Groundwater-Surface Water Interactions: Implications for ... › ... › 2 ›...

299
Groundwater-Surface Water Interactions: Implications for Nutrient Transport to Tropical Rivers Prachi Dixon-Jain May 2008 A thesis submitted for the degree of Doctor of Philosophy of The Australian National University

Transcript of Groundwater-Surface Water Interactions: Implications for ... › ... › 2 ›...

Page 1: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Groundwater-Surface Water Interactions:

Implications for Nutrient Transport to Tropical Rivers

Prachi Dixon-Jain

May 2008

A thesis submitted for the degree of Doctor of Philosophy of

The Australian National University

Page 2: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to
Page 3: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

iii

I certify that this thesis does not incorporate without acknowledgment any material previously

published. This work has not previously been submitted for a degree or diploma in any

institution of higher education.

Prachi Dixon-Jain

May 2008

Page 4: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to
Page 5: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

v

Acknowledgements

The journey of a PhD is filled with moments of excitement, frustration, uncertainty, satisfaction

and relief. I dearly thank my family and friends for their encouragement during all of these

phases. In particular, I thank my husband Stephen, as well as Mum, Dad and Jij for their faith in

my abilities and unfailing support to see me through to the end.

I would like to acknowledge my supervisors Professor Tony Jakeman, Dr Rebecca Letcher, Dr

Barry Croke (ANU), Dr Richard Cresswell (CSIRO Land and Water) and Dr John Sims (Bureau

of Rural Sciences) for their contributions to different aspects of my research. I’d particularly

like to thank Rebecca for her guidance, flexibility and encouragement. I am also grateful to

Barry for his technical expertise and willingness to act as a sounding board to work through

analytical problems. Support from Richard has been invaluable for guiding me through the

hydrochemical aspects of my research. I especially thank Richard for agreeing to provide

supervision at an advanced stage in the project. Thanks also to John for his encouragement to

pursue a PhD in the first place and supporting me with the appropriate arrangements at BRS to

make it a reality.

I acknowledge Ray Evans (Salient Solutions) for helpful discussions in the initial phases of my

research. I’d also like to thank Dirk Kirste and Bear McPhail (ANU) for sharing their

hydrochemical expertise. Furthermore, I am grateful to Ian White (ANU), Jon Olley (CSIRO

Land and Water) and Peter Baker (BRS) for their interest and enthusiasm in my project.

I am extremely grateful to John Spring and Grahaem Chiles (BRS) for their professional

technical assistance in the field. Thanks also to John Charles (QDNRW) for his assistance

during a reconnaissance of the study area. Laboratory analyses were undertaken by Aleksandra

Plazinska (BRS), John Pengelly (Murray Darling Freshwater Research Centre), staff at CSIRO

Land and Water Laboratories (South Australia) and ECOWISE Environmental, to whom I am

grateful. In particular I’d like to thank Fred Leaney and Megan LeFourner for their assistance

with radon sampling.

The invaluable technical advice of QDNRW staff, especially Ray McGowan, Ian Baker, Geoff

Pocock and Phil Kerr is gratefully acknowledged. For logistical support and supply of

subsidiary data I also thank Peter Martin and Glen Romano (Hinchinbrook Shire Council);

Andrew Wood and Tony Marino (CSR); Anna Forrest and Raymond De Lai (Herbert Resource

Information Centre); Anthony McLoughlin and Ed Stephens (QDNRW); David Post, Peter

Fitch and John Dighton (CSIRO Land and Water); Caroline Coppo (Herbert River Catchment

Page 6: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

vi

Group), Rod Collins (National Parks); Greg Shannon (Bureau of Sugar Experiment Stations);

and Ron Kerkwyk (Herbert Cane Productivity Services).

This research would not have been possible without the kind support of members of the local

community in the catchment. I really appreciate their generous hospitality and willingness to

help out where they could. I’d especially like to acknowledge the landholders who assisted with

sampling and allowed access to their properties, including Margaret and Doug Matthews, Vince

Vitale, Tony Palmas, Michael and Lynn Cristaudo, John Gollogly, Ian Morley, John and Pam

Schmidt, and Trevor Pallanza. In addition, I extend my gratitude to Norm and Karyn Bliesner

for patiently searching through drilling records and advising me on access to key sampling sites.

I would like to acknowledge the support and friendship of the students that have accompanied

me at various stages of the PhD journey, particularly Karen Ivkovic, Beth Rickwood, Celina

Smith, Amir Sadoddin, Birte Schoettker, John Drewry, Wendy Welsh, Geoff Adams and Sue

Powell. Thanks to Karen and Wendy for their feedback on chapter 4. A special thanks to Karen,

who has not only been a colleague and a friend, but a valuable mentor.

Other technical support was provided by various staff at the ANU. I acknowledge Jason

Sharples for his assistance with manipulating the groundwater database, Paul Sjoberg for his

help with scanning cross-sections, Debbie Claridge and Clive Hilliker for assistance with

graphics, and Karl Nissen and Steve Leahy for IT support. I also thank Sue Kelo for her caring

nature and administrative support.

This research was jointly funded by the Australian National University and the Bureau of Rural

Sciences, to whom I extend my gratitude. I especially thank BRS for their generous funding

towards field work and for allowing flexible work arrangements.

Page 7: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

vii

Abstract

The interaction between groundwater and surface water systems is a key component of the

hydrological cycle and an understanding of their connectivity is fundamental for sustainable

water resource management. Water is a vehicle for mobilising dissolved constituents, including

nutrients, between surface and subsurface waters and between terrestrial and marine systems.

Therefore, knowledge of surface-subsurface linkages is critical not only for water quantity

allocation, but also for water quality and its implications for ecosystem health. In particular,

ascertaining the significance of groundwater fluxes for river nitrogen budgets is an important

motivation for characterising river-groundwater connectivity. This overarching theme is

developed through the course of the thesis.

The marked seasonality of tropical river systems provides a unique opportunity to investigate

groundwater contributions to surface waters, especially when there are minimal overland flows.

The Herbert River in northeast Queensland represents a useful case study in the Australian

tropics for assessing the potential for transport of agricultural contaminants, such as dissolved

forms of nitrogen, between surface and subsurface waters, and between terrestrial and marine

systems, including the ecologically significant Great Barrier Reef World Heritage Area. Whilst

the lower Herbert River catchment, dominated by sugarcane production, is the focus for this

thesis, the research methodology and policy implications for nutrient monitoring and

management are applicable to other tropical catchments.

An extensive water quality sampling program was instigated to collect river and groundwater

samples during low flow conditions, for analysis of a range of conservative and non-

conservative environmental tracers including major ions, stable isotopes of water, radon, and

dissolved inorganic forms of nitrogen. Grab samples were collected during months representing

the beginning and end of the dry season to compare connectivity relationships at contrasting

stages of the stream hydrograph. Hydrochemical data at the end of the dry season is particularly

useful for isolating the groundwater signal in the river and its tributaries. Existing physical and

chemical datasets are also an important source of high temporal resolution information to

supplement the more detailed water quality data collected specifically for this investigation.

An understanding of the dynamics of water movement between river and aquifer storages is

critical for assessing the mobility of dissolved nitrogen between them. A combination of

hydrogeological, hydrometric, hydrological and hydrochemical tools are applied to characterise

the interaction between the alluvial aquifers and the lower Herbert River at a catchment scale.

Page 8: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

viii

Specifically, the potential for hydraulic connection and the direction of flux between the aquifer

system and the river are evaluated through qualitative hydrometric approaches, including: depth

relationships of the river channel with that of the underlying alluvial sediments; historical

groundwater elevation-stream stage relationships; and groundwater flow patterns around the

river. Hydrological techniques such as stream hydrograph and flow duration curve analysis are

utilised to assess the temporal characteristics of flow in the river; the groundwater flux to the

river is also quantified by hydrograph separation. Physical understanding of river-aquifer

linkages is verified and enriched through analysis of surface water chemistry data, in

conjunction with the conceptual hydrogeological model developed from physical and chemical

assessment of the aquifers. The significance of groundwater as a vector for nitrogen is then

evaluated in light of a conceptual process understanding of the river-aquifer system. This

provides a platform for undertaking future catchment-scale nutrient budget studies based on

detailed investigations of nitrogen sources and transformations.

The research approach used in this thesis highlights the value of combining analytical

techniques, not provided by any one method, to inform and verify different aspects of a complex

water resource problem involving both surface and groundwater systems. The application of

multiple environmental tracers, at varied spatial and temporal resolution, is particularly

instructive for distinguishing between the key processes that influence the chemistry of the river

in space and time. Furthermore, the spectrum of tracer techniques provides both qualitative and

quantitative information regarding the flux of groundwater along the length of the lower Herbert

River. Whilst the absolute groundwater fluxes determined have a degree of uncertainty, mass

balances of radon and selected solutes highlight the value of quantitative estimates in

combination with qualitative trends to characterise river-aquifer relationships. The analyses

demonstrate that discharge of groundwater from the alluvial aquifers is a dominant influence on

both the flow and chemistry of the lower Herbert River in the dry season. In particular,

groundwater is a key vector for the delivery of nitrate to the river during low flow conditions.

This provides a new perspective for monitoring and management of nutrients in tropical rivers

where there is good connectivity with the underlying groundwater system.

Key recommendations arising from this research include: (1) water quality sampling should be

undertaken at recognised periods on the stream/groundwater hydrograph, with an understanding

of temporal and spatial river-aquifer connectivity relationships; (2) surface and subsurface

sources of water and dissolved nutrients must be considered, including identification of nutrient

hotpots in both surface water and groundwater systems; (3) sampling locations should capture

the longitudinal variation in river nutrient concentrations, not simply end-of-river monitoring;

(4) appropriate water quality guideline values must be set to account for seasonal changes in

both the sources and forms of nutrients transported to surface waters.

Page 9: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

ix

Table of Contents

ACKNOWLEDGEMENTS ....................................................................................................... v

ABSTRACT............................................................................................................................... vii

TABLE OF CONTENTS .......................................................................................................... ix

LIST OF FIGURES .................................................................................................................. xv

LIST OF TABLES ................................................................................................................. xxiii

CHAPTER 1 RESEARCH CONTEXT .................................................................................... 1

1.1 Introduction ............................................................................................................... 1

1.2 Integrated Catchment Management ........................................................................ 2

1.3 Water Reform in Australia....................................................................................... 4

1.3.1 Water quality management...................................................................................... 4

1.3.2 Groundwater policy and conjunctive management ................................................. 4

1.3.3 National Water Initiative......................................................................................... 5

1.4 Motivation .................................................................................................................. 5

1.4.1 Nitrogen in surface water and groundwater ............................................................ 5

1.4.2 Tropical Australia and nutrient delivery ................................................................. 6

1.5 Scope of the Thesis..................................................................................................... 8

1.6 Thesis Outline ............................................................................................................ 9

CHAPTER 2 CONNECTED WATER RESOURCES .......................................................... 11

2.1 Introduction ............................................................................................................. 11

2.2 Mechanisms of Interaction ..................................................................................... 11

2.2.1 Physical interactions.............................................................................................. 12

2.2.2 Chemical interactions............................................................................................ 15

2.3 Framework for a Nutrient Budget ......................................................................... 18

2.4 Assessment Methods................................................................................................ 20

2.4.1 Characterising river-groundwater interactions...................................................... 20

2.4.1.1 Hydrogeological approaches ........................................................................ 20

2.4.1.2 Hydrometric approaches............................................................................... 21

2.4.1.3 Hydrological approaches .............................................................................. 21

2.4.1.4 Hydrochemical approaches .......................................................................... 22

2.4.1.5 GIS-based approaches .................................................................................. 22

2.4.1.6 Modelling approaches .................................................................................. 23

2.4.2 Characterising nutrient mobility in water.............................................................. 24

Page 10: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

x

2.4.2.1 Field-based studies........................................................................................24

2.4.2.2 Nutrient modelling ........................................................................................25

2.4.3 Implications for the research approach..................................................................26

2.5 Chapter Summary....................................................................................................27

CHAPTER 3 RIVER-AQUIFER INTERACTIONS IN THE WET TROPICS .................29

3.1 Introduction..............................................................................................................29

3.1.1 Climate...................................................................................................................29

3.1.2 Features of streamflow and groundwater...............................................................30

3.1.3 Water quality and nutrients....................................................................................31

3.2 Selection of Case Study Area ..................................................................................33

3.2.1 Catchment water quality issues..............................................................................34

3.2.1.1 Previous N studies on surface water .............................................................35

3.2.1.2 Previous N studies on subsurface water........................................................37

3.3 Catchment Characteristics......................................................................................38

3.3.1 Climate...................................................................................................................41

3.3.2 Rivers and aquifers ................................................................................................42

3.4 Research Approach..................................................................................................43

3.5 Data Collection .........................................................................................................44

3.5.1 Existing data and applicability...............................................................................44

3.5.2 Sample types for this study....................................................................................46

3.5.3 Site selection..........................................................................................................47

3.5.4 Timing of sampling................................................................................................48

3.5.5 Logistics, materials and methods...........................................................................50

3.5.5.1 Sampling technique.......................................................................................50

3.5.5.2 Sample preparation and preservation ............................................................50

3.5.5.3 Analytical techniques....................................................................................51

3.6 Chapter Summary....................................................................................................51

CHAPTER 4 HYDROGEOLOGICAL FRAMEWORK ......................................................53

4.1 Introduction..............................................................................................................53

4.1.1 Key concepts and definitions.................................................................................53

4.2 Geologic Characterisation .......................................................................................55

4.2.1 General depositional environment .........................................................................55

4.2.2 Lithostratigraphic interpretation ............................................................................56

4.2.2.1 Relationships with the river ..........................................................................61

4.2.3 Boundary of the study area ....................................................................................61

4.3 Hydraulic Properties of the Aquifers .....................................................................62

4.3.1 Summary of Cox’s interpretation ..........................................................................62

4.3.2 Interpretation based on lithostratigraphy ...............................................................63

Page 11: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

xi

4.4 Vertical Flow in the Subsurface ............................................................................. 63

4.4.1 Data preparation .................................................................................................... 64

4.4.2 Inter-aquifer connectivity...................................................................................... 67

4.4.3 Rainfall response................................................................................................... 74

4.4.4 Classification of the aquifers................................................................................. 76

4.5 Lateral Flow in the Subsurface .............................................................................. 76

4.5.1 Deep aquifer system.............................................................................................. 77

4.5.1.1 Flow pattern.................................................................................................. 77

4.5.1.2 Recharge-discharge characteristics............................................................... 77

4.5.2 Shallow aquifer system ......................................................................................... 81

4.5.2.1 Flow pattern.................................................................................................. 81

4.5.2.2 Recharge-discharge characteristics............................................................... 81

4.5.3 Relationships between aquifers ............................................................................. 85

4.6 Chapter Summary ................................................................................................... 87

CHAPTER 5 HYDROGEOCHEMICAL FRAMEWORK .................................................. 89

5.1 Introduction ............................................................................................................. 89

5.1.1 General principles ................................................................................................. 90

5.1.1.1 Environmental tracers................................................................................... 90

5.1.1.2 Ion chemistry ................................................................................................ 90

5.1.1.3 Isotope chemistry.......................................................................................... 92

5.1.1.4 Nitrogen chemistry ....................................................................................... 94

5.1.2 Methods of interpretation ...................................................................................... 95

5.2 Hydrochemical Patterns.......................................................................................... 96

5.2.1 Compositional groups ........................................................................................... 98

5.2.2 Linear trends........................................................................................................ 103

5.2.2.1 Deep aquifer ............................................................................................... 104

5.2.2.2 Shallow aquifer........................................................................................... 107

5.2.3 Spatial trends....................................................................................................... 111

5.2.3.1 Deep aquifer ............................................................................................... 111

5.2.3.2 Shallow aquifer........................................................................................... 113

5.3 Vertical Relationships Between Aquifers ............................................................ 115

5.3.1 Unconfined and confined systems....................................................................... 115

5.3.2 Saturation Indices................................................................................................ 116

5.3.3 Intra-aquifer relationships ................................................................................... 117

5.3.4 Inter-aquifer mixing trends.................................................................................. 119

5.3.5 Relationship with the bedrock............................................................................. 126

Page 12: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

xii

5.4 Spatial Relationships Within Aquifers.................................................................126

5.4.1 Lateral hydrochemical evolution .........................................................................127

5.4.1.1 Deep aquifer................................................................................................127

5.4.1.2 Shallow aquifer ...........................................................................................130

5.4.2 Seawater intrusion ...............................................................................................132

5.5 Nitrogen in Groundwater ......................................................................................135

5.5.1 Spatial distribution of N.......................................................................................136

5.5.1.1 Shallow aquifer ...........................................................................................136

5.5.1.2 Deep aquifer................................................................................................137

5.5.2 Speciation of N ....................................................................................................139

5.5.2.1 Shallow aquifer ...........................................................................................139

5.5.2.2 Deep aquifer................................................................................................142

5.5.3 Nitrogen transport ................................................................................................144

5.5.3.1 Shallow aquifer ...........................................................................................144

5.5.3.2 Deep aquifer................................................................................................145

5.6 Chapter Summary..................................................................................................147

CHAPTER 6 PHYSICAL RIVER-GROUNDWATER INTERACTIONS .......................149

6.1 Introduction............................................................................................................149

6.1.1 Data availability and preparation.........................................................................150

6.2 Potential for Hydraulic Connection .....................................................................154

6.3 Direction of Interaction .........................................................................................156

6.3.1 Groundwater elevation - river stage relationships ...............................................156

6.3.1.1 Reach A: river mouth to gauge 116001 ......................................................157

6.3.1.2 Reach B: gauge 116001 to Stone River junction ........................................159

6.3.1.3 Reach C: upstream of Stone River junction to Long Pocket.......................160

6.3.1.4 Reach D: Long Pocket to Abergowrie ........................................................161

6.3.1.5 Interpretation of flux direction ....................................................................162

6.3.2 Flow characteristics .............................................................................................164

6.3.2.1 Stream hydrographs ....................................................................................164

6.3.2.2 Flow duration ..............................................................................................166

6.3.2.3 Hydrograph separation................................................................................167

6.4 Implications for N Transport................................................................................170

6.5 Chapter Summary..................................................................................................173

Page 13: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

xiii

CHAPTER 7 CHEMICAL RIVER-GROUNDWATER INTERACTIONS..................... 175

7.1 Introduction ........................................................................................................... 175

7.1.1 Factors that influence river chemistry ................................................................. 175

7.1.1.1 Climatic factors .......................................................................................... 177

7.1.1.2 Geologic factors.......................................................................................... 178

7.1.1.3 Biogeochemical factors .............................................................................. 178

7.1.1.4 Mixing of waters......................................................................................... 179

7.1.2 Assessment principles and methods.................................................................... 179

7.1.3 Study site & terminology .................................................................................... 181

7.2 General Hydrochemistry ...................................................................................... 183

7.2.1 Compositional groups ......................................................................................... 184

7.2.2 Stable isotopes..................................................................................................... 185

7.3 Temporal Data ....................................................................................................... 186

7.3.1 Field parameters .................................................................................................. 187

7.3.1.1 Electrical conductivity................................................................................ 188

7.3.1.2 pH ............................................................................................................... 191

7.3.1.3 Temperature................................................................................................ 192

7.3.2 Major ions ........................................................................................................... 193

7.3.2.1 Inter-seasonal trends................................................................................... 193

7.3.2.2 Intra-seasonal trends................................................................................... 196

7.3.2.3 Tributaries of the Herbert River ................................................................. 197

7.3.3 Nitrogen............................................................................................................... 199

7.3.3.1 Particulate vs dissolved N .......................................................................... 199

7.3.3.2 Dissolved organic vs inorganic N............................................................... 201

7.3.3.3 Inorganic species ........................................................................................ 202

7.4 Longitudinal Data.................................................................................................. 202

7.4.1 Salinity ................................................................................................................ 203

7.4.2 Major ions in the tidal zone................................................................................. 205

7.4.3 Processes in the freshwater zone......................................................................... 207

7.4.3.1 Evaporation................................................................................................. 207

7.4.3.2 Overland flow............................................................................................. 209

7.4.3.3 Tributary inflow.......................................................................................... 210

7.4.3.4 Groundwater discharge............................................................................... 212

7.5 Tracing Groundwater ........................................................................................... 214

7.5.1 Radon distribution in groundwater...................................................................... 215

7.5.2 Temporal trends in radon along the river ............................................................ 215

7.5.3 Relative flux of groundwater along the river ...................................................... 217

7.5.3.1 Comparison with hydrochemistry .............................................................. 220

Page 14: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

xiv

7.5.4 Uncertainty in groundwater flux..........................................................................222

7.5.4.1 Solute mass balance ....................................................................................222

7.5.4.2 Local baseflow contribution........................................................................224

7.6 Transport of Nitrogen by Groundwater ..............................................................228

7.6.1 Longitudinal trends in DIN..................................................................................229

7.6.1.1 Comparisons with other tracers...................................................................230

7.6.1.2 Comparisons within the dry season.............................................................233

7.6.2 Environmental significance .................................................................................234

7.7 Chapter Summary..................................................................................................237

CHAPTER 8 RESEARCH CONCLUSIONS .......................................................................239

8.1 Introduction............................................................................................................239

8.2 Key Findings...........................................................................................................240

8.2.1 Hydrogeological framework ................................................................................240

8.2.2 River-groundwater interactions ...........................................................................241

8.2.3 The significance of groundwater for river N budgets..........................................242

8.3 Research Approach................................................................................................245

8.3.1 Data collection .....................................................................................................245

8.3.2 Characterising river-aquifer interactions .............................................................246

8.3.3 Characterising nutrient mobility ..........................................................................248

8.4 Implications for the Use of Nutrient Budgets and Models .................................249

8.5 Monitoring and Management Implications for the Tropics...............................250

8.6 Further Research ...................................................................................................251

8.7 Concluding Statement............................................................................................253

REFERENCES ........................................................................................................................255

APPENDIX A LABORATORY ANALYSES AND FIELD DATA....................................271

APPENDIX B RADON SAMPLING PROCEDURE ..........................................................273

APPENDIX C UNCERTAINTY ANALYSIS.......................................................................275

Page 15: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

xv

List of Figures

Figure 1-1 Framework for Integrated Catchment Management ................................................... 2

Figure 1-2 Framework for Integrated Catchment Management illustrating the focus in this study

on surface-groundwater interactions and water resource quality.................................................. 3

Figure 2-1 The basic types of river-groundwater interactions: (a) gaining stream, (b) losing

stream, (c) disconnected stream, and (d) bank storage (after Winter et al., 1998). .................... 13

Figure 2-2 Groundwater and stream channel interactions: (a) parallel-flow and (b) flow-through

dominated streams (after Woessner, 2000, in REM 2002). ........................................................ 14

Figure 2-3 Processes that influence subsurface hydrochemistry (after Back et al. 1993, in

Herczeg and Edmunds 2000). ..................................................................................................... 16

Figure 2-4 The hyporheic zone as the interface between local and regional groundwater flow

systems and surface waters (Winter et al., 1998)........................................................................ 17

Figure 2-5 A conceptual representation of a nitrogen budget..................................................... 19

Figure 2-6 Spatial connections of individual subsystems represented in Figure 2-5.................. 19

Figure 3-1 Location of the Herbert River catchment showing the Herbert River and its major

tributaries .................................................................................................................................... 34

Figure 3-2 Land cover in the lower Herbert River catchment (adapted from Bramley and Muller

1999) ........................................................................................................................................... 39

Figure 3-3 1:250,000 geological map comprising the lower Herbert River catchment (BMR,

1965) ........................................................................................................................................... 40

Figure 3-4 Mean monthly rainfall in the lower Herbert River catchment .................................. 41

Figure 3-5 Cumulative residual rainfall at station 32045 in the lower catchment ...................... 42

Figure 3-6 Daily rainfall versus the cumulative deviation of residual rainfall from the mean at

station 32045............................................................................................................................... 42

Figure 3-7 Location of surface water, groundwater, and rainfall collection sites during the three

sampling periods. ........................................................................................................................ 48

Figure 3-8 The beginning of each sample collection period in 2004-2005, in relation to stream

discharge and rainfall in the lower catchment ............................................................................ 49

Figure 4-1 Bedrock contours in the Herbert River delta............................................................. 56

Page 16: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

xvi

Figure 4-2 Cross-sections constructed from lithological logs of bores in the study area ............57

Figure 4-3 Representative lithologic cross-sections in the lower Herbert River catchment........59

Figure 4-4 Fence diagram for the alluvial aquifer system in the lower Herbert River catchment

.....................................................................................................................................................60

Figure 4-5 Digital Elevation Model (DEM) for the lower Herbert River catchment. .................65

Figure 4-6 Historical groundwater elevations for bores sampled automatically and manually. .66

Figure 4-7 Summary of the degree of vertical hydraulic connectivity (strong, good, poor) and

direction of head gradient between the shallowest and deepest aquifers ....................................67

Figure 4-8 Hydrographs for nested monitoring bores screened in the deepest (A-pipe) and

shallowest (B-pipe) water-bearing units in the upper half of the catchment ...............................68

Figure 4-9 Hydrographs for nested monitoring bores screened in the deepest (A-pipe) and

shallowest (B-pipe) water-bearing units in the lower half of the catchment ...............................69

Figure 4-10 Similarity in hydrographic response for selected nested bores displaying an

upwards potential.........................................................................................................................71

Figure 4-11 Hydrographs for nested monitoring bores screened in an intervening sand unit as

well as in the deepest and/or shallowest water-bearing units ......................................................73

Figure 4-12 Cross-correlation of groundwater levels against rainfall at gauge 32091................74

Figure 4-13 Cross-correlation of groundwater levels against stage height at gauge 116001

(shown as a 14-day moving average) ..........................................................................................75

Figure 4-14 Potentiometric surfaces for the deep aquifer system during (a) the dry season of

1976 and (b) consecutive wet season of 1977 .............................................................................78

Figure 4-15 Head difference contours for the deep aquifer system, comparing groundwater

levels during the 1977 wet season with previous dry season levels in November 1976 .............80

Figure 4-16 Watertable maps for the shallow aquifer system during (a) the dry season of 1976

and (b) consecutive wet season of 1977 ......................................................................................82

Figure 4-17 Head difference contours for the shallow aquifer system comparing groundwater

levels during the 1977 wet season with previous dry season levels in November 1976 .............84

Figure 4-18 Contours of groundwater elevation superimposed for the shallow and deep aquifers

during (a) the dry season and (b) the wet season.........................................................................86

Figure 5-1 The nitrogen cycle, modified after Pidwirny (2005) .................................................94

Figure 5-2 Piper diagram for deep and shallow aquifer samples collected during three sampling

periods: May 2004, October 2004, June 2005.............................................................................97

Page 17: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

xvii

Figure 5-3 Oxygen-18 and deuterium stable isotope data for deep and shallow aquifer samples

and a coastal rainfall event in May 2004 .................................................................................... 98

Figure 5-4 Piper diagram for (a) HSs and (b) HSd groundwater samples displayed by month of

collection..................................................................................................................................... 99

Figure 5-5 Schoeller plots illustrating the two dominant hydrochemical groups observed in

shallow aquifer samples in June 2005....................................................................................... 100

Figure 5-6 Piper diagram for shallow aquifer samples in June 2005........................................ 101

Figure 5-7 Schoeller plots illustrating the two dominant hydrochemical groups observed in deep

aquifer samples in June 2005.................................................................................................... 102

Figure 5-8 Piper diagram for deep aquifer samples in June 2005............................................. 103

Figure 5-9 TDS vs Cl for all deep and shallow aquifer samples .............................................. 104

Figure 5-10 Bivariate plots of major ions against Cl for deep groundwater samples collected

during the three sampling periods in 2004-2005....................................................................... 105

Figure 5-11 Oxygen-18 and deuterium stable isotope data for HSd samples and a coastal rainfall

event in May 2004 .................................................................................................................... 107

Figure 5-12 Bivariate plots of major ions against Cl for shallow groundwater samples collected

during the three sampling periods in 2004-2005....................................................................... 108

Figure 5-13 Bivariate plot of Br/Cl against Cl for shallow aquifer samples collected during the

three sampling periods in 2004-2005........................................................................................ 109

Figure 5-14 Oxygen-18 and deuterium stable isotope data for HSs samples and a coastal rainfall

event in May 2004 .................................................................................................................... 110

Figure 5-15 Pie charts illustrating the spatial distribution of (a) major anions and (b) major

cations in HSd in June 2005 as a percentage of total meq/L ..................................................... 112

Figure 5-16 Monovalent ions plotted against divalent ions, normalised to Cl for all HSd bores

and seawater.............................................................................................................................. 113

Figure 5-17 Pie charts illustrating the spatial distribution of (a) major anions and (b) major

cations in HSs in June 2005 as a percentage of total meq/L ..................................................... 114

Figure 5-18 Field measurements for shallow and deep aquifer samples collected in May and

October 2004............................................................................................................................. 117

Figure 5-19 Schoeller plots for two screened intervals within the same aquifer based on

available random measurements (1975 - 2005) ........................................................................ 118

Figure 5-20 Plots of saturation indices (logarithmic form) for intervals screened within (a) the

deep aquifer and (b) the shallow aquifer................................................................................... 119

Page 18: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

xviii

Figure 5-21 Schoeller plots for groundwater samples from shallow (blue) and deep (black)

nested bores in June 2005..........................................................................................................122

Figure 5-22 Plots of saturation indices (logarithmic form) for selected nested bores based on

samples collected in October 2004............................................................................................124

Figure 5-23 (a) Hydrochemical facies and (b) distribution of total dissolved solids determined at

bores in the deep aquifer in June 2005. .....................................................................................128

Figure 5-24 Stable isotopic values along a flowpath for the deep aquifer based on samples

collected in May 2004. ..............................................................................................................129

Figure 5-25 (a) Hydrochemical facies and (b) distribution of total dissolved solids (b)

determined at bores in the shallow aquifer in June 2005...........................................................131

Figure 5-26 Transects for estimating the theoretical saltwater wedge below freshwater according

to the Ghyben-Herzberg relation ...............................................................................................132

Figure 5-27 Theoretical seawater interface during two periods in 2004 and slotted depths of

bores in HSs and HSd as a function of distance from the coast along three transects................134

Figure 5-28 Distribution of mapped soil types in the study area (Wood et al., 2003) ..............136

Figure 5-29 Spatial distribution of dissolved inorganic nitrogen in bores screened in the (a)

shallow and (b) deep aquifers in June 2005 ..............................................................................138

Figure 5-30 Spatial distribution of (a) NO3- and (b) NH4

+ in bores screened in the shallow

aquifer in June 2005, with selected flowlines depicted .............................................................140

Figure 5-31 Bivariate plots for shallow groundwater samples (2004 samples only), displayed

according to compositional groups............................................................................................141

Figure 5-32 Spatial distribution of (a) NO3- and (b) NH4

+ in bores screened in the deep aquifer

in June 2005...............................................................................................................................143

Figure 5-33 Conceptual diagram summarising the movement of water and N in the alluvial

aquifer system and potentially to the Herbert River..................................................................146

Figure 6-1 Selected QDNRW stream gauges along the lower (116006, 116001) and upper

(116004) Herbert River .............................................................................................................150

Figure 6-2 (a) Surveyed riverbed and (b) estimated river width (June 2005) as a function of

distance from the mouth of the Herbert River ...........................................................................152

Figure 6-3 Derived historical stage height in the Herbert River at Trebonne ...........................153

Figure 6-4 Surveyed topographic features of the lower Herbert River channel ........................154

Figure 6-5 Map showing geographical features which relate to the analyses of groundwater

elevations and river topography/stage along the lower Herbert River ......................................155

Page 19: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

xix

Figure 6-6 Comparison of time series groundwater elevation at bore 1160048 (reach A) with the

corresponding surveyed streambed and bank height in the Herbert River. .............................. 156

Figure 6-7 Comparison of historical groundwater elevations in selected (a) shallow bores and

(b) deep bores with corresponding river stage (adjusted) along reach A.................................. 158

Figure 6-8 Comparison of historical groundwater elevations in selected (a) shallow bores and

(b) deep bores with corresponding river stage heights (adjusted) along reach B ..................... 159

Figure 6-9 Comparison of groundwater elevation in shallow (11600068B) and deep

(11600068A) bores with river stage (adjusted) at Lannercost. ................................................. 160

Figure 6-10 Comparison of groundwater elevation at bore 11600070 and stage height at

adjacent gauge 116006.............................................................................................................. 161

Figure 6-11 Daily flow at the two lower Herbert River stream gauges during 1995-2005 ...... 165

Figure 6-12 Flow duration curves for stream gauges in the lower (116001, 116006) and upper

(116004) Herbert River catchment (refer to Figure 6-1 for gauge locations) ........................... 167

Figure 6-13 Baseflow separation using the Lyne-Hollick algorithm at the lower Herbert River

stream gauges............................................................................................................................ 169

Figure 6-14 Time series of calculated baseflow at stream gauges 116001 and 116006 and

observed total flow at gauge 116001 ........................................................................................ 171

Figure 6-15 Calculated daily baseflow as a percentage of observed streamflow at gauge 116001

.................................................................................................................................................. 172

Figure 7-1 Gibbs diagram depicting the key processes that control the chemistry of surface

waters. ....................................................................................................................................... 176

Figure 7-2 Surface water sampling sites during the three collection periods and locations of

rainfall samples ......................................................................................................................... 182

Figure 7-3 Piper diagram for surface water and groundwater samples collected during three

sampling periods: May 2004, October 2004, June 2005........................................................... 183

Figure 7-4 Modified Gibbs diagram with logarithmic plot of TDI against Na/(Na+Ca) for water

samples collected along the lower Herbert River ..................................................................... 184

Figure 7-5 Schoeller plots of (a) samples in the Herbert River and (b) Ca-Mg enriched

groundwaters of the shallow aquifer ......................................................................................... 185

Figure 7-6 Oxygen-18 and deuterium stable isotope data for surface water and groundwater

samples and a coastal rainfall event in May 2004..................................................................... 186

Figure 7-7 Selected QDNRW stream gauges along the lower (116006, 116001) and upper

(116004) Herbert River............................................................................................................. 187

Page 20: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

xx

Figure 7-8 Time series water quality data relative to flow at gauge 116001.............................188

Figure 7-9 Electrical conductivity versus flow at gauge 116001 ..............................................188

Figure 7-10 Continuous daily electrical conductivity at gauge 116001 relative to streamflow and

rainfall .......................................................................................................................................189

Figure 7-11 Electrical conductivity versus flow at gauge 116001, depicting theoretical

evaporation lines starting from different flow/EC combinations ..............................................190

Figure 7-12 Continuous daily pH at gauge 116001 relative to streamflow ...191

Figure 7-13 Time series of mean air temperature, mean river temperature and streamflow at

gauge 116001 in the lower Herbert River .................................................................................192

Figure 7-14 Correlation between mean air and river temperatures at gauge 116001 in the lower

Herbert River .............................................................................................................................193

Figure 7-15 Piper plot of historical major ion compositions (1973-2004) of the Herbert River at

three stream gauges ...................................................................................................................194

Figure 7-16 Schoeller plots of historical (a) wet season and (b) dry season water quality samples

collected at gauge 116001 since the 1970’s ..............................................................................194

Figure 7-17 HCO3 and Cl concentrations against streamflow at gauge 116001 during months of

the wet and dry seasons .............................................................................................................195

Figure 7-18 Major ion chemistry for samples collected in the lower Herbert River during the

beginning (May 2004, June 2005) and end (October 2004) of the dry season..........................196

Figure 7-19 Schoeller plots for tributaries of the lower Herbert River during the beginning (May

2004, June 2005) and end (October 2004) of the dry season ....................................................198

Figure 7-20 Saturation indices (logarithmic form) for bores and tributaries in the upper part of

the catchment based on samples collected at the end of the dry season (October 2004). .........199

Figure 7-21 Time series concentrations of (a) total, (b) particulate and (c) soluble N at various

sites along the lower Herbert River ...........................................................................................200

Figure 7-22 Concentrations of dissolved N as inorganic and organic components at (a) upstream

and (b) downstream sampling sites along the lower Herbert River...........................................201

Figure 7-23 Nitrate and ammonium concentrations at (a) upstream and (b) downstream

sampling sites along the lower Herbert River ...........................................................................202

Figure 7-24 Measurements of field electrical conductivity along the lower Herbert River and its

tributaries during the beginning (June) and end (October) of the dry season............................204

Figure 7-25 Longitudinal comparison of major ion and oxide concentrations along the lower

Herbert River during months representing the beginning and end of the dry season................205

Page 21: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

xxi

Figure 7-26 Mg/Ca ratio for samples collected along the lower Herbert River during October

2004 and June 2005 .................................................................................................................. 206

Figure 7-27 Salinity, Na/Cl and oxygen-18 (δ18O) along the lower Herbert River .................. 208

Figure 7-28 Longitudinal trends for selected major ions along the lower Herbert River in June

2005 .......................................................................................................................................... 209

Figure 7-29 Radon activities along the freshwater reaches of the lower Herbert River and

sampled tributaries in October 2004 ......................................................................................... 210

Figure 7-30 Longitudinal plots for selected ions along the freshwater reaches of the lower

Herbert River and sampled tributaries in October 2004 ........................................................... 211

Figure 7-31 Longitudinal trends for selected major ions along the lower Herbert River in

October 2004............................................................................................................................. 213

Figure 7-32 Concentration of radon along the lower Herbert River during periods representing

the end (October 2004) and beginning (June 2005) of the dry season...................................... 216

Figure 7-33 Measured radon concentrations and estimated groundwater flux along the lower

Herbert River in October 2004 and at selected reaches in June 2005....................................... 219

Figure 7-34 Ions ratios, estimated groundwater flux and radon along the lower Herbert River in

October 2004............................................................................................................................. 221

Figure 7-35 Speciation of DIN along the lower Herbert River during months representing the

beginning (May) and end (October) of the dry season ............................................................. 229

Figure 7-36 Comparison of NO3- with ion ratios and radon along the lower Herbert River in

October 2004............................................................................................................................. 231

Figure 7-37 Longitudinal plots of NO3- during months representing the beginning (May, June)

and end (October) of the dry season. ........................................................................................ 233

Figure 7-38 Conceptual diagram summarising the movement of water and N between the

aquifers and the lower Herbert River at the end of the dry season ........................................... 236

Figure 8-1 Conceptual diagram summarising the movement of water and N in the alluvial

aquifer system and potentially to the Herbert River ................................................................. 243

Figure 8-2 Conceptual diagram summarising the movement of water and N between the aquifers

and the lower Herbert River at the end of the dry season ......................................................... 244

Page 22: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to
Page 23: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

xxiii

List of Tables

Table 3–1 Available datasets for the lower Herbert River catchment ........................................ 45

Table 3–2 Summary of laboratory and field measurements ....................................................... 46

Table 4–1 Summary of hydraulic characteristics of the water-bearing alluvial stratigraphic units

in the Herbert River valley, after Cox (1979) ............................................................................. 62

Table 5–1 Hydrochemical groups of the alluvial aquifers ........................................................ 110

Table 6–1 Classification system for stream-aquifer interactions relevant to conjunctive use

management (REM, 2002)........................................................................................................ 149

Table 6–2 Features of the QDNRW stream gauging stations in the lower catchment and selected

upper catchment gauges. ........................................................................................................... 151

Table 6–3 Comparison between gauged and derived stage heights on the same day of

measurement during the dry season. ......................................................................................... 152

Table 6–4 Bores and comparison stream gauges at sites along the four river reaches ............. 157

Table 6–5 Comparison of derived river stage at Timrith and Lannercost and actual stage heights

at gauging stations 116001 and 116006 in the lower Herbert River. ........................................ 160

Table 6–6 Summary of the dominant river-groundwater elevation relationships along the lower

Herbert River during wet and dry seasons in the historical record. .......................................... 162

Table 7–1 Distinctive chemical characteristics of processes that influence the chemistry of

surface waters............................................................................................................................ 181

Table 7–2 Required and estimated flow rates in tributaries of the lower Herbert River and

groundwater based on changes in total solute loads ................................................................. 212

Table 7–3 Measured input parameters for modelling radon activities in the river ................... 218

Table 7–4 Best estimates of the percentage of local baseflow that contributes to streamflow

along reach 2 during months in the dry season, based on a solute mass balance. .................... 224

Table 7–5 Calculated values for xi, Δxi and the fraction of error (equation 7-14) during the

beginning (May) and end (October) of the dry season in 2004 ................................................ 227

Table 7–6 Required groundwater discharge to account for observed NO3- concentrations in the

Herbert River in October 2004.................................................................................................. 232

Table 7–7 Concentration range for species of DIN for all samples collected in the lower Herbert

River during selected months of the dry season ....................................................................... 235

Page 24: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to
Page 25: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

1

Chapter 1 Research Context

1.1 INTRODUCTION

Water is a key natural resource that is required for the environment and to support the nation’s

social and economic structures. Sustainable water resources can be defined as ‘those designed

and managed to fully contribute to the objectives of society, now and in the future, while

maintaining their ecological, environmental and hydrological integrity’ (Loucks and Gladwell,

1999). The continued availability of water is open to change, not only through natural variation,

but also through human impacts on the water cycle (ANZECC and ARMCANZ, 1994).

Therefore, the way in which we use and manage water today can influence the availability and

quality of water for all users of the resource in the future.

The notion of water management is concerned with maintaining water quality and quantity in

both surface and sub-surface reservoirs. In many parts of the landscape, sub-surface water

(including groundwater) and surface water (streams, lakes, dams, wetlands and estuaries)

interact such that the quality and quantity of one affects the quality and quantity of the other

(Winter et al., 1998). For example, surface water features can gain water and solutes from

groundwater systems, while in other situations surface waters can be a source of groundwater

recharge and cause changes in groundwater quality. In addition, pumping of groundwater can

affect surface water volumes, or conversely, withdrawal of water from streams can deplete

groundwater reserves. Therefore, double accounting of water resources, based on independent

allocations of surface waters and groundwaters, can have serious consequences for future water

availability.

Water management has traditionally focused on surface water and groundwater as if they were

separate entities rather than connected resources. This can largely be attributed to factors such

as: differing timescales of surface water and groundwater movement in the landscape; lack of

data and methodology; disciplinary differences between surface and groundwater hydrologists;

and legislation. However, increased understanding of the importance of the linkages between

groundwater and surface water is directing more water managers throughout the world to adopt

conjunctive management principles (REM, 2002).

Page 26: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 1

2

A new awareness of the linkages that exist between groundwater and surface water has been

brought about because of:

• the increased demand for water, resulting in the need to maximise the water available

for use;

• the realisation of the ecological importance of groundwater and the need to consider its

environmental values; and

• the obvious disastrous effects that an ‘out of balance’ system can have in terms of

drawdown, salinity and water quality (REM, 2002).

Although recent developments in water policy in Australia have begun to reflect this increased

awareness, the scientific research and knowledge to underpin national policy, and hence inform

effective management practices, is still in relative infancy. Whilst this thesis is primarily

concerned with the scientific aspects of surface-groundwater interactions, the implications of the

research findings for water policy are also considered.

1.2 INTEGRATED CATCHMENT MANAGEMENT

Integrated Catchment Management (ICM) involves a systems approach to managing a

catchment. It is directed towards achieving a balance between natural environmental values,

economic activity and social concerns. The connections between each element in an ICM

problem can be represented within a framework such as that depicted in Figure 1-1.

POLICY

CLIMATE

LAND USE

WATER RESOURCES

OUTCOMES

LAND MANAGEMENT

PRACTICE

POLICY

CLIMATE

LAND USE

WATER RESOURCES

OUTCOMES

LAND MANAGEMENT

PRACTICE

Figure 1-1 Framework for Integrated Catchment Management

Page 27: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Research Context

3

Considering land use as the start of the ICM loop, then the type of activity carried out on a

parcel of land will govern the land management practices undertaken. For example, in an area of

cropping activity, a landholder has particular management practices with regard to fertiliser

regime: the type, timing and amount applied to their crop. Due to the interconnectedness of land

use and water, land management practices such as crop fertilisation or irrigation will necessarily

have impacts on the quality and quantity of water resources. Impacts on water resources lead to

outcomes which have biophysical, ecological, economic and social implications (Ticehurst et

al., 2007; Letcher et al., 2005). Within such a system, change can be brought about through

voluntary measures or through levers such as policy intervention, at catchment, state or national

level. Such policies for example, can direct how land is used, set rules for how land should be

managed or establish thresholds for the quality and quantity of water in a catchment. Through

the notion of adaptive management, policies can be revised as experience reveals where

amendments are required. An external driver to the framework loop is climate, an overarching

factor in the system, but over which there is no direct control.

The focus of this research is only on one aspect within the framework of ICM: water resource

quality and the interaction between groundwater and surface water (Figure 1-2). Although the

thesis does not deal specifically with the complexities of ICM, the cause and effect relationships

related to water quality are considered within the broader ICM framework.

POLICY

CLIMATE

LAND USE

WATER RESOURCES

OUTCOMES

LAND MANAGEMENT

PRACTICE

fertilisertillagechannelling

croppinggrazingforestryhorticulturenational park

rainfalltemperatureseasons

biophysicalecologicaleconomicsocial

qualityquantity

adaptivemanagement surface water

groundwater

e.g. COAG

POLICY

CLIMATE

LAND USE

WATER RESOURCES

OUTCOMES

LAND MANAGEMENT

PRACTICE

fertilisertillagechannelling

croppinggrazingforestryhorticulturenational park

rainfalltemperatureseasons

biophysicalecologicaleconomicsocial

qualityquantity

adaptivemanagement surface water

groundwater

e.g. COAG

Figure 1-2 Framework for Integrated Catchment Management illustrating the focus in this study on surface-groundwater interactions and water resource quality in an agricultural-based region.

Page 28: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 1

4

1.3 WATER REFORM IN AUSTRALIA

Australia is the driest inhabited landmass and has the world’s highest rainfall variability

(McMahon et al., 1992); thus, there is an inherent limitation to the water resource-base. Water is

crucial to Australia’s natural and economic wealth. Inefficient and inappropriate water use has

created widespread degradation of Australia’s natural resources, such as a decline in the quality

and quantity of surface and groundwater systems, increased land salinisation, a contraction in

wetlands and diminishing populations of native fish, flora and fauna.

In February 1994, all State and Territory governments in Australia agreed that the management

and regulation of Australia’s water resources required significant policy and institutional

change, as it was recognised that inefficient and inappropriate water use was responsible for

widespread degradation of Australia’s natural resources. This agreement resulted in the

endorsement of a national policy by the Council of Australian Governments (COAG), known as

the COAG Water Reform Framework, to achieve an efficient, economically viable and

environmentally sustainable urban and rural water industry. The framework recognised the

unique characteristics of Australia’s water resources and their economic, social and

environmental importance for Australia’s wellbeing.

1.3.1 Water quality management

An important component of water reform is the management of water quality. In 1992 a

National Water Quality Management Strategy (NWQMS) was introduced by the

Commonwealth, States and Territories and was subsequently included in the COAG Water

Reform Framework (ARMCANZ and ANZECC, 1994). The main policy objective of the

NWQMS is based on the philosophy of ecologically sustainable development: to achieve

sustainable use of the nation’s water resources by protecting and enhancing their quality while

maintaining economic and social development. Guidelines for groundwater quality protection

were included as a module of the NWQMS (ARMCANZ and ANZECC, 1995).

1.3.2 Groundwater policy and conjunctive management

Although the management of groundwater was included in the provisions of the 1994 reforms,

the framework had a focus on surface water resources, and was not explicit about which aspects

applied to surface water and groundwater. It was not until 1996 that the issues surrounding

groundwater management were formally accepted into the reform framework by COAG, and

thus introduced into national water policy. Recommendation 3 drew attention to the problems of

managing interlinked groundwater and surface water resources through different programs, and

recommended that ‘groundwater and surface water resource management should be better

integrated’ (ARMCANZ, 1996). According to the policy paper by the Framework’s Taskforce:

Page 29: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Research Context

5

‘In many situations, groundwater and surface water are interconnected and interchangeable

resources where decisions made in one area affect the other’ (ARMCANZ, 1996). In addition,

Resolution 6 stated that: ‘Progress has been made in some limited areas in meeting

environmental needs of groundwater, but further progress has been constrained by a poor

understanding of the location, extent and processes associated with groundwater/surface water

interactions and associated ecosystems’ (ARMCANZ, 1998).

1.3.3 National Water Initiative

In 2003, COAG agreed to refresh the 1994 water reform agenda and develop a National Water

Initiative (NWI) in order to increase the productivity and efficiency of water use, sustain rural

and urban communities, and to ensure the health of river and groundwater systems. A National

Water Commission was established in 2004 to assess progress in implementing the NWI and to

advise on actions required to better realise the objectives of the Agreement. In relation to

conjunctive water management, one of the key objectives of the intergovernmental agreements

on a NWI is to achieve ‘recognition of the connectivity between surface and groundwater

resources and for connected systems to be managed as a single resource’ (COAG, 2004).

1.4 MOTIVATION

In light of the above policy framework, the interaction between groundwater and surface water

is nationally regarded to be an important water resource management consideration. This thesis

is concerned with the quality aspect of connected water resources and particularly, how the

interaction with groundwater affects in-stream water quality. In regards to water quality, river-

aquifer interactions are important in situations where a stream can be degraded by discharge of

saline or other low-quality groundwater, or where groundwater is polluted by a contaminated

stream. Pollution of surface and groundwater resources by nutrients is of growing concern

worldwide due to potentially harmful effects on both ecosystem and human health. The

movement of water between groundwater and surface water provides a major pathway for

chemical transfer between terrestrial and aquatic systems (Winter et al., 1998).

1.4.1 Nitrogen in surface water and groundwater

Nitrogen (N) inputs in intensive-agricultural catchments have been identified as a major causal

factor in the trends of increased nutrient concentrations in surface, ground and coastal waters

(Heathwaite et al., 2000). In recognition, European directives have been introduced to reduce

water pollution caused or induced by nitrates from agricultural sources and to prevent further

pollution (e.g. Statutory Instrument 1289, 2006; Council Directive 91/676/EEC, 1991). Diffuse

sources of N in agriculture are generally derived from fertilisers and manure. Imbalances in

nutrient budgets, such as for N, have shifted in scale from local to regional and continental

Page 30: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 1

6

dimensions during the last decade (Oenema et al., 2003). Nitrogen in surface and groundwater

systems is a concern for both human and environmental health. Elevated levels of nitrate in

drinking water can be harmful to humans, especially infants (Lawrence, 1983), as well as

livestock, while nitrogen inputs into surface waters can cause problems of eutrophication.

Nitrogen can be exported from agricultural areas in many forms and through various pathways,

including leaching losses to groundwater. Unlike salt, which is transported conservatively

throughout the landscape, nitrogen has a complex cycle because of the various forms it can take,

depending on the hydrochemical environment. Nutrient mobilisation from non-point sources

depends on the coincidence of source and transport controls (Heathwaite et al., 2000). Land

management factors largely control the magnitude of potential N loss (source control), while the

rate of N loss through leaching (transport control) depends on soil properties and the amount of

water percolating through the soil profile. Nitrogen loss from agricultural land generally occurs

at the catchment scale (Heathwaite et al., 1989). This is especially the case under oxidising

conditions, where nitrate is the dominant species of N, as it is soluble and is hence transported

with the movement of water. The high mobility of N as nitrate in leaching water means that a

significant proportion of the nitrate created by source factors is translated into N loss

(Heathwaite et al., 2000).

Nutrient export from catchments is a concern for Australian river systems for a number of

reasons, one of which is the risk of algal bloom formations. It is now recognised that in addition

to exports from surface runoff, groundwater can be a significant source of nitrogen to

freshwater, estuarine and coastal environments (Lamontagne et al., 2003; Lamontagne et al.,

2002; Linderfelt and Turner, 2001; Smith and Turner, 2001). Therefore, understanding river-

aquifer connectivity is an important aspect of nutrient management in landscapes where

nutrient-bearing surface waters and groundwaters interact.

1.4.2 Tropical Australia and nutrient delivery

Terrestrial runoff of pollutants has led to concerns over the threat to tropical Australian

waterways, wetlands and coastal ecosystems, such as the Great Barrier Reef World Heritage

Area (GBRWHA) (Brodie and Mitchell, 2005; Furnas, 2003). Water borne contaminants have

the potential to be delivered to the GBRWHA via direct farm runoff into streams and rivers;

leaching into groundwater systems with subsequent discharge into rivers, wetlands and near

shore marine waters; and submarine groundwater discharge via paleochannels (Stewart et al.,

2005; Stieglitz, 2005). Runoff incorporating sediment, nutrients and pesticides is increasing,

with loads for most pollutants being many times higher than the natural amount discharged e.g.

150 years ago (Brodie and Mitchell, 2005). The principal land uses in northern Australia

contributing to this pollution are cropping and cattle grazing, with lesser contributions from

industrial areas, mining and urban developments. Furthermore, Brodie and Mitchell (2005) note

Page 31: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Research Context

7

that due to agricultural development, nutrient inputs (of N and P) have changed from

dominantly dissolved organic forms in ‘natural’ systems to dissolved inorganic forms that are

more bioavailable than organic species.

Although remnant forest exists in parts of northern Australian, most catchments are now used

for agricultural or pastoral purposes. Sugarcane is the dominant cultivated crop in northern

Australia, harvested primarily on the coastal plain south of the Daintree, around the Atherton

Tablelands in Queensland, and in the Ord River irrigation area of Western Australia (Brodie and

Mitchell, 2005). Off-site impacts due to nitrogen and other nutrient/chemical losses from

sugarcane production have come under increasing public scrutiny, particularly when they occur

adjacent to sensitive wetlands, marine, and estuarine environments, such as the Great Barrier

Reef (Bohl et al., 2001). It is well documented that high concentrations of N, especially

inorganic nitrate, are found particularly in surface waters draining land under sugarcane (Bartley

et al., 2003; Furnas, 2003; Bramley and Roth, 2002; Bramley and Muller, 1999). In addition to

surface waters, groundwater studies have shown there to be subsurface losses of nitrate beneath

sugarcane growing areas (Bohl et al., 2000b; Weier, 1999). Despite the recognition of nitrate

leaching into both surface and subsurface waters, the role and contribution of nitrate in

groundwater on surface water quality is not well documented.

High nutrient loadings, especially of N and P, can disrupt the nutrient balance and cause

widespread eutrophication in both in-stream and near-shore surface waters. Coral reefs are

particularly susceptible to damage from increased nutrient discharge from terrestrial sources;

reef communities degraded by terrestrial inputs are documented globally (Fabricius, 2005). Due

to the ecological significance of the GBRWHA, this study will focus on one of the thirty-four

GBR coastal catchments, the Herbert River catchment in north Queensland, to explore linkages

between groundwater and surface water systems. The Herbert is an important river system of

the wet-dry tropics because it drains directly into the area of the GBR lagoon considered under

greatest threat from terrestrial runoff (Productivity Commission, 2003). Sugarcane production is

widespread in the catchment; therefore, terrestrial activities can potentially impact on the quality

of water both within the river system and discharged to the marine environment. Further

characteristics of the case study catchment are deferred to Chapter 3.

In general, Australia’s tropical river systems are poorly understood in comparison with

Australia’s temperate freshwater and tropical marine systems (Hamilton and Gehrke, 2005).

Although these tropical catchments are considered to have greater freshwater biodiversity and

aquatic ecosystem health than found in temperate Australia, eutrophication of estuaries and

coastal waters has been documented and is attributed primarily to increased nutrient loading

from rivers. Based on a review of tropical research in Australia, Hamilton and Gehrke (2005)

noted that hydrological linkages between groundwater and surface waters are inadequately

Page 32: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 1

8

understood in tropical Australia. Therefore, this research aims to contribute to the knowledge

base of river-aquifer interactions in tropical river systems of Australia, including the

significance of groundwater for river nitrogen budgets. Whilst there has been a greater

geographic focus of research centred on catchments draining into the GBR than elsewhere in the

Australian tropics, it is anticipated that the research methodology, implications for monitoring,

and management considerations presented in this thesis, will have greater application than to the

GBR coastal catchments alone.

1.5 SCOPE OF THE THESIS

The overarching aim of the thesis is to characterise the interaction between surface water and

aquifer systems and thereby consider the potential implications of these interactions for river

nitrogen budgets. Emphasis is on investigating surface-subsurface linkages rather than

understanding nutrient sources, processes and loads. Water is a vehicle for mobilising and

transporting nutrients and other dissolved constituents; therefore, knowledge of the dynamics of

water movement, including connectivity between surface and subsurface reservoirs, is central to

understanding the transport of such components. In order to assess large-scale river-

groundwater interactions and the application to nutrient transport, a catchment scale

investigation is undertaken. This scale is considered appropriate for the management of

catchments, rather than individual river-reaches, which also has relevance for water policy. In

addition, potential transferability of the thesis findings to other regions is enhanced by

conducting a broad scale assessment. The selected case study catchment, located in the tropical

climate zone of northeastern Australia, provides a unique opportunity to study baseflow

conditions (largely groundwater contributions) with minimal overland flows. Given the focus on

quality aspects rather than quantity, an extensive water quality sampling program was instigated

in the case study catchment during the dry season, as described in Chapter 3. A range of

hydrogeological, hydrological and hydrochemical analytical tools are applied to existing and

new data in order to address three key questions:

(1) What is the nature of river-aquifer interactions in the Herbert River catchment,

particularly during the dry season?

(2) What is the significance of river-aquifer interactions for the nitrogen budget of the river?

(3) What are the implications of these interactions for nutrient monitoring, management,

and policies relating to water quality at a catchment scale?

Whilst this research primarily focuses on (1) and (2), recommendations for management (3) are

raised in the concluding chapter of the thesis.

Page 33: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Research Context

9

1.6 THESIS OUTLINE

This thesis has been structured into eight chapters. A summary of each chapter is outlined

below:

Chapter 1: Research Context

This chapter provides a brief policy context for this research, outlining the most significant

historical developments in water reform in Australia, and highlights the need to integrate surface

and groundwater management as it relates to water quality. Pollution of connected water

resources by nutrients arising from agricultural activities is discussed in the context of northern

Australian catchments, with particular attention drawn to the significance of the Great Barrier

Reef and terrestrial activities in adjacent catchment areas. This provides a motivation for the

chosen case study catchment and scope of the thesis.

Chapter 2: Connected Water Resources

This chapter reviews current scientific understanding of physical and chemical interactions

between surface and groundwaters in the landscape and summarises methodologies for

investigating interaction processes. Approaches for characterising nutrient transport and fluxes

are also described. The merits of understanding physical processes such as river-aquifer

linkages as an aid to interpreting model outputs are also discussed. This provides a framework

for the methodology employed in this research.

Chapter 3: River-Aquifer Interactions in the Wet Tropics

This chapter discusses the key distinguishing features of tropical climates in relation to

hydrological processes and water quality. In addition, the chapter introduces the characteristics

of the case study catchment and summarises previous water quality studies relating to nitrogen.

The general research approach of the thesis and available datasets are also outlined. Additional

data requirements are highlighted: these underpin the extensive water quality sampling program

undertaken in the catchment. Data collection and details of the sampling methodology are also

described.

Chapter 4: Hydrogeological framework

This chapter is the first of two chapters focused on characterising the hydrogeology in the case

study area. Existing bore logs are interpreted and physical datasets are analysed to conceptualise

the hydrogeology of the alluvial aquifer system.

Page 34: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 1

10

Chapter 5: Hydrogeochemical framework

This chapter provides a validation and extension of the physical hydrogeological framework,

based primarily on analysis of hydrogeochemical data collected as part of this study. The spatial

distribution and speciation of nitrogen in groundwater are also examined in order to establish

whether there is the potential for nitrogen in groundwater to contribute to the river system.

Chapter 6: Physical River-Groundwater Interactions

This chapter considers physical hydraulic relationships between groundwater and surface water,

in order to assess potential for hydraulic connection and the nature of stream-aquifer interaction

processes along the river. Based on the interaction characteristics, implications for the transport

of nitrogen from groundwater to surface waters are also considered.

