow periods may also

result in increased inputs of labile organic matter, which in turn supports bacterial

respiration and production (Bertilsson & Jones, 2003). Where inputs of terrestrial

organic matter are low, stream or pool respiration can therefore be highly correlated

with algal production (Jones, Fisher & Grimm, 1995b; Fellows et al., 2007). However,

cessation of surface flow can also increase retention of allochthonous inputs, which may

also lead to higher heterotrophic respiration rates (Acuña et al., 2004). Within streams

or pools productivity and respiration during inter-flood periods can also be affected by

the transport of organic matter and nutrients in subsurface flowpaths (Fisher et al.,

1998a) or autotrophic community changes along drying fronts (Stanley et al., 1997).

The results of Chapter 3 suggest that nutrient inputs in alluvial water (AW) throughflow

may maintain autotrophic communities even when pools are contracting. Diel cycles of

DIC in dryland streams may therefore be largely controlled by ecosystem metabolism

during inter-flood periods, and differences between pools may also be attributable to

increased autotrophic production and respiration rates in pools with alluvial water

connections.

The stable isotopic composition of DIC (δ13CDIC) is a widely used tracer of aquatic

ecosystem system processes. In net autotrophic and highly productive rivers and

streams, δ13CDIC variation over diel cycles is largely controlled by discrimination

against the heavier 13C isotope during photosynthesis (Farquhar, Ehleringer & Hubick,

1989), resulting in regular daily increases of δ13CDIC values that are proportional to rates

of primary productivity (Parker et al., 2005; Parker et al., 2010). Increases of δ13CDIC

due to photosynthetic CO2 uptake can also be enhanced by carbonate precipitation

associated with the consequent increase in pH (de Montety et al., 2011). Respiration

typically produces CO2 with more negative δ13C values, reflecting highly 13C-depleted

Page 71: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

61

organic matter sources and resulting in night-time decreases in δ13CDIC. When primary

productivity is low, δ13CDIC may therefore show little variation at diel timescales

(Waldron et al., 2007). In addition, diel cycles of δ13CDIC can be associated with

fluctuations in temperature, as streams with carbonate-rich source water experience CO2

outgassing as calcite solubility decreases during the day (Drysdale, Lucas & Carthew,

2003). In open systems, this can result in increases in δ13CDIC as discrimination occurs

during the chemical equilibrium of CO2(aq) with HCO3- (Doctor et al., 2008), and may

create diel cycles of δ13CDIC independent of primary productivity and respiration.

However, photosynthetic uptake of CO2 may cause regular diel cycles of δ13CDIC even

when hydrological inputs of DIC are large (Tobias & Böhlke, 2011). In intermittent

streams of the Pilbara, pools in the constrained gorge systems where AW connectivity is

likely to occur often have substantial proportions of underlying carbonaceous geology

(Fellman et al., 2011a; Dogramaci et al., 2015). Weeli Wolli Creek in particular

dissects an extensive calcrete surface (Dogramaci et al., 2015) suggesting carbonate-

rich source water. In such systems (see Chapters 2 and 3), this raises the question as to

what contributes most to diurnal cycles of δ13CDIC and to what extent variation in

ecosystem metabolism among pools is related to AW connectivity.

Here, I investigate the δ13CDIC signature in pools of contrasting hydrology (connectivity

to AW) during an inter-flood period to (i) evaluate whether δ13CDIC is an effective tracer

of diel cycles of ecosystem metabolism, and (ii) if ecosystem metabolism was higher in

pools with AW throughflow.

Methods

Study sites and sampling design

Study sites were at Coondiner Creek and Weeli Wolli Creek in the Pilbara region of

northwest Australia, as described previously in Chapters 2 and 3. All samples were

collected from seven pools along Coondiner Creek (pools 1-7; see Chapters 2 and 3)

and Weeli Wolli Creek (four pools in a reach of the creek known as Ben’s Oasis; see

Chapter 2). Both creeks were sampled during inter-flood periods approximately 9

months after the cessation of flow following the last major rainfall event, Cyclone Heidi

(Chapter 2). At Coondiner Creek, four pools (pools 1-4; here referred to as CND1-4)

were monitored for ecosystem metabolism from 19th to 21st October, 2012. At Ben’s

Page 72: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

62

Oasis on Weeli Wolli Creek, two pools (BO3, BO4) were monitored from the 15th to

17th of November (Fig. 4.1). Climatic conditions were similar across sampling periods.

At Coondiner Creek, pools 1 & 2 (CND1, CND2) receive consistent throughflow of

alluvial water (AW), while pools 3 & 4 (CND3, CND4) were likely isolated from AW

input shortly after flooding events (see Chapter 2; Fellman et al., 2011a). At Ben’s

Oasis, pools BO3 and BO4 are suspected to receive year-round groundwater inputs:

BO3 is the largest and deepest in the reach, and BO4 is a smaller and shallower pool

directly downstream of BO3. Owing to logistic constraints, a subset of pools (CND2,

CND3, and BO3) was chosen to assess diel variation in AW throughflow and

evaporation, dissolved inorganic carbon and nutrients over a total of 26 to 28 hours at

each pool. CND2 was sampled from 3 pm on the 19th to 7 pm on the 20th Oct 2012,

CND3 was sampled from 10 am on the 20th to 2 pm on the 21st of Oct 2012 and BO3

was sampled from 6 am on the 16th to 8 am on the 17th of November 2012.

At the time of diel sampling, pools differed substantially in their autotrophic

communities. In Coondiner Creek, CND1 and CND2 were dominated by assemblages

of charophytes (Chara spp.), emergent (Typha orientalis and Schoenoplectus subulatis)

and floating (Potamogeton tricarinatus) aquatic macrophytes, and attached filamentous

Figure 4.1: Location of the study pools along a) Coondiner Creek and b) Ben’s Oasis in

northwest Australia.

Page 73: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

63

algae in 1-3 m wide bands along the margins of each pool. CND3 was shallow (average

0.14 m deep), but charophytes and filamentous algae were restricted to very narrow

(<1m) patches along the margins. CND4 had a water column that was opaque with

green algae. In all of these pools bar CND4, there was conspicuous build up of

terrestrial leaf litter (predominantly Eucalyptus and Melaleuca spp.) on the pool bed,

particularly so in CND3. At Ben’s Oasis, the water column of BO3 was dark with

tannins, with aquatic macrophytes confined to very small (<1m2), sporadic patches

along the shallowest margins of the pool. The water column of BO4 was similarly

stained, but the pool was shallow (average 0.4 m deep) and aquatic macrophytes more

abundant. BO3 was too deep and dark to estimate terrestrial organic matter build up at

the bottom of the pool, but leaf litter was abundant along the pool bed of BO4, and both

pools were fringed by dense riparian woodland.

Field methods

Dissolved oxygen (DO) and temperature were monitored in surface water at 15 minute

intervals for approximately 48-50 hours in CND1-4 and BO3-4, using a TPS WP-82Y

DO-temperature meter and autologger in combination with an EDYSI DO sensor (YSI

Inc., Yellow Springs, OH, USA). The sensor was placed in surface water, at a depth of

20 cm, at the approximate thalweg of each pool. Cloud cover was absent in the

mornings, but increased to low amounts (<30%) from midday to late afternoon during

both the October 2012 and November 2012 sample periods. Water samples for δ13CDIC,

δ18OH2O and δ2HH2O, dissolved organic matter (DOM), and nutrient analyses were taken

from the same area of each pool selected for diel monitoring of DO measurements. Two

40 mL samples of water were collected from 5cm below the surface, filtered

immediately through 0.2μm sterile PES syringe filters, and stored without headspace in

airtight, sterile glass vials at 4 ⁰C in the dark until analysis. Samples were taken every 2

hours over the 26-28 hour sampling periods. These δ18OH2O and δ2HH2O samples were

compared to those taken at the other sampling periods in these catchments described in

Chapters 2 and 3 in order to compare diel cycles in the stable isotopic composition of

pool water to long-term trends.

Pools were mapped for length, width, depth, aspect, cover of terrestrial and aquatic

vegetation, and distance to and height of adjoining gorge walls while in the field in

order to assess local differences in light exposure. Conductivity and salinity were

Page 74: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

64

measured using a YSI 85 handheld probe (Xylem Inc., Yellow Springs, OH, USA),

while pH was measured using a handheld ecoTestr pH 2 probe (Oakton Instruments,

Vernon Hills, IL, USA). In the Coondiner catchment, photosynthetically active

radiation (PAR), wind speed and direction, air temperature, relative humidity, and

cumulative rainfall were measured at 15 minute intervals using a HOBO weather station

(Onset Computer Corporation, Cape Cod, MA, USA), with data extending continuously

back from the sampling period to October 2010.

Analytical methods

The stable isotope composition of water (δ18OH2O and δ2HH2O) and water chemistry

(DOC, DON, DIN and SRP) were analysed as described in earlier chapters (Chapters 2

and 3). In addition, the fluorescence index (FI) of dissolved organic matter (DOM) was

calculated as the ratio of fluorescence intensity at emission wavelengths of 450 and 500

nm, at an excitation wavelength of 370 nm, on a Varian Cary Eclipse fluorescence

spectrometer (Varian Inc., Australia), as an indicator of the possible source of DOM

(McKnight et al., 2001). Samples for δ13CDIC were acidified under He atmosphere with

orthophosphoric acid, and the produced CO2(gas) was analysed using a GasBench II

coupled with a Delta XL Mass Spectrometer (Thermo-Fisher Scientific Inc., Bremen,

Germany). δ13CDIC values are given in per mil (‰, VPDB) according to delta notation,

following three-points normalization based on international standards provided by

IAEA (L-SVEC, NBS19, and NBS18) used in order to reduce raw values to the

international scale (Skrzypek, 2013). Values of international standards for carbon (δ13C)

are based on Coplen et al. (2006). The analytical error of δ13C analysis is <0.10‰.

Data analysis

I calculated pool evaporative losses between sampling periods in the Coondiner Creek

catchment using Hydrocalculator ver. 1.02 (Skrzypek et al., 2015), following the model

of Craig and Gordon (1965) as modified by Gonfiantini (1986). CND1 and CND2 were

calculated assuming a steady-state, whereas CND3 and CND4 were calculated

assuming a non-steady state. The local meteoric water line (LMWL; the relationship

between δ18O and δ2H in local precipitation events) and local evaporation line (LEL; the

relationship describing variation in δ18O and δ2H under local evaporation conditions) for

Page 75: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

65

δ18OH2O and δ2HH2O were taken from Dogramaci et al.’s (2012) hydrological

characterization of the Hamersley Basin. For the Coondiner Creek sites, I did not adjust

the model for impacts of rainfall as between June and October 2012 only 9.6 mm of

rainfall was recorded, with no daily cumulative rainfall above 1.2 mm (Chapter 2). I

assumed groundwater to be derived from the last major rainfall (Cyclone Heidi, 10-13th

Jan, 2012; approximately 120 mm) for which isotopic values (δ2H = -55.5, δ18O = -

9.15) were obtained from samples collected by rainfall collector at the Hope Downs

mine site (12 km to Ben’s Oasis, 62 km to Coondiner Creek).

At Ben's Oasis, the δ18OH2O and δ2HH2O values for the November 2012 sampling did not

all fall along the LEL. The Hope Downs rainfall gauge recorded 25 mm of rainfall over

the 24th October to 12th November, with a daily maximum of 12.8 mm. Thus, I could

not assume that rainfall did not affect the catchment between the October and

November samplings. As a result, I constructed a two-stage mixing model to infer the

hydrological regime of individual pools (Kendall & McDonnell, 2012), with the

assumption that pools could connect unidirectionally along longitudinal gradients (i.e.,

pools could be affected by pools upstream but not downstream of them) and that

decreases in δ18OH2O reflected mixing of water from rainfall or pools upstream with

separate and relatively depleted sources (i.e. groundwater). To summarise, BO4 (-7.14

‰) likely had high throughflow; the δ18OH2O values of BO3 (-3.05 ‰) and BO1 (-1.23

‰) suggest some seepage of subsurface water but also evaporation; and BO2 (5.79 ‰)

was likely disconnected from any subsurface flow and acting as an evaporation pan for

rainfall or floodwater. However, there was some shallow surface water connection

between BO3 and BO4 that suggests BO4 was acting as overflow from BO3. It is

therefore likely that seepage from BO2 infiltrates through alluvium and mixes at the

head of a spring or upwelling originating at BO3.

A malfunction in the data logger at CND2 meant there were insufficient DO and

temperature measurements to construct precise metabolic models. Consequently, I

reconstructed DO curves in CND2 using the few measurements which were logged and

the strong relationship between DO and relative PAR (PAR x relative area of pool

shaded at each time point; r2 = 0.75-0.98, p < 0.01) at each of the other sites. Shading

models were created for each of the pools sampled by cross-checking field

measurements of pool habitat and landscape features with GPS coordinates taken on

site, satellite photos, digital elevation models, 1:250,000 topographical maps and prior

mapping of pool habitat (see Chapter 3). 3D models of each pool and its surrounding

Page 76: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

66

landscape were built in Autodesk Ecotect Analysis 2011 (Autodesk Inc., Waltham, MA,

USA), and maps of shading over each pool were calculated using sun track data from

the specific coordinates and times sampled. Unless otherwise stated, DO estimates for

CND2 were not included in any statistical tests involving their values or derivatives.

For each pool, gross primary productivity (GPP; g C L-1 d-1) and ecosystem respiration

(ER; g C L-1 d-1) were calculated from diel DO measurements using the single station

method (Odum, 1956), from which I also utilized estimates for each 15-minute period

of night time respiration (Rnight, mg O2 L-1 min-1) and daytime GPP flux (mg O2 L-1 min-

1). I estimated reaeration flux (k) using regression between the night time DO saturation

deficit and change in DO (Hornberger & Kelly, 1975). Mean k across all pools was 0.11

d-1 and was modified for temperature at each time step (Thyssen et al., 1983). GPP and

ER measurements were converted to areal data (mg C m-2 d-1) by multiplication with

average depth.

Change in δ13CDIC over time (Δ13Ct2-t1, ‰ hour-1) was calculated from average δ13CDIC

values at each 2-hour time step, and changes in GPP, ER, and temperature were

standardized to 2-hour intervals to aid in comparison. Unless otherwise stated,

regressions between Δ13Ct2-t1 and other variables were linear least-squares. All statistical

analyses (linear regression, ANOVA, T-test) were conducted in MATLAB v7.14

(MathWorks, Natick, MA, USA) at a confidence level of 94%, and all r2 values given

are adjusted for sample size.

Results

Diel cycles of δ13CDIC

δ13CDIC showed regular diel patterns across all three pools sampled, with increasing

values during the day balanced by decreasing values during the night (Fig. 4.2c).

Coondiner and Ben's Oasis sites differed in average δ13CDIC values (CND2 = -10.66‰,

CND3 = -10.29‰, BO3 = -12.53‰); similarly, diel ranges in δ13CDIC values differed

slightly between sites (CND2 = 0.90‰, CND3 = 0.91‰ compared to BO3 = 0.64‰).

Change in δ13CDIC over time (Δ13Ct2-t1) was not correlated with changes in δ18OH2O

(Δ18Ot2-t1) or δ2HH2O (Δ2Ht2-t1) over diel cycles in any pool.

Page 77: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

67

Figure 4.2: Diel variation in a) dissolved oxygen (DO) concentrations, b) temperature,

and c) δ13CDIC values of dissolved inorganic carbon across Coondiner Creek pools 2

(CND2) and 3 (CND3) and Ben’s Oasis pool 3 (BO3). Shaded areas represent night-

time measurements (approx 18:05 – 05:20). Note: points of temperature and DO for

CND2 are partly modelled due to missing data.

Page 78: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

68

Times of δ13CDIC maxima and minima were generally consistent with patterns of DO,

but maxima of δ13CDIC did not always occur close to times of maximum DO

concentrations (Fig. 4.2a, c). Rate of change in δ13CDIC (Δ13Ct2-t1) increased with higher

GPP flux in CND3 but not in BO3 (Fig. 4.3a), while Δ13Ct2-t1 increased with higher

Figure 4.3: Relationships between rates of change in dissolved oxygen due to a)

ecosystem metabolism during the day (approx. 05:20 – 18:05; GPP flux) and b) night

(Rnight) on rates of change in δ13CDIC values (Δ13Ct2-t1) in pool 3 of Coondiner Creek

(CND3) and pool 3 of Ben’s Oasis (BO3). Regression lines indicate significant

relationships for a) CND3 and b) BO3.

Page 79: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

69

Figure 4.4: Relationship between rates of change in temperature (ΔTt2-t1) and rates of

change in δ13CDIC values (Δ13Ct2-t1) in pool 3 of Coondiner Creek (CND3) and pool 3 of

Ben’s Oasis (BO3). Regression line indicates a significant relationship containing data

from both pools.

Rnight values in BO3 but not CND3 (Fig. 4.3b). Diel temperature curves more closely

resembled those of δ13CDIC (Fig. 4.2a, c), and more rapid changes in temperature (ΔTt2-

t1) were associated with higher rates of Δ13Ct2-t1 (Fig. 4.4). The slope of this relationship

was consistent across both CND3 and BO3 (T-test P = 0.46; Fig. 4.4).

Diel patterns in ecosystem metabolism

Concentrations of dissolved oxygen (DO) exhibited regular diel variation across all

pools. However, the range in DO values differed among pools, from 0 to 8 mg L-1 in

CND4 to 0.1 to 1.4 mg L-1 in BO3 (Table 4.1). Timing of maxima varied slightly

among pools and was generally between 2 pm and 8 pm (Fig. 4.2a). Water temperatures

also exhibited consistent diel variation, with warmest temperatures (25.1 to 29.5 ⁰C) at

4 pm decreasing to 17.3 to 23.3 ⁰C just before dawn (Table 4.1). Pools were therefore

more consistent in timing of temperature maxima than DO (Fig. 4.2b). Gross primary

Page 80: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

70

Table 4.1: Physical, chemical, and metabolic characteristics of pools in Coondiner

Creek (October 2012) and Ben’s Oasis (November 2012). Notes: * indicates that results

are partly based on modelled values due to missing data; - indicates values were not

measured.

pH Temp (⁰C)

δ13CDIC (‰) DO

(mg L-1) FI GPP (mg C m-2 d-1)

ER (mg C m-2 d-1)

GPP:ER

Coondiner Creek

Pool 1 (CND1) 7.8 22.0 to

29.5 -

0.5 to 7.8 1.43 296.14 364.48 0.81

Pool 2 (CND2) 7.9 20.4 to

25.1† -10.93 to

-10.03 1.1 to 5.1† 1.38 391.01* 478.94* 0.78*

Pool 3 (CND3) 7.6 20.3 to

26.3 -10.86 to

-9.96 0.9 to 4.4 1.56 87.92 127.17 0.69

Pool 4 (CND4) 7.7 17.3 to

26.3 -

0.0 to 8.0 1.68 379.12 482.55 0.79

Ben’s Oasis

Pool 3 (BO3) 7.9 23.3 to

27.0 -12.80 to

-12.16 0.1 to 1.4 1.36 250.15 760.42 0.33

Pool 4 (BO4) 8.0 21.0 to

25.6 -

1.1 to 4.6 1.36 155.28 176.20 0.88

production (GPP) and ecosystem respiration (ER) estimates also varied across pools

(Table 4.1). GPP and ER were strongly correlated (r2 = 0.99, p < 0.01) except at BO3.