Chapter 7: Chemical River-Groundwater Interactions

In this chapter, analyses of an extensive database of hydrochemical data provide a powerful tool

to characterise the hydrochemistry of the river during the wet and dry seasons. In particular,

river-aquifer relationships established in the previous chapter are verified and enhanced through

the application of qualitative and quantitative techniques. The potential for a groundwater

source of nitrate and the environmental significance for in-stream and marine ecosystems are

also raised.

Chapter 8: Conclusions

The final chapter provides a summary of conclusions of the thesis, including the key research

contributions. In light of the conceptual understanding developed throughout the thesis on river-

aquifer interactions, the chapter discusses implications and recommendations for nutrient

monitoring, management, and water policy, particularly in tropical catchments. Suggested

future research is also outlined.

Page 35: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

11

Chapter 2 Connected Water Resources

2.1 INTRODUCTION

Groundwaters are fed by rain and surface waters. Groundwaters ultimately discharge to

surface waters or the sea. Surface waters are fed by groundwaters, and feed groundwaters.

Surface and groundwaters form parts of one interlinked system (p. 98, Nevill et al., 2001).

Groundwater and surface water are not isolated components of the hydrological cycle; they

interact in a range of topographic, geologic and climatic landscapes (Sophocleous, 2002; Winter

et al., 1998). Effective management of water requires an understanding of the components of

the hydrological cycle as well as the linkages between those components. One important

connection that has traditionally been overlooked in water resource management in Australia is

the interaction between surface water and groundwater. Direct observation and measurement of

surface-subsurface relationships is complex, as it requires an understanding of processes that

occur over different timescales and how they interact both temporally and spatially. Whereas

water movement on the surface can be quite rapid in response to rainfall events, the movement

of water beneath the land surface can be slow and variable, hence more difficult to predict. As

outlined in Chapter 1, the focus of this thesis is on water quality aspects of river-groundwater

interactions and the significance of groundwater as a vector for nitrogen to the stream. This

chapter reviews current scientific understanding of mechanisms by which surface water and

groundwater interact in the landscape. A conceptual framework for a nitrogen budget is also

presented, as a context for assessing river-aquifer connectivity and the consequent mobilisation

of N. Furthermore, methods for investigating interaction processes and broad approaches for

characterising nutrient transport and fluxes are reviewed. This provides a framework for the

research approach of the thesis.

2.2 MECHANISMS OF INTERACTION

Comprehensive reviews by Sophocleous (2002), Woessner (2000), Winter (1999) and Winter et

al. (1998), outline the key developments in scientific understanding of groundwater-surface

water interactions based on theoretical and field-based studies. Recent research has focussed on

the ecological significance of groundwater-surface water interactions; specifically, in

understanding the biogeochemical processes that occur at the interface. Literature by Boulton et

al. (1998), Dahm et al. (1998) and Brunke and Gonser (1997) provide a review of this emerging

Page 36: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 2

12

research. The following sections describe the main principles and mechanisms by which surface

water and groundwater interact, with an emphasis on river-groundwater interaction processes.

Whilst the discussion assumes aquifers are simple porous media, it is noted that much of

Australia is underlain by fractured aquifers which are more difficult to characterise and model.

Therefore, many of the conceptual models are oversimplified.

2.2.1 Physical interactions

The movement of surface water and groundwater is largely controlled by the physiography

(topography and geologic framework) of an area, while the sources and losses of water in the

hydrological cycle are controlled by climate (Winter, 1999). These controls in turn influence the

flowpath along which groundwater moves, known as the groundwater flow system, and hence

determine the nature of the interaction with surface water features. Streams, lakes and wetlands

are integral parts of groundwater flow systems whereby fluxes of water and chemicals from and

to groundwater reflect the positions of surface water bodies with respect to different-scale

groundwater flow systems (Winter, 1999).

Groundwater moves along flowpaths of varying length and travel time, from areas of recharge

to areas of discharge. Based on their spatial extent and influence, Tóth (1963) identified three

distinct types of flow systems: local, intermediate and regional. In local flow systems water is

recharged at watertable highs and flows to nearby discharge areas such as streams. Groundwater

in regional flow systems travels a greater distance and often discharges to major rivers, large

lakes or oceans. As local flow systems are the most dynamic and the shallowest flow systems,

they have the greatest interchange with surface water (Winter et al., 1998). In general, areas of

high topographic relief tend to have dominant local flow systems, whereby groundwater moves

in systems of predictable patterns; conversely, areas of nearly flat relief tend to have dominant

intermediate and regional flow systems, in which the scale of groundwater movement is much

larger and less predictable. In nature, a region with irregular topography contains multiple flow

systems of differing sizes and depths that can be superimposed on one another within a

groundwater basin (Sophocleous, 2002).

Surface water features such as lakes, wetlands and streams can have multiple sources of water,

including direct precipitation, surface runoff and groundwater. Although groundwater

interactions with lakes and wetlands are similar to that of river systems, there are subtle

differences that will not be discussed here (refer to Winter, 1999; Winter et al., 1998). Streams

interact with groundwater by: (1) gaining inflow of water from groundwater through the

streambed (gaining stream, Figure 2-1a); (2) losing water to groundwater by outflow through

the streambed (losing stream, Figure 2-1b); or (3) variably gaining or losing water depending

on the season and/or physiographic conditions. A connected stream is in direct contact with the

underlying aquifer via a zone of saturated material, or is separated by a narrow unsaturated zone

Page 37: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Connected Water Resources

13

generally less than twice the stream width (Bouwer and Maddock, 1997). Where these

conditions are not met, disconnection from the groundwater system can result, with losses

occurring to seepage (Figure 2-1c). Bank storage is another type of river-groundwater

interaction whereby a rapid rise in river stage, higher than adjacent groundwater levels, causes

water to move from the river into the streambank (Figure 2-1d). The return of this water to the

stream can be on the order of a few days (in the case of local bank storage) or years (in response

to flooding and aquifer recharge) (Winter et al., 1998).

a

b

c

d

Figure 2-1 The basic types of river-groundwater interactions: (a) gaining stream, (b) losing stream, (c) disconnected stream, and (d) bank storage (after Winter et al., 1998). Arrows show the direction of water movement.

Page 38: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 2

14

Woessner (2000) additionally defines fluvial plain-groundwater interactions as: parallel-flow

dominated, where groundwater flows parallel to the stream but essentially does not discharge

into it (Figure 2-2a); and flow-through dominated, where a reach both gains and loses water as

groundwater flows through (Figure 2-2b).

a

b

Figure 2-2 Groundwater and stream channel interactions: (a) parallel-flow and (b) flow-through dominated streams. The watertable (between the lightest and intermediate shades), stream (darkest shade), direction of groundwater flow (arrows) and equipotential lines (dashed) are shown (after Woessner, 2000, in REM 2002).

River-aquifer interactions occur at a variety of scales. Whilst groundwater exchange with the

stream occurs by processes of discharge, recharge and flow-through at the fluvial plain scale,

hyporheic exchange occurs at the channel-bed scale, whereby local, shallow surface water

circulation in the underlying sediments creates areas of groundwater recharge and discharge

within river reaches that are characterised as either gaining or losing. Hyporheic exchange itself

occurs at many scales, ranging from centimetres to tens of metres, depending on sediment type,

bed geometry and hydraulic gradients in the adjacent groundwater system (Woessner, 2000;

Winter et al., 1998).

The direction of interaction in hydraulically connected systems depends on the relative

elevations of the watertable and stream-water surface. For instance, in order for groundwater to

discharge into a stream (gaining condition), the altitude of the watertable (groundwater head) in

the vicinity of the stream must be higher than that of the river stage; the converse applies for

surface water discharging to groundwater (losing condition). Therefore, parallel-flow occurs

when the channel stage and groundwater head are equal, while flow-through occurs when the

channel stage is less than the groundwater head on one bank and greater than the head at the

opposite bank (Woessner, 2000). Importantly, whilst the relative hydraulic heads influence the

direction of exchange, the actual river-groundwater flux depends on the hydraulic conductivity

of both the channel sediments and the aquifer (Sophocleous, 2002).

The interaction between surface water and groundwater can vary both spatially and temporally.

While in some environments the river-groundwater relationship may remain fairly constant, in

other environments flow direction may vary along the stream length (gaining and losing in

different reaches), and also change in response to rainfall events or seasonal climatic patterns

(gaining or losing in the same reach at different times). Water that enters a surface water body

Page 39: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Connected Water Resources

15

rapidly in response to a water input event such as rainfall is known as quick flow or event flow.

This water is distinguished from baseflow, or subsurface water (mostly groundwater flow), that

contributes to the stream and maintains streamflow between water input events. Discharge and

recharge of an aquifer has a buffering effect on the flow regimes of rivers (Brunke and Gonser,

1997). For instance, under low precipitation, discharge in many streams is attributed to

baseflow. In contrast, under conditions of high rainfall, increased surface runoff and interflow

lead to higher hydraulic pressures in the lower stream reaches, causing the river to change from

gaining to losing as it infiltrates its banks and recharges groundwater. Australian rivers are in

general fed by surface aquifers, not by overland flow, snowmelt, or direct rainfall. Except for

those areas in Australia with exceptionally high rainfall and surface runoff, the existence of

permanent standing water indicates substantial dependence on groundwater inflows (Nevill et

al., 2001).

In addition to environmental controls, anthropogenic factors such as groundwater pumping and

land use can have significant impacts on the movement of water between groundwater and

surface water resources. This raises an important issue for the management of water availability.

As this thesis is specifically concerned with water quality aspects of connected water resources,

issues associated with site-specific surface water or groundwater extraction will not be

discussed further. On the basis of the physical principles summarised above, the following

section examines factors that affect chemical river-aquifer interactions.

2.2.2 Chemical interactions

The chemistry of surface and subsurface waters is largely influenced by the geology and the

contact time that water has with the geological materials. Microorganisms in the soil, sediments

and water also affect the chemical characteristics of groundwater and surface water. In local

flow systems, groundwater has a relatively short time in contact with aquifer minerals, which

can result in minimal chemical changes prior to surface water discharge (Winter et al., 1998).

However, in the shallow environment groundwaters can be in their most aggressive state or be

exposed to a large store of salts; in addition, being closest to the land surface, shallow

groundwater is the most vulnerable to contamination from anthropogenic sources. Therefore,

the interaction between local flow systems and surface waters is important for water quality

management of connected resources. In contrast, water in deeper flow systems has longer

flowpaths and hence a longer contact time with subsurface materials for products of

geochemical weathering to accumulate. The longer residence time of deep groundwater can also

mean that the effects of point source or diffuse pollution to the groundwater system can take

longer to detect if the groundwater is ultimately discharged to a stream (and can persist for

longer after a change is made). Therefore, in connected systems where shallow and/or deep

Page 40: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 2

16

groundwaters are connected to a stream, the chemistry of receiving waters can be influenced by

groundwater or surface water having different histories and distinctive chemical signatures.

As water infiltrates through the unsaturated zone of the subsurface and into the saturated zone it

undergoes biogeochemical transformations. Although the most dramatic changes in

hydrochemistry occur in the soil and unsaturated zones, water quality in the saturated zone

progressively evolves along flowpaths towards areas of discharge, recording a chemical

signature that is often closely related to aquifer characteristics and residence time (Herczeg and

Edmunds, 2000). The main biogeochemical reactions that affect the transport of chemicals in

surface water and groundwater include: acid-base reactions; precipitation and dissolution of

minerals; sorption and ion exchange; oxidation-reduction (redox) reactions; biodegradation; and

dissolution and exsolution of gases (Winter et al., 1998). Figure 2-3 summarises the

hydrological cycle and the major processes that influence chemical evolution of water in the

subsurface as it progresses towards surface water features. Many of these processes also directly

influence surface water chemistry, although the relative influence of each process may differ

from that in the subsurface.

Figure 2-3 Processes that influence subsurface hydrochemistry (after Back et al. 1993, in Herczeg and Edmunds 2000).

The hyporheic zone, as defined by Winter et al. (1998), is the subsurface zone where stream

water flows through short segments of its adjacent bed and bank sediments, thereby creating a

mixing zone between subsurface and surface waters (Figure 2-4). Due to this mixing, the

chemical composition of the intervening hyporheic zone may differ significantly from that in

Page 41: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Connected Water Resources

17

both ground and surface waters. The hyporheic zone is typically a region of intensified

biogeochemical activity; these biogeochemical processes can in turn affect the movement of

nutrients and other chemical constituents, including contaminants, in both directions across the

streambed (Sophocleous, 2002; Winter et al., 1998). This has important implications for stream

health and ecological function. For example, microbes and algae attached to sediment in

hyporheic zones can uptake nutrients such as nitrate and hence lower the concentration of

dissolved nitrogen in the stream (Winter et al., 1998). Consequently, the hyporheic zone can be

viewed as a biological filter which can affect the chemistry of groundwater entering surface

water and also that of surface water entering groundwater.

Figure 2-4 The hyporheic zone as the interface between local and regional groundwater flow systems and surface waters (Winter et al., 1998).

The hyporheic zone is one example of an environment that can attenuate nitrogen

concentrations in groundwater. Lamontagne et al. (2001) note that there are two additional

hydrogeological environments: riparian zones and suboxic/anoxic aquifers, which have the

potential to remove a large fraction of nitrogen from polluted groundwater before it reaches

sensitive ecosystems. Issues surrounding the role of riparian zones in nitrogen attenuation have

been discussed by several authors (Rassam et al., 2006; Lamontagne et al., 2005; Malard et al.,

2002). Effective riparian buffers are generally characterised by a shallow impermeable

geological layer, which forces groundwater to move laterally through shallow root zones and

organic-rich deposits (Hill, 1996). Similarly, shallow unconfined aquifers with significant inputs

of surface-derived dissolved organic carbon, or aquifers containing organic matter/reactive

reduced material, can also attenuate nitrate from groundwater travelling through them

(Lamontagne et al., 2001).

The distinction between the three hydrogeological environments for nitrogen attenuation

discussed above is related to scale, ranging from whole of catchment (for reduced aquifers),

floodplain (for riparian zones), and river bed (for hyporheic processes) (Lamontagne et al.,

2001). However, as noted by Winter et al. (1998), an important characteristic of hyporheic

zones is that they represent an area ‘where groundwater that drains much of the subsurface of

landscapes interacts with surface water that drains much of the surface of landscapes’.

Page 42: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 2

18

Therefore, although hyporheic exchange can be viewed on the one hand as a small-scale

process, it also represents the interaction of surface and subsurface waters influenced by

catchment-scale processes. Woessner (2000) argues that stream-groundwater exchange should

be conceptualised and characterised more holistically at both the fluvial plain and channel-bed

scale, by multidisciplinary researchers, not hydrogeologists alone.

The significance of riparian and hyporheic zones for nitrogen attenuation is not well understood

in tropical and semi-arid river systems, as much of the research has taken place in temperate

regions, generally outside of Australia (Lamontagne et al., 2001). Although the focus of this

study is not specifically on riparian or hyporheic zones, or the complex hydrochemical reactions

at the groundwater-surface water interface, these environments are important controls on what

ultimately reaches the river system. Therefore, it is important to bear in mind the potential role

of these biological systems in river nutrient budgets.

2.3 FRAMEWORK FOR A NUTRIENT BUDGET

The overarching theme of this thesis is concerned with the implications of surface-groundwater

interactions for river nutrient budgets. A nutrient budget for nitrogen can be conceptualised as a

series of interconnected compartments or storages, each of which contributes and/or mobilises

N in its various forms to surface water. As illustrated in Figure 2-5, the key storages are:

atmosphere, rainfall, agriculture, soil, aquifer (shallow and deep), riparian/ hyporheic zone, and

surface water. Whilst point source inputs are not explicitly included, they are potential sources

of N to surface and subsurface storages. Upstream contributions also constitute an important

storage entity as an input into the system. The agriculture storage incorporates vegetation,

including crops; therefore, N can be added to the soil storage through applied fertilisers and/or

organic decay, or removed from the soil storage through uptake by crops and other vegetation.

Similarly, as depicted by the double arrows in Figure 2-5, there is the potential for N exchange

in both directions between other storages, such as between the aquifer storages and surface

water. For the schematic purposes of Figure 2-5, no distinction is drawn between dissolved

inorganic and organic forms of N, with both types represented as dissolved N. In addition, it is

assumed that particulate forms of N are not mobilised to groundwater.

The framework depicted in Figure 2-5 is a description of each subsystem, or node, at a

particular location along a river; the stretch of river between two nodes is referred to as a reach,

which has a corresponding contributing subcatchment area. The sum of each N contribution

(source term) arising from the various storages within a subsystem determines the net load of N

delivered to the downstream node of a particular river reach. The calculated N load for a reach

in turn provides the upstream input load for the adjacent downstream reach (except for the most

upstream node for which the input load is only from the upper catchment area). The river N

Page 43: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Connected Water Resources

19

budget is thus conceptualised as a series of subsystems that are linked spatially by the

contributing N load from the surface water storage in the adjacent upstream subsystem (Figure

2-6). The identification of source strength (of dissolved or particulate N), both within storages

of individual reaches and between river reaches, is important for prioritising management

interventions in a catchment. Within the budget framework the storages of agriculture and

riparian/ hyporheic zone (N attenuation) represent two controls that can be varied in order to

achieve different outcomes (N loads) in the river.

DEEP AQUIFER

UPSTREAM LOADS

SHALLOW AQUIFER

SOIL

AGRICULTURESURFACE

WATER

RAINFALLATMOSPHERE

RIPARIAN / HYPORHEIC

ZONE

dissolved N

particulate N

dissolved N

dissolved N

dissolved N

dissolved N

dissolved N

dissolved N

dissolved N particulate N

particulate N

dissolved N

gaseous N

dissolved N

particulate N

particulate N

particulate N

DEEP AQUIFER

UPSTREAM LOADS

SHALLOW AQUIFER

SOIL

AGRICULTURESURFACE

WATER

RAINFALLATMOSPHERE

RIPARIAN / HYPORHEIC

ZONE

dissolved N

particulate N

dissolved N

dissolved N

dissolved N

dissolved N

dissolved N

dissolved N

dissolved N particulate N

particulate N

dissolved N

gaseous N

dissolved N

particulate N

particulate N

particulate N

Figure 2-5 A conceptual representation of a nitrogen budget. Note that atmospheric losses of N are in the gaseous state.

Upper catchment

N load

DownstreamUpstream Nitrogen transport along the lower Herbert River

Upper catchment

N load

DownstreamUpstream Nitrogen transport along the lower Herbert River

Figure 2-6 Spatial connections of individual subsystems represented in Figure 2-5. Note that adjacent subsystems or nodes are connected by a river reach, which has a corresponding contributing subcatchment area.

Page 44: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 2

20

This thesis is concerned with the potential movement of N between the surface water and

aquifer storages. Therefore, the nitrogen budget framework provides a context for assessing

river-aquifer connectivity and the consequent transport of N.

2.4 ASSESSMENT METHODS

As outlined in Chapter 1, there are two main analytical components to the thesis: (1)

characterising river-aquifer interactions, and (2) ascertaining the role of groundwater in the

nitrogen budget of river systems. Accordingly, this section describes techniques for

investigating river-groundwater interactions and an overview of approaches used to characterise

nutrient mobility, particularly in surface and subsurface waters. This discussion forms the basis

for the research approach proposed in Chapter 3.

2.4.1 Characterising river-groundwater interactions

There are a variety of techniques for assessing the nature and degree of stream-aquifer

interaction processes. As summarised by Brodie et al. (2007), the range of methods include:

seepage measurement, field observations, ecological indicators, hydrogeological mapping,

geophysics and remote sensing, hydrographic analysis, hydrometric analysis, hydrochemistry

and environmental tracers, artificial tracers, temperature studies, water budgets, and modelling.

These methods can be described in the context of spatial scale, temporal scale, cost, ease of use,

advantages, limitations, and application. As this thesis is concerned with the catchment scale,

this section provides an overview of approaches that are applicable to regional scale studies.

However, it is noted that the findings of large scale studies have implications at the local scale,

such as for water quality protection and ecosystem health. The approaches are broadly grouped

as hydrogeological, hydrometric, hydrological, hydrochemical, GIS-based and modelling.

2.4.1.1 Hydrogeological approaches

Knowledge of the hydrogeology surrounding a surface water feature is critical to characterising

connectivity. This includes an understanding of the factors that control groundwater flow

(Section 2.2.1), such as aquifer geometry, host geology, stratigraphy, hydraulic properties,

geological structures and river morphology (Brodie et al., 2007). Standard techniques for

conceptualising the hydrogeology involve interpretation of borehole (lithological logs and water

levels) and pump test data using a variety of methods. These include construction of cross

sections and three dimensional stratigraphic maps; bore hydrograph analyses; and mapping of

groundwater level contours and flowlines (Domenico and Schwartz, 1990; Fetter, 1988; Heath,

1987; Freeze and Cherry, 1979). Limitations with hydrogeological approaches are that

compilation and analysis of the data can be time consuming and complex; limited borehole data

can also lead to misinterpretation (Brodie et al., 2007).

Page 45: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Connected Water Resources

21

2.4.1.2 Hydrometric approaches

Hydrometric methods are based on Darcy’s Law for fluid movement in a porous medium, and

are thus concerned with the hydraulic gradient between groundwater and surface water systems

and the hydraulic conductivity of the intervening aquifer (Brodie et al., 2007). Darcy’s Law can

be applied to estimate the rate and direction of groundwater flux to or from a surface water body

by measuring the head difference between the stream level and groundwater level; the vertical

distance between the measuring point in the aquifer and the streambed; and the vertical

hydraulic conductivity of the material along the vertical flowpath between the groundwater

measuring point and the stream bed. However, the approach can be an oversimplification of the

groundwater flow conditions (Woessner, 2000), in that the flux between the stream and shallow

aquifer are assumed to be entirely vertical. Flownet analysis is another approach for estimating

groundwater seepage rates in the horizontal or vertical plane (Loaiciga and Zekster, 2002).

While this approach can provide a simple and cost-effective way to estimate seepage flux, the

method can not account for spatial variability and local groundwater factors. Quantitative

methods for estimating seepage fluxes also rely on reasonable estimates of the hydraulic

conductivity, which can vary along a flowpath (Brodie et al., 2007).

A qualitative approach for determining the hydraulic gradient is to compare stream and

groundwater levels and hence establish the potential direction of seepage. Furthermore,

comparison of stream and bore hydrographs can indicate temporal changes in the hydraulic

gradient and whether a reach is likely to be gaining or losing at different periods in time

(Section 2.2.1). Groundwater level contours in the vicinity of a stream or lake can also indicate

the relationship between an aquifer and surface water feature (Winter et al., 1998). Whilst these

qualitative methods provide a rapid assessment of potential seepage directions between an

aquifer and a river, their effectiveness relies on having stream level and groundwater level data

at close proximity.

2.4.1.3 Hydrological approaches

Historical streamflow data form the basis of hydrologic methods for characterising surface

water outflow from a catchment. Hydrograph separation techniques aim to separate the stream

hydrograph into slowflow and quickflow components, and thus isolate the low-frequency signal

of a stream. This is useful for characterising the magnitude and timing of groundwater discharge

to a gaining stream. Baseflow separation techniques include digital filtering (Furey and Gupta,

2001; Chapman, 1999; Nathan and McMahon, 1990; Lyne and Hollick, 1979), graphical

methods (McNamara et al., 1997; Linsley et al., 1958), and chemical separation approaches

(Laudon and Slaymaker, 1997; Hooper and Shoemaker, 1986). Strictly speaking, baseflow

separation techniques that are based purely on the observed streamflow are actually estimating

the slowflow component. Whether this is actually baseflow depends on the assumptions used

Page 46: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 2

22

and whether the filter is appropriate for the system being studied. In this thesis, the output from

the filtering techniques is assumed to be baseflow. Analyses of flow duration curves are another

common method for determining the significance of baseflow contributions to a stream

(Smakhtin, 2001). Whilst these hydrological approaches are useful desktop tools, they rely on

the availability of accurate stream gauging information, which may be of limited spatial extent.

In addition, they do not provide information on the spatial distribution of groundwater input

along the stream.

Water budgets such as river-reach water balances are an additional hydrological approach,

whereby estimates are made of the unaccounted difference in the water balance for a specified

river reach, commonly between two gauging stations (Bratten and Gates, 2003; REM, 2002).

Whilst this method allows the flow volume associated with stream-aquifer interactions to be

inferred, it relies on the accurate measurement of surface water flow, which can have

considerable measurement errors. In addition, the flow volume is a net measurement; therefore,

it does not account for all the other potential gains and losses for the reach (Brodie et al., 2007).

2.4.1.4 Hydrochemical approaches

Dissolved constituents in water have the potential to retain a ‘memory’ of the movement and

interactions of water through surface and subsurface storages. The application of environmental

tracer methods for characterising river-aquifer interactions is based on the principle that surface

waters and groundwaters may have different tracer contents, which allows differentiation of

their sources (REM, 2002). Furthermore, the analysis of groundwater chemistry data in

conjunction with groundwater hydraulic information is useful for building a conceptual model

of the hydrogeological system. There are a variety of tracer techniques to identify gaining and

losing-river reaches and to quantify the groundwater-surface water flux. Commonly used tracers

include field parameters, major ions, stable and radioactive isotopes, trace elements, and

industrial chemicals. Different tracers provide information on different processes; therefore, a

multi-tracer approach is commonly used (Cresswell and Herczeg, 2004; Thayalakumaran et al.,

2004; Cook et al., 2003; Lamontagne et al., 2003; Herczeg et al., 2001). The spatial and

temporal resolution, reliability, and degree of quantification may also vary between tracers;

thus, a combination of techniques provides a more robust process understanding. A major

drawback of hydrochemical approaches is the generally high cost associated with sampling

logistics and laboratory analyses.

2.4.1.5 GIS-based approaches

Geographic information systems (GIS) can be used as a mapping tool for spatial representation

of aspects of the hydrogeology, hydrology and hydrochemistry, and hence aid with the

conceptualisation of processes at a catchment scale. A more sophisticated application of a GIS

is to characterise groundwater-surface water connectivity by integrating geophysical and remote

Page 47: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Connected Water Resources

23

sensing imagery with other spatial datasets. Remote sensing surveys allow for rapid, non-

invasive mapping of landscape parameters that either indicate or control groundwater-surface

water interactions. In addition, they can provide good spatial resolution in the vicinity of surface

water features, including information on changes through time. However, undertaking and

interpreting these surveys can be complex, requiring expensive equipment, technical expertise

and logistical support (Brodie et al., 2007).

2.4.1.6 Modelling approaches

Merritt et al. (2003) review the three basic model types: empirical, conceptual and physically-

based. Empirical models are generally the simplest of the model types; they are primarily based

on the analysis of observations and seek to characterise response from these data.

Computational and data requirements for such models are usually less than for conceptual and

physics-based models, often being capable of being supported by coarse measurements. As a

result, this class of model tends to produce outputs at high spatial and temporal aggregation. In

comparison, conceptual or physics-based models tend to be more dynamic, capturing greater

spatio-temporal resolution. Conceptual models are typically based on the representation of a

catchment as a series of internal storages and thus include a general description of catchment

processes without including the specific details of process interactions. Whilst they tend to be

aggregated, the distinction between conceptual and empirical models is that they still reflect the

hypotheses about the processes governing system behaviour. Physics-based models are based on

the solution of fundamental physical equations and therefore can involve large numbers of

parameters. These models tend to have the greatest spatial and temporal resolution; however,

the physical significance of model outputs due to scale issues is questionable. In addition,

detailed model approaches (i.e. mechanistic, spatially distributed) require large efforts in model

development, calibration and validation for a range of conditions (Wolf et al., 2005). Despite

their relative simplicity, empirical models are frequently used in preference to more complex

models as they can be implemented in situations with limited data and parameter inputs and are

particularly useful as a first step in source identification (Merritt et al., 2003).

Numerical modelling approaches can provide a valuable tool for combining into a consistent

framework the information obtained from the approaches described above. Existing models that

simulate river-aquifer interactions fall into the broad categories of surface water models,

groundwater models, combined models, and fully integrated models (REM, 2002). Numerical

models are useful as a predictive tool for management and policy. However, oversimplified

models may not be adequately robust, while complex models can increase data requirements,

costs and time (Brodie et al., 2007).

Page 48: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 2

24

2.4.2 Characterising nutrient mobility in water

Approaches for characterising and predicting the movement of nutrients between surface water

and groundwater range from qualitative field-based studies to quantitative nutrient modelling.

With a particular focus on N as the nutrient of interest, this section provides a brief overview of

these contrasting approaches.

2.4.2.1 Field-based studies

A plethora of documented studies have focused on nutrient delivery to and export from surface

waters (Brodie and Mitchell, 2005; Eyre and Pont, 2003; Mitchell et al., 2001). Similarly, there

are numerous studies on nitrate leaching to groundwater, particularly in agricultural catchments

such as below sugarcane (Thorburn et al., 2003; Bohl et al., 2001; Rasiah and Armour, 2001;

Weier, 1999; Verburg et al., 1998; Lawrence, 1983). Other field-based studies consider nutrient

balances in the air-soil-water system of agricultural catchments to determine the efficiency of

fertiliser management practices (Meier et al., 2006; Thorburn, 1999; Bristow et al., 1998; Prove

et al., 1997; Freyney et al., 1994). More recently, there has been heightened research interest in

the interactions between surface and groundwater systems and the implications of this

connectivity for N movement, particularly from groundwater to rivers or coastal waters.

Investigations range from purely groundwater N studies that infer implications for transport to

surface waters (Ahern et al., 2006; Merrill and Benning, 2006; Rasiah et al., 2003) to more

comprehensive studies of N cycling between groundwater and adjacent surface waters during

different time periods (Lamontagne et al., 2005; Lamontagne et al., 2002). Characterising the

hydrogeochemical environment using tracers in addition to N, have also been the focus, or a

component of, studies concerned with N cycling (Grimaldi et al., 2004; Thayalakumaran et al.,

2004; Korom, 1992).

As discussed in Section 2.2.1, surface water-groundwater interactions occur at various spatial

scales (Woessner, 2000). Furthermore, key biogeochemical reactions involving N, such as

nitrate reduction, occur in distinct hydrogeological environments that differ in scale

(Lamontagne et al., 2001). Accordingly, studies concerning N transformations and movement

between surface and groundwaters have been represented as regional (catchment), floodplain

(riparian zone) or riverbed (hyporheic zone) scale investigations, or a combination. The scale of

investigation depends largely on the underlying research questions, for example, whether there

is an interest in small scale biogeochemical interactions at the surface-groundwater interface

(Hefting et al., 2006; Rassam et al., 2006; Young and Briggs, 2005; McKergow et al., 2004;

Lamontagne et al., 2003; Malard et al., 2002; Boulton et al., 1998; Cirmo and McDonnell,

1997), or whether large scale mechanisms such as discharge of regional groundwater and

nutrients to surface waters are relevant (Linderfelt and Turner, 2001). Exchanges between near-

channel and in-channel water are critical to stream restoration and riparian management efforts

Page 49: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Connected Water Resources

25

(Sophocleous, 2002). Hence, given the ecological implications of N cycling at the surface

water-groundwater interface, there has in general been greater research attention towards reach

scale rather than catchment scale investigations.

2.4.2.2 Nutrient modelling

Due to the capacity of a catchment to store N in soil, vegetation and groundwater, it may take

considerable time for the full impacts of anthropogenic activity to be reflected in river nitrate

quality. Therefore, models are considered essential for predicting how changes in deposition,

land use, management and climate will affect N loading to rivers (Whitehead et al., 1998).

Efforts to model diffuse nutrient losses within catchments are intensifying, with the types of

model used varying from simple lumped and GIS representations to fully distributed models

(Neal and Heathwaite, 2005). There is an extensive body of literature on different nutrient

models that have been used for a variety of applications. As reviewed by Wolf et al. (2005),

existing nutrient models differ with respect to: (1) modelling approach (e.g. mechanistic,

empirical, equilibrium); (2) input data aggregation (e.g. spatially distributed, lumped); (3)

calculation time step (e.g. hourly, yearly, average annual); and (4) spatial unit or scale of

modelling (e.g. plot, catchment). According to Wade et al (2005), ‘whilst purely empirical

approaches may fit the observed patterns, only dynamic, process-based models can also

represent the likely future changes in the driving factors, the catchment N stores and the

hydrological pathways, and therefore predict the change in the catchment hydrology and stream

water nitrate concentrations’.

For some studies, there may be a requirement for integrating multiple variables and modelling

scenarios (such as potential impacts) at different spatial and temporal scales. Detailed numerical

modelling may have a role in these situations. In contrast, nutrient budgets, or mass balance

approaches, summarise nutrient inputs and outputs of a defined system over a defined period of

time. Hence, budget approaches, which are an example of an empirical type of model, are often

static or produce outputs at large temporal scales such as annually. Budgets are a valuable tool

for summarising large amounts of information in transparent and easily understood input-output

diagrams and are flexible in that they allow for approximations and revisions where there are

gaps in data. However, there are various sources of uncertainty in the budget approach due to

biases (e.g. personal, sampling, measurement, data choice and manipulation) and errors related

to spatial and temporal variability (Oenema et al., 2003). The conceptual framework presented

in Figure 2-5 illustrates the input-output structure of a nutrient budget. Input data for nutrient

budgets can be classified according to: (1) type (primary, estimate or assumptions); (2) source

(field/ laboratory measurement, observation, computation); and (3) frequency (continuous,

seasonal, annually). Nutrient budgets are commonly based on a combination of data types which

are derived from various sources and are collected at different frequencies. Ultimately, the

Page 50: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 2

26

purpose of the study defines the budgeting approach, scale, data acquisition strategy and

required accuracy and precision of the nutrient budget (Oenema et al., 2003).

2.4.3 Implications for the research approach

Many of the techniques for characterising groundwater-surface water interactions (Section

2.4.1) are standard methods historically used by separate disciplines; however, it is the

combination of approaches applied to a connected water resource problem that provides a

powerful toolbox to qualitatively or quantitatively assess surface water-groundwater linkages.

Therefore, whilst there is no method, formula or computer model that can be universally

applied, different datasets and techniques can be explored at different stages in the assessment

to inform the various aspects of the investigation (REM, 2002). Factors to consider in the choice

of method(s) include: the water resource question and objectives; availability of data and other

resources; location and characteristics of the study area; assessment time; and scale/type of

output required. In light of the above considerations, a range of hydrogeological (Chapters 4 and

5), hydrometric and hydrological (Chapter 6), and hydrochemical (Chapters 5 and 7) techniques

are regarded as appropriate for the objectives of this study, with use of a GIS as a mapping tool.

The chosen techniques complement one another such that together they provide a more

complete picture of groundwater-surface water interactions, not provided by any one method.

Based on standard reference texts and applications in other studies, the general principles

underlying the methods employed in the thesis are outlined in the relevant analytical chapters.

As outlined in Section 2.4.2, field-based and modelling approaches represent the two broad

categories for investigating the mobility of nutrients in the environment. Nutrient modelling

approaches are concerned with how much N is ultimately delivered to the river; they do not

explicitly require a detailed understanding of the underlying hydrological or biogeochemical

processes for N gains/losses from source to sink. Whilst a nutrient budget, such as the

conceptual nitrogen budget (Figure 2-5 and Figure 2-6), is a simple approach for identifying

source strength within and between reaches, of interest to this study are the physical processes

that underpin the transport of N to surface waters, particularly in relation to river-aquifer

interactions. Due to the complex transformation reactions that are associated with N transport in

the environment, characterising the dynamics of water movement between river and aquifer

storages is a first step to understanding the mobility of dissolved N between them. Measured

concentrations of N in groundwater and along the river are hence interpreted in light of this

process understanding. Thus, a broad scale field-based approach is considered appropriate for

investigating the significance of groundwater N fluxes to the river, which is founded on an

understanding of the hydrology, hydrogeology, hydrometrics and hydrochemistry of the river-

aquifer system. This is the basis for the research approach, as outlined further in Chapter 3

(Section 3.4). Whilst it is considered that completely parameterising a nutrient budget or other

Page 51: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Connected Water Resources

27

models is outside the scope of this thesis, the interpretation of outputs from these approaches

can be enhanced with reference to a conceptual framework based on process understanding of

the system.

2.5 CHAPTER SUMMARY

Chapter 1 provided a context for the importance of considering integrated water resources,

whereby groundwater and surface water are managed as a single resource. Based on a review of

studies concerning groundwater-surface water interactions, it was also highlighted that the role

and significance of groundwater for river nitrogen budgets is not well understood. Therefore,

two key analytical components to this research were identified: (1) characterisation of river-

aquifer interactions and (2) assessment of the potential role of groundwater for the nitrogen

budget of river systems. This chapter introduced the key physical concepts surrounding

connected water resources and highlighted the implications of surface-groundwater interactions

for water quality. Assessment tools and methods available to characterise these interactions

were also examined: many of these techniques are utilised in this thesis. In addition, approaches

for understanding nutrient transport and fluxes were reviewed, which ranged from field-based

studies at various spatial scales to complex numerical modelling at different spatial and

temporal resolutions. Given the focus on process understanding and the extensive data

requirements which could not be met with existing data, a modelling approach was considered

beyond the scope of the thesis. However, it was noted that the interpretation of model outputs

can be improved with process knowledge of key components such as river-aquifer linkages.

Therefore, based on considerations such as the key research questions, data availability,

characteristics of the study area and the scale of the study, it was established that hydrochemical

techniques, combined with hydrogeological, hydrometric and hydrological methods, are

appropriate for characterising river-aquifer interactions at a catchment scale. The implications

for N cycling will hence be examined with regards to process understanding of how water

moves between surface and subsurface systems.

The following chapter reviews tropical hydrological processes and issues concerning water

quality. The case study area is presented for the research, with previous N studies outlined. The

broad research approach is described in the context of the overarching research questions of the

thesis. Based on the characteristics of the case study catchment and existing datasets, the

methodology for further data collection is also presented.

Page 52: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to
Page 53: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

29

Chapter 3 River-Aquifer Interactions in the Wet Tropics

3.1 INTRODUCTION

Since the 1960’s there has been considerable progress in process hydrology research in the

temperate latitudes, while the humid tropics have received comparatively less attention. As a

result, findings from temperate regions have tended to be transferred to tropical environments

(Bonell and Balek, 1993). In the Australian context, hydrology studies have concentrated on

sub-humid and semi-arid climates and landscapes (Fleming, 1993). Due to world-wide concerns

regarding the impact of tropical forest clearance, international research specifically related to

hydrology in the tropics has generally focused on the environmental and human impacts of

deforestation and land cover change in large humid catchments of South East Asia, Africa and

South America (Bonell and Bruijnzeel, 2005; Bonell et al., 1993). In comparison, hydrological

studies in the Australian tropics have largely been undertaken in agricultural catchments that no

longer preserve large areas of native vegetation, where the issues addressed are largely

concerned with agricultural impacts on water resources (Rasiah et al., 2003; Thorburn et al.,

2003). Despite the different foci of overseas and Australian studies, the international literature

provides insight into key features that differentiate hydrological conditions in the tropics from

those in temperate climates. The review that follows draws on the available literature to shed

light on river-aquifer interactions and nutrient transport, specifically in tropical climates. This

subsequently leads to a discussion of the case study catchment in the wet/dry tropics of north

Queensland.

3.1.1 Climate

Geographically, the tropical regions are approximately bounded by the Tropics of Cancer

(23o27’N) and Capricorn (23o27’S). The humid tropic regions are defined as areas where the

mean temperature of the coldest month is above 18oC, the duration of the wet season exceeds

4½ months and a wet month has on average 1000 mm of rainfall (Chang and Lau, 1983). Under

this definition, approximately 12% of Australia can be classified as humid tropics, which

includes the wet-dry tropical sub-region of northern Australia (Fleming, 1993; Stewart, 1993).

The remainder of tropical Australia is considered to be in the dry tropics, where the duration of

the wet season is less than 4½ months. A great amount of solar energy in the tropics creates a

climate without harsh winters; therefore, solar energy influences the hydrological cycle more

directly in the tropics than in other climatic zones (Latrubesse et al., 2005). In tropical areas,

Page 54: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 3

30

rainfall is considered to be the main factor that determines the seasons and is also the most

variable element of tropical climate (Callaghan and Bonell, 2005; Latrubesse et al., 2005).

According to Bonell and Balek (1993), the magnitude of rainfall is one of the driving forces in

differentiating between the response of hydrological processes in the humid tropics compared

with processes in higher latitudes.

Rainfall in tropical northern Australia is distinctly seasonal. Due to the geographic location, the

Australian tropics are subject to monsoonal influences and tropical storms and cyclones

(Fleming, 1993). The highest rainfall areas are associated with the Great Dividing Range in the

east, exceeding 3200 mm annually along the wet tropical coast (Stewart, 1993).

3.1.2 Features of streamflow and groundwater

Tropical rivers, in general, show high but variable peak discharges during the rainy season and

periods of low flow when rainfall decreases. These rivers can be grouped into two main types:

(1) rivers with well defined high and low discharges corresponding to unimodal rainy periods;

and (2) rivers with two flood peaks per year, consistent with bimodal rainy periods. Tropical

rivers in Australia tend to have highly seasonal flow regimes, with much of their annual flow

between November to May (Hamilton and Gehrke, 2005). Individual rivers in northern

Australian catchments can have multiple major flows each year (such as the Tully), one major

annual flow (such as the Herbert) or major flows that are separated by several years (such as the

Burdekin). In both the wet and dry tropical catchments of Australia, the two distinct flow

conditions are relatively well separated in time, with only short periods dominated by high

flows compared to the otherwise dominant low flow conditions (Brodie and Mitchell, 2005).

Higher rainfall intensities in the tropics lead to differences in runoff generation (especially

during tropical storm events) compared to non-tropical areas (Bonell and Balek, 1993). Based

on research in northeast tropical Queensland, Bonell et al. (1998) found that ‘at low rainfall

intensities, precipitation is routed to streamflow via the groundwater/soil system. However, at

high precipitation rates, the flow capacity of these pathways is exceeded and alternative

pathways involving rapid flow are invoked which effectively short-circuit the former pathways’.

Hence, while vertical movement of soil water prevails between rain events and during most

small events, under heavy and prolonged rain the horizontal transfer of water develops to

produce lateral flow (Douglas and Guyot, 2005). This is in contrast to other climatic and

geographic regions where the event hydrograph consists of pre-event or ‘old’ water

(groundwater/soil water), rather than event or ‘new’ water from rainfall/overland flow. It is

suggested that any shift in the delicate balance between rainfall intensity, soil hydraulic

properties and topography can provide widely different runoff processes, both within and

between various tropical forests (Bonell and Balek, 1993).

Page 55: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

River-Aquifer Interactions in the Wet Tropics

31

In parallel with differences in surface hydrological processes, there are fundamental differences

in groundwater behaviour with regard to mechanisms of recharge and discharge under humid

and sub-humid conditions. Aquifers in the humid tropics tend to fill up rapidly in the wet

season, with the watertable virtually reaching the land surface; further excess rainfall is rejected

because of the absence of storage space, which leads to overland flow (Bonell and Balek, 1993).

In addition, due to a shallow watertable, the volume of aquifer discharge may increase rapidly

following groundwater recharge from excess rainfall, and hence contribute to the peak runoff

response of tropical catchments (Foster and Chilton, 1993). Callaghan and Bonell (2005)

indicate that groundwater remains a neglected area of research in the humid tropics and

highlight the need for better coupling of surface water-groundwater interactions in future

assessments. Geographically, large areas of the humid tropics are underlain by aquifers of

varying geological character. Most of the groundwater system types are characterised by

shallow watertables directly connected to surface waters, which makes groundwater highly

vulnerable to pollution from a range of human activities (Foster and Chilton, 1993). This is

exacerbated by the high rates of precipitation in the humid tropics which cause rapid leaching of

pollutants from urban, agricultural and industrial wastes. According to Bonell (2005), despite

recognition of the connectivity between the surface hydrology with the hydrogeology, globally

there are only a few studies that have parameterised these groundwater systems.

3.1.3 Water quality and nutrients

There have been numerous studies concerned with the physical processes of erosion and

sedimentation in the humid tropics (Rose, 1993); however, little effort has been devoted to

specifically understanding the transport of dissolved constituents through surface and

groundwater pathways in tropical climates. Based on a review of water quality issues in the

humid tropics, Roche (1993) concluded that managing the microbiological aspects of water

quality control is the major issue in terms of direct effects on human health. Bonell and Balek

(1993) note that ‘there is a need for environmental isotope and other conservative tracer studies

in the humid tropics… for assisting studies in water-borne nutrient cycling’. With regard to

water quality, temperature is one of the parameters that differs the most between the temperate

and tropical zones, which has a direct influence on the physical, chemical and biological

properties of an aquatic medium (Roche, 1993). Although the magnitude of water quality

problems, particularly in relation to human health, is perhaps less in northern Australia

compared to other tropical countries, there are similar causes of water pollution. For example,

one of the most urgent water quality problems in the humid tropics is associated with the change

in salt and nutrient cycles and contents, because of land and water resources management.

Tropical systems have an important role in sediment and nutrient transfer to the oceans and

coastal areas (Latrubesse et al., 2005). In relation to northern Australian conditions, discharge of

Page 56: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 3

32

terrestrial material to the coast occurs predominantly during the major river floods generally

associated with cyclonic rainfall events between November and May (Furnas and Mitchell,

2001). Given the seasonal differences in discharge from monsoonal catchments, it is important

to distinguish between water quality in flow event conditions (flood pulse) from baseflow

conditions (Brodie and Mitchell, 2005). Flow events in northern Australian rivers are

characteristically short and energetic, with water residence times in the river of about one week.

Therefore, water quality measurements during flow events provide information on catchment

contaminant loads (suspended sediment, nutrients and pesticide residue) discharged to

downstream environments. In contrast, nutrient concentrations measured during baseflow

conditions indicate the water quality status which persists for much of the year, which

influences the health of in-stream ecosystems (Brodie and Mitchell, 2005). According to Harris

(2001), the baseflow period is when in-stream interactions between macrobiota, microbiota and

water chemistry have adequate time to fully progress; hence, there is a tight coupling between

water chemistry and water biology.

Given the focus on nitrogen (N) in this thesis, it is pertinent to discuss N inputs into tropical

rivers. Based on a review of studies in northern Australia, Brodie and Mitchell (2005) found that

waters draining pristine rainforest and woodlands have moderate concentrations of dissolved

organic nitrogen (DON), low to moderate concentrations of particulate N and low

concentrations of dissolved inorganic nitrogen (DIN); the dominant form of N being DON.

Similarly, in savannah woodlands and grasslands with low grazing intensity, N speciation is

dominated by DON. However, the authors note that with land clearing for agricultural and

urban development, N concentrations in receiving waters have increased and the form of N has

changed from organic (DON) to inorganic (nitrate, ammonia) forms that are more bioavailable

than DON. Furthermore, it has been suggested that although the concentrations of DIN are

generally low under ‘natural’ conditions, occasional high concentrations of nitrate may be

associated with groundwater discharge to the stream after the main peak flow (Brodie and

Mitchell, 2005). Furnas (2003) observed that in most of the rivers of the Great Barrier Reef

(GBR) catchment, nitrate concentrations are generally highest during wet season flood events,

usually during the first flow or flood event of the season (first flush) when large amounts of

water wash across and through catchment soils. However, Furnas (2003) concluded that in some

rivers such as the Herbert, characterised by having some rainfall throughout the year, the highest

nitrate concentrations occur at the end of the dry season when high nitrate groundwater inputs

make a larger relative contribution to water in the river. Other comparisons between the wet and

dry tropical rivers of the GBR catchment in regards to nutrient export characteristics are

summarised in Furnas (2003).

Page 57: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

River-Aquifer Interactions in the Wet Tropics

33

In a review of water and N balances, Bristow et al. (1998) found that deep drainage rates are

higher in both natural and agricultural systems in the Wet Tropics of north Queensland

compared to other regions of Australia. It is suggested that this characteristic, coupled with high

inputs of N in fertilisers, results in considerable N loss below the rootzone. In addition, Bristow

et al. (1998) propose that the intensity of the N cycle in the humid tropics is driven by constant

high temperatures and rainfall which enable year-round biomass production together with high

rates of decomposition and hence nutrient release; this high rate of nutrient release increases the

potential for leaching through tropical soils. Therefore, the role of N in groundwater is

potentially an important consideration in the wet/dry tropics of northern Australia.

3.2 SELECTION OF CASE STUDY AREA

The Herbert River in north Queensland (Figure 3-1) represents a useful study area in the tropics

to develop linkages between land-based activities, stream water quality and potential impacts on

the near-shore marine environment. The catchment is one of thirty-one that drain into the Great

Barrier Reef Marine Park, a marine ecosystem that is recognised internationally for its unique

biological and physical features (Johnson et al., 2000). In addition to its environmental value,

the Park has economic significance, supporting a 4 billion dollar tourism industry. It is believed

that diffuse pollution from cropping and grazing lands in adjacent catchments pose a significant

threat to the Reef. The Herbert is one of the Wet Tropics catchments that drains directly into the

area of the Great Barrier Reef lagoon considered under greatest threat from terrestrial runoff

(Productivity Commission, 2003).

Page 58: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 3

34

Figure 3-1 Location of the Herbert River catchment showing the Herbert River and its major tributaries. The lower catchment target area for this study is circled.

3.2.1 Catchment water quality issues

Concerns over sediment, nutrient and contaminant export to the Great Barrier Reef World

Heritage Area (GBRWHA) have driven much research into water quality in the Herbert River

catchment. The major water quality issues relate to the transport of nutrients and sediment; it

has been suggested that land management can be improved in the catchment to minimise off-

site exports (Bramley and Roth, 2002). Phosphorus (P) and nitrogen (N) are the nutrients of

most concern. Research has shown that P tends to be largely bound to sediment particles,

whereas the N budget is dominated by dissolved N from both runoff and subsurface flows

(Bartley et al., 2003). Given that N inputs into the system are correlated with fertiliser

application (Bramley and Roth, 2002), N is a link between land use/management and water

resource quality in both surface and subsurface resources. According to Mitchell et al. (1997)

and Brodie et al. (2001), excess N derived from the major land uses in the catchment has the

potential to impact on in-stream and near-shore water quality.

It is well documented that high concentrations of N, especially inorganic nitrate, are found

particularly in surface waters draining land under sugarcane (Bartley et al., 2003; Furnas, 2003;

Bramley and Roth, 2002; Bramley and Muller, 1999). Groundwater studies have also shown

there to be subsurface losses of nitrate beneath sugarcane growing areas (Weier, 1999). Given

an average fertiliser input of 160 kg N/ha/yr and an average output of roughly 80 kg N/ha/yr in

Page 59: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

River-Aquifer Interactions in the Wet Tropics

35

millable cane, is suggested that about 50 percent of the N applied in the sugarcane industry can

potentially be lost through denitrification, surface runoff and/or leaching to groundwater (Bohl

et al., 2001). Under current management practices of minimal soil erosion (e.g. minimum tillage

and green cane harvesting/trash blanketing), losses of nitrogen from cane land are attributed to

fertiliser as the major source of nutrients, rather than a sediment source (Brodie and Mitchell,

2005). Based on surface water studies, it is considered that the tendency for nitrate

concentrations to be elevated in the Herbert River during low flow periods is due to inputs of

high-nitrate groundwater (Furnas, 2003; Furnas and Mitchell, 2000). Furthermore, it has been

proposed that the quantity of nitrate introduced into the Herbert River is comparable to the

estimated loss of fertiliser N to groundwaters (Furnas and Mitchell, 2000). Based on a study of

radium isotopes in riverine muds in the Herbert River estuary, Brunskill (2000) also suggested

that groundwater inputs are important during medium to low river discharge periods. Whilst

these studies raise the possibility of groundwater discharge to the river, the interaction of

groundwater and surface water in the lower catchment has not been investigated in any detail. In

addition, despite concerns over N levels found in the Herbert River and in some of the shallow

aquifers, no previous work has explicitly assessed the role of subsurface N on surface water

quality in the catchment.

The target area for this study is the lower Herbert River catchment, due to the occurrence of

sugarcane farming which is considered to be the major contributor of N to surface waters

(Bartley et al., 2003; Bramley and Muller, 1999). In addition, several key datasets from

previous studies and ongoing monitoring are available for the lower catchment (Section 3.5).

3.2.1.1 Previous N studies on surface water

Water quality studies previously undertaken in the lower catchment have generally been

concerned with both N and P. Below is a summary of key findings from past research relating

specifically to N, for both surface water and groundwater resources.

Two major water quality studies in the lower catchment were previously undertaken by the

Australian Institute of Marine Science (AIMS) in collaboration with the Bureau of Sugar

Experimental Stations (BSES) (Furnas et al., 1995), and the Commonwealth Scientific and

Industrial Research Organisation (CSIRO) (Bramley and Muller, 1999). The AIMS study was

primarily concerned with determining riverine export of nutrients and fine sediment to the Great

Barrier Reef lagoon. The CSIRO sampling programme involved extensive water and soil

sampling to resolve source areas for the nutrients and sediment leaving the catchment.

Sampling by AIMS took place during 1989-1994 at three sites along the lower Herbert River.

Key findings from the study were that:

Page 60: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 3

36

• the concentrations of DIN and particulate N in river source waters are likely to be

similar throughout the catchment, or that chemical processes stabilise concentrations in

river and soil waters;

• the concentration of DON decreases downstream, indicating dilution with low-DON

water on the floodplain or in-stream consumption (bacterial mineralisation) of DON in

the lower reaches of the river;

• the concentration of nitrate is generally higher at the downstream sampling site than that

measured upstream, particularly during low flow periods;

• most of the nitrate exported is considered to come from a floodplain source, with

agricultural fertilisers the most likely source.

Of direct relevance to the current research is that according to Furnas (2003), the tendency for

nitrate concentrations to be elevated during low flow conditions and diluted during floods

suggests that inputs of high-nitrate groundwater are responsible for the higher concentrations;

dilution occurs because groundwater inputs from aquifers are relatively constant and not closely

coupled to surface runoff.

The CSIRO water quality sampling was undertaken during October 1992 – May 1995 at 33

surface water sites (reduced to 19 by the end of the study) around the lower catchment: the sites

were selected to reflect the major land uses, soil types and subcatchments. The CSIRO data

showed that:

• N concentrations tend to be greater downstream than upstream, indicating an export of

nutrients from land draining into the Herbert River between these areas;

• the concentration of N in streams draining land under sugarcane tends to be greater than

in streams draining other land uses (grazing and forestry);

• peak wet season events dominate the annual riverine flux of nutrients.

In addition, Bramley and Roth (2002) found that with respect to total nitrogen in the lower

catchment, approximately 31%, 9% and 3% of samples collected from streams predominantly

draining cane land, grazing and forestry, respectively, were above the ‘interim trigger levels for

assessing possible risk of adverse effects due to nutrients’ (ANZECC and ARMCANZ, 2000).

Bramley and Johnson (1996) further suggested that nutrient concentrations in the Herbert were

generally below ANZECC (2000) target levels for the protection of freshwater ecosystems,

except during high flow conditions when the trigger values were exceeded; hence, the authors

concluded that nutrient loss in the catchment is event based and insignificant outside the wet

season months. Although nutrient losses from intensively managed cane lands exceeded those

Page 61: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

River-Aquifer Interactions in the Wet Tropics

37

from other land uses, Bramley and Johnson (1996) also noted that these losses might be

expected to occur in these areas irrespective of crop type due to strongly seasonal climate and

high rainfall intensities. Nonetheless, Bramley and Roth (2002) concluded that land

management can be improved in the catchment to minimise this off-site export of nutrients.

A recent modelling study for the entire Herbert River catchment (Bartley et al., 2003) using a

sediment budget, SedNet (Prosser et al., 2001), combined with a nutrient budget component,

ANNEX (Young et al., 2001), was undertaken to spatially identify the major sediment and

nutrient sources. With regard to N, the study showed that dissolved forms of N predominate and

that the N budget is dominated by losses from canelands on the floodplain, followed by open

forests and cultivated areas of the middle and upper catchment. Nutrient scenario results

suggested that reduced fertiliser application rates on cane lands (reduced from 200 kg/ha/yr to

130 kg/ha/yr) could produce a decrease of 27% in dissolved inorganic N levels and a 10%

decrease in the overall N budget.

3.2.1.2 Previous N studies on subsurface water

Although the majority of water quality research in the catchment has focussed on surface water

rather than groundwater, the studies summarised below of Bohl et al. (2000a), Bohl et al.

(2000b) and Bohl et al. (2001), which are specific to the lower Herbert River catchment,

provide some important background information for this research in regards to subsurface N

losses under different soil types. A further study by Thorburn et al. (1999) modelled the impact

of trash retention on soil nitrogen, using the lower catchment as a case study site. The

magnitude of N losses to groundwater from sugarcane have previously been investigated for

various soils and areas in the Australian sugar industry such as: the freely drained basaltic soils

in the South Johnstone catchment (Rasiah et al., 2003; Hunter and Walton, 1997; Reghenzani et

al., 1996); under irrigated cane in the Bundaberg and Burdekin delta regions (Stewart et al.,

2005; Kuhanesan et al., 1998; Verburg et al., 1998); and in several coastal areas of northeast

Australia (Thorburn et al., 2003). Bristow et al (1998) summarised research relating to water

and nitrogen balances in natural and agricultural systems in the Wet Tropics of north

Queensland; recent modelling/experimental studies by Meier et al. (2006) and Thorburn et al.

(2005) add to this body of research.

Rasiah and Armour (2001) noted that nitrate leached below the crop-root zone may be adsorbed

onto soil, move laterally to discharge into streams and rivers, enter deep groundwater, and/or

denitrify in the profile. Furthermore, the amount of N adsorbed in soil depends on: anion

exchange capacity, net negative charge, ionic strength of the soil bulk solution, nitrate

concentration, competition with other anions, pH2O, pH, anionic composition of the soil

solution, cation exchange capacity, and leaching. Given the range of factors that potentially

affect N leaching, the transferability of findings from one study site to another in the Wet

Page 62: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 3

38

Tropics is questionable. For this reason only literature specific to the lower Herbert River

catchment is summarised below.

Based on field sampling over two wet seasons during 1997-1999, Bohl et al. (2000b)

determined the components of the water and nitrogen balance on a range of soils under

sugarcane in the Ripple Creek area in the lower catchment. Gaseous losses (such as

denitrification) or losses via surface runoff were found to be the major loss pathways for N

under wet conditions (up to 40% of the amount of applied nitrogen), rather than nitrate leaching

via subsurface flow (groundwater and interflow). However, on highly permeable soils, such as

the sandy soils of the riverbank, N losses to groundwater from fertiliser were found to be up to

45% of that applied. Large variations in N loss were observed between bore sites as well as

between sampling years, with the highest losses to groundwater under plant cane (compared to

ratoon) due to the less developed root system. Bohl et al. (2000b) suggested that the

distribution, size and intensity of rainfall after fertiliser application are important factors in

dictating N losses.

Bohl et al. (2001) expanded on the previous work by Bohl et al. (2000b) to examine the spatial

distribution of N leaching losses based on pedological (mapped soil types) and hydrological

features. The study found that the more freely draining soils of the alluvial fans and sandy river

banks had the highest leaching losses, while the lowest losses were estimated for the heavy soils

on the plain. The authors concluded that, in general, the risk of nitrate leaching is comparatively

low, with the exception of pedohydrological units characterised by high internal permeability

and high drainage in relation to landscape position.

3.3 CATCHMENT CHARACTERISTICS

The Herbert River catchment (Figure 3-1) can be divided into four distinct physiographic

sections (Bartley et al., 2003). The upper part is characterised by extensive cattle production and

other minor agricultural activities such as horticulture and dairy. The central part of the

catchment, comprising the deep Herbert River gorge, is predominantly Wet Tropics World

Heritage Area, State Forest and timber reserves. The lower catchment comprises the Herbert

River floodplain and the southern coastal section, containing a network of streams that drain

directly to the coast. Sugarcane farming dominates the banks of the major rivers in the lower

catchment, while the remaining area supports cattle grazing, native vegetation and plantation

forestry (Figure 3-2). Agricultural and pastoral production are the largest users of land in the

entire catchment, with less than 1% of the area allocated to industrial and urban uses (Johnson et

al., 2000).