However, neither GPP nor ER, or the ratio between them, correlated with any indicator

of the hydrological regime (δ18OH2O, δ2HH2O, or modelled evaporative loss) or with any

biogeochemical parameter (DIN, FI, DOC, or DON).

Diel cycles in nutrients and DOM

Concentrations of nutrients and DOM did not exhibit strong diel patterns. Dissolved

inorganic P (SRP) was below detection limits (0.025 µg L-1) in all samples. DIN

Page 81: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

71

concentrations did not differ significantly across diel time periods (NO3-N, P = 0.19;

NH4-N, P = 0.68). The FI also did not change over diel time periods but differed among

pools (Table 4.1). Sites at Ben's Oasis had predominantly terrestrially-derived DOM

(1.36 to 1.38 for pools BO3, BO4, CND2) while pools at Coondiner Creek had more in-

stream derived DOM (1.43 to 1.68 for pools CND1, CND3, CND4; McKnight et al.,

2001). Pools with higher δ18OH2O values had higher FI values (r2 = 0.72, P = 0.02),

indicating an increase in the proportion of autochthonous to allochthonous organic

matter within more highly evaporated pools, unlike previous observations (Chapters 2

and 3).

Diel patterns in AW throughflow and evaporation

Diel variation of δ18OH2O and δ2HH2O values differed widely among pools (Fig. 4.5).

δ18OH2O and δ2HH2O values in CND2 did not decrease overnight, but did increase during

the day (Fig. 4.5a). In contrast, δ18OH2O and δ2HH2O values in CND3 decreased

overnight but did not consistently increase during the day (Fig. 4.5b). In BO3, δ18OH2O

and δ2HH2O values increased and then decreased during both the day and night (Fig.

4.5c). The overall range of variation in diel δ18OH2O values (approximately 0.3-0.8 ‰)

and δ2HH2O values (approximately 2-4 ‰) was low compared to the analytical

precisions of 0.1 ‰ for δ18O and 1.0 ‰ for δ2H, and both δ2HH2O and δ18OH2O were

highly inconsistent between consecutive samplings, particularly in CND2 and CND3

(Fig. 4.5). There was therefore little evidence for any consistent diel patterns in

evaporation or AW throughflow.

Discussion

This study revealed clear and consistent diel cycles of δ13CDIC in pools of contrasting

alluvial water (AW) connectivity. Variation in δ13CDIC values across pools of these

dryland streams were not consistently correlated with δ18O, δ2H, or dissolved oxygen

measurements but instead clearly reflected diel variation in temperature. These results

suggest that physical processes, rather than ecosystem metabolism, are the primary

control on inorganic carbon cycling in these systems, regardless of AW connectivity.

Page 82: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

72

Figure 4.5: Diel variation in δ18O (closed circles) and δ2H (open circles) of surface pool

water for a) Coondiner Creek pool CND2 and b) CND3, and c) pool BO3 of Ben’s

Oasis. Shaded areas represent night-time measurements (approx 18:05 – 05:20).

Page 83: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

73

The δ13CDIC values measured here, of -9.9 to -10.9 ‰ in Coondiner Creek and -12.2 to -

12.8 ‰ in Ben’s Oasis, are similar to values for relatively large lakes with neutral pH

and low primary productivity (Bade et al., 2004) or mid-size streams with significant

groundwater contributions (Finlay, 2003). In the case of Coondiner Creek, this likely

represents significant contributions from alluvial water (measured from -13.7 to -17.7

‰ in the alluvial bore at the Coondiner Creek gorge outlet; Skrzypek pers.

communication July 2015) to pool water budgets, although the lack of difference in

δ13CDIC between the AW-fed CND2 (-10 to -10.9 ‰) and isolated CND3 (-9.9 to -10.5

‰) suggests that other factors are contributing to these values. One possibility is the

higher rates of GPP in CND2 enriching DIC (Farquhar et al., 1989; Finlay, 2003).

Alternatively, as these sites were sampled at least ten months after cessation of surface

flow all pools may simply have equilibrated closer to atmospheric CO2 (approx -8.2 ‰;

Keeling et al., 2010), especially given the low rates of both GPP and ER.

Diel patterns of δ13CDIC were highly consistent and much less variable across pools than

those of dissolved oxygen, contrasting with studies of eutrophic, net autotrophic, and

perennial streams (Parker et al., 2005; Parker et al., 2010; Tobias & Böhlke, 2011)

which typically observe corresponding variation in the diel range of both δ13CDIC and

DO across different sites. Rates of ecosystem metabolism in these pools may therefore

be too low to significantly affect δ13C values over diel time periods, despite the strong

correlation between GPP flux and change in δ13CDIC values in pool CND3 of Coondiner

Creek. However, under low nutrient concentrations primary productivity is likely to be

limited by light availability (Karlsson et al., 2009) and therefore highly correlated with

temperature, especially in the shallow (average 0.14 m depth) and clear water CND3.

The consistent diel δ13CDIC patterns across both pools therefore suggest that diurnal

temperature cycles are the driver of δ13C patterns, especially given that the rate of

Δ13Ct2-t1 is consistent with each step in temperature across both BO3 and CND3. The

most likely mechanism is through changes in CO2 solubility, where rising temperatures

and reduced solubility during the day result in outgassing. Changes in CO2

concentration due to varying solubility should discriminate against heavier isotopes in

the water column, especially in open, pH neutral systems where discrimination also

occurs in the chemical equilibrium between CO2(aq) and the more abundant HCO3-

(Doctor et al., 2008), generating the consistent daily increases in δ13C values seen in

these pools. However, CO2 gas dissolution associated with cooling temperatures during

the night is not associated with significant discrimination, and should result in isotopic

Page 84: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

74

values moving closer to equilibrium with atmospheric CO2, which does not explain the

nightly decreases in δ13C values across these pools.

Regular diel cycles in δ13CDIC may instead reflect the overall metabolic state of these

pools, rather than daily fluctuations in GPP and ER. If respiration of both terrestrial and

autochthonous organic matter is enhanced by primary production, then respiration of

organic matter with low δ13C values (approx. -28 to -30 ‰ for both allochthonous and

autochthonous; see Chapter 5) may result in consistent production of CO2 with low δ13C

(McCallister & Giorgio, 2008) over entire diel cycles due to the predominance of ER

over GPP, but with peaks during the day. If CO2 concentrations are also maintained

near super-saturation due to net heterotrophy (Sobek, Tranvik & Cole, 2005) and warm

temperatures (Marotta et al., 2009), this may result in greater rates of outgassing during

the day. Differences in uptake of CO2(aq) (or possibly HCO3-) by photosynthetic

organisms between pools may also be buffered by associated increases in respiration if

ER is highly linked to GPP (Tobias, Böhlke & Judson, 2007), perhaps obscuring any

effect of photosynthesis on δ13CDIC. Temperature-driven changes in CO2 solubility

would therefore enable retention of low-δ13C CO2 respired during the night while

enhancing outgassing and CO2(aq) ↔ HCO3- fractionation during the day, creating

regular daily increases and nightly decreases in δ13CDIC.

Relationships between ecosystem metabolism and hydrology

The rates of GPP (0.09 – 0.37 g C m-2 d-1) and ER (0.13 – 0.78 g C m-2 d-1) observed

here are within the lower ranges of Bunn et al.’s (2006a) review of ecosystem

metabolism in dryland rivers, and comparable to those in central Australian, lowland

river waterholes (Fellows et al., 2007). However, those systems tend to be highly turbid

and dominated by benthic algal production restricted to narrow photic zones (Bunn et

al., 2003), whereas the macrophyte-dominated Coondiner Creek is probably more

analogous to the low-nutrient, groundwater-maintained Daly River in tropical Northern

Australia; not only in its water chemistry, hydrology, and rates of ecosystem

metabolism, but also in the close relationship between GPP and ER under net

heterotrophic conditions (Webster et al., 2005).

This relationship between GPP and ER is unusual compared to most other stream

ecosystems, and may be reflective of a priming effect; an increased mineralization of

Page 85: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

75

recalcitrant organic matter in the presence of labile C inputs (Guenet et al., 2010). If

algal growth rates are restricted by oligotrophic conditions, high photosynthetic rates

(such as those expected in the high light and temperature environments of warm dryland

streams) may result in GPP fuelling the production and release of readily metabolised

DOM (Townsend, Webster & Schult, 2011), priming increased

respiration of other organic material by non photosynthetic microorganisms. Increasing

GPP may therefore contribute to an increasing abundance of autochthonous DOM after

cessation of flow (Fellman et al., 2011a; Vazquez et al., 2011). An increase in

autochthonous DOM is not, however, reflected in the DOM chemistry of these pools,

where neither the FI or DOC:DON ratios have any correlation with GPP, ER, or the

ratio between them, and is consistent with the conjecture earlier in this thesis (Chapter

2) that autochthonous processes have little effect on DOM concentration or composition

in pools of the Pilbara. Rates of ER were also low (0.4 – 2.4 g O2 m-2 day-1) compared

to other dryland streams with high accumulations of terrestrially-derived benthic

organic matter (Acuña et al., 2004). A possible explanation may be that differences in

morphology and concentrations of coloured DOM (CDOM) among pools may also

create high variation in primary productivity (e.g. Karlsson et al. 2009). Preliminary

data on light attenuation within Coondiner Creek suggests that attenuation coefficients

are variable across pools but can be as high as 0.97 even within the first 0.5 to 0.6 m of

water.

While pools differed in their AW connectivity, these differences were not apparent over

diurnal periods, as δ18O and δ2H values did not appear to display consistent diel cycles.

While pool 3 of Ben’s Oasis did show some consistent patterns in both δ18O and δ2H,

the pattern of increase and decrease observed during both the day and night is

inconsistent with discrimination against heavier isotopes through evaporation during the

day or any consistent mixing from groundwater with lower δ18O and δ2H values. As a

result, changes in δ18O and δ2H could not be related to diel patterns in either ecosystem

metabolism or δ13CDIC, and it is inconclusive whether substantial hydrological variation

occurred over the study period. However, the hydrological regime of pools may still

contribute to a close relationship between GPP and ER in a wider sense. Cessation of

surface flow and subsequent loss of upstream subsidies (e.g. Wiegner et al., 2005) is

likely to increase tight internal cycling of nutrients and organic matter within microbial

communities (Mulholland et al., 1995) regardless of production rates, particularly under

oligotrophic conditions (Mulholland et al., 1991; Scott et al., 2008). The charophyte

Page 86: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

76

communities which become abundant in Coondiner Creek after cessation of flow may

also act as nutrient sinks and sites of increased microbial metabolism (Kufel & Kufel,

2002). In this sense, drought may facilitate higher correlations between photosynthesis

and respiration than at times when downstream transport is a major influence on

resource dynamics, as respiration of organic matter should occur close to where it is

generated as opposed to various points along downstream flowpaths.

Conclusions

In pools of intermittent streams of the Pilbara, changes in water temperature were the

most consistent predictor of diel δ13CDIC patterns, presumably due to solubility-driven

changes in exchange of CO2 between the atmosphere and water column and low rates of

primary productivity and ecosystem respiration. However, the predominance of net

heterotrophy, together with strong links between primary productivity and respiration,

may create the conditions under which temperature-driven changes in CO2 solubility

can influence the diurnal patterns of δ13CDIC. Similarly, while there was little evidence

for short-term influences of hydrological variation on δ13CDIC cycles, extended periods

without surface flow may increase the occurrence of significant priming effects under

net heterotrophic conditions.

While differences in the δ13CDIC signatures of rivers and streams undoubtedly reflect

variation in the balance between primary productivity, respiration, and groundwater

inputs (Finlay, 2003), these results show that δ13CDIC values may vary over short time

periods not only due to high productivity but also temperature-induced changes in CO2

solubility. While the changes in δ13CDIC due to temperature cycles were relatively small

here, this further reinforces the unreliability of estimating ecosystem processes based on

single, instantaneous measures of δ13CDIC (Parker et al., 2005). The possibility that diel

changes in δ13CDIC within highly productive rivers and streams may be amplified by

temperature-solubility relationships therefore also bears investigation. In a wider

context, these results emphasize the role of the hydrological regime as a template for

ecological processes in dryland rivers and streams.

Page 87: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

77

CHAPTER FIVE: Basal carbon sources maintaining

macroinvertebrate food webs in a dryland stream do not differ among

pools of varying hydrology

Introduction

Stable isotope analysis of consumers and their possible food sources (δ13C and δ15N)

across a wide range of aquatic habitats has shown that autochthonous organic matter

sources, particularly algae, generally provide most of the C and N assimilated by

metazoan consumers (Thorp et al., 2006). The high levels of autochthonous primary

productivity supported by intermittent and ephemeral streams during periods of low or

no flow have also been shown to translate into a dominance of algal carbon in the

trophic base of metazoan production (e.g. Dekar, Magoulick & Huxel, 2009).

Exceptions to this may mostly occur when autotrophic communities are scoured by high

flows or floods, when allochthonous organic matter may provide the bulk of C and N

for secondary production until in-stream primary productivity can recover (e.g. Magaña,

2013). However, there is also increasing evidence of considerable variability in the

trophic base of food webs within intermittent and ephemeral systems during inter-flood

periods. For example, the fauna in waterholes of wide, lowland intermittent rivers in

central Australia obtain most of their energy from highly productive bands of algae in

thin littoral zones, despite high turbidity and the accumulation of large amounts of

terrestrial organic matter (Bunn et al., 2003). Conversely, the fauna of intermittent

streams of more temperate southeast Australia are thought to obtain the majority of their

carbon from terrestrial particulate organic matter (POM) and associated biofilms,

despite autochthonous sources of higher lability becoming more prevalent as habitats

dried (Reid et al., 2008). Macroinvertebrates in intermittent floodplain rivers of the wet-

dry tropics of northern Australia may obtain a majority of their energy from algae, but

also make extensive use of detritus from riparian vegetation during the dry season

(Leigh et al., 2010; Hunt et al., 2012). This variability in the trophic base of metazoan

production may be a response to high overall hydrological variability, with the ability to

vary diet in response to changing environmental conditions a key factor in species

persistence within intermittent or ephemeral systems (Leigh et al., 2010; Blanchette et

al., 2013). The mechanisms by which this variation occurs are not, however, currently

well understood.

Page 88: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

78

In streams with regular and sustained periods of drying and contraction, such as those

that occur in dryland regions, fauna appear to have generally evolved life history and

behavioural characteristics which provide a strong resistance and resilience to drying

(Lake, 2003). After cessation of surface flow and fragmentation of wetted area into

isolated and drying habitats, macroinvertebrates adapted to drying tend to concentrate

into dense populations within pools (e.g. Boulton & Lake, 1992; Stanley et al., 1994).

Drying should therefore theoretically lead to higher competition for food resources

unless autochthonous primary productivity or allochthonous organic matter inputs

increase as well, or if there are substantial food resources present before drying occurs.

Terrestrial litterfall inputs tend to accumulate in intermittent streams in the absence of

downstream transport (Acuña et al., 2004; Reid et al., 2008). However, autotrophic

communities within pools may decline as they are stranded and desiccate above drying

fronts, or become less productive as community composition shifts to more drying-

tolerant species (Stanley et al., 1997). Consequently, a shift in the source of carbon

inputs as streams and then individual pools contract is less likely to limit organisms that

rely on an allochthonous trophic base (Acuña et al., 2004) while macroinvertebrates

competing for diminishing autochthonous resources may adopt an increasingly

generalist use of food resources (e.g. Blanchette et al., 2013). Previous studies of the

seasonal tropics in Australia have also demonstrated that biophysical variables and

macroinvertebrate assemblages are highly spatially and temporally variable, and that

local factors were as important as the regional setting in determining assemblage

composition and temporal trajectories (Blanchette et al., 2013; Warfe et al., 2013).

Similarly, macroinvertebrates increased in density during a drought in an intermittent

forested Mediterranean stream, but with markedly different patterns between cobbles

and leaves compared with sandy substrates (Acuña et al., 2005). Given the above, we

might expect that the degree of drying and contraction within aquatic habitats in

intermittent streams is directly relatable to shifts in the trophic base of

macroinvertebrate production over the same period.

In this study, I investigated the structure and trophic base of macroinvertebrate food

webs in pools along an intermittent stream in the Pilbara region of sub-tropical

northwest Australia. I used stable isotope analysis (δ13C and δ15N) to: i) identify the

trophic position of macroinvertebrates in relation to basal C and N sources, and ii) to

compare food web reliance on different basal sources among pools of varying

hydrology. Pools sampled had previously been characterised as either disconnected or

Page 89: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

79

maintained by shallow alluvial groundwater (see Chapter 2; Fellman et al., 2011a) and

were sampled twice; once near the end of a dry inter-flood period and then again in the

early-mid dry season following recharge and flooding after a major summer cyclone. I

predicted that macroinvertebrate densities would be higher in more highly evaporated

pools as well as later in the dry season. I thus also expected that macroinvertebrates

would derive most of their carbon from algae in the early-mid dry season but that the

trophic base of macroinvertebrate production would be derived equally from

allochthonous and autochthonous sources in the late dry season, especially in highly

evaporated and drying pools that had become disconnected from AW.

Methods

Study sites and sampling

Sampling occurred at Coondiner Creek within a chain of consecutive pools (1 – 7)

during September 2011 and June 2012 (sites and climatic conditions are as described in

previous Chapters). Sampling was therefore spread across four pools with persistent

alluvial water (AW) connectivity (pools 1, 2, 5 and 6) and three others that were

disconnected from AW (pools 3, 4 and 7) and subsequently evaporating. The September

2011 sampling period occurred approx. seven months after cessation of surface flow

resulting from heavy rainfall in Jan 2011 (the late dry season) while the June 2012

sampling was four months after flooding from Cyclone Heidi (early-mid dry season;

Fig. 5.1; see also earlier Chapters). Pools were therefore relatively larger in the early-

mid dry season when compared to the late dry season (Table 5.1).

At each pool, samples were collected at two or three transects (~15 m long) aligned at

right angles to the channel to ensure representative sampling of the entire pool.

Transects were placed consistently to capture the heads of pools (starting 5 m from the

most upstream point) and another section about mid-way along the length of the pool.

Where space permitted, a third transect was placed at the end of the pool (ending 5 m

from the most downstream point). Due to contraction following the cessation of surface

flow and subsequent disconnection from AW, the wetted length only allowed one 15 m

transect at Pool 4 in September 2011. Pool 7 was not sampled in September 2011 owing

to difficulties with site access due to time constraints.

Page 90: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

80

Figure 5.1: Monthly cumulative rainfall at Coondiner Creek from January 2011 until

October 2012. • indicates periods when macroinvertebrate sampling was undertaken.

Samples of the dominant primary producers (both terrestrial and aquatic) were collected

from within the pools and surrounding vegetation. Whole conditioned (i.e., abscised,

wet and partly decomposed) leaves of Eucalyptus and Melaleuca spp. (where present)

were collected from the beds of pools as the main component of terrestrial detritus. Live

specimens of aquatic macrophytes were collected from the littoral zone of pools.

Filamentous algae (Spirogyra spp.) were also separated and collected from sediments

and aquatic macrophytes in the littoral zone, while benthic biofilms were collected from

rocky substrate by scrubbing with a sterile sponge (Speci-Sponge; VWR International).

Samples of fine particulate organic matter (FPOM) analysis were collected from each

transect by filtering the maximum possible amount of water through a 0.7 μm pre-

combusted glass fibre or quartz filters before clogging occurred (approx. 0.5 – 2.5 L).

All plant, algae, biofilm and FPOM samples were stored at 4 ⁰C until analysis.

Macroinvertebrates were sampled by sweeping the length of each 15 m transect with a

200 μm kick-net, taking care to include all possible substrates and habitats within the

Page 91: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

81

Table 5.1: Maximum size and depth of pools within Coondiner Creek during the late

dry season (Sept 2011) and early-mid dry season (June 2012), with respect to alluvial

water (AW) connectivity. Pool 7 was not sampled in September 2011.

Pool

number

AW

connectivity

Late dry season

(September 2011)

Early-mid dry season

(June 2012)

Length

(m)

Width (m) Depth

(m)

Length

(m)

Width (m) Depth

(m)

1 Connected 170 23 1 460 31 1.5

2 Connected 270 25 2 280 27 2.5

3 Isolated 60 8 1.4 63 9 2

4 Isolated 45 16 0.5 98 26 2

5 Connected 90 8 0.8 165 10 1.5

6 Connected 60 22 2.5 78 31 3

7 Isolated - - - 64 21 2.5

transect. Samples were sieved in the field to remove larger detritus, with larger

invertebrates handpicked from the resulting debris and returned to the samples. All

invertebrate samples were frozen on the day of collection and stored at -20 ⁰C until

analysis.

Isotopic analyses

Filamentous algae and plant samples were rinsed in distilled water, oven-dried at 60 ⁰C

for 3-4 days and then ground to fine powder in a ball mill grinder. Speci-sponges were

rinsed with distilled water and the resulting solution then oven-dried at 60 ⁰C for 3-4

days. FPOM samples were scraped from filters and then oven-dried at 60 ⁰C for 3-4

days.

Macroinvertebrates were hand-sorted under x10 magnification. All invertebrates were

counted and identified to genera for most taxonomic groups, and to family for mites

(Acarina). Only a few caddis fly larval cases (Trichoptera: possibly Leptoceridae) were

found and did not contain individuals; these were not included in final counts. Only 13

specimens (out of c. 9800 found) could not be identified to family and were not

Page 92: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

82

included in final counts. Macroinvertebrate samples were bulked by genera (or family

for Acarina) and then oven-dried at 60 ⁰C for 3-4 days.

Primary source and macroinvertebrate samples were analysed for total C content, total

N content, δ13C and δ15N using a continuous flow system consisting of a Delta V Plus

mass spectrometer connected via Conflo IV with a Thermo Flush 1112 (Thermo-

Finnigan, Germany). Multi-point normalization was used to reduce raw values to

international scale (Skrzypek, 2013) using reference materials provided by the

International Atomic Energy Agency (Coplen et al., 2006): for δ13C, NBS22, USGS24,

NBS19 and LSVEC; and for δ15N, N1, N2, USGS32 and laboratory standards. δ13C

values are reported in per mil (‰) after normalization to VPDB while δ15N values are

reported in per mil (‰) after normalization to Air. Analytical precision was 0.1‰ for

both δ13C and δ15N.

Data for δ13C and δ15N were unable to be fully obtained for several samples. Gastropod

biomass was not high enough (< 1 mg) to obtain readings without including shell

material; I determined this would be an unreliable indicator of their food source, as the

δ13C signature of mollusc shells is controlled largely by the aqueous carbon species

rather than diet (Fritz & Poplawski, 1974) and I had a lack of detailed information on

the δ13C signature of dissolved inorganic carbon within pools. Only one specimen of

Crustacea was found, and was too small to analyse. In addition, the total content of C

(approx. 1 %) and N (approx. 0.1 %) of FPOM collected were, unfortunately,

insufficient for reliable stable isotope analyses (see Chapter 3).

Data analysis

Similarities among the macroinvertebrate communities of different pools were

identified by non-metric multidimensional scaling (nMDS) using Bray-Curtis

dissimilarities (Clarke, 1993). Genera (or families, in the case of Acarina) found in less

than 5 % of the samples were not included in the nMDS analysis to avoid undue

emphasis on rare species. However, I did create an exception to this rule for adult

predaceous diving beetles (Coleoptera: Dytiscidae), as the individuals found had much

greater mass than the more common invertebrates and are thus likely to play a key

functional role in macroinvertebrate assemblages (Cao, Larsen & Thorne, 2001; Poos

& Jackson, 2012). Consequently, all adult diving beetles were collectively included in

Page 93: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

83

the analysis under the family Dytiscidae. In order to determine similarities among pools

based on abundance, data were log10(x+1) transformed to avoid bias from extremely

high densities of Acarina within Pool 4 in September 2011 (when highly evaporated).

Similarities among pools based on species richness were based on presence/absence

data. Data were also assessed for evidence of spatial autocorrelations in

macroinvertebrate assemblage composition, by comparing resemblance matrices of

macroinvertebrate community composition with the resemblance matrices of distances

between pools (i.e. the Euclidean distance between spatial coordinates) using Mantel

coefficient tests with 9999 permutations (Clarke, 1993).

Spearman rank correlations were used to assess the influence of different

macroinvertebrate orders on the ordination. Differences in the abundance or

composition of macroinvertebrate communities between sampling periods were

assessed using one-way ANOSIM analysis, while differences between AW-connected

and isolated pools were assessed using two-way nested ANOSIM (where hydrology was

nested within the sampling date). I also assessed changes in the density and species

composition of macroinvertebrate communities in relation to changes in hydrology by

calculating the linear least-squares regression between differences in δ18O values

between the late dry and early-mid dry season samplings (Δ18Ot1-t2) and the Euclidean

distance between paired pools in 2D nMDS space. Multivariate analyses were

conducted using Primer 6 (PRIMER-E ltd., UK) and univariate analyses were

conducted using MATLAB v7.14 (MathWorks, Natick, MA, USA).

As primary producer and macroinvertebrates were analysed as bulk tissue samples for

δ13C, I applied correction factors to the data to adjust for variable lipid content (Post et

al., 2007). Primary producer δ13C values were adjusted by using the equations defined

by Post et al. (2007), where the difference between δ13C and δ13C(corrected) was -3.02 +

0.09 x % C for samples where % C was ≤ 40 % and -5.83 + 0.14 x % C for samples

where % C was > 40 %. Macroinvertebrate δ13C values were adjusted by using the

equations defined by Logan et al. (2008), where the difference between δ13C and

δ13C(corrected) was -2.056 + 1.907 x ln (C:N). The corrected δ13C and δ15N values of

primary producers and macroinvertebrate orders were then compared across pool AW

connectivity (AW-connected versus isolated) and sampling periods (late dry season and

early-mid dry season) by single factor ANOVA, with a confidence level of 95 %.

Page 94: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

84

I characterised the trophic diversity of macroinvertebrate food webs between the late

dry and early-mid dry season by using the isotopic diversity indices proposed by

Layman et al. (2007): the total range in δ15N (NR) and δ13C (CR); the total convex hull

area, or area of the smallest possible convex polygon enclosing all paired δ13C and δ15N

points in 2D space (TA); the mean distance to centroid for all samples (CD); and the

mean distance to nearest neighbour for each order (NND) with its standard deviation

(SDNND).

I used the Stable Isotope Analysis in R (SIAR; Parnell et al., 2010) package for R 3.1.3

(R Project for Statistical Computing, Austria) to produce isotopic mixing models via

Bayesian inference for estimating the contributions of different food sources to

macroinvertebrate consumers over each sampling period. The trophic enrichment factor

(TEF) for δ13C does not generally vary widely across different organisms and habitats

(Peterson & Fry, 1987) and was therefore set at 0.4 ± 1.3 ‰ as per Post (2002b) for all

organisms. The TEF for δ15N is considered more variable across both environments and

organisms. I therefore used Bunn, Leigh and Jardine’s (2013) characterisation of river

and stream macroinvertebrates in Australia and New Guinea to define the TEF for δ15N

as 0.6 ± 1.7 ‰ for primary consumers, and 1.8 ± 1.7 ‰ for secondary consumers. Each

model was the result of 500,000 iterations, with the first 50,000 iterations discarded, and

is presented as 95 % confidence intervals of the estimated population mean.

Results

Macroinvertebrate community composition

Macroinvertebrates found within Coondiner Creek were mainly adult (Hemiptera,

Coleoptera) and larval insects (Ephemeroptera, Diptera, Odonata), although water mites

(Acarina) and snails (Gastropoda) were also commonly found (Table 5.2). Acarina were

the most abundant individuals across all samples (592 ± 491), followed by

Ephemeroptera (59 ± 33) and Hemiptera (30 ± 6). All other orders generally had fewer

than ten individuals per pool. Differences in overall community composition among

pools were therefore strongly driven by the abundances of Acarina, Ephemeroptera and

Hemiptera, and to a lesser extent by abundance of Gastropoda, Odonata and Diptera

(Fig. 5.2a). Differences in species composition (presence/absence) among pools were

Page 95: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

85

Table 5.2: Families and descriptions of macroinvertebrate orders found within pools of

Coondiner Creek in September 2011 and June 2012, with predicted feeding strategies.

Order Family Description Feeding strategy

Ephemeroptera Baetidae Mayfly nymphs Gathering collector

Diptera Chironimidae Midge nymphs Gathering collector

Gastropoda Planorbidae Semi-aquatic snails Grazer

Lymnaeidae Semi-aquatic snails Grazer

Crustacea Triopsidae Shield shrimp Grazer

Hemiptera Corixidae Aquatic adult bugs Piercing herbivore

Notonectidae Aquatic adult bugs Predator

Pleidae Aquatic adult bugs Predator

Nepidae Semi-aquatic adult

bugs Predator

Mesoveliidae Semi-aquatic adult

bugs Predator

Coleoptera Hydraenidae Semi-aquatic adult

beetles Grazer

Hydrochidae Semi-aquatic adult

beetles Shredder

Dytiscidae Adult diving beetles Predator

Odonata Lestidae Dragonfly nymphs Predator

Libellulidae Dragonfly nymphs Predator

Trichoptera Polycentropodidae Caddisfly nymphs Predator

Acarina Hygrobatidae Aquatic mites Predator/parasitic

Limnesiidae Aquatic mites Predator/parasitic

Unionicolidae Aquatic mites Predator/parasitic

Pionidae Aquatic mites Predator/parasitic

Hydrachnidae Aquatic mites Predator/parasitic

Oribatidae Terrestrial mites Predator/parasitic

driven mainly by the presence or absence of Ephemeroptera, Gastropoda, Odonata,

Coleoptera and Diptera (Fig. 5.2b).

ANOSIM analysis identified that differences in macroinvertebrate abundance among

pools were not clearly associated with AW connectivity in either the late or early-mid

Page 96: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

86

Figure 5.2: Similarities among pool macroinvertebrate communities in Coondiner

Creek during the late dry season (Sept 2011) and early-mid dry season (June 2012)

shown by nMDS ordinations of a) average abundance within pools, and b)

presence/absence data for each identified genus. Data labels indicate pool number.

Vectors show Spearman rank correlations between the ordinations and

macroinvertebrate orders with correlations ≥ 0.3.

Page 97: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

87

Figure 5.3: Similarities in macroinvertebrate communities in the late dry season (Sept

2011) and early-mid dry season (June 2012) of the same pool (Euclidean distance

between paired pools) based on nMDS analysis of a) average abundance of

macroinvertebrates, and b) macroinvertebrate community species composition, in

relation to the difference in pool water δ18O values between sampling times (Δ18Ot1-t2).

dry season (P = 0.20). However, pools with greater enrichment in δ18O values between

the late and early-mid dry season samplings had the greatest differences in overall

macroinvertebrate abundance, suggesting that isolation from AW and subsequent

evaporation was directly related to higher macroinvertebrate abundance (Fig. 5.3a). In

contrast, ANOSIM analysis indicated that species composition was not driven by AW

connectivity in either sampling period (P = 0.35), and the distance between paired pools

Page 98: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

88

in nMDS space and changes in δ18O values between the late and early-mid dry season

were not significantly different (Fig. 5.3b).

Macroinvertebrates were overall more than 13 times more abundant in the late dry

season compared to the early-mid dry season (P = 0.002). The largest differences were

in Acarina (74.5 x higher), Gastropoda (5.3 x higher) and Hemiptera (2.13 x higher) in

the late dry season, especially in highly evaporated pools (Fig. 5.2a). Ephemeroptera

(1.2 x higher) and Coleoptera (0.6 x higher) were also more abundant in the late dry

season (Fig. 5.2a). In contrast, Odonata (7.4 x lower) and Diptera (1.3 x lower) were

less abundant in the late dry season. While macroinvertebrate species composition did

not differ significantly between the two sampling periods (P = 0.11), pools 1 and 2 were

more likely to have Odonata, Ephemeroptera, Diptera, and Gastropoda, while pools 4

and 5 were less likely to contain Coleoptera (Fig 5.2b). However, there was no evidence

that macroinvertebrate species composition was more similar in pools that were closer

to each other, whether in the early-mid (P = 0.40) or late dry season (P = 0.18).

Food web size and trophic base

The sampled basal sources for food webs had δ13C values ranging from -40.9 to -23.1

‰ and δ15N values ranging from -0.1 to 10.2 ‰, but with little clear differentiation

between allochthonous and autochthonous primary sources (Fig. 5.4). Filamentous

algae had the most negative δ13C (-40.8 ‰) but also the most positive δ15N values (10.2

‰), perhaps indicating that epiphytic cyanobacteria were also present in the samples.

Biofilms, in contrast, had the highest δ13C (-23.1 ‰) and lowest δ15N values (-0.1 ‰;

Fig. 5.4). "Conditioned" or partly decomposed Eucalyptus and Melaleuca leaves (i.e.,

terrestrial detritus) had the least variability of any OM source (δ13C -29.3 to -28.6 ‰;

δ15N 0.0 to 1.8 ‰) but in general had similar δ13C and δ15N values to biofilms. There

was also substantial overlap in the δ13C and δ15N values of biofilms with those of

charophytes (δ13C -34.2 to -30.8 ‰; δ15N 1.8 to 5.5 ‰) and Potamogeton tricarinatus

(δ13C -30.4 to -27.9 ‰; δ15N 0.6 to 3.0 ‰; Fig. 5.4).

Macroinvertebrates had δ13C values ranging from -42.2 to -22.9 ‰ and δ15N values

ranging from 1.1 to 9.0 ‰ (Fig. 5.4). On average, Coleoptera had the highest δ13C

values (-29.2 ‰) and Ephemeroptera the lowest (-36.8 ‰), while the highest δ15N

values (11.2 ‰) were observed in the Trichoptera (although only one Trichopteran dual

Page 99: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

89

Figure 5.4: Dual isotope plot of macroinvertebrates and basal sources within Coondiner

Creek in a) the late dry season (Sept 2011) and b) the early-mid dry season (June 2012).

Macroinvertebrates are shown as separate samples, whereas basal sources are shown as

averages with error bars displaying the total range of lipid-corrected values.

Page 100: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

90

isotope sample was obtained; the highest average δ15N value over multiple samples was

7.2 ‰ in the Acarina). The lowest δ15N values (2.4 ‰) were observed in the Diptera.

The δ13C and δ15N of primary consumers (Ephemeroptera, Diptera) appeared closest to

those of charophytes (Fig. 5.4a) and filamentous algae (Fig. 5.4b). Mixed primary and

secondary consumers (Hemiptera, Coleoptera) and solely parasitic or predatory

consumers (Acarina) had higher δ13C and δ15N values relative to basal sources and

primary consumers, consistent with their expected trophic positions.

The δ13C and δ15N values of basal sources did not differ significantly between AW-

connected and isolated pools for either sampling period (P = 0.4 to 0.7). Patterns in δ13C

and δ15N of macroinvertebrates were also quite similar across AW-connected and

isolated pools (P = 0.3 to 0.9), but with two exceptions. Ephemeroptera had

significantly lower δ15N values (P < 0.001) in AW-connected pools (1.9 to 3.4 ‰) than

in evaporating pools (5.4 to 6.5 ‰) in the early-mid dry season, while Acarina had

lower δ13C values (P < 0.001) in AW-connected pools (-29.9 to -32.5 ‰) than in

evaporating pools (-25.8 to -29.1 ‰) during the late dry season.

The δ13C and δ15N values of basal sources also did not differ significantly between the

early-mid dry season and the late dry season (P = 0.1 to 0.8). However, the

Ephemeroptera (P = 0.02), Hemiptera (P = 0.01), and Diptera had lower δ13C values on

average in the early-mid dry season than in the late dry season, and were closer to the

δ13C signatures of filamentous algae (Fig. 5.4b). δ13C and δ15N values for the

Coleoptera (P = 0.87) and Acarina (P = 0.50) were similar across both sampling periods.

Macroinvertebrate communities in the early-mid dry season also included members of

the Odonata, which had isotopic signatures close to those of the Diptera and the

Ephemeroptera (δ13C -38.7 to -33.1 ‰; δ15N 2.2 to 4.4 ‰; Fig. 5.4b). Overall,

macroinvertebrate trophic structure in the early-mid dry season was characterised by

more trophic levels, less overlap in trophic niche and higher diversity in basal resources

compared with the late dry season (Table 5.3).

However, the δ13C and δ15N values of the SIAR mixing models revealed that the trophic

base of macroinvertebrate food webs did not differ between sampling periods.

Furthermore, there was also little to no difference between the contributions of primary

sources for any order (Table 5.4). Only three results indicated contributions of over 20

% of any basal source to any macroinvertebrate order: in the early-mid dry season,

filamentous algae for the Ephemeroptera (20.7 %) and predaceous Hemiptera (26.6 %),

Page 101: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

91

Table 5.3: Macroinvertebrate trophic structure as defined by isotopic diversity indices

(Layman et al., 2007) derived from consumer δ15N and δ13C values at Coondiner Creek

in September 2011 (late dry season) and June 2012 (early-mid dry season).

Ephemeroptera and Diptera in the early-mid dry season were sometimes too low to have

been derived solely from their measured potential food sources, suggesting the

possibility of an (unmeasured) additional basal source (Fig. 5.4b).