Page 63: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

River-Aquifer Interactions in the Wet Tropics

39

Figure 3-2 Land cover in the lower Herbert River catchment (adapted from Bramley and Muller 1999). Note that the Herbert River runs through the sugarcane growing areas.

Land cover and land use in the catchment has changed substantially since European settlement

in the early 1860’s, particularly in the lower catchment (Johnson et al., 2000, 1999). Prior to

settlement, natural vegetation in the lower catchment was dominated by grassland, riparian

forests and freshwater wetlands. However, extensive land use change for grazing and

particularly sugarcane production has since led to significant losses in riparian and wetland

areas, which prior to clearing, are considered to have provided buffer strips protecting coastal

river systems, estuaries and shorelines (Johnson et al., 1999).

The geology of the region comprises undifferentiated Paleozoic quartzite, Carboniferous acid

volcanics, Carboniferous granites, Cainozoic basalts and Quaternary alluvium and beach sands

(Figure 3-3). A detailed account of the geology of the Ingham district in the lower catchment is

provided in Rienks et al. (2000). Soils are closely related to geology and in the lower catchment

are heavily influenced by fluvial processes. Soils are mostly coarse textured in the upper

catchment and fine textured clay/loam soils in the lower catchment (Johnson and Murray,

1997). Wilson and Baker (1990) defined three major geomorphological units for the Herbert

River floodplain: (1) soils of the alluvial fans derived from granite and acid volcanic rocks; (2)

heavier textured (duplex) soils of low permeability of the alluvial plain characterised by perched

watertables (waterlogging); and (3) sandy, freely drained soils of the riverbank bordering the

Herbert River.

Page 64: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Figure 3-3 1:250,000 geological map comprising the lower Herbert River catchment (BMR, 1965). The town of Ingham is marked for reference (refer to Figure 3-1).

INGHAM■

Page 65: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

River-Aquifer Interactions in the Wet Tropics

41

Based on detailed mapping in the sugarcane areas of the lower catchment, 24 soil types have

been identified, which can be broken down into seven broad categories based on similarity of

parent materials (Wood et al., 2003).

3.3.1 Climate

The catchment is characteristically tropical, experiencing warm humid summers and mild dry

winters. Mean annual rainfall is 1370 mm and ranges from greater that 3000 mm in the north-

east of the lower catchment, to 750 mm in the extreme west of the upper catchment.

Approximately 74% of the mean annual rainfall occurs from December to March (Johnson and

Murray, 1997). The marked seasonal rainfall pattern is illustrated in Figure 3-4 for five rainfall

stations in the lower catchment. The upland areas to the northwest and southwest have lower

average monthly rainfall compared to closer to the coast, particularly during the wet season. The

cumulative residual rainfall curve provides a measure of the accumulated deficit or surplus of

rainfall at a particular time, relative to average rainfall. The curve can be interpreted at different

scales: a positive slope indicates a cumulative period of above average monthly rainfall while a

negative slope indicates the reverse. As illustrated in Figure 3-5, the early 1900’s to 1980’s was

an extended period of above average monthly rainfall, while the following period to 2004

represents a cumulative deficit. Fluctuations in the residual mass curve are also apparent within

the gross wet and dry trends, representing shorter duration rainfall surplus and deficit periods

(Figure 3-6).

%U

%U

%U

%U

%U

32091

32043

32045

32023 32031

10 0 10 KmN

%U Rainfall stations

Streams

Herbert R

0

100

200

300

400

500

1 2 3 4 5 6 7 8 9 10 11 12

Month

Ave

rage

rain

fall

(mm

)

3202332031320433204532091

%U

%U

%U

%U

%U

32091

32043

32045

32023 32031

10 0 10 KmN

%U Rainfall stations

Streams

Herbert R

0

100

200

300

400

500

1 2 3 4 5 6 7 8 9 10 11 12

Month

Ave

rage

rain

fall

(mm

)

3202332031320433204532091

Figure 3-4 Mean monthly rainfall in the lower Herbert River catchment. Source: BoM

Page 66: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 3

42

-6000

-4000

-2000

0

2000

4000

6000

1900

1904

1909

1914

1919

1924

1929

1934

1939

1944

1949

1954

1959

1964

1969

1974

1979

1984

1989

1994

1999

2004

Cum

ulat

ive

resi

dual

rain

fall

(mm

)

Figure 3-5 Cumulative residual rainfall at station 32045 in the lower catchment. The arrows depict periods of above (blue) and below (red) average monthly rainfall based on records since 1900. By definition, zero residual rainfall represents the overall average monthly rainfall over the averaging period.

0

50

100

150

200

250

300

350

400

450

1975

1977

1979

1981

1983

1985

1987

1989

1991

1993

1995

1997

1999

2001

2003

Dai

ly ra

infa

ll (m

m)

0

1000

2000

3000

4000

5000

Res

idua

l rai

nfal

l (m

m)

rainfallresidual

Figure 3-6 Daily rainfall versus the cumulative deviation of residual rainfall from the mean at station 32045. Years are labelled as the approximate start of the wet season (November). Source: BoM (rainfall)

3.3.2 Rivers and aquifers

The entire catchment drains an area of approximately 10,000 km2 to the Coral Sea and is the

largest of the river systems in the sub-humid to humid tropical region of northeast Australia

(Johnson and Murray, 1997). The main drainage systems in the catchment are the Herbert, Wild

and Stone Rivers. Mean annual flow in the Herbert River is 3440 GL, with the highest monthly

flows occurring from November to May. Mean annual runoff for the catchment is 4991 GL (493

Page 67: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

River-Aquifer Interactions in the Wet Tropics

43

mm), with a runoff to rainfall ratio of around 37%. Flooding of the Herbert and Stone Rivers is

confined to the wet season, generally from December to March. Tropical cyclones and

associated heavy rainfall are usually responsible for major flooding (Johnson and Murray,

1997). Further physical characteristics, as well as chemical trends in the Herbert River, are

provided in Chapters 6 and 7.

In the upper and middle catchment, groundwater supplies are available from alluvial, fractured

basalt and other fractured rock aquifers, with the majority of aquifers in fractured material.

Groundwater quality in the upper and mid-sections of the catchment is considered to be good.

Investigations of the sub-surface hydrology have only been carried out in detail for the lower

catchment, which represents an alluvial aquifer system (Cox, 1979). Reinterpretation of the

hydrogeology as part of this study is provided in Chapter 4. The shallowest sandy aquifer is of

greatest economic significance as it is very permeable; however, pumping rates are limited

because of the shallow depth. The aquifer is mainly used to supply town water and for domestic

purposes, with only minimal use for irrigating crops. Although the quality of groundwater in the

lower catchment is considered to be good, high nitrate concentrations and high salinities have

been found in isolated areas. The hydrogeochemistry of the alluvial aquifer system is examined

in detail in Chapter 5. The threat of salt water intrusion on the coastal fringes has traditionally

been the main consideration for groundwater management in the area, as groundwater extraction

in excess of recharge may cause groundwater levels to decline, with a subsequent reversal in the

groundwater flow gradient. Although seawater intrusion into the coastal aquifers can cause

significant deterioration in water quality, formal licensing and allocation systems have not been

adopted in the catchment because the volume of groundwater extracted is significantly less than

the long term yield (Johnson and Murray, 1997).

3.4 RESEARCH APPROACH

As outlined in Chapter 1, specific questions that this research will address include:

(1) What is the nature of river-aquifer interactions in the lower Herbert River catchment,

particularly during the dry season?

(2) What is the significance of river-aquifer interactions for the nitrogen budget of the river?

(3) What are the implications of these interactions for nutrient monitoring, management,

and policies relating to water quality at a catchment scale?

Water is a vehicle for mobilising and transporting nutrients and other dissolved constituents;

therefore, knowledge of the dynamics of water movement, including river-groundwater

connectivity, is central to understanding the transport of such components. In addition, an

Page 68: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 3

44

understanding of groundwater dynamics is an essential element of characterising the

relationships between surface and subsurface waters. In light of these considerations, including

the research aims, the thesis has been divided into four analytical chapters. Chapters 4 and 5

examine the hydrogeology of the alluvial aquifer system; Chapters 6 and 7 specifically

investigate river-groundwater interactions, given the hydrogeological framework. Physical

datasets are derived from existing sources, and provide a background for understanding the

hydrogeology and hydrology in the catchment. Verification and extension of concepts is

accomplished through analysis of an extensive database of hydrochemical information which

was collected for the purposes of this research. Hence, a conceptual understanding of river-

aquifer interactions in the catchment is developed through the course of the thesis. Drawing on

the conceptual framework for river-groundwater interactions, the implications for N transport in

groundwater and potentially to the river are also progressed. Establishing the significance of

groundwater as a vector for dissolved inorganic forms of N to the river is the culmination of the

core analytical components of the thesis. The environmental significance of the key research

outcomes for in-stream and marine ecosystems provides a context for recommendations

regarding nutrient monitoring, management and water policy.

Whilst the broad research approach outlined above is driven by the research questions, specifics

of the methodology are governed by the available datasets and resources to collect additional

data. Attributes of the case study catchment necessarily influence the design of the data

collection program, as discussed in the following section. In reality, the research approach and

methodology are interrelated, whereby decisions about each are the result of an iterative

process.

3.5 DATA COLLECTION

3.5.1 Existing data and applicability

As discussed in Section 3.2.1.1, river water monitoring in the lower catchment has previously

been undertaken by the research agencies of AIMS and CSIRO (Bramley and Muller, 1999;

Furnas et al., 1995). These extensive water sampling programs had a focus on surface water

quality and therefore complementary groundwater data was not collected. Additionally, water

quality samples in these studies were collected upstream of one of the Herbert River tributaries

that drains an intensive sugarcane farming area (Ripple Creek), and were not collected in the

lower tidal reaches of the river. Water quality data is collected by the Queensland Department of

Natural Resources and Water (QDNRW) at least twice a year at each of their monitoring bores

(approximately 70 bore locations, some of which have multiple pipes) and also at two stream

gauges in the lower catchment. Determining the correlation between in-stream water chemistry

and adjacent groundwater sites is difficult with few surface water monitoring sites and when the

Page 69: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

River-Aquifer Interactions in the Wet Tropics

45

timing of water quality sampling does not necessarily coincide for surface water and

groundwater. In addition, QDNRW groundwater samples are not routinely tested for other

dissolved species of N such as nitrite and ammonium, which are potentially important

components of total DIN. The reliability of existing groundwater quality data is also uncertain.

Whilst previous groundwater studies in a subcatchment of the study area (Bohl et al., 2001;

Bohl et al., 2000a; Bohl et al., 2000b) provide valuable background information, all

groundwater samples in these studies were collected from within the upper 10 cm of the aquifer.

Although the concentration of N at this depth is considered to provide an estimate of the

nitrogen which had passed through the profile after leaving the root zone (Bohl et al., 2000b),

similar nitrogen samples were not collected for the deeper aquifer which may also be an

important nitrogen store. In addition, the studies were performed during the wet season only,

were of limited spatial extent, and did not explicitly look at interactions with surface water

resources.

Existing water quality data accessed for this study are summarised in Table 3-1. Although

incomplete for the aims of this study, the available data is particularly useful for observing

temporal variations in water chemistry in the Herbert River. In addition to water quality

information, supplementary data specific to the lower catchment was obtained from numerous

sources (Table 3-1).

Table 3–1 Available datasets for the lower Herbert River catchment

Data type Source

time series river flow

time series stage height

surveyed river profiles

river water chemistry

groundwater chemistry

bore logs

time series groundwater levels

hydrogeological mapping

time series rainfall

GIS data

QDNRW1

QDNRW

HSC2

QDNRW; Bramley and Muller (1999)

QDNRW

QDNRW

QDNRW

Cox (1979)

BoM3

HRIC4; CSIRO5; QDNRW 1 Queensland Department of Natural Resources and Water 2 Hinchinbrook Shire Council 3 Bureau of Meteorology 4 Herbert Resource Information Centre 5 CSIRO Land and Water

Page 70: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 3

46

To address the concerns outlined above, a sampling program was designed to collect consistent

river and groundwater samples in the lower Herbert River catchment (Dixon-Jain et al., 2005).

Key considerations included: types of samples required; scale and spatial distribution of

sampling sites; timing and logistics of sample collection; and technical aspects of sample

preparation and analysis.

3.5.2 Sample types for this study

Based on the aims of the study, water quality samples were collected from surface waters and

groundwater, and of rainfall, for laboratory analysis of major ions, stable isotopes of water and

dissolved N species. In addition, radon samples were collected from surface waters and

groundwaters. Field measurements of water quality parameters and other physical

characteristics were also recorded. Table 3-2 summarises the primary reasons for analysing

samples for environmental tracers and measuring other parameters. Further details on the

principles of environmental tracers and ion, isotope, and nitrogen chemistry, are provided in

Chapters 5 and 7. Field and laboratory data generated during the sampling programs are

summarised in Appendix A.

Table 3–2 Summary of laboratory and field measurements

Measurements Description Purpose

major ions standard cations, anions water sources, evolution, mixing

nutrients TON+, nitrite, ammonium* hotspots, transformations, transport

stable isotopes δ18O, δ 2H water sources, mixing, evaporation

radon 222Rn groundwater discharge

other lab TDS#, alkalinity, pH, EC@ hydrochemical environment, water sources, evolution

field water quality pH, EC, T, Eh^, DO' hydrochemical environment, stable water samples (groundwater), detect hydrochemical change

other field water column depth, river width, flow, depth to groundwater, GPS location, sampling method, date and time

groundwater discharge calculations (radon), water level elevation calculations, mapping

+ Total Oxidised Nitrogen (dissolved) * Nitrate concentration is calculated as the difference TDN - nitrite # Total Dissolved Solids (evaporation to dryness) @ Electrical Conductivity at 25 oC (specific electrical conductance) ^ Redox potential ' Dissolved oxygen

Page 71: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

River-Aquifer Interactions in the Wet Tropics

47

As described below in Section 3.5.4, three sampling trips were undertaken. Major ion and

nutrient data were collected during each field campaign; stable isotopes of water were only

analysed during May 2004, while radon samples were collected in October 2004 and June 2005.

The spatial coverage of river and groundwater samples was greatest during the final sampling

program, largely due to resolution of logistical issues through increased familiarity with the

area. Despite differences in the types and distribution of samples, it is shown in subsequent

chapters that data from each trip has its merits for different aspects of the research.

3.5.3 Site selection

An important consideration for data collection is the scale of the required output, as this affects

the spatial distribution of sampling sites. This thesis is concerned with processes at the

catchment scale; therefore, sampling was undertaken at an appropriate level of detail in order to

detect major hydrochemical changes in the river and groundwater. The density of groundwater

sites (and hence the aquifers sampled) was largely determined by the availability of QDNRW

monitoring bores. Groundwater was sampled from bores as close as possible, and on both sides,

of the Herbert River. Where there were obvious gaps in spatial distribution, samples were also

collected from selected private bores (owned by landholders and the Hinchinbrook Shire

Council). Note that all groundwater samples were from the alluvial aquifer system.

Surface water samples were collected along the entire length of the lower Herbert River, at

intervals determined by access, the location of tributaries, and the distribution of groundwater

sites. Where possible, samples were collected immediately downstream of the entry point of a

tributary: in some locations it was also feasible to sample upstream of tributary inflows. Nash’s

Crossing, within Yamanie National Park, represents the upstream extent of sampling in the

river: this location was chosen as it is situated above the region of sugarcane farming. Hence,

the chemistry of the river at Nash’s Crossing reflects inputs from the upper catchment. At the

downstream extent, samples were collected towards the river mouth in order to examine the

hydrochemistry downstream of the main tributaries and to assess the quality of water potentially

entering the marine environment. Note that samples within the tidal zone were collected further

downstream than in previous studies. In addition to the Herbert River, sampling was undertaken

in selected tributaries, depending on accessibility (refer to Section 3.5.5). A map of river and

groundwater sampling sites is provided in Figure 3-7.

Page 72: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 3

48

U

U

%U%U%U%U

%U%U

%U%U

%U%U%U

%U

%U%U

%U

%U

%U%U

%U

%U

%U

%U

%U%U

%U

%U%U

%U

%U%U

%U%U%U%U%U%U

%U

%U%U

%U%U%U

%U

%U%U

%U

%U%U%U

%U

%U

%U%U%U%U%U

%U%U

%U%U

%U%U

%U

%U%U%U%U%U

%U%U

%U

%U%U

%U%U%U%U

%U

%U

%U

%U

%U

%U%U

%U

%U%U

%U

%U%U

%U%U%U

%U

%U

%U%U

%U%U%U

%U

%U%U

%U%U

%U%U

%U%U%U

%U%U%U%U%U

%U

%U%U

%U%U%U%U

%U%U

%U

%U%U

%U%U

%U%U

%U

%U

%U

%U%U%U%U

%U

%U

%U%U%U

%U

%U

%U

%U

%U

%U

%U%U

%U%U

%U

%U

%U%U%U

#S#S

#S

#S

#S #S

#S

#S#S

#S

#S

#S

#S#S #S

#S

#S#S#S #S #S

#S

#S

#S

#S

#S

#S

#S

#S

#S

#S

#S #S

#S

#S#S

#S

#S

#S

#S #S

#S#S

#S #S#S #S

#S

#S #S#S

#S#S #S

#S #S#S

#S#S

#S

#S

#S

#S

#S

#S

#S

#S

#S

#S#S

#S

#S

#S

#S#S

#S

#S#S

#S

#S

#S

#S

#S#S

#S

#S

#S

#Sr

r $T

Elphinstone Ck

Gow

rie Ck

Stone R

Lannercost Ck

Ripple C kH awk i

ns C

k

Dal

rym

ple

Ck

#

Herbert RGorge

Palm C k

Tre bonne Ck

Seymour RHERBERT RIVER

HinchinbrookIsland116006

116001

INGHAM

Surface water samples#S

Groundwater samples%U

Rainfall samplesr

Gauges%U

N

5 0 5 Km

Figure 3-7 Location of surface water, groundwater, and rainfall collection sites during the three sampling periods. The two QDNRW stream gauges (116001, 116006) and major town of Ingham are also indicated.

3.5.4 Timing of sampling

The tropical climate zone of Australia provides a unique opportunity to study baseflow without

(or with minimal) surface water flows. Therefore, given the specific interest of this thesis on

groundwater contributions to surface waters, water quality sampling was purposely undertaken

during the dry season, when surface runoff is minimal and baseflow dominates streamflow. Due

to the more stable water quality conditions in the dry season, compared with the rapid changes

in water quality parameters during an event, the sampling frequency required for collecting

water quality data during low flow conditions is much less than during event flows (Brodie and

Mitchell, 2005). In the lower catchment, baseflow (low flow) conditions persist for much of the

year; hence, it was considered that grab-samples, during months representing the beginning and

end of the dry season, would be adequate for comparing differences in river-aquifer connectivity

relationships between the extremes of the dry season.

Note that baseflow is comprised of numerous potential sources: groundwater discharge; bank

discharge; unsaturated zone flow (interflow); delayed surface water (e.g. wetlands, lakes); and

delayed groundwater (e.g. perched aquifers) (Evans, 2005). As groundwater discharge is

Page 73: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

River-Aquifer Interactions in the Wet Tropics

49

generally the major process, it is assumed for the purpose of this study that all baseflow, unless

otherwise stated, represents discharge from groundwater sources.

Three sampling trips were undertaken during the dry seasons in 2004-2005 (Figure 3-8). Water

levels in the aquifers of the catchment are at their highest at the cessation of the wet season,

generally by the end of April. Therefore, the start of May 2004 (11/5-23/5) was selected to

conduct an initial sampling program, to capture water quality during conditions of high aquifer

recharge and potentially high physical river-groundwater connectivity. To complement this

sampling, a subsequent collection was undertaken in October-September 2004 (23/10-4/11),

before the start of the 2004-05 wet season. This period was characterised by low groundwater

levels and low river stage. A final field session was undertaken in May-June 2005 (29/5-11/6) to

repeat sampling procedures and hence verify earlier results. Note that for simplicity, the last two

sampling periods are henceforth referred to as October 2004 and June 2005.

0

25

50

75

100

125

150

Jan

04

Feb

04M

ar 0

4

Apr

04

May

04

Jun

04Ju

l 04

Aug

04

Sep

04

Oct

04

Nov

04

Dec

04

Jan

05

Feb

05M

ar 0

5

Apr

05

May

05

Jun

05

Rai

nfal

l (m

m)

0.1

1

10

100

1000

Stre

am d

isch

arge

(GL/

day)

rainfall stream discharge

October2004

May2004

June2005

0

25

50

75

100

125

150

Jan

04

Feb

04M

ar 0

4

Apr

04

May

04

Jun

04Ju

l 04

Aug

04

Sep

04

Oct

04

Nov

04

Dec

04

Jan

05

Feb

05M

ar 0

5

Apr

05

May

05

Jun

05

Rai

nfal

l (m

m)

0.1

1

10

100

1000

Stre

am d

isch

arge

(GL/

day)

rainfall stream discharge

October2004

May2004

June2005

Figure 3-8 The beginning of each sample collection period in 2004-2005, in relation to stream discharge and rainfall in the lower catchment. Source: QDNRW (stream discharge); BoM (rainfall)

Page 74: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 3

50

3.5.5 Logistics, materials and methods

3.5.5.1 Sampling technique

Access to the rivers in the lower catchment was hampered by thick riparian rainforest as well as

hazards associated with wildlife. Consequently, river samples were collected either from

bridges, in the middle of the river by boat, or from the riverbank. Different water collection

devices were used in these different scenarios. For bridge samples, water was collected by

hauling a bucket attached to a rope over the bridge; in the case of low crossings or riverbanks,

an extendable pole was used with an attached sample bottle or by use of a small submersible

pump (Amazon) attached to the pole. Samples were collected in the middle of the river or in the

deepest water, upstream of the bridge/crossing and away from bridge support structures. This

was to ensure that water collected was not influenced by any accumulated debris. Samples from

a boat were collected in the middle of the river using a plastic scoop attached to a short pole or

the submersible pump. Where there was adequate control of the sampling depth, sampling was

from a depth of around 10 cm below the stream surface. The majority of groundwater samples

were collected from QDNRW monitoring bores using a Grundfos MP1 submersible pump.

Bores were purged to remove stagnant water before obtaining a representative sample. At least

three casing volumes of water were removed and pH, temperature and electrical conductivity

(EC) were allowed to stabilise before sampling (MDBC, 1997). Other groundwater samples

were collected from domestic bores equipped with a pump.

Basic field measurements of pH, temperature, EC, redox potential and dissolved oxygen were

determined at all sampling locations using probes connected to a field meter, either directly

from the water source (e.g. in the river or flowing through a hose attached to a pump) or in a

bucket of unfiltered sample immediately after collection. All water collection vessels were pre-

rinsed with the new sample three times to decontaminate between samples. Sample bottles were

also pre-rinsed three times with unfiltered sample, followed by a final rinse with filtrate before

filling. In order to check the accuracy of field sampling procedures and laboratory analyses, a

set of duplicate and spiked1 duplicate samples were collected for major ion and nutrient analyses

at the beginning, middle and end of each field program.

3.5.5.2 Sample preparation and preservation

Samples were either filtered on site or collected in glass bottles and filtered later on the same

day. All filtration equipment was rinsed with deionised water between samples, followed by

rinsing with filtrate. River and groundwater was filtered through a 0.45 μm membrane filter

(nitrate-free for N samples).

1 Containing a known concentration of major ions or nitrate (for N samples)

Page 75: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

River-Aquifer Interactions in the Wet Tropics

51

Samples were preserved in distinct bottles2 for analysis of different chemical parameters,

including major ions, nutrients, stable isotopes and radon. The procedure used for radon

sampling is outlined in Appendix B. Cation samples were acidified to < pH 2 with 1 mL nitric

acid. All samples were refrigerated after collection, with nutrient samples frozen at the end of

each day (and transported frozen to the laboratory).

3.5.5.3 Analytical techniques

Nutrient analyses were performed by flow injection colorimetric methods to determine the

concentrations of dissolved total oxidised nitrogen, nitrite and ammonium3. The concentration

of dissolved nitrate was calculated as the difference between dissolved total oxidised nitrogen

and nitrite. Analyses of major ions were undertaken by ICP-OES (Inductively Coupled

Plasma Optical Emission Spectrometry) for cation and ion chromatography for anion

determinations4. Radon was counted in the laboratory5 by liquid scintillation, on a LKB Wallac

Quantulus counter using the pulse shape analysis program to discriminate alpha and beta decay

(Herczeg et al., 1994). Note that corrections were made for radioactive decay that occurred

between the time of sampling and the time of analysis. Alkalinity, total dissolved solids (TDS),

pH and EC determinations were performed in the laboratory. Alkalinity of the samples was

determined by potentiometric titration; TDS was measured by evaporation to dryness at 180°C.

3.6 CHAPTER SUMMARY

This chapter progressed the theme of groundwater-surface water interactions which was

introduced in Chapters 1 and 2, with a focus on tropical hydrology. A review of research in

tropical systems indicates that there has been far greater emphasis worldwide, including in

Australia, on hydrology in temperate climates. While temperate and tropical regions share many

similarities, there are some fundamental differences related to rainfall patterns and other

climatic factors that affect both stream hydrology and groundwater characteristics. These

differences in turn affect the way in which surface and subsurface waters interact. River systems

in tropical catchments have distinct high and low flow periods that are largely influenced by

seasonal rainfall patterns. Groundwaters are commonly recharged to levels close to the land

surface during the wet season, and thus can contribute to the peak runoff response of rivers.

2 Sample bottled were pre-rinsed with unfiltered sample, followed by a final rinse with filtrate before filling 3 Nutrient analyses were undertaken by ECOWISE Environmental and The Murray Darling Freshwater Research Centre 4 Major ions analyses were performed by The Bureau of Rural Sciences and CSIRO Land and Water Laboratories, South Australia 5 CSIRO Land and Water Laboratories, South Australia

Page 76: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 3

52

Given the shallow depth of the watertable, groundwater is vulnerable to leaching of pollutants

from the land surface. This in turn has the potential to impact on the quality of surface water

resources.

The water quality status of tropical rivers can vary markedly between seasons due to differences

in the residence time of water and flow pathways. While sediment and nutrient transport during

flow events provide an indication of catchment contaminant loads, water quality measurements

during low flow conditions reflect more typical conditions that persist for much of the year.

Therefore, it might be expected that seasonal variations in stream water quality are more

pronounced in tropical catchments. The potential importance of groundwater as a vector for

dissolved species, such as nitrate, to surface waters was also highlighted.

The chosen case study area in the wet/dry tropics of north Queensland was introduced,

including a discussion of water quality issues and a summary of previous research related to N

in the lower Herbert River catchment. Nitrogen in surface and subsurface waters is associated

with fertiliser inputs onto cane land, thus N is a link between land management practices and

water resource quality. Despite concern over nitrate concentrations found in both the Herbert

River and in the shallow aquifer systems, the explicit role of groundwater as a vector for N has

not previously been explored. Therefore, this thesis will characterise river-groundwater

connectivity in the lower catchment and hence elucidate the significance of subsurface N for

surface water quality.

The final part of the chapter outlined the general research approach of the thesis and discussed

data availability and additional requirements. Details of the sampling methodology employed

for this research were also outlined. Key considerations for data collection included: types of

samples required; scale and spatial distribution of sampling sites; timing and logistics of sample

collection; and technical aspects of sample preparation and analysis. This chapter concludes the

introductory chapters of the thesis. The following chapters (4-7) have an emphasis on

examining the dynamics of water and nutrient movement through surface and subsurface

waters, drawing on a range of physical and chemical-based techniques. Whilst data analysed is

specific to the lower Herbert River catchment, the methodology is generic.

Page 77: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

53

Chapter 4 Hydrogeological Framework

4.1 INTRODUCTION

Central to this thesis is the role of groundwater as a possible transport vector for dissolved

forms of nitrogen to surface waters. A first step to understanding this role is the development of

a hydrogeological framework to describe the way in which water, and its dissolved constituents,

are transported through the subsurface (Section 2.4.1.1). This chapter is the first of two chapters

centered on characterising the hydrogeology of the alluvial aquifer system in the case study

area. Whilst this chapter considers the physical aspects, Chapter 5 validates and extends the

physical model through the analysis of hydrogeochemical data. Key questions that the current

chapter addresses include:

• what is the spatial extent of the aquifers?;

• what is the nature of vertical hydraulic connection between the aquifers?;

• how does groundwater move laterally through the lower catchment?

Whilst subsequent chapters (6 and 7) are specifically aimed at characterising river-aquifer

connectivity, the relationship between surface water and groundwater is briefly examined in this

chapter as evidence of potential for interaction. The analyses presented in this chapter provide a

framework to examine the speciation, concentrations, and spatial variability of dissolved

nitrogen in groundwater, and hence to explore the prospect of a subsurface source of nitrogen to

the river.

4.1.1 Key concepts and definitions

Detailed discussions of the fundamental physical principles of hydrogeology are provided in

classic textbooks such as Freeze and Cherry (1979) and Heath (1987). The aim here is to

summarise the main aspects that are of direct relevance to the analyses presented in this chapter.

Darcy’s Law is fundamental to hydrogeology as it describes the factors controlling the

movement of groundwater. According to Darcy’s Law, Q = KA × dh/dl where Q is the rate of

groundwater flow (volume per unit of time), K is the hydraulic conductivity (dependent on the

size and arrangement of pore spaces and on the dynamic characteristics of the fluid), A is the

cross-sectional area (at right angles to the flow direction, through which flow Q occurs), and

dh/dl is the hydraulic gradient (the change in head per unit distance). An aquifer can be defined

Page 78: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 4

54

as a saturated permeable geologic unit that can transmit significant quantities of water under

ordinary hydraulic gradients (Freeze and Cherry, 1979). This is in contrast to a confining layer,

which is a less-permeable geological unit or layer that retards the movement of water in and out

of an adjacent aquifer. Aquifers tend to have relatively high hydraulic conductivities, and are

usually comprised of unconsolidated sands and gravels, permeable sedimentary rocks and

heavily fractured crystalline rocks. In contrast, confining beds possess a very low hydraulic

conductivity and are typically comprised of clays, shales and dense crystalline rocks.

Aquifers can be confined or unconfined: this important distinction affects hydraulic behaviour.

A confined aquifer has a confining bed above and below, and is fully saturated with respect to

water in the pore spaces of the aquifer material. As water in a confined aquifer is pressurised,

the measured water level in a well tapping this type of aquifer will be above the top of the

aquifer, but not necessarily above the land surface. If the latter, the well is considered to be a

free-flowing artesian well. In an unconfined aquifer, the watertable forms the upper boundary of

the saturated zone. As the aquifer is not fully saturated, the saturated thickness can vary over

time, for example, with changes in recharge or groundwater use.

The measured level at which water stands in a well tapping either a confined or unconfined

aquifer is referred to as the hydraulic head, which is the sum of the pressure head, elevation

head, and velocity head. Given that groundwater movement is generally slow, the velocity head

is often disregarded. Moreover, for a watertable aquifer, the pressure head at the watertable is

zero because it is equal to atmospheric pressure, and thus the hydraulic head is equal to the

elevation head at that point. Measurement of depth to groundwater and subtraction from the

elevation of the top of the well casing provides a measure of total head. A groundwater surface

map can be generated by contouring water level elevations (referenced to a common datum such

as sea level): for a confined system this represents the potentiometric surface, while for an

unconfined aquifer this represents the watertable surface. As groundwater flows in the direction

of decreasing head, these surfaces provide an indication of the direction of groundwater flow.

As per Darcy’s Law, the rate of flow depends on the hydraulic gradient and the permeability.

The hydraulic properties of an aquifer can be determined through pumping tests and other

methods such as laboratory measurements of aquifer materials.

Recharge to a groundwater system represents the addition of water to the body of water already

stored in the ground. As the volume of water stored is increased, it is reasonable to expect a rise

in water level in the area receiving recharge. This change in water level will only occur when

the recharge water reaches the watertable, taking a few minutes to hours in a shallow

groundwater system within a permeable unsaturated zone, or months to years if the watertable is

overlain by relatively impermeable material or is very deep (Armstrong and Narayan, 1998).

Therefore, the response of an aquifer to recharge processes provides key information about

Page 79: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeological Framework

55

aquifer properties in the unsaturated zone. In conceptualising groundwater flow, it is important

to be explicit about the type of recharge model being considered and the associated

assumptions. For example, a one dimensional recharge model assumes that recharge operates

largely in the vertical direction, while a three-dimensional model considers both vertical and

lateral movement of water, including overland flow, shallow throughflow and deeper

groundwater redistribution (Hatton, 1998). Although recharge by rainfall is only one of many

possible causes of water level rise (e.g. water-related activities such as recovery from pumping

in a nearby well or the application of irrigation water can also influence water levels), the

following analyses assume that all rises in the watertable are due to natural recharge events.

This assumption is reasonable given that irrigation is not a common practice in the case study

catchment.

Groundwater moves from areas of recharge to discharge sites under the influence of a hydraulic

gradient. Therefore, an understanding of discharge processes, in addition to recharge, is

important in formulating a conceptual model for groundwater flow. This is particularly relevant

when considering the transport, potential residence time, and ultimate fate of dissolved

constituents that are mobilised in groundwater, as is central to this thesis for dissolved species

of nitrogen.

4.2 GEOLOGIC CHARACTERISATION

The following section describes the depositional environment for the unconsolidated alluvial

sediment of the Herbert River valley, and provides a lithostratigraphic interpretation for the

sedimentary sequences identified. This forms the basis for the defined lateral extent of the study

area, together with identifying confined and unconfined layers and their spatial (vertical and

lateral) relationships.

4.2.1 General depositional environment

Of particular interest to this study is the depositional environment of the sediments that

comprise the Herbert River valley, as this affects the composition, characteristics and spatial

extent of the alluvial aquifers. According to Rienks et al. (2000), sediment was deposited during

the Quaternary Period (i.e. less than 2 million years ago) above a predominantly granitic

basement. The largest structural unit in the area is the Ingham Batholith, which comprises

Carboniferous-Permian igneous rocks of the Kennedy Province. The province represents at least

two episodes of granitoid emplacement and possibly three of felsic volcanism (Rienks et al.,

2000). Granitic and volcanic landforms also crop out above the deltaic sediments (Figure 3-3).

Sedimentation occurred from both terrestrial and marine deposition, with five broad

depositional settings recognised: alluvial-plain, estuarine-plain, strand-plain, dune-field and

marine (Holmes et al., 1991). A bedrock contour map clearly delineates a palaeochannel of the

Page 80: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 4

56

Herbert River that was probably formed in response to a relative fall in sea level (Figure 4-1).

Importantly, the map also shows that there is a bedrock-high centered on Trebonne Creek,

which defines the southern margin of the palaeovalley. The trend of the Creek and

palaeochannel of the Herbert River approximately parallel the Palmerville Fault, which is

considered to be a major crustal structure of north Queensland (De Keyser, 1963). Major

movement of the fault is thought to predate the late Carboniferous rocks comprising the Ingham

Batholith. As a result, the fault is difficult to recognise on the ground in the study area (Rienks

et al., 2000).

streams

# bores

? inferred fault

bedrock contours

(59)

53

(40)

(27)

(29)29(17)

(27)(23)

+24+19

+12+19+17+6+13 +2

+81 13

1517

+6+7

+1724

22

(38)

(34)25

(33)51

2930(36)

3438

31 35

54(48)

(30)39

(33)43

(66)65

2937

61 80

77 (81)84

86

90

76

90

92

53

7258 68

72

(71)(79)62

(45)35

5745

41

(1.5)

2210

+10

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

##

##

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

##

##

##

#

#

##

#

##

#

# #

##

## # #

#

#

#

#

#24

FAULT

?

?

?

?

N

Stone R

Herbert R

Trebonne Ck

- 90

- 80

-60

-70

-50

0

-30

-20

-10

-10

0

-2 0

-90-80

-30

-50

Mt Cordelia

-70

5 0 5 Km

-30

-40

Figure 4-1 Bedrock contours in the Herbert River delta referenced to the Australian Height Datum (AHD). Depths to bedrock (negative numbers unless shown with a ‘+’) were converted to AHD by subtraction from the bore elevation, surveyed or estimated from a DEM (enclosed in parentheses). The approximate location of the Palmerville Fault and Mt Cordelia (approximately to scale) are also shown. Source: QDNRW groundwater database (depths to bedrock and surveyed bore elevations); CSIRO (DEM).

4.2.2 Lithostratigraphic interpretation

In order to identify the main lithostratigraphic units of the deltaic sediments, lithologic cross-

sections have been constructed from logs of the bores described by Cox (1979), as well as

selected QDNRW water bores in the study area (Figure 4-2). Bores with complete logs, known

screened intervals and surveyed elevations were selected for the lithostratigraphic interpretation.

The lithologic descriptions have been simplified such that units are distinguished as mud, sand,

sand and gravel, or bedrock (weathered granite). Although sand units vary in their proportion of

Page 81: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeological Framework

57

mud, they are nonetheless grouped together as one type based on the dominance of sand.

Delineation of the main lithostratigraphic units is important for understanding the spatial extent

and vertical relationships between the alluvial aquifers. Selected cross-sections and a fence

diagram are displayed in Figure 4-3 and Figure 4-4: bores have been projected onto the section

lines shown in Figure 4-2. Surveyed riverbed and bank heights (estimated from river profiles

provided by the Hinchinbrook Shire Council) and major outcrops are also included on the cross-

sections. For convenience bores are referred to by the last two or three digits in the QDNRW

numbering system, omitting the 116000 prefix.

$T$T

$T$T

$T$T

$T

$T$T

$T

$T

$T

$T$T

$T

$T$T$T$T

$T

$T$T

$T

%U%U

%U

%U

%U

%U

%U

%U%U

%U

%U

%U

%U

%U

%U

%U

%U

%U

%U%U

%U

%U

%U

%U

%U

%U

%U

%U

%U

%U %U

%U

%U%U

%U

%U

%U

%U

%U

%U

%U%U %U

%U

%U

%U %U

%U

%U

%U

%U%U%U%U

%U%U

%U

%U

%U%U

%U

%U

%U

%U %U

%U

%U

%U

%U

%U

%U

%U

%U%U

%U%U %U %U

%U%U

%U%U %U

%U

%U%U%U

%U

%U

%U%U

%U

%U

%U

%U

%U

%U

%U

%U

%U%U

%U

%U

%U

%U

%U

%U

%U%U

%U

%U

%U

%U

%U

%U

%U

%U

%U%U%U%U

%U%U%U

%U

%U

%U

%U%U%U%U%U%U

%U%U

%U

%U

%U%U%U%U

%U

%U %U%U%U%U%U %U

%U

%U

%U

%U

%U

%U

%U

%U

%U

%U

%U

%U

%U

%U%U

%U

%U

%U

%U

%U

%U

%U

%U

%U

%U

%U

%U

%U

%U

%U

%U

%U

%U

%U

%U%U

%U%U%U

%U

%U%U

%U%U

%U%U

%U

%U

%U

%U

%U

%U

%U

%U

%U

%U

%U

%U

%U

%U%U

%U%U

%U

%U

%U%U

%U

%U%U

%U

%U

%U %U

%U%U

%U

%U %U %U %U %U

%U

%U

%U

%U

%U

%U

%U%U

%U

$T$T$T$T$T$T$T$T$T$T$T$T$T$T

$T$T$T$T$T$T$T$T$T$T$T$T$T

$T$T$T$T$T$T

$T$T$T

$T$T$T

$T$T$T$T$T$T$T$T

$T$T$T

$T$T$T$T$T$T$T

$T$T$T$T$T$T

$T

$T$T

$T

$T

$T$T

$T

$T

#S

#S

#S

#S

#S

#S

#S

#S

#S

#S

#S#S

#S

#S#S

#S

#S

#S

#S

#S

#S

#S

#S

#S

#S#S

#S

#S#S

#S

#S

#S

#S

#S#S#S I

A'

B'

C'

E'

J'

H

D'

36

37

38

39

4041

4243

44

45

47

48

49

50 51

52

53

54

55

56

5758

59

6061

62

6364

65

66

67

68

69

70

46

71

A

DBC

EJ

K

N

%U QDNRW bores#S Cox's bores

River sections$T

5 0 5 Km

Mt Cordelia

Outcrop

Herbert R

Stone R

Trebonne Ck

Palm Ck

boundary ofstudy area

Figure 4-2 Cross-sections constructed from lithological logs of bores in the study area. Outcrops, including Mt Cordelia are shown approximately to scale. Source: bore logs from Cox (1979) and QDNRW groundwater database; river cross-sections from the Hinchinbrook Shire Council.

Interpretation of the lithologic sections is complex because the identified sedimentary units

show considerable interfingering. However, within individual bores, sequences of gravel-sand-

mud can be identified, which are interpreted to represent discrete sedimentary cycles. At least

four sedimentary cycles are recognised, with a unit of mud defining the top of each cycle. In

bores where sedimentary cycles appear to be incomplete, this could either be due to erosion and

reworking or due to a transition in deposition. As depicted in Figure 4-3 and Figure 4-4, lateral

correlation of some units is possible based on the lithologic descriptions and relative depths (in

Page 82: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 4

58

m AHD); however, where there is insufficient evidence to indicate spatial continuity, units are

represented as lenses.

Interpretation of the cross-sections and fence diagram indicates that from the upper part of the

valley to the coast there is a dominant sedimentary unit that overlies the bedrock in the

palaeochannels. This sand and gravel unit varies in thickness from approximately 15 m to 80 m

at the coast. Lenses of sand containing varying degrees of mud are also present within the unit.

Based on lateral comparisons with adjacent bores, the sand and gravel unit is depicted as

continuous along the section H-I-J-K (Figure 4-4). However, radiating outward from this

section, particularly south of Trebonne Creek, spatial correlations are less apparent. This may

reflect a difference in depositional environment on either side of the Trebonne Creek bedrock-

high (Figure 4-1 and Figure 4-2). Sediment deposited along the middle to upper Stone River is

also lithologically different to that of the Herbert River valley (cross-section J-J', Figure 4-4).

Thick mud layers dominate the sediment in these areas and the sand/gravel units are generally

not as thick or laterally continuous. This is consistent with a different source of sediment and/or

a different depositional environment.

The basal sand/gravel layer in the Herbert River valley is generally overlain by a unit of mud of

varying thickness. In the upper parts of the valley the mud contains a thin lens of sand; further

down-valley additional sand/gravel lenses are apparent to the east of bore 67 (Figure 4-4), that

interfinger in the vicinity of bores 52 and 53 (cross-section C-C', Figure 4-3). It is unclear how

laterally continuous these lenses are to the south of the study area (e.g. between bores 54 and 55

in cross-section C-C', Figure 4-3). Towards the coast there is considerable thickening of the

sedimentary strata compared to the upper reaches of the valley; a change in composition to sand

and sand/gravel dominated units with minor intervening muds is also evident. A distinctive

organic-rich mud is found towards the top of the sequence in the coastal bores (cross-section A-

A', Figure 4-3). It has previously been suggested that this represents a marine mud (Cox, 1979).

Figure 4-3 Representative lithologic cross-sections in the lower Herbert River catchment (continued next page). Refer to Figure 4-2 for cross-section locations.

Page 83: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeological Framework

59

cont.

Figure 4-3 Representative lithologic cross-sections in the lower Herbert River catchment (refer to Figure 4-2 for cross-section locations).

Page 84: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Figure 4-4 Fence diagram for the alluvial aquifer system in the lower Herbert River catchment (refer to Figure 4-2 for cross-section locations).

Page 85: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeological Framework

61

Based on the litholostratigraphic analysis presented in this research, the sediments of the

Herbert River valley are predominantly comprised of interbedded mud, sand and sand/gravel

sequences. Whilst sedimentary sequences corresponding to different cycles of deposition can be

identified in individual bores, the lateral discontinuity of these layers and interfingering nature

of the strata make it difficult to define distinct aquifer units. In contrast to this interpretation,

Cox (1979) considered that there were four lithologically distinct sand units, as well as two mud

units. From a hydrogeological perspective, the characterisation of discrete aquifer units is only

important if hydraulically significant at the scale of interest. Hydraulic characteristics and water

levels in nested boreholes are analysed below in order to determine the hydraulic significance of

the observed lithological relationships.

4.2.2.1 Relationships with the river

The locations of river cross-sections available along the Herbert River and other tributaries in

the lower catchment are shown in Figure 4-2. The historical minimum riverbed elevation and

maximum riverbank elevation were estimated from surveyed river profiles and included on the

lithologic cross-sections. Figure 4-3 and Figure 4-4 illustrate the depth at which the Herbert

River and the underlying lithology physically intersect. In the upland areas in the northwest of

the study area, the sedimentary profile has been deeply incised by the Herbert River to a depth

of at least 20 m (e.g. cross-section H-I, Figure 4-4). Hence, the riverbed intersects the basal sand

and gravel unit. Further downstream, the depth of incision gradually declines, such that the base

of the riverbed penetrates one of the upper sand lenses, to a depth of approximately 10 m (e.g.

cross-section J-E, Figure 4-4). These observations have implications for the connectivity of

groundwater with surface water, as discussed in detail in Chapters 6 and 7.

4.2.3 Boundary of the study area

It was noted above that the observed lithology of bores located along the southern extent of the

lower catchment (along the upper Stone River and approximately south of Trebonne Creek) was

different to that observed for bores further north, making lateral comparisons difficult (refer to

the most southern bores in cross-sections B-B', C-C', D-D' and J-J', Figure 4-4). In addition,

lithological logs for some of the southern bores are absent from the database or poorly

described. For these reasons, the southern boundary of the study area is defined by Trebonne

Creek and the middle Stone River valley in the hydraulic analyses that follow. Disregarding the

southern/southwestern bores is considered reasonable given the emphasis in this thesis on

groundwater interactions with the Herbert River. The approximate extent of the study area for

the purposes of this research is depicted in Figure 4-2.

Page 86: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 4

62

4.3 HYDRAULIC PROPERTIES OF THE AQUIFERS

No pumping tests were performed as part of this research. However various tests were

previously carried out by Calvert (1959) and Cox (1979) in the alluvial sediments of the Herbert

River valley. The tests of Calvert varied from the testing of properly screened bores to testing of

open bores cased without screens; recovery readings were used to determine transmissivity. In

contrast, the tests performed by Cox were more comprehensive, comprising measurements of

drawdown and recovery over a period of 24 hours (constant discharge tests) to determine the

aquifer characteristics. Cox defined four alluvial aquifers, referred to as S1, S2, S3 and S4 (from

deepest to shallowest). The results from the pumping tests are summarised in Table 4-1 based

on Cox’s terminology; Cox’s interpretation of the data are summarised below. In order to

distinguish aquifer units based on hydraulic properties, the existing pumping test data are also

assessed in the context of the lithostratigraphic interpretation presented as part of the current

study (Section 4.2.2).

Table 4–1 Summary of hydraulic characteristics of the water-bearing alluvial stratigraphic units in the Herbert River valley, after Cox (1979). A summary of transmissivity values from Calvert (1959) is also provided.

Stratigraphic unit

Aquifer Number of testing sites+

Hydraulic conductivity*

(m/day)

Transmissivity (m2/day)

Cox

Transmissivity (m2/day) Calvert

shallow S4 2 61 - 86 400 - 578 1634-6530

intervening S3 1 2.9 25 170-6114

intervening S2 1 13.9# 117 6-63

basal S1 4 0.5 - 50 50 - 1675 64

+ Cox’s tests * Calculated values by Cox # Results considered by Cox to be of low credibility

4.3.1 Summary of Cox’s interpretation

Based on the pumping tests, Cox concluded that the deepest aquifer (which coincides with the

basal sand and gravel unit described in Section 4.2.2) is mainly confined, but semi-confined in

the upper parts of the valley. The hydraulic characteristics of the aquifer were considered to be

generally uniform, except for in the upper part of the valley where the permeability of the

sediments is higher. The S2 aquifer was interpreted to be confined, although discontinuous and

of limited areal extent. Cox also considered the S3 aquifer to be confined as there was no

evidence of leakage from overlying or underlying units; the transmissivity was observed to be

Page 87: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeological Framework

63

variable. Cox suggested that the S4 aquifer represented an unconfined shallow aquifer of

consistently high to very high transmissivity.

Cox summarised that units S1 and S3 were more permeable on their western margins, with

permeability in each unit decreasing towards the coast due to the increasing mud content to the

east. In terms of vertical variation through the sedimentary sequence, it was noted that with the

exception of the up-valley area, permeability in the S1 unit was low and a little higher in unit

S2. In unit S3 the permeability was substantially higher again, with isolated areas of very high

values. The upper S4 aquifer was considered to be very permeable, although infiltration was

slow due to the veneer of fine grained material covering the more permeable sand.

4.3.2 Interpretation based on lithostratigraphy

In the absence of lithological logs for several bores for which pumping tests were performed, it

is not possible to reinterpret some of the test results. Of the bores with known lithologies, the

results indicate that the hydraulic properties of the aquifer sediments are quite variable, even

within the stratigraphic units defined by Cox. This is most likely a reflection of the complex

stratigraphy in the Herbert River valley typical of fluvial environments. Based on the

lithostratigraphic interpretation presented in Section 4.2.2, the S1 bores of Cox are screened

within the basal sand and gravel layer, while the other bores are screened in the overlying

sand/gravel units (S2-S4). From the available pumping test data it is difficult to delineate

hydraulic characteristics which typify a particular stratigraphic unit; however, in general, bores

screened in the shallower sand/gravel layers have a higher hydraulic conductivity and hence

transmissivity compared to the deeper screened intervals. The exception is in the upper part of

the valley, where the transmissivity in the deep aquifer is high. This is in agreement with Cox

(Section 4.3.1 and Table 4-1). Only two of Cox’s pumping tests were at nested sites; the results

for these tests also highlight the contrast in hydraulic properties between the deepest and

shallowest sedimentary units. However, given the limited number of pumping tests over the

study area and lack of conclusive data, it is considered that further evidence is required in order

to confirm or eliminate the existence of separate confined/unconfined aquifers. Analysis of

water level data (Sections 4.4 and 4.5) and hydrogeochemistry (Chapter 5) will assist in the

characterisation of the alluvial aquifer system in the study area.

4.4 VERTICAL FLOW IN THE SUBSURFACE

Having explored the lithostratigraphic relationships of the sediments in Section 4.2, the aim of

the following sections is to conceptualise how water moves through the subsurface. Water level

data are interpreted in order to describe recharge and discharge processes in the alluvial aquifers

of the Herbert River valley and establish the potential routes along which groundwater flows

between these areas. Bore hydrographs and water level contours are examined in order to

Page 88: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 4

64

characterise the vertical and lateral connectivity between aquifers. The resulting conceptual

framework for recharge, discharge and flow through the aquifer system is validated and

enhanced in Chapter 5 through the interpretation of hydrogeochemical data.

The key characteristics related to vertical flow in the subsurface include: (i) the degree of

vertical aquifer connectivity; (ii) direction of hydraulic gradient between the aquifers; and (iii)

recharge-discharge dynamics. In order to examine these attributes, bore hydrographs are plotted

for nested monitoring bores in the study area based on the historical record of groundwater level

measurements. Bore hydrographs reflect local groundwater behaviour, which will vary

throughout an aquifer in response to local variations in hydraulic properties and distance to

recharge/discharge sites. Importantly, the analyses of hydrographs provide a basis for

identifying discrete aquifers, based on hydraulic behaviour as opposed to lithostratigraphic

units, which may not necessarily correspond to individual aquifers. Nevertheless, water level

behaviour between nested bores has been interpreted within the lithostratigraphic framework,

using the limited spatial extent of monitoring bores. An additional limitation with the bore

hydrograph analysis is that, whilst groundwater is found at various depths in the stratigraphic

profile, the majority of nested bores are screened at only two of those depths in any location:

generally the shallowest and deepest intervals. Therefore, the interpretation is biased towards

the behaviour between the upper and basal sandy units. The analyses in this section build on the

hydraulic characteristics evident from the pumping test data (Section 4.3).

4.4.1 Data preparation

Groundwater level (head) data were extracted from the QDNRW database for all registered

bores (approximately 200) in the study area (Figure 4-2). Of these bores, only around 150 have

historical water level data (of varying length of record), with stratigraphic and screened depth

information at 100 of these bores. Water level records date back to 1967, however some of these

early bores were abandoned and the longest records are generally from 1976 to the present.

Monitoring of water levels averaged four times per year in the 1970’s, but has since declined to

about twice per year: once before the wet season (October-November) and once soon after the

wet season (March-April-May). A limited number of bores have also been monitored daily

using data loggers.

In the database, depth to groundwater and elevation of the bore (in m AHD) are recorded from

the top of the bore casing, known as the reference point. In the absence of an elevation record

for a particular bore, an estimate was made using a 100 m digital elevation model (DEM) grid

(Error! Reference source not found.). Groundwater elevations (in m AHD) were obtained by

subtracting the depth to groundwater from the bore elevation. Screened intervals have mostly

been recorded in the database as depth to the bottom of the slots: in general the slotted interval

extends 3 m above this level (QDNRW technical staff, pers. comm.). Nested monitoring bores

Page 89: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeological Framework

65

comprise multiple pipes (within the same bore) that are screened in different water-bearing

units. The deepest pipe is denoted as the ‘A’ pipe, while ‘B’ and ‘C’ refer to progressively

shallower pipes.

Figure 4-5 Digital Elevation Model (DEM) for the lower Herbert River catchment. Cox’s bores are shown for reference. Source: CSIRO.

Many groundwater monitoring networks similar to that in the study area are read manually and

infrequently (monthly/seasonally), with the objective of measuring the water levels when they

are at their highest and lowest. Although infrequent monitoring does not guarantee to pick the

true maximum and minimum levels, it does provide an idea of the general range of fluctuations

(Armstrong and Narayan, 1998). Illustrated in Figure 4-6 are bore hydrographs for shallow and

deep bores based on groundwater levels measured by an automatic recorder (daily records but

with a significant number of gaps) and manually (from piezometers screened in the same

intervals and located at the same site as each of the recorders) to highlight the difference in bore

hydrographs obtained by the two methods of measurement. Clearly, the high temporal

resolution of the continuous data is useful for analysing dynamic processes such as the rainfall

response of the aquifers, particularly in the wet season. However, these bores are limited in

spatial extent compared to the majority of bores in the catchment which have historically been

sampled manually. It is considered that due to the strong seasonal contrast in groundwater

elevations, this lower temporal resolution data is still adequate for establishing inter-aquifer

relationships in the area.

Page 90: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 4

66

6

8

10

12

14

16

1978

1979

1980

1981

1982

Gro

undw

ater

ele

vatio

n (m

AH

D)

0

100

200

300

Rai

nfal

l (m

m)

61B 80A 79A 61A rainfall 32091

#S

#S

#S#S

78A, 69A

79A, 61A80A, 61B

Herbert R

75A, 46B

4

6

8

1978

1979

1980

1981

1982

Gro

undw

ater

ele

vatio

n (m

AH

D)

0

100

200

300Ra

infa

ll (m

m)

46B 75A rainfall 32045

12

14

16

18

20

22

24

1978

1979

1980

1981

1982

1983

1984

1985

1986

Gro

undw

ater

ele

vatio

n (m

AHD

)

0

100

200

300

Rain

fall

(mm

)

69A 78A rainfall 32091

Figure 4-6 Historical groundwater elevations for bores sampled automatically (labelled) and manually (dots). Groundwater elevations are from the shallowest (80A, 61B, 75A, 46B) and deepest (79A, 61A, 78A, 69A) water-bearing units. Daily rainfall is from sites closest (up-valley) to the monitoring bores. Source: QDNRW (historical groundwater depths) and BoM (rainfall). Refer to inset map for general bore locations in the catchment.

shallow

shallow

deep

deep

automatic

automatic

automatic

Page 91: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeological Framework

67

4.4.2 Inter-aquifer connectivity

Nested bores that have identical hydrographs can be considered to have strong vertical

connectivity. Conversely, where there is head separation between nested intervals, retarded

vertical flow between semi-confined units (in-phase hydrographs) or poor vertical connection

due to intervening confining layers (different hydrograph patterns) is suggested. Note that

similar hydrograph patterns may also occur for reasons other than connectivity; therefore,

reference to bore logs is also important where possible. The vertical head gradient (potential

direction of vertical flow) is deduced by comparing the relative groundwater elevations between

screened intervals. A downwards pressure results if the groundwater elevation for the shallower

screened interval is greater than that of the deeper interval, reflecting a positive head gradient.

Conversely, a negative head gradient indicates an upwards potential. The hydrographs of 12

nested bores screened in the shallowest and deepest water-bearing stratigraphic units are

analysed below based on visual inspection. These nested bores were selected for analysis on the

basis of the length of historical record (at least 10 years) and the presence of bore logs and

screened intervals at known depths. A map of bores referred to in this section is provided in

Figure 4-7, including a spatial representation of the degree and direction of vertical hydraulic

connection between the shallowest and deepest aquifers. Inter-aquifer relationships are further

examined in Chapter 5 through the analysis of hydrochemical data. Hydrographs representing

the main patterns observed in the study area are provided in Figure 4-8 and Figure 4-9.

#S

#S

#S

#S

#S

#S

#S

#S

#

#S#S

#S

#S68

62

6160

53 46 (74)

38

36

48

51 (101)

49

Herbert R47

(54)

#S

#S

Good (downwards potential)#S

Strong (indeterminant)#S

Vertical hydraulic connectivity

Poor (downwards potential)

Poor (upwards potential)

Strong (upwards potential)#S

5 0 5 10 KmN

Figure 4-7 Summary of the degree of vertical hydraulic connectivity (strong, good, poor) and direction of head gradient between the shallowest and deepest aquifers. Note that the locations of bores 54, 101 and 74, screened in an intervening sand unit and the shallowest and/or deepest aquifers, are also shown.

Page 92: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 4

68

8

12

16

20

24

2819

7619

77

1979

1981

1983

1985

1987

1989

1991

1993

1995

1997

1999

2001

2003

Elev

atio

n (m

AHD

)68A (-21.5 m AHD) 68B (13.3 m AHD)

#S#S

#S68

6160

Herbert R

6

10

14

1976

1977

1979

1981

1983

1985

1987

1989

1991

1993

1995

1997

1999

2001

2003

Elev

atio

n (m

AHD

)

61A (-36.1 m AHD) 61B (8.3 m AHD)

11

12

13

14

1976

1977

1979

1981

1983

1985

1987

1989

1991

1993

1995

1997

1999

2001

2003

Elev

atio

n (m

AHD

)

60A (-27.7 m AHD) 60B (-4.2 m AHD)

Figure 4-8 Hydrographs for nested monitoring bores screened in the deepest (A-pipe) and shallowest (B-pipe) water-bearing units in the upper half of the catchment. The elevation of the base of the screened interval in each pipe is also indicated in parentheses. Source: QDNRW (historical groundwater depths). Refer to inset map for general bore locations in the study area.

shallow

deep

deep

shallow

shallow

deep

Page 93: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeological Framework

69

4

6

8

10

12

1976

1977

1979

1981

1983

1985

1987

1989

1991

1993

1995

1997

1999

2001

2003

Elev

atio

n (m

AHD

)

53A (-76.8 m AHD) 53B (3.5 m AHD)

#S

#S

#S53

48

36

Herbert R

0

1

2

3

4

5

6

1976

1977

1979

1981

1983

1985

1987

1989

1991

1993

1995

1997

1999

2001

2003

Elev

atio

n (m

AH

D)

48A (-52.6 m AHD) 48B (-2.2 m AHD)

0.0

0.5

1.0

1.5

2.0

2.5

1976

1977

1979

1981

1983

1985

1987

1989

1991

1993

1995

1997

1999

2001

2003

Elev

atio

n (m

AHD

)

36A (-92.7 m AHD) 36B (-5.9 m AHD)

Figure 4-9 Hydrographs for nested monitoring bores screened in the deepest (A-pipe) and shallowest (B-pipe) water-bearing units in the lower half of the catchment. The elevation of the base of the screened interval in each pipe is also indicated in parentheses. Source: QDNRW (historical groundwater depths). Refer to inset map for general bore locations in the study area.

shallow

deep

shallow

deep

shallow

deep

Page 94: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 4

70

Based on the available data, nested bores in the upper half of the catchment display in-phase

hydrograph patterns (Figure 4-8). The similarity in patterns between nested intervals indicates

good vertical connectivity; however, the head separation at bores 68 and 61 suggests that the

screened intervals are separated by a semi-confining layer and/or have significantly different

hydraulic properties such that vertical flow, although not inhibited, is retarded. The positive

head gradient indicates a downwards potential between the shallow and deep intervals during

both wet and dry periods. Bore logs illustrate the presence of a mud unit of approximately 8 m

in thickness underlying the shallow interval in these bores (Figure 4-4). In addition, the

pumping test results (Section 4.3) indicated that the transmissivity of the shallowest

stratigraphic unit far exceeds that of the deeper unit. Therefore, the observed head separation is

considered to be the result of an increase in storage in the shallow aquifer due to the slight

impediment of downward flow resulting from the lower permeability of the underlying

sediments. The magnitude of head separation declines from the northwest towards the centre of

the catchment, such that at bore 60 the nested hydrographs are near-coincident and indicate

strong vertical connection. These bore hydrograph interpretations are supported by

hydrochemical evidence presented in Chapter 5.