Metric Interpretation Late dry season

(Sept 2011)

Early-mid dry

season (June 2012)

δ15N range (NR) higher = more trophic

levels, more trophic

diversity

7.4 9.3

δ13C range (CR) higher = more basal

resource diversity

11.8 18.0

Total area (TA) higher = more trophic

diversity overall

6.9 30.6

Mean distance to

centroid (CD)

higher = more trophic

diversity

2.8 4.3

Mean nearest

neighbour distance

(NND)

higher = species have

less overlapping trophic

niches

1.2 1.8

Standard deviation of

NND (SDNND)

higher = species have

less overlapping trophic

niches

0.5 1.4

and emergent macrophytes for the Acarina (27.5 %) in the late dry season. As

mentioned above, the low δ13C and δ15N values of the Ephemeroptera and Diptera

suggested that an unmeasured food source was also contributing to macroinvertebrate

diets in the early- mid dry season. Consequently, I repeated the SIAR mixing models

with the inclusion of an additional 13C and 15N-depleted food source, for which I set

δ13C at -66 ± 4.4 ‰ and δ15N at -13 ± 2.6 ‰; this range of isotopic values were based

on conservative values for methane oxidising bacteria across the literature and the

possibility that this potential food source was utilizing methane oxidation as a source of

carbon (Jones & Grey, 2011). However, the results of the SIAR models indicated that

Page 102: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

92

Table 5.4: Results of stable isotope mixing models (SIAR) showing estimated percentage contributions of each primary source to

macroinvertebrate consumers (as 95 % confidence intervals of the population mean) in pools of Coondiner Creek in the late dry season

(Sept 2011) and early-mid dry season (June 2012). There was insufficient material of Gastropoda and Crustacea for stable isotope analyses.

Values in bold indicate sources with the highest estimated mean contribution for each order at each sampling, as determined by SIAR.

Note: – indicates consumers were not found in the specified sampling period.

Order

(trophic level)

Terrestrial detritus Emergent

macrophytes

Potamogeton

tricarinatus

Charophytes Filamentous algae Biofilm

Sept June Sept June Sept June Sept June Sept June Sept June

Ephemeroptera

(1⁰) 0-34 0-31 0-31 0-33 0-31 0-31 0-36 0-34 0-32 1-37 0-34 0-30

Diptera (1⁰) 0-33 0-34 0-33 0-30 0-33 0-31 0-33 0-35 0-33 0-30 0-33 0-33

Hemiptera (1⁰) 0-32 0-32 0-33 0-35 0-33 0-33 0-33 0-31 0-33 0-35 0-33 0-33

Hemiptera (2⁰) 1-33 0-28 0-31 0-36 0-32 0-31 0-33 0-33 0-26 4-49 1-31 0-28

Coleoptera (1⁰) 0-33 0-32 0-34 0-35 0-34 0-35 0-31 0-30 0-32 0-33 0-31 0-33

Coleoptera (2⁰) 0-33 - 0-32 - 0-33 - 0-32 - 0-32 - 0-33 -

Odonata (2⁰) - 0-34 - 0-32 - 0-32 - 0-34 - 0-32 - 0-34

Trichoptera

(2⁰) - 0-33 - 0-33 - 0-33 - 0-33 - 0-32 - 0-33

Acarina (2⁰) 0-23 0-31 2-52 0-35 1-47 0-35 0-19 0-30 1-38 0-33 0-27 0-32

Page 103: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

93

no single food source was likely to provide more than 25 % of the diet of any organism,

with only low contributions (means of 8.5 to 13.9 ‰) from the potential methanotrophic

food source to either Ephemeroptera or Diptera.

Discussion

Consistent with my original hypotheses, I found that macroinvertebrate abundance was

greatest in more evaporated or contracting pools, i.e., in pools isolated from AW and

with increasing time since recharge. Previous, widespread sampling of the Pilbara has

found little seasonality in occurrence of individual macroinvertebrate taxa but that

community composition changes with habitat permanence (Pinder et al., 2010).

Macroinvertebrates that increase in abundance when drying and contraction of habitat

occurs are usually those with low dispersal abilities, such as flightless and fully aquatic

taxa (Lake, 2003). In Coondiner Creek, this was reflected in the higher abundances of

the fully aquatic or semi-aquatic Gastropoda, Acarina, and Hemiptera in the late dry

season. In contrast, abundances of Odonata nymphs appeared to be negatively affected

by evaporation despite the high dispersal ability of the adults, which may reflect these

large predators’ need for bigger habitats (Welborn, Skelly & Werner, 1996). Otherwise,

the abundance of taxa with highly mobile adults or the ability to survive on land were

more variable amongst pools rather than between sampling periods. These findings

suggest that, as in other intermittent streams, the ability to disperse to more favourable

habitats is the main control on changes in macroinvertebrate abundance as pools

contract. Consequently, any competition for food resources may also increase directly

along with evaporation and contraction of wetted area.

Macroinvertebrate food webs were larger and more diverse in the early-mid dry season

when compared with the late dry season on Coondiner Creek. While the overall number

of taxa (genera plus Acarina families) identified here (22) was less than that found by

Pinder et al. (2010) in a previous sampling of the catchment (52), this study does

include all of the more common taxa found in that earlier sampling. Larger ecosystems

may support larger food webs through increased resource availability, increased trophic

diversity, and provision of space for larger predators (Post, 2002a). Macroinvertebrates

may therefore also be responding to changes in habitat quality. Higher densities of

aquatic macrophytes, especially structurally complex ones like the charophytes that

dominate Coondiner Creek, may support larger and more diverse food webs by

Page 104: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

94

providing substrata for epiphytes and refuge from predators (Warfe, Barmuta &

Wotherspoon, 2008; Pettit et al., 2011; Ziegler et al., 2015). As pools contract, habitat

quality as well as quantity may decrease as macrophyte biomass is lost along drying

fronts and autotrophic community composition shifts towards biofilms or algae (Stanley

et al., 1997). The Layman metrics of food web size here suggest that larger pools (e.g.,

pools in the early-mid dry season) and therefore larger food webs may also lead to more

diverse resource use by macroinvertebrates in Coondiner Creek.

However, the SIAR models contrast directly with the Layman metrics in indicating that

there was little change in resource use by either primary or secondary consumers

between sampling periods, with no preferentially assimilated food source for any

organism. In particular, basal resource use was very similar across the primary

consumers (Ephemeroptera, Diptera, Hemiptera and Coleoptera), which is consistent

with some recent studies in the Australian seasonal tropics and also in more temperate

systems. For example, in the intermittent rivers of the Burdekin catchment of northeast

Australia baetid mayflies (Ephemeroptera) have exhibited largely generalist feeding

strategies during the dry season (Blanchette et al., 2013). Similarly, in intermittent

streams of the Goulburn and Loddon River catchments of southeast Australia, where

allochthonous organic matter provides most of the trophic base for food webs, Diptera

may also obtain a substantial proportion of their energy from biofilms (Reid et al.,

2008). However, studies elsewhere have shown that different primary consumers,

particularly filter feeders and decapods, may be more likely to specialise on algal food

sources. For example, snails, mussels, shrimp and crayfish made up the majority of

primary consumers in food webs of central Australian waterholes supported by algal

carbon (Bunn et al., 2003). Decapods and caddisfly nymphs were among the primary

consumers supported by algal carbon in lentic sections of intermittent rivers in the

Gregory and Flinders River systems in northeast Australia (Leigh et al., 2010). These

disparate observations among different intermittent systems suggest that the generalist

trophic base of macroinvertebrate food webs in Coondiner Creek may be reflective of

the primary consumers present there, rather than the result of high competition for

autochthonous resources.

We might expect organisms adapted to exploiting ephemeral habitats to exhibit

generalist feeding strategies as an inherited adaptation, rather than adapting their

behaviour in response to changing hydrological conditions. The generalist assimilation

of C and N by primary consumers observed at Coondiner Creek likely reflects their high

Page 105: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

95

dispersal ability and broad adaptations to a relatively dry environment. All of the

primary consumers either had fully terrestrial and flying adults (Ephemeroptera,

Diptera) or were semi-aquatic adults with the ability to fly (Hemiptera) or survive on

land (Coleoptera). Flying adults are able to exploit more ephemeral aquatic habitats

when they become available; thus it may be advantageous for their aquatic offspring to

have adapted to utilise whatever resources are available in case their habitat dries and

competition for resources becomes higher (Kassen, 2002). Specialist feeding strategies

in macroinvertebrates may therefore also become more prevalent when food sources are

more predictable, much like in the fish species of intermittent rivers (Bunn et al., 2006a;

Pusey et al., 2010).

Evidence for a methanotrophic food source

The δ13C and δ15N values of Ephemeroptera and Diptera were too negative at times to

have been derived solely from the measured primary producers, suggesting that they

may be deriving a substantial portion of their carbon from an additional food source.

There is growing evidence that methane-derived carbon provides a significant fraction

of the trophic base of freshwater food webs given that biogenic methane is typically

much more highly 13C-depleted (-60 to -80 ‰) than other carbon sources (Jones &

Grey, 2011). Diptera are usually the most reliable indicator that methane-oxidising

bacteria (MOB) are contributing to macroinvertebrate food webs, especially in taxa that

form burrows in the sediment along which anoxic-oxic boundaries can develop and

encourage MOB to grow (Deines et al., 2007b). In pools of central Australian

intermittent rivers, Diptera may utilise MOB even when the majority of secondary

production is supported by algal carbon (Bunn et al., 2003). Diptera utilizing a

methanotrophic food source can be much more highly depleted in 13C (as low as -72 ‰)

than observed here, although these extremely low values are usually only seen in

profundal lake taxa (Jones et al., 2008b). Conversely, direct evidence that grazers such

as Ephemeroptera utilise MOB as a food source is rare; but may occur when methane is

produced under anaerobic decomposition of detritus in stagnant water, or when stream

water is saturated with groundwater-derived methane which can fuel biofilms (Jones &

Grey, 2011).

In a study of three shallow water bodies in the Murray-Darling catchment of southeast

Australia, Bunn & Boon (1993) observed more depleted δ13C values in

Page 106: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

96

macroinvertebrates, including those for baetid mayflies (approx. -35 ‰), compared to

those of epiphytes, detritus, or macrophytes. Under lentic conditions in these billabongs,

decaying aquatic macrophytes may be decomposed under anaerobic conditions in the

sediment, leading to large fluxes of methane which could be utilised by methanotrophic

bacteria (Boon, 1991; Bunn & Boon, 1993). Shallow, net heterotrophic, and lentic

water bodies may also develop small-scale and intermittent development of anoxia at

the sediment-water interface which can stimulate biological methane oxidation (Jones &

Grey, 2011). This may lead to opportunistic rather than specialist use of MOB as a food

source. Methane-derived carbon could then enter macroinvertebrate food webs from

various sources, such as biofilms on the cases of caddisfly larvae or through crustaceous

zooplankton, where the carbon signatures of consumers are relatively higher (approx. -

40 to -35 ‰) compared to burrow-forming Chironimidae (Jones et al., 1999; Trimmer

et al., 2009). Additional analyses based on other isotopes such as 34S, which can be

indicative of anaerobic conditions due to its sensitivity to sulphate reduction (Finlay &

Lajtha, 2007), may therefore help resolve the complexity of when methane enters

intermittent stream food webs.

With regard to other alternate carbon sources, it is unlikely that carbon in the seston is a

significantly 13C-depleted food source for macroinvertebrates in Coondiner Creek, given

that fine particulates from the water column had very low C and N content. Previous

measurements of FPOM in the seston of Coondiner Creek pools has also shown that it is

relatively enriched (approx. – 30.2 ‰; see Appendix A) compared to the low values

seen in Ephemeroptera and Diptera here. Alternatively, dissolved organic matter (DOM)

may be providing a relatively 13C-depleted (approx. -40 ‰; see Appendix A) source of

carbon for bacteria. DOM may represent the greatest flux of carbon in Coondiner Creek

if it is net heterotrophic (but see Chapter 4), like other northern Australian rivers

(Townsend et al., 2011). In my study of Coondiner Creek biofilm collection was limited

to epilithic and epiphytic sources and may have not encompassed methane-oxidising

bacteria on the sediment that could contribute to the diet of Diptera and Ephemeroptera.

Surface water biofilms can also contain a mixture of algae in addition to heterotrophic

and chemoautotrophic bacteria, and their δ13C signatures may therefore be the result of

complex mixing of many different carbon production and utilisation processes. For

example, bacteria in sediment biofilms can strongly reflect the 13C signature of diatom

exudates or extracellular polysaccharides, suggesting a closely coupled algae-bacteria

system (Miyatake et al., 2014), but the specialization of microbes on other forms of

Page 107: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

97

carbon also increases with depth and the development of localised oxic and anoxic

zones (Freixa et al., 2015). These examples illustrate that the role of microbial

community composition in organic matter cycling within biofilms is only just beginning

to be understood (e.g. Logue et al., 2015), especially with respect to how community

composition may differ over short spatial or temporal scales. In particular, the diversity

of methanogenic archaea may show considerable heterogeneity in sediment cores but

little correlation with depth (Brablcová et al., 2015). It therefore remains a possibility

that MOB do contribute to macroinvertebrate food webs in Coondiner Creek, although

if they do then it is likely as part of a generalist diet rather than specialised feeding by

particular taxa.

Conclusions

This study has shown that some of the inherent variability observed in intermittent

rivers of northern Australia (Leigh et al., 2010; Blanchette et al., 2013) might be

explained by the hydrologic status of individual pools. In pools of Coondiner Creek,

evaporation and contraction of wetted area were associated with higher densities of

macroinvertebrates, which coincided with lower overall trophic diversity. However, the

assimilation of C and N sources by macroinvertebrates suggests that primary consumers

within Coondiner Creek are adapted to generalist feeding strategies regardless of the

degree to which pools have evaporated. These feeding strategies likely include some

assimilation of C and N from methanotrophic microbial communities that form during

intermittent periods of anoxia at the sediment, although there are probably no

macroinvertebrates within Coondiner Creek that specialise on methane-derived carbon.

Overall this study demonstrates that evaporation and contraction in intermittent streams

is a primary control on the densities of fauna within them but that changes in abundance

do not necessarily lead to increased competition for autochthonous carbon. I suggest

that long-term hydrological characteristics of intermittent or ephemeral rivers and

streams may instead restrict the species available to colonise them and thus play a large

role in the determination of trophic structure.

Page 108: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

98

CHAPTER SIX: General discussion: organic matter dynamics and

ecosystem functioning of intermittent streams in the semi-arid Pilbara

The findings of this thesis expand the knowledge of the different sources and

significance of organic matter to intermittent dryland streams and the role they play in

sustaining the productivity and biodiversity of aquatic ecosystems in the Pilbara. More

generally, this research also provides new understanding of the ways in which the

varying environmental conditions within and among intermittent streams may interact

with their flow regimes to create distinct ecological patterns. This research thus

contributes to the growing recognition of the significance of intermittent flows to

understanding biogeochemical processes across rivers and streams around the world. In

particular, this research emphasises the importance of understanding the connectivity of

streams below the surface and thus the role alluvial water plays in sustaining pools

through prolonged drought periods.

The principle objectives of this research were to investigate i) how the source and

availability of organic matter changes as pools evaporate and contract, ii) how microbial

and metazoan utilisation of organic matter changes as pools evaporate and contract, and

(iii) what influence AW connectivity has on these relationships. In this respect, organic

matter within the focus study catchment, Coondiner Creek, clearly concentrates as pools

evaporate and dry but without any substantial and lasting change in the amount of

organic matter being produced within the stream environment. Throughflow of shallow

alluvial groundwater (AW) appears to moderate the effects of evaporative

concentration, either through dilution or compensation for evaporative losses; but only

affects the source and composition of organic matter on rare occasions, such as when

substantial proportions of DOM inputs from AW throughflow are highly degraded.

Consequently, patterns of organic matter utilisation by both microbial and

macroinvertebrate food webs do not appear to change as pools dry, despite the fact that

macroinvertebrates also concentrate in pools as they evaporate. However, this may be a

result of general adaptations of microbial and metazoan communities to the ephemeral

nature of their habitat, rather than being directly related to the degree of contraction and

drying within pools in any given year. In addition, δ18O values are an indicator of

consistent AW connectivity, not of differences in pool morphology, rates of AW

throughflow, or rates of expansion and contraction of wetted area. The relationship

Page 109: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

99

between AW inputs and ecosystem structure or processes may therefore be more

complex than is immediately apparent.

My research demonstrates that understanding how variation in groundwater

connectivity affects ecological processes in pools will help illuminate how they fit into

overall conceptual models of intermittent streams. Here, I discuss the patterns of organic

matter source and utilisation within pools of Pilbara streams and how intermittent flow

regimes may have created or maintained them. I also consider how these intermittent

stream ecosystems may also fit into wider paradigms emerging within freshwater

ecology as a whole.

Research outcomes: ecological consequences of intermittent flow

Intermittent rivers and streams can be seen as “punctuated biogeochemical reactors”,

systems in which organic matter transport and processing occurs quickly during times

of inundation and then slows as surface water contracts and dries (Larned et al., 2010):

an ultimate adaptation of the nutrient spiralling concept which sees organic matter and

nutrients taken up by organisms and processed only to be released, transported

downstream, taken up and processed further; with processing lengths conditional to the

interacting effects of flow and disturbance (Fisher et al., 1998b). Conversely, pools or

waterholes that persist through dry periods within intermittent rivers and streams may

well be better represented by conceptual models adapted from lakes or wetlands (e.g.

Fellows et al., 2007; Heffernan, 2008), as the presence of surface water means rapid

processing of organic matter continues almost indefinitely (bar complete drying). The

intermittent streams of the Pilbara, and in particular Coondiner Creek, would seem to fit

this category; with shallow, macrophyte dominated and nutrient-poor lakes likely to

provide a useful comparison (Scheffer & Nes, 2007). However, the constant pressures

of evaporation and contraction may also place the pools of the Pilbara close to those of

temporary ponds and lakes (e.g. Sahuquillo et al., 2012; Holgerson, 2015). Regardless

of how they function in inter-flood periods, pools and waterholes that form along

intermittent streams are also units within that system, and when floods or flow pulses

occur the biogeochemical changes associated with the rapid inundation of previously

dry areas (Larned et al., 2010; Datry et al., 2014b) will still be a major driver of their

structure and function (e.g. Burford et al., 2008). Understanding how these interacting

influences affect the pools of intermittent streams is obviously a major challenge.

Page 110: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

100

In pools of intermittent streams in the Pilbara, high insolation, warm and shallow water,

and infrequent flood disturbance should produce a high potential for algal growth.

However, my research has shown that pools are strongly net heterotrophic (Chapter 4)

and that algae are not the dominant source of energy supporting macroinvertebrate

production (Chapter 5). Low autochthonous production is also indicated by the

dominance of allochthonous DOM in pools across multiple catchments (Chapters 2 and

3). This may in part be attributable to limitation of algal production by low

concentrations of inorganic nutrients in intermittent streams of the Pilbara (Pinder et al.,

2010; Fellman et al., 2011a), which is in turn reflective of the ancient, nutrient depleted

landscape of the Pilbara (Ford et al., 2007; Pillans, 2007; McIntyre et al., 2009).

Strong uptake of inorganic nutrients from alluvial water inputs by aquatic macrophytes,

particularly charophytes, may also limit algal production (Chapter 3). However, net

heterotrophy may simply be a general consequence of a lack of surface flow, which

leaves allochthonous organic matter to be decomposed fully at the same location where

it enters stream ecosystems rather than being sequentially transporting to downstream

locations (Fisher et al., 1998b; Webster, 2007). Intermittent rivers and streams

exemplify this phenomenon, particularly where pool ecosystems have high levels of

autochthonous primary productivity in surface water or along littoral fringes but are net

heterotrophic overall (e.g. Grimm & Fisher, 1984; Fellows et al., 2009).