Nested hydrographs in the lower half of the catchment display different trends from bores

screened in the similar shallow and deep sandy units up-valley (Figure 4-9). South of the

Herbert River, in the vicinity of bores 53 and 46 (not shown), the hydrograph of the deeper

interval represents a subdued trend of the hydrograph for the shallower interval. The relative

groundwater elevations indicate a downwards head gradient. The broad similarity in

hydrographic response between the shallow and deep intervals suggests poor vertical

connectivity. Retardation of vertical flow is not unexpected in this area given the presence of a

thick mud unit (e.g. 30 m thick at bore 53) in the middle of the sedimentary profile (e.g. cross-

section C, Figure 4-3). The significance of the observed poor vertical connection is that recharge

to the deep aquifer is likely to be from lateral sources rather than via the shallow aquifer.

The hydrographs of nested bores to the north and northeast are characteristically non-parallel

and sometimes coincide, in contrast to the bores further up-valley which display parallel

hydrographs and a consistent seasonal head separation. Importantly, some bores displaying this

type of nested hydrograph pattern have a dominantly downwards head gradient (e.g. bore 51,

Figure 4-7), while further east the dominant direction of flow is upwards (e.g. bores 48, 49, 36,

38, Figure 4-7). The non-parallel hydrograph patterns between the shallowest and deepest

screened intervals at bore 51 indicate poor downwards vertical connectivity. This is supported

by bore log information, which highlights a 26 m thick mud unit below the shallow aquifer

(cross-section C-C', Figure 4-3). Bore logs illustrate the presence of mud units between 9-12 m

in thickness at bores 48 and 49, and 4-7 m at bores 36 and 38 below the shallowest interval

(cross-sections B-B' and A-A', Figure 4-3).

Page 95: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeological Framework

71

Although these nested bores generally display a head separation, as a consequence of the

intervening muds, there is similarity in hydrographic response between screened intervals

(Figure 4-10). However, whilst the hydraulic information is consistent with good upwards

vertical connection, it is shown in Chapter 5 from hydrochemical evidence, that there is a degree

of confinement between the deep and shallow aquifers at bores 48 and 49 and hence poor

connectivity. Strong upwards vertical connectivity is however demonstrated using

hydrochemistry at bores 36 and 38, indicative of a large component of vertical discharge from

HSd to HSs (Section 5.3.4, Chapter 5).

1

2

3

4

5

6

1976

1977

1979

1981

1983

1985

1987

1989

1991

1993

1995

1997

1999

2001

2003

Ele

vatio

n (m

AH

D)

A-p

ipe

0

1

2

3

4

5

648A (-52.6 m AHD) 48B (-2.2 m AHD)

Ele

vatio

n (m

AH

D)

B-p

ipe

1.5

2.0

1976

1977

1979

1981

1983

1985

1987

1989

1991

1993

1995

1997

1999

2001

2003

Elev

atio

n (m

AHD

)A-

pipe

0.0

0.5

1.0

1.5

2.036A (-92.7 m AHD) 36B (-5.9 m AHD)

Ele

vatio

n (m

AHD

)B-

pipe

Figure 4-10 Similarity in hydrographic response for selected nested bores displaying an upwards potential. Note the different vertical scales for the A and B pipes in each bore.

shallow

deep

shallow

deep

Page 96: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 4

72

As depicted in Figure 4-4, east of cross-section J-J' there are additional sand units between the

shallowest and deepest aquifers. Three nested bores in the study area, bores 54, 46 and 101, are

screened in one of these intervening sands, in addition to the shallowest and/or deepest aquifers

(cross-sections B-B' and C-C', Figure 4-3). Hydrograph analysis of bore 54 (A and B pipes)

indicates that there is a slight impediment to vertical flow between the nested sand units, due to

separation by a 2 m thick mud unit. However, the similarity in hydrograph patterns indicates a

high degree of vertical connectivity between the two upper sand units (Figure 4-11). Bores 46

(with co-located bore 74) and 101 are screened in the two lower sand units, as well as in the

shallowest aquifer. Note that data from bore 101 is analysed rather than co-located bore 51 (bore

101 is the replacement of 51), as 51 is not screened in an intervening sand unit. Analysis of bore

46 highlights the close similarity in hydrographic response between the deepest and intervening

sand units (compare A pipe and bore 74 in Figure 4-11). Whilst these units are separated by at

least 15 m of mud, the head separation is less than 40 cm. Thus, it is considered that vertical

connectivity is high, particularly in comparison to the shallowest water-bearing unit (B-pipe),

which has a different hydrograph pattern (compare A and B pipes in Figure 4-11). Similarly,

analysis of bore 101 highlights the difference in hydrograph pattern between the shallowest unit

(C-pipe) compared to the two lower sand units (A and B pipes). These deeper units display

almost identical hydrographs (maximum of 10 cm head separation); however, it is noted that

there is a consistent upwards pressure between the sand units due to a thin intervening mud

layer (Figure 4-11).

Based on hydrograph analysis of the limited nested bores screened in the intervening sand units,

it is concluded that there is limited retardation of vertical flow between the water-bearing sand

units in the lower stratigraphic layers. This is consistent with their being overlap in the

hydraulic properties of Cox’s S2 and S1 aquifers (Table 4-1). In comparison, the degree of

confinement is larger between the two upper sand units, consistent with the differences in

hydraulic properties between Cox’s S3 and S4 aquifers. However, as noted in Section 4.2.2, the

characterisation of discrete aquifers is only important at the scale of interest. Therefore, in terms

of large scale processes, the greatest hydraulic differences are observed between the shallowest

and deepest water bearing units; the intervening sand units are hence considered as components

of the shallowest (bore 54) or deepest (bores 46 and 101) aquifers. This simplification is also

supported by hydrochemical evidence presented in Chapter 5 (Section 5.3.3).

Page 97: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeological Framework

73

7

8

9

10

11

1976

1977

1979

1981

1983

1985

1987

1989

1991

1993

1995

1997

1999

2001

2003

Elev

atio

n (m

AHD

)

54A (-15 m AHD) 54B (3.7 m AHD)

shallow

intermediate

#S

#S

#S

101 (51)

54

Herbert R

46 (74)

2

4

6

8

1976

1977

1979

1981

1983

1985

1987

1989

1991

1993

1995

1997

1999

2001

2003

Elev

atio

n (m

AHD

)

46A (-89 m AHD) 74 (-48 m AHD) 46B (-0.3 m AHD)

shallow

intermediate

deep

2

3

4

5

6

1993

1995

1997

1999

2001

2003

Ele

vatio

n (m

AHD

)

101A (-64 m AHD) 101B (-36 m AHD) 101C (1 m AHD)

intermediate

shallow

deep

Figure 4-11 Hydrographs for nested monitoring bores screened in an intervening sand unit (green) as well as in the deepest (blue) and/or shallowest (red) water-bearing units. The elevation of the base of the screened interval in each pipe is also indicated in parentheses. Source: QDNRW (historical groundwater depths). Refer to inset map for general bore locations in the study area. Note that bore 101 is the replacement of bore 51, while bores 46 and 74 are co-located.

Page 98: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 4

74

4.4.3 Rainfall response

Visual comparison of available bore hydrographs with historical rainfall indicates that

groundwater elevations display a distinct seasonal pattern of maximum head during the wet

season declining to a minimum during the dry season. Hydrographs for automatically sampled

bores (daily data but with gaps) screened in the shallow interval, represented by bores 80A and

75A, display a rapid response to rainfall (Figure 4-6). Given the shallow depth to water, rapid

recharge response and highly fluctuating water level, the shallow screened unit is considered to

represent an unconfined watertable aquifer. Whilst a similar hydrograph pattern is noted for the

basal aquifer, represented by bores 78A and 79A (Figure 4-6), there is less fluctuation in water

levels compared to the watertable aquifer. This observation is consistent with a longer time for

rainfall to recharge the deeper aquifer. The more subdued hydrograph pattern of the deep

aquifer is particularly evident when comparing the hydrographs of co-located deep bore 79A

with shallow bore 80A (Figure 4-6). Note that this difference in hydrographic response between

aquifers was not apparent when comparing the nested hydrographs of an infrequently sampled

bore at the same location (A and B pipes of bore 61, Figure 4-8). This highlights the importance

of continuous data for establishing recharge dynamics.

Cross-correlation analysis (Croke, in prep.) of groundwater levels in co-located bores 79A and

80A (screened at -36.1 and 8.3 m AHD, respectively) with rainfall clearly shows that there is a

rapid groundwater level response (zero lag) to rainfall in the shallow aquifer, consistent with a

shallow watertable aquifer (Figure 4-12).

-0.2

0

0.2

0.4

0.6

0.8

1

-20 -10 0 10 20

Lag (daily time step)

Corr

elat

ion

coef

ficie

nt

autocorrelation of rainfall

cross-correlation of groundwaterlevel (deep aquifer) with rainfall

cross-correlation of groundwaterlevel (shallow aquifer) with rainfall

rapid groundwater level response at

zero lag

rapid drainage from shallow aquifer

Figure 4-12 Cross-correlation of groundwater levels against rainfall at gauge 32091. The autocorrelation of rainfall provides a reference for the lag in groundwater level response. The deep and shallow aquifers are represented by co-located bores 79A and 80A, respectively (refer to Figure 4-6 for bore hydrographs).

Page 99: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeological Framework

75

However, there is almost no correlation between groundwater levels in the deep aquifer and

rainfall at zero lag. This suggests that unlike the shallow aquifer, the deep aquifer does not

instantaneously respond to rainfall events. Furthermore, the analysis indicates that drainage

from the shallow aquifer after a rainfall event is rapid (around 2 days).

Stage height data (at gauge 116001, approximately 7 km down-valley from the bores), used as a

surrogate for rainfall, gives a better correlation with groundwater levels at non-zero lag. This is

possibly due to the smoothing effect of catchment response in a stream compared to much

higher frequencies and gaps in the rainfall signal. Cross-correlation analysis of stream stage and

groundwater levels highlights a strong wet season-dry season signal (Figure 4-13). In addition

to the instant rainfall response observed in the shallow aquifer (Figure 4-12), a delayed response

is evident in both aquifers relative to streamflow, as shown by the cross-correlation peaks for

each aquifer at approximately 20 and 40 days for the shallow and deep aquifers, respectively

(Figure 4-13). This reflects delayed recharge to the aquifers, such as from lateral flow. Whilst

the effect of a streamflow event persists in both aquifers for longer than in the river, which

drains more rapidly, there is greater persistence in the basal aquifer, as shown by broadening of

the deep aquifer trend after the peak response. This indicates a lag in drainage of the deep

aquifer compared to the shallow aquifer, consistent with differences in aquifer properties.

-0.6

-0.4

-0.2

0

0.2

0.4

0.6

0.8

1

-500 -400 -300 -200 -100 0 100 200 300 400 500

Lag (daily time step)

Cor

rela

tion

coef

ficie

nt

autocorrelation of river stage

cross-correlation of groundwater level(shallow aquifer) with river stage

cross-correlation of groundwater level(deep aquifer) with river stage

peak at ~40 days

peak at ~20 days

aquifer drainage

Figure 4-13 Cross-correlation of groundwater levels against stage height at gauge 116001 (shown as a 14-day moving average). The autocorrelation of stream stage provides a reference for the lag in groundwater level response. Deep and shallow aquifers are represented by bores 79A and 80A, respectively (refer to Figure 4-6 for bore hydrographs).

Page 100: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 4

76

4.4.4 Classification of the aquifers

Based on the interpretation of hydraulic characteristics (pumping tests and bore hydrographs) as

well as lithostratigraphic units, it is concluded that the alluvial subsurface of the lower Herbert

River valley can be conceptualised as a two-aquifer system. The shallow, or upper aquifer,

represents an unconfined watertable aquifer, while the deep, or basal aquifer, is semi-confined.

The alluvial aquifers are bounded at depth by weathered granitic bedrock. Although Cox (1979)

distinguished additional confined units, it is considered that based on the available hydraulic

information, the intervening water-bearing units can be lumped together with the upper and

basal aquifers. Thus, a simplified interpretation is considered reasonable. In the west, the

shallow aquifer extends to a depth of approximately 15 m below the surface, while in the east,

the depth varies between 10 m to at least 35 m below ground level. Although the alluvial aquifer

system is conceptualised to have two aquifers, the degree of vertical hydraulic connection varies

spatially. Given that classification of the aquifers is based on a hydro-stratigraphic

interpretation, the aquifers distinguished as part of this study are henceforth referred to as HSs

(shallow aquifer) and HSd (deep aquifer).

4.5 LATERAL FLOW IN THE SUBSURFACE

Potentiometric surfaces and watertable maps have been constructed in this section to establish

an understanding of lateral groundwater flow through the HSd and HSs aquifer systems. These

analyses build on the recharge-discharge relationships that were developed in the previous

section based on bore hydrograph interpretation. In addition, groundwater level differences

between consecutive dry and wet periods are contoured to illustrate the seasonal response of the

aquifers to rainfall and hence identify potential areas of recharge. Based on the availability of

groundwater level measurements across a number of bores, representative months during 1976-

77 have been selected for the analysis. The wet season in 1977 corresponds to one of the

historically largest flooding events in the catchment (Johnson and Murray, 1997) and therefore

the recharge response is more clearly evident during this period.

Page 101: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeological Framework

77

4.5.1 Deep aquifer system

4.5.1.1 Flow pattern

Potentiometric surfaces show very similar patterns during consecutive wet and dry seasons

(Figure 4-14). Although the potentiometric surface in the deep aquifer system is higher in the

wet season than the dry, the general direction of lateral groundwater flow throughout the year is

from the upper parts of the Herbert River and Stone River valleys towards the coast

(groundwater flow is perpendicular to the contours). The potentiometric surfaces also indicate

that in the dry season there is a preferential flowpath in the vicinity of bore 48 (Figure 4-2) that

has a northeasterly trend. Notably, groundwater flow is towards the Herbert River along a large

section of the river: this has implications for the relationship between groundwater and surface

water systems, as discussed in Chapters 6 and 7. The preferential flowpath depicted in Figure

4-14 (flowline 1) that parallels the western half of the Herbert River, departs from the river in

the middle of the catchment and thereafter follows an eastward trend towards the coast. The

geological control imposed on this flowpath is clearly evident with reference to the bedrock

contour map (Figure 4-1); in particular, the movement of groundwater through the palaeovalley.

In addition, a flow divide is apparent in the south, such that the direction of groundwater flow in

the deep aquifer (e.g. flowline 2) is away from the Herbert River; this may indicate that

groundwater flow is also affected by the bedrock high (Figure 4-1).

4.5.1.2 Recharge-discharge characteristics

Based on the analysis of changes in piezometric level between the wet season and the end of the

previous dry season, there is a major recharge zone for the deep aquifer system in the upper,

northwestern part of the catchment (Figure 4-15a). This zone, in the vicinity of bores 69 and 70

(Figure 4-2), is where the greatest rise in piezometric level is observed in the study area (up to 9

m by March 1977). Less recharge is observed in the southwest, along the upper Stone River

valley (1.8 m by March 1977), which influences water levels in bores along the Stone River and

further to the east. Recharge also occurs in the northeast (1-3 m rise, shown in Figure 4-15a), in

the vicinity of bores 48 and 49 (Figure 4-2). In the months following the peak of the 1977 wet

season, the dominant recharge front shifts further down the Herbert River valley as water levels

decline and ultimately approach their dry season values. The volume of recharge to the deep

aquifer has previously been estimated at 540 megalitres (Cox, 1979).

Page 102: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 4

78

#

#

#

#

#

#

#

#

#

#

#

#

#

##

#

#

#

#

#

#

#

#

#

#

## 28 24 20 16

1.8

2.5

2.2

3.4

5.0

5.4

3.6

3.7

2.0

1.0

3.2

5.9

6.0

5.67.6

8.0

17.3

10.1

12.0

12.5

19.318.0

24.0

10.5

15.426.131.9

261022 1826 14

1632

N

Stream

Groundwater elevationcontour

5 0 5 Km

November 1976

# Groundwater elevation

Outcrop

a

#

#

#

#

#

#

#

#

#

#

#

#

##

#

#

#

#

#

#

#

#

#

#

## 32 28

1.9

2.5

2.6

3.6

5.7

4.0

3.9

3.0

3.9

3.8

6.4

6.3

7.1

9.9

17.8

10.8

12.9

13.020.0

19.0

25.8

11.5

14.9

24.434.5

35.2

261022 1826 14

24

N

Stream

Groundwater elevationcontour

5 0 5 Km

March 1977

# Groundwater elevation

Outcrop

b

Figure 4-14 Potentiometric surfaces (m AHD) for the deep aquifer system during (a) the dry season of 1976 and (b) consecutive wet season of 1977. The direction of groundwater flow is shown by selected flowlines, with flowlines 1 and 2 labelled. Source: QDNRW (groundwater depths)

1

2

1

2

Page 103: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeological Framework

79

The association of the northwest and southwest recharge areas with the topographically elevated

parts of the catchment (Figure 4-5) suggests that lateral recharge to the deep aquifer is sourced

from rainfall towards the catchment boundary (Cardwell Range and Seaview Range, Figure

4-15b) during the wet season. For the recharge zone in the northwest, this is consistent with the

observed correlation between rainfall and groundwater response illustrated in Figure 4-6 at bore

78. In addition, a local recharge area associated with the Mount Leach Range to the north of the

river is evident (Figure 4-15b). Strong vertical connectivity between the aquifers was also

observed in this area (Section 4.4.2), consistent with vertical leakage to the deep aquifer. For the

identified recharge zone in the northeast, the upwards head gradient suggests that recharge to

the deep aquifer is unlikely to be from direct rainfall; however, a combination of recharge from

the southwest and west via lateral flow is plausible (refer to flowlines in Figure 4-14a).

Visual inspection of the potentiometric surfaces for the deep aquifer highlights a decrease in the

hydraulic gradient (increase in spacing between equipotential lines) nearer to the coast (Figure

4-14), indicative of lower hydraulic pressure and hence faster lateral movement of groundwater

due to an increase in hydraulic conductivity. This is consistent with lithological evidence at the

coast, whereby thinning (or absence) of mud units is observed at depth (cross-section A-A',

Figure 4-3). Given the direction of lateral groundwater flow and reversal in vertical head

gradient towards the coast, these observations are consistent with groundwater discharge. Whilst

discharge from the deep aquifer is ultimately out to sea, it was noted in Section 4.4.2 that a

component of vertical leakage from the deep to the shallow aquifer is also plausible towards the

coast. Groundwater pumping is generally from the shallow rather than the deep aquifer and

therefore is unlikely to contribute to losses from the groundwater system. However, given the

deep incision of the Herbert River into HSd in the northwest (Section 4.2.2.1) and the

groundwater flow pattern, this raises the possibility of groundwater discharge from the deep

aquifer to the river. This relationship is explicitly examined in Chapters 6 and 7.

Page 104: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 4

80

#

#

#

#

#

#

#

#

#

#

#

#

##

#

#

#

#

#

#

#

#

#

#

##

0.1

0.1

0.3

0.2

0.5

0.4

0.3

0.2

1.0

3.0

0.7

0.5

0.20.7

1.5

2.3

0.50.7

0.9

3.5

4.4

9.08.53.3

0.89 7

3

1.52 1 0.5

68

21

N

Stream

Head differencecontour

5 0 5 Km

March 1977-November 1976

# Water level difference

Outcrop

a

N10 0 10 20 Km

Catchment boundaryStreamHead difference contoursOutcropCardwell Range

Seaview R

ange

RECHARGEZONE

RECHARGEZONE

LOCALRECHARGE

DISCHARGEZONE

#

Mount LeachRange

b

Figure 4-15 Head difference contours (m AHD) for the deep aquifer system, comparing groundwater levels during the 1977 wet season with previous dry season levels in November 1976. An enlarged area of the catchment is shown in (b): the main recharge areas associated with the mountain ranges and the coastal discharge zone are highlighted.

Page 105: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeological Framework

81

4.5.2 Shallow aquifer system

Watertable maps have been constructed for the shallow aquifer system for selected months

during the dry and wet seasons in 1976-1977 (Figure 4-16). Due the scarcity of monitoring

bores, data from private wells were used in conjunction with monitoring bore data. Manual data

selection from private wells were based on comparisons with groundwater levels recorded in

surrounding monitoring bores. Manual selection was necessary because of the absence of

information on the total depth of some wells: groundwater levels that did not appear consistent

with surrounding bore measurements were discarded. Therefore, in the first instance, contours

were based on water levels in bores with known screened depths, with gaps filled in using

private well data. The addition of the private well data resulted in more detailed flow patterns

downstream of the junction of the Stone River with the Herbert River. Although the watertable

contours are based on a degree of interpretation, the flow patterns observed in Figure 4-16 are

plausible based on the corresponding maps constructed for the deep aquifer system, particularly

where there is good vertical connectivity (compare with Figure 4-14).

4.5.2.1 Flow pattern

Although the watertable is higher in the wet season compared to the dry, the groundwater flow

pattern in the shallow aquifer system is essentially the same in both seasons (Figure 4-16). In

general, it is observed that groundwater flows either from the west towards the Herbert River

and thereafter to the coast, or from the west to the east/southeast towards the coast. As for the

deep aquifer system, a preferential flowpath coincident with the Herbert River is evident;

however, in the centre of the catchment the flowpath is deflected to the northeast towards bore

48 (Figure 4-2).

4.5.2.2 Recharge-discharge characteristics

As for the deep aquifer system, water level difference maps have been constructed for the

shallow aquifer system to compare groundwater levels in the 1976 dry season (November) to

the subsequent wet season in 1977. Comparison of wet season versus dry season groundwater

levels indicates a major recharge area in the northwest: from November 1976 to February 1977

water levels in bore 68B increased by 6.9 m (Figure 4-17a), and by 7.2 m in March (not shown).

In the months following the peak of the wet season, groundwater levels in the main recharge

area declined, while levels in bores further down valley increased in response to lateral

movement of the northwesterly recharge front. Dry season water levels were approached again

by September (not shown).

Page 106: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 4

82

#

#

# #

# #

#

#

#

#

#

#

##

#

#

#

#

#

#

#

#

#

#

#

#

##

#

#

#

#

#

#

#

#

#

#

#

#

24612 1018 8

5.5

3.0 2.1

1.01.0

0.5

1.9

1.4

5.84.4

0.8

0.72.8

3.3

9.2

9.2

8.4

8.7

9.9

7.08.7

7.2

6.7

3.9

3.6

1.0

1.1

3.9

0.6

6.9

16.0 10.1

11.510.2

17.6

15.3

13.012.4

12.8

14.3

1416

N

Stream

Groundwater elevationcontour

5 0 5 Km

November 1976

# Groundwater elevation

Outcrop

a

#

#

# #

# #

#

#

#

#

#

#

##

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

# ##

#

#

24612 1018 81416

7.0

4.4 3.2

2.52.3

1.6

3.0

2.8

6.95.5

3.8

2.44.5

4.9

9.0

8.19.7

8.7

8.9

5.22.5

2.55.3

9.7

17.8 10.8

10.210.5

10.8

13.2

12.1

24.8

12.4

20.114.5

15.7

16.1

16.1

20222426

N

Stream

Groundwater elevationcontour

5 0 5 Km

March 1977

# Groundwater elevation

Outcrop

b

Figure 4-16 Watertable maps (m AHD) for the shallow aquifer system during (a) the dry season of 1976 and (b) consecutive wet season of 1977. Groundwater elevations are from both monitoring bores and private wells. The direction of groundwater flow is shown by selected flowlines. Source: QDNRW (groundwater depths)

Page 107: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeological Framework

83

The watertable difference contours also indicate potential small areas of local recharge to the

east (adjacent to the river), whereby water levels in some bores increased by around 3 m from

November 1976 to February 1977 (Figure 4-17a). This is in contrast to bores recharged from the

northwest that attained maximum increases in water levels from March onwards, suggesting that

different recharge mechanisms (with different timings) operate in the two areas. The cross-

correlation analysis presented in Section 4.4.3 also demonstrated that there is a rapid and a

delayed groundwater level response to rainfall events in the shallow aquifer.

It was established in Section 4.4.3 that the shallow aquifer is a watertable aquifer that is

recharged directly by rainfall. However, the analysis of watertable difference contours indicates

that lateral recharge from the northwest is also an important source of recharge, at least in some

parts of the aquifer. Therefore, there are at least two major recharge sources to HSs, with the

importance of each source varying spatially. Similar to the deep aquifer, there is also evidence

of a local recharge area in the north (Figure 4-17b). In addition, it has previously been suggested

by Herbert (1994) that the northeasterly trending flowline (near bore 48, Figure 4-16) is a

preferential flowpath for the inflow of seawater to the shallow aquifer. As discussed in Section

4.4.2, the upwards vertical head gradient towards the coast also indicates the potential for

leakage of deep groundwater to the shallow aquifer.

The groundwater flow pattern, in combination with the incision depth of the river bed into the

shallow aquifer (Section 4.2.2.1), indicates the potential for exchange of water between HSs and

the Herbert River. For instance, the shape of the contour lines, pointing upstream (concave)

where they cross the western half of the river, is indicative of a gaining stream due to the

discharge of shallow groundwater (Winter et al., 1998). In contrast, in the eastern half of the

river, watertable elevation contours point slightly downstream (convex) in the vicinity of the

stream, such as where the 4-10 m watertable contours cross the Herbert River in the wet season

(Figure 4-16b). This may indicate a losing river reach at that time of year and hence the

possibility of the river recharging the shallow aquifer. Further analyses are provided in Chapters

6 and 7 to examine river-aquifer interactions. Given the inter-aquifer connectivity relationships

established in Section 4.4.2 (Figure 4-7), vertical leakage from HSs to HSd is likely in the west.

Other plausible groundwater losses from the shallow aquifer system could result from

evaporation, discharge to sea, and pumping for town water supplies.

Page 108: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 4

84

#

#

# #

# #

#

#

#

#

#

#

##

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

#

2.0 1.7

1.0 0.91.41.3

0.7

1.0

1.2

1.4

0.8

3.1

1.41.7

1.5

1.6

1.0

0.5

1.30.2

1.70.9

1.9

6.9

1.9

1.10.9

2.0

1.3

5.2

1.8

1.8

1.5 1.4

1.5

1.5

2.33.4

1.3

246

23

1

1

N

Stream

Head differencecontour

5 0 5 Km

February 1977-November 1976

# Water level difference

Outcrop

a

N10 0 10 20 Km

Catchment boundaryStreamHead difference contoursOutcropCardwell Range

Seaview R

ange

RECHARGEZONE

RECHARGEZONE

LOCALRECHARGE

DISCHARGEZONE

#

Mount LeachRange

DIFFUSERECHARGE

b

Figure 4-17 Head difference contours (m AHD) for the shallow aquifer system comparing groundwater levels during the 1977 wet season with previous dry season levels in November 1976. An enlarged area of the catchment is shown in (b), depicting recharge areas associated with the mountain ranges, diffuse rainfall sources, and the coastal discharge zone.

Page 109: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeological Framework

85

4.5.3 Relationships between aquifers

Superposition of groundwater elevation contours for HSs and HSd during the same time periods

(Figure 4-18) illustrates that that there is a dominant preferential pathway for groundwater flow

from the northwest and parallel to the Herbert River in both aquifer systems. As noted above,

this flowpath follows the major palaeovalley delineated by the bedrock (Figure 4-1). The

concave shape of the groundwater elevation contours, where they cross the western half of the

river, suggests that the river acts as a groundwater sink for both aquifers. However, in the east,

where the main flowpath diverges from the river, a subtle difference between the aquifers is

observed: groundwater elevation contours for HSs are slightly convex in the vicinity of the river,

consistent with discharge from the river to the shallow aquifer, while those of HSd are more

concave. Further discussion of river-aquifer relationships is deferred to Chapters 6 and 7. In

general, the flowpaths in each aquifer follow a similar trajectory from recharge to discharge

areas, consistent with good vertical connectivity and hence a semi-confined alluvial aquifer

system. However, minor deviations between the HSd and HSs flowlines in the east (south of the

river) are probably due to the presence of localised thick mud sequences which restrict vertical

hydraulic connection, as shown by the area of poor vertical connectivity in Figure 4-7.

In addition to illustrating the direction of lateral groundwater movement in each aquifer, Figure

4-18 depicts the areas (shaded) where the direction of vertical flow is from the shallow aquifer

to the deep aquifer in representative wet and dry seasons. To derive these areas, groundwater

elevation contours between each aquifer were compared at the same period in time, to delineate

where the elevation of HSs was equal to HSd and hence where the elevation of HSs exceeded

HSd. This approach is similar to that used to compare bore hydrographs at nested sites to

determine the potential direction of flow between aquifers (Section 4.4). Note that not all areas

outside of the shaded area represent an upward head gradient between aquifer units, for

example, the northwestern extent of the shaded area is constrained by the available water level

data for the shallow aquifer (refer to Figure 4-16). Comparison of dry and wet season maps

clearly shows that the area of downwards hydraulic head expands to the east during the wet

season. Therefore, there is a seasonal switch in the vertical head gradient in bores towards the

coast and a resultant shift in the extent of potential discharge from HSd to HSs.

Page 110: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 4

86

24612 1018 81416

2

610

1620242832 30 26

N

Stream

HSd groundwaterelevation contour

Outcrop

5 0 5 Km

November 1976

#

#

# #

#

##

#

elevation HSs = HSd#

#

#

#

#

# # # #

#

#

#

elevation HSs > HSd

HSs groundwaterelevation contour

HSs flowlineHSd flowline

24612 1018 81416202224262610

24263034

N

Stream

HSd groundwaterelevation contour

Outcrop

HSs groundwaterelevation contour

5 0 5 Km

March 1977

# elevation HSs = HSdelevation HSs > HSd

#

#

#

#

#

#

#

##

#

#

## # # # #

##

# #

#

HSs flowlineHSd flowline

Figure 4-18 Contours of groundwater elevation superimposed for the shallow (HSs) and deep aquifers (HSd) during (a) the dry season and (b) the wet season. The approximate area where the groundwater elevation is greater in the shallow aquifer compared to the deep aquifer at the same period in time (i.e. the region of downwards hydraulic head) is shaded.

Page 111: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeological Framework

87

4.6 CHAPTER SUMMARY

Lithological bore logs and hydraulic data were analysed in this chapter in order to conceptualise

the hydrogeology of the lower Herbert River catchment. Two alluvial aquifers are distinguished:

HSs, a shallow unconfined aquifer, and HSd, a semi-confined deep aquifer. Although the

aquifers are distinct, the extent of vertical hydraulic connection varies spatially. On balance, it is

considered that there is generally good vertical connectivity between HSs and HSd, with the

exception of an area in the middle of the catchment, where there is poor connection.

Groundwater in each aquifer follows approximately the same flowpaths: from the upland

recharge areas in the northwest and southwest towards (and parallel to) the Herbert River, and

then eastward to the coastal discharge area. A component of vertical discharge from the deep

aquifer to the shallow is also evident towards the coast, where the head gradient reverses from a

downwards to an upwards potential. In addition to lateral recharge at the catchment boundary, a

local recharge zone is evident in the middle of the catchment, to the north of the river. Direct

rainfall is also an important component of recharge to the shallow aquifer, with vertical leakage

to the deep aquifer likely in some areas. Importantly, the analyses in this chapter highlight the

relationship between the aquifers and the Herbert River. Whilst the direction of groundwater

flow in the deep aquifer is towards the river along much of its length, the depth of incision of

the river into HSd decreases down the valley. Based on the available information, potential

groundwater discharge from the deep aquifer to the river is restricted to the upper reaches of the

valley, approximately west of the junction with the Stone River. In the east, where incision of

the river bed is only shallow, the movement of deep groundwater is through the sediments of the

palaeochannel of the Herbert River. In contrast, there is the potential for direct exchange of

water between HSs and the Herbert River along its entire length. Available evidence suggests

that the river acts as a sink for shallow groundwater in the west and perhaps a source in the east.

Given the vertical hydraulic connection between HSd and HSs over a large part of the study area,

dissolved constituents that penetrate the groundwater system may be found in the two aquifers.

Therefore, in reaches where one or both of the aquifers interact with the Herbert River there is

the potential for groundwater to influence in-stream water quality and vice versa. In addition,

groundwater discharged to the river or directly to the sea has the potential to impact on water

quality in the marine environment. The following chapter examines hydrogeochemical data to

verify and develop the hydrogeological framework, including analysis of the distribution of

nitrogen in the subsurface. Subsequent chapters evaluate the physical and chemical relationships

between groundwater and surface water and assess the potential for a groundwater source of

nitrogen to the Herbert River.

Page 112: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to
Page 113: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

89

Chapter 5 Hydrogeochemical Framework

5.1 INTRODUCTION

A conceptual framework for the movement of water in the alluvial aquifer system was

developed in Chapter 4 from existing lithologic and hydraulic information. Based on available

evidence, it was concluded that the subsurface can be conceptualised as a two-aquifer system

comprising a shallow unconfined aquifer (HSs) and semi-confined deep aquifer (HSd). The

nature of hydraulic connection (i.e. degree and direction) between the aquifers varies spatially.

Groundwater in each aquifer follows approximately the same flowpath, including flow towards

a section of the lower Herbert River enroute to the coastal discharge zone. Hydrogeochemical

data are examined in this chapter to verify and further develop these concepts. In accordance

with the interpretation based on hydraulics, hydrochemical analyses in the following sections

consider groundwater samples as representative of one of the two aquifer systems. Field data

collected for this study (Table 3-2) as well as existing sources are analysed (Table 3-1). Details

of the sampling program were outlined in Chapter 3 (Section 3.5).

There are numerous techniques used to analyse hydrochemical data, which take the form of

qualitative (graphical) and /or quantitative (numerical) procedures (Herczeg and Edmunds,

2000; Mazor, 1991; Hem, 1985; Freeze and Cherry, 1979; Zaporozec, 1972). The chosen

techniques are a function of the research questions and the nature of the available data (Section

2.4.1.4). This chapter specifically considers groundwater chemistry, while surface water

chemistry is introduced in Chapter 7 in the context of river-aquifer interactions. Given the

overarching interest in this thesis in the mobility of dissolved N, hydrochemical data relating to

species of N are examined in groundwater in this chapter and in surface waters in Chapter 7. A

conceptual diagram summarising water and N movement within the alluvial aquifer system is

presented at the end of this chapter (Figure 5-33). Whilst this chapter does not aim to provide a

comprehensive hydrogeochemical analysis, hydrochemical data is specifically analysed to:

• examine hydrochemical signatures within the lower catchment aquifers;

• establish the degree of vertical mixing between aquifers;

• examine lateral hydrochemical relationships within each aquifer; and

• explore the distribution and speciation of N within the hydrogeologic framework.

Page 114: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

90

Therefore, hydrochemistry is used as a tool to inform physical processes. Furthermore, by

characterising the chemistry of the groundwater, this provides a hydrochemical signature for the

groundwater component in surface waters, as discussed in Chapter 7.

5.1.1 General principles

5.1.1.1 Environmental tracers

Environmental tracers constitute natural or anthropogenic isotopes, elements or compounds that

are widely-distributed in the near-surface environment of the earth, such that variations in their

abundance can be used to infer environmental processes (Cook et al., 2002). Useful tracers

include (1) common dissolved constituents such as major cations and anions; (2) stable isotopes

of oxygen and hydrogen; (3) radioactive isotopes such as tritium; and (4) physical properties

such as water temperature (Winter et al., 1998). As most of these constituents are found in a

dissolved state in water, they have the potential to track the evolution of water as it moves

through the landscape. Tracers fall into the categories of conservative (inert) and non-

conservative (reactive) types, each type having their relative merits for particular applications.

Within each type there are: timescale, fingerprinting, reaction, pollution, and artificial tracers

(Cook, 2005). Central to this research is the application of pollution tracers, particularly

inorganic species of nitrogen, as a way of linking land-based agricultural activities to water

quality outcomes in both surface water and groundwater resources. Also of interest are

fingerprinting tracers which rely on the fact that where different water sources have different

chemical compositions, measurement of the chemistry of a water sample can provide

information on its origin.

A particular tracer can be conservative or non-conservative depending on the chemical

environment; for example, nitrate may be relatively inert under aerobic conditions and reactive

under reducing conditions. Thus, in contrast to Cl, which is generally considered to be

conservative, the concentration and speciation of nitrogen along a flowpath can vary

dramatically due to chemical interactions (Section 5.1.1.4). Therefore, it is important to

characterise the hydrochemical environment in order to inform the choice of environmental

tracer for the research problem, and to assist with interpretation of the results.

5.1.1.2 Ion chemistry

The solutes present in groundwater are primarily derived from rainfall and weathering,

including water-rock interaction. Water that infiltrates through the soil and unsaturated zones is

altered by evapotranspiration and the production of carbon dioxide from decay of organic matter

and plant respiration. The decay of organic matter also consumes dissolved oxygen and

produces water that is low or deficient in oxygen. The formation of carbon dioxide leads to the

production of H2CO3 which promotes mineral-water reactions. Other organic acids produced in

Page 115: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

91

the soil zone also aid mineral dissolution and hence affect the chemistry of water prior to its

passage to the saturated zone. Water that infiltrates into the saturated zone evolves

geochemically towards areas of discharge; chemical modification reflects time-dependent

processes such as water-rock/mineral interactions, dissolution-precipitation, and mixing.

Therefore, analysis of the hydrochemistry of groundwater along its flowpaths provides insight

into chemical processes that have influenced its lateral hydrochemical development.

In general, the bicarbonate ion is the dominant anion in recharge areas, derived from soil zone

CO2 and mineral weathering, especially dissolution of calcite and dolomite. Therefore, the

presence of bicarbonate-type waters is often indicative of proximity to a rainfall-recharge zone.

In addition, as groundwater moves along flowpaths there is generally an increase in total

dissolved solids (TDS) and the concentration of major ions. Hence, shallow groundwater is

typically lower in dissolved solids in recharge areas than in discharge areas (Freeze and Cherry,

1979). According to Chebotarev (1955), groundwater changes chemically towards the

composition of seawater, with anion evolution from bicarbonate (HCO3- ) to sulfate (SO4

2-) and

ultimately to chloride (Cl-) with increasing age and distance along the flowpath. Back (1960)

also related the chemical composition of groundwater to its flowpath; specifically, Ca-Mg-

HCO3 facies are predominantly found in recharge areas, whilst Na-Cl-SO4 facies are more

commonly found in discharge areas. Hydrochemical changes along the flowpath are attributed

to the processes of ion exchange and sulfate-reducing bacteria. An additional hydrochemical

evolution sequence is that of electrochemical evolution, concerning the tendency for the redox

potential of groundwater to decrease as water moves along its flowpath (Germanov et al., 1958).

Of particular relevance to this study is that in a closed system, the oxidation of organic matter

accompanied by consumption of O2 is followed by reduction of NO3-. Thus, in some

groundwater systems NO3- occurs at shallow depth and diminishes in concentration as the water

moves deeper into the flow system (Freeze and Cherry, 1979). It is important to note that the

hydrochemical evolution sequences described above are broad generalisations and each system

must be examined for peculiarities and local variability.

The notion of conservative and non-conservative tracers was discussed above (Section 5.1.1.1).

Chloride is generally assumed to behave conservatively in groundwater due to its high solubility

and inability to precipitate in minerals except at very high concentrations. This means that the

mobility of Cl is very similar to water molecules, noting that an ‘enriched’ Cl signature relative

to rainfall is carried to groundwater after leaving the zone of evapotranspiration. Relationships

between major ions (including non-conservative types) and chloride (as an inert reference ion)

are commonly used in hydrogeochemical investigations to differentiate compositional groups,

explore mixing relationships and calculate groundwater recharge (Cresswell and Herczeg, 2004;

Herczeg and Edmunds, 2000; Mazor, 1991; Hem, 1985; Zaporozec, 1972). Caution, however,

must be applied for groundwaters known to be influenced by seawater intrusion or in areas

Page 116: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

92

where Cl-bearing evaporite minerals are present, as in such situations Cl is no longer

conservative. Variations in the concentrations and ratios of non-conservative ions provide

information on geochemical reactions along a flowpath, and hence indicate maturity and relative

residence time of groundwater down-gradient.

5.1.1.3 Isotope chemistry

In addition to major and minor ion chemistry, isotopes (stable and radioactive) provide

information on hydrogeological characteristics of aquifers including origins, age and rate of

recharge, and aquifer interactions. Given the large number of isotopes and variety of

applications, the following discussion provides a brief introduction to stable isotopes of water

and selected radioactive isotopes. The application of 222Rn, a radioactive isotope useful for

examining river-groundwater interactions, is described in Chapter 7. Comprehensive reviews of

the principles and applications of isotopes in hydrogeology are provided in Clark and Fritz

(1997) and Gonfiantini et al. (1998).

Isotopes are atoms of the same element but with different masses, due to different numbers of

neutrons in the nucleus. Isotopes may be either stable (i.e. they do not change over geological

time under ambient conditions), such as 18O and 16O, or may be radioactive (i.e. are unstable and

decay by emitting a high-energy fragment, or fragments, until a stable isotope is created), such

as 14C. If the rate at which radio-isotopes decay is known, they can be used to measure process

rates and date events. Stable isotopes are generally measured as the ratio of the two most

abundant isotopes. The isotopes of the water molecule are particularly useful for hydrological

studies as they trace the actual movement of water (Cook and Herczeg, 1998). For a water

molecule, the oxygen isotopic ratio is represented by 18O/16O, while that for hydrogen is 2H/1H.

Measuring the ratio of heavy to light isotopes relative to a standard reference material is easier

and more accurate than determining absolute concentrations. Stable isotope ratios are generally

expressed using the delta (δ) notation:

δ (o/oo) ×−

=std

stdsample

RRR

1000

where R is the isotope ratio and δ is expressed as parts per thousand (o/oo). For δ18O and δ2H, R = 18O/16O or 2H/1H and the standard (std) is Standard Mean Ocean Water (SMOW). Thus,

seawater is assigned the δ values for oxygen and hydrogen equal to 0 o/oo (although local

variations are possible). Given this definition of δ, water samples that are relatively depleted in

the heavy stable isotopes are represented by negative values.

The different masses of isotopes result in different rates of mobility and reactions during

physical and chemical processes such as evaporation or water-rock exchange, a process known

Page 117: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

93

as fractionation. Fractionation processes affect the relative abundance of the isotopes of a given

element, which is central to their utility in hydrochemical investigations.

These processes and their implications for the stable isotope composition of rainfall can be

summarised as (Gat, 1980):

• the elevation effect: depletion of heavy isotopes in higher altitude rainfall;

• the amount effect: depletion of heavy isotopes with increasing quantity of rainfall;

• the continental effect: depletion of heavy isotopes with increasing distance from the coast; and

• the temperature effect: depletion of heavy isotopes at cooler temperatures.

In general, stable isotopes of oxygen and deuterium of most fresh waters around the world

conform to the relationship approximated by δ2H = 8δ18O + 10 (Craig, 1961), known as the

Global Meteoric Water Line (GMWL). While the GMWL is a useful reference for comparing

stable isotopic data, δ18O and δ2H for a given region may reflect different meteoric conditions

and be described by a local meteoric water line (LMWL) of a slightly different slope and

intercept (Ingraham, 1998). Hence, the LMWL may be the more relevant reference line for

detailed stable isotopic studies. Isotope fractionation occurs during evaporation of water such

that the residual water phase becomes relatively enriched in heavy isotopes (more positive δ18O

and δ2H). Thus, an evaporation line is an additional trend that may be observed in a collection of

water samples, depicted in a δ2H- δ18O plot as a series of points to the right of the local water

line with a slope determined by temperature and humidity (Mazor, 1991).

Radioactive isotopes play an important role for establishing the relative age of groundwater

(Gonfiantini et al., 1998). The naturally-occurring radio-isotope of tritium, 3H, is another

isotope of water that is useful for hydrochemical investigations. Large increases in the

concentration of 3H in rainfall occurred in the 1950’s and 1960’s due to thermonuclear weapons

testing. Prior to testing, atmospheric levels were 5-10 tritium units (TU), where one TU

represents one atom of 3H in 1018 atoms of hydrogen. In the southern hemisphere mean annual

tritium concentrations exceeded 50 TU in the early 1960’s before declining to natural levels

(Cook and Herczeg, 1998). The half-life of tritium is 12.43 years and thus can only be used to

study systems in which the residence or transit time of groundwater is a few to 100 years

(Gonfiantini et al., 1998). The most useful environmental radio-isotopes for determining

groundwater age in confined aquifers are those with a half-life ranging from 103 to 106 years.

For example, carbon-14 has a half-life of 5730 years, while chlorine-36 has a much longer half-

life of 301 000 ± 2000 years. Further discussion of the theory and application of these isotopes

can be found in Cook and Herczeg (1998) and Gonfiantini et al. (1998).

Page 118: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

94

5.1.1.4 Nitrogen chemistry

Nitrogen can exist in many forms. In order of decreasing oxidation state, the dissolved species

of interest to the soil/water environment include: nitrate (NO3-), nitrite (NO2

-), nitrogen gas (N2),

ammonium (NH4+), and ammonia (NH3). Dissolved organic N (DON) may also be present in

freshwaters. Transformation of N compounds can occur through several mechanisms: fixation,

ammonification, synthesis, nitrification, and denitrification (Canter, 1997). Transformations that

occur between the different forms of N in the biosphere are summarised by the nitrogen cycle

(Figure 5-1).

Leaching to groundwater

(N2)

Nitrification

Mineralisation

Nitrification

LightningFixation

BiologicalFixation

IndustrialFixation

Denitrification

PlantConsumption

Adsorption

Leaching to surface water

INTERACTION

Leaching to groundwater

(N2)

Nitrification

Mineralisation

Nitrification

LightningFixation

BiologicalFixation

IndustrialFixation

Denitrification

PlantConsumption

AdsorptionDenitrification

PlantConsumption

Adsorption

Leaching to surface water

INTERACTION

Figure 5-1 The nitrogen cycle, modified after Pidwirny (2005). The interaction between groundwater and surface water and the potential transport of N between them is also indicated.

Given an inorganic source of N, such as fertiliser, the key transformation reactions of direct

relevance to this study include:

• Nitrification: biological oxidation of ammonium ions to nitrite and then nitrate,

involving specific chemoautotrophic bacteria and an inorganic carbon source

i.e. NH4+ + O2 → NO2

- + O2 → NO3- (5-1)

• Denitrification: biological reduction of nitrate to nitrogen gas, involving heterotrophic

bacteria and an organic carbon source

i.e. NO3- + organic carbon → NO2

- + organic carbon → N2(g) + CO2 + H2O (5-2)

Page 119: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

95

Whilst nitrification leads to the formation of nitrate that is readily leached to soils and

groundwater, denitrification produces nitrogen gases that can escape from the aqueous

environment. Changes in environmental conditions that can affect N transformations include

redox potential, temperature, pH, microbial population, and the concentrations of electron

donors (e.g. dissolved organic carbon) and electron acceptors (e.g. oxygen, nitrate)

(Thayalakumaran et al., 2004; Canter, 1997).

5.1.2 Methods of interpretation

Groundwater investigations can produce large volumes of chemical data. Therefore, effective

methods for organising, classifying and presenting the data are important for assisting with

analysis and hydrochemical interpretation. Methods utilised specifically in this chapter are

summarised below. In this chapter, ionic concentrations are generally expressed in the units of

milliequivalents per litre (meq/L), while concentrations of other chemical species are expressed

in milligrams per litre (mg/L).

Trilinear plots, such as the Piper diagram (Piper, 1944), provide a useful tool for summarising a

large number of chemical analyses in a single plot. The Piper diagram (e.g. Figure 5-2)

combines two trilinear plots, each representing the relative proportions (in meq/L) of the major

cations and anions. In addition to comparing relationships between cations and anions, Piper

diagrams can also highlight distinct chemical groups, trends along flowpaths and mixing

between waters. Fingerprint diagrams, such as the Schoeller plot (Schoeller, 1955), also have

the ability to represent numerous major ion analyses on one diagram (e.g. Figure 5-5); unlike

trilinear diagrams they have the advantage of representing actual ion concentrations rather than

normalised values. This feature is particularly useful when exploring chemical similarities and

differences between aquifers and hence ascertaining the degree of inter-aquifer groundwater

interactions (Section 5.3).

Bivariate plots allow interpretation of the relationships between two variables. In particular,

expressing the concentrations of ions as ratios of one ion to another, or of one constituent to the

total concentration, can be helpful for highlighting similarities and differences between waters

(Hem, 1985). Ion/chloride bivariate plots (e.g. Figure 5-10) are commonly used to show the

relationships between major or minor ions relative to the conservative Cl ion. Compositional

clusters and mixing relationships can also be ascertained from these types of graphical analyses.

Stable isotope plots comparing δ18O and δ2H (e.g. Figure 5-11) are another example of a

bivariate plot.

The graphical techniques discussed above lack the ability to easily represent the geographical

distribution of the hydrochemistry. For this reason, a combination of methods is desirable in any

hydrogeochemical investigation. To explore the spatial hydrochemical evolution of groundwater

Page 120: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

96

along flowpaths, water quality maps (e.g. Figure 5-23) can be constructed to illustrate the

distribution of particular ions or their ratios, TDS, water types, isotope compositions and other

measured/calculated parameters at each sampling location. The determination of recharge and

discharge areas is an important aspect of hydrogeological studies and therefore hydrochemical

maps are a useful tool for conceptualising flow within the groundwater system. Inclusion of

other geographical attributes such as surface topography, flowpaths and surface drainage can

also aid with spatial interpretation.

In addition to the above qualitative methods for analysing hydrochemical data, semi-

quantitative/quantitative techniques include geochemical modelling, such as the calculation of

saturation indices i.e. the degree of saturation with respect to specified minerals in a

groundwater sample (Section 5.3.2). This is useful for comparing interactions between aquifers

and tracing groundwater evolution along flowpaths. Mass balance calculations also allow

mixing proportions between waters to be estimated, where the end-member compositions are

well constrained and mixing is conservative.

Having outlined the key hydrochemical methods that will be utilised in this chapter, it is

important to note that in general, different graphical/numerical techniques will be relevant to

different hydrogeological settings. Therefore, hydrochemical analyses have greatest value when

analysed and interpreted within a sound and reliable hydrogeological context (Herczeg and

Edmunds, 2000). Moreover, interpretation of hydrochemical data for tracing water flowpaths is

best done by collecting data along hydraulic transects, or where hydraulic connection can be

assured (Woessner, 2000).

5.2 HYDROCHEMICAL PATTERNS

Groundwaters of the lower Herbert River catchment cover a broad compositional range in

relation to major ion chemistry. As illustrated in Figure 5-2, groundwaters are dominated by

sodium (Na+), calcium (Ca2+), bicarbonate (HCO3-) and chloride (Cl-) ions and are relatively

depleted in magnesium (Mg2+) and sulfate (SO42-). Potassium (K+) is not included on the Piper

diagram due to its low concentration compared to other major ions. Note that for convenience,

ions are herein referred to without their charge, with the exception of species of nitrogen. A

summary of major and minor inorganic chemistry for all groundwater samples is provided in

Appendix A.

Page 121: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

97

80 60 40 20 20 40 60 80

20

40

60

80 80

60

40

20

20

40

60

80

20

40

60

80

Ca Na HCO3 Cl

Mg SO4

RainfallHSsHSdSeawater

Figure 5-2 Piper diagram for deep (HSd) and shallow (HSs) aquifer samples collected during three sampling periods: May 2004, October 2004, June 2005. Note that only samples along the Herbert River valley and lower Stone River are shown.

Analysis of stable isotopes of water indicates that deep and shallow groundwaters primarily lie

on the LMWL (Figure 5-3). Differences in isotopic values along the LMWL result from a

combination of the elevation, amount and continental effects (Section 5.1.1.3). The amount

effect and differences in the sources of recharge can account for the observed isotope

differences between the two aquifers. For example, in general, deep aquifer waters are more

depleted (more negative) in heavy isotopes, consistent with large recharge events. In contrast,

shallow aquifer waters are more enriched in heavy isotopes, representative of smaller rainfall

events and diffuse recharge. The most depleted groundwaters depicted in Figure 5-3 (enclosed)

are from the upper Stone River valley. Note that the uncertainty in δ2H and δ18O is 1 and 0.15 o/oo, respectively; thus, the departure of some samples from the LMWL, outside of this

uncertainty, is indicative of other processes influencing the groundwater isotopic composition.

Further discussion of isotope data is deferred to Section 5.2.2. Note that although the rainfall

sample for major ion analysis was collected inland, stable isotopes were only determined for a

coastal rainfall sample.

Page 122: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

98

-7 -5 -3 -1 1-40

-30

-20

-10

0

10

δ2H (o/oo, SMOW)

δ18O (o/oo, SMOW)

SMOW

coastal rainfall MoretonBay

GMWL

Local meteoricwater line

Brisba

ne W

L

RainfallHSsHSdSeawater

-7 -5 -3 -1 1-40

-30

-20

-10

0

10

δ2H (o/oo, SMOW)

δ18O (o/oo, SMOW)

SMOW

coastal rainfall MoretonBay

GMWL

Local meteoricwater line

Brisba

ne W

L

RainfallHSsHSdSeawater

-7 -5 -3 -1 1-40

-30

-20

-10

0

10

δ2H (o/oo, SMOW)

δ18O (o/oo, SMOW)

SMOW

coastal rainfall MoretonBay

GMWL

Local meteoricwater line

Brisba

ne W

L

RainfallHSsHSdSeawater

Figure 5-3 Oxygen-18 (δ18O) and deuterium (δ2H) stable isotope data for deep and shallow aquifer samples and a coastal rainfall event in May 2004. The isotopic composition of seawater at Moreton Bay (QLD) (Cresswell 2006, pers. comm.) and SMOW are also shown. Trend lines represent the LMWL (blue); GMWL (black solid); and Brisbane water line6 (black dashed). Groundwaters from the upper Stone River valley are enclosed (red).

5.2.1 Compositional groups

Hydrochemical groups within each aquifer can be distinguished based on similarities in the

relative concentrations of major ions and salinities. As outlined in Chapter 3 (Section 3.5),

groundwater samples were collected during three sampling periods: May 2004, October 2004

and June 2005. For the majority of bores sampled in at least two of these periods, there is little

variation in groundwater composition, especially in the deep aquifer (Figure 5-4). Therefore, as

samples from June 2005 have the best spatial coverage, they are used as a basis for identifying

broad hydrochemical groups within each aquifer.

6 Data for the Brisbane water line was obtained from the Isotope Hydrology Section database maintained by the International Atomic Energy Agency (IAEA, 2006). The Brisbane WL was calculated from monthly amount-weighted mean δ18O and δ2H values.

Page 123: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

99

80 60 40 20 20 40 60 80

20

40

60

80 80

60

40

20

20

40

60

80

20

40

60

80

Ca Na HCO3 Cl

Mg SO4

HSs samples

May 2004October 2004June 2005

a

80 60 40 20 20 40 60 80

20

40

60

80 80

60

40

20

20

40

60

80

20

40

60

80

Ca Na HCO3 Cl

Mg SO4

HSd samples

May 2004October 2004June 2005

b

Figure 5-4 Piper diagram for (a) HSs and (b) HSd groundwater samples displayed by month of collection. Note that only bores sampled in at least two of the three sampling periods are shown.

Schoeller plots (Figure 5-5) illustrate that HSs waters can be grouped according to the relative

concentrations of HCO3 and Cl. Within the HCO3-dominated type there are two distinct patterns

based primarily on the relative proportions of the three major cations: one group is enriched in

Na (Figure 5-5a), while the other group has a greater proportion of Ca and Mg (Figure 5-5b).

While Cl-dominated waters overlap in their relative proportions of cations, they are

distinguished by their HCO3/Cl ratio and relative concentrations of Na to Cl. For instance, one

group has a greater proportion of HCO3 to Cl and higher concentration of Na compared to Cl

(Figure 5-5c). Conversely, the other group has a hydrochemical pattern that resembles seawater

Page 124: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

100

and thus has a lower HCO3/Cl ratio and virtually equal concentrations of Na and Cl (Figure

5-5d). The hydrochemical relationships between the identified groups are illustrated in Figure

5-6.

a

b

c

d

Figure 5-5 Schoeller plots illustrating the two dominant hydrochemical groups observed in shallow aquifer samples in June 2005: (a) and (b) represent HCO3-enriched samples while (c) and (d) represent Cl-enriched samples, including seawater (red) (refer to Figure 5-6).

Page 125: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

101

Figure 5-6 Piper diagram for shallow aquifer samples in June 2005: shaded areas highlight waters with similar hydrochemical fingerprints. The blue and green shaded areas represent HCO3-dominated waters whilst the purple and red areas represent Cl-dominated waters (refer to Figure 5-5).

Similar to HSs, two dominant compositional groups are distinguished in HSd based on the

relative concentration of anions (Figure 5-7). HCO3-dominated waters are the most dilute of the

HSd waters: within the group there are samples that are relatively enriched in Na compared to

the other major cations (Figure 5-7a) and other samples enriched in Ca and Mg (Figure 5-7b).

The second compositional group is more enriched in Cl relative to HCO3, with both high and

low salinity trends. There is considerable variation in relative cation and anion concentrations

between samples in the low salinity subgroup; however, these waters define a distinct

hydrochemical cluster (Figure 5-8). Similarly, whilst the concentration of cations is variable

within the high salinity subgroup, particularly with respect to Ca (Figure 5-7d), the waters are

nonetheless grouped as one type due to the overlap in relative anion concentrations (Figure 5-8).

Note that while the Schoeller plots for some bores in both Cl enriched subgroups are similar,

high salinity waters generally have a greater concentration of Mg, Na and Cl; are enriched in Cl

relative to HCO3; and have Na < Cl (compare Figure 5-7 c and d).

Comparison of Schoeller plots indicates that there are similar hydrochemical patterns between

the aquifers. The Na-enriched and Ca-Mg enriched trends of HSs are similar to those in HSd,

while the Na = Cl trend of HSs resembles the high salinity trend of HSd. There is also overlap

between the low salinity and Na > Cl trends in the deep and shallow aquifers, respectively.

Although there are chemical variations within the hydrochemical groups identified, particularly

in regards to relative cation concentrations, these discrepancies are small compared to the major

Page 126: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

102

shifts in anion concentrations between the groups. The spatial distribution of ions in HSs and

HSd is examined in Section 5.2.3.

a

b

c

d

Figure 5-7 Schoeller plots illustrating the two dominant hydrochemical groups observed in deep aquifer samples in June 2005: (a) and (b) represent HCO3-dominated samples enriched in Na and Ca-Mg, respectively, while (b) and (c) represent the low and high salinity Cl-enriched samples, respectively, including seawater (red).

Page 127: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

103

Figure 5-8 Piper diagram for deep aquifer samples in June 2005: shaded areas highlight waters with similar hydrochemical fingerprints. The blue and green shaded areas represent HCO3-dominated waters whilst the purple and red shaded areas represent Cl-enriched waters (refer to Figure 5-7).