The lack of downstream transport and retention of allochthonous organic matter may

also be a controlling factor in the development of priming effects (Chapter 4). However,

the cessation of surface flow may not be the only factor involved in the development of

priming effects. Priming is the increased mineralization of previously recalcitrant

organic matter in the presence of new, labile carbon sources and has been widely

reported in soils, particularly in the rhizosphere (Guenet et al., 2010; Cheng et al.,

2014). Theoretically, it makes more sense for priming effects to occur where

autochthonous production is low, as under high levels of primary productivity an

additional source of organic matter and nutrients (in this case, allochthonous organic

matter) would not need to be mineralised to meet the demands of bacterial metabolism.

Accordingly, autochthonous primary production may provide enough energy to support

the majority of bacterial metabolism in pools of other intermittent systems which are

highly productive but, overall, net heterotrophic (e.g. Jones et al., 1995b; Fellows et al.,

2007). The predominant autotrophs in these systems are algae or cyanobacteria in the

surface stream or pools. In contrast, the charophytes which dominate the autotrophic

Page 111: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

101

communities of Coondiner Creek can shade out algae within benthic biofilms (Gette-

Bouvarot et al., 2015) or exude allelopathic compounds into the water column which

inhibit algal growth (van Donk & van de Bund, 2002). Charophytes do promote

microbial metabolism and, possibly, chemoautotrophic processes such as nitrification

by oxygenating sediments (Kufel & Kufel, 2002). However, the amount of labile carbon

produced by root exudates, for example, is likely to be much less than that exuded if

there were highly productive algae in the surface stream; thereby perhaps promoting the

mineralisation of additional sources of carbon. The organic matter dynamics of

Coondiner Creek therefore suggest that priming effects may be concentrated in the

sediments of low productivity intermittent rivers, and particularly in the rhizosphere of

aquatic plants, rather than the surface stream (Kuzyakov, 2002). Labile carbon

production being concentrated within sediments, rather than the surface pool, may also

help explain why DOM fluorescence indicators of autochthonous production within the

water column do not appear to change along with the hydrological variation of pools

(Chapter 3).

Importance of groundwater connectivity in contracting intermittent streams

Pools with δ18O signatures indicative of sustained alluvial water connectivity may still

contract through the dry season (Chapter 3). While this contraction is partly due to

transmission losses, it is also probably associated with falling alluvial water tables as

the dry season progresses (e.g. Dahm et al., 2003). In intermittent systems where little

to no connectivity exists between remnant surface water and groundwater during inter-

flood periods (e.g. Hamilton et al., 2005), changes in ecological processes as pools dry

are likely to be directly related to the degree of wetted area contraction. However,

studies of intermittent streams elsewhere that have observed considerable variation in

indicators of biogeochemical processes amongst pools during contraction or

fragmentation of surface water (e.g. von Schiller et al., 2011) may likely find that much

of this variation is attributable to differences in groundwater connectivity. More

specifically, when concentrations of organic matter do not occur with contraction, it is

likely to be attributable to the dilution effect of alluvial throughflow (Chapters 2 and 3).

Whether these differences in the concentration of organic matter among pools have a

direct influence on ecological processes is still unclear, as AW connectivity appears to

have little functional effect on basal carbon sources and trophic structure (Chapter 5).

Page 112: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

102

The possible direct exception may be support of aquatic macrophyte growth by

inorganic nutrient inputs (Chapter 3). However, ecosystem metabolism appears to be

unrelated to AW connectivity, while fluorescence indicators of DOM source show little

variation between AW-connected and isolated pools (Chapters 2 and 4).

Macroinvertebrates also show little evidence of differing dietary habits in pools with

different AW connectivity (Chapter 5). These results are surprising in many ways. For

example, the concentration of coloured DOM within nutrient-poor lakes largely controls

benthic productivity through light attenuation, and therefore overall ecosystem

metabolism (Ask et al., 2009; Karlsson et al., 2009). While the pools within Coondiner

Creek were shallow, they also appear to have high light attenuation (Chapter 4). Part of

the difficulty in explaining these patterns is due to the complexity of both DOM

composition and the effects of different DOM fractions on aquatic ecosystems, where,

for example, the types of humic compounds present within DOM can alternatively

inhibit (through interference with electron transport mechanisms) or enhance the

photosynthetic productivity (through greater inhibition of competitors) of different

species of algae (Steinberg et al., 2006).

In the case of macroinvertebrate communities, some illumination may be gained in the

disconnection between δ18O values and the observed degree of contraction through the

dry season. Macroinvertebrate abundances were higher in more highly evaporated pools

as indicated by δ18O (Chapter 5), but the pools identified as having AW connectivity

were still more highly contracted in the late compared with the early-mid dry season

(Chapter 3). This indicates that it may be an associated ecological effect of AW

connectivity, rather than just a slowdown or prevention of contraction, which creates

lower macroinvertebrate abundance. Pools with AW throughflow tend to maintain their

relative, percentage cover of aquatic macrophytes through contraction while AW-

isolated pools experience a reduction in percentage cover (Chapter 3).

Macroinvertebrates in AW-connected pools may therefore have a wider area of high-

quality habitat to exploit (Warfe et al., 2008; Pettit et al., 2011; Ziegler et al., 2015).

This relationship may have been obscured by the overall temporal trend in

macroinvertebrate abundance caused by contraction of pools (Chapter 5).

These interactions between generalised ecological adaptations to intermittent flow and

hydrological variation at a more local level are likely to be an important feature of

intermittent stream ecology. For example, the “dynamic stability” exhibited by some

intermittent stream food webs is increasingly being related to shifting mosaics of

Page 113: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

103

terrestrial and aquatic habitats created by intermittent flow (Leigh et al., 2015), but the

ability to respond to these changing conditions may ultimately be determined by

species’ innate resilience or resistance to drying (Datry et al., 2014a). As discussed

above, the occurrence of close links between primary productivity and ecosystem

respiration may be due to the interrelated factors of oligotrophy, a dominance of

allochthonous over autochthonous sources of carbon, and the predominance of aquatic

macrophytes over algae; but that overall, increased internal recycling of organic matter

and nutrients within biofilms (Mulholland et al., 1995) or sediments (Scheffer & Nes,

2007) as an adaptation to decreased or absent flow is required to create these patterns.

Biogenic methane in intermittent stream food webs

The results of Chapter 5 support the premise that methane-derived carbon may provide

more of the trophic base for secondary production in freshwater environments than has

generally been acknowledged (Jones & Grey, 2011). Methane-oxidising bacteria

(MOB) can be highly productive at oxic-anoxic interfaces when the production of large

amounts of methane from organic matter occurs in anoxic sediments. However, there is

also limited evidence that methane-derived carbon may provide a significant energy

source for a wider range of taxa within stream backwaters and dystrophic, or eutrophic,

shallow lakes. The stable isotope analysis of macroinvertebrates within Coondiner

Creek (Chapter 5) therefore raises the question of whether intermittent or ephemeral

flow regimes may promote the utilisation of methane-derived carbon by metazoan food

webs.

A possible pathway for the promotion of MOB utilisation in Coondiner Creek may be

through the predominance of aquatic vegetation, and therefore indirectly through the

lack of flow disturbance. Extensive aquatic macrophyte growth may prevent the

resuspension of particulate matter through a reduction in turbulence at the sediment-

water interface, as in other shallow lentic environments (Scheffer & Nes, 2007).

Accordingly, the proportion of fine particulate organic matter (FPOM) in the seston of

Coondiner Creek pools appears to be very low, especially given the absence of

predominant filter feeders such as the bivalves common to other Australian pools or

waterholes (Bunn et al., 2003; Leigh et al., 2010). Despite the prevalence of aquatic

vegetation in Coondiner Creek, pools during inter-flood periods appear to be largely net

heterotrophic and also experience low oxygen concentrations during the night (Chapter

Page 114: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

104

4). FPOM settling out of the water column may therefore be trapped and metabolised by

bacterial communities under anaerobic conditions at the sediment-water interface, rather

than by macroinvertebrate consumers. If electron acceptors other than oxygen

(predominantly NO3-, Fe3+, and sulphate) are low in concentration, and light organic

compounds produced by anaerobic fermentation accumulate in the final stages of

organic matter decomposition, then methanogenesis should be promoted.

Aquatic macrophytes also oxygenate the sediments that they grow in, but it is likely that

these oxygenated zones do not extend far past the root environment or rhizosphere. In

some cases aeration of roots themselves may not even be complete (Segers, 1998).

Large surface areas of oxic-anoxic interface within the sediment may therefore be

created along fine rooting systems such as the extensive branching rhizoids of

charophytes, enhancing the biological oxidation of methane (Gerard & Chanton, 1993).

Conditions that promote methane-derived carbon entering the food webs of intermittent

streams may therefore be 1) lack of or low flow and 2) proliferation of aquatic

vegetation, leading to 3) a lack of resuspension of POM together with 4) widespread

zones of oxic-anoxic interface which create high surface areas for methanogenesis,

methane oxidation, and access to MOB for grazers. The last condition may be further

enhanced by the accumulation of large amounts of terrestrial detritus together with net

heterotrophy or intermittent development of low dissolved oxygen (DO) concentrations

(e.g. over the nightly phase of diel DO curves), given that anaerobic carbon

mineralisation is a major control on methane production (Segers, 1998).

The specificity of these conditions is emphasised by stable isotope analyses of

intermittent and ephemeral rivers and streams in Australia showing varying evidence of

methane contribution to food webs. While usually net heterotrophic overall, food webs

in waterholes of central Australia typically show dependence on thin but highly

productive bands of littoral algae or on pulses of floodplain production (Bunn et al.,

2003; Burford et al., 2008). In these systems, only Chironimidae larvae may show

evidence of methanotrophy (Bunn et al., 2003). However, these pools are typically

highly turbid and aquatic macrophytes are scarce. Similarly, the macroinvertebrates of

pools within intermittent rivers of the Australian tropics studied by Leigh et al. (2010)

showed little evidence of utilising methane-derived carbon, but the pools along these

waterways also tend to be highly turbid after cessation of surface flow. The most similar

systems to Coondiner Creek appears to be the three waterholes of southeastern Australia

studied by Bunn & Boon (1993), where much of the macroinvertebrate taxa had

Page 115: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

105

depleted δ13C values relative to measured primary sources. However, their sites contrast

with Coondiner Creek in that while they were extensively vegetated, they did not appear

to be oligotrophic (supporting frequent algal blooms along with high water column

chlorophyll a concentrations) and filter feeders were common within the

macroinvertebrate fauna. Additionally, the intermittent streams of southeastern

Australia studied by Reid et al. (2008), while heavily vegetated, rich in detrital organic

matter, and relatively low in turbidity (6 – 36 NTU), had a macroinvertebrate fauna

without any of the characteristically depleted 13C-signatures of methane-derived carbon.

Clearly, contributions of methane-derived carbon to intermittent stream food webs are

not strongly linked to any single one of the conditions listed above.

A simple first step for disentangling the complexity of methanotrophy in intermittent

systems is to make measurement of methane δ13C (and δ15N) values standard in studies

of food web trophic structure. Estimates of contributions of methane-derived carbon are

also largely based on dual-isotope natural abundance studies, whereas labelling

experiments may be able to more clearly define how much carbon is transferred from

MOB to other organisms (e.g. Deines, Bodelier & Eller, 2007a; Kip et al., 2010).

However, it is worth noting that while methane is easily measured in most aquatic

environments, determining the δ13C of the vital intermediate between methane and

metazoan food webs – MOB – is much more technically difficult due to their difficulty

to culture (Jones & Grey, 2011). A productive approach may be to measure the δ13C of

lipid biomarkers specific to methanotrophs (Nichols et al., 1985; Crossman, Ineson &

Evershed, 2005), particularly in conjunction with the measurement of redox state and

methane concentrations around possible zones of oxic-anoxic interface such as the

rooting systems of aquatic macrophytes. Again, labelling experiments may provide

increased clarity on the incorporation of methane into microbial biomarkers (e.g. Deines

et al., 2007a; Dumont et al., 2011). However, the current lack of information on how

the δ13C of MOB differs from their lipid biomarkers under different environmental or

cell growth conditions limits their usefulness at present (Jones & Grey, 2011).

Nevertheless, measurement of the spatial distribution, concentration and δ13C of

methane is likely to become an important part of the investigation of intermittent and

ephemeral stream food webs.

Page 116: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

106

The source and lability of groundwater carbon in intermittent systems

My research has demonstrated that while AW throughflow usually dilutes the organic

matter pools available within intermittent streams (Chapter 2), at times the contribution

of tyrosine-like fluorescence components possibly linked to highly degraded DOM is

enhanced by AW connectivity. Recent studies of streams with highly seasonal flow

regimes have indicated that DOM inputs from groundwater, particularly those

correlated with tyrosine-like fluorescence, may be both highly bioavailable and

extremely old (Fellman et al., 2014; Helton et al., 2015). Radiocarbon dating of DOM

within Coondiner Creek has revealed that it may also have high percentages of ancient

carbon (Appendix A). Previously, the increase in the inorganic:organic nutrient ratio

associated with inputs of low-DOM concentration groundwater within intermittent

dryland streams was hypothesised to promote autochthonous primary production over

heterotrophic metabolism (Dahm et al., 2003). However, if groundwater inputs of old

DOM are low but highly labile, then they may be an important key to the development

of net heterotrophic conditions in intermittent streams, particularly those with low

nutrient enrichment (e.g. Chapter 4; Townsend et al., 2011). Crucial questions about the

inputs of groundwater carbon to freshwater ecosystems are, therefore: first, where does

this apparently ancient carbon originate; second, how and why is it transported to

streams so much later; and finally, why is it so labile, given that there is often high

competition for carbon along groundwater flowpaths?

Soil organisms are constrained by their environment towards low metabolic rates and

specialisation to particular microhabitats (Ekschmitt et al., 2005). Old stores of organic

carbon, possibly as particulate organic carbon (POC) buried beneath alluvial sediments,

may not be decomposed if accessing them is metabolically costly for microbial and

metazoan fauna. These “partial refuges” of organic matter, which are stabilised in soil

not only by substrate quality or soil conditions but also by the biology of soil organisms,

are increasingly being recognised as major components of organic matter persistence

within soils (Ekschmitt et al., 2005). The utilisation of ancient carbon by stream

ecosystems may therefore be enhanced within intermittent or highly seasonal flow

regimes. Alluvial groundwater flowpaths can experience abrupt and substantial

reorganization in rivers and streams with high variation in geomorphology and flow

regime, particularly when previously dry areas are inundated (Poole et al., 2006). This

may expose previously buried carbon stores to leaching, transport and transformations

along hydrologic flowpaths (Fig. 6.1).

Page 117: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

107

Figure 6.1: A conceptual model of the input of old, buried carbon into pool ecosystems

through alluvial water flowpaths and the possible implications for organic matter

cycling within pools, based on their connections to alluvial water.

When finally transported to surface water, this older carbon may become more

bioavailable upon cleavage of photoreactive DOM (McCallister & del Giorgio, 2012).

However, older DOM has been shown to be more bioavailable in both surface water

(Fellman et al., 2014) and groundwater (Helton et al., 2015), suggesting that

environmental factors within aquifers control its lability. In particular, the correlation

between tyrosine-like fluorescence and labile groundwater DOM suggests that it is

either autochthonous in origin or heavily degraded (Yamashita & Tanoue, 2004; Elliott

et al., 2006). Helton et al. (2015) suggest that chemoautotrophic production of organic

carbon is a possible source of labile carbon within aquifers. Production of new, labile

DOM within aquifers may be the controlling factor in the decomposition of buried,

ancient POC (Fontaine et al., 2007); which, if conducted under anaerobic conditions,

may result in methane being produced. Microbial community structure could therefore

play a major role in shaping the composition and bioavailability of groundwater DOM

(Shen et al., 2015). If methane produced from ancient POC decomposition is oxidised

by methanotrophs as a carbon source, then metabolites produced by methane-oxidising

bacteria could well provide a new source of labile carbon; consequently stimulating

more decomposition of POC. If this is the case, then labile carbon produced via these

pathways would carry the 14C signature of the older POC source. A lack of buried POC

to enable this methane-methanotrophy-POC-decomposition-methane loop may explain

why, in contrast, the bioavailability of DOC tends to decrease along shallower

hyporheic flowpaths (Jones et al., 1995c; Sobczak & Findlay, 2002).

Page 118: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

108

Methane produced from ancient carbon decomposition within aquifers may also enter

surface water via transport through upwellings. If so, then it may be another mechanism

by which incorporation of biogenic methane into stream food webs occurs (Jones &

Grey, 2011). Measurement of 14C within freshwater food webs which have shown

evidence of methanotrophy may therefore also help resolve the source and ecological

role of ancient groundwater carbon, although separating groundwater inputs of methane

from those produced within streams is likely to be a significant logistical challenge. In

addition, measurement of the spectral properties of methanotrophic bacterial metabolites

may well show high proportions of tyrosine-like fluorescence, providing a rapidly

measurable indicator that methanotrophy might be occurring within aquatic

environments. Measurements of 14C in methanotrophic biomarkers (as discussed above)

may also prove productive in this regard. Ultimately, further chemical characterisation

of groundwater DOM should provide clarification of the possible source, age and roles

of groundwater carbon. Incubation experiments which typically measure bulk DOC loss

(as in Fellman et al., 2014; Helton et al., 2015) can be expanded to measure the

composition of DOM before and after incubation. Conducting anaerobic as well as

aerobic incubations may also provide insight into the processes controlling the lability

of groundwater DOM if they involve processes such as methanogenesis or biological

sulphate reduction, particularly in the case of bioassays which specifically measure the

methane output of anaerobic microbial communities given a range of possible carbon

sources (e.g. Jones et al., 2008a). Comparisons of ancient groundwater DOM with those

of possibly analogous systems, such as the food webs of proglacial streams supported

by release of ancient carbon from glaciers, may also be important (Fellman et al., 2015).

Finally, tracing the variability of groundwater flowpaths may prove useful in identifying

the mechanisms leading to release of ancient DOM into streams, particularly in

intermittent systems.

Conclusions

Organic matter dynamics within intermittent streams of the Pilbara are clearly complex

and closely tied to patterns of both evaporation and alluvial water connectivity, and are

therefore likely to be highly affected by altered dynamics of flow, whether natural or

anthropogenic. Strikingly, my research indicates that streams of the Pilbara, and

intermittent streams in general, may be ideal environments to investigate several recent,

Page 119: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

109

novel and interconnected paradigms of freshwater ecology: the importance of priming

effects in the breakdown of relatively recalcitrant carbon; the widespread incorporation

of methane into freshwater food webs; and the support of freshwater ecosystems by

ancient carbon. As this thesis presents some of the first information on organic matter

cycling within stream pools of northwest Australia, particularly with regard to

ecological stoichiometry, the trophic base of food webs, and ecosystem metabolism,

there is obviously much more that can be learned from additional research that

combines empirical and experimental approaches. Some of the most important

questions to be addressed include:

Is variation in microbial community composition linked to differences in DOM

composition, especially with regard to changes measured during decomposition

of different POM sources?