5.2.2 Linear trends

The previous analyses identified the main compositional groups in each aquifer based on major

ion concentrations in each sample. Relationships between solutes are examined in this section,

with data presented in terms of the main hydrochemical groups identified above. Plots of TDS

against Cl for each aquifer display a good linear relationship at high salinity (Figure 5-9);

however, at low salinity Cl is not a good indicator of total solutes in solution. Nevertheless, Cl

is used as a conservative reference ion in the following analyses in order to examine changes in

solute concentrations with increasing salinity and to identify hydrochemical clusters. An

increase in the concentration of solutes with Cl is indicative of groundwater evolution due to

processes such as water-rock/mineral interactions, evaporation and ion exchange. However, in

order to infer groundwater evolution with confidence, spatial trends along groundwater

flowpaths must also be examined (Section 5.4).

Page 128: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

104

1 10 100 1000 10000 10000010

100

1000

10000

100000TDS

Cl (mg/l)

HSd samples

SeawaterHigh salinityLow salinityCa-Mg enrichedNa enriched

1 10 100 1000 10000 1000001

10

100

1000

10000

100000TDS

Cl (mg/l)

HSs samples

Na = ClNa > ClCa-Mg enrichedNa enrichedSeawater

Figure 5-9 TDS vs Cl (mg/L) for all deep and shallow aquifer samples (refer to Figures 5-5, 5-6, 5-7 and 5-8 for classification into hydrochemical groups).

5.2.2.1 Deep aquifer

Bivariate plots of major ions against Cl show that groundwater compositions in HSd lie between

rainfall and seawater (Figure 5-10). Na enrichment relative to rainfall is indicative of cation

exchange reactions and interactions with Na-containing minerals. Whilst there is generally a

positive correlation of major ions with Cl, there are distinct deviations from linearity which

correspond to particular hydrochemical groups. Notably, at low salinity there are two clusters of

HCO3-dominated samples that are distinguished in the Ca vs Cl and Mg vs Cl plots: these

clusters correspond to the Ca-Mg enriched and Na enriched hydrochemical groups identified

(Section 5.2.1). Spatial comparison of these water types with the distribution of major soil types

in the area (up to 1 m below ground surface) highlights that Na-enriched waters are associated

with silty clays/loams that have a medium to high cation exchange capacity (CEC) (between 3.9

- 7.9 meq/100g) and have acid volcanic parent material (Wood et al., 2003). In contrast, Ca-Mg

enriched waters are associated with well drained soils (red loam and river bank soils) with a low

CEC of around 3.1 meq/100g (Wood et al., 2003). Calcite (CaCO3) and dolomite (CaMg(CO3)2)

dissolution in the presence of CO2-charged water infiltrating through the soil zone can account

for the Ca-Mg-HCO3 groundwater; coupled with the presence of clay minerals with

exchangeable Na+, Na-HCO3-dominated waters can also be produced through ion exchange.

Page 129: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

105

0 0 1 10 100 10000.01

0.1

1.

10.

100.

1000.Na (meq/l)

Cl (meq/l)0.01 0.1 1. 10. 100. 1000.

0.001

0.01

0.1

1.

10.

100.K (meq/l)

Cl (meq/l)

0.01 0.1 1. 10. 100. 1000.0.001

0.01

0.1

1.

10.

100.Ca (meq/l)

Cl (meq/l)0.01 0.1 1. 10. 100. 1000.

0.01

0.1

1.

10.

100.

1000.Mg (meq/l)

Cl (meq/l)

0.01 0.1 1. 10. 100. 1000.0.1

1.

10.HCO3 (meq/l)

Cl (meq/l)0.01 0.1 1. 10. 100. 1000.

0.001

0.01

0.1

1.

10.

100.SO4 (meq/l)

Cl (meq/l)

HSd samples

RainfallSeawaterHigh salinityLow salinityCa-Mg enrichedNa enriched

Figure 5-10 Bivariate plots of major ions against Cl for deep (HSd) groundwater samples collected during the three sampling periods in 2004-2005. Note that the rainfall sample was collected approximately 20 km inland from the coast.

Page 130: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

106

Geochemical processes that can produce the two HCO3 waters are represented by the reactions:

CaCO3 + H2CO3 → 2HCO3- + Ca2+ (5-3)

Ca2+ + 2Na(ad) ↔ 2Na+ + Ca(ad) (5-4)

where (ad) denotes cations adsorbed on clays (Freeze and Cherry, 1979).

In addition to the two clusters of waters at low salinity, there is a cluster of high salinity samples

in HSd with a different HCO3:Cl relationship to the remaining samples (e.g. HCO3 vs Cl, Figure

5-10). Some of these high salinity groundwaters have anomalous concentrations of Ca relative

to the linear trend, above that of seawater. Stable isotope compositions (for a subset of bores)

illustrate that the high salinity waters plot to the right of the LMWL (Figure 5-11). Although

evaporation prior to recharge can account for these compositions, the linear trend depicted in

Figure 5-11 is consistent with mixing between freshwater and seawater. Further evidence for

this is provided in Section 5.2.3.1 with reference to the spatial distribution of the main water

types. Note that bore 101A lies on the LMWL; however, allowing for the uncertainty in stable

isotope measurements (Section 5.2) it is plausible that the stable isotopic composition of this

bore also corresponds to the seawater mixing line. Stable isotopes for all bores except those in

the high salinity group indicate a purely meteoric origin, with no evidence of evaporation prior

to recharge. This is consistent with lateral recharge to the deep aquifer as opposed to slow

infiltration due to diffuse recharge (e.g. Figure 4-15b, Chapter 4).

The group of low salinity waters show considerable scattering with respect to the major ion-Cl

trends. However, based on visual inspection of the series of bivariate plots presented in this

section, a degree of hydrochemical evolution from HCO3-dominated waters (Ca-Mg and/or Na

enriched types) to low salinity Cl-dominated waters is possible. Lateral hydrochemical trends

along groundwater flowlines in the deep aquifer are specifically examined in Section 5.4 in

order to confirm this observation.

Page 131: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

107

-7 -5 -3 -1 1-40

-30

-20

-10

0

10 HSd samples

RainfallSeawaterHigh salinityLow salinityCa-Mg enrichedNa enriched

δ2H (o/oo, SMOW)

δ18O (o/oo, SMOW)

Seawater mixing line

SMOW

coastal rainfall

Local meteoricwater line

MoretonBay

49A

101B

48A36A

101A

-7 -5 -3 -1 1-40

-30

-20

-10

0

10 HSd samples

RainfallSeawaterHigh salinityLow salinityCa-Mg enrichedNa enriched

δ2H (o/oo, SMOW)

δ18O (o/oo, SMOW)

Seawater mixing line

SMOW

coastal rainfall

Local meteoricwater line

MoretonBay

49A

101B

48A36A

101A

Figure 5-11 Oxygen-18 (δ18O) and deuterium (δ2H) stable isotope data for HSd samples and a coastal rainfall event in May 2004. The isotopic composition of seawater at Moreton Bay (QLD) (Cresswell 2006, pers. comm.) and SMOW are also shown. Trend lines represent the LMWL (blue) and mixing with seawater (red). Bores common to Figure 5-16 are labelled.

5.2.2.2 Shallow aquifer

Bivariate plots of major ions against Cl in HSs are displayed in Figure 5-12. While there is a

linear relationship observed between Na and Cl, there are no dominant trends with respect to the

other ions and Cl. However, the ion-Cl plots in combination with other bivariate plots highlight

distinct groups of waters in HSs that can be attributed to various natural processes. Similar to

groundwaters in HSd, a population of dilute waters in HSs has elevated Ca and Mg

concentrations (e.g. Ca vs Cl and Mg vs Cl, Figure 5-12); this cluster corresponds to the Ca-Mg

enriched hydrochemical group. As established for the deep aquifer, Ca-Mg enriched waters are

associated with sandy soils that have a low CEC of ~3 meq/100g. In contrast, the Na enriched

waters are associated with silty clays/loam and thus have a higher clay content, with CEC ~4

meq/100g.

Page 132: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

108

0.01 0.1 1. 10. 100. 1000.0.01

0.1

1.

10.

100.

1000.Na (meq/l)

Cl (meq/l)0.01 0.1 1. 10. 100. 1000.

0.001

0.01

0.1

1.

10.

100.K (meq/l)

Cl (meq/l)

0.01 0.1 1. 10. 100. 1000.0.001

0.01

0.1

1.

10.

100.Ca (meq/l)

Cl (meq/l)0.01 0.1 1. 10. 100. 1000.

0.01

0.1

1.

10.

100.

1000.Mg (meq/l)

Cl (meq/l)

0.01 0.1 1. 10. 100. 1000.0.1

1.

10.HCO3 (meq/l)

Cl (meq/l)0.01 0.1 1. 10. 100. 1000.

0.01

0.1

1.

10.

100.SO4 (meq/l)

Cl (meq/l)

HSs samples

Na = ClNa > ClCa-Mg enrichedNa enrichedRainfallSeawater

Figure 5-12 Bivariate plots of major ions against Cl for shallow (HSs) groundwater samples collected during the three sampling periods in 2004-2005.

Page 133: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

109

As noted for the deep aquifer, Na = Cl waters in HSs display a different HCO3-Cl relationship to

other samples, with lower HCO3 for the concentration of Cl (e.g. HCO3 vs Cl, Figure 5-12).

Schoeller plots in Figure 5-5d illustrated that these waters resemble dilute seawater. Br-Cl

relationships (Figure 5-13) are also consistent with a dominant seawater origin. Note that while

these waters have a seawater signature, this does not necessarily imply mixing with seawater

per se. Coastal rainfall would also be expected to resemble a seawater source; for example, as a

starting point, the Br/Cl ratio of rainfall can be assumed to be similar to that of seawater

(Herczeg and Edmunds, 2000). Stable isotopes indicate that unlike the deep aquifer, there is no

evidence of mixing with seawater in the shallow aquifer (Figure 5-14). Therefore, samples lying

below the LWML may indicate secondary fractionation processes such as evaporation prior to

infiltration (e.g. bore 61B). The Na > Cl hydrochemical group is only represented by a few

samples. Available evidence from major ions suggests that these waters display similar ion-Cl

relationships to the Na = Cl group, although Na > Cl waters are more enriched in HCO3 relative

to Cl.

0.1 1. 10. 100. 1000.0.0001

0.001

0.01Br/Cl

Cl (meq/l)

HSs samples

Na = ClNa > ClCa-Mg enrichedNa enrichedSeawater

Figure 5-13 Bivariate plot of Br/Cl against Cl for shallow aquifer samples collected during the three sampling periods in 2004-2005.

Based on the above analyses, a summary of the main hydrochemical groups in the deep and

shallow aquifers and possible explanations for these waters types is provided in Table 5-1.

Page 134: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

110

-7 -5 -3 -1 1-40

-30

-20

-10

0

10 HSs samples

Na = ClNa > ClCa-Mg enrichedNa enrichedRainfallSeawater

δ2H (o/oo, SMOW)

δ18O (o/oo, SMOW)

Seawater mixing line

SMOW

coastal rainfall

Local meteoricwater line

MoretonBay

61B

-7 -5 -3 -1 1-40

-30

-20

-10

0

10 HSs samples

Na = ClNa > ClCa-Mg enrichedNa enrichedRainfallSeawater

δ2H (o/oo, SMOW)

δ18O (o/oo, SMOW)

Seawater mixing line

SMOW

coastal rainfall

Local meteoricwater line

MoretonBay

-7 -5 -3 -1 1-40

-30

-20

-10

0

10 HSs samples

Na = ClNa > ClCa-Mg enrichedNa enrichedRainfallSeawater

δ2H (o/oo, SMOW)

δ18O (o/oo, SMOW)

Seawater mixing line

SMOW

coastal rainfall

Local meteoricwater line

MoretonBay

61B

Figure 5-14 Oxygen-18 (δ18O) and deuterium (δ2H) stable isotope data for HSs samples and a coastal rainfall event in May 2004. The isotopic composition of seawater at Moreton Bay (QLD) (Cresswell 2006, pers. comm.) and SMOW are also shown. Trend lines represent the LMWL (blue) and mixing with seawater (red) (derived from deep aquifer samples). Bore 61B is labelled as representative of evaporation prior to recharge.

Table 5–1 Hydrochemical groups of the alluvial aquifers Observation Possible explanation

HCO3-dominated compositions (HSd and HSs)

Ca-Mg enriched

Na enriched

calcite/dolomite dissolution

associated with sandy soils (low CEC)

associated with clay soils (high CEC)

HSd: Cl-dominated compositions

low salinity

high salinity

hydrochemical evolution

mixing with seawater

HSs: Cl-dominated compositions

Na = Cl

Na > Cl

seawater origin

insufficient evidence, but similar to Na = Cl type

Page 135: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

111

5.2.3 Spatial trends

Hydrochemical relationships between the deep and shallow aquifers are examined in Section

5.3, while the lateral chemical evolution of groundwater through HSd and HSs is examined in

Section 5.4. The aim of this section is to place the hydrochemical patterns observed from the

Piper, Schoeller and bivariate plots into a spatial context. Nitrate (NO3-) is also included in the

anion plots to observe relationships with the main water types.

5.2.3.1 Deep aquifer

Spatial representation of the identified hydrochemical groups indicates that HSd waters that

display a high salinity, Cl-enriched trend (Figure 5-7d) are located within 10 km of the coast.

Anion and cation pie charts in Figure 5-15 clearly illustrate that these waters (red shaded area)

characteristically have a high TDS and relatively minor concentrations of anions other than Cl.

Consistent with stable isotope trends (Figure 5-11), a plot of monovalent ions against divalent

ions normalised to chloride provides further evidence of interactions with seawater (Figure

5-16). For example, when seawater intrudes into a coastal freshwater aquifer (trend 1, Figure

5-16), Ca-Na exchange on clays results in an increase in divalent ions in the groundwater

(increase in (Ca + Mg)/Cl). This is balanced by a relative depletion in monovalent ions relative

to seawater (decrease in Na/Cl) such that Δ(Ca + Mg) = -2ΔNa molar units, as depicted by the

theoretical mixing lines in Figure 5-16 (Vengosh, 2004). In contrast, seawater displaced by

freshwater results in an opposite reaction such that there is an increase in Na/Cl and decrease in

(Ca + Mg)/Cl (Appelo, 1994). Bores along trend 2 in Figure 5-16, characterised as low salinity

waters, may be influenced by this process.

Groundwaters in the middle of the catchment, down-flow of the main recharge area for HSd,

belong to the low salinity group (Figure 4-15). Therefore, as suggested from the linear trends

(Figure 5-10), it is possible that these waters represent evolved groundwater compositions.

However, it is noted that low salinity groundwaters (bores 36 and 38) are also observed at the

coast, which define trend 2 in Figure 5-16. Therefore, there may be different processes that give

rise to this hydrochemically diverse group of low salinity waters. It was previously noted that

HCO3-dominated Ca-Mg and Na enriched waters correlate with the distribution of soil type in

the lower catchment (Table 5-1). Ca-Mg enriched waters in the upper part of the catchment, to

the northwest, are characteristically elevated in NO3-. This has significance for the mobility of N

in groundwater, as discussed in Section 5.6.

Page 136: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

112

a

b

Figure 5-15 Pie charts illustrating the spatial distribution of (a) major anions and (b) major cations in HSd in June 2005 as a percentage of total meq/L. The size of the inner circle is proportional to TDS (mg/L). Screened depths are shown in metres below ground level at each bore (italics). Refer to Figure 5-7, Figure 5-8 and Figure 5-10 for corresponding Schoeller, Piper and bivariate plots. Note that the pie chart at bore 101 is displayed for the deeper screened interval (101A or 51A); bore 101B (not shown but in the same water group) is screened at 40-46 m below ground.

Page 137: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

113

0

1

2

3

4

5

6

7

8

0 1 2 3

(Ca+Mg)/Cl (molar)

Na/

Cl (m

olar

)

HSd boresSeawater

49A

101A

36A

48A

38A

53A

1

2

45A46A

47A44A

101B

0

1

2

3

4

5

6

7

8

0 1 2 3

(Ca+Mg)/Cl (molar)

Na/

Cl (m

olar

)

HSd boresSeawater

49A

101A

36A

48A

38A

53A

1

2

45A46A

47A44A

101B

Figure 5-16 Monovalent ions plotted against divalent ions, normalised to Cl for all HSd bores (purple) and seawater (blue). Inset: samples are scaled by TDS (mg/L) and depict two trends away from seawater composition (compare with bores labelled in Figure 5-11). Dashed red lines depict idealised mixing lines for seawater intruding into sediments.

5.2.3.2 Shallow aquifer

Spatial relationships between major anions and cations in HSs are depicted in Figure 5-17, with

the main compositional groups highlighted. Based on the samples collected, the spatial

distribution of hydrochemical groups shows that the Na = Cl waters are restricted to the

northeast. Interestingly, two of the three bores that define this water type have low TDS (150-

300 mg/L), while the third (bore 36B) displays elevated TDS (around 1000 mg/L). High TDS

waters comprising the Na > Cl hydrochemical group do not show an obvious spatial trend.

These observations are further examined in Section 5.4.2.

As illustrated in Figure 5-17, there is one bore that is screened at two depths within HSs: whilst

the shallowest bore (54B) is a Ca-Mg enriched type (pie chart displayed), the deeper interval

(54A) is a Na-enriched type. It was noted above that the spatial distribution of the Ca-Mg and

Na enriched waters can be attributed to soil type (Table 5-1). Furthermore, with reference to the

lithologic cross-sections in Chapter 4 (e.g. Figure 4-4), the two types of HCO3-dominated

waters relate to different sandy units within the shallow aquifer. Hydrochemical relationships

between these units within HSs and justification for their inclusion in one aquifer are discussed

in Section 5.3.3. Importantly, the relative concentration of NO3- is highest in HCO3-dominated

waters that are enriched in Ca and Mg. Factors that influence the spatial distribution of NO3- are

examined in Section 5.6.

Page 138: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

114

a

b

Figure 5-17 Pie charts illustrating the spatial distribution of (a) major anions and (b) major cations in HSs in June 2005 as a percentage of total meq/L. The size of the inner circle is proportional to TDS (mg/L). Screened intervals are shown in metres below ground level at each bore (italics). Refer to Figure 5-5, Figure 5-6 and Figure 5-12 for corresponding Schoeller, Piper and bivariate plots. Note that the pie chart at bore 54 is displayed for the shallowest screened interval (54B), which represents a Ca-Mg enriched composition.

Page 139: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

115

5.3 VERTICAL RELATIONSHIPS BETWEEN AQUIFERS

This section examines hydrochemical relationships between HSs and HSd in order to establish

the degree of vertical interaction between the aquifers and hence verify the conceptual model

developed in Chapter 4. The previous section showed that there is considerable overlap in the

chemistry of shallow and deep groundwaters with respect to major and minor ions, thus it is

difficult to define a unique chemical signature for each aquifer. However, similar to the bore

hydrograph analysis presented in Chapter 4 (Section 4.4), hydrochemical comparisons between

nested bores enable site-specific vertical relationships between the aquifers to be examined.

5.3.1 Unconfined and confined systems

The chemistry of subsurface waters is a complex function of many variables, including: the

composition of groundwater recharge; the mineralogical composition of subsurface rocks; and

the hydrogeologic properties of rocks, that influence the extent of water/rock interaction

(Langmuir, 1997). These factors result in chemical differences between shallow unconfined and

deeper semiconfined/confined aquifers. As discussed in Section 5.1.1, there are numerous key

processes that influence the chemistry of groundwater, namely: evapotranspiration, dissolution

and precipitation, ion exchange, sorption, redox reactions, and gas generation and consumption

(also refer to Figure 2-3). Whilst many of these processes are common to the unsaturated and

saturated zones, the relative importance of each process in each zone may vary. Evaporation, for

instance, is likely to influence the chemistry of a watertable aquifer; however, it is not a

dominant process in deep confined aquifer systems. In addition, the degree and extent of

weathering affects the residence time of water, which is generally shorter in shallow aquifer

systems compared to deeper systems. For example, chemical weathering is very active in

shallow unconfined aquifers due to the aggressiveness of recharge waters, relatively high

water/rock ratios and high groundwater velocities. Conversely, the groundwater chemistry of

confined systems tends to be rock-dominated (many minerals at saturation) due to low

water/rock ratios and long residence times. In general, groundwater systems in sedimentary

rocks become more confined with depth and the water becomes more saline (high TDS),

anaerobic, and isolated from fresh recharge (Langmuir, 1997). An important corollary of the

decline in dissolved oxygen with depth is that the concentration of NO3- also declines, or is

absent, due to microbial activity under anaerobic conditions. This is discussed further in Section

5.6.

Page 140: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

116

5.3.2 Saturation Indices

The calculation of saturation indices is a purely thermodynamic analysis of departure of a

solution from equilibrium, predicting whether samples are undersaturated, saturated or

supersaturated with respect to a particular mineral phase. This type of analysis can be useful for

understanding vertical and spatial hydrochemical relationships between groundwater samples

due to mineral dissolution and precipitation reactions. Whilst calculation of saturation indices is

theoretical, it can constrain the direction of reaction or the potential of a water to dissolve or

precipitate a mineral. The method provides a useful way to combine element concentrations as

well as chemical parameters such as pH, temperature and Eh into plausible minerals and

examine relative differences between samples. The saturation index, SI, is defined as the ratio of

the reaction quotient Q (or ion activity product) and the thermodynamic equilibrium condition

represented by the equilibrium constant Keq at the given temperature (Freeze and Cherry, 1979).

Therefore, SI = Q/ Keq for the particular mineral of interest. If SI > 1 the water contains an

excess of ionic constituents and thus mineral precipitation potentially occurs; conversely, SI < 1

is indicative of undersaturation and hence mineral dissolution is possible. If SI = 1 the reaction

is at equilibrium, which means that the solution is saturated with respect to the mineral of

interest. In this chapter SI’s are expressed in logarithmic form (i.e. SI = log[Q/ Keq]), in which

case a value of zero denotes the equilibrium condition.

Based on the ionic constituents present, SI’s were derived using the program PHREEQC

(Parkhurst, 1995), using the AquaChem7 graphical interface for the common weathering

minerals expected in an alluvial system derived from a granitic source. These minerals included:

chalcedony, halite, calcite, dolomite, gypsum, magnesite, siderite, albite, anorthite, K-mica,

phlogopite, annite, gibbsite, illite, kaolinite, pyrophyllite, Ca-montmorillonite, and goethite. A

phase diagram for the common aluminosilicates (i.e. log([K+]/[H+]) versus log(H4SiO4)

indicates that the sampled shallow and deep groundwaters primarily plot in the montmorillonite

stability field (Langmuir, 1997). Considering the assumptions inherent in phase diagrams, and

their applicability to natural systems, this suggests that smectite clays are likely to be stable.

Given the low permeability and poor drainage of smectite clays, this observation has

implications for the redox condition of groundwater and N leaching potential (Section 5.6).

Input values for the SI calculations included solute concentrations for each sample, field pH,

field Eh (pe) and temperature. Figure 5-18 depicts the Eh-pH distribution for HSs and HSd

groundwaters. Although there is a greater spatial distribution of samples collected in June 2005,

SI’s were only computed for samples from May 2004 and October 2004 due to unreliable Eh

measurements in June 2005. The calculations indicate that waters of HSs and HSd are generally

undersaturated with respect to the majority of minerals examined. However, all samples are

7 AquaChem v.3.7 for Windows 95/98/NT

Page 141: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

117

essentially saturated with respect to chalcedony (SiO2) (log SI ≥ -0.4) and generally

supersaturated with respect to goethite (FeOOH). Saturation indices for both saturated and

unsaturated phases are compared between nested bores in the following section as a semi-

quantitative tool to examine hydrochemical interaction between aquifers. In Section 5.4, the

extent of lateral groundwater evolution along the flowpath is also verified by the application of

these indices.

5 6 7 8100

200

300

400

500Eh (mV)

pH (field)

HSs samples

Na = ClNa > ClCa-Mg enrichedNa enriched

5 6 7 8100

200

300

400

500Eh (mV)

pH (field)

HSd samples

High salinityLow salinityCa-Mg enrichedNa enriched

Figure 5-18 Field measurements for shallow and deep aquifer samples collected in May and October 2004.

5.3.3 Intra-aquifer relationships

In Chapter 4 it was established that based on the available stratigraphic and hydraulic

information, the subsurface can be conceptualised as a two-aquifer system. In order to support

this interpretation, hydrochemical signatures are examined for bores screened at different

intervals within the same aquifer. Three bores in the study area allow this type of comparison:

bore 46 (intervals 74 and 46A) and bore 51 (intervals 101B and 51A/101A) are screened at two

depths in HSd, whilst bore 54 (54B and 54A) is screened at two depths in HSs (refer to Figure

4-3). With reference to the hydrochemical groups identified in Section 5.2, Schoeller plots

indicate that at the respective bores, the two intervals screened in HSd fall within the envelope

of major ion concentrations of the high salinity chemical group (compare Figure 5-19a and b

with Figure 5-7d). The main hydrochemical difference between the nested intervals at each bore

is largely in the Ca/Mg ratio, which is greater in the deeper interval of HSd. Saturation indices

for the two intervals in bore 101 (pipes A and B) suggest that the deeper interval is more

Page 142: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

118

saturated with respect to both dolomite and calcite but otherwise has identical SI’s for other

minerals (Figure 5-20a). At bore 54, the two screened intervals in HSs represent the Ca-Mg

enriched (shallower) and Na enriched (deeper) chemical groups (compare Figure 5-19c with

Figure 5-5a and b). Whilst these groups of waters are distinct, the two intervals have very

similar SI’s for the calculated minerals, suggestive of a common source of water (Figure 5-20b).

Nonetheless, as highlighted in Section 5.2.3.2, the distinction between the Ca-Mg and Na

enriched groups is important from the perspective of tracking NO3- in groundwaters. This is

discussed further in Section 5.6. Therefore, while the hydraulics indicate that the different sand

units behave as one aquifer, the hydrochemistry is a more sensitive variable at small scale.

On balance it is considered that the hydrochemistry of the three bores examined in this section

provide evidence of good hydraulic continuity between the screened intervals within the same

aquifer, consistent with the bore hydrograph trends (Chapter 4).

a

b

c

Figure 5-19 Schoeller plots for two screened intervals within the same aquifer based on available random measurements (1975 - 2005). (a) and (b) represent screened intervals in HSd while (c) represents screened intervals in HSs (refer to Figure 4-3 in Chapter 4). The red trend lines highlight samples from the shallower screened interval in each aquifer. Source: QDNRW.

shallower

deeper shallower

deeper

deeper

shallower

Page 143: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

119

-6

-2

2

6

Mag

nesi

te

Side

rite

Goe

thite

Cha

lced

ony

Gyp

sum

Cal

cite

Dol

omite

log

SI 101B

101A

a

-10

-6

-2

2

Mag

nesi

te

Cha

lced

ony

Cal

cite

Dol

omite

Gyp

sum

log

SI

54B

54A

b

Figure 5-20 Plots of saturation indices (logarithmic form) for intervals screened within (a) the deep aquifer and (b) the shallow aquifer. The A-pipe is the deeper interval (compare with Figure 5-19). The dotted line denotes the equilibrium condition where SI = 1.

5.3.4 Inter-aquifer mixing trends

Hydrochemical comparisons between nested bores enable site-specific vertical relationships

between the aquifers to be examined. Three broad patterns emerge from Schoeller plots of

nested bores screened in the shallow and deep aquifers (Figure 5-21):

Pattern 1: HSs and HSd have ion concentrations of a different magnitude but the separation

is common for all ions i.e. parallel Schoeller plots;

Pattern 2: HSs and HSd have ion concentrations of the same magnitude for all ions i.e.

coincident Schoeller plots; or,

Pattern 3: HSs and HSd have different concentrations of major ions and different proportions

of each ion i.e. non-parallel Schoeller plots.

Page 144: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

120

As discussed in Section 5.3.1, these trends between nested bores arise because of a combination

of factors, including differences/similarities in aquifer material and reactivity of the rocks;

groundwater residence time; volume of water transmitted; water source; and chemical processes

in shallow versus deep aquifers. Therefore, while parallel Schoeller plots indicate that the

groundwater in each aquifer is of a similar composition and non-parallel trends indicate the

converse, these trends do not necessarily correlate with the degree of inter-aquifer connectivity.

However, in combination with SI’s (Figure 5-22) it is possible to infer whether the observed

hydrochemical trends can be attributed to similarities or differences in source waters or

chemical conditions. Therefore, although calculated SI’s are a function of the numerous factors

listed above, similar trends in SI’s at nested bores are indicative of similar processes and/or

similar water-rock interactions. Whilst the potential degree of vertical connectivity between

aquifers was assessed in Chapter 4 (Section 4.4.2) based on hydraulic information, the

interpretation of hydrochemistry can provide further insight into inter-aquifer processes and

interactions than aquifer hydraulics alone.

Schoeller plots for nested bores 68, 61, 60, 36 and 38 display more-or-less parallel trends for the

majority of ions (Figure 5-21). Whilst the deeper aquifer generally represents the more saline

water type (greater upwards vertical displacement of Schoeller plot), in some bores the trends

overlap and give rise to pattern 2, a special case of pattern 1. The remaining nested bores

depicted in Figure 5-21 display non-parallel trends and represent pattern 3.

Hydrographs for bores in the western half of the catchment indicated a downwards vertical head

gradient and strong to good vertical connectivity between HSs and HSd (Figure 4-7). Schoeller

plots for bores 68, 61 and 60 are consistent with this interpretation: the separation between

trends also indicates that the degree of vertical leakage is greatest in the vicinity of bores 60 and

61. The separation between SI trends (between nested bores for the same minerals) provides

further evidence of the same relative degree of connectivity at each location (e.g. Figure 5-22a

and b). Given the similarity in SI’s, the observed mismatch between Schoeller plots at bore 61

in regards to SO4 and HCO3 is indicative of Eh-pH controls between aquifers, such as the

reduction of SO4 as groundwater moves to the more reduced environment of HSd compared to

HSs. Although the disparate ionic concentrations at bore 62 result in non-parallel Schoeller

trends, SI’s illustrate that there is a similar trend of enrichment and depletion of the calculated

minerals in each aquifer (Figure 5-22c). Thus, while there is a degree of confinement between

aquifers in the vicinity of bore 62, a common recharge source is probable. The greater

enrichment in bicarbonate and other solute concentrations in the deep aquifer is consistent with

longer residence times and/or mixing with older water.

Page 145: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

121

For bores in the eastern half of the catchment, the interpretation of non-parallel Schoeller trends

in relation to the degree of vertical connectivity is complicated by secondary processes and/or

seawater influences, that particularly enrich the deep aquifer in NaCl relative to the shallow

aquifer. Where there is an upwards head gradient, there is the additional influence of direct

rainfall to the shallow aquifer that may dilute the HSd signal from vertical leakage. Based on

hydrograph analysis, poor downwards vertical connection between aquifers was established at

bores 46, 51 (101) and 53. Hydrochemical trends between nested bores are consistent with the

observed hydraulic relationships. The large difference in ionic concentrations between aquifers,

particularly at bores 51 (101) and 46, are indicative of different source waters to each aquifer. It

was shown in Section 5.2.3.1 (Figure 5-16) that the hydrochemistry of these bores is consistent

with seawater intruding into the aquifer sediments.

Based on their historical water level elevations, upwards vertical connectivity between HSs and

HSd was evident at bores 48, 49, 36 and 38. Schoeller and SI plots (Figure 5-21 h-k and Figure

5-22 d-f) illustrate that while there is the potential for exchange of water between aquifers, there

is clearly a degree of confinement in the vicinity of bores 48 and 49: poor connectivity is

established. In contrast, the near-coincident Schoeller plots at bores 36 and 38 and identical SI’s

imply that the same waters are present in each aquifer at these locations. Given that the head

gradient is upwards in these bores, the hydrochemistry is consistent with a large component of

vertical discharge from HSd to HSs.

The hydrochemical analyses highlight that differences between Schoeller plots for most nested

bores provide a good indication of the relative degree of vertical hydraulic continuity between

aquifers. This is useful given that the calculation of SI’s requires reliable field measurements.

However, as exemplified by bore 62, disparate Schoeller plots in isolation of SI calculations

may lead to misinterpretation. Hydraulic relationships are also important for interpreting

hydrochemical trends.

Page 146: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

122

a b

c d

e f

Figure 5-21 Schoeller plots for groundwater samples from shallow (blue) and deep (black) nested bores in June 2005. The corresponding bore hydrographs indicate that vertical connectivity is: (a-d) strong/good (downwards head gradient); (e-g) poor (downwards head gradient), (h-i) poor (upwards head gradient); and (j-k) good (upwards head gradient) (refer to Figure 4-7). Note the change in vertical scale in the plots.

Page 147: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

123

cont.

g h

i j

k

#S

#S

#S

#S

#S

#S

#S

#S#S

#S

#S6861

60

Herbert R

4662

3853

51

48

49 36

Figure 5-21 Schoeller plots for groundwater samples from shallow (blue) and deep (black) nested bores in June 2005. Refer to inset map for bore locations.

Page 148: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

124

-10

-6

-2

2

Cha

lced

ony

Gyp

sum

Cal

cite

Mag

nesi

te

Dol

omite

log

SI 68B

68A

a

-10

-6

-2

2

Cha

lced

ony

Gyp

sum

Mag

nesi

te

Cal

cite

Dol

omite

log

SI 61B

61A

b

-10

-6

-2

2

Cha

lced

ony

Mag

nesi

te

Gyp

sum

Cal

cite

Dol

omite

log

SI

62B

62A

c

Figure 5-22 Plots of saturation indices (logarithmic form) for selected nested bores based on samples collected in October 2004 (compare with Figure 5-21). The dotted line denotes the equilibrium condition (SI = 1). Graphs a-c represent bores with a downwards vertical head gradient.

Page 149: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

125

cont.

-10

-6

-2

2

6

Cha

lced

ony

Cal

cite

Gyp

sum

Mag

nesi

te

Dol

omite

log

SI

48B

48A

d

-10

-6

-2

2

6

Cha

lced

ony

Goe

thite

Side

rite

Mag

nesi

te

Cal

cite

Gyp

sum

Dol

omite

log

SI

49B

49A

e

-10

-6

-2

2

6

Mag

nesi

te

Cha

lced

ony

Goe

thite

Side

rite

Gyp

sum

Cal

cite

Dol

omite

log

SI

36B

36A

f

Figure 5-22 Plots of saturation indices (logarithmic form) for selected nested bores based on samples collected in October 2004 (compare with Figure 5-21). The dotted line denotes the equilibrium condition (SI = 1). Graphs d-f represent bores with an upwards vertical head gradient.

Page 150: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

126

In addition to the graphical methods used in this section to examine vertical connectivity

between the aquifers, semi-quantitative estimates of the proportion of vertical mixing can be

determined by simple mass balance calculations based on solute concentrations. As this

approach assumes conservative mixing between two end-members, the reliability of the

estimates is a function of how well the end-member chemical compositions can be constrained

(Herczeg and Edmunds, 2000). The hydrochemistry of the most dilute HSd waters (in the

northwest of the catchment, Figure 5-15) and that of HSs at the nested bore of interest are

plausible end-members. However, given the semi-confined nature of the deep aquifer, mixing

proportions at any bore represent an overestimate. Therefore, due to difficulties in defining the

end members, the percentage of mixing could not be estimated by this method. More

sophisticated geochemical modelling techniques (Parkhurst, 1995) are considered beyond the

scope of the thesis.

5.3.5 Relationship with the bedrock

A comprehensive search of the QDNRW groundwater database was carried out to select bores

screened in the bedrock which contain stratigraphic logs, screened intervals of known depth and

water chemistry data. The majority of deep bores in the study area and surrounding region are

screened in the alluvial aquifer, with only a few bores screened in weathered granitic bedrock:

no bores are screened exclusively in unweathered bedrock. Schoeller plots indicate that

although groundwater in the weathered bedrock is spatially variable, there is a similar

hydrochemical pattern to the deep alluvial aquifer at each location. The lack of a distinct

hydrochemical signature suggests there is vertical connectivity between the alluvial aquifer and

weathered zone of granitic bedrock; however, the relationship with underlying unweathered

bedrock is unclear.

5.4 SPATIAL RELATIONSHIPS WITHIN AQUIFERS

Lateral relationships in each aquifer are examined in this section to verify the inferred flow

paths established in Chapter 4 (Figure 4-18). The spatial distribution of the main hydrochemical

groups in each aquifer was illustrated in pie charts in Figure 5-15 and Figure 5-17. In Chapter 4

it was noted that with the exception of a marine mud at the coastal fringe, sedimentation in the

case study area occurred from terrestrial deposition (Section 4.2). Given the geological setting

and basement rock, terrestrial sediments are likely to have originated from an igneous source.

As such, the chemical evolution of groundwater is likely to be dominated by the dissolution of

feldspars, micas and other silicate minerals, with the release of Na+, K+, Mg2+ and Ca2+ ions to

the water (Freeze and Cherry, 1979). The key processes that influence the chemistry of

groundwater in both unconfined and confined aquifers were outlined in Sections 5.1.1 and 5.3.1.

Page 151: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

127

5.4.1 Lateral hydrochemical evolution

5.4.1.1 Deep aquifer

A shift from Ca-Mg enriched groundwaters dominated by HCO3 in the northwest of the

catchment, to high salinity waters with seawater-type signatures towards the coast is observed in

the deep aquifer (Figure 5-15). Low salinity group waters with Cl > HCO3 dominate the middle

section of the catchment. By considering the major ions which each comprise greater than 20 %

of the total cation and anion concentrations respectively (expressed in meq/L), the spatial anion

and cation pie charts can be combined into a single hydrochemical facies map (Figure 5-23a).

Analysis of water level contours in Chapter 4 indicated that a major rainfall-recharge zone for

the deep aquifer is present in the northwest of the catchment, while discharge occurs towards the

coast (Figure 4-15b). In accordance with the principles of hydrochemical evolution discussed in

Section 5.1.1.2, there is evidence of a general trend from HCO3-dominated waters with low

TDS values in the main recharge area (e.g. bores 71A, 69A) to Cl-dominated waters with high

TDS values towards the discharge area (e.g. bores 48A, 38A, 36A) during both the beginning

and end of the dry season (e.g. Figure 5-15). However, examination of hydrochemical trends

along groundwater flowpaths in HSd, such as along the main flowpath depicted in Figure 5-23,

indicates that there are clearly other processes occurring that interrupt lateral hydrochemical

development. For example, it is apparent that while there is an increase in TDS away from the

main recharge area in the northwest, waters with TDS less than 300 mg/L are observed further

down the flowpath from bore 68A (Figure 5-23b). In light of the inter-aquifer relationships

established in Section 5.3.4, these dilute waters are consistent with enhanced vertical leakage

from the shallow aquifer and/or proximity to a local recharge area (e.g. north of the Herbert

River, refer to Figure 4-15b). Furthermore, an increase in TDS between bores 61A and 59A is

consistent with groundwater contributions from the southwest.

Evidence of both vertical and lateral recharge sources is provided by stable isotope

measurements. As illustrated in Figure 5-24, there is a progressive enrichment in stable isotopes

from the beginning of the main flowpath (bore 69A) to bore 61A. Given the generally enriched

isotopic signature of the shallow aquifer over the deep (Figure 5-3), this trend along the

flowpath is consistent with isotope enrichment due to mixing between the shallow and deep

aquifers (as discussed in Section 5.3.4). However, a decline in heavy isotopes between bores

61A and 59A is considered to be due to groundwater from the upper Stone River valley, which

is more depleted compared to waters from the northwestern recharge area (Figure 5-3). Note

that the concentration of stable isotopes in groundwater depends mainly on the origin of water:

effects of water-rock interactions on water isotopic composition generally become important

only at high temperature (Gonfiantini et al., 1998).

Page 152: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

128

a

#S

#S

#S#S

#S

#S#S#S

#S

#S

#S

#S

#S

#S

#S

#S#S #S#S

#S

#S

#S

#S

#S

#S

#S

#S

216652

166138

392

260

578

220 794

128

134

136

704

634

384

1076

1050

2416

2662

B: 3644

24941670

1048

3434

14120

A: 4924

> 3000#S

< 300#S

300 - 750#S

2000 - 3000#S750 - 2000#S

TDS (mg/L)

5 0 5 Km

N

HSd TDS

71A69A

68A

67A

61A

62A

63A 58A

59A

60A

52A53A

45A

44A38A

37A46A

47A

36A48A

49A

101A

130A

127B

101B

Stone R

Herbert R

b

Figure 5-23 (a) Hydrochemical facies and (b) distribution of total dissolved solids determined at bores in the deep aquifer in June 2005. Selected flowpaths from the northwest to the coast as well as from the Stone River are also depicted. Note that A and B refer to different screened intervals within the same aquifer (interval A is the deepest). The shaded areas in (a) correspond to the major water types identified: Ca-Mg enriched (green); Na enriched (blue); low salinity (purple); high salinity (red) (refer to Figure 5-15). QDNRW monitoring bores are labelled in (b) (domestic bores are not labelled).

Page 153: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

129

Given that HSd groundwaters are of relatively low temperature (maximum 30oC) it can be

assumed that stable isotopes of water are unaffected by secondary processes. Therefore, the

observed trends are consistent with contributions from isotopically different waters.

-34

-32

-30

-28

-26

-24

-22

69 68A 67 61A 59

HSd bores down the flowpath

Deut

eriu

m (

o / oo,

SM

OW

)

-5.6

-5.4

-5.2

-5

-4.8

-4.6

Oxy

gen-

18 (

o / oo,

SM

OW

)

Figure 5-24 Stable isotopic values along a flowpath for the deep aquifer based on samples collected in May 2004.

Hydrochemical evolution is further obscured in the eastern part of the catchment, where there is

a dramatic increase in TDS values. The corresponding Na-Cl type waters are indicative of a

different aquifer material or mixing with higher salinity waters. From the analysis of ion ratios

(e.g. Figure 5-10 and Figure 5-16) and stable isotope evidence (Figure 5-11), the observed TDS

increase is attributed to mixing with seawater. However, as illustrated in Figure 5-23b, there is a

decrease in TDS at the coast; as such, the spatial distribution of TDS does not correspond with

the present-day coastline. Possible mechanisms to account for these observations are proposed

in Section 5.4.2.

In agreement with the observed TDS distribution, SI’s for the calculated minerals (Section

5.3.2) increase immediately down-flow from the main recharge area (e.g. between bore 69A and

68A). In addition, in the high TDS zone (e.g. bore 47A) there is a marked increase in SI’s,

particularly with respect to gypsum and magnesite: this is consistent with a seawater influence.

Although groundwaters are generally undersaturated with respect to the calculated mineral

phases, fluctuations in SI’s generally mimic major ion and TDS trends in response to influences

from dilute and enriched groundwaters contributing along the flowpath. Importantly, the main

flowpath depicted in Figure 5-23 is a sink for groundwater from different areas in the

catchment. Therefore, whilst there is overall an increase in solute concentrations and TDS from

the main recharge area to the coast, the degree of hydrochemical evolution is suppressed

because of the vertical connectivity between aquifers and the convergence of flowpaths

contributing groundwaters of different compositions.

Page 154: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

130

5.4.1.2 Shallow aquifer

A hydrochemical facies map for the shallow aquifer highlights the distribution of Na-HCO3-Cl

facies groundwater in the west, interspersed with Na-Cl-HCO3 waters in the east; Na-Cl facies

groundwaters are restricted to the northeast (Figure 5-25a). Note that groundwater at bore 101C

(e.g. Figure 5-25b) is classified as belonging to the Na-HCO3-Cl facies (with low TDS) during

other sampling periods, rather than the high salinity Na-Cl type depicted in Figure 5-25a.

Similarly, low TDS measurements have also previously been obtained at bore 57A (QDNRW

database). Given the instability of EC readings during pumping, contamination from the deep

aquifer is possible at both of these bores, although further sampling would be required to

confirm this.

Analysis of water level contours in Chapter 4 indicated that recharge to HSs occurs from both

diffuse and lateral sources (Figure 4-17b). The interpretation of both diffuse and lateral recharge

is supported by available stable isotope data for the shallow aquifer (Figure 5-14). Waters lying

below the LMWL reflect evaporation and therefore slow infiltration during diffuse recharge. In

contrast, waters lying on the LMWL represent rapid infiltration to the shallow aquifer (lateral

and/or diffuse recharge). Given these recharge mechanisms the lack of obvious hydrochemical

evolution is not unexpected. For example, there is no systematic increase in TDS along the main

flowpath depicted in Figure 5-25b. Although groundwater from the northwest becomes more

saline down the flowpath, there is a decrease in TDS at bore 53B. However, the slight increase

in TDS between bores 53B and 46B, as well as between 53B to 48B (and further to the

northeast), are indicative of minor hydrochemical development. It was noted in Section 5.3.3

that the Ca-Mg and Na enriched waters represent different sandy units within the shallow

aquifer. Hydrochemical evolution is more pronounced within the Na enriched unit, as

highlighted by a hypothetical flowpath between bores 68B-60B-50A-49B (Figure 5-25b). This

unit representing Na enriched waters is at a greater depth and has a higher clay content than the

unit representing Ca-Mg waters. These attributes may result in a slightly longer residence time

and hence enhanced TDS evolution along the flowpath. However, along the main flowpath

depicted in Figure 5-25b there is little evidence of hydrochemical evolution. Given that HSs is

characterised as a watertable aquifer, evaporation may affect the hydrochemistry and thus there

may be localised areas of high TDS. The absence of strong evidence for hydrochemical

evolution in the upper sandy unit is consistent with relatively short residence times and

therefore less opportunity for water-rock interactions and other processes to influence the

chemistry of the groundwater.

Page 155: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

131

a

#S

#S

#S#S

#S#S

#S

#S

#S

#S

#S

#S#S

#S

#S

#S

#S

#S

#S#S#S

#S

#S

#S

#S

82

258140

120710

144

154176

114

142

B: 152A: 292

122206

320

172

112

210218

282

576

256

1152

1000

< 150#S

150 - 300#S

800 - 1200#S300 - 800#S

TDS (mg/L)

5 0 5 Km

N

HSs TDS

68B

61B

62B

60B

53B

38B

46B

36B

48B

49B

130B

101C

57A54B54A

50A

b

Figure 5-25 (a) Hydrochemical facies and (b) distribution of total dissolved solids determined at bores in the shallow aquifer in June 2005. Potential flowpaths are depicted. Note that A and B refer to different screened intervals within the same aquifer (interval A is the deepest). The shaded areas in (a) correspond to the major water types identified: Ca-Mg enriched (green); Na enriched (blue); Na > Cl (purple); Na = Cl (red) (refer to Figure 5-15). QDNRW monitoring bores are labelled in (b) (domestic bores are not labelled).

Page 156: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

132

5.4.2 Seawater intrusion

The preceding analyses have identified water types in each aquifer that have high salinity and/or

seawater-type signatures. Based on major element relationships and stable isotope evidence

(Section 5.2), seawater intrusion into the deep aquifer is proposed, while meteoric influences are

suggested for the shallow aquifer. In coastal areas, fresh groundwater derived from precipitation

on the land comes in contact with and discharges into the sea or into estuaries containing

brackish water (Heath, 1987). Therefore, depending on the height of the watertable and

thickness of the freshwater lens, saltwater intrusion can influence the salinity of an aquifer. An

estimate of depth of the saltwater interface as a function of distance from the coast can be made

by application of the Ghyben-Herzberg relation; this represents a theoretical relationship that

generally leads to an underestimate of the depth to the saltwater interface due to the assumption

that the freshwater and seawater zones are static (Freeze and Cherry, 1979).

Groundwater elevations of bores screened in the unconfined aquifer (HSs) were used to depict

the theoretical saltwater wedge below freshwater along three transects (Figure 5-26). Two

periods were chosen, May and October 2004, to approximately represent maximum and

minimum watertable elevations: this translates to minimum and maximum elevations of the

interface, respectively. Comparison of the saltwater interfaces with slotted depths of bores along

transects 1, 2 and 3 indicates that HSd bores within approximately 10 km from the coast are

screened within the seawater zone, and thus have the potential to be intruded by seawater

(Figure 5-27). Whilst this includes bores 37A, 48A and 49A, the estimates suggest that the other

high salinity bores such as 46A, 47A and 101A are screened within the freshwater zone.

However, given that in reality the interface separating freshwater and saltwater is not a sharp

boundary, there may be mixing of saltwater and freshwater in a zone of diffusion around the

interface (Freeze and Cherry, 1979). This is consistent with trend 1 in Figure 5-16.

Figure 5-26 Transects for estimating the theoretical saltwater wedge below freshwater according to the Ghyben-Herzberg relation (Figure 5-27). The red line encloses high salinity type waters. Bores are labelled without their 116000 prefix.

#S

#S

#S

#S

#S

#S#S

#S#S

#S

#S

#S

#S

4737Herbert

R

46

4438

101

48

49 36

53

61

45

62

Transect 1

Transect 2

Transect 3

Page 157: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

133

The spatial distribution of TDS in HSd at the coast (Figure 5-23b), as decreasing from north to

south, is consistent with seawater intrusion from a source from the north, rather than from the

east. Water level contours presented in Chapter 4 indicated a northeasterly trending groundwater

flowpath near the coast (e.g. Figure 4-14a) that may also be a preferential flowpath for the

inflow of seawater to the deep aquifer. Alternative explanations for the source of salinity in the

deep aquifer could be relic seawater in the aquifer sediments or a seawater mixing zone around

the river. Characteristics of residual evaporated seawater include: high TDS (> 35 g/L); Ca-

chloride composition (Ca/(SO4 + HCO3) > 1); molar Na/Cl ratio below that of modern seawater

(0.86); Br/Cl ≥ seawater ratio (1.5 x 10-3) and relative depletion of sulfate (SO4/Cl < 0.05)

(Vengosh, 2004). Whilst not all factors are satisfied by any of the bores in the study area, the

high TDS, low Na/Cl ratio (average 0.6) and Ca-Cl character of bores 101A (51A) and 49A

(Figure 5-23) are indicative of relic seawater in at least these locations. Note that evaporites

have not been described in the aquifer sediments, nor are there characteristically low Br/Cl

ratios to suggest groundwater flow through evaporites (Herczeg and Edmunds, 2000).

Membrane effects such as salt filtering generally occur at depths of greater than 500-1000 m

below the ground surface (Freeze and Cherry, 1979) and are therefore considered unlikely in

HSd, which has a maximum depth of approximately 100 m below the surface.

The estimates of the saltwater interface suggest that coastal bores 36A and 38A are also

intruded by seawater. While available stable isotopic evidence at bore 36A does not exclude this

possibility (e.g. Figure 5-11), these bores have a characteristically low TDS compared to bores

further west. Note that there is an upwards vertical head gradient towards the coast, which

would preclude leakage of freshwater from the shallow aquifer to the deep. As illustrated in

Figure 5-23a, there are two compositional groups distinguished within the Na-Cl facies.

Specifically, bores 53A, 36A and 38A are low salinity-type waters (purple shading) with a

Na/Cl molar ratio ranging from 1.1-1.5. In contrast, the other high TDS bores in the Na-Cl

facies (red shading) have a lower Na/Cl ratio, between 0.8-0.96. As discussed in Section 5.2.3.1,

two trends can be observed as a result of interactions with seawater, arising from the intrusion

of seawater into aquifer sediments or displacement of seawater by freshwater. Based on the ratio

between Na/Cl and (Ca+Mg)/Cl, it is therefore plausible that the low salinity trend can be

explained by the latter process, consistent with trend 2 in Figure 5-16. Furthermore, given the

upwards vertical head gradient and strong vertical connectivity between the shallow and deep

aquifers in the coastal bores (Section 5.3.4), the possibility exists for sites of local groundwater

discharge in the vicinity of bores 36 and 38, where seawater is forced back by freshwater.

Page 158: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

134

Theoreticaldepth May 2004

TheoreticaldepthOctober 2004

Slotted depthHSd

Slotted depthHSs Transect 1

-500

-400

-300

-200

-100

0

051015202530

Dept

h to

inte

rface

(m A

HD)

-500

-400

-300

-200

-100

0

Slot

ted

dept

h of

bor

es (m

AHD

)

seawater zone

freshwater zone

48B101C

36A

36B

49A48A

49B

101B

101A

61B

high salinity bores

-650

-550

-450

-350

-250

-150

-50

50

051015202530

Dep

th to

inte

rface

(m A

HD)

-650

-550

-450

-350

-250

-150

-50

50

Slo

tted

dept

h of

bor

es (m

AH

D)

seawater zone

freshwater zone

37A

38B

38A

46B47B46A

47A

53B

53A

Transect 2

high salinity bores

-250

-200

-150

-100

-50

0

50

051015202530

Distance from coast (km)

Dept

h to

inte

rface

(m A

HD)

-250

-200

-150

-100

-50

0

50

Slot

ted

dept

h of

bor

es (m

AH

D)

seawater zone

freshwater zone

49A

48B

48A

46B

47B

46A

47A

44A 45A

Transect 3

high salinity bores

49B

Figure 5-27 Theoretical seawater interface during two periods in 2004 (Ghyben-Herzberg relation) and slotted depths of bores in HSs and HSd as a function of distance from the coast along three transects (refer to map in Figure 5-26).

W E

W E

S N

Page 159: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

135

Comparison of the theoretical saltwater interface with the slotted depth of HSs bores suggests

that the shallow aquifer is not affected by seawater intrusion. Therefore, the high TDS and

seawater-type signature of bores 36B and 38B (Figure 5-25) must be accounted for by other

mechanisms. It was discussed above (Section 5.3.4) that there is evidence of vertical discharge

from the deep aquifer; the influence of saline coastal rainfall/sea spray are also plausible sources

of salinity. The hydrograph patterns for nested intervals at bores 36 and 38 (Figure 4-9, Chapter

4) indicated that in addition to recharge from the deep aquifer there is an enhanced signal in the

shallow aquifer due to direct rainfall recharge. Therefore, the hydrochemistry of HSs in these

coastal bores is a reflection of both coastal rainfall and groundwater from HSd. Two bores

located adjacent to bore 36B belong to the same hydrochemical facies but have a much lower

TDS (Figure 5-25). As these bores are located within a few hundred metres of the Herbert River

estuary, they are considered to be influenced by the chemistry of the river and the tidal pattern.

Evidence of this is provided by one of the bores which varied from a fresh Ca-Mg type at low

tide to a more saline Na = Cl type water at high tide.

5.5 NITROGEN IN GROUNDWATER

As discussed in Chapter 1 (1.4.1), nitrogen (N) is commonly found in groundwaters beneath

sugarcane growing areas, due largely to fertiliser inputs. Given that sugarcane farming is a

major land use in the study area and that all bores sampled were located in these cropping areas,

it is plausible that N is an important anion in the groundwater. This section explores the

concentration and speciation of N in the two aquifers in order to characterise the N signal of

groundwater. Subsequent chapters examine the N signal in surface waters and consider potential

N contributions from groundwater based on river-aquifer interactions. General principles

regarding nitrogen chemistry and transformation reactions were described in Section 5.1.1.4.

Nitrogen leaching to the subsurface has previously been investigated in the lower Herbert River

catchment, based on recharge estimates and measurements of nitrate in the upper part of the

shallow aquifer along a transect covering the three main geomorphological units (Bohl et al.,

2000b). The spatial distribution of N leaching losses was further investigated based on

pedological and hydrological features in a subcatchment of the Herbert River valley (Bohl et al.,

2001). These studies found that the highest leaching losses are expected to occur on the more

freely draining soils of the alluvial fans and the sand riverbank soils, while the lowest losses are

estimated for the heavy clay soils on the floodplain. While these previous studies yielded

estimates of N leaching potential, the estimates were based on limited spatial coverage of

measured N just below the watertable. Furthermore, potential leaching to deeper groundwater

was not considered as part of the mass balance. Since the time of these previous studies, more

detailed soil mapping of the Herbert valley has been undertaken (Wood et al., 2003). Seven

broad categories based on landscape position and formation are identified, as depicted in Figure

Page 160: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

136

5-28. It can be observed that clay soils dominate the central part of the study area, while alluvial

soils dominate the riverbank of the Herbert River. Based on this 2003 soil data and

understanding of the hydrogeological and hydrogeochemical characteristics, the spatial

distribution of N within each aquifer and partitioning of N between the aquifers are examined

below.

Streams

Terrace loamy soils

Hillslope soils

Sandy soils

Seymour soils

Clay soilsAlluvial soils

5 0 5 Km

N

Figure 5-28 Distribution of mapped soil types in the study area (Wood et al., 2003). Note that the Seymour soils have a high proportion of fine sand derived from acid volcanic parent material. Source: HRIC

5.5.1 Spatial distribution of N

The spatial distribution of total dissolved inorganic forms of nitrogen (DIN) is examined in this

section within the shallow and deep aquifers. Ammonium (NH4+), nitrite (NO2

-) and total

dissolved oxides of nitrogen (NOX) were analysed in all groundwater samples collected: nitrate

(NO3-) is calculated by difference from NOX and NO2

-. DIN is calculated as the sum of the

concentrations of dissolved NO3-, NH4

+ and NO2- (each as mg N/L).

5.5.1.1 Shallow aquifer

As shown in Figure 5-29a, the concentration of DIN is variable throughout the shallow aquifer.

Although the spatial distribution appears random, in general, bores recording the highest

concentrations of DIN are associated with Ca-Mg enriched groundwaters (Figure 5-17).

Comparison with the distribution of soil types shows that bores with DIN > 3 mg N/L are

located on either the alluvial or terrace loamy soils. In contrast, shallow bores with the lowest

Page 161: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

137

measured concentrations of DIN are mostly located on clay soils and are associated with Na

enriched groundwaters (Table 5-1). Note that these shallow groundwaters also plot within the

stability field for smectite clays (Section 5.3.2). The alluvial soils are generally sandy and

permeable, while the clay soils range from silty to heavy clays; terrace loamy soils are similar to

the clay soils although they are situated higher in the landscape and have better structure and

drainage (Wood et al., 2003). The alluvial soils and sandy terrace loams are considered to have

a high risk of leaching due to their well draining nature. Conversely, nitrogen losses by

denitrification due to intermittent waterlogging (anaerobic conditions) are characteristic of the

clay dominated soils (Wood et al., 2003). Therefore, given the soil hydraulic properties and

spatial distribution of DIN in HSs, it is considered that soil type has a major influence on the

amount of N leached to the shallow aquifer. This is consistent with previous studies (Bohl et al.,

2001).

5.5.1.2 Deep aquifer

Figure 5-29b shows that there are three distinct zones in the deep aquifer in regards to the

concentration of DIN. The upper part of the Herbert River valley has Ca-Mg enriched

groundwaters with the highest concentration of DIN, as also shown in Figure 5-15. Further

down-valley, in the middle section of the catchment, groundwater has very little DIN. Whilst

good vertical connection with HSs is evident in some bores, including shallow bores with high

DIN, the near-absence of DIN at depth indicates N loss due to transformations to other species

of nitrogen. For example, denitrification results in the formation of nitrous oxides and nitrogen

gas (Section 5.1.1.4), which were not measured in the samples collected.

Towards the coast, the concentration of DIN is observed to increase in the deep aquifer,

corresponding with Cl-dominated groundwaters. Similar to the TDS distribution (Figure 5-23b),

the highest concentration of DIN in this area is observed in the northeast at bore 49 (0.81 mg

N/L) and progressively declines in groundwater to south. This suggests there may be a

hydrochemical control on the concentration and spatial distribution of DIN in the deep aquifer.

Page 162: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

138

a

b

Figure 5-29 Spatial distribution of dissolved inorganic nitrogen (DIN) (mg N/L) in bores screened in the (a) shallow and (b) deep aquifers in June 2005. A and B refer to different screened intervals considered to be part of the same aquifer at that bore location. The shaded areas in (a) correspond to the major water types identified in Section 5.2.1 (refer to Figure 5-15 and Figure 5-17).

Page 163: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

139

5.5.2 Speciation of N

Whilst the concentrations of the three inorganic nitrogen forms were determined, results

indicate that NO2- is below detection (< 0.002 mgN/L) in groundwaters from HSd and just above

the detection limit in some samples from HSs. Therefore, hydrochemical trends regarding the

other two forms of DIN (NO3- and NH4

+) are the focus of this section.