What is the role of POM in driving microbial community composition and

metabolism, and is the main source of DOM in stream pools decomposition of

terrestrial POM?

Is methane produced via decomposition of buried organic matter and subsequent

transport to pools in groundwater flowpaths, or through in-pool decomposition

in intermittent or stratified anoxic zones?

Does the eutrophic nature of pools within northwest Australia limit the

production and cycling of labile organic matter, or does the lack of downstream

transport associated with cessation of surface flow lead to tightly coupled

internal recycling within pool ecosystems even under nutrient additions?

How do the patterns of organic matter production and utilization change in these

pools in the times immediately following flooding, when I was unable to

sample?

Further understanding of the effects of variation in alluvial water connectivity on

ecological processes in intermittent streams should therefore inform both the effective

management and restoration of intermittent systems, as well as continuing to advance

the science of freshwater ecology as a whole.

Page 120: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

110

Page 121: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

111

REFERENCES

Acuña V., Giorgi A., Muñoz I., Uehlinger U. & Sabater S. (2004) Flow extremes and

benthic organic matter shape the metabolism of a headwater mediterranean

stream. Freshwater Biology, 49, 960-971.

Acuña V., Muñoz I., Giorgi A., Omella M., Sabater F. & Sabater S. (2005) Drought and

postdrought recovery cycles in an intermittent mediterranean stream: structural

and functional aspects. Journal of the North American Benthological Society,

24, 919-933.

Ågren G.I. (2004) The C:N:P stoichiometry of autotrophs – theory and observations.

Ecology Letters, 7, 185-191.

Aitkenhead-Peterson J.A., McDowell W.H. & Neff J.C. (2003) Sources, production,

and regulation of allochthonous dissolved organic matter inputs to surface

waters. In: Aquatic Ecosystems: Interactivity of Dissolved Organic Matter. (Eds

R.L. Sinsabaugh & S. Findlay), pp. 25-70. Academic Press, San Diego.

Amalfitano S., Fazi S., Zoppini A., Barra Caracciolo A., Grenni P. & Puddu A. (2008)

Responses of benthic bacteria to experimental drying in sediments from

mediterranean temporary rivers. Microbial Ecology, 55, 270-279.

Ask J., Karlsson J., Persson L., Ask P., Byström P. & Jansson M. (2009) Terrestrial

organic matter and light penetration: effects on bacterial and primary production

in lakes. Limnology and Oceanography, 54, 2034-2040.

Austin A.T. & Vivanco L. (2006) Plant litter decomposition in a semi-arid ecosystem

controlled by photodegradation. Nature, 442, 555-558.

Bade D.L., Carpenter S.R., Cole J.J., Hanson P.C. & Hesslein R.H. (2004) Controls of

d13C-DIC in lakes: geochemistry, lake metabolism, and morphometry.

Limnology and Oceanography, 49, 1160-1172.

Baillie M.N., Hogan J.F., Ekwurzel B., Wahi A.K. & Eastoe C.J. (2007) Quantifying

water sources to a semiarid riparian ecosystem, San Pedro River, Arizona.

Journal of Geophysical Research, 112, G03S02.

Bärlocher F. (1985) The role of fungi in the nutrition of stream invertebrates. Botanical

Journal of the Linnean Society, 91, 83-94.

Battin T.J., Kaplan L.A., Findlay S., Hopkinson C.S., Marti E., Packman A.I., Newbold

J.D. & Sabater F. (2008) Biophysical controls on organic carbon fluxes in fluvial

networks. Nature Geoscience, 1, 95-100.

Page 122: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

112

Beesley L.S. & Prince J. (2010) Fish community structure in an intermittent river: the

importance of environmental stability, landscape factors and within-pool habitat

descriptors. Marine and Freshwater Research, 61, 605-614.

Belnap J., Welter J.R., Grimm N.B., Barger N. & Ludwig J.A. (2005) Linkages between

microbial and hydrologic processes in arid and semiarid watersheds. Ecology,

86, 298-307.

Benfield E.F. (1997) Comparison of litterfall input to streams. Journal of the North

American Benthological Society, 16, 104-108.

Bengtsson G., Bengtson P. & Månsson K.F. (2003) Gross nitrogen mineralization,

immobilization, and nitrification rates as a function of soil C/N ratio and

microbial activity. Soil Biology and Biochemistry, 35, 143-154.

Bentley D., Grierson P.F., Bennett L.T. & Adams M.A. (1999) Evaluation of anion

exchange membranes to estimate bioavailable phosphorus in native grasslands

of semi‐arid northwestern Australia. Communications in Soil Science & Plant

Analysis, 30, 2231-2244.

Bertilsson S. & Jones J.B. (2003) Supply of dissolved organic matter to aquatic

ecosystems: autochthonous sources. In: Aquatic Ecosystems: Interactivity of

Dissolved Organic Matter. (Eds S.E.G. Findlay & R.L. Sinsabaugh), pp. 3-25.

Academic Press, San Diego.

Blanchette M.L., Davis A.M., Jardine T.D. & Pearson R.G. (2013) Omnivory and

opportunism characterize food webs in a large dry-tropics river system.

Freshwater Science, 33, 142-158.

Boon P.I. (1991) Bacterial assemblages in rivers and billabongs of southeastern

Australia. Microbial Ecology, 22, 27-52.

Boon P.I., Mitchell A. & Lee K. (1997) Effects of wetting and drying on methane

emissions from ephemeral floodplain wetlands in south-eastern Australia.

Hydrobiologia, 357, 73-87.

Boulton A.J. & Lake P.S. (1992) The ecology of two intermittent streams in victoria,

australia: iii. temporal changes in faunal composition. Freshwater Biology, 27,

123-138.

Brablcová L., Buriánková I., Badurová P., Chaudhary P.P. & Rulík M. (2015)

Methanogenic archaea diversity in hyporheic sediments of a small lowland

stream. Anaerobe, 32, 24-31.

Page 123: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

113

Brooks P.D., Haas P.A. & Huth A.K. (2007) Seasonal variability in the concentration

and flux of organic matter and inorganic nitrogen in a semiarid catchment, San

Pedro River, Arizona. Journal of Geophysical Reearch., 112, G03S04.

Bunn S.E. & Boon P.I. (1993) What sources of organic carbon drive food webs in

billabongs? A study based on stable isotope analysis. Oecologia, 96, 85-94.

Bunn S.E., Davies P.M. & Winning M. (2003) Sources of organic carbon supporting the

food web of an arid zone floodplain river. Freshwater Biology, 48, 619-635.

Bunn S.E., Balcombe S.R., Davies P.M., Fellows C.S. & Mckenzie-Smith F.J. (2006a)

Aquatic productivity and food webs of desert river ecosystems. In: Ecology of

Desert Rivers. (Ed R.T. Kingsford). Cambridge University Press, Cambridge.

Bunn S.E., Thoms M.C., Hamilton S.K. & Capon S.J. (2006b) Flow variability in

dryland rivers: boom, bust and the bits in between. River Research and

Applications, 22, 179-186.

Bunn S.E., Leigh C. & Jardine T.D. (2013) Diet-tissue fractionation of δ15N by

consumers from streams and rivers. Limnology and Oceanography, 58, 765-773.

Burford M.A., Cook A.J., Fellows C.S., Balcombe S.R. & Bunn S.E. (2008) Sources of

carbon fuelling production in an arid floodplain river. Marine and Freshwater

Research, 59, 224-234.

Busch D.E. & Fisher S.G. (1981) Metabolism of a desert stream. Freshwater Biology,

11, 301-307.

Callisto M. & Graça M.a.S. (2013) The quality and availability of fine particulate

organic matter for collector species in headwater streams. International Review

of Hydrobiology, 98, 132-140.

Cammack W.K.L., Kalff J., Prairie Y.T. & Smith E.M. (2004) Fluorescent dissolved

organic matter in lakes: relationships with heterotrophic metabolism. Limnology

and Oceanography, 49, 2034-2045.

Cao Y., Larsen D.P. & Thorne R.S.-J. (2001) Rare species in multivariate analysis for

bioassessment: some considerations. Journal of the North American

Benthological Society, 20, 144-153.

Catalán N., Obrador B., Alomar C. & Pretus J. (2013) Seasonality and landscape factors

drive dissolved organic matter properties in mediterranean ephemeral washes.

Biogeochemistry, 112, 261-274.

Chan M., Moser K., Davis J., Southam G., Hughes K. & Graham T. (2005) Desert

potholes: ephemeral aquatic microsystems. Aquatic Geochemistry, 11, 279-302.

Page 124: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

114

Cheng W., Parton W.J., Gonzalez-Meler M.A., Phillips R., Asao S., Mcnickle G.G.,

Brzostek E. & Jastrow J.D. (2014) Synthesis and modeling perspectives of

rhizosphere priming. New Phytologist, 201, 31-44.

Clarke K.R. (1993) Non-parametric multivariate analyses of changes in community

structure. Australian Journal of Ecology, 18, 117-143.

Coble P.G. (1996) Characterization of marine and terrestrial DOM in seawater using

excitation-emission matrix spectroscopy. Marine Chemistry, 51, 325-346.

Cole J.J., Prairie Y.T., Caraco N.F., McDowell W.H., Tranvik L.J., Striegl R.G., Duarte

C.M., Kortelainen P., Downing J.A., Middelburg J.J. & Melack J. (2007)

Plumbing the global carbon cycle: integrating inland waters into the terrestrial

carbon budget. Ecosystems, 10, 172-185.

Collins S.L., Sinsabaugh R.L., Crenshaw C., Green L., Porras-Alfaro A., Stursova M. &

Zeglin L.H. (2008) Pulse dynamics and microbial processes in aridland

ecosystems. Journal of Ecology, 96, 413-420.

Coplen T.B. (1994) Reporting of stable hydrogen, carbon, and oxygen isotopic

abundances. Pure and Applied Chemistry, 66, 273-276.

Coplen T.B. (1995) Discontinuance of SMOW and PDB. Nature, 375, 285-285.

Coplen T.B., Brand W.A., Gehre M., Gröning M., Meijer H.a.J., Toman B. &

Verkouteren R.M. (2006) New guidelines for δ13C measurements. Analytical

Chemistry, 78, 2439-2441.

Cory R.M. & Kaplan L.A. (2012) Biological lability of streamwater fluorescent

dissolved organic matter. Limnology and Oceanography, 57, 1347-1360.

Craig H. & Gordon L.I. (1965) Deuterium and oxygen 18 variations in the ocean and

the marine atmosphere. In: Stable Isotopes in Oceanographic Studies and

Paleotemperatures. (Ed E. Tongiorgi), pp. 9-130. Laboratorio di Geologia

Nucleare, Pisa, Italy.

Creed I.F., Mcknight D.M., Pellerin B.A., Green M.B., Bergamaschi B.A., Aiken G.R.,

Burns D.A., Findlay S.E.G. et al. (2015) The river as a chemostat: fresh

perspectives on dissolved organic matter flowing down the river continuum.

Canadian Journal of Fisheries and Aquatic Sciences.

Cross W.F., Benstead J.P., Frost P.C. & Thomas S.A. (2005) Ecological stoichiometry

in freshwater benthic systems: recent progress and perspectives. Freshwater

Biology, 50, 1895-1912.

Page 125: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

115

Crossman Z.M., Ineson P. & Evershed R.P. (2005) The use of 13C labelling of bacterial

lipids in the characterisation of ambient methane-oxidising bacteria in soils.

Organic Geochemistry, 36, 769-778.

Dahm C.N., Baker M.A., Moore D.I. & Thibault J.R. (2003) Coupled biogeochemical

and hydrological responses of streams and rivers to drought. Freshwater

Biology, 48, 1219-1231.

Datry T., Larned S.T., Fritz K.M., Bogan M.T., Wood P.J., Meyer E.I. & Santos A.N.

(2014a) Broad-scale patterns of invertebrate richness and community

composition in temporary rivers: effects of flow intermittence. Ecography, 37,

94-104.

Datry T., Larned S.T. & Tockner K. (2014b) Intermittent rivers: a challenge for

freshwater ecology. BioScience 64, 229-235.

De Montety V., Martin J.B., Cohen M.J., Foster C. & Kurz M.J. (2011) Influence of

diel biogeochemical cycles on carbonate equilibrium in a karst river. Chemical

Geology, 283, 31-43.

Deines P., Bodelier P.L.E. & Eller G. (2007a) Methane-derived carbon flows through

methane-oxidizing bacteria to higher trophic levels in aquatic systems.

Environmental Microbiology, 9, 1126-1134.

Deines P., Grey J., Richnow H.H. & Eller G. (2007b) Linking larval chironomids to

methane: seasonal variation of the microbial methane cycle and chironomid

δ13C. Aquatic Microbial Ecology, 46, 273-282.

Dekar M., Magoulick D. & Huxel G. (2009) Shifts in the trophic base of intermittent

stream food webs. Hydrobiologia, 635, 263-277.

Del Giorgio P.A. & Davis J. (2003) Patterns in dissolved organic matter lability and

consumption across aquatic ecosystems. In: Aquatic Ecosystems: Interactivity of

Dissolved Organic Matter. (Eds S. Findlay & R.L. Sinsabaugh), pp. 399-423.

Academic Press, San Diego.

Department of Mines and Petroleum (2014) Western Australian mineral and pertoleum

statistics digest 2013-14. Government of Western Australia, Perth, Australia.

Department of Water (2010) Pilbara regional water plan 2010-2030. Department of

Water, Perth, Western Australia.

Dieter D., Frindte K., Krüger A. & Wurzbacher C. (2013) Preconditioning of leaves by

solar radiation and anoxia affects microbial colonisation and rate of leaf mass

loss in an intermittent stream. Freshwater Biology, 58, 1918-1931.

Page 126: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

116

Doctor D.H., Kendall C., Sebestyen S.D., Shanley J.B., Ohte N. & Boyer E.W. (2008)

Carbon isotope fractionation of dissolved inorganic carbon (DIC) due to

outgassing of carbon dioxide from a headwater stream. Hydrological Processes,

22, 2410-2423.

Dogramaci S., Skrzypek G., Dodson W. & Grierson P.F. (2012) Stable isotope and

hydrochemical evolution of groundwater in the semi-arid Hamersley Basin of

subtropical northwest Australia. Journal of Hydrology, 475, 281-293.

Dogramaci S., Firmani G., Hedley P., Skrzypek G. & Grierson P.F. (2015) Evaluating

recharge to an ephemeral dryland stream using a hydraulic model and water,

chloride and isotope mass balance. Journal of Hydrology, 521, 520-532.

Drysdale R., Lucas S. & Carthew K. (2003) The influence of diurnal temperatures on

the hydrochemistry of a tufa-depositing stream. Hydrological Processes, 17,

3421-3441.

Dumont M.G., Pommerenke B., Casper P. & Conrad R. (2011) DNA, rRNA and

mRNA-based stable isotope probing of aerobic methanotrophs in lake sediment.

Environmental Microbiology, 13, 1153-1167.

Ekschmitt K., Liu M., Vetter S., Fox O. & Wolters V. (2005) Strategies used by soil

biota to overcome soil organic matter stability — why is dead organic matter left

over in the soil? Geoderma, 128, 167-176.

Elliott S., Lead J.R. & Baker A. (2006) Characterisation of the fluorescence from

freshwater, planktonic bacteria. Water Research, 40, 2075-2083.

Elser J.J. & Urabe J. (1999) The stoichiometry of consumer-driven nutrient recycling:

theory, observations, and consequences. Ecology, 80, 735-751.

Enríquez S., Duarte C.M. & Sand-Jensen K. (1993) Patterns in decomposition rates

among photosynthetic organisms: the importance of detritus C:N:P content.

Oecologia, 94, 457-471.

Environmental Protection Authority (2002) Hydrogeological research programme at

Marandoo – trial dewatering and re-injection test, Karijini National Park. The

Government of Western Australia, Perth.

Farquhar G.D., Ehleringer J.R. & Hubick K.T. (1989) Carbon isotope discrimination

and photosynthesis. Annual Review of Plant Physiology and Plant Molecular

Biology, 40, 503-537.

Fellman J.B., Dogramaci S., Skrzypek G., Dodson W. & Grierson P.F. (2011a)

Hydrologic control of dissolved organic matter biogeochemistry in pools of a

subtropical dryland river. Water Resources Research, 47, W06501.

Page 127: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

117

Fellman J.B., Hood E. & Spencer R.G.M. (2011b) Fluorescence spectroscopy opens

new windows into dissolved organic matter dynamics in freshwater ecosystems:

A review. Limnology and Oceanography, 55, 2452-2462.

Fellman J.B., Petrone K.C. & Grierson P.F. (2013) Leaf litter age, chemical quality, and

photodegradation control the fate of leachate dissolved organic matter in a

dryland river. Journal of Arid Environments, 89, 30-37.

Fellman J.B., Spencer R.G.M., Raymond P.A., Pettit N.E., Skrzypek G., Hernes P.J. &

Grierson P.F. (2014) Dissolved organic carbon biolability decreases along with

its modernization in fluvial networks in an ancient landscape. Ecology, 95,

2622-2632.

Fellman J.B., Hood E., Raymond P.A., Hudson J., Bozeman M. & Arimitsu M. (2015)

Evidence for the assimilation of ancient glacier organic carbon in a proglacial

stream food web. Limnology and Oceanography, 60, 1118-1128.

Fellows C.S., Wos M.L., Pollard P.C. & Bunn S.E. (2007) Ecosystem metabolism in a

dryland river waterhole. Marine and Freshwater Research, 58, 250-262.

Fellows C.S., Bunn S.E., Sheldon F. & Beard N.J. (2009) Benthic metabolism in two

turbid dryland rivers. Freshwater Biology, 54, 236-253.

Findlay S., Tank J., Dye S., Valett H.M., Mulholland P.J., Mcdowell W.H., Johnson

S.L., Hamilton S.K. et al. (2002) A cross-system comparison of bacterial and

fungal biomass in detritus pools of headwater streams. Microbial Ecology, 43,

55-66.

Findlay S. (2003) Bacterial response to variation in dissolved organic matter. In:

Aquatic Ecosystems: Interactivity of Dissolved Organic Matter. (Eds S. Findlay

& R.L. Sinsabaugh), pp. 363-379. Academic Press, San Diego.

Finlay J. (2003) Controls of streamwater dissolved inorganic carbon dynamics in a

forested watershed. Biogeochemistry, 62, 231-252.

Finlay J.C. & Kendall C. (2007) Stable isotope tracing of temporal and spatial

variability in organic matter sources to freshwater ecosystems. In: Stable

Isotopes in Ecology and Environmental Science. (Eds R. Michener & K. Lajtha),

pp. 283-333. Blackwell Publishing, Oxford.

Fisher S.G., Gray L.J., Grimm N.B. & Busch D.E. (1982) Temporal succession in a

desert stream ecosystem following flash flooding. Ecological Monographs, 52,

93-110.

Page 128: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

118

Fisher S.G., Grimm N.B., Marti E. & Gómez R. (1998a) Hierarchy, spatial

configuration, and nutrient cycling in a desert stream. Australian Journal of

Ecology, 23, 41-52.