5.5.2.1 Shallow aquifer

As illustrated in Figure 5-30a, the concentration of NO3- in the shallow aquifer ranges from

below detection up to almost 90 mg/L. In contrast, NH4+ is a relatively minor component of

HSs; the majority of bores have NH4+ concentrations of less than 0.1 mg/L (Figure 5-30b).

Therefore the distribution of DIN observed in Figure 5-29a is largely due to NO3-, which is the

dominant species of N in the shallow aquifer. In contrast to nitrate, the ammonium form of

nitrogen is less mobile in the subsurface environment, and is therefore less likely to be

transported through the unsaturated zone into groundwater. Key processes that inhibit this

transport are adsorption, cation exchange, incorporation into microbial biomass or release to the

environment as a gas. Adsorption onto soil particles is considered to be the major mechanism of

removal of NH4+ in the subsurface environment (Canter, 1997).

In general, an inverse relationship between the two dominant forms of inorganic N is observed

in HSs, such that high NO3- waters have a low concentration of NH4

+ and high NH4+ waters have

low NO3- (compare Figure 5-30a and b). The redox control on the concentration of NO3

- and

NH4+ in the shallow aquifer is clearly illustrated in Figure 5-31: high NO3

- is associated with

more oxidised waters (high Eh), whilst more reduced waters (low Eh) are dominated by NH4+.

These trends are consistent with the respective oxidation states of the nitrate ion (+5) versus that

of the ammonium ion (-3). Where there is sufficient oxygen, an inorganic carbon source, and

specific chemoautotrophic bacteria, biological oxidation (nitrification) of NH4+ produces NO3

-

that is leached to shallow groundwater (equation 5-1, Section 5.1.1.4). In contrast, low NO3-

groundwater is consistent with only minor leaching of NO3- and/or denitrification due to

reduced conditions. Therefore, while soil type influences the amount of DIN leached to the

shallow aquifer (as discussed in Section 5.6.1.1), the redox condition of the groundwater is a

major control on the speciation of N in HSs. Speciation of N is also dependent on the pH

(Appelo and Postma, 1994); however, over the limited pH range of HSs waters (5 - 7) there is no

obvious relationship.

Page 164: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

140

#S

#S

#S#S

#S#S

#S

#S#S

#S

#S

#S#S

#S

#S

#S

#S

#S

#S#S#S

#S

#S

#S

#S

#S#S

#S#S

#S

#S#S

#S

#S

#S#S

#S

#S

#S

#S

#S

#S#S#S

#S

#S

#S

0.03

8.23

0.330.01

A: 2.21

4.020.01

0.31

8.452.57

0.01

27.44

40.06

46.46

38.51

15.71

B: 20.14

87.8715.43

19.26

22.35< 0.01

< 0.01

< 0.01

< 0.3#S0.3 - 4#S

15 - 20#S4 - 15#S

NO3 (mg/L)

5 0 5 Km

N

HSs NO3

#S >20

a

%U

%U

%U%U

%U%U

%U

%U%U

%U

%U

%U%U

%U

%U

%U

%U

%U

%U%U%U

%U

%U

%U

%U

%U

%U

%U%U

%U

%U%U

%U

%U

%U

%U

%U

%U%U

%U

%U

%U

0.34

1.42

0.23

0.020.09

0.010.08

0.01

A: 0.01

0.04

0.02

0.01

0.010.20

0.01

0.47

0.03

< 0.01

< 0.01

< 0.01< 0.01

< 0.01

< 0.01

B: < 0.01

0 - 0.01%U0.01 - 0.1%U

0.5 - 1.5%U0.1 - 0.5%U

NH4 (mg/L)

5 0 5 Km

N

HSs NH4

b

Figure 5-30 Spatial distribution of (a) NO3- (mg/L) and (b) NH4

+ (mg/L) in bores screened in the shallow aquifer in June 2005, with selected flowlines depicted. A and B refer to the two screened intervals considered to be part of the same aquifer at bore 54.

Page 165: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

141

0.0001 0.001 0.01 0.1 1. 10.100

200

300

400

500Eh (mV)

NO3 (meq/l)0.0001 0.001 0.01 0.1

100

200

300

400

500Eh (mV)

NH4 (meq/l)

HSs samples

Na = ClNa > ClCa-Mg enrichedNa enriched

Figure 5-31 Bivariate plots for shallow groundwater samples (2004 samples only), displayed according to the compositional groups determined in Section 5.2.1.

It was noted in Chapter 4 (Section 4.2.2) that there are two sandy units within the shallow

aquifer of varying lateral extent. Whilst the two units display similar hydraulic behaviour, they

are distinct hydrochemically (Section 5.3.3), including with respect to N. Thus, bores screened

in the upper unit, associated with oxidised Ca-Mg enriched waters, are observed to have a high

concentration of NO3- (Figure 5-31). In contrast, groundwater in the deeper unit of HSs,

associated with more reduced Na enriched waters, generally has less than 1 mg/L NO3- and a

detectable concentration of NH4+ (Figure 5-30). Given the high mobility of NO3

- and high

vertical connectivity between the two sandy units in the shallow aquifer (Section 5.3.3),

leaching of NO3- to the deeper unit of HSs would be expected. The near-absence of NO3

- and

NO2- in this deeper unit is consistent with reduction and ultimate denitrification to nitrogen gas,

N2. Vertical movement of N is discussed further in Section 5.6.3.

Page 166: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

142

5.5.2.2 Deep aquifer

As shown in Figure 5-32, N speciation in the deep aquifer is characterised by NO3- in the upland

area to the northwest and NH4+ towards the east. The same trend is noted during other sampling

periods (May and October 2004). In the intervening zone, the concentrations of both NO3- and

NH4+ are low, with no particular dominant species. Therefore, the observed spatial pattern of

DIN (Figure 5-29b) is due to different N species dominating in different parts of the catchment.

The spatial extent of NO3- and NH4

+ enriched waters are associated with Ca-Mg and Na-Cl type

waters respectively (compare Figure 5-32 with Figure 5-29b). Therefore, as suggested in

Section 5.6.1.2, a change in hydrochemical conditions is a plausible explanation for the

speciation difference up-valley compared to on the floodplain.

As for the shallow aquifer, high NO3- concentrations in the northwest are consistent with

oxidising conditions (nitrification) and hence leaching of NO3- to the deep aquifer. Given the

general direction of groundwater flow in HSd (Figure 4-14, Chapter 4), the speciation trends

down gradient of the northwestern recharge area are consistent with reduction of N. Increases in

HCO3 and pH in the intervening zone (low NO3- and NH4

+) indicate reduction of nitrate by

organic matter (denitrification), while elevated concentrations of ferrous iron and sulfate in HSd

in the eastern area (low NO3- and higher NH4

+) are consistent with reduction of N coupled with

pyrite oxidation or reduction by other Fe2+ bearing minerals (Appelo and Postma, 1994).

Nitrogen gas is a product of these reduction processes, which was not specifically measured in

the groundwater samples. Therefore, at least in the middle of the catchment, it is plausible that

DIN is present as N2, which remains in solution until the groundwater discharges to a surface

water body and equilibrates with the atmosphere (Thayalakumaran et al., 2004). Given that N2

is the end product of denitrification, an additional source of N to HSd is required to account for

the observed increase in NH4+ in the lower part of the catchment. It was established above that

seawater intrusion influences the hydrochemistry of HSd in this area. Based on the

hydrochemistry of seawater at Moreton Bay, where the concentration of NO3- was measured at

50 mg/L (Cresswell 2006, pers. comm.), it is therefore plausible that seawater provides the

additional source of N. Dissimilatory nitrate reduction to ammonium (DNRA) (Silver et al.,

2001; Korom, 1992) may thus have a role in the transformation of N. Alternatively, an increase

in ammonium could be attributed to desorption from clays under reduced conditions.

Verification of the factors or processes responsible for the observed speciation/concentration

trends of DIN is beyond the scope of this research. However, it is noted that measurement of

gaseous forms of N; redox indicators; microbiology; and availability of electron acceptors and

donors, could assist in future investigations. Nitrogen isotopes can also provide insight into N

processes (Kendall, 1998).

Page 167: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

143

#S

#S

#S#S

#S

#S#S#S

#S

#S

#S#S

#S

#S

#S#S

#S #S#S

#S

#S

#S#S

#S

#S

#S

#S

#S

#S

#S

#S#S

#S

#S

#S#S

#S #S#S

#S#S

#S

A: < 0.01

0.05

0.040.02

0.04

0.01

0.010.02 0.01

0.02

5.84

0.030.010.05

0.040.02

0.04

0.01

0.030.01

0.020.02

5.84

24.3911.24

< 0.01 < 0.01

< 0.01

< 0.01

< 0.01

< 0.01

< 0.01B: < 0.01< 0.01

< 0.01< 0.1#S

0.1 - 10#S

> 20#S

10 - 20#S

NO3 (mg/L)

5 0 5 Km

N

HSd NO3

a

%U

%U

%U%U

%U

%U%U%U

%U

%U

%U%U

%U

%U

%U%U

%U %U%U

%U

%U

%U%U

%U

%U

%U

%U

%U

%U

%U%U

%U

%U%U%U

%U

%U

%U%U

%U

%U

%U%U

%U %U%U

%U

%U

%U

%U

%U

%U

0.24

0.14

0.26

1.04

0.39

A: 0.11B: 0.90

0.110.01

0.070.02

0.02

0.02

0.020.17

0.03 0.22

0.08

0.14

0.06

0.01

0.07

0.10

0.12

< 0.01

< 0.01

< 0.01%U

0.01 - 0.1%U

0.5 - 1.5%U

0.1 - 0.5%U

NH4 (mg/L)

5 0 5 Km

N

HSd NH4

b

Figure 5-32 Spatial distribution of (a) NO3- (mg/L) and (b) NH4

+ (mg/L) in bores screened in the deep aquifer in June 2005. A and B refer to different screened intervals considered to be part of the same aquifer at that bore location. A selected flowline is depicted.

Page 168: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

144

5.5.3 Nitrogen transport

As discussed in Chapter 3 (Section 3.2.1), fertiliser applied to sugarcane is considered to be a

major source of DIN in groundwaters of the study area. Local hot spots of DIN in bores

adjacent to garden beds also provide evidence of point source contributions. Urea ([NH2]2CO)

and/or ammonium sulfate (NH4SO4) are commonly applied crop fertilisers in the study area,

which are mineralised by microorganisms in the soil to NH4+ (Wood et al., 2003). As depicted

in the nitrogen cycle (Figure 5-1), NH4+ has a number of fates: volatilisation, adsorption onto

clay minerals or nitrification to NO3-. In the absence of N measurements in soils of the study

area, it is not possible to determine the proportion of NH4+ bound to soil particles versus

atmospheric losses or transformations to other forms such as NO3-. However, it is sufficient to

note that consistently high concentrations of NO3- are observed in parts of the shallow and deep

aquifer during different sampling periods. Therefore, there is evidence of N leaching to the

groundwater system. It is beyond the scope of this study to carry out a detailed mass balance for

the inputs and outputs of N within and between aquifers and hence to resolve a component

(aquifer storage) of the nitrogen budget framework described in Chapter 2 (Figure 2-5).

However, in light of the recharge-discharge characteristics established in Chapter 4 (Section

4.5) and the vertical and spatial relationships established in Sections 5.3 and 5.4, the following

analysis considers the implications of the hydrogeological characterisation for the transport of N

to groundwaters and potentially to surface waters. Figure 5-33 summarises these concepts.

Visual comparisons between Figure 5-29, Figure 5-30 and Figure 5-32 form the basis of the

following interpretation.

5.5.3.1 Shallow aquifer

It was established in Chapter 4 that recharge to the shallow aquifer is from both lateral and

diffuse sources (Figure 4-17b). Therefore, given the widespread fertiliser source of DIN in the

landscape, the main hydrogeologic control on the distribution of N in HSs is the aquifer

composition and hence redox condition. As discussed above (Section 5.6.2), oxygenated

groundwaters associated with sandy aquifer material result in the observed high NO3-

concentrations, while reducing groundwaters (associated with a higher clay content) result in

low DIN, with NH4+ as the dominant species. Conversion of NO3

- to N2 by denitrification may

also occur where reducing conditions are encountered. Due to the upwards head gradient and

strong vertical connection between HSs and HSd at the coastal bores, the distribution of N in the

shallow aquifer is also influenced by the composition of the deep aquifer. For example, given

that HSd discharges vertically to HSs towards the coast (Section 5.3.4), a component of DIN in

the shallow aquifer may also be from the deep aquifer. Of direct relevance to this study is that

high NO3- groundwaters are found adjacent to the Herbert River: the location of these high DIN

waters has potential implications for the transport of nitrogen to and from surface waters, as

discussed further in Chapters 6 and 7. It was also established in Chapter 4 that in addition to

Page 169: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical Framework

145

discharge to the Herbert River, discharge from the shallow aquifer occurs out to sea. Therefore,

based on the available measurements at the coastal bores, there is the potential for relatively low

concentrations of DIN (present as NH4+) to contribute directly from the shallow aquifer to the

offshore marine environment.

5.5.3.2 Deep aquifer

The highest concentrations of DIN are found in the northwest of the catchment, which coincides

with the main recharge area for the deep aquifer. Given that lateral recharge is the dominant

supply of water to HSd, N sourced from this area is a major contributor to the deep aquifer.

Although the shallow aquifer has a high concentration of NO3-, and strong/good vertical

connectivity between HSs and HSd is maintained away from the recharge zone, there is a

dramatic decrease in DIN in the deep aquifer along the flowpath. Therefore, despite there being

multiple sources of DIN (lateral and vertical) and appropriate hydraulics to enable leaching to

occur, the observed concentrations in the deep aquifer are indicative of denitrification arising

from more reducing conditions. Further measurements would be required to verify whether the

decrease in DIN is matched by an increase in the concentration of N2 (Bohlke and Denver,

1995). Alternatively, it is plausible that DIN concentrations are lower in the deep aquifer

compared to the shallow due to HSd groundwaters being older and recharged prior to more

recent anthropogenic inputs.

The potential contribution of DIN from groundwater to surface waters is of particular interest to

this study. Whilst river-aquifer relationships in relation to N are explicitly examined in Chapters

6 and 7, it is noted that high NO3- groundwaters are observed at the beginning of a preferential

pathway that runs parallel to the lower Herbert River. In addition, it was proposed in Chapter 4

that the potential for groundwater discharge from the deep aquifer to the river exists in the upper

reaches of the valley. Therefore, a contribution of DIN from deep groundwater to the river is

plausible. Furthermore, given that a component of lateral discharge from the deep aquifer is out

to sea (Section 4.5.1), there exists an additional pathway for the transport of DIN to surface

waters. In addition to diffuse groundwater discharge to the ocean, point discharge of

groundwater from confined aquifer systems, several kilometres offshore, has been described

within the Great Barrier Reef region (Stieglitz and Ridd, 2000). It has been suggested that a

small net flux of submarine groundwater discharge can deliver a comparatively large flux of

nutrients to sea. Therefore, diffuse and/or point discharge mechanisms may have ecological

significance for the delivery of nutrients to the intertidal zone or inner/mid shelf of the GBR

(Stieglitz, 2005).

Page 170: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Figure 5-33 Conceptual diagram summarising the movement of water and N in the alluvial aquifer system and potentially to the Herbert River (HR)

Page 171: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Hydrogeochemical framework

147

5.6 CHAPTER SUMMARY

A range of hydrogeochemical data have been analysed in this chapter in order to extend the

conceptual model for the hydrogeology in the alluvial aquifer system of the lower Herbert River

catchment. Specifically, hydrochemical analyses have been used to verify key attributes and

processes relating to the distinction of the main aquifers; the degree of interaction between

aquifers; and the spatial relationships within the aquifers. Based on this enhanced understanding

of the hydrogeological framework, the distribution of N in groundwaters of the two aquifers and

the potential implications for contributions to surface waters have also been assessed.

HSs groundwaters in the upper and middle sections of the study area are characterised by Na-

HCO3-Cl facies, with Ca-Mg enriched and Na enriched groups related to clay content and hence

aquifer lithology. Whilst sandy units within the shallow aquifer are chemically distinct, they

nonetheless share a common recharge source and are considered to represent the same aquifer at

the scale of interest. The dominance of HCO3 is consistent with recharge from proximity to a

rainfall-recharge zone: recharge is rapid and, in general, not associated with evaporation prior to

recharge. Na-Cl ± HCO3 facies groundwaters are found in the lower (eastern) section of the

catchment, with high and low salinity groups observed. The highest salinities are associated

with contributions from the deep aquifer, either from upward vertical recharge at the coast or

contamination during pumping. Lower salinity trends are indicative of minor hydrochemical

evolution in some areas and mixing with water derived from the Herbert River in the tidal zone.

The absence of strong evidence for hydrochemical evolution is consistent with relatively short

groundwater residence times and therefore less opportunity for water-rock interactions.

The spatial separation of water types in the deep aquifer is more distinct than in the shallow

aquifer. The main recharge area in the northwest is characterised by Na-HCO3 facies

groundwaters, whilst the central part of the study area is dominated by Na-HCO3-Cl facies. The

lower section of the catchment is represented by Na-Cl facies waters with high salinities. Whilst

there is evidence of hydrochemical evolution in HSd, lateral development is interrupted by

enhanced vertical leakage from the shallow aquifer and contributions of higher salinity waters

from the Stone River valley. Therefore, whilst there is overall a general increase in solute

concentrations and TDS down-flow of the main recharge area, the degree of hydrochemical

evolution is suppressed because of the vertical connectivity between aquifers and the

convergence of flowpaths contributing groundwater of different compositions. In the eastern

part of the catchment, the hydrochemistry of the deep aquifer is influenced by seawater from

past and/or present-day intrusion. A preferential pathway for intrusion in the northeast, rather

than the eastern coastline is suggested. Hydrochemical evidence supports local vertical

groundwater discharge at the coast, where seawater is effectively forced back by freshwater.

Page 172: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 5

148

The analyses have highlighted the role of hydrochemistry to infer relative degrees of inter-

aquifer mixing. While bore hydrographs indicate the direction of vertical head gradient between

aquifers and the potential degree of connectivity, the hydrochemistry provides evidence of

actual exchange of water and the relative extent. On balance, the hydrochemistry supports the

conceptualisation of the subsurface into a two-aquifer system, with there being spatial

variability in hydraulic connectivity between aquifers. Geochemical modelling, which is beyond

the scope of this research, would allow the degree of inter-aquifer mixing to be quantified.

Connectivity relationships between the aquifers are important for understanding the partitioning

of N in groundwater. In general, concentrations of DIN are greatest in the shallow aquifer, with

the distribution and speciation influenced by aquifer composition, including soil type and redox

state. Nitrification, associated with sandy soils and oxidising conditions, produces NO3- that is

readily leached to HSs. In contrast, clay soils, prone to water-logging, result in denitrification

and hence low DIN groundwaters. High DIN in HSd is restricted to the main recharge area in the

northwest, which is the main source area for N leached to the deep aquifer. Despite good

vertical connectivity between the aquifers in some areas, there is strong partitioning of DIN due

to the redox control on the fate of N in groundwater. Thus, whilst oxidising conditions in the

main recharge area favour NO3-, a dramatic decline in DIN is evident away from the recharge

zone due to N reduction processes. Mass balance calculations, including measurements of other

forms of N, would be required to verify the proposed mechanisms for the observed speciation of

DIN in both aquifers. The application of nitrogen isotopes is also an area for future research.

Whilst it is beyond the scope of the thesis to undertake a detailed mass balance for the inputs

and outputs of DIN within and between aquifers, the hydrochemical analyses provide evidence

of N leaching to the groundwater system. Furthermore, the spatial distribution of DIN in each

aquifer allows the potential risk of N transport from groundwater to surface waters to be

considered. In particular, the observed high concentration of DIN in both aquifers in locations

adjacent to the Herbert River indicates the potential for N in groundwater to contribute to the

river system or vice versa. Given the ultimate discharge of groundwater from both aquifers to

the sea, there is an additional pathway for the movement of N offshore. The following two

chapters explicitly examine the connectivity between groundwater and the Herbert River, from

both a physical (Chapter 6) and chemical (Chapter 7) perspective. Therefore, while this chapter

has highlighted areas of high DIN in the alluvial aquifers, subsequent chapters examine (i)

whether exchange of water from groundwater to surface water is a plausible mechanism and (ii)

whether there is hydrochemical evidence of N transport to the Herbert River via this

mechanism.

Page 173: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

149

Chapter 6 Physical River-Groundwater Interactions

6.1 INTRODUCTION

A framework for subsurface water movement in the lower Herbert River catchment was

developed in Chapters 4 and 5 through interpretation of physical and chemical hydrogeological

datasets. The following two chapters incorporate surface water into the conceptual model by

considering river-groundwater interactions. Resource and Environmental Management (REM,

2002) devised a classification system for defining the nature of stream-aquifer interactions in

order to assist with targeting management in areas where these interactions are important.

Although the classification system was originally proposed for use in the Murray-Darling Basin,

where the issues are largely concerned with managing conjunctive water use, the principles are

also relevant to the current study, in which the nature of these interactions can have important

water quality implications (refer to introductory chapters 1-3).

Table 6–1 Classification system for stream-aquifer interactions relevant to conjunctive use management (REM, 2002).

(1) Hydraulic connection (2) Stream-aquifer interaction process

(3) Potential impacts of poor quality groundwater on

surface water quality

connected gaining stream high

connected losing stream no impact

connected variable gaining/losing stream low

disconnected losing stream no impact

As shown in Table 6-1, the three levels of classification are aimed at distinguishing whether (1)

there is hydraulic connection; (2) the direction of interaction; and (3) the likely impact of

groundwater on stream water quality. This chapter specifically addresses (1) through

comparisons of groundwater elevation to the topography (bed and floodplain) of the Herbert

River, while (2) is assessed through groundwater elevation-river stage relationships. These

techniques represent qualitative hydrometric approaches (Section 2.4.1.2). In addition, various

hydrological methods (e.g. Section 2.4.1.3) are applied to temporal streamflow data to provide

further insight into the direction of flux between the aquifers and the river. Whilst the aim of

Page 174: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 6

150

this chapter is to characterise the physical hydraulic relationships between groundwater and

surface water along the lower Herbert River, Chapter 7 considers chemical interactions and

hence addresses classification (3).

6.1.1 Data availability and preparation

Numerous gauging stations have been installed in the Herbert River catchment to collect flow

and/or water quality data. Two flow gauges are currently operational along the lower Herbert

River (Figure 6-1). Although the lower catchment is the focus of this section, measurements

from gauge 116004, immediately upstream of the Herbert River gorge, are referred to where

appropriate. The length of historical data available at each of the stream gauges is variable: a

long time series of flow and stage data (from 1915) is available at gauge 116001. Water quality

data is also available at some gauges during part of the gauging period (analysed in Chapter 7).

A summary of gauge characteristics is provided in Table 6-2.

%U

%U

%U

116004

116006

116001

stream gauges%U

lower Herbert Riverstreams

20 0 20 kmN

#

Herbert RGorge

#S

Figure 6-1 Selected QDNRW stream gauges along the lower (116006, 116001) and upper (116004) Herbert River which are referred to in the following text.

Page 175: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Physical River-Groundwater Interactions

151

In order to compare groundwater and river heights, the data were converted to a common

datum, in this case the Australian Height Datum (AHD). River stage data provided by QDNRW

were corrected for the gauge zero value and then a general correction factor (advised by

hydrographers at QDNRW) was applied to convert the stage height data from the State Datum

to AHD. Groundwater elevations (in m AHD) were derived by subtracting groundwater levels

(depth from reference point) from the elevation of the bore reference point (in m AHD) (as

discussed in Section 4.4.1).

Table 6–2 Features of the QDNRW stream gauging stations in the lower catchment and selected upper catchment gauges.

Stationa Streamb Start date End date Typec

116001 (A-E) L Herbert River 1/8/1915 open F, WQ

116006 (A-B) L Herbert River 2/2/1968 open F, WQ

116004 (A-C) U Herbert River 31/5/1922 open F, WQ

a Letters attached to the station number after ‘A’ correspond to minor shifts in the gauge position over time: the start date is given for the first gauge to be installed and end date for the last b L and U refer to Lower and Upper extents of the streams c F and WQ refer to Flow and Water Quality (WQ is available for at least part of the gauging period).

Surveyed river cross-sections were also provided by the Hinchinbrook Shire Council at a

number of sites along the lower Herbert River. The lowest point of the riverbed and the

maximum heights of the left and right banks (all in m AHD) were recorded from the most recent

cross-section available at each site. The distance from the river outlet of each survey location

was also determined from a GIS. As can be observed in Figure 6-2a, there is a close distribution

of surveyed cross-sections for approximately 60 km upstream from the river mouth. Within this

stretch of river there is a stream gauge (116001) at around 22 km upstream; at around 70 km

from the river mouth there is an additional stream gauge (116006). Both of these gauges have

stage height records in metres relative to AHD. In the absence of extensive gauge data along the

entire length of the lower Herbert River, stage height is calculated at locations away from the

gauges by adding the measured water column at the sampling sites to the riverbed depth. Note

that measurement of the water column allows for changes in river stage due to variation in the

width of the river along its length but not to variations in height with discharge (Figure 6-2b).

Where sampling sites and cross-section locations are not coincident, the riverbed depth is

derived from the average between adjacent (upstream and downstream) surveyed cross-sections.

Table 6-3 shows that there is reasonably good agreement (within 1 m) between derived and

gauged stage heights at the two QDNRW stream gauges on the same day of measurement. This

provides confidence in using derived river stage heights (where necessary) to compare with

groundwater elevations, particularly during the dry season. This approach is examined in

Section 6.3.

Page 176: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 6

152

-10

0

10

20

30

0 10 20 30 40 50 60 70 80 90 100Distance from river mouth (km)

Sur

veye

d riv

erbe

d (m

AH

D)

a

0

50

100

150

200

250

0 10 20 30 40 50 60 70 80 90 100Distance from river mouth (km)

Rive

r w

idth

(m)

116006116001

b

Figure 6-2 (a) Surveyed riverbed and (b) estimated river width (June 2005) as a function of distance from the mouth of the Herbert River. The location of the two stream gauges along the river is also indicated. Source: Hinchinbrook Shire Council (surveyed riverbed).

Table 6–3 Comparison between gauged (QDNRW) and derived stage heights on the same day of measurement during the dry season.

Date Gauge Gauged (m AHD)

Derived (m AHD)

Difference (m AHD)

26/10/2004 116001 1.4 0.6 0.8

3/06/2005 116001 1.6 1.8 -0.2

2/11/2004 116006 16.6 16.0 0.6

8/06/2005 116006 17.0 16.1 0.9

Page 177: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Physical River-Groundwater Interactions

153

Time series stage heights are derived at locations away from the gauges by adjusting the gauged

data by the difference between the gauged and derived stage (away from the gauge) at the same

period in time. As water column measurements were recorded during two dry season months

(early and late in the season), an average value is used. For example, the surveyed depth of the

riverbed at Trebonne (located approximately 7 km upstream of gauge 116001) is 2.7 m AHD,

while the measured water column in October 2004 (late dry season) was 0.8 m. Therefore, stage

height at Trebonne is estimated as the summation i.e. 3.5 m AHD. The recorded stage height at

the gauge on the same day was 1.36 m AHD; hence, the difference between the gauged and

calculated stage is approximately 2 m. A similar difference is obtained in June 2005 (early dry

season). This translates to a 2 m vertical shift upwards of historical stage heights at gauge

116001 to derive the time series at Trebonne (Figure 6-3). Note that this approach assumes that

a constant factor can be applied to the gauge data for the entire record based on shifts in stage

heights at Trebonne during the dry season. While this assumption is plausible during the dry

season due to the low variability in river stage, there is greater uncertainty in the derived stage

heights during high flow periods.

0

5

10

15

20

1975

1977

1979

1981

1983

1985

1987

1989

1991

1993

1995

1997

1999

2001

2003

2005

m A

HD

Stage @gauge116001

Stage @Trebonne

HR @Trebonne

Riverbed @Trebonne

2 m vertical shift

Figure 6-3 Derived historical stage height (m AHD) in the Herbert River at Trebonne. Circles represent the calculated river stage in October 2004 and June 2005 based on surveyed riverbed and measured water column depth, which were used in conjunction with gauge 116001 values to derive the river stage time series at Trebonne.

Page 178: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 6

154

6.2 POTENTIAL FOR HYDRAULIC CONNECTION

As outlined in Chapter 2 (Section 2.2.1), a connected river-aquifer system arises where there is

direct contact between a stream reach and an underlying aquifer via a zone of saturated material

or a narrow unsaturated zone (Bouwer and Maddock, 1997). The potential for hydraulic

connection was discussed briefly in Chapter 4 (Section 4.2.2.1) based on the construction of

lithologic cross-sections and river profiles, which illustrated that in the upper part of the valley

the Herbert River intersects the sediment comprising the deep aquifer, while further downvalley

the river incises the shallow aquifer only (Figure 4-4). Hydraulic connection is further assessed

in this section by comparing groundwater elevations with stream topography: connection is

assumed to potentially exist where the groundwater elevation lies within the elevation of the

channel, defined as between the bed and bank of the river. Given the availability of surveyed

river profiles and the relatively small number of bores along the Herbert River, site-based

comparisons are considered to be feasible. In order to simplify the analysis, the lower Herbert

River is divided into four distinct reaches: (A) from the river mouth to gauge 116001; (B) from

gauge 116001 upstream to the junction with the Stone River; (C) upstream of the Stone River

junction to Long Pocket; and (D) from Long Pocket to Abergowrie, towards the western extent

of the lower catchment (Figure 6-4 and Figure 6-5). These reaches are chosen because of the

locations of the stream gauges along the lower Herbert River and consideration of the tidal zone

(downstream of gauge 116001) and bore distribution. Bores selected for the analysis are located

within 3 km of the Herbert River (at the closest point) and correspond to one of the four river

reaches (Figure 6-5).

-10

0

10

20

30

40

50

0 10 20 30 40 50 60 70 80 90

Distance from river mouth (km)

m A

HD

River bed

Max elevation left bank

Max elevation right bank

Figure 6-4 Surveyed topographic features of the lower Herbert River channel in m AHD. Vertical lines mark the division of the lower Herbert River into the four reaches (labelled). Note that left and right banks are defined from the perspective of looking downstream.

A

B

C

D

Page 179: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Physical River-Groundwater Interactions

155

%U

%U

%U

%U

%U%U

%U

%U

#

#S#S

#S

#S#S

#S

#S#S

#S

#S

#S

#S

#S

#S

#S

#S#S

%U

#S#S

#S #S

#S

#S

#S#S #S

#S

#S#S #S #S

#S

#S

#S

#S

#S

#S

#S

#S

$T$T$T$T

$T$T

$T

$T$T

$T

$T

$T$T$T

$T

$T$T$T$T

$T

$T$T

$T

Halifax

Trebonne

Lannercost

#

Abergowrie bridge

Timrith

Ingham

#

Gairloch bridge

99025 99047

36

48

51

52

59

61

62

67

68

69

7071

%U

McNade

58

53

49

#

Gedges Xing

River mouth

Abergowrie

LongpocketD

C

B

A

Stone Riverjunction 116001

116006

Stream gauges%UWater column#S

River x-sections$TBores#S

StreamsPlaces%U

5 0 5 KmN

Ripple Ck

Ston

e R

Herbert R

Figure 6-5 Map showing geographical features which relate to the analyses of groundwater elevations and river topography/stage along the lower Herbert River. Bores are labelled without their 11600 prefix. Also indicated are locations of water column measurements and where river cross-sections are available. The four river reaches are defined as: (A) river mouth to gauge 116001; (B) gauge 116001 to Stone River junction; (C) Stone River junction to Long Pocket; and (D) Long Pocket to Abergowrie. Note that reach A is tidal.

An example of the approach used to compare historical groundwater elevations to the bed and

bank height of the Herbert River is shown in Figure 6-6 for bore 11600048. A similar approach

is applied to each of the bores within the four reaches; where the bore and cross-section

locations do not coincide, an average of bank heights (on the relevant side of the river) and

riverbed measurements from the nearest upstream and downstream cross-sections are used. Note

that the relevant bank is that which is on the same side of the river as the bore, where the bank is

defined from the perspective of looking downstream. Thus, for bore 11600048, the height of the

right bank is applicable, while for bore 11600061, the left bank is appropriate for comparison

with groundwater elevation (Figure 6-5). The analyses are performed on both shallow and deep

groundwaters because of the vertical connectivity between HSs and HSd in some parts of the

catchment and incision of the river channel into both aquifers (Figure 4-4, Chapter 4).

The analyses indicate that throughout the historical record, groundwater elevations for the

shallow and deep aquifer are within the corresponding elevation of the river channel within each

of the four reaches i.e. above the elevation of the base of the stream and below the level of the

stream bank at each comparison site. Based on the available information this indicates that there

is potential for stream-aquifer connectivity along the entire length of the lower Herbert River in

both wet and dry season months. According to the framework for classifying stream-aquifer

interactions (Table 6-1) the next step is to establish the potential direction of flux.

Page 180: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 6

156

-5

0

5

10

1976

1978

1980

1982

1984

1986

1988

1990

1992

1994

1996

1998

2000

2002

2004

m A

HD

11600048B

Riverbed

Riverbank

Stre

am c

hann

el

Figure 6-6 Comparison of time series groundwater elevation at bore 1160048 (reach A) with the corresponding surveyed streambed and bank height in the Herbert River.

6.3 DIRECTION OF INTERACTION

Having established the potential for river-aquifer connectivity, the direction of flux between the

aquifer and the stream remains to be ascertained in order to characterise the interaction process

(Table 6-1). In Chapter 4, the shape and upstream orientation of groundwater elevation contours

where they cross the river suggested a potential gaining stream system over a large section of

the lower Herbert River during selected wet and dry periods (Figure 4-14 and Figure 4-16).

Detailed comparisons of groundwater elevation and river stage throughout the historical record

provide further insight into stream-aquifer relationships. Whilst height comparisons provide an

indication of the potential direction of interaction at a small scale, examination of flow

characteristics provides evidence of actual volumetric flux over a broad area. Therefore,

hydrological approaches such as streamflow analysis, flow duration curves and hydrograph

separation are also examined.

6.3.1 Groundwater elevation - river stage relationships

If the groundwater elevation is above the elevation of the stream, groundwater potentially

moves into the stream, which is termed gaining; the converse gives rise to a losing stream

situation (Section 2.2.1, Chapter 2). This is the underlying principle behind the following

analyses. Although the catchment has only two stream gauges, river stage is estimated at

intervening sites with reference to the measured water column and surveyed depths to the

riverbed (as discussed in Section 6.1.1). To aid with the analysis, the river is divided into the

four reaches defined in Figure 6-4 and Figure 6-5. A list of bores and the relevant comparison

Page 181: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Physical River-Groundwater Interactions

157

stream gauge at the closest sites along each reach is provided in Table 6-4 (also refer to Figure

6-5). An explanation of features specific to the analyses along each reach is provided below,

with examples illustrating the approach used to compare groundwater elevations and river stage.

Table 6–4 Bores and comparison stream gauges at sites along the four river reaches defined as: (A) river mouth to gauge 116001; (B) gauge 116001 to Stone River junction; (C) Stone River junction to Long Pocket; and (D) Long Pocket to Abergowrie (refer to Figure 6-5).

Reach Bore number1 Stream distance2 (km)

Comparison stream gauge

A

101A, 101C3

48A, 48B, 49A, 49B

36A, 36B

52A, 53A, 53B

0.82

0.15; 2.6

1.8

0.73; 2.7

adjusted4 gauge 116001 at Ripple Ck

adjusted gauge 116001 at McNade

adjusted gauge 116001 at Halifax

adjusted gauge 116001 at Gairloch

B

61A, 61B, 62A, 62B

99025, 99047

59A, 58A

0.97; 1.7

0.65; 0.45

0.42; 2.2

adjusted gauge 116001 at Trebonne

adjusted gauge 116001 at Trebonne

adjusted gauge 116001 at Gedges Crossing

C

67A

68A, 68B

0.18

0.65

adjusted gauge 116001 & 116006 at Timrith

adjusted gauge 116001 & 116006 at Lannercost

D

69A

70A

71A

2.1

0.81

1.7

adjusted gauge 116006 at Abergowrie Bridge

gauge 116006

adjusted gauge 116006 at Abergowrie 1 Bores are denoted without their 116000 prefix 2 Closest distance to the Herbert River

3 Bore 101 is the replacement for 51: 51A = 101A and 51B = 101C 4 River stage has been adjusted from the gauge value based on the measured water column and surveyed depths to the riverbed (refer to Section 6.1.1).

6.3.1.1 Reach A: river mouth to gauge 116001

The river reach downstream of gauge 116001 is tidal and therefore stage heights measured at

the gauge do not necessarily represent stage heights further downstream. However, in the

absence of detailed tidal information, river stage has been derived (from the measured water

column, riverbed depth and gauge data) at selected locations along the reach and compared with

groundwater elevations (Figure 6-7). The analyses indicate that all bores screened in the shallow

and deep aquifers have groundwater elevations above the corresponding river stage

measurements during the dry season. In addition, the analyses highlight that the river-

groundwater relationship generally reverses during streamflow events (peaks) in the wet season;

however, as groundwater heads (for both aquifers) increase further upstream along the reach,

such as at bores 11600052 and 11600053, this reversal is less frequent.

Page 182: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 6

158

-5

0

5

10

15

1975

1977

1979

1981

1983

1985

1987

1989

1991

1993

1995

1997

1999

2001

2003

2005

m A

HD

HR @ CSR

HR @ Gairloch

Stage @ CSR

Stage @Gairloch

11600048B

11600053B

a

-5

0

5

10

15

1975

1977

1979

1981

1983

1985

1987

1989

1991

1993

1995

1997

1999

2001

2003

2005

m A

HD

HR @ CSR

HR @ Gairloch

Stage @ CSR

Stage @Gairloch

11600048A

11600052A

b

Figure 6-7 Comparison of historical groundwater elevations in selected (a) shallow bores and (b) deep bores with corresponding river stage (adjusted) along reach A of the Herbert River. Circles represent calculated river stage based on the riverbed elevation and measured water column. Bore 11600048 is compared with river stage at CSR (blue), while bores 11600052 and 11600053 are compared with river stage at Gairloch (orange).

Page 183: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Physical River-Groundwater Interactions

159

6.3.1.2 Reach B: gauge 116001 to Stone River junction

Comparison between historical groundwater elevations and derived stage heights for HSs and

HSd bores along reach B indicates that river stage remains below the groundwater elevation in

both aquifers during the dry season (Figure 6-8). However, in the wet season river stage

sometimes exceeds the groundwater elevation, especially at bores on the left bank of the river

(e.g. bores 11600061, 99047 and 11600059) which tend to have a lower groundwater elevation

than on the right bank (e.g. bores 99025 and 11600062).

0

5

10

15

20

1975

1977

1979

1981

1983

1985

1987

1989

1991

1993

1995

1997

1999

2001

2003

2005

m A

HD

HR @Trebonne

Stage @Trebonne

11600061B

11600062B

99025

99047

a

0

5

10

15

20

1975

1977

1979

1981

1983

1985

1987

1989

1991

1993

1995

1997

1999

2001

2003

2005

m A

HD

HR @ Gedges Xing

HR @Trebonne

Stage @Gedges Xing

Stage @Trebonne

11600061A

11600062A

11600059

b

Figure 6-8 Comparison of historical groundwater elevations in selected (a) shallow bores and (b) deep bores with corresponding river stage heights (adjusted) along reach B of the Herbert River. Circles represent calculated river stage based on the riverbed elevation and measured water column. Bores 11600061, 11600062, 99025 and 99047 are compared with river stage at Trebonne (orange), while bore 11600059 is compared with river stage at Gedges Crossing (blue). Note that bores 99025 and 99047 have only partial records.

Page 184: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 6

160

6.3.1.3 Reach C: upstream of Stone River junction to Long Pocket

There are only two monitoring bores along reach C. Given that there are no stream gauges along

the reach, historical stage height data at gauges 116001 and 116006 have been adjusted based on

field measurements (Section 6.1.1). Table 6-5 shows that in the Herbert River at Timrith, the

derived stage height is 1.4 m above gauge 116001 and 13.9 m below gauge 116006 during both

the end (October 2004) and beginning (June 2005) of the subsequent dry season. In the Herbert

River at Lannercost, the measured stage height is 4.1 m above gauge 116001 and 11.2 m below

gauge 116006 during both dry season measurements. Note that both stream gauges give similar

stage height time series when adjusted for measurements at Timrith or Lannercost.

Table 6–5 Comparison of derived river stage at Timrith and Lannercost and actual stage heights at gauging stations 116001 and 116006 in the lower Herbert River.

Location Date

Derived river stage (m AHD)

Stage height gauge 116001 (m AHD)

Stage height gauge 116006 (m AHD)

Timrith 28/10/04 2.8 1.4 16.7

5/06/05 3.0 1.6 16.9

Lannercost 27/10/04 5.5 1.4 16.7

1/06/05 5.7 1.6 16.9

0

5

10

15

20

25

30

1975

1977

1979

1981

1983

1985

1987

1989

1991

1993

1995

1997

1999

2001

2003

2005

m A

HD

HR @Lannercost

Stage @Lannercost(116001)

Stage @Lannercost(116006)

11600068A

11600068B

Figure 6-9 Comparison of groundwater elevation in shallow (11600068B) and deep (11600068A) bores with river stage (adjusted) at Lannercost. Similar stage heights are attained when adjusted using the records at gauge 116001 or 116006. Circles represent calculated river stage based on the riverbed elevation and measured water column.

Page 185: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Physical River-Groundwater Interactions

161

Comparison of groundwater and adjusted stream elevations indicate that for the deep aquifer

(e.g. bore 11600068A) the groundwater elevation is above river stage during the dry season and

generally less during the wet season. However, for the shallow aquifer (e.g. bore 11600068B)

the groundwater elevation generally exceeds river stage throughout both wet and dry season

months (Figure 6-9).

6.3.1.4 Reach D: Long Pocket to Abergowrie

Stream gauge 116006 is located on the most upstream reach of the lower Herbert River. As for

the other river reaches, stage height has been adjusted at locations distant from the gauge (close

to the monitoring bores) based on derived stage heights (Table 6-4). As there was no river depth

measurement available for the Herbert River at Abergowrie, a conservative estimate of stage

height was made from a surveyed cross section of the riverbed (assuming a zero water column

as a minimum in the dry season). Stage height and groundwater elevation comparisons indicate

that the elevation of groundwater in the deep aquifer (e.g. bore 11600070) is generally above

river stage throughout the year, except during some wet season events when there is a large rise

in river stage (Figure 6-10).

15

20

25

30

35

1975

1977

1979

1981

1983

1985

1987

1989

1991

1993

1995

1997

1999

2001

2003

2005

m A

HD

Stage @116006

11600070

Figure 6-10 Comparison of groundwater elevation at bore 11600070 and stage height at adjacent gauge 116006.

Page 186: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 6

162

6.3.1.5 Interpretation of flux direction

The above analyses indicate that the direction of flux between the Herbert River and the two

alluvial aquifers varies both seasonally and spatially. Based on the available information, the

relationships of the four reaches of the Herbert River with groundwater are discussed below.

As summarised in Table 6-6, river reaches in the floodplain area (reaches A and B) display a

seasonal pattern: the dominant direction of potential flux is towards the river during the dry

season and from the stream to the aquifers during high flow periods in the wet season. In the dry

season, stream-aquifer relationships along reach C display a similar trend to the downstream

reaches. However, while the direction of flux reverses with respect to HSd in periods of the wet

season, the elevation of groundwater in HSs is maintained above stage height. The potential

flow direction is thus from the shallow aquifer to the stream throughout the year. Similarly, the

high elevation of groundwater in reach D means that groundwater levels are generally above

river stage throughout the year: the dominant direction of potential flux is from the deep aquifer

to the river.

Table 6–6 Summary of the dominant river-groundwater elevation relationships along the lower Herbert River during wet and dry seasons in the historical record.

Reach Season Aquifer(s) River-groundwater elevation relationship

Theoretical direction of flow

A dry shallow & deep gw elevation > stage gw → river

wet shallow & deep stage > gw elevation+ gw ← river

B dry shallow & deep gw elevation > stage gw → river

wet shallow & deep stage > gw elevation+ gw ← river

C dry shallow & deep gw elevation > stage gw → river

wet shallow gw elevation > stage+ gw → river

wet deep stage > gw elevation+ gw ← river

D dry deep gw elevation > stage gw → river

wet deep gw elevation > stage+ gw → river

+ This relationship is observed during the majority of wet season periods in the historical record.

Page 187: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Physical River-Groundwater Interactions

163

On balance, the analyses indicate that the Herbert River can be characterised as a dominantly

gaining stream with respect to the shallow and deep aquifers, consistent with the water level

contours constructed in Chapter 4 (Figure 4-14 and Figure 4-16). However, during streamflow

events in the wet season the lower reaches (A and B) of the river potentially represent a losing

system, noting that the direction of flux switches from the aquifers back to the stream as river

stage relaxes. While a reversal in flow direction is consistently observed at some locations in the

deep and shallow aquifers, at other locations this relationship depends on the extent of river

stage rise during a particular streamflow event. As there is greatest uncertainty in the river stage

data during high flow events, there is less confidence in the river-aquifer relationships

established during the wet season. However, assuming that the relative error between

groundwater elevations and river stage is small, in general there is a transition from a

dominantly gaining system to variably gaining-losing system from the upper to lower reaches of

the lower Herbert River.

An important observation from the analysis of paired stream-bore hydrographs is that even

though the direction of stream-aquifer flux is towards the river during the dry season,

groundwater elevation is always maintained above minimum river stage. This suggests that

while there is evidence of potential leakage from the aquifer to the stream, there is a constriction

in flow such that the groundwater head remains elevated. This is particularly evident in the

upper reaches of the catchment where there is up to a 10 m difference between the elevation of

the stream and shallow groundwater at a distance of approximately 800 m from the river (e.g.

bore 11600068B in Figure 6-9). It is beyond the scope of the thesis to examine in detail the

reasons for this observation; however, it was noted in Chapter 4 that mud units present

throughout the lithologic sequence may act as semi-confining/confining layers and hence retard

flow either vertically or laterally. Therefore, the hydraulic properties of the aquifers or

surrounding material may provide an explanation. Whilst a zero point error in either the

groundwater elevations or river stage is a possibility, a difference of up to 10 m seems unlikely.

The approach taken in this section has enabled the potential direction of flux between the lower

Herbert River and the adjacent aquifers to be determined. However, the potential direction of

flux does not necessarily equate to actual movement of water between the aquifers and the

stream, as this will depend on factors such as aquifer hydraulic properties and the infiltration

characteristics of the soils and riverbed sediment. In addition, the approach has assumed that

stage height and the elevation of the riverbed, bank and groundwater have been accurately

determined. Therefore, the following section considers the characteristics of streamflow to

establish whether there are volumetric gains and/or losses to the Herbert River that can be

attributed to stream-aquifer fluxes.

Page 188: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 6

164

6.3.2 Flow characteristics

There are a number of standard hydrological techniques used to study the flow characteristics of

rivers (Section 2.4.1.3). Several of these methods are applied in the following sections based on

the availability of historical streamflow data. Given that this thesis is primarily concerned with

interactions during the dry season, the low-flow regime of the river is the focus of the analyses.

6.3.2.1 Stream hydrographs

A stream hydrograph shows the fluctuations in stream discharge through time and can be used

to gain insight into the relationships between rainfall and runoff in a catchment as well as the

role of baseflow. As depicted in Figure 6-11a, streamflow in the river displays a distinct

seasonal pattern. In the wettest months, generally from December - March, flow in the river is at

least 200 GL/day and greater than 900 GL during large flood events in the historical record. In

comparison, dry season flows are generally less than 50 GL/day in April/May and decline to

less than 200 ML/day by October/November, before the next wet season commences. Although

streamflow declines by the end of the dry season, flow remains above zero despite little or no

rainfall (Figure 6-11b). In the absence of lakes (or glaciers) in the area this indicates that

releases from groundwater storage (± bank storage) must sustain streamflow in the lower

Herbert River during low-flow periods. Given the topographic relationships between the

streambed and aquifers and the relative river-groundwater elevations (Sections 6.2 and 6.3.1) it

is considered that a ‘true’ groundwater body is drained rather than just near-channel/bank

storages (Smakhtin, 2001).

In contrast to the lower reaches of the Herbert River, dry season flows at gauge 116004 in the

upper catchment (Figure 6-1) are commonly less than 30 ML/day and zero-flow has been

recorded several times in the historical record. This suggests that unlike in the lower catchment,

groundwater does not always sustain flow in the upper Herbert River during the dry season.

Page 189: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Physical River-Groundwater Interactions

165

0

100

200

300

400

500

600

700

800

1995

1996

1997

1998

1999

2000

2001

2002

2003

2004

2005

Flow

(GL/

day)

Flow @ 116001Flow @ 116006

a

1

10

100

1000

10000

100000

1000000

1995

1996

1997

1998

1999

2000

2001

2002

2003

2004

2005

Flow

(ML/

day)

0

100

200

300

400

Rai

nfal

l (m

m)

Flow @ 116001Flow @ 116006Rainfall

b

Figure 6-11 Daily flow at the two lower Herbert River stream gauges during 1995-2005. Source: QDNRW

Page 190: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 6

166

6.3.2.2 Flow duration

In addition to hydrograph analysis of streamflow, the characteristics of flow can be further

assessed by constructing flow duration curves. These plots show the percentage of time during

which the flow of a stream is equal to or greater than any specified discharge, regardless of

chronological order. If a stream is dominated by baseflow, the flow duration curve (FDC) is

characterised by a low slope, which means that for most of the year the stream has a fairly

uniform discharge rate. Conversely, if the FDC is steep, discharge is more variable throughout

the year and the stream is likely to be dominated by direct runoff: these streams are also likely

to show long periods of no flow. Therefore, unregulated streams that exhibit a low slope on the

FDC, and/or do not show a cease-to-flow point, indicate potential for strong linkages to

groundwater systems (REM, 2002).

Figure 6-12 illustrates the FDCs for gauges in the lower catchment as well as for one in the

upper catchment based on daily historical streamflow data. Comparison of the three FDCs

indicates that the probability of attaining or exceeding a particular discharge increases from

upstream to downstream (Figure 6-12a). A log-normal representation of the FDCs delineates the

low-flow and high-flow ends of the curves: Figure 6-12b clearly illustrates the different

characteristics of flow in the upper and lower catchments, particularly at low flow. Whilst the

Herbert River in the lower catchment flows throughout the historical record and can be

considered to be perennial, flow in the upper catchment is not always sustained. This is

consistent with a source of baseflow to the stream below the gorge, which divides the upper and

lower catchments.

In order for low-flows in the river to be maintained by groundwater storages, (i) the draining

aquifer must be recharged seasonally with sufficient volumes of water; (ii) the watertable must

be shallow enough to be intersected by the stream; and (iii) the aquifer’s size and hydraulic

properties must be sufficient to maintain flows throughout the dry season (Smakhtin, 2001).

While the lower catchment comprises alluvial aquifers, the majority of aquifers in the upper

catchment are in fractured rock (Johnson and Murray, 1997). Therefore, the difference in flow

characteristics between the upper and lower catchments could be due to differences in aquifer

properties as well as in the other factors listed above.

Page 191: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Physical River-Groundwater Interactions

167

0.01

0.1

1

10

100

1000

10000

100000

1000000

0 10 20 30 40 50 60 70 80 90 100

Probability (%)

Stre

amflo

w (M

L/da

y)gauge 116001gauge 116006gauge 116004

a

0.01

0.1

1

10

100

1000

10000

100000

1000000

-4 -3 -2 -1 0 1 2 3 4

Normal variate

Stre

amflo

w (M

L/da

y)

gauge 116001gauge 116006gauge 116004

(0.1) (2.3) (15.9) (50.0) (84.1) (97.7) (99.9)

b

Figure 6-12 Flow duration curves for stream gauges in the lower (116001, 116006) and upper (116004) Herbert River catchment (refer to Figure 6-1 for gauge locations). Note that the FDCs are represented in (b) as a log-normal plot, with probabilities (%) shown in brackets on the x-axis.

6.3.2.3 Hydrograph separation

Baseflow is defined as that part of streamflow that is not attributable to direct runoff from

precipitation or snowmelt and is usually contributed by groundwater storage or other delayed

sources (e.g. shallow subsurface storage). In many hydrological applications it is useful to be

able to separate the baseflow, or slow flow component, from quick flow (runoff, interflow and

direct rainfall), and hence attempt to isolate the low-frequency signal of a stream. In a wet

season discharge comprises baseflow and quickflow, which represent the catchment response to

rainfall; conversely, stream discharge in a dry season is dominated by baseflow (Smakhtin,

2001). The steady decline of stream discharge during periods without rainfall is referred to as

recession, representing the gradual drainage from subsurface storages.

Page 192: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 6

168

In the context of this study, it would be useful to quantify the groundwater input to the lower

Herbert River, especially during the recession period when the proportion of subsurface runoff

to surface runoff is considered to be greatest. At different times along a stream hydrograph the

baseflow contribution is comprised of varying proportions of groundwater discharge and

streambank storage. Contributions to baseflow from bank storage generally decline as stream

levels fall; therefore, by the end of the dry season the baseflow contribution should be largely

composed of groundwater discharge (Ward, 1967).

Baseflow separation techniques include digital filtering, graphical methods and chemical

separation approaches (Sponberg, 2000). Although digital filtering can be a simple and robust

method for evaluating baseflow for a large range of unregulated catchments, the results are very

sensitive to the filter parameter value, which needs calibration before the results can be

considered to be numerically valid (REM, 2002; SKM, 2001). The Lyne-Hollick digital filter

(Lyne and Hollick, 1979) is a one-parameter mathematical algorithm that separates the runoff

signal, or quick response; baseflow is then calculated by subtracting the filtered quick response

from total streamflow (Sponberg, 2000). Although there are numerous documented baseflow

filters (Furey and Gupta, 2001; Chapman, 1999; Nathan and McMahon, 1990; Lyne and

Hollick, 1979), the Lyne-Hollick filter was considered to be appropriate for the purposes of this

study as it has been widely applied to daily data and has a recommended filter parameter for

daily data (Grayson et al., 1996). Hence, the filter was applied to historical streamflow data at

the two lower Herbert River gauges in order to generate baseflow time series.

Figure 6-13 shows the relationships between total (observed) streamflow at the stream gauges

and the corresponding calculated baseflow over a 1-year period. Visual inspection indicates that

the contribution of baseflow as defined by this algorithm is greater in absolute terms during the

wet season (November-April); however, the proportion of baseflow to streamflow is greater

during the dry season (May-October). Consistent with the flow duration and hydrograph

analyses above, modelled baseflow at the end of the dry season correlates with total flow i.e.

stream discharge is maintained by subsurface storages. The marked seasonality of the baseflow

signal as calculated by this methodology is clearly evident in Figure 6-14, consistent with the

seasonal trends observed from the bore hydrographs in Chapter 4 (Figure 4-6).

Page 193: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Physical River-Groundwater Interactions

169

0

100

200

300

400

500

600

700

800

Jan

00

Feb

00

Mar

00

Apr

00

May

00

Jun

00

Jul 0

0

Aug

00

Sep

00

Oct

00

Nov

00

Dec

00

Jan

01

Flow

(GL/

day)

116001 total flow116006 total flow

116001 baseflow116006 baseflow

a

100

1000

10000

100000

1000000

Jan

00

Feb

00

Mar

00

Apr

00

May

00

Jun

00

Jul 0

0

Aug

00

Sep

00

Oct

00

Nov

00

Dec

00

Jan

01

Flow

(ML/

day)

116001 total flow116006 total flow

116001 baseflow116006 baseflow

b

Figure 6-13 Baseflow separation using the Lyne-Hollick algorithm at the lower Herbert River stream gauges. Note that the filter was applied to the entire historical record of streamflow data but for clarity is only shown for 2000-01. Source: daily streamflow data from QDNRW.

Page 194: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 6

170

The baseflow index (BFI) is an important index for low-flow studies and is defined as the

volume of baseflow divided by the volume of total streamflow (Smakhtin, 2001). As shown in

Figure 6-15, there is a range in the proportion of baseflow contributed to streamflow from

approximately 10-100% on a daily basis. However, on a bimonthly basis, the contribution is

generally within 50-90% throughout the course of a year at gauge 116001. Calculation of the

seasonal BFI (volumetric ratio of historical mean baseflow to historical mean streamflow for

months within each season) at gauge 116001 expressed as a percentage, gives 34% and 67% for

the wet and dry seasons, respectively. Similar percentages of 31% and 66% for average wet and

dry season baseflow contributions are calculated at gauge 116006. Although the BFI calculated

at the two gauges along the lower Herbert River are very similar, the volume of baseflow is

between 20-100 % greater at the downstream gauge at any time (Figure 6-14). Given that

calculated baseflows represent the aggregated contributions from the entire catchment area

upstream of each gauge, these observations indicate that groundwater contributions to the river

from the additional catchment area between the two gauges are significantly large. Based on

analysis of the baseflow contribution, there is evidence that groundwater discharges to both the

upper and lower reaches of the lower Herbert River. Therefore, the large subsurface

contribution between the gauges may reflect enhanced river-aquifer connectivity due to

aquifer/soil hydraulics, recharge sources and volumes, as well as physical river-aquifer

relationships such as relative elevations of the stream-aquifer topography and water levels.

6.4 IMPLICATIONS FOR N TRANSPORT

Based on the available information, the analyses in this chapter have shown that there is

evidence of groundwater contributions to the lower Herbert River throughout the year. Whilst

there is potential flux from the river to the aquifers along some reaches during periods of high

river stage, the dominant direction of flux is from the aquifers to the river. These interaction

characteristics have important implications for the transport of N from groundwater to surface

waters. It was established in Chapter 5 that high concentrations of DIN are found in both

shallow and deep aquifers, including adjacent to the Herbert River. In HSd a hotspot for NO3- is

found in bores along reach D, while for HSs high concentrations of NO3- are observed in bores

along reaches A, B and C. Hence, given the hydraulic relationships, the potential exists for NO3-

to be contributed to the lower Herbert River from groundwater sources along the entire length of

the river. However, near-stream processes such as dentrification, resulting from contact with

reducing sediments or riparian vegetation prior to groundwater discharge, may reduce or

prevent the emergence of DIN in surface waters (Bohlke and Denver, 1995). River chemistry

data are thus evaluated in the following chapter to observe whether there is evidence of N in the

Herbert River derived from the alluvial aquifers.

Page 195: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

10

100

1000

10000

100000M

ay 9

1

May

92

May

93

May

94

May

95

May

96

May

97

May

98

May

99

May

00

Bas

eflo

w (M

L/da

y)

0

100

200

300

400

500

600

700

800

900

Stre

amflo

w (G

L/da

y)

116001 baseflow

116006 baseflow

116001 streamflow

Figure 6-14 Time series of calculated baseflow at stream gauges 116001 and 116006 and observed total flow at gauge 116001 (for reference). Note the baseflow recession period during the dry season (May to October). Note the larger volume of baseflow at the downstream gauge (116001) compared to the upstream gauge (116006).

Page 196: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

0

10

20

30

40

50

60

70

80

90

100

1981

1982

1983

1984

1985

1986

1987

1988

1989

1990

1991

1992

1993

1994

1995

1996

1997

1998

1999

2000

2001

Bas

eflo

w p

ropo

rtio

n

0

100

200

300

400

500

600

700

800

900

Stre

amflo

w (G

L/da

y)

baseflowproportionmovingaverage

streamflow116001

Figure 6-15 Calculated daily baseflow as a percentage of observed streamflow at gauge 116001. The red trend represents a bimonthly moving average through the baseflow proportions. Flow at the gauge highlights the wet (November-April) and dry (May-October) seasons. The baseflow filter was applied to the entire historical record of streamflow data but for clarity is only shown for a 10-year period. Note the trend of increasing baseflow proportion from the beginning to end of the dry season, which corresponds with progressively decreasing streamflow.

Page 197: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Physical River-Groundwater Interactions

173

6.5 CHAPTER SUMMARY

Physical relationships between the alluvial aquifers of the lower catchment and the Herbert

River have been examined in this chapter. In accordance with the classification scheme for

stream-aquifer interactions outlined in the introduction to this chapter, there is evidence of

potential for hydraulic connection along the entire lower Herbert River. The dominant direction

of potential flux is from the aquifers to the river, although during short periods of high river

stage in the wet season the direction reverses along some reaches in the floodplain area. Flow

characteristics provide an indication of actual flux between the aquifers and the stream.

Baseflow is an important source of recharge to the river, particularly during the dry season when

contributions from groundwater can be up to 100 %. While groundwater contributions to both

the upper and lower reaches of the lower Herbert River are apparent, there is an enhanced

contribution of baseflow between the two gauges in the study area (i.e. below gauge 116006) as

indicated by digital filtering. This may reflect enhanced river-aquifer connectivity due to

aquifer/soil hydraulics, recharge sources and volumes, as well as physical river-aquifer

relationships. On balance, the upper reaches of the lower Herbert River can be considered to be

a dominantly gaining system, while towards the coast the river is characterised as a variably

gaining/losing stream. Given that similar physical river-groundwater relationships have been

identified in this chapter for each aquifer, it is not possible based on the available information to

determine the relative proportion of groundwater flux from each aquifer to the river.

Hydrochemical data is thus examined in Chapter 7 to build on the conceptual model for river-

groundwater interactions. The physical relationships established in this chapter provide a

platform for the subsequent hydrochemical analysis and assessment of the potential impacts of

stream-aquifer interactions on surface water quality, particularly in relation to species of N.

Based on the spatial distribution of DIN in both shallow and deep aquifers, the stream-aquifer

hydraulics indicate that there is potential for N in groundwater to contribute to the lower Herbert

River. Evidence for this is provided in the following chapter.

Page 198: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to
Page 199: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

175

Chapter 7 Chemical River-Groundwater Interactions

7.1 INTRODUCTION

In the previous chapter it was established that based on physical hydraulic relationships there is

evidence of stream-aquifer interaction along the lower Herbert River. In addition, it was

demonstrated that the dominant direction of interaction is from groundwater to the river,

especially during the dry season. This chapter explores the chemistry of the surface water

system in order to verify the physical river-aquifer linkages and enrich the conceptual model

that the previous three analytical chapters (4-6) have developed. Temporal river chemistry data

are analysed in order to identify hydrochemical signatures and seasonal patterns. Longitudinal

trends are also examined, from data collected for this study, in order to identify variations in

water sources to the stream during months at the beginning and end of the dry season. A variety

of dissolved constituents are examined, both conservative and non-conservative tracers,

including field parameters, major ions, and stable isotopes. Given the focus of the thesis on

river-groundwater interactions, radioactive isotopes are also analysed in order to trace

groundwater discharge and estimate the flux along the river. The synthesis of the extensive

hydrochemical database provides a powerful tool to characterise the chemistry of the river in

space and time, and for the key processes that influence its chemistry to be identified. Based on

the observed concentrations of DIN in the shallow and deep aquifers adjacent to the river, it was

proposed in Chapter 6 that there is the potential for N transport to the river from groundwater

sources. Therefore, in accordance with the objectives of the thesis, this chapter evaluates the

role of groundwater as a transport vector for dissolved inorganic forms of N to the river.

Furthermore, the environmental significance of the results for in-stream and marine ecosystems

is considered. A conceptual diagram summarising water and N movement between the aquifers

and the lower Herebert River is presented at the end of this chapter (Figure 7-38).

7.1.1 Factors that influence river chemistry

Surface water and groundwater contain dissolved solutes and gases derived through a variety of

processes within the hydrologic cycle. As discussed in Chapter 2 (Section 2.2.2), climatic,

geological and biochemical factors influence the chemistry of natural waters; mixing between

waters from different sources also impacts on stream and groundwater chemistry. Compared to

groundwater, the chemistry of a stream is generally more dynamic due to faster water

movement; rapid response to rainfall events; and potential for multiple sources which contribute

Page 200: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

176

water to the stream, including groundwater. In addition, the direct exposure of surface waters to

the atmosphere means that climatic effects such as evaporation in general have a greater

influence on surface water chemistry than subsurface waters, especially in comparison to deep

groundwaters.

Based on the compositions of the world’s major rivers, Gibbs (1970) presented a diagram which

relates total dissolved solids (TDS) and the Na/(Na+Ca) ratio to predict whether a stream

sample is dominated by: (1) the chemistry of precipitation (rain dominated); (2) rock-weathering

reactions (rock dominated); (3) evaporative concentration-crystallisation; or by a combination of

these influences (Figure 7-1). Seawater plots in the top right hand corner of the diagram, while

rainfall generally plots in the lower right hand corner at low TDS.

1

10

100

1000

10000

100000

0.0

0.5

1.0

Na+/(Na++ Ca2+) (meq/L)

TDS

(mg/

L)

Precipitation dominance

Evaporation/ fractional crystallisation dominance

seawater

Rock dominance

Figure 7-1 Gibbs diagram depicting the key processes that control the chemistry of surface waters. The world’s major rivers plot within the ‘boomerang’ envelope (modified after Gibbs 1970).

In addition to environmental factors are the effects of human activities. For instance,

agricultural development can impact on stream water quality through the use of fertilisers and

pesticides as well as pumping of surface and/or groundwaters for irrigation. As the case study

area is largely comprised of dryland cropping, the potential effects of irrigation will not be

discussed further. However, as the catchment is dominated by sugarcane farming, nitrogen is an

important diffuse contaminant to the surface water network (Section 3.2.1, Chapter 3).

Superimposed on the processes which affect the transport of chemical constituents to a stream

are in-stream biogeochemical reactions such as acid-base reactions; mineral precipitation and

Page 201: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

177

dissolution; sorption and ion exchange; oxidation-reduction (redox) reactions; dissolution and

exsolution of gases; and biodegradation (Winter et al., 1998). These reactions affect the fate and

concentration of chemical species from their point of entry to further downstream.

A comprehensive discussion of the key thermodynamic principles and chemical processes that

ultimately control the composition of natural waters is provided in Hem (1985). The effects of

each of the natural processes and anthropogenic activities outlined above on the observed

hydrochemistry of a stream are summarised below.

7.1.1.1 Climatic factors

The amount and rate of rainfall, runoff and evaporation, as well as temperature, are the key

components of climate that influence surface water composition. Solutes introduced through

rainfall reflect inputs from seawater, dust, and atmospheric gases, which in turn impart a

characteristic chemistry to the rainfall that hits the land surface. The ultimate chemistry of a

stream largely depends on the pathways that this rainfall moves along and hence the time it

takes for the water to reach the stream. While overland flow can have an important effect on

stream quality following storm events, groundwater inflow can dominate the hydrochemistry

during dry periods. In general, the solute concentration of river water tends to be inversely

related to flow rate, such that at very high flow rates runoff may be nearly as dilute as rainwater

(Hem, 1985). This observation stems from the fact that direct runoff has a relatively short

contact time with soil or vegetation. Thus, an increase in streamflow is generally accompanied

by a decrease in solute concentration due to dilution by waters containing a lower concentration

of particular ions than originally present in the stream. An exception is where the chemical

constituent is insoluble and present within the stream sediment load, such that a positive

correlation between stream discharge and concentration may be found (Langmuir, 1997). In

contrast, water infiltrating through the soil zone and interacting with rocks undergoes more

extensive reactions due to the longer residence time, resulting in baseflow that has a higher

concentration of dissolved solids. Therefore, groundwater discharge to a stream would be

expected to increase the concentration of solutes derived from the soils and/or geology.

Temperature is another aspect of climate that can influence stream chemistry through different

rates of weathering, biogeochemical reactions and evaporation. While weathering and

biogeochemical reactions differentially affect the concentration of the solutes in a stream,

evaporation results in an increase in the concentrations of all dissolved components in the

stream, constant ion ratios, and enrichment in stable isotopes of water. These are important

diagnostic features for distinguishing between processes. Climates characterised by distinct wet

and dry periods may favour weathering reactions and therefore produce larger amounts of

soluble inorganic matter during particular seasons of the year (Hem, 1985). Also related to the

seasons are flow and river stage, which influence the variability and overall concentration of

Page 202: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

178

solutes in-stream (McNeil et al., 2005). Therefore, in catchments such as the Herbert that are

distinctly seasonal in regards to temperature, rainfall and flow, river chemistry would be

predicted to change between the seasons due to one or more of these factors.

7.1.1.2 Geologic factors

Dissolution and leaching of minerals are major sources of dissolved solutes reaching surface

waters. Therefore, the hydrochemical signature of a stream is characteristic of the dominant

source(s) of water, such as a particular aquifer system. Reactions between streambed sediments

and the stream can have an additional impact on the observed hydrochemistry.

The noble gas radon (222Rn) is derived from radioactive decay of uranium-series isotopes;

therefore, aquifers containing uranium-bearing minerals will also contain 222Rn. Due to the short

half-life of 3.8 days and loss via gas exchange with the atmosphere, high radon concentrations

are only present in surface waters in the vicinity of points of groundwater discharge (inflow)

and at relatively short distances downstream of such locations (Cook et al., 2003). For these

reasons 222Rn is a particularly useful environmental tracer for identifying zones of preferred

groundwater discharge into surface water bodies. In contrast to dissolved inorganic tracers, 222Rn is also conservative and is not modified by biogeochemical reactions. Whilst the

concentration of 222Rn in groundwater is significantly different to that in surface waters, radon

data must be interpreted with caution, as variations in radon concentration along a river may be

due to numerous factors. For example, aquifers comprised of different geology may have an

inherently different composition of uranium and hence radon. In addition, tributaries supplied

by groundwater may also contribute radon to the stream that is not related to direct inflow from

an aquifer. A multi-tracer approach is thus desirable, coupled with a good understanding of the

hydrogeology. In addition to examining qualitative trends, 222Rn data can be used to derive

estimates of groundwater discharge to surface waters, as described in Stieglitz (2005) and Cook

et al. (2004). The approach is applied to data collected as part of this study (Section 7.5).

7.1.1.3 Biogeochemical factors

Biochemical processes that affect the chemistry of surface waters include processes that require

a net energy input, such as photosynthesis; redistribute chemically stored energy; convert

chemically stored energy; or do not involve energy transfer (Hem, 1985). Given the focus in this

thesis on nitrogen in surface water and groundwater, bacteria-assisted transformation reactions

such as nitrification and denitrification (Section 5.1.1.4, Chapter 5) represent important

biogeochemical mechanisms that influence the concentration and speciation of nitrogen in

water. The concentrations of other solutes such as silica and trace elements such as iron may

also be controlled by biological processes (Hem, 1985).

Page 203: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

179

7.1.1.4 Mixing of waters

Groundwater is generally more enriched in dissolved solids than runoff due to the longer time

for mineral-water reactions to occur. This means that discharge of baseflow to a stream typically

results in mixing between hydrochemically distinct waters, leading to a shift in hydrochemical

signature i.e. a change in solute concentrations, ion ratios, water quality indicators (e.g. pH, Eh,

temperature), and isotope composition. In addition to groundwater, a stream receives water from

tributaries that may also have a distinct chemical signature due to contributing sources and in-

stream reactions. This results in mixing relationships between different waters. A further

hydrochemical mixing trend occurs where fresh and saline waters meet in the estuary of a river.

This mixing zone fluctuates in position depending on river flow, winds and ocean tides (Hem,

1985). Mixing relationships in the tidal zone are complex because of the greater density of

seawater compared to freshwater and hence vertical and horizontal variability. Nonetheless, as

shown in this chapter, the tidal zone represents an important part of the river system that

warrants hydrochemical investigation because of its junction between terrestrial and marine

waters. Hydrochemical consequences of mixing between fresh groundwater and saline water,

both in the Herbert River estuary and the sea, were also discussed in Chapter 5 (Sections 5.4.2

and 5.6.3).

The potential mixing relationships within a stream can generally be distinguished if the end

member compositions are known. For instance, if the hydrochemical signatures of groundwater

and runoff have been identified, then the proportion of mixing can be calculated by a simple

mass balance relationship for solutes that differ in concentration between the two waters. The

inherent assumption is that the solute behaves conservatively upon mixing. Therefore, if Cl is

the constituent of interest, the mass balance relationship is described by:

[Cl]M = x[Cl]A + (1-x)[Cl]B (7-1)

where [Cl]M is the Cl concentration of the mixture; [Cl]A and [Cl]B are the Cl concentrations of

water A and B, respectively; x is the fraction of water A; and (1-x) is the fraction that water B

contributes to the mixture.

7.1.2 Assessment principles and methods

Given the variety of factors that can potentially influence the chemistry of the river, different

tracers are required to distinguish between processes (refer to Section 5.1.1.1 for a summary of

tracer types). An environmental tracer is particularly useful for estimating groundwater inflows

to rivers where the concentration of the tracer in groundwater is relatively uniform and

significantly different to that in the river (Cook et al., 2003). Importantly, when surface water

chemistry is used to identify stream-aquifer interactions, the groundwater component in the

river is the cumulative result of the hydrogeological and hydrological processes along the entire

Page 204: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

180

upstream watercourse length. Therefore, the interpretation of environmental tracer data requires

knowledge of upstream processes that affect both surface and groundwater quality. Based on a

review of tracer techniques, REM (2002) conclude that methods for investigating groundwater-

surface water interactions fall into two main groups:

• Stream routing approaches: where the concentration of the tracer of interest is measured

along the entire stream reach, allowing identification of zones of groundwater

discharge.

• Groundwater approaches: where the concentration of a tracer in (nested) piezometer

transects is measured and compared to stream water concentrations, allowing

determination of stream water contributions to groundwater.

Using a variety of tracers, a combination of these approaches is used in this chapter in order to

characterise the dynamics of the system (Section 2.4.1.4).

Table 3-2 (Chapter 3) summarised the field and laboratory parameters that were measured in

surface water samples and the primary reason(s) for their measurement. In order to assess

processes in detail this extensive dataset of hydrochemical information must be analysed in a

targeted manner. Various methods for organising groundwater data in order to assist with

hydrochemical interpretation were presented in Section 5.1.2 (Chapter 5). Similar techniques are

also applicable to surface water samples, such as Piper, Schoeller and bivariate plots. In

addition, spatial relationships are examined in this chapter through longitudinal plots, while

temporal hydrochemical trends are explored through time series analyses.

The physical and chemical factors outlined in Section 7.1.1 result in: (i) a change in the

concentration of one or more constituents in the river; and/or (ii) a change in chemical signature,

which may be at a single location in the stream or follow a trend along the river. Therefore, the

challenge is to select appropriate chemical species and relationships with other parameters that

will reveal conclusive information about specific processes that impact on the chemistry of

surface waters. The key natural physicochemical processes identified above were: overland

flow, evaporation, groundwater discharge, and mixing with tributaries and seawater.

Denitrification and nitrification were also noted as important biogeochemical processes that

specifically influence stream nitrogen concentrations. Summarised in Table 7-1 are the

distinctive chemical characteristics observed in a stream arising from the range of processes that

can influence the stream’s chemistry. Salinity is an important chemical measure in the stream

for numerous processes and can be represented by the total dissolved solids (TDS) content as

well as total dissolved ions (TDI) and electrical conductivity (EC). While each of these

measures represents the majority of dissolved constituents, there are important differences. As

summarised by McNeil and Cox (2000), TDS is the concentration of dissolved substances in

Page 205: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

181

water, including mineral and organic matter, whether or not in ionic form: this includes SiO2. In

contrast, TDI is the total number of ions in solution, whether they are dissociated or not: the

TDI is calculated as the sum of the major ions expressed in mg/L. EC is the ability of a solution

to conduct an electrical current: it is not affected by dissolved silica or undissociated ions such

as H2CO3 which do not carry an electric charge. McNeil and Cox (2000) note that TDI is the

most useful salinity measure in hydrological studies, although EC is the most convenient to

measure. Both of these measures are analysed in this chapter.

Table 7–1 Distinctive chemical characteristics of processes that influence the chemistry of surface waters. Key attributes used in the analyses are highlighted. Process Influence on stream chemistry Overland flow Decrease in salinity (EC, TDI) and dilution of ion concentrations; increase in

Cl/HCO3

Isotope composition of the stream may reflect mixing between different waters (baseflow-dominant and rainfall)

Inverse relationship between solute concentration and flow

Generally low pH (4-6)+

Evaporation Enrichment in stable isotopes of water, particularly in δ18O relative to δ2H

Increase in salinity (e.g. EC, TDI)

Constant ion ratios

Groundwater discharge

Receiving stream has a hydrochemical signature (ions/ratios/isotopes) similar to that of groundwater (Chapter 5) or is consistent with a mixing trend

Presence of radon

Greater salinity than surface waters

Often more reducing (anaerobic) than surface waters (lower Eh)

Tributary inflow

Receiving stream has a hydrochemical signature (ions/ratios/isotopes) similar to that of the contributing tributary or is consistent with a mixing trend (downstream of the tributary entry point)

Seawater mixing

Increase in salinity (TDS, TDI, EC) and other solute concentrations except silica

Decrease in radon concentration

Increase in pH (8.1)

Mixing trends between fresh and oceanic water (ions/ion ratios/isotopes) +pH varies from 4-6 (continental) to 5-6 (coastal) based on global atmospheric precipitation, with initial rainfall from a given event tending to be the most acid (Langmuir, 1997)

7.1.3 Study site & terminology

As discussed in Chapter 3, water quality data were collected as part of this study soon after

cessation of the wet season (May) and at the end of the dry season (October) in 2004, as well as

in the early part of the dry season (June) in 2005. The Herbert River gorge marks the upstream

extent of the lower catchment area. The most upstream samples were collected just below the

Page 206: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

182

gorge at Nash’s Crossing (within Yamanie National Park), whilst the furthest downstream

samples were collected below gauge 116001 within the tidal zone of the river (Figure 7-2). Note

that Nash’s Crossing is located upstream of sugarcane production areas. To aid with the

analysis, the Herbert River is divided into a series of reaches (1 - 24): the downstream extent of

each reach is marked by the location at which a full water quality sample was collected (for

major ion analysis). As this thesis is concerned with river-groundwater interactions, river

reaches are numbered from upstream to downstream, consistent with the general direction of

groundwater flow (Chapters 4 and 5). Therefore, reach 1 refers to the reach downstream of the

Herbert River gorge, to Nash’s Crossing (2.9 km). Locations along the river are also referred to

in the text as distance in km downstream from the gorge, as this provides a better indication of

distances between observed hydrochemical changes in the stream than reach numbers (Figure

7-2).

U

U

#S#S

#S#S

#S#S

#S

#S #S

#S

#S#S

#S#S#S#S

#S#S#S #S #S#S#S#S

#S

#S

#S#S

#S

#S

#S

#S #S

#S#S #S

#S

#S

#S

#S #S

#S#S

#S #S#S #S

#S

#S #S#S

#S#S #S

#S #S#S

#S#S

#S

#S

#S

#S

#S

#S

#S

#S

#S

#S#S

#S

#S#S

$T

$T

$T$T

$T

$T$T

$T

$T

$T

$T

$T$T

$T

$T

$T

$T

$T$T

$T$T

r

r

Elphinstone Ck

Gow

rie Ck

Stone R

Lannercost Ck

Ripple C kH aw

k ins

Ck

Dal

rym

ple

Ck

#

Herbert RGorge

Palm C k

Tre bonne Ck

Seymour R

HERBERT RIVER

2.9 km

34.0 km

57.4 km

68.3 km

76.4 km

84.5 km 93.8 km

100.0 km

24.4 km 44.5 km

HinchinbrookIsland116006

116001

May 2004$T

October 2004#S

June 2005#SRainfall 2004r

N

5 0 5 Km

Figure 7-2 Surface water sampling sites during the three collection periods and locations of rainfall samples. Selected reach numbers and distances downstream of the Herbert River gorge are also indicated. Note that the two QDNRW stream gauges are located at approximately 33 km (116006) and 76 km (116001) downstream of the gorge.

In this chapter the months of May-October are referred to as dry season months, while

November-January are months of the wet season, noting that the beginning of the wet season

varies from year to year and that there is a transition period between seasons. The beginning and

end of the dry season are interchangeably referred to as higher flow and lower flow periods,

respectively.

1

2

3

4

5

6 8

910

11

16

18

20 14

24

22

Nash’s Crossing

Page 207: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

183

7.2 GENERAL HYDROCHEMISTRY

Hydrochemical data are analysed in this section to establish general chemical characteristics of

the Herbert River and some of its tributaries. Given the focus in this chapter on river-

groundwater interactions, surface water compositions are also compared with that of

groundwater (Chapter 5).

The Piper diagram depicted in Figure 7-3 illustrates that the Herbert River is relatively enriched

in HCO3, except in the tidal zone, where Cl is the dominant anion. Na, Ca and Mg comprise the

major cations. Note that while the Herbert River samples appear to define a linear trend in the

Piper diagram, this trend does not translate to a systematic evolution in chemistry along the

river. The exception is for samples in the estuary with Cl > 60 % of the major anions, which

correspond to progressively increased proportions of seawater towards the river mouth. Overlap

in relative major ion concentrations is observed between the Herbert River and groundwaters as

well as with seawater, consistent with waters contributing to the river from multiple sources.

However, given the narrow trend for surface waters compared to groundwaters, multiple sources

might be expected to give rise to considerable scatter if the groundwater inputs are as variable as

indicated in Figure 7-3. Surface waters appear to be relatively more enriched in Mg and

depleted in SO4 compared to groundwaters. A summary of major and minor inorganic chemistry

for all surface water samples is provided in Appendix A.

80 60 40 20 20 40 60 80

20

40

60

80 80

60

40

20

20

40

60

80

20

40

60

80

Ca Na HCO3 Cl

Mg SO4

Herbert RTributariesRainfallSeawaterHSsHSd

Figure 7-3 Piper diagram for surface water and groundwater samples collected during three sampling periods: May 2004, October 2004, June 2005.

Page 208: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

184

A modified Gibbs diagram (Figure 7-4), whereby TDI is plotted instead of TDS (compare with

Figure 7-1), shows that samples collected along the lower Herbert River are generally of low

salinity (TDI) and are Na-dominated, similar to the Type D water types of McNeil et al. (2005)

only with a dominance of HCO3 over Cl. Type D waters are characteristic of the steep, high

rainfall, northern coastal streams of Queensland which are typically associated with granitic

terrains that release Na, K and Ca upon weathering of feldspar. Low salinity samples also have a

Na:Cl ratio (meq) of greater than 1, indicative of Na released during silicate weathering

(Meybeck, 1987). Similarly, according to the classification scheme proposed by Meybeck

(2004), based on the sum of cations as an indicator of weathering, the freshwater reaches of the

Herbert River generally correspond to dilute or medium dilute waters, indicative of rainfall and

silicate weathering controls. The higher TDI estuarine samples, which have Na/Cl (molar) equal

to 0.86 and Na/(Na+Ca) approximating that of seawater, reflect an oceanic origin of salts

(Figure 7-4). These waters correspond to the Type A water type of McNeil et al. (2005)

associated with evaporation and/or tidal influences.

1

10

100

1000

10000

100000

0.0

0.5

1.0

Na/(Na + Ca) (meq/L)

TDI (

mg/

L)

Seawater

Rainfall

Groundwater

HR - QDNRW

HR - this study

Evaporation

Dilution

Weathering

Figure 7-4 Modified Gibbs diagram with logarithmic plot of TDI against Na/(Na+Ca) for water samples collected along the lower Herbert River (HR): estuarine samples are circled. Groundwater (both aquifers), seawater and rainfall (collected inland from the coast) are also shown (adapted from McNeil et al. 2005).

7.2.1 Compositional groups

Samples were collected along the entire length of the lower Herbert River, including within the

tidal reaches. Schoeller plots indicate that there are two main compositional groups, with some

samples representing mixtures between these end members (Figure 7-5a).

Page 209: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

185

Consistent with the Piper and Gibbs diagrams (Figure 7-3 and Figure 7-4), surface water in the

tidal zone has a hydrochemical signature that resembles seawater. In contrast, freshwater

compositions are generally more depleted in Cl and SO4 and relatively enriched in HCO3 and

Ca, similar to the Ca-Mg groundwaters of the shallow and deep aquifers (Figure 7-5b). In

addition to Schoeller plots, saturation indices provide evidence of hydrochemical similarities

between the Herbert River and Ca-Mg enriched groundwaters. Note the two distinct groups

within the low salinity trend depicted in Figure 7-5a which are further examined in Section

7.3.2.2.

a

b

Figure 7-5 Schoeller plots of (a) samples in the Herbert River, with seawater highlighted in red and (b) Ca-Mg enriched groundwaters of the shallow aquifer (compare also with the Ca-Mg group of the deep aquifer, Figure 5-7b).

7.2.2 Stable isotopes

The theory behind the use of stable isotopes and factors that cause isotopic fractionation were

outlined in Chapter 5 (Section 5.1.1.3). Stable isotope compositions of samples collected in May

2004 in the Herbert River and two of its tributaries (Stone River and Ripple Creek, Figure 7-2)

are provided in Figure 7-6, with groundwater compositions also indicated. Surface waters

generally cluster around the LMWL, derived from analyses of groundwaters in the lower

catchment (Figure 5-3, Chapter 5). Although isotopic enrichment is observed between some

consecutive reaches (e.g. reaches 9-11), as well generally from upstream (reach 1) to

downstream (e.g. reach 22), there is no evidence of a continuous evaporation trend (inset Figure

7-6). Longitudinal isotopic trends are further analysed in Section 7.4.3.1. Depletion in heavy

isotopes in the upper reaches of the Herbert River (reaches 1-3) is consistent with the elevation

effect, while the isotope composition of Ripple Creek, which lies on the LMWL, is consistent

with enrichment due to rainfall from a short-duration event (the amount effect). Note that the

isotope composition of some shallow and deep groundwaters is similar to that of the Herbert

River, indicative of common source waters.

Herbert River Ca-Mg enriched (HSs)

Page 210: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

186

-7 -5 -3 -1 1-40

-30

-20

-10

0

10Herbert RStone RRipple CkRainfallSeawaterHSsHSd

δ2H (o/oo, SMOW)

δ18O (o/oo, SMOW)

SMOW

coastal rainfall MoretonBay

GMWL

Local meteoricwater line

Brisba

ne

WL

12

3

5

9

18

11

20

16

24

22

14

-7 -5 -3 -1 1-40

-30

-20

-10

0

10Herbert RStone RRipple CkRainfallSeawaterHSsHSd

δ2H (o/oo, SMOW)

δ18O (o/oo, SMOW)

SMOW

coastal rainfall MoretonBay

GMWL

Local meteoricwater line

Brisba

ne

WL

-7 -5 -3 -1 1-40

-30

-20

-10

0

10Herbert RStone RRipple CkRainfallSeawaterHSsHSd

δ2H (o/oo, SMOW)

δ18O (o/oo, SMOW)

SMOW

coastal rainfall MoretonBay

GMWL

Local meteoricwater line

Brisba

ne

WL

12

3

5

9

18

11

20

16

24

22

14

12

3

5

9

18

11

20

16

24

22

14

12

3

5

9

18

11

20

16

24

22

14

Figure 7-6 Oxygen-18 (δ18O) and deuterium (δ2H) stable isotope data for surface water and groundwater samples and a coastal rainfall event in May 2004. The isotopic composition of seawater at Moreton Bay (QLD) (Cresswell 2006, pers. comm.) and SMOW are also shown. Trend lines represent the LMWL (blue); GMWL (black solid); and Brisbane water line (black dashed) (see Figure 5-3). The inset shows isotopic compositions of samples along the Herbert River (labelled by reach number) and the Stone River (near the confluence with the Herbert River), as well as groundwaters with similar isotopic values.

7.3 TEMPORAL DATA

Historical measurements of field parameters and a suite of major and minor elements are

available from QDNRW at several stream gauges in the catchment. These datasets, combined

with data collected for this study, are analysed below to examine temporal hydrochemical trends

in surface waters in the catchment. Although the lower catchment, including its two gauges

(116001 and 116006), is the focus of the analysis, data from an upper catchment gauge

(116004) are included for comparison where applicable. Figure 7-7 illustrates the location of the

stream gauging stations in relation to the entire catchment. Given the emphasis of this thesis on

nitrogen, temporal trends for different forms of N in the lower Herbert River are also examined

based on data from CSIRO (Bramley and Muller, 1999).

Page 211: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

187

%U

%U

%U

116004

116006

116001

stream gauges%U

lower Herbert Riverstreams

#S upstream sampling site

20 0 20 kmN

#

Herbert RGorge

#S

Figure 7-7 Selected QDNRW stream gauges along the lower (116006, 116001) and upper (116004) Herbert River which are referred to in the following text. Also shown is the location of the most upstream sampling site below the gorge, at Nash’s Crossing.

7.3.1 Field parameters

Automatic recorders installed by QDNRW have collected daily water quality data at gauge

116001 since November 1999. Plots of each parameter relative to streamflow highlight distinct

seasonal patterns, especially with respect to electrical conductivity (EC) and temperature (T)

(Figure 7-8). The following sections examine trends within and between seasons for these field

parameters based on available time series data.

Page 212: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

188

1

10

100

1000

Nov

-99

May

-00

Nov

-00

May

-01

Nov

-01

May

-02

Nov

-02

May

-03

Nov

-03

May

-04

Nov

-04

May

-05

Fiel

d pa

ram

eter

s

0

100

200

300

400

500

600

700

800

Flow

(GL/

day)

EC @ 116001 pH @ 116001T @ 116001 Flow @ 116001

Figure 7-8 Time series water quality data relative to flow at gauge 116001. Vertical lines approximate the beginning of the dry season (May). Source: QDNRW

7.3.1.1 Electrical conductivity

Electrical conductivity displays a marked negative correlation with streamflow at gauge 116001

(Figure 7-9). Rapid declines in river salinity correspond with the rising stage of individual

streamflow events, with minimum EC values correlating with streamflow peaks (Figure 7-10a).

In contrast, there is an observed salinity rise after a streamflow event (falling limb of the stream

hydrograph). In addition to salinity fluctuations during streamflow events, the salinity in the

river shows a distinct seasonal trend of low EC during the wet season, rising steadily to a

maximum value by the end of the dry season (Figure 7-10b).

0

40

80

120

160

1 10 100 1000 10000 100000 1000000

Flow (ML/day)

EC

(uS/

cm)

Figure 7-9 Electrical conductivity versus flow at gauge 116001. Source: QDNRW

EC

pH

T

Flow

Page 213: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

189

0

50

100

150

200

250

Jan-

02M

ar-0

2

May

-02

Jul-0

2

Sep

-02

Nov

-02

Jan-

03

Mar

-03

May

-03

Jul-0

3

Sep

-03

Nov

-03

Jan-

04

Mar

-04

May

-04

Jul-0

4

Sep

-04

Nov

-04

Jan-

05M

ar-0

5

EC (u

S/cm

)R

ainf

all (

mm

)

1

10

100

1000

10000

100000

1000000

Flow

(ML/

day)

EC @ 116001Rainfall @ 32045Flow @ 116001

a

0

50

100

150

200

250

Jan-

02M

ar-0

2

May

-02

Jul-0

2

Sep

-02

Nov

-02

Jan-

03M

ar-0

3

May

-03

Jul-0

3

Sep

-03

Nov

-03

Jan-

04M

ar-0

4

May

-04

Jul-0

4

Sep

-04

Nov

-04

Jan-

05M

ar-0

5

EC

(uS/

cm)

Rai

nfal

l (m

m)

1

10

100

1000

10000

100000

1000000

Flow

(ML/

day)

EC @ 116001

Rainfall @ 32045Seasonal flow @ 116001

b

Figure 7-10 Continuous daily electrical conductivity at gauge 116001 relative to streamflow and rainfall: (a) depicts the actual stream hydrograph based on daily data, while (b) shows the approximate seasonal stream hydrograph (60-day moving average). Dotted circles in (b) show the three distinct clusters of EC values at particular stages along the stream hydrograph. Source: BoM (rainfall); QDNRW (flow)

The declining EC trends are consistent with flushing and dilution of dissolved salts in the river

as relatively less saline water (runoff) is introduced into the river system during streamflow

events. In contrast, concentration of dissolved solutes can occur as a result of evaporation and/or

baseflow contributions (including groundwater). Evaporation is especially plausible throughout

the course of the dry season, characterised by a progressive increase in river temperature (Figure

falling limb

peak

falling limb peakrisinglimb

WET DRY

Page 214: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

190

7-13). However, assuming that EC is proportional to TDS, the theoretical trajectory that EC

would follow under different rates of evaporation is defined by the parallel trends depicted in

Figure 7-11. Given that these evaporation lines are much steeper than the observed EC trend at

different flow rates (Figure 7-9), evaporation cannot be the dominant mechanism driving the

increase in EC as flow decreases. However, evaporation is evident in Figure 7-11 where EC

values parallel the evaporation trends. Therefore, while evaporative effects cannot be

discounted, the most likely explanation for the increasing trends in river EC, within individual

events and seasonally, is an increase in the proportion of baseflow discharged to the river

relative to surface runoff. This is consistent with the large baseflow flux determined by

hydrograph separation (Section 6.3.2.3, Chapter 6). Assuming that the EC of baseflow remains

approximately constant throughout the year, seasonal EC values in the river can be interpreted

as tracking particular stages within the seasonal stream hydrograph. As illustrated in Figure

7-10b for 2004, the lowest EC values occur during the streamflow peak, when streamflow is

dominated by surface runoff. After the peak, a higher EC is maintained in the river due to

increased proportions of baseflow contributed by streambank and groundwater discharges. The

highest EC in the river is attained by the end of the dry season, when low flows in the river are

sustained by groundwater. A similar pattern is observed for other years, although the three

stages are particularly evident in 2004 because of the dilution effect caused by high rainfall

during the 2004 wet season. Figure 7-11 further illustrates that EC in the river during

intermediate flows reflects mixing between baseflow-dominated and runoff-dominated waters.

Note that the minimum EC measured in shallow and deep groundwater samples is

approximately 100 μS/cm; above this value there is negligible surface flow contribution (Figure

7-11). EC can thus be used as an indicator of the pure baseflow component.

0

40

80

120

160

1 10 100 1000 10000 100000 1000000

Flow (ML/day)

EC (u

S/cm

)

baseflow

mixing

runoff

evaporation lines

minimum EC (gw)

Figure 7-11 Electrical conductivity versus flow at gauge 116001, depicting theoretical evaporation lines starting from different flow/EC combinations. The minimum EC measured in groundwater (gw) is indicated. Distinct clusters of EC values corresponding to baseflow-dominated, runoff-dominated and mixing between these waters, are also highlighted. Source: flow and EC data from QDNRW

Page 215: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

191

7.3.1.2 pH

River pH can be influenced by factors such as the pH of the source water, turbidity (affecting

the dissolved CO2 concentration) and mineral/weathering reactions in the river-bed/water

column. In addition to natural factors, human influences such as stream pollution can influence

the chemistry of a stream (Hem, 1985). Although pH at gauge 116001 does not display marked

seasonal fluctuations compared to the other water quality parameters (e.g. Figure 7-8), time

series data indicate that large streamflow events produce a rapid decline in pH (more acidic)

followed by a return to neutral values within a few days of the event (Figure 7-12). The pH of

rainfall generally varies from 4-6 (continental) to 5-6 (coastal) (Langmuir, 1997). Therefore,

depressed pH values during large streamflow events are consistent with a high rainfall input. In

addition, the rapid pH response to streamflow events is consistent with the contribution of low

pH waters to the Herbert River from overland flow. As the lower catchment is dominated by

sugarcane farming, which uses nitrogenous fertilisers, soil acidity can result (Wood et al.,

2003); acids can hence be leached out of the soils during runoff events. High concentrations of

ammonia are also likely to be present in the atmosphere during crop fertilisation (Freyney et al.,

1994) that can lower the pH of precipitation (Hem, 1985). The low correlation between pH and

streamflow in the dry season suggests that factors other than streamflow influence river pH

during low flows. Further analysis is beyond the scope of the thesis.

6.0

6.5

7.0

7.5

8.0

Jan-

02

Mar

-02

May

-02

Jul-0

2

Sep

-02

Nov

-02

Jan-

03

Mar

-03

May

-03

Jul-0

3

Sep

-03

Nov

-03

Jan-

04

Mar

-04

May

-04

Jul-0

4

Sep

-04

Nov

-04

Jan-

05

Mar

-05

pH

10

100

1000

10000

100000

Flow

(ML/

day)

pH @ 116001

Flow @ 116001

Figure 7-12 Continuous daily pH at gauge 116001 relative to streamflow. Source: QDNRW

Page 216: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

192

7.3.1.3 Temperature

River temperature can be influenced by numerous interrelated factors such as air temperature;

temperature of the source water (eg. rainfall, groundwater); and volume of water in the water

column. Time series temperature data highlight that the mean temperature of the Herbert River

closely tracks the mean air temperature. Troughs in stream temperature coincide with minimum

air temperatures in July. The temperature of the stream increases in the latter half of the dry

season, consistent with higher air temperatures and lower volumes of water in the river.

10

20

30

40

Nov

-99

May

-00

Nov

-00

May

-01

Nov

-01

May

-02

Nov

-02

May

-03

Nov

-03

May

-04

Nov

-04

May

-05

Tem

pera

ture

(o C)

1

10

100

1000

10000

100000

1000000

Flow

(ML/

day)

Mean air T @ 32045 River T @ 116001 Flow @ 116001

DRY SEASON

WET SEASON

Figure 7-13 Time series of mean air temperature, mean river temperature and streamflow at gauge 116001 in the lower Herbert River. Source: BoM (air T); QDNRW (river T and flow)

Given the strong positive correlation between river and air temperatures (Figure 7-14a), climate

is considered to be a dominant influence on the temperature of the Herbert River. Furthermore,

as groundwater generally has a temperature range between 25-29oC, an increase in baseflow

contribution to the stream (as discussed in Section 7.3.1.1) could be an additional factor that

influences river temperatures in the latter part of the dry season. The split regressions displayed

in Figure 7-14b for river temperatures less than 25oC, between 25-29oC, and greater than 25oC,

are consistent with groundwater buffering the mean river temperature relative to the mean air

temperature towards the end of the dry season.

Page 217: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

193

y = 0.92x + 0.82R2 = 0.82

10

20

30

40

15 20 25 30 35

Mean water T (oC)M

ean

air T

(o C)

a

y = 1.2x - 6.0R2 = 0.61

y = 0.59x + 10R2 = 0.21

y = 1.0x - 2.4R2 = 0.53

10

20

30

40

15 20 25 30 35

Mean water T (oC)

Mea

n ai

r T

(o C)

< 25 degrees

> 25-29 degrees

>29 degrees

b

Figure 7-14 Correlation between mean air and river temperatures at gauge 116001 in the lower Herbert River: (a) regression over all samples, (b) regression separated for river temperatures < 25 oC, 25-29 oC and > 26 oC. Source: BoM (air T); QDNRW (river T).

7.3.2 Major ions

7.3.2.1 Inter-seasonal trends

Trilinear plots based on historical major ion chemistry indicate that there is overlap in the

proportions of major anions in the Herbert River at each of the stream gauges (Figure 7-15).

However, the composition of the stream in the upper catchment (gauge 116004) is distinctly

more enriched in Mg compared to the lower catchment gauges (116006 and 116001). Rivers in

Queensland generally reflect geographical factors such as geology and climate (McNeil et al.,

2005); hence, the observed hydrochemical clusters are consistent with differences in geology

and soil types, which are basaltic in origin in the upper catchment and granitic in the lower

catchment. Different flow characteristics of the Herbert River in the two catchments were also

noted in Chapter 6 (Section 6.3.2.2). Figure 7-15 also highlights that the composition of the

river at gauge 116001 is more variable than at the gauges upstream. Comparison of wet and dry

season data indicate that this variation in hydrochemistry occurs during months of the wet

season. For example, Schoeller plots of samples at the downstream gauge (116001) illustrate

that during the wet season there are both Cl-enriched and Cl-depleted signatures relative to

Page 218: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

194

HCO3 (Figure 7-16a). Comparison of HCO3 and Cl with streamflow indicates that while the

concentrations of both anions decrease with increasing flow, the decrease in HCO3 is

proportionally greater (Figure 7-17). Hence, the composition of the Herbert River during the

wet season reflects the timing of sample collection relative to the flow event, with an increase in

Cl/HCO3 close to the flow peak. An increase in the Cl/HCO3 ratio is therefore a diagnostic for

overland flow. Note that Schoeller plots for dry season samples at gauge 116001 resemble the

Cl-depleted signature of the wet season (Figure 7-16b), consistent with a constant source of

water to the river throughout the year.

80 60 40 20 20 40 60 80

20

40

60

80 80

60

40

20

20

40

60

80

20

40

60

80

Ca Na HCO3 Cl

Mg SO4

Gauge 116001Gauge 116006Gauge 116004

a

Figure 7-15 Piper plot of historical major ion compositions (1973-2004) of the Herbert River at three stream gauges: 116001 (downstream) and 116006 (upstream) are located in the lower catchment while 116004 is located in the upper catchment. Source: QDNRW

a b

Figure 7-16 Schoeller plots of historical (a) wet season and (b) dry season water quality samples collected at gauge 116001 since the 1970’s. Samples highlighted in red correspond to waters that are relatively enriched in Cl relative to HCO3. Source: QDNRW

Page 219: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

195

y = -3.88Ln(x) + 33.05R2 = 0.70

0

10

20

30

40

50

60

0.01

0.10

1.00

10.0

0

100.

00

1000

.00

Flow (GL/day)

HCO

3 (m

g/L)

wet season

dry season

a

y = -0.78Ln(x) + 9.66R2 = 0.57

0

5

10

15

20

0.01

0.10

1.00

10.0

0

100.

00

1000

.00

Flow (GL/day)

Cl (m

g/L) wet season

dry season

b

Figure 7-17 HCO3 and Cl concentrations against streamflow at gauge 116001 during months of the wet and dry seasons. Source: QDNRW

Although the frequency of collection of major ion data at gauge 116001 is in general biannual,

examination of time series data highlight that anions of HCO3, Cl and SO4 display a consistent

increase in concentration during the dry season, followed by a decrease in response to wet

season rainfall. This seasonal trend is particularly apparent for HCO3, which is the dominant

anion in the river. Major cations (Na, Ca, Mg) and SiO2 display a similar seasonal trend to the

anions. The seasonal patterns are consistent with the EC trends identified in Section 7.3.1.1:

dilution occurs during runoff events as relatively fresh surface waters are introduced to the river,

while the subsequent increase in concentrations can be attributed to increased baseflow.

Page 220: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

196

7.3.2.2 Intra-seasonal trends

Similar to the inter-seasonal trends, within the freshwater zone of the river (upstream of tidal

influence), Schoeller plots illustrate that months at the beginning (May/June) and end (October)

of the dry season also have distinct chemical traces. For example, by the end of the dry season

the composition of the Herbert River shifts to higher concentrations of the major cations and

anions, except for Cl (Figure 7-18a). Although the absolute concentration of Cl does not change

significantly, there is a change in the proportion of Cl relative to HCO3 such that the Cl/HCO3

ratio decreases by the end of the dry season.

a

b

Figure 7-18 Major ion chemistry for samples collected in the lower Herbert River during the beginning (May 2004, June 2005) and end (October 2004) of the dry season. Schoeller plots in (a) are for all freshwater compositions except at Nash’s Crossing (reach 1), with end of dry season samples highlighted in red; (b) samples collected at the most upstream sampling site (Nash’s Crossing, Figure 7-2) during months at the extremes of the dry season.

Whilst a change in hydrochemistry can be attributed to numerous factors (as outlined in Section

7.1.1), the EC data examined in Section 7.3.1.1 indicated a transition from surface runoff-

dominated to baseflow-dominated flow during the course of the dry season. Furthermore, based

on physical relationships it was established in Chapter 6 that groundwater is an important source

of recharge to the river, particularly during low flow conditions. Therefore, the shift in

hydrochemical signature is consistent with a change in the composition of water discharging

into the Herbert River. Figure 7-18b illustrates that in contrast to the other reaches, the

hydrochemical signature at reach 1 (Nash’s Crossing) is similar at the two extremes of the dry

season, particularly in regards to relative anion concentrations. This suggests that baseflow

contributions downstream of Nash’s Crossing are the dominant influence on the chemistry of

the lower Herbert River at the end of the dry season, consistent with the baseflow analysis in

Chapter 6 (6.3.2.3).

May 2004October 2004June 2005

end of dry season

beginning of dry season

Page 221: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

197

7.3.2.3 Tributaries of the Herbert River

Schoeller plots for selected tributaries of the lower Herbert River are provided in Figure 7-19.

Similar to the Herbert River (e.g. Figure 7-18a), the plots indicate a shift in hydrochemical

signature at the end of the dry season compared to the beginning. An increase in the

concentration of major ions and a decrease in the Cl/HCO3 ratio are generally observed,

consistent with a change in tributary source waters during the course of the dry season. With the

exception of Hawkins Creek, the hydrochemical trends at the end of the dry season (and

beginning for most tributaries) resemble the Ca-Mg enriched groundwaters of the shallow and

deep aquifers (Figure 5-5 and Figure 5-7, Chapter 5). This observation is supported by

saturation indices, which highlight the hydrochemical similarity between groundwater

compositions and tributaries in the upper part of the catchment (Figure 7-20). Hence, as noted

for the Herbert River, groundwater is a likely source. Furthermore, the intra-seasonal

hydrochemical trends of the tributaries suggest that the change in composition of the lower

Herbert River during the dry season can be attributed to a change in the contribution of baseflow

to the river from both indirect (via tributaries) and direct groundwater discharge. Note that the

hydrochemical trends observed particularly in Hawkins Creek and Ripple Creek (at the

beginning of the dry season) resemble the Na > Cl water type in HSs (Figure 5-5c, Chapter 5).

Given the uncertainty in the origin of this groundwater type (Table 5-1, Chapter 5) it is beyond

the scope of the thesis to further assess these trends.

Page 222: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

198

Figure 7-19 Schoeller plots for tributaries of the lower Herbert River during the beginning (May 2004, June 2005) and end (October 2004) of the dry season (refer to Figure 7-2 for tributary locations). Red trend lines highlight measurements at the end of the dry season.

Page 223: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

199

-7

-6

-5

-4

-3

-2

-1

0

1

Mag

nesi

te

Cal

cite

Gyp

sum

Cha

lced

ony

Dol

omite

log

SI

71A

69A

Gow rie Ck

Elphinstone Ck

Dalrymple Ck

Figure 7-20 Saturation indices (logarithmic form) for bores and tributaries in the upper part of the catchment based on samples collected at the end of the dry season (October 2004).

7.3.3 Nitrogen

Temporal trends for nitrogen in the lower Herbert River are examined in this section based on

data from a water quality project conducted by CSIRO (Bramley and Muller, 1999). Although

dissolved inorganic nitrogen (DIN) is of particular interest to this study, total particulate N and

dissolved organic nitrogen (DON) data are also analysed in order to determine the relative

contribution of DIN to the total N load of the river. Sampling sites along the Herbert River

include Nash’s Crossing, Abergowrie Bridge, John Row Bridge and Gairloch Bridge, which lie

approximately 2 km, 44 km, 78 km and 81 km downstream of the gorge, respectively (Figure

7-2). A summary of N chemistry was presented in Chapter 5 (Section 5.1.1.4).

7.3.3.1 Particulate vs dissolved N

Analysis of time series nitrogen data (total, soluble and particulate) shows that the highest total

N concentrations in the river generally correspond to high flow events in the wet season, which

are dominated by particulate N (Figure 7-21). As observed in Figure 7-21b and also noted by

Mitchell et al. (1997), the concentration of particulate N increases with rising flow and reaches

a maximum close to the flow peak. In contrast, the dry season is generally dominated by

dissolved forms of N. Based on the available data, the proportion of soluble N in the Herbert

River relative to total N is generally between 80-90 % during the dry season months.

Page 224: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

200

0

1500

3000

4500

Jan-

1992

Jan-

1993

Jan-

1994

Jan-

1995

Tota

l N (u

g N/

L)

1

100

10000

1000000

Flow

(ML/

day)

Nash's Crossing Abergowrie bridgeJohn Row bridge Gairloch bridgeFlow @ 116001 Flow @ 116006

a

0

1500

3000

4500

Jan-

1992

Jan-

1993

Jan-

1994

Jan-

1995

Tota

l par

ticul

ate

N (u

g N/

L)

1

100

10000

1000000

Flow

(ML/

day)

Nash's Crossing Abergowrie bridgeJohn Row bridge Gairloch bridgeFlow @ 116001 Flow @ 116006

b

0

1500

3000

4500

Jan-

1992

Jan-

1993

Jan-

1994

Jan-

1995

Tota

l sol

uble

N (u

g N/

L)

1

100

10000

1000000

Flow

(ML/

day)

Nash's Crossing Abergowrie bridgeJohn Row bridge Gairloch bridgeFlow @ 116001 Flow @ 116006

c

Figure 7-21 Time series concentrations of (a) total; (b) particulate and (c) soluble N at various sites along the lower Herbert River. Flow at gauges 116006 and 116001 is also shown. Source: N data from Bramley and Muller (1999); flow data from QDNRW

Page 225: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

201

7.3.3.2 Dissolved organic vs inorganic N

Examination of dissolved N at each sampling location along the lower Herbert River indicates

that the concentration of DON is generally higher than DIN, particularly at the upstream

sampling site (Figure 7-22). DON does not show a strong seasonal pattern but is highly variable.

In contrast, there is a noticeable trend in DIN characterised by a marked increase at the

beginning of the wet season followed by a decline to a base concentration during the dry season.

Accordingly, the proportion of DIN relative to total dissolved N ranges from a maximum of 70

% during the start of the wet season, to less than 2 % by the end of the dry season: this

represents a maximum of 30% and down to 1% of total N. As illustrated in Figure 7-22 and also

observed by Mitchell et al. (1997), high DIN in the wet season corresponds to the first major

flow event (first flush) following the dry season (high availability of DIN in the catchment),

while a decrease in DIN occurs around the flow peak due to dilution by rainwater.

0

100

200

300

400

500

600

Jan-

1992

Jan-

1993

Jan-

1994

Jan-

1995

ug N

/L

1

100

10000

1000000

Flow

(ML/

day)

DIN Abergowrie bridge DON Abergowrie bridge

Flow @ 116001

a

0

200

400

600

800

1000

Jan-

1992

Jan-

1993

Jan-

1994

Jan-

1995

ug N

/L

1

100

10000

1000000

Flow

(ML/

day)

DIN John Row bridge DON John Row bridge

Flow @ 116001

b

Figure 7-22 Concentrations of dissolved N as inorganic (DIN) and organic (DON) components at (a) upstream and (b) downstream sampling sites along the lower Herbert River. Flow at gauge 116001 is also shown. Source: N data from Bramley and Muller (1999); flow data from QDNRW

Page 226: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

202

7.3.3.3 Inorganic species

The inorganic components of dissolved nitrogen are comprised of nitrate (NO3-), ammonium

(NH4+) and nitrite (NO2

-), which differ in their oxidation states (Section 5.1.1.4, Chapter 5). As

depicted in Figure 7-23, NO3- is consistently the dominant inorganic species over NH4

+ (NO2- is

generally present in very low concentrations in surface waters). Whilst NO3- in the river varies

during the wet season, the concentration is relatively constant during the dry season. This

observation is consistent with NO3- inputs from groundwater, as discussed in Section 7.6.

0

100

200

300

Jan-

1992

Jan-

1993

Jan-

1994

Jan-

1995

ug N

/L

1

100

10000

1000000

Flow

(ML/

day)

NO3 Abergowrie bridge NH4 Abergowrie bridge

Flow @ 116001

a

0

300

600

900

Jan-

1992

Jan-

1993

Jan-

1994

Jan-

1995

ug N

/L

1

100

10000

1000000

Flow

(ML/

day)

NO3 John Row bridge NH4 John Row bridge

Flow @ 116001

b

Figure 7-23 Nitrate and ammonium concentrations at (a) upstream and (b) downstream sampling sites along the lower Herbert River. Flow at gauge 116001 is also shown. Source: N data from Bramley and Muller (1999); flow data QDNRW

7.4 LONGITUDINAL DATA

The above analyses have highlighted that there are differences in the hydrochemistry of the

Herbert River and its tributaries both inter-seasonally (between the wet and dry seasons) as well

as intra-seasonally (between the beginning and end of the dry season). Notably, as streamflow

declines the:

Page 227: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

203

• salinity of the river (and tributaries) increases due to an increase in major ion

concentrations, not due to evaporation alone;

• Cl/HCO3 ratio of the river (and tributaries) decreases due to relative enrichment in

HCO3; and

• mean river temperature decreases relative to air temperature.

In addition, there is similarity between the hydrochemical signature of surface waters and Ca-

Mg-HCO3 enriched groundwaters of the shallow and deep aquifers. These collective

observations provide strong evidence of baseflow contributions to the Herbert River and its

tributaries. Moreover, consistent with the flow characteristics examined in Chapter 6 (Section

6.3.2), the available hydrochemical data demonstrate that the contribution of groundwater

increases from the wet season to the end of the dry season. Hence, there is a seasonal shift from

runoff-dominated to baseflow-dominated waters contributing to the Herbert River.

Given the above hydrochemical characteristics of surface waters, the aim of this section is to

examine chemical trends along the length of the lower Herbert River that may indicate variation

in water sources to the stream. Field parameters as well as other solutes are analysed for

longitudinal trends, both within particular sampling periods, and between different periods in

the dry season (refer to Appendix A for the entire dataset). The range of parameters analysed

provide different levels of information. Thus, EC is useful for examining broad hydrochemical

trends, whilst individual solutes yield information on small variations due to particular

processes/influences. Isotopes such as 222Rn are a particularly powerful tool for identifying

recently added groundwater sources contributing to the Herbert River (Section 7.1.1.2). Note

that no sampling was undertaken during the wet season. However, samples collected at the

beginning (May, June) and end (October) of the dry season provide an indication of

hydrochemical changes in the river as flow declines.

7.4.1 Salinity

Electrical conductivity was recorded in the field at regular intervals along the length of the

lower Herbert River and selected tributaries. As observed in Figure 7-24a, there is a marked

increase in EC within the tidal zone, which extends to a maximum of approximately 25 km

upstream from the mouth of the river (75 km downstream from the gorge). Compared to the

tidal reach there is little variation in EC along the freshwater section of the river. However,

detailed examination highlights that EC progressively increases downstream from the top of the

catchment, particularly during lower flow conditions (Figure 7-24b). This is indicative of a

source of dissolved constituents downstream from the gorge and/or evaporation.

Page 228: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

204

Note that there is a strong positive relationship between EC and TDI, hence EC is considered to

be a suitable indicator of river salinity for this population of waters during the sampling periods

(McNeil and Cox, 2000). Although the longitudinal variations are similar, there is a distinct

difference in the magnitude of salinity in the river and its tributaries between the beginning and

end of the dry seasons. These observations are discussed in the sections below in the context of

the individual constituents of salinity.

10

100

1000

10000

100000

0 20 40 60 80 100

Distance downstream from gorge (km)

EC (u

S/c

m)

October 2004June 2005

tidal limit

a

40

80

120

160

200

240

0 20 40 60 80

Distance downstream from gorge (km)

EC (u

S/cm

)

Oct-04

Jun-05Tribs - Oct

Tribs - June

21 3 4 6 8 10 12 15 17

b

Figure 7-24 Measurements of field electrical conductivity along the lower Herbert River and its tributaries (Tribs) during the beginning (June) and end (October) of the dry season. Note that (b) excludes the high salinity section of the river; selected reach numbers corresponding to sample locations are provided on the upper horizontal axis.

Page 229: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

205

7.4.2 Major ions in the tidal zone

In general, there is little longitudinal variation observed in the concentrations of major ions

within the freshwater zone of the river. However, consistent with the EC measurements, there is

a marked increase in the concentrations of most constituents within the tidal zone, except HCO3

and SiO2 (Figure 7-25). Similarities with the relative proportions of major ions in seawater (e.g.

Figures 7-3, 7-4, 7-5) confirm that the high solute concentrations are due to mixing with

seawater. In addition, the dramatic increase in the Mg/Ca ratio is consistent with a seawater

influence (Figure 7-26) (Vengosh, 2004).

June 2005

0.1

1.0

10.0

100.0

1000.0

10000.0

0 20 40 60 80 100

Distance downstream from gorge (km)

Con

cent

ratio

n (m

g/L) HCO3

ClSO4NaCaMgKSiO292 km

HCO3

SiO2

tidal limit

October 2004

0.1

1.0

10.0

100.0

1000.0

10000.0

0 20 40 60 80 100

Distance downstream from gorge (km)

Con

cent

ratio

n (m

g/L) HCO3

ClSO4NaCaMgKSiO286 km

HCO3

SiO2

tidal limit

Figure 7-25 Longitudinal comparison of major ion and oxide concentrations along the lower Herbert River during months representing the beginning (June) and end (October) of the dry season. Seawater influence is observed in samples downstream of the vertical dotted lines. The longitudinal trends for HCO3 (green) and SiO2 (dark blue) are also labelled.

Page 230: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

206

Within the tidal zone, the extent of seawater influence on the chemistry of the Herbert River is

determined by the river stage and sampling time in relation to the tidal patterns (relative sea

level). Hence, the location and chemistry of the mixing zone varies seasonally as well as daily.

For example, in samples from October 2004 mixing is evident downstream of 86 km (reach 20),

while in June 2005 seawater influence is apparent at 92 km (reach 23) downstream from the

gorge (Figure 7-25 and Figure 7-26).

The near-constant concentrations of HCO3 and SiO2 upstream of, and within the tidal zone, are

due to the relatively small difference in concentration of these dissolved constituents in seawater

compared to freshwater. For instance, the amount of SiO2 in seawater is approximately 5 mg/L

compared to around 6-7 mg/L in freshwaters of the river; therefore, in the zone of seawater

mixing there is only a minor decrease (dilution) in SiO2. Conversely, HCO3 is slightly enriched

in seawater (142 mg/L) compared to freshwater (20-40 mg/L); therefore, a minor increase in

concentration is observed in the tidal zone. In comparison, the concentrations of the other major

ions are at least two orders of magnitude greater in seawater than in freshwater of the Herbert

River, resulting in a marked change in their concentrations where fresh and saline waters mix.

Although longitudinal variations in major anions, cations and SiO2 within the freshwater

reaches are relatively small in magnitude, they are important for detecting variations in source

waters contributing to the river. Concentration differences between higher flow and lower flow

periods also provide insight into different processes influencing the lower Herbert River during

the extremes of the dry season, as examined below.

0.0

1.0

2.0

3.0

4.0

5.0

6.0

1 2 3 4 5 6 7 8 9 10 11 12 14 15 16 17 18 20 21 23 24

Reach number

Mg/

Ca (m

eq/L

)

October 2004June 2005

Figure 7-26 Mg/Ca ratio for samples collected along the lower Herbert River during October 2004 and June 2005. Reaches are numbered from the gorge to the river mouth (refer to Figure 7-2).

Page 231: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

207

7.4.3 Processes in the freshwater zone

As discussed in Section 7.1.1, there are a number of processes that can influence the chemistry

of the Herbert River. Along the freshwater reaches, the key processes include evaporation,

overland flow, tributary inflow, and groundwater discharge (Table 7-1). The analyses below aim

to distinguish between these processes based on major ion and isotope trends. Whilst

concentrations were measured for a number of dissolved constituents, longitudinal trends are

illustrated for selected solutes and their ratios that demonstrate the main hydrochemical

variations along the river attributed to the key processes. Note that although the downstream

extent of the freshwater zone is variable between sampling periods (e.g. Figure 7-25 and Figure

7-26), for ease of comparison longitudinal plots representing freshwater compositions are

depicted from the gorge (reach 1) to 86 km downstream (reach 18).

7.4.3.1 Evaporation

An increase in salinity and individual ion concentrations, constant ion ratios, and isotopic

enrichment, are indicative of evaporation (Table 7-1). It was noted in Section 7.2.2 that

enrichment in stable isotopes is observed between some consecutive reaches in May,

particularly between reaches 1-2 and 9-14 (Figure 7-27). Although the Na/Cl ratio is virtually

constant between reaches 1-2, the δ2H/δ18O slope between these reaches is -1.3, compared to a

typical evaporation slope from open waters in the range 4-6 (Gat, 1980). In contrast, between

reaches 9-11, the δ2H/δ18O slope is 4.6, the corresponding TDI increases, and Na/Cl is virtually

constant over the reach (Figure 7-27). However, the δ2H/δ18O slope between reaches 11-14 is

10.8, which is incompatible with evaporation. Therefore, whilst there is evidence for

evaporation along at least one section of the river in May (e.g. between reaches 9-11), based on

the available data evaporation is not considered to be a major influence on the chemistry of the

river at the beginning of the dry season. It was noted in Section 7.3.1 that evaporative effects

would be expected to increase towards the end of the dry season as a result of an increase in

river and air temperatures. However, while TDI generally increases from upstream to

downstream in October, ion ratios such as Na/Cl are not constant between consecutive reaches

displaying large salinity increases (e.g. between reaches 1-2 in October) (Figure 7-27). This is

consistent with the conclusions from the temporal EC analyses in Section 7.3.1.1: whilst

evaporation cannot be discounted, it is not the major mechanism driving the increase in river

salinity through the dry season. The collection of complementary stable isotopic data at the end

of the dry season would assist in quantifying the evaporative influence along different reaches

of the river.

Page 232: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

208

40

50

60

70

80

90

100

0 20 40 60 80

Distance downstream from gorge (km)

TDI (

mg/

L)

May-04

Oct-04

21 3 4 6 8 10 12 15 17

a

0.5

1.0

1.5

2.0

2.5

3.0

0 20 40 60 80

Distance downstream from gorge (km)

Na/

Cl (

meq

/L)

May-04

Oct-04

21 3 4 6 8 10 12 15 17

b

-5.5

-5.0

-4.5

-4.0

-3.5

0 20 40 60 80

Distance downstream from gorge (km)

18O

(o / oo S

MO

W)

May-0421 3 5 9 11 14 16 18

c

Figure 7-27 Salinity, Na/Cl and oxygen-18 (δ18O) along the lower Herbert River. Reach numbers corresponding to each sample location are provided on the upper horizontal axis.

Page 233: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

209

7.4.3.2 Overland flow

Key diagnostic features of overland flow include a decrease in salinity and dilution of ions

present in the stream prior to the runoff event. In addition, it was noted in Section 7.3.2 that the

Cl/HCO3 ratio is related to flow such that the ratio increases close to the flow peak. Whilst

sampling was undertaken during periods in the dry season, with little or no expected overland

flow, rainfall occurred during the sampling period in June 2005: mean discharge increased by

approximately 200 ML/day at the upstream gauge (116006) and 2000 ML at the downstream

gauge (116001). In contrast, in May and October 2004 sampling was undertaken during the

declining phase of the stream hydrograph, with there being no major runoff events throughout

the periods of sample collection. The effect of the flow event in June 2005 on the chemistry of

the Herbert River is clearly evident as a trough (dilution) in longitudinal plots between reaches

7-10, whereby the concentrations of all constituents except for Cl and SO4 decline compared to

adjacent reaches (Figure 7-28). In contrast, all solute concentrations increase within this section

of the river during May and October 2004. As Cl and SO4 do not similarly decrease, this

suggests that there is a source of these solutes in the contributing catchment area; the ions are

thus readily mobilised during a runoff event. The ratio of Cl/HCO3 also increases over this reach

in June which provides further evidence of overland flow.

0.2

0.3

0.4

0.5

0 20 40 60 80

Distance downstream from gorge (km)

meq

/L

Na (June)

HCO3 (June)

Cl (June)

21 3 4 6 8 10 12 15 17

reaches 7-10

Figure 7-28 Longitudinal trends for selected major ions along the lower Herbert River in June 2005. Reach numbers corresponding to each sample location are provided on the upper horizontal axis, with reaches 7-10 highlighted.

Page 234: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

210

7.4.3.3 Tributary inflow

Numerous tributaries enter the Herbert River, particularly in the upper reaches which

collectively drain a significant area of the catchment. Based on saturation indices and Schoeller

plots it was observed that there is similarity between the hydrochemistry of groundwater and

most tributaries, especially at the end of the dry season (Section 7.3.2.3). In addition, high radon

activities in measured tributaries (Figure 7-29) indicate a groundwater supply during low flow

conditions (Section 7.1.1.2). Therefore, where a tributary is groundwater-fed, it is difficult to

distinguish between direct groundwater discharge and tributary inflow as the source of water to

the Herbert River. Nonetheless, visual comparisons between solute concentrations within the

tributaries and upstream and downstream of the entry point into the river indicate that

decreasing Cl between reaches 1-4 is consistent with inflow of low Cl tributaries such as

Gowrie Creek and Elphinstone Creek (Figure 7-30a). However, additional sources (other than

Gowrie Creek) must be invoked to account for the concentrations of solutes such as HCO3 and

Mg between reaches 1-2 of the river (Figure 7-30b, c). These sources could include the other

tributaries south of the river (Figure 7-2) and/or direct groundwater discharge.

0.0

1.0

2.0

3.0

4.0

5.0

6.0

Reach number

222 Rn

(Bq/

L)

21 3 4 6 8 10 11 12 145 7 9 15 16 17 18

October 2004

Dal

rym

ple

Ck

Haw

kins

Ck

Ston

e R

Rip

ple

Ck

Figure 7-29 Radon (222Rn) activities along the freshwater reaches of the lower Herbert River and sampled tributaries (blue) in October 2004. Reach numbers on the horizontal axis relate to the river, while sampled tributaries are individually labelled.

Assuming that the solute changes in the Herbert River are due entirely to tributary inflow, the

required salt load (TDI x flow) contributions from the tributaries can be quantified to test

whether additional sources must be invoked. Based on the TDI in measured tributaries and the

change in TDI of the river (upstream and downstream of each tributary), the required discharge

from the tributaries to account for the river solute loads is estimated (Table 7-2). In the absence

of flow gauges on the tributaries (with the exception of Gowrie Creek), tributary flow rates are

estimated based on field measurements of flow velocity (assuming 0.1 m/s if not visibly

flowing) and river depth and width (assuming a triangular cross-sectional area of the river).

Page 235: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

211

0.1

0.2

0.3

0.4

0.5

0.6

Reach number

Cl (

meq

/L)

21 3 4 6 8 10 11 12 145 7 9 15 16 17 18

October 2004

Gow

rie C

k

Elph

inst

one

Ck

Dal

rym

ple

Ck

Haw

kins

Ck

Ston

e R

Rip

ple

Ck

a

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

1.1

Reach number

HCO

3 (m

eq/L

)

21 3 4 6 8 10 11 12 145 7 9 15 16 17 18

October 2004

Gow

rie C

k

Elph

inst

one

Ck

Dal

rym

ple

Ck

Haw

kins

Ck

Ston

e R

Rip

ple

Ck

b

0.0

0.1

0.2

0.3

0.4

0.5

Reach number

Mg

(meq

/L)

21 3 4 6 8 10 11 12 145 7 9 15 16 17 18

October 2004

Gow

rie C

k

Elph

inst

one

Ck

Dal

rym

ple

Ck

Haw

kins

Ck

Ston

e R

Rip

ple

Ck

c

Figure 7-30 Longitudinal plots for selected ions along the freshwater reaches of the lower Herbert River and sampled tributaries (blue) in October 2004. Reach numbers on the horizontal axis relate to the river, while sampled tributaries are individually labelled.

Page 236: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

212

Table 7–2 Required and estimated flow rates in tributaries of the lower Herbert River and groundwater based on changes in total solute loads upstream and downstream of selected tributaries in October 2004

Tributary TDI of tributary

(mg/L)

Δ solute load in Herbert R (106 mg/day)+

Required tributary discharge (ML/day)

Estimated tributary discharge

(ML/day)*

Required gw input

(ML/day)#

Gowrie Ck 71 5100 70 20 38

Elphinstone Ck 62 8300 130 80 36

Dalrymple Ck 91 -300 -3 10 0

Stone R 140 3500 30 60 0

Ripple Ck 120 300 3 4 0 + Defined as downstream minus upstream load estimated from Herbert River gauges (Figure 7-7) * Estimated from field measurements of flow velocity and river width and depth (gauge value assumed for Gowrie Ck) # TDI of contributing aquifer (HSd) is 96 mg/L (near Gowrie Ck) and 93 mg/L (near Elphinstone Ck)

The calculations indicate that tributaries such as Gowrie Creek and Elphinstone Creek most

likely contribute to the solute load in the Herbert River; however, additional sources are

required to explain the total solute loads in the river. Assuming a groundwater contribution, the

required groundwater input to account for the solute loads in the river is estimated to be around

40 ML/day. While there are additional tributaries above and below the confluence of Gowrie

Creek within reach 2 (Figure 7-2) that could be supplied by groundwater, there are no major

tributaries other than Elphinstone Creek within reach 4; hence, direct groundwater discharge to

the river is plausible. The mass balance indicates that inflow of Dalrymple Creek does not

contribute to the change in total solute load in the river. Note that Hawkins Creek was not

included in the calculations as the TDI is well below that of the upstream and downstream

concentrations in the river; the tributary also has a very small catchment area (Figure 7-2) and

would be unlikely to contribute to noticeable flow (or solutes) in the river at the end of the dry

season. However, the calculations indicate that while the estimated discharge rates are small,

solute loads from the Stone River and Ripple Creek are consistent with loads in the Herbert

River. Furthermore, in the absence of tributaries in between reaches 8-17, the increase in solute

loads must be attributed to other sources, such as direct groundwater discharge or evaporation.

7.4.3.4 Groundwater discharge

Based on hydrochemical similarities between surface water and groundwater (Section 7.2) and

temporal chemical trends in the river (Section 7.3) it was established that groundwater is a

source of water to the river, with an increasing relative contribution towards the end of the dry

season. Furthermore, longitudinal plots along the freshwater reaches of the river highlight a

marked change in the concentration of solutes between reaches 1 and 2 (3-34 km downstream),

particularly in October, which cannot be accounted for by evaporation alone (e.g. Figure 7-27

Page 237: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

213

and Figure 7-31). In the absence of overland flow during this period, other processes must be

invoked. Similarly, it has been shown that variations in ion concentrations and ratios along other

reaches are not consistent with evaporation as the dominant processes influencing the chemistry

of the river at the end of the dry season. Based on the available information, groundwater

discharge is the most likely mechanism, either directly or via groundwater-fed tributaries. In

addition, longitudinal trends in the concentration of radon (refer to Section 7.5) demonstrate that

groundwater contributes over the entire length of the Herbert River during months at the

extremes of the dry season. Evidence for changes in the groundwater inflow rate along the river

is examined in Section 7.5.

Variations in ion concentrations in the river are evident between reaches 6-11, particularly in

October, for Na, Mg and HCO3 (Figure 7-31). Similar oscillations in cation concentrations are

also observed in June; however, due to the flow event the trends are subdued. In addition, the

concentration of Cl gradually increases downstream of reach 5 in May, June and October. These

subtle changes in chemistry are indicative of changes in source waters contributing to the river.

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0 20 40 60 80

Distance downstream from gorge (km)

Na

and

HC

O 3 (

meq

/L)

0.1

0.2

0.3

0.4 Na (Oct)

HCO3 (Oct)

Cl (Oct)

Mg (Oct)

21 3 4 6 8 10 12 15 17

Cl a

nd M

g (m

eq/L

)

reaches 6 -11

Figure 7-31 Longitudinal trends for selected major ions along the lower Herbert River in October 2004. Reach numbers corresponding to each sample location are provided on the upper horizontal axis, with reaches 6-11 highlighted.

Due to the complex stratigraphy and variable degree of interconnectivity between the aquifers

(Chapter 4), it is difficult to isolate which aquifer(s) is the dominant influence on the river along

particular reaches. However, given that HSd is the main aquifer in the northwest upland area,

where it is also incised by the Herbert River (Figure 4-4) and has the potential to discharge to

the river (Table 6-6), it can be inferred that the deep aquifer discharges into the upper reaches of

the river. However, further downstream, in the vicinity of reach 6, the channel depth decreases

such that the river only intersects the shallow aquifer. Moreover, as discussed in Chapter 5

(Section 5.2.3.2), there is evidence of an additional deeper sand unit in the shallow aquifer that

Page 238: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

214

is more enriched in Na, Mg, HCO3 and Cl than the upper unit of HSs. Hence, the enriched and

depleted concentrations of these ions within reach 6-11 are consistent with relative changes in

the contributing sand units of the shallow aquifer. Based on the available evidence, the changes

in river chemistry downstream of reach 5 are consistent with spatial variation in the

hydrogeology and hence hydrogeochemistry.

7.5 TRACING GROUNDWATER

The application of radon as a tracer of groundwater discharge to surface waters was discussed in

Section 7.1.1.2. Due to potential losses of radon through processes such as gas exchange with

the atmosphere and radioactive decay, the radon content of a stream represents the balance of

additions and losses. Using the formulation of Cook et al. (2004) and under the assumption that

radon contributions to the stream are solely from groundwater inflow (either directly or via

groundwater-fed tributaries), changes in radon content in the stream can be expressed as:

dwckwcIcxc

i λ−−=∂∂Q

(7-2)

where c is the radon activity in the stream; ci is the activity in groundwater inflow; Q is the

streamflow rate (m3/day); I is the groundwater inflow rate per unit of stream length (m3/day/m);

k is the gas transfer velocity across the water surface (m/day); λ is the radioactive decay

constant (day-1); w is the width of the river surface (m); d is the mean stream depth; and x is the

distance in the direction of flow. In the absence of surface water inflow or direct rainfall input,

change in flow with distance is given by:

EwIx

−=∂∂Q

(7-3)

where E is the evaporation rate (m/day). Hence, the activity of radon with distance becomes:

cdwkwcwEcccIxcQ i λ−−+−=∂∂ )( (7-4)

The last three terms in equation 7-4 represent changes in activity due to evaporation (increases

radon activity in residual water), gas exchange (decreases radon activity), and radioactive decay

(decreases radon activity). Assuming that for a particular reach of the river the groundwater

inflow rate, evaporation rate and gas transfer velocity do not change with distance, the change in

radon activity over the reach can be expressed as:

λλ dwkwwEIIc

wExIxQQ

dwkwwEIIc

cxc i

p

i

++−+⎟⎟

⎞⎜⎜⎝

⎛−+

⎟⎠⎞

⎜⎝⎛

++−−=

o

oo)( (7-5)

Page 239: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

215

where wEI

dwkwwEIp−

++−=

λ and Qo is the river flow at the start of the river reach (x = 0).

The above mass balance equations highlight that there are a number of parameters that influence

the net radon flux in the river. The following analyses examine longitudinal trends in radon in

light of these considerations.

7.5.1 Radon distribution in groundwater

Samples were collected at selected bores in the study area for the measurement of 222Rn in

groundwater. Whilst the concentration of 222Rn is relatively constant between sampling periods

(October 2004 and June 2005) there is considerable spatial variation. The ultimate source of 222Rn to groundwater is likely to be uranium-bearing minerals in the granite bedrock that

underlies the alluvial aquifer system. As 222Rn is a gas, the flux to groundwater is related to the

travel time (diffusion rate) of radon from the source. Greater travel distances and retardation

caused by impeding layers (e.g. clays) would be expected to decrease the vertical 222Rn flux due

to radioactive decay. Therefore, aquifer thickness and characteristics such as the degree and

direction of vertical connectivity are plausible factors that could influence the measured radon

content in groundwater. In general it is observed that the average concentration of radon in

groundwater of the shallow aquifer decreases from an average of around 58 Bq/L in the upper

part of the catchment to 9 Bq/L towards the coast. In addition, there is an increase in thickness

of the alluvial profile (Section 4.2.2, Chapter 4) as well as strong upwards vertical flow (Section

5.3, Chapter 5,) towards the coast. Whilst it is beyond the scope of the thesis to quantify in

detail the competing processes that can account for the spatial distribution of radon in

groundwater, the observed decline in concentration is important for interpreting longitudinal

changes in the activity of 222Rn. For example, low concentrations in the river do not necessarily

translate to a proportionately low flux of groundwater (or vice versa), as the input concentration

varies along the river.

7.5.2 Temporal trends in radon along the river

Longitudinal measurements of radon in the lower Herbert River indicate that months at the two

extremes of the dry season display different trends (Figure 7-32). In June there is a continual

decline in the concentration of 222Rn along the entire length of the river, with a marked decrease

downstream of reach 6 (62 km). In contrast, the October trend is characterised by an increase in

the flux of 222Rn for approximately 58 km downstream from the gorge (reaches 1-5), with a

progressive decline below reach 10 (70 km) towards the mouth of the Herbert River.

Furthermore, the concentration of 222Rn just below the gorge is greatly elevated in June

compared to October.

Page 240: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

216

0.0

0.5

1.0

1.5

2.0

0 20 40 60 80 100

Distance downstream from gorge (km)

222 R

n (B

q/L)

OctoberJune

9 10 12 16 18 20 22 2421 3 5 6 8

prior to flow event

dilution after flow event

peak flow at gauge 116006

rising stage at gauge

peak flow at gauge 116001

prior toflow event

gauge 116006

gauge 116001

(tidal limit)

Figure 7-32 Concentration of radon (222Rn) along the lower Herbert River during periods representing the end (October 2004) and beginning (June 2005) of the dry season. The approximate position of the stream gauges in reaches 2 and 12 are also indicated. Reaches in June are labelled with reference to the stages of the streamflow event that occurred during the sampling period.

Disparity between the June and October radon trends can largely be explained by the difference

in river discharge at the extremes of the dry season. As noted in Section 7.4.3.2, samples in June

were also influenced by the streamflow event that occurred during the sampling period. Prior to

the event, mean daily discharge at gauge 116006 (reach 2) in June was 770 ML/day compared

to 116 ML/day in October 2004; at gauge 116001 (reach 12) mean daily discharge was 1380

ML/day in June compared to 450 ML/day in October. Whilst higher discharge and

corresponding river width may result in greater loss of radon due to gas exchange (equation 7-

4), there is a greater reach contribution of radon at higher flow rates (due to the extra distance

that a molecule of radon can travel prior to decay). Therefore, the elevated concentration of 222Rn in the upper and lower reaches (reaches 1-3 and 12-24) at the beginning of the dry season

(June) compared to the end (October) is consistent with differences in stream discharge. A

greater presence of clays in the riverbed, contributing additional radon to the stream, would also

be expected in June (following the wet season) compared to October. Furthermore, samples in

June corresponding to reaches 1 and 2 were collected just after the peak flow recorded at gauge

116006, which may have additionally enhanced the in-stream radon concentration. Note that the

downstream samples (reaches 12-24) in June were collected during the rising stage of the stream

hydrograph and at peak flow recorded at gauge 116001 (Figure 7-32). Therefore, whilst river

discharge in the absence of rainfall is inherently higher at the beginning of the dry season

compared to the end, it is not conclusive whether the observed in-stream concentrations of 222Rn

in June are elevated in the tidal zone due to the flow event. Further sampling downstream of

Page 241: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

217

reach 11 at the beginning of the dry season (in the absence of a flow event) would be required to

confirm this.

In the middle reaches of the Herbert River the longitudinal radon trends cross over in June and

October. Samples in June corresponding to reaches 7-10 were collected two days after the peak

of the flow event, which resulted in an increase in mean discharge of approximately 2000

ML/day at gauge 116001 (reach 12). Note that the increase was of far greater magnitude at the

downstream gauge compared to the upstream gauge (increase of 300 ML/day at gauge 116006).

Therefore, as also observed with major ions (Figure 7-28), the radon concentrations between

reaches 7-10 reflect dilution from overland flow. The reversal in seasonal trends at reach 5 is

consistent with the higher 222Rn concentration of the inflowing tributary (Dalrymple Creek) in

October (4.5 Bq/L) than June (1.9 Bq/L). Additional tributary sampling would be required to

verify if the 222Rn concentration of Elphinstone Creek (between reach 3 and 4) is also elevated

at the end of the dry season.

7.5.3 Relative flux of groundwater along the river

Whilst 222Rn is a useful tracer of groundwater discharge to surface waters, there are numerous

factors that can influence the measured in-stream concentration, that are not necessarily related

to river-aquifer interactions. However, based on the approach of Cook et al. (2004), the

groundwater flux along a reach can be estimated under the assumptions outlined for equation 7-

5. Whilst several of the parameters are not accurately known, the estimates provide an

indication of the relative contribution over the specified reaches and whether a groundwater

source can account for the observed radon concentrations at the beginning and end of a reach.

This semi-quantitative approach is particularly useful given that the measured radon activity in

surface waters is not necessarily proportional to the flux of groundwater. Given that radon

samples in June were influenced by a significant flow event, measurements of radon from

October are considered to be more reliable for examining changes in groundwater discharge

along the river. However, for comparison, estimates in June are also generated for reaches

sampled prior to the flow event.

Inputs into equation 7-5 include a combination of assumed and measured values. The

evaporation rate, gas transfer velocity and decay are assumed to be constant at E = 7mm/day,

k = 1 m/day and λ = 0.18 day-1, respectively (Cook et al., 2004); mean river width, mean depth,

and in-stream radon concentrations (initial and final) for each reach were determined in the

field. The distance between sampling points was determined from a GIS. Given the variability

of 222Rn in groundwater, the input concentration is taken as an average value for bores centered

on a particular reach. Although HSd is the dominant aquifer in the upper reaches of the river, for

simplicity a similar flux is assumed from each aquifer. Therefore, an average radon

concentration of HSd and HSs is used. However, in the lower reaches, only HSs concentrations

Page 242: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

218

are used. Given the available data these assumptions are considered reasonable, which also

allow, in a relative sense, for the dramatic decline in the concentration of radon potentially

discharging to the lower reaches of the river (Section 7.5.1). The streamflow rate was only

recorded at the two gauges along the river (Figure 7-2). Hence, the input value is the average of

the gauge discharge values (on the days of radon sampling for a particular reach), unless the

reach is upstream or downstream of the gauge, in which case the nearest corresponding gauge

value is assumed. Note that within the tidal zone, radon activities have been corrected where

mixing with seawater (dilution of 222Rn) is apparent from major ion analysis. The mass balance

method described in Cook et al. (2004) using electrical conductivity measurements was applied

to determine the proportion of seawater mixing and hence the undiluted concentration of radon

in the stream. Input parameters for the determination of groundwater fluxes are displayed in

Table 7-3.

Table 7–3 Measured input parameters for modelling radon activities in the river (equation 7-5). Refer to text for assumed input values for E, k and λ.

Reach co (Bq/L) ci (Bq/L) Qav (m3/day)+ w (m) d (m)

1-2 0.57 55 120000 53 0.5

2-5 0.81 58 270000 50 0.3

5-6 1.5 35 300000 50 0.4

6-8 1.1 35 290000 50 0.4

8-9 1.4 35 410000 80 0.5

9-10 1.1 35 390000 30 0.4

10-14 1.4 33 390000 50 0.5

14-17 0.74 30 390000 100 1.7

17-19 0.28 16 390000 100 2.4

19-24 0.43 9.2 390000 100 1.6 + Note that Qav is an average of the gauge values on the day(s) that radon samples were collected along a particular reach, which is not necessarily the stream discharge rate at the start of the reach.

An approximate groundwater inflow rate was determined for each reach (using equation 7-5) in

order to minimise the difference between the calculated and measured downstream radon

activities. Discharge estimates compared with measured radon concentrations in October

(Figure 7-33) illustrate that the observed radon trend can be accounted for by changes in the flux

of groundwater. Within the freshwater zone, increases in radon concentration are consistent with

an increase in groundwater flux, while a decrease in radon concentration along a reach

corresponds with a decrease in the groundwater discharge rate. However, downstream of reach

6, the magnitude of the change in groundwater flux between reaches is greater than that for

radon. This is largely due to the fact that a large change in groundwater discharge is required to

Page 243: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

219

compensate for a small change in the concentration of radon. i.e. the estimated groundwater flux

is sensitive to the in-stream radon activity. Note that the groundwater flux estimated in June for

selected reaches sampled prior to the flow event (Figure 7-32) is similar to October estimates

between reaches 2-5; however, between reaches 5-6 the flux is greatly elevated in June

compared to October.

0

1

2

3

4

5

0 20 40 60 80 100

Distance downstream from gorge (km)

222 R

n (B

q/L)

0

1

2

3

4

5

Gro

undw

ater

flux

(m3 /d

ay/m

)

Rn (Oct)

groundwater flux (Oct)

groundwater flux (June)

9 10 1214 17 19 2421 5 6 8tidal limit

seawater mixing

Figure 7-33 Measured radon concentrations and estimated groundwater flux along the lower Herbert River in October 2004 and at selected reaches (2-5 and 5-6) in June 2005 based on the approach of Cook et al. (2004). Reach numbers are displayed on the horizontal axis, with the tidal limit and zone of seawater mixing (determined from major ions in October 2004) indicated.

In the tidal zone the reverse trend is observed compared to upstream, whereby the concentration

of radon declines, while the flux of groundwater increases. This trend arises because of the

increase in river width and depth (at least double) downstream of reach 14 compared to

upstream; a large increase in groundwater flux is thus required to account for the increased

losses from gas exchange and decay (equation 7-4). Sensitivity analysis indicates that the

estimated groundwater inflow is particularly sensitive to river width. In addition, the radon

concentration of groundwater declines within the tidal zone, especially downstream of reach 17

(Table 7-3); sensitivity analysis indicates that the groundwater flux is also responsive to this

parameter. In a study by Cook et al. (2004) it was also found that estimated groundwater

inflows are most sensitive to changes in radon activity of groundwater inflow, followed by the

gas exchange velocity and river width.

The groundwater flux estimates in the freshwater zone can be compared with gauged stream

discharge measurements during the October 2004 sampling period. Gauges 116006 and 116001

are located at approximately the beginning of reaches 2 and 12 (Figure 7-2). During the period

of interest, the difference in recorded discharge between the gauges is approximately

Page 244: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

220

240 ML/day. Given minimal overland flow during October 2004, this discharge difference

represents the baseflow contribution. Correspondingly, the summation of groundwater flux

estimates between reaches 2-12 is around 10 m3/day/m, which equates to around 400 ML/day

(40 km distance between the gauges). Whilst the estimated groundwater contribution from the

radon mass balance is greater than that determined from the difference in streamflow between

the gauges, the two results are of a similar magnitude.

The results of the above radon mass balance approach indicate that groundwater discharges

along the entire length of the lower Herbert River at the end of the dry season, either directly or

via groundwater-fed tributaries. This is consistent with the physical river-groundwater

relationships established in Chapter 6 (Table 6-6), which showed that there is potential for

hydraulic connection along the four main reaches (A-D) and that the dominant direction of flux

is from the aquifers to the river. Whilst the estimated discharge rate is underpinned by numerous

assumptions, the approach highlights the relative changes in groundwater inflow along the river,

with there being a considerably greater flux between reaches 6-10.

7.5.3.1 Comparison with hydrochemistry

Hydrochemical data coupled with the groundwater discharge estimates provide a powerful tool

to verify and describe the river-groundwater relationships along the length of the river. For

example, over the first 58 km of the river downstream from the gorge (reaches 1-5), there is a

marked decrease in Mg/Ca and increase in Na/Cl in the river in October, coinciding with an

increasing trend for 222Rn activity (Figure 7-34). These trends are indicative of increasing

contributions from groundwater sources below the gorge, which have a lower Mg/Ca ratio and

higher Na/Cl ratio than water contributing from the upper catchment. Furthermore, the large

decrease in Mg/Ca is consistent with dilution of Mg/Ca enriched waters associated with basalts

in the upper catchment (e.g. gauge 116004 in Figure 7-15) by lower catchment alluvial

groundwaters. The increase in estimated groundwater flux below the gorge is thus consistent

with the observed hydrochemistry. Similarly, in the middle reaches of the river, changes in ion

ratios (particularly Mg/Ca) coincide with changes in groundwater flux. As noted in Section

7.4.3.4 these hydrochemical variations are indicative of different contributing aquifers. The

corresponding groundwater flux estimates further indicate that there is a change in the river-

aquifer hydraulics such that discharge from groundwater increases downstream of reach 6. Note

that in Chapter 6 (Section 6.3.2.3) it was also established from baseflow filtering that there is an

enhanced contribution of baseflow between the two gauges in the study area. These collective

observations are consistent with a switch from HSd to HSs as the dominant contributing aquifers.

Page 245: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

221

0.8

1.0

1.2

1.40 20 40 60 80 100

Distance downstream from gorge (km)

Mg/

Ca (m

eq/L

)

0

1

2

3

4

5

Gro

undw

ater

flux

(m3 /d

ay/m

)22

2 Rn a

ctiv

ity (B

q/L)

Mg/Cagroundwater fluxRn (Oct)

tidal limit

seawater mixing

15 17 19 249101221 5 6 83 21

a

0.5

1.0

1.5

2.0

2.5

3.0

0 20 40 60 80 100

Distance downstream from gorge (km)

Na/C

l (m

eq/L

)

0

1

2

3

4

5

Gro

undw

ater

flux

(m3 /d

ay/m

)22

2 Rn a

ctiv

ity (B

q/L)

Na/Clgroundwater fluxRn (Oct)

tidal limit

seawater mixing

15 17 19 249101221 5 6 83 21

b

Figure 7-34 Ions ratios, estimated groundwater flux and radon along the lower Herbert River in October 2004. Note that in (a) the vertical scale is reversed for Mg/Ca and that only concentrations along the freshwater reaches are depicted (Mg/Ca increases dramatically in the zone of seawater mixing).

Based on major ion trends it is difficult to verify whether the estimated groundwater flux in the

tidal zone is plausible, as mixing with seawater dramatically alters the solute concentrations in

the river and hence masks the groundwater signal. However, in Chapter 4 (Figure 4-7) it was

established that the vertical head gradient between the aquifers reverses from downwards to

dominantly upwards in the vicinity of reach 17. Hence, the increase in estimated groundwater

discharge in the tidal zone is consistent with the change in inter-aquifer hydraulic relationships.

Page 246: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

222

7.5.4 Uncertainty in groundwater flux

Longitudinal analyses presented above for a variety of environmental tracers have highlighted

that the chemistry of the lower Herbert River changes markedly from the most upstream

location studied (Nash’s Crossing) to approximately 34 km downstream (reach 2), particularly

at the end of the dry season. Moreover, the above analyses have shown that the dominant

process leading to this shift in hydrochemistry is discharge of groundwater to the river. The

groundwater flux along reach 2 of the river was estimated in Section 7.5.3 based on a mass

balance approach using radon. A different mass balance, incorporating the concentrations of

major and minor elements, is examined in this section in order to: (1) quantitatively assess

whether groundwater can explain the observed changes in ion concentrations; (2) compare the

relative differences in baseflow contributions to the river at the extremes of the dry season; and

(3) compare the results of two independent mass balance approaches (involving solute

concentrations or radon), in order to examine the uncertainty in the estimated groundwater flux.

7.5.4.1 Solute mass balance

Applying the principle of conservation of mass and assuming conservative mixing between two

end-members (Section 7.1.1.4), the relationship between river and groundwater concentrations

for the same chemical constituent can be expressed as:

[solute]down = x [solute]gw + (1- x )[solute]up (7-6)

which can be rearranged to give:

[solute] - [solute] [solute] - [solute]

upgw

updown=x (7-7)

where [solute]down, [solute]up and [solute]gw are the respective downstream, upstream and

groundwater concentrations of the chemical constituent of interest. The variable x is defined as

the fraction of groundwater in the stream. Re-writing equation 7-7 gives:

b-c b- a

=x (7-8)

where down[solute]=a , up[solute]=b and gw[solute]=c

i.e. z

yx = (7-9)

where by −= a and bcz −=

Page 247: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

223

Numerous dissolved chemical species were measured in groundwater and the Herbert River

during the sampling periods in May 2004 and June 2005 (beginning of dry season) and October

2004 (end of dry season) (Table 3-2, Chapter 3). Therefore, a mass balance can be applied to all

of the common constituents at the same sampling time. As the approach assumes conservative

mixing between two end-members, the reliability of the estimates is a function of how well the

end-member chemical compositions can be constrained (Herczeg and Edmunds, 2000). An

important aspect of the calculation is capturing the uncertainty in the measurements of

individual solutes and propagating these errors through to the estimation of the contribution

from groundwater. Uncertainty can be introduced at numerous stages throughout the process of

collecting and analysing water samples. Random and systematic errors in the field occur in the

collection, preparation, and preservation of samples, while in the laboratory there are errors

associated with sample processing and the analytical instruments. Superimposed on this is the

variability in the composition of groundwater, which creates an uncertainty in its average value.

While it is difficult to quantify all of the possible error sources, the mass balance presented here

allows for uncertainty in the concentrations of the solutes due to: (1) instrument limits of

detection; and (2) the spatial heterogeneity of groundwater. The limits of detection vary between

solutes and between different laboratories: these differences were taken into account in the

uncertainty for each constituent during the three sampling periods. The concentration of HCO3

was calculated from the alkalinity, which was determined by titration against an acid;

measurement uncertainties were propagated through the calculation to define a limit of detection

for this solute. Under the assumption of local baseflow contributions to the stream along the

reach of interest, the chemistry of nearby bores (deep and shallow) was averaged to give a

representative groundwater composition. Note that the river intersects the deep aquifer in this

area. As there were only three groundwater measurements, the uncertainty in the groundwater

composition was defined as the range (i.e. ± ½ range) between the maximum and minimum

concentrations for each chemical constituent. This uncertainty is a reflection of the spatial

heterogeneity of groundwater in the upper part of the catchment. Assuming that the errors

accounted for are random, the errors on y and z (equation 7-9) can be expressed as:

22 )(a)( by Δ+Δ=Δ and 22 )(c)( bz Δ+Δ=Δ (7-10)

where aΔ and bΔ are the detection limits of the analytical instrument for each chemical

constituent (or error calculated for HCO3) and cΔ is the greater of the two types of error in the

measurement of each solute in groundwater.

Page 248: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

224

The fractional uncertainty in the calculated proportion of groundwater is thus:

222

⎟⎠⎞

⎜⎝⎛ Δ+⎟⎟

⎞⎜⎜⎝

⎛ Δ=⎟

⎠⎞

⎜⎝⎛ Δ

zz

yy

xx

(7-11)

A best estimate of x based all chemical constituents can thus be computed as:

=

== N

ii

N

iii

best

w

xwx

1

1 (7-12)

for solutes i = 1, 2, ….., N where the weight is defined as 2

1

ii x

= (Taylor, 1982).

Therefore, chemical constituents with the greatest amount of uncertainty contribute less to the

final estimate of the proportion of groundwater. The uncertainty in xbest can therefore be

calculated as:

2/1

1

=

⎟⎠

⎞⎜⎝

⎛=Δ ∑

N

iibest wx (Appendix C) (Taylor, 1982) (7-13)

where 2

1

ii x

= as before.

7.5.4.2 Local baseflow contribution

Based on the approach presented above, a mass balance was undertaken over reach 2. The

approach assumes that other than local baseflow, there are no additional sources of solutes to the

stream. The results of these computations for the three sampling periods are summarised in

Table 7-4.

Table 7–4 Best estimates of the percentage of local baseflow that contributes to streamflow along reach 2 during months in the dry season, based on a solute mass balance (refer to text for details).

May 2004 October 2004 June 2005

% groundwater (xbest) 5.2 30 8.8

uncertainty (Δxbest) 0.8 2 0.9

Page 249: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

225

Where available, the constituents included in the mass balance were F, Cl, SO4, HCO3, Ca, Mg,

Na, K, Si, Zn, NO3, δ18O, and δ2H; solutes with no concentration difference between

downstream and upstream in the river did not contribute to the calculations for that particular

time period. As shown in Table 7-4, the estimated local contribution of groundwater to

streamflow over reach 2 is less than 10 % during the beginning of the dry season and rises to

around 30 % by the end of the dry season. The solutes that contribute most to the estimates (as

determined by the weight wi) include HCO3, Ca, Mg, Na, Si and NO3 towards the beginning of

the dry season, with Cl becoming an important solute and Mg less so, at the end of the dry

season.

The fraction of the error,

i

ibest

xxx

Δ−

(7-14)

for the main contributing solutes is generally around 1 or slightly beyond the error range, except

for NO3, which has a much higher deviation in October 2004 and June 2005 (Table 7-5). This

either suggests that the uncertainty has been underestimated or that there is an additional source

that contributes NO3 to the stream. Given that there were only three bore measurements used to

represent the composition of the groundwater, it is possible that the spatial variability in the

aquifer was underestimated, particularly with respect to NO3. On balance, the generally low

fraction of the error for most constituents suggests that the mass balance approach and the

associated assumptions are reasonable for estimating the local baseflow contribution. While not

all possible water sources have been considered, the change in river chemistry along reach 2 in

each sampling period is consistent with discharge from a local groundwater source in the upper

reaches of the river. In the absence of chemistry data for inflowing tributaries over the reach, it

is not possible to differentiate between direct groundwater discharge versus inflow from

groundwater-fed tributaries. However, there is strong evidence that there is an increase in the

local contribution of groundwater to the stream from the beginning to the end of the dry season.

Given that estimates of groundwater flux using a radon mass balance could not be determined

along reach 2 in June (due to the flow event), results from the solute and radon mass balance

approaches are compared for October only. Along reach 2, the groundwater flux was estimated

at 0.85 m3/day/m using radon, which equates to approximately 17 % of total streamflow

(approximately 200 ML/day at gauge 116006) during that period in October 2004. In

comparison, using a solute mass balance, the proportion of baseflow contribution along the

same reach is around 30 %. Assuming that inflowing tributaries are supplied by groundwater in

October, a comparable proportion of groundwater input of 38 % was estimated along reach 2

from a mass balance of solute loads (Gowrie Creek, Table 7-2).

Page 250: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

226

Note that these estimates represent the local baseflow contribution along the reach, not the total

groundwater flux contributing to streamflow above reach 2 (as determined in Chapter 6). Whilst

instrument errors and the spatial heterogeneity in groundwater composition were considered in

the solute mass balance approach, errors were not considered in the radon mass balance due to

the number of parameters with unknown uncertainties. Therefore, the discrepancy in the

estimated flux between the two approaches can be attributed to an underestimation of the errors.

In relation to the radon approach, sensitivity analysis indicates that the estimated groundwater

flux is sensitive to the radon activity of groundwater inflow as well as the gas exchange velocity

and river width. Therefore, improvement in the measurement of these parameters and capturing

their uncertainty would be important for refining the estimation of groundwater flux using a

radon mass balance. Similarly, additional sources of error could be included in the solute mass

balance. Whilst refinements to both approaches would provide greater confidence in the

estimated groundwater flux, this is beyond the scope of the thesis and is an area for future work.

Page 251: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Table 7–5 Calculated values for xi, Δxi and the fraction of error (equation 7-14) during the beginning (May) and end (October) of the dry season in 2004 where xbest (May) = 5.2±0.8, xbest (October) = 30±2 and xbest (June) = 8.8±0.9 (refer to Table 7-4).

F Cl SO4 HCO3 Ca K Mg Na S Si Zn NO3 δ18O δ2H

May 2004

xi - -22 3.3 17 10 75 15 6.5 14 4.2 50 6.0 -41 5

Δxi - 44 120 6.7 3.6 260 11 1.6 1800 0.91 160 3.9 71 38

fraction of the error - 0.62 0.02 -1.7 -1.3 -0.27 -0.93 -0.78 0 1.1 -0.28 -0.20 0.66 0

Oct 2004

xi 44 15 120 48 49 34 110 32 6700 27 100 13 - -

Δxi 680 17 210 10 11 56 77 2.7 1100000 5.8 280 5.5 - -

fraction of the error -0.02 0.83 -0.43 -1.8 -1.7 -0.09 -1.0 -0.88 -0.01 0.41 -0.26 3.0 - -

June 2005

xi - 37 24 11 14 - 15 11 - 7.8 100 2.9 - -

Δxi - 66 100 3.6 4.2 - 11 1.4 - 1.7 200 1.9 - -

fraction of the error - 0.70 -0.15 -0.69 -1.3 - -0.58 -2.0 - 0.58 -0.46 3.1 - -

Page 252: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

228

7.6 TRANSPORT OF NITROGEN BY GROUNDWATER

The analyses presented in Sections 7.2-7.5 have demonstrated that:

• discharge of groundwater is the major process that influences the chemistry of the lower

Herbert River and its tributaries, particularly at the end of the dry season;

• groundwater discharges along the entire length of river, either directly or via

groundwater-fed tributaries;

• the flux of groundwater varies along the length of the river, with a considerably greater

flux between reaches 6-10 due to a switch from HSd to HSs as the dominant contributing

aquifer; and

• a large flux of groundwater is evident in the tidal zone at the end of the dry season.

In addition, based on temporal trends in species of N (Section 7.3.3):

• dry season months are dominated by dissolved forms, between 80-90 % of total N;

• DIN ranges from 70 % (start of the wet season) to less than 2 % (end of the dry season)

of total dissolved N, representing a maximum of 30% and down to 1% of total N; and

• NO3- is the dominant inorganic species, with a relatively constant concentration during

the dry season.

In this section the distribution of DIN along the lower Herbert River is examined. The above

analyses have provided evidence of groundwater discharge along the river throughout the year;

therefore, the aim of the following analyses is to establish whether groundwater is also a

transport vector for DIN. The concentrations and spatial distribution of DIN in groundwater

were analysed in Chapter 5 (Section 5.6), and summarised in Figure 5-33. Similarly, a

conceptual diagram highlighting the spatial distribution of NO3- and sources to the lower

Herbert River at the end of the dry season is presented in Figure 7-38 at the conclusion of this

section.

Page 253: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

229

7.6.1 Longitudinal trends in DIN

In addition to field parameters and major ions, samples were collected as part of this study for

the three species of DIN along the river and its tributaries. Comparison of longitudinal sections

indicates that the concentration of NO3- does not vary markedly between months representing

the extremes of the dry season. However, speciation in regards to the other forms of DIN differs

between higher and lower flow periods (Figure 7-35). Based on the available data, it is evident

that NO3- and NH4

+ dominate at the beginning of the dry season (NO2- below the detection

limit), while NO3- and NO2

- are the main species at the end of the dry season. Whilst it is

beyond the scope of the thesis to examine N speciation in detail, the observation is consistent

with differences in the environmental conditions of the river, such as in the pH, redox,

temperature and microorganism activity. In addition, due to the transition from surface-

dominated to groundwater-dominated discharge, changes in the sources of N to the river are

plausible.

0.001

0.010

0.100

1.000

10.000

0 20 40 60 80 100

Conc

entr

atio

n (m

g/L)

NO3 (May)

NH4 (May)

16 18 20 249 1121 5

tidal limit

3

seawater mixing

a

0.001

0.010

0.100

1.000

10.000

0 20 40 60 80 100

Distance downstream from gorge (km)

Conc

entra

tion

(mg/

L)

NO3 (Oct)

NO2 (Oct)

NH4 (Oct)

15 18 20 249 10 1221 5 6 8

tidal limit

3

seawater mixing

b

Figure 7-35 Speciation of DIN along the lower Herbert River during months representing the beginning (May) and end (October) of the dry season. The zone of seawater influence is indicated based on major ion chemistry.

Page 254: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

230

7.6.1.1 Comparisons with other tracers

Based on the analysis of major ions and radon in Sections 7.4 and 7.5, it was established that

groundwater discharges along the entire length of the river, with contributions to the upper

reaches from the deep aquifer and from the shallow aquifer downstream of reach 5. Figure

7-36a illustrates that the large increase in NO3- concentration between reach 1-2 in October

corresponds to a decrease in the Mg/Ca ratio of a similar relative magnitude, which is indicative

of increasing groundwater contributions below the gorge and hence evolution of the chemistry

of the river. In addition, oscillations in NO3- downstream of reach 5 are generally paralleled by

shifts in both the Mg/Ca and Na/Cl ratios, consistent with discharge from different units within

the shallow aquifer (Figure 7-36a and b). Mismatch between major ions and NO3- are observed

at approximately 58 km (reach 5) and between 64-68 km downstream (reaches 7-9), where the

pronounced spikes (declines) represent NO3- loss due to denitrification or in-stream biological

reactions. In addition to major ions, 222Rn concentrations along the river display a similar trend

to NO3- (Figure 7-36c). On balance, these observations provide evidence that NO3

-, major ions,

and radon share a common groundwater source, at least at the end of the dry season. Measured

concentrations of NO3- in the shallow and deep aquifers (Section 5.6.2, Chapter 5) are also

consistent with a groundwater origin.

Downstream of reach 10 and upstream of the seawater mixing zone (reaches 10-18), the decline

in both NO3- and 222Rn (Figure 7-36c) corresponds with a decrease in the estimated groundwater

flux (refer to Figure 7-33). The decline in radon can be explained by a combination of

radioactive decay and degassing at a higher rate than the supply of 222Rn in the inflowing

groundwater. Similarly, the decrease in NO3- can be attributed to losses (such as through

denitrification or other in-stream transformations) in excess of the declining flux from

groundwater and/or dilution from NO3--depleted groundwater. There are a range of

environmental factors that can potentially influence the fate of NO3- in inflowing groundwater,

as well in-stream. However, given the similarity in trends between NO3- and 222Rn, the decline

in groundwater flux is a plausible explanation for the decline in the concentrations of both of

these dissolved species.

Page 255: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

231

0

1

2

0 20 40 60 80 100

Distance downstream from gorge (km)

NO

3- (mg/

L)

1

Mg/

Ca

(meq

/L)

NO3

Mg/Ca

9 10 12 14 17 20 2421 5 6 8tidal limit

seawater mixing

3 0.8

1.2

1.4

a

0

1

2

0 20 40 60 80 100

Distance downstream from gorge (km)

NO3- (m

g/L)

0

1

2

3

Na/C

l (m

eq/L

)

NO3

Na/Cl

9 10 12 14 17 20 2421 5 6 8tidal limit

seawater mixing

3

b

0

1

2

0 20 40 60 80 100

Distance downstream from gorge (km)

NO3- (m

g/L)

0

1

2

222 Rn

(Bq/

L)

NO3

Rn (Oct)

9 10 1214 17 20 2421 5 6 8tidal limit

seawater mixing

3

c

Figure 7-36 Comparison of NO3- with ion ratios and radon along the lower Herbert River

in October 2004. The zone of seawater influence is indicated based on major ion chemistry.

Page 256: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

232

Analysis of major elements showed that there is an abrupt change in the hydrochemistry of the

river in the zone of seawater mixing (Figure 7-25). Similarly, the concentration of NO3-

increases in this zone in October (downstream of reach 18). Given that the flux of groundwater

is also estimated to increase in this zone (at least in October, Figure 7-33), NO3- measured in the

river could be the result of direct groundwater discharge. NO3- concentrations in the shallow

aquifer, adjacent to the estuarine reaches of the river (Figure 5-30, Chapter 5), are consistent

with a groundwater source of NO3-. However, mixing with seawater cannot be discounted,

especially given that NO3- levels in seawater can be quite high e.g. 50 mg/L NO3

- has previously

been measured at Moreton Bay (QLD) (Cresswell, 2006, pers. comm.). Whilst the data are

inconclusive with regards to the source of NO3- in the zone of seawater mixing, previous studies

have found evidence of submarine discharge to the near-shore environment adjacent to the

catchment (Gagan et al., 2002; Stieglitz and Ridd, 2000). Therefore, an indirect groundwater

source of NO3- to the river, via mixing with seawater NO3

- sourced from submarine discharged

groundwater, is also possible.

Mass balance calculations can be used to test whether a potential groundwater contribution of

NO3- to the river is plausible (Table 7-6). Similar to the solute load balances in Section 7.4.3.3

(Table 7-2), a NO3- mass balance along reach 2, which displays the greatest increase in NO3

-

(Figure 7-36), highlights a surplus NO3- load in the river which cannot be accounted for by

Gowrie Creek alone. Assuming that other tributaries along the reach are supplied by

groundwater, the required groundwater discharge rate to contribute the NO3- load is

approximately 40 ML/day (given the concentration of NO3- in nearby HSd groundwater). This is

comparable with the estimated groundwater input required to account for the observed change

in solute load along reach 2 (Table 7-2). Hence, groundwater contributes to both the salinity and

NO3- load of the river.

Table 7–6 Required groundwater discharge to account for observed NO3- concentrations

in the Herbert River in October 2004

Tributary NO3- in

tributary (mg/L)

Estimated tributary NO3

-

load (106 mg/day)+

Δ NO3- load in

Herbert R (106 mg/day)*

Unaccounted NO3

- load in Herbert R

(mg/L)#

Required gw input

(ML/day)@

Gowrie Ck 2.7 54 300 250 40

Elphinstone Ck 1.3 110 100 0 0

Dalrymple Ck 1.0 10 -130 0 0

Stone R 1.9 110 30 0 0

Ripple Ck 12 50 -50 0 0 + Estimated tributary discharge from Table 7-2 * Defined as downstream minus upstream concentrations in the Herbert River # NO3

-concentrations in contributing aquifer (HSd upstream of Dalrymple Ck and HSs downstream) @ In addition to inflow from measured tributary (compare with Table 7-2)

Page 257: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

233

Similarly, the other tributaries of Elphinstone Creek and Stone River, which are most likely

groundwater-fed in October, contribute to the NO3- load of the Herbert River. Consistent with

the solute load calculations, Dalrymple Creek does not contribute NO3- to the river; however,

while Ripple Creek contributes other solutes (Table 7-2), it is not a source of NO3- (Table 7-6).

The low Eh of Ripple Creek compared to the river is consistent with denitrification and hence

transformation of NO3- prior to inflow or upon mixing with the river.

7.6.1.2 Comparisons within the dry season

Comparison of longitudinal data at the beginning and end of the dry season indicates that in

general there is an increase in NO3- concentration downstream of reach 1 and a gradual decrease

downstream of reach 10, noting that as for major ions, the concentration of NO3- between

reaches 7-10 is diluted by overland flow during the June sampling period (Figure 7-37).

0.0

0.4

0.8

1.2

1.6

2.0

0 20 40 60 80 100

Distance downstream from gorge (km)

NO

3- (mg/

L)

May 2004

Oct 2004

June 2005

15 18 20 249 1221 5 6 83 10

tidal limit

seawater mixing

seawater mixing

Figure 7-37 Longitudinal plots of NO3- during months representing the beginning (May,

June) and end (October) of the dry season. The zone of seawater influence is indicated based on major element chemistry in the corresponding sampling periods.

Whilst the concentration of NO3- at reach 1 is similar in each sampling period, the rate of

increase for 45 km downstream of the gorge is markedly higher at the end of the dry season

(October) compared to the beginning (May/June). As established from major ions and EC, this

intra-seasonal variation can be attributed to differences in the dominant transport vectors at the

extremes of the dry season i.e. a combination of surface runoff and groundwater discharge at the

beginning of the dry season transitioning to groundwater dominance by the end of the dry

season. Whilst the concentration of groundwater NO3- is likely to remain fairly constant

throughout the dry season, inflow of tributaries between reaches 1-5, such as Gowrie Creek,

may dilute the NO3- concentration in the river during higher flow conditions. Further tributary

sampling during the beginning of the dry season, in the absence of rainfall events, would be

required to test this.

Page 258: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

234

Within the tidal zone, different NO3- trends are also observed during each of the sampling

periods. Disregarding the tidal reaches of the June data, which were possibly affected by the

flow event, the decline in NO3- downstream of the freshwater zone is observed at reach 14 in

May and at reach 18 in October. These differing trends within the estuary indicate that different

processes influence the concentration of NO3- during the extremes of the dry season. There is

insufficient data to explain why these discrepancies are observed; however, it is noted that the

chemistry in the tidal zone at different times of the year is a reflection of the complex balance

between competing hydraulic pressures imposed by the river, the groundwater, and the sea. The

collection of complementary radon data and hence groundwater flux modelling at the beginning

of the dry season (in the absence of flow events) could be an avenue for future research.

The analyses in this section highlight that there are many unanswered questions regarding the

distribution of DIN in the river. However, the analyses have shown that the concentration of

NO3- at the downstream extent of the freshwater zone is distinctly elevated over the

concentration measured below the gorge during both the beginning and end of the dry season.

This demonstrates that NO3- originates from a source in the lower catchment. Furthermore, the

results provide strong evidence that groundwater is an important transport vector for NO3- to the

lower Herbert River, at least during months of the dry season. In addition, the concentration of

NO3- is highly variable along the river; therefore, NO3

- data must be interpreted with regard to

the location of sample collection as well as the timing. The data presented above clearly

demonstrate that measurements at the river mouth do not necessarily represent the maximum

levels of NO3- attained in the river system. Therefore, while end-of-river sampling can be useful

for monitoring downstream impacts, such as the potential transport of nutrients to the Great

Barrier Reef, longitudinal sampling is important for river management and hence identifying

hotspots of high nutrient availability.

7.6.2 Environmental significance

Water quality guidelines have been developed specifically for different regions and water types

within Queensland (EPA, 2006) as an extension of the National Water Quality Management

Strategy and the ANZECC8 guidelines for fresh and marine water quality in tropical Australia

(ANZECC and ARMCANZ, 2000). The regional guideline values for the defined Wet Tropics

drainage division, which includes the Herbert River catchment, summarise the trigger values for

oxidised N (NO3- + NO2

-) and ammonium N to be 30 μgN/L and 10-15 μgN/L, respectively, in

lowland streams and mid-estuarine/tidal canals in slightly-moderately disturbed systems.

Comparison of these trigger values with measurements in the lower Herbert River (Table 7-7)

indicate that concentrations of NH4+ are close to the guidelines values for aquatic ecosystem

protection. However, downstream of Nash’s Crossing (reach 1), the concentration of oxidised N

8 Australian and New Zealand Environment and Conservation Council

Page 259: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

235

far exceeds the recommended trigger value at both the beginning and end of the dry season.

This suggests that N contributed from the lower catchment is of greater environmental concern

than upstream contributions, at least during the dry season. The concentration of NO3-, the main

form of oxidised N, is particularly high within the first 70 km downstream of the gorge.

Table 7–7 Concentration range for species of DIN for all samples collected in the lower Herbert River during selected months of the dry season. EPA and ANZECC guideline values for tropical Australia are also indicated.

Sampling time NO3- (μg/L) NO2

- (μg/L) NH4+ (μg/L)

May 2004 66 - 1549 - 3 - 10

October 2004 35 - 1709 7 - 16 8 - 12

June 2005 18 - 1062 - 8 - 22

EPA+ 133 13 - 19

ANZECC+ 44-133 13 - 19 + the oxidised N trigger value includes both NO3

- and NO2-

Importantly, longitudinal analyses highlight that the concentration of NO3- can vary

considerably along the river. Therefore, depending on where samples are collected, comparisons

against guideline values may or may not reflect the magnitude of potential impacts on

ecosystem health. For instance, as illustrated for measurements in October 2004 (Figure 7-37),

if samples were collected just below the gorge and downstream at the end of the freshwater

zone, this would not reveal that much higher concentrations are obtained in the stream in

between. As reviewed in Chapter 3 (Section 3.2.1.1), a previous study suggested that nutrient

concentrations in the Herbert River are below ANZECC (2000) target levels for the protection

of freshwater ecosystems, except in high flow conditions when the trigger values are exceeded.

Furthermore, it was concluded that nutrient loss in the catchment is event based and

insignificant outside the wet season months (Bramley and Johnson, 1996). Whilst peak wet

season events dominate the annual riverine export of nutrients (Bramley and Muller, 1999), the

results presented in this chapter clearly demonstrate that the dry season is also characterised by

elevated in-stream concentrations of NO3- within the freshwater reaches of the river.

Measurements in the estuary during the extremes of the dry season also highlight that the

concentration of NO3- is well above the Wet Tropics trigger value. This indicates the potential

for transport of high NO3- waters to the near-shore environment during the dry season, which

may also have implications for ecosystem health in the Great Barrier Reef.

Page 260: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

236

Figure 7-38 Conceptual diagram summarising the movement of water and N between the aquifers and the lower Herbert River at the end of the dry season

Page 261: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chemical River-Groundwater Interactions

237

7.7 CHAPTER SUMMARY

Based on the availability of temporal water quality data in the lower Herbert River, the analyses

in this chapter have shown that there are distinct hydrochemical differences between the wet and

dry seasons, as well as between the extremes of the dry season. Although the availability of fine

resolution temporal data is limited, daily EC measurements clearly illustrate a progressive

increase in salinity of the river throughout the dry season, followed by dilution during months of

the subsequent wet season. Whilst evaporative effects are not discounted, a progressive increase

in baseflow contributions is the dominant mechanism driving the increase in river EC as flow

decreases. Hence, there is a seasonal shift from runoff-dominated to baseflow-dominated waters

contributing to the Herbert River. A decrease in the Cl/HCO3 ratio of surface waters between

the wet and dry seasons, coupled with similarities between the hydrochemical signature of

surface waters and Ca-Mg-HCO3 enriched groundwaters, provide further evidence of aquifer

discharge to the Herbert River and its tributaries during the dry season.

Consistent with the temporal analyses, longitudinal analysis of a range of parameters in the dry

season has demonstrated that groundwater discharge is the major process that influences the

chemistry of the lower Herbert River, particularly during lowest flow conditions. Although

tributary inflow is also important within some river reaches, these tributaries are supplied by

groundwater, especially at the end of the dry season. Therefore, groundwater discharge to the

river represents a combination of direct contributions from an aquifer and inflow from

groundwater-fed tributaries. Whilst there is evidence to support groundwater discharge to the

stream along the entire freshwater section throughout the dry season, the flux of groundwater

varies along the river. Considerably greater discharge is evident between approximately 60-70

km downstream from the gorge in October and June. Supported by major element

concentrations, this change in flux is due to a switch from HSd to HSs as the dominant

contributing aquifer. Whilst discharge from the deep aquifer is the main influence on river

chemistry in the upper reaches, discharge from the shallow aquifer dominates the lower reaches.

This shift in river-aquifer relationship, based on hydrochemical evidence, is consistent with

spatial variations in the incision depth of the river into the aquifers (Chapter 4), together with

baseflow estimates that indicate an enhanced contribution of baseflow between the two gauges

in the study area (Chapter 6). Whilst the groundwater flux declines towards the estuary, a

dramatic increase is estimated in the zone of seawater mixing. Although strong discharge of

groundwater is plausible in the tidal zone, there is lower confidence in the flux as mixing with

seawater masks the groundwater signal.

Uncertainty analysis of the estimated groundwater flux using radon, by comparison with a

solute mass balance, provides reasonable confidence in the estimated local baseflow

contribution along the reach characterised by the greatest change in solute concentrations.

Page 262: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 7

238

Furthermore, comparison of the estimated groundwater discharge between the two gauges along

the river, using a radon mass balance, is of a similar magnitude to the difference in recorded

discharge between the gauges during the same period. Improvements to measurements of the

most sensitive parameters and their uncertainties would provide greater confidence in the

estimated groundwater fluxes determined by each of the mass balance approaches. This is an

area for future work.

The analyses presented in this chapter provide strong evidence that groundwater is an important

transport vector for NO3- to the lower Herbert River during months of the dry season. NO3

- and

NH4+ dominate the river at the beginning of the dry season, whilst NO3

- and NO2- are the main

species during lowest flow conditions. These speciation differences are due to the transition

from surface-dominated to groundwater-dominated discharge to the river through the course of

the dry season. An increase in NO3- below the gorge is attributed to discharge from the deep

aquifer and inflow from groundwater-fed tributaries, while in the middle reaches of the river

NO3- is sustained by groundwater fluxes from the shallow aquifer. Although some tributaries

influence both the salinity and NO3- load of the river, others contribute to the solute load alone

with no input of NO3-. Therefore, the interpretation of non-conservative tracers such as NO3

- is

strengthened by complementary analysis of conservative tracer data. Towards the estuary, a

decline in NO3- concentration corresponds with a decrease in groundwater flux, coupled with

losses due to in-stream processes and/or denitrification. Whilst there is an estimated increase in

groundwater flux in the zone of seawater mixing, the observed concentrations of NO3- in the

river could be from groundwater or seawater sources.

Analysis of particulate and dissolved species of N has shown that dissolved forms are an

important component of total N in the stream during the dry season. Although the proportion of

DIN in the dry season is relatively minor compared