Fisher S.G., Grimm N.B., Martí E., Holmes R.M. & Jones J.J.B. (1998b) Material

spiraling in stream corridors: a telescoping ecosystem model. Ecosystems, 1, 19-

34.

Fisher S.G., Welter J., Schade J. & Henry J. (2001) Landscape challenges to ecosystem

thinking: creative flood and drought in the american southwest. Scientia Marina,

65, 181-192.

Fisher S.G., Sponseller R.A. & Heffernan J.B. (2004) Horizons in stream

biogeochemistry: flowpaths to progress. Ecology, 85, 2369-2379.

Fontaine S., Barot S., Barre P., Bdioui N., Mary B. & Rumpel C. (2007) Stability of

organic carbon in deep soil layers controlled by fresh carbon supply. Nature,

450, 277-280.

Ford D.J., Cookson W.R., Adams M.A. & Grierson P.F. (2007) Role of soil drying in

nitrogen mineralization and microbial community function in semi-arid

grasslands of north-west Australia. Soil Biology and Biochemistry, 39, 1557-

1569.

Freixa A., Rubol S., Carles-Brangarí A., Fernàndez-Garcia D., Butturini A., Sanchez-

Vila X. & Romaní A.M. (2015) The effects of sediment depth and oxygen

concentration on the use of organic matter: an experimental study using an

infiltration sediment tank. Science of The Total Environment,

doi:10.1016/j.scitotenv.2015.04.007.

Fritz P. & Poplawski S. (1974) 18O and 13C in the shells of freshwater molluscs and

their environments. Earth and Planetary Science Letters, 24, 91-98.

Gat J.R. (1996) Oxygen and hydrogen isotopes in the hydrologic cycle. Annual Review

of Earth and Planetary Sciences, 24, 225-262.

Gerard G. & Chanton J. (1993) Quantification of methane oxidation in the rhizosphere

of emergent aquatic macrophytes: defining upper limits. Biogeochemistry, 23,

79-97.

Gette-Bouvarot M., Mermillod-Blondin F., Lemoine D., Delolme C., Danjean M.,

Etienne L. & Volatier L. (2015) The potential control of benthic biofilm growth

by macrophytes—a mesocosm approach. Ecological Engineering, 75, 178-186.

Page 129: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

119

Gonfiantini R. (1986) Environmental isotopes in lake studies. In: Handbook of

Environmental Isotope Geochemistry Vol. 2: The Terrestrial Environment. (Ed

P. Fritz), pp. 113-163. Elsevier, Amsterdam.

Green S.A. & Blough N.B. (1994) Optical absorption and fluorescence properties of

chromophoric dissolved organic matter in natural waters. Limnology and

Oceanography, 39, 1903-1916.

Grimm N. & Fisher S. (1984) Exchange between interstitial and surface water:

implications for stream metabolism and nutrient cycling. Hydrobiologia, 111,

219-228.

Grimm N.B. & Fisher S.G. (1986) Nitrogen limitation in a Sonoran Desert stream.

Journal of the North American Benthological Society, 5, 2-15.

Grimm N.B. & Fisher S.G. (1989) Stability of periphyton and macroinvertebrates to

disturbance by flash floods in a desert stream. Journal of the North American

Benthological Society, 8, 293-307.

Grimm N.B. & Petrone K.C. (1997) Nitrogen fixation in a desert stream ecosystem.

Biogeochemistry, 37, 33-61.

Guenet B., Danger M., Abbadie L. & Lacroix G. (2010) Priming effect: bridging the

gap between terrestrial and aquatic ecology. Ecology, 91, 2850-2861.

Hamilton S.K., Bunn S.E., Thoms M.C. & Marshall J.C. (2005) Persistence of aquatic

refugia between flow pulses in a dryland river system (Cooper Creek, Australia).

Limnology and Oceanography, 50, 743-754.

Heffernan J.B. (2008) Wetlands as an alternative stable state in desert streams. Ecology,

89, 1261-1271.

Hein T., Baranyi C., Herndl G.J., Wanek W. & Schiemer F. (2003) Allochthonous and

autochthonous particulate organic matter in floodplains of the River Danube: the

importance of hydrological connectivity. Freshwater Biology, 48, 220-232.

Helton A.M., Wright M.S., Bernhardt E.S., Poole G.C., Cory R.M. & Stanford J.A.

(2015) Dissolved organic carbon lability increases with water residence time in

the alluvial aquifer of a river floodplain ecosystem. Journal of Geophysical

Research: Biogeosciences, 120, 2014JG002832.

Hill W. (1996) Effects of light. In: Algal Ecology. (Eds R.J.Sinsabaugh & B.L. Lowe),

pp. 121-148. Academic Press, San Diego.

Hladyz S., Watkins S.C., Whitworth K.L. & Baldwin D.S. (2011) Flows and hypoxic

blackwater events in managed ephemeral river channels. Journal of Hydrology,

401, 117-125.

Page 130: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

120

Hogg A.G., Hua Q., Blackwell P.G., Niu M., Buck C.E., Guilderson T.P., Heaton T.J.,

Palmer J.G. et al. (2013) SHCal13 Southern Hemisphere calibration, 0–50,000

years cal BP. Radiocarbon, 53, 1889-1903.

Holgerson M. (2015) Drivers of carbon dioxide and methane supersaturation in small,

temporary ponds. Biogeochemistry, 124, 305-318.

Hornberger G.M. & Kelly M.G. (1975) Atmospheric reaeration in a river using

productivity analysis. Journal of the Environmental Engineering Division of the

American Society of Civil Engineers, 101, 729-739.

Hornberger J.M., Raffensperger J.P., Wiberg P.L. & Eshleman K.N. (1998) Elements of

Physical Hydrology. Johns Hopkins University Press, Baltimore.

Howarth R.W., Marino R. & Cole J.J. (1988) Nitrogen fixation in freshwater, estuarine,

and marine ecosystems. 2. Biogeochemical controls. Limnology and

Oceanography, 33, 688-701.

Howitt J.A., Baldwin D.S., Rees G.N. & Hart B.T. (2008) Photodegradation, interaction

with iron oxides and bioavailability of dissolved organic matter from forested

floodplain sources. Marine and Freshwater Research, 59, 780-791.

Hudson N., Baker A. & Reynolds D. (2007) Fluorescence analysis of dissolved organic

matter in natural, waste and polluted waters—a review. River Research and

Applications, 23, 631-649.

Humphries P., Keckeis H. & Finlayson B. (2014) The river wave concept: integrating

river ecosystem models. BioScience, 64, 870-882.

Hunt R.J., Jardine T.D., Hamilton S.K. & Bunn S.E. (2012) Temporal and spatial

variation in ecosystem metabolism and food web carbon transfer in a wet-dry

tropical river. Freshwater Biology, 57, 435-450.

Jacobson P.J., Jacobson K.M., Angermeier P.L. & Cherry D.S. (2000) Variation in

material transport and water chemistry along a large ephemeral river in the

Namib Desert. Freshwater Biology, 44, 481-491.

Jaffé R., McKnight D., Maie N., Cory R., McDowell W.H. & Campbell J.L. (2008)

Spatial and temporal variations in DOM composition in ecosystems: the

importance of long-term monitoring of optical properties. Journal of

Geophysical Research: Biogeosciences, 113, G04032.

Jansson M., Persson L., De Roos A.M., Jones R.I. & Tranvik L.J. (2007) Terrestrial

carbon and intraspecific size-variation shape lake ecosystems. Trends in Ecology

& Evolution, 22, 316-322.

Page 131: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

121

Jones E.J.P., Voytek M.A., Warwick P.D., Corum M.D., Cohn A., Bunnell J.E., Clark

A.C. & Orem W.H. (2008a) Bioassay for estimating the biogenic methane-

generating potential of coal samples. International Journal of Coal Geology, 1-

2, 138-150.

Jones J.B. (1995) Factors controlling hyporheic respiration in a desert stream.

Freshwater Biology, 34, 91-99.

Jones J.B., Jr., Fisher S.G. & Grimm N.B. (1995a) Nitrification in the hyporheic zone of

a desert stream ecosystem. Journal of the North American Benthological

Society, 14, 249-258.

Jones J.B., Jr., Fisher S.G. & Grimm N.B. (1995b) Vertical hydrologic exchange and

ecosystem metabolism in a Sonoran Desert stream. Ecology, 76, 942-952.

Jones J., Holmes R., Fisher S., Grimm N. & Greene D. (1995c) Methanogenesis in

Arizona, USA dryland streams. Biogeochemistry, 31, 155-173.

Jones J.B., Jr., Schade J.D., Fisher S.G. & Grimm N.B. (1997) Organic matter dynamics

in Sycamore Creek, a desert stream in Arizona, USA. Journal of the North

American Benthological Society, 16, 78-82.

Jones J.B. (2002) Groundwater controls on nutrient cycling in a Mojave Desert stream.

Freshwater Biology, 47, 971-983.

Jones R.I., Carter C.E., Kelly A., Ward S., Kelly D.J. & Grey J. (2008b) Widespread

contribution of methane-cycle bacteria to the diets of lake profundal chironomid

larvae. Ecology, 89, 857-864.

Jones R.I., Grey J., Sleep D. & Arvola L. (1999) Stable isotope analysis of zooplankton

carbon nutrition in humic lakes. Oikos, 86, 97-104.

Jones R.I. & Grey J. (2011) Biogenic methane in freshwater food webs. Freshwater

Biology, 56, 213-229.

Junk J.W., Bayley P.B. & Sparks R.E. (1989) The flood pulse concept in river-

floodplain systems. In: Proceedings of the International Large River

Symposium. (Ed D.P. Dodge), pp. 110-127. Canadian Special Publications of

Fisheries and Aquatic Sciences.

Karlsson J., Bystrom P., Ask J., Ask P., Persson L. & Jansson M. (2009) Light

limitation of nutrient-poor lake ecosystems. Nature, 460, 506-509.

Kassen R. (2002) The experimental evolution of specialists, generalists, and the

maintenance of diversity. Journal of Evolutionary Biology, 15, 173-190.

Keeling R.F., Piper S.C., Bollenbacher A.F. & Walker S.J. (2010) Monthly atmospheric 13C/12C isotopic ratios for 11 SIO stations. In: Trends: A Compendium of Data

Page 132: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

122

on Global Change. (Eds. Carbon Dioxide Information Analysis Center, Oak

Ridge National Laboratory, U.S. Department of Energy), Oak Ridge, Tenn.,

U.S.A.

Kellerman A.M., Kothawala D.N., Dittmar T. & Tranvik L.J. (2015) Persistence of

dissolved organic matter in lakes related to its molecular characteristics. Nature

Geoscience, 8, 454-457.

Kendall C. & Mcdonnell J.J. (2012) Isotope Tracers in Catchment Hydrology. Elsevier.

Kennard M.J., Pusey B.J., Olden J.D., Mackay S.J., Stein J.L. & Marsh N. (2010)

Classification of natural flow regimes in Australia to support environmental

flow management. Freshwater Biology, 55, 171-193.

Kip N., Van Winden J.F., Pan Y., Bodrossy L., Reichart G.-J., Smolders A.J.P., Jetten

M.S.M., Damste J.S.S. & Op Den Camp H.J.M. (2010) Global prevalence of

methane oxidation by symbiotic bacteria in peat-moss ecosystems. Nature

Geoscience, 3, 617-621.

Knighton A.D. & Nanson G.C. (1997) Distinctiveness, diversity and uniqueness in arid

zone river systems. In: Arid Zone Geomorphology: Process, Form and Change

in Drylands. (Eds D.S.G. Thomas), pp. 185-203. John Wiley and Sons,

Chichester.

Koerselman W. & Meuleman A.F. (1996) The vegetation N: P ratio: a new tool to

detect the nature of nutrient limitation. Journal of Applied Ecology, 1441-1450.

Kufel L. & Kufel I. (2002) Chara beds acting as nutrient sinks in shallow lakes—a

review. Aquatic Botany, 72, 249-260.

Kuzyakov Y. (2002) Review: Factors affecting rhizosphere priming effects. Journal of

Plant Nutrition and Soil Science, 165, 382-396.

Lake P.S. (2003) Ecological effects of perturbation by drought in flowing waters.

Freshwater Biology, 48, 1161-1172.

Langhans S. & Tockner K. (2006) The role of timing, duration, and frequency of

inundation in controlling leaf litter decomposition in a river-floodplain

ecosystem (Tagliamento, northeastern Italy). Oecologia, 147, 501-509.

Larned S.T., Datry T., Arscott D.B. & Tockner K. (2010) Emerging concepts in

temporary-river ecology. Freshwater Biology, 55, 717-738.

Lavender S. & Abbs D. (2013) Trends in Australian rainfall: contribution of tropical

cyclones and closed lows. Climate Dynamics, 40, 317-326.

Page 133: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

123

Layman C.A., Arrington D.A., Montaña C.G. & Post D.M. (2007) Can stable isotope

ratios provide for community-wide measures of trophic structure? Ecology, 88,

42-48.

Leigh C., Burford M.A., Sheldon F. & Bunn S.E. (2010) Dynamic stability in dry

season food webs within tropical floodplain rivers. Marine and Freshwater

Research, 61, 357-368.

Leigh C., Boulton A.J., Courtwright J.L., Fritz K., May C.L., Walker R.H. & Datry T.

(2015) Ecological research and management of intermittent rivers: an historical

review and future directions. Freshwater Biology, DOI: 10.1111/fwb.12646.

Logan J.M., Jardine T.D., Miller T.J., Bunn S.E., Cunjak R.A. & Lutcavage M.E.

(2008) Lipid corrections in carbon and nitrogen stable isotope analyses:

comparison of chemical extraction and modelling methods. Journal of Animal

Ecology, 77, 838-846.

Logue J.B., Stedmon C.A., Kellerman A.M., Nielsen N.J., Andersson A.F., Laudon H.,

Lindstrom E.S. & Kritzberg E.S. (2015) Experimental insights into the

importance of aquatic bacterial community composition to the degradation of

dissolved organic matter. The ISME Journal, doi: 10.1038/ismej.2015.131.

Macauley B.J. & Fox L.R. (1980) Variation in total phenols and condensed tannins in

eucalyptus: leaf phenology and insect grazing. Australian Journal of Ecology, 5,

31-35.

Magaña H. A. (2013). Flood pulse trophic dynamics of larval fishes in a restored arid-

land, river-floodplain, Middle Rio Grande, Los Lunas, New Mexico. Reviews in

Fish Biology and Fisheries, 23, 507-521.

Maie N., Scully N.M., Pisani O. & Jaffé R. (2007) Composition of a protein-like

fluorophore of dissolved organic matter in coastal wetland and estuarine

ecosystems. Water Research, 41, 563-570.

Maranger R. & Pullin M.J. (2003) Elemental complexation by dissolved organic matter

in lakes: Implications for Fe speciation and the bioavailability of Fe and P. In:

Aquatic Ecosystems: Interactivity of Dissolved Organic Matter. (Eds S. Findlay

& R.L. Sinsabaugh), pp. 185-214. Academic Press, San Diego.

Marotta H., Duarte C.M., Sobek S. & Enrich-Prast A. (2009) Large CO2 disequilibria in

tropical lakes. Global Biogeochemical Cycles, 23, GB4022.

Marti E., Fisher S.G., Schade J.D. & Grimm N.B. (2000) Flood frequency and stream-

riparian linkages in arid lands. In: Streams and Ground waters. (Eds B.J. Jeremy

& J.M. Patrick), pp. 111-136. Academic Press, San Diego.

Page 134: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

124

McCallister S.L. & Giorgio P.A.D. (2008) Direct measurement of the δ13C signature of

carbon respired by bacteria in lakes: linkages to potential carbon sources,

ecosystem baseline metabolism, and CO2 fluxes. Limnology and Oceanography,

53, 1204-1216.

McCallister S.L. & Del Giorgio P.A. (2012) Evidence for the respiration of ancient

terrestrial organic C in northern temperate lakes and streams. Proceedings of the

National Academy of Sciences, 109, 16963-16968.

McClain M.E., Boyer E.W., Dent C.L., Gergel S.E., Grimm N.B., Groffman P.M., Hart

S.C., Harvey J.W. et al. (2003) Biogeochemical hot spots and hot moments at

the interface of terrestrial and aquatic ecosystems. Ecosystems, 6, 301-312.

McIntyre R.E.S., Adams M.A., Ford D.J. & Grierson P.F. (2009) Rewetting and litter

addition influence mineralisation and microbial communities in soils from a

semi-arid intermittent stream. Soil Biology and Biochemistry, 41, 92-101.

McKnight D.M., Boyer E.W., Westerhoff P.K., Doran P.T., Kulbe T. & Andersen D.T.

(2001) Spectrofluorometric characterization of dissolved organic matter for

indication of precursor organic material and aromaticity. Limnology and

Oceanography, 46, 38-48.

Middleton N. & Thomas D. (1997) World Atlas of Desertification. Arnold, London.

Miyatake T., Moerdijk-Poortvliet T.C., Stal L.J. & Boschker H.T. (2014) Tracing

carbon flow from microphytobenthos to major bacterial groups in an intertidal

marine sediment by using an in situ 13C pulse‐chase method. Limnology and

Oceanography, 59, 1275-1287.

Mooshammer M., Wanek W., Schnecker J., Wild B., Leitner S., Hofhansl F., Blöchl A.,

Hämmerle I. et al. (2011) Stoichiometric controls of nitrogen and phosphorus

cycling in decomposing beech leaf litter. Ecology, 93, 770-782.

Mulholland P.J., Steinman A.D., Palumbo A.V., Elwood J.W. & Kirschtel D.B. (1991)

Role of nutrient cycling and herbivory in regulating periphyton communities in

laboratory streams. Ecology, 72, 966-982.

Mulholland P.J., Marzolf E.R., Hendricks S.P., Wilkerson R.V. & Baybayan A.K.

(1995) Longitudinal patterns of nutrient cycling and periphyton characteristics in

streams: a test of upstream-downstream linkage. Journal of the North American

Benthological Society, 14, 357-370.

Murphy J. & Riley J.P. (1962) A modified single solution method for the determination

of phosphate in natural waters. Analytica Chimica Acta, 27, 31-36.

Page 135: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

125

Nichols P.D., Glen A S., Antworth C.P., Hanson R.S. & White D.C. (1985)

Phospholipid and lipopolysaccharide normal and hydroxy fatty acids as potential

signatures for methane-oxidizing bacteria. FEMS Microbiology Ecology, 1, 327-

335.

Nimick D.A., Gammons C.H. & Parker S.R. (2011) Diel biogeochemical processes and

their effect on the aqueous chemistry of streams: a review. Chemical Geology,

283, 3-17.

O'Donnell A.J., Cook E.R., Palmer J.G., Turney C.S.M., Page G.F.M. & Grierson P.F.

(2015) Tree rings show recent high summer-autumn precipitation in northwest

Australia is unprecedented within the last two centuries. PLoS ONE, 10,

e0128533.

Odum H.T. (1956) Primary production in flowing waters. Limnology and

Oceanography, 1, 102-117.

Parker S.R., Poulson S.R., Gammons C.H. & Degrandpre M.D. (2005) Biogeochemical

controls on diel cycling of stable isotopes of dissolved O2 and dissolved

inorganic carbon in the Big Hole River, Montana. Environmental Science &

Technology, 39, 7134-7140.

Parker S.R., Gammons C.H., Poulson S.R., Degrandpre M.D., Weyer C.L., Smith M.G.,

Babcock J.N. & Oba Y. (2010) Diel behavior of stable isotopes of dissolved

oxygen and dissolved inorganic carbon in rivers over a range of trophic

conditions, and in a mesocosm experiment. Chemical Geology, 269, 22-32.

Parnell A.C., Inger R., Bearhop S. & Jackson A.L. (2010) Source partitioning using

stable isotopes: coping with too much variation. PLoS ONE, 5, e9672.

Perkins A., Santos I., Sadat-Noori M., Gatland J. & Maher D. (2015) Groundwater

seepage as a driver of CO2 evasion in a coastal lake (Lake Ainsworth, NSW,

Australia). Environmental Earth Sciences, 74, 779-792.

Peterson B.J. & Fry B. (1987) Stable isotopes in ecosystem studies. Annual Review of

Ecology and Systematics, 18, 293-320.

Pettit N.E., Latterell J.J. & Naiman R.J. (2006) Formation, distribution and ecological

consequences of flood-related wood debris piles in a bedrock confined river in

semi-arid South Africa. River Research and Applications, 22, 1097-1110.

Pettit N.E., Bayliss P., Davies P.M., Hamilton S.K., Warfe D.M., Bunn S.E. & Douglas

M.M. (2011) Seasonal contrasts in carbon resources and ecological processes on

a tropical floodplain. Freshwater Biology, 56, 1047-1064.

Page 136: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

126

Pillans B. (2007) Pre-quaternary landscape inheritance in Australia. Journal of

Quaternary Science, 22, 439-447.

Pinder A.M., Halse S.A., Shiel R.J. & Mcrae J.M. (2010) An arid zone awash with

diversity: patterns in the distribution of aquatic invertebrates in the Pilbara

region of Western Australia. Records of the Western Australian Museum, 78,

205-246.

Poole G.C., Stanford J.A., Running S.W. & Frissell C.A. (2006) Multiscale geomorphic

drivers of groundwater flow paths: subsurface hydrologic dynamics and

hyporheic habitat diversity. Journal of the North American Benthological

Society, 25, 288-303.

Poos M.S. & Jackson D.A. (2012) Addressing the removal of rare species in

multivariate bioassessments: the impact of methodological choices. Ecological

Indicators, 18, 82-90.

Post D.M. (2002a) The long and short of food-chain length. Trends in Ecology &

Evolution, 17, 269-277.

Post D.M. (2002b) Using stable isotopes to estimate trophic position: models, methods,

and assumptions. Ecology, 83, 703-718.

Post D., Layman C., Arrington D.A., Takimoto G., Quattrochi J. & Montaña C. (2007)

Getting to the fat of the matter: models, methods and assumptions for dealing

with lipids in stable isotope analyses. Oecologia, 152, 179-189.

Puckridge J.T., Sheldon F., Walker K.F. & Boulton A.J. (1998) Flow variability and the

ecology of large rivers. Marine and Freshwater Research, 49, 55-72.

Pusey B.J., Arthington A.H., Stewart-Koster B., Kennard M.J. & Read M.G. (2010)

Widespread omnivory and low temporal and spatial variation in the diet of

fishes in a hydrologically variable northern Australian river. Journal of Fish

Biology, 77, 731-753.

Reid D.J., Quinn G.P., Lake P.S. & Reich P. (2008) Terrestrial detritus supports the

food webs in lowland intermittent streams of south-eastern Australia: a stable

isotope study. Freshwater Biology, 53, 2036-2050.

Romaní A., Vázquez E. & Butturini A. (2006) Microbial availability and size

fractionation of dissolved organic carbon after drought in an intermittent stream:

biogeochemical link across the stream–riparian interface. Microbial Ecology, 52,

501-512.

Rouillard A., Skrzypek G., Dogramaci S., Turney C. & Grierson P.F. (2015) Impacts of

high inter-annual variability of rainfall on a century of extreme hydrologic

Page 137: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

127

regime of northwest australia. Hydrology and Earth System Sciences, 19, 2057-

2078.

Sahuquillo M., Miracle M.R., Morata S.M. & Vicente E. (2012) Nutrient dynamics in

water and sediment of mediterranean ponds across a wide hydroperiod gradient.

Limnologica - Ecology and Management of Inland Waters, 42, 282-290.

Sardans J., Rivas-Ubach A. & Peñuelas J. (2012) The elemental stoichiometry of

aquatic and terrestrial ecosystems and its relationships with organismic lifestyle

and ecosystem structure and function: a review and perspectives.

Biogeochemistry, 111, 1-39.

Scheffer M. & Nes E.H. (2007) Shallow lakes theory revisited: Various alternative

regimes driven by climate, nutrients, depth and lake size. In: Shallow Lakes in a

Changing World. (Eds R.D. Gulati & E. Lammens & N. Pauw & E. Donk), pp.

455-466. Springer Netherlands.

Schlesinger W.H. & Melack J.M. (1981) Transport of organic carbon in the world's

rivers. Tellus, 33, 172-187.

Scott J.T., Back J.A., Taylor J.M. & King R.S. (2008) Does nutrient enrichment

decouple algal–bacterial production in periphyton? Journal of the North

American Benthological Society, 27, 332-344.

Segers R. (1998) Methane production and methane consumption: a review of processes

underlying wetland methane fluxes. Biogeochemistry, 41, 23-51.

Sheldon F. & Fellows C.S. (2010) Water quality in two Australian dryland rivers:

spatial and temporal variability and the role of flow. Marine and Freshwater

Research, 61, 864-874.

Shen Y., Chapelle F.H., Strom E.W. & Benner R. (2015) Origins and bioavailability of

dissolved organic matter in groundwater. Biogeochemistry, 122, 61-78.

Shurin J.B., Gruner D.S. & Hillebrand H. (2006) All wet or dried up? Real differences

between aquatic and terrestrial food webs. Proceedings of the Royal Society B:

Biological Sciences, 273, 1-9.

Sinsabaugh R.L. & Findlay S. (2003) Dissolved organic matter: out of the black box

and into the mainstream. In: Aquatic Ecosystems: Interactivity of Dissolved

Organic Matter. (Eds S. Findlay & R.L. Sinsabaugh), pp. 479-498. Academic

Press, San Diego.

Skrzypek G. (2013) Normalization procedures and reference material selection in stable

hcnos isotope analyses: an overview. Analytical and Bioanalytical Chemistry,

405, 2815-2823.

Page 138: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

128

Skrzypek G., Dogramaci S. & Grierson P.F. (2013) Geochemical and hydrological

processes controlling groundwater salinity of a large inland wetland of

northwest Australia. Chemical Geology, 357, 164-177.

Skrzypek G., Mydłowski A., Dogramaci S., Hedley P., Gibson J.J. & Grierson P.F.

(2015) Estimation of evaporative loss based on the stable isotope composition of

water using Hydrocalculator. Journal of Hydrology, 523, 781-789.

Smith S. & Hollibaugh J. (1997) Annual cycle and interannual variability of ecosystem

metabolism in a temperate climate embayment. Ecological Monographs, 67,

509-533.

Sobczak W.V. & Findlay S. (2002) Variation in bioavailability of dissolved organic

carbon among stream hyporheic flowpaths. Ecology, 83, 3194-3209.

Sobek S., Tranvik L.J. & Cole J.J. (2005) Temperature independence of carbon dioxide

supersaturation in global lakes. Global Biogeochemical Cycles, 19, GB2003.

Solorzano L. & Sharp J.H. (1980) Determination of total dissolved phosphorus and

particulate phosphorus in natural waters. Limnology and Oceanography, 25,

754-758.

Stanley E.H., Buschman D.L., Boulton A.J., Grimm N.B. & Fisher S.G. (1994)

Invertebrate resistance and resilience to intermittency in a desert stream.

American Midland Naturalist, 131, 288-300.

Stanley E.H., Fisher S.G. & Grimm N.B. (1997) Ecosystem expansion and contraction

in streams. BioScience, 47, 427-435.

Stedmon C.A., Markager S. & Bro R. (2003) Tracing dissolved organic matter in

aquatic environments using a new approach to fluorescence spectroscopy.

Marine Chemistry, 82, 239-254.

Stedmon C.A. & Bro R. (2008) Characterizing dissolved organic matter fluorescence

with parallel factor analysis: a tutorial. Limnology and Oceanography: Methods,

6, 572-579.

Steinberg C.E.W., Kamara S., Prokhotskaya V.Y., Manusadžianas L., Karasyova T.A.,

Timofeyev M.A., Jie Z., Paul A. et al. (2006) Dissolved humic substances –

ecological driving forces from the individual to the ecosystem level? Freshwater

Biology, 51, 1189-1210.

Sterner, Clasen, Lampert & Weisse. (1998) Carbon:Phosphorus stoichiometry and food

chain production. Ecology Letters, 1, 146-150.

Sterner R.W. & Elser J.J. (2002) Ecological Stoichiometry: The Biology of Elements

from Molecules to the Biosphere. Princeton University Press, Princeton.

Page 139: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

129

Stromberg J.C., Beauchamp V.B., Dixon M.D., Lite S.J. & Paradzick C. (2007)

Importance of low-flow and high-flow characteristics to restoration of riparian

vegetation along rivers in arid south-western United States. Freshwater Biology,

52, 651-679.

Stubbins A., Lapierre J.F., Berggren M., Prairie Y.T., Dittmar T. & Del Giorgio P.A.

(2014) What’s in an EEM? Molecular signatures associated with dissolved

organic fluorescence in Boreal Canada. Environmental Science & Technology,

48, 10598-10606.

Tank J.L., Rosi-Marshall E.J., Griffiths N.A., Entrekin S.A. & Stephen M.L. (2010) A

review of allochthonous organic matter dynamics and metabolism in streams.

Journal of the North American Benthological Society, 29, 118-146.

Thorp J.H. & Delong M.D. (2002) Dominance of autochthonous autotrophic carbon in

food webs of heterotrophic rivers. Oikos, 96, 543-550.

Thorp J.H., Thoms M.C. & Delong M.D. (2006) The Riverine Ecosystem Synthesis:

biocomplexity in river networks across space and time. River Research and

Applications, 22, 123-147.

Thygesen L.G., Rinnan Å., Barsberg S. & Møller J.K.S. (2004) Stabilizing the

PARAFAC decomposition of fluorescence spectra by insertion of zeros outside

the data area. Chemometrics and Intelligent Laboratory Systems, 71, 97-106.

Thyssen N., Erlandsen M., Jeppesen E. & Holm T.F. (1983) Modelling the reaeration

capacity of low-land streams dominated by submerged macrophytes. In:

Analysis of Ecological Systems: State of the Art in Ecological Modelling. (Eds

W.K. Lauenroth & V. Skogerboe & M. Flug), pp. 861-867. Elsevier,

Amsterdam, The Netherlands.

Tobias C.R., Böhlke J.K. & Judson W.H. (2007) The oxygen-18 isotope approach for

measuring aquatic metabolism in high-productivity waters. Limnology and

Oceanography, 52, 1439-1453.

Tobias C. & Böhlke J.K. (2011) Biological and geochemical controls on diel dissolved

inorganic carbon cycling in a low-order agricultural stream: implications for

reach scales and beyond. Chemical Geology, 283, 18-30.

Tooth S. (2000) Process, form and change in dryland rivers: a review of recent research.

Earth-Science Reviews, 51, 67-107.

Townsend S.A. (2002) Seasonal evaporative concentration of an extremely turbid

water-body in the semiarid tropics of Australia. Lakes & Reservoirs: Research

& Management, 7, 103-107.

Page 140: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

130

Townsend S.A., Webster I.T. & Schult J.H. (2011) Metabolism in a groundwater-fed

river system in the Australian wet/dry tropics: tight coupling of photosynthesis

and respiration. Journal of the North American Benthological Society, 30, 603-

620.

Trimmer M., Hildrew A.G., Jackson M.C., Pretty J.L. & Grey J. (2009) Evidence for

the role of methane-derived carbon in a free-flowing, lowland river food web.

Limnology and Oceanography, 54, 1541-1547.

Valett H.M., Fisher S.G., Grimm N.B. & Camill P. (1994) Vertical hydrologic exchange

and ecological stability of a desert stream ecosystem. Ecology, 75, 548-560.

Valett H.M., Morrice J.A., Dahm C.N. & Campana M.E. (1996) Parent lithology,

surface–groundwater exchange, and nitrate retention in headwater streams.

Limnology and Oceanography, 41, 333-345.

Valett H.M., Baker M.A., Morrice J.A., Crawford C.S., Molles M.C., Dahm C.N.,

Moyer D.L., Thibault J.R. & Ellis L.M. (2005) Biogeochemical and metabolic

responses to the flood pulse in a semiarid floodplain. Ecology, 86, 220-234.

Van Donk E. & Van De Bund W.J. (2002) Impact of submerged macrophytes including

charophytes on phyto- and zooplankton communities: allelopathy versus other

mechanisms. Aquatic Botany, 72, 261-274.

Van Vreeswyk A.M.E., Payne A.L., Leighton K.A. & Hennig P. (2004) An inventory

and condition survey of the Pilbara region, Western Australia. Department of

Agriculture, Western Australia.

Vanni M.J. (2002) Nutrient cycling by animals in freshwater ecosystems. Annual

Review of Ecology and Systematics, 33, 341-370.

Vannote R.L., Minshall G.W., Cummins K.W., Sedell J.R. & Cushing C.E. (1980) The

River Continuum Concept. Canadian Journal of Fisheries and Aquatic Sciences,

37, 130-137.

Vazquez E., Amalfitano S., Fazi S. & Butturini A. (2011) Dissolved organic matter

composition in a fragmented mediterranean fluvial system under severe drought

conditions. Biogeochemistry, 102, 59-72.

Von Schiller D., Acuña V., Graeber D., Martí E., Ribot M., Sabater S., Timoner X. &

Tockner K. (2011) Contraction, fragmentation and expansion dynamics

determine nutrient availability in a mediterranean forest stream. Aquatic

Sciences - Research Across Boundaries, 73, 485-497.

Von Schiller D., Graeber D., Ribot M., Timoner X., Acuña V., Martí E., Sabater S. &

Tockner K. (2015) Hydrological transitions drive dissolved organic matter

Page 141: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

131

quantity and composition in a temporary mediterranean stream.

Biogeochemistry, 1-18.

Waldron S., Scott E.M. & Soulsby C. (2007) Stable isotope analysis reveals lower-order

river dissolved inorganic carbon pools are highly dynamic. Environmental

Science & Technology, 41, 6156-6162.

Warfe D.M., Barmuta L.A. & Wotherspoon S. (2008) Quantifying habitat structure:

surface convolution and living space for species in complex environments.

Oikos, 117, 1764-1773.

Warfe D.M., Pettit N.E., Magierowski R.H., Pusey B.J., Davies P.M., Douglas M.M. &

Bunn S.E. (2013) Hydrological connectivity structures concordant plant and

animal assemblages according to niche rather than dispersal processes.

Freshwater Biology, 58, 292-305.

Webster J.R., Benfield E.F., Ehrman T.P., Schaeffer M.A., Tank J.L., Hutchens J.J. &

D’angelo D.J. (1999) What happens to allochthonous material that falls into

streams? A synthesis of new and published information from Coweeta.

Freshwater Biology, 41, 687-705.

Webster I.T., Rea N., Padovan A.V., Dostine P., Townsend S.A. & Cook S. (2005) An

analysis of primary production in the Daly River, a relatively unimpacted

tropical river in northern Australia. Marine and Freshwater Research, 56, 303-

316.

Webster J.R. (2007) Spiraling down the river continuum: stream ecology and the U-

shaped curve. Journal of the North American Benthological Society, 26, 375-

389.

Webster J.R. & Meyer J.L. (1997) Organic matter budgets for streams: a synthesis.

Journal of the North American Benthological Society, 16, 141-161.

Welborn G.A., Skelly D.K. & Werner E.E. (1996) Mechanisms creating community

structure across a freshwater habitat gradient. Annual Review of Ecology and

Systematics, 27, 337-363.

Wetzel R.G. (2001) Limnology: Lake and River Ecosystems. Academic Press, San

Diego.

Whitford W.G. (2006) Ecology of Desert Systems. Elsevier, San Diego.

Wiegner T.N., Kaplan L.A., Newbold J.D. & Ostrom P.H. (2005) Contribution of

dissolved organic C to stream metabolism: a mesocosm study using 13C-

enriched tree-tissue leachate. Journal of the North American Benthological

Society, 24, 48-67.

Page 142: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

132

Wood D.J., Fisher S.G. & Grimm N.B. (1992) Pools in desert streams: limnology and

response to disturbance. Journal of the Arizona-Nevada Academy of Science, 26,

171-179.

Wyatt K.H., Tellez E., Woodke R.L., Bidner R.J. & Davison I.R. (2014) Effects of

nutrient limitation on the release and use of dissolved organic carbon from

benthic algae in Lake Michigan. Freshwater Science, 33, 557-567.

Yamashita Y. & Tanoue E. (2004) Chemical characteristics of amino acid-containing

dissolved organic matter in seawater. Organic Geochemistry, 35, 679-692.

Yuan F. & Miyamoto S. (2008) Characteristics of oxygen-18 and deuterium

composition in waters from the Pecos River in American southwest. Chemical

Geology, 255, 220-230.

Zektser S., Loáiciga H.A. & Wolf J.T. (2005) Environmental impacts of groundwater

overdraft: selected case studies in the southwestern United States.

Environmental Geology, 47, 396-404.

Ziegler J.P., Solomon C.T., Finney B.P. & Gregory-Eaves I. (2015) Macrophyte

biomass predicts food chain length in shallow lakes. Ecosphere, 6, art5.

Page 143: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

133

Page 144: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

134

Page 145: Organic matter dynamics of hydrologically variable pools ... · investigation of a single catchment indicated that changes in pool volume due to evaporation and contraction can control

135

APPENDIX A: Stable isotope analysis of fine particulate organic

matter and dissolved organic matter within Coondiner Creek

Table A-1: δ13C values and estimated radiocarbon ages (derived from δ14C values;

calibrated using SHCal 13 (Hogg et al., 2013) at >95 % probability using OxCal 4.2,

Oxford Radiocarbon Accelerator Unit, Oxford) for fine particulate organic matter

(FPOM) and dissolved organic matter (DOM) within pools 14 and 16 of Coondiner

Creek, October 2010 (unpub. data).

δ13C (‰) Probable date of origin

(radiocarbon age)

Percent modern carbon (%)

FPOM Pool 14 -30.2 387 A.D. to 574 A.D. 81.7 (± 0.4)

Pool 16 -30.2 1396 A.D. to 1645 A.D. 94.4 (± 0.8)

DOC Pool 14 -40.1 8740 B.C. to 8236 B.C. 31.2 (± 0.3)

Pool 16 -40.1 1383 B.C. to 1308 B.C. 69.8 (± 1.0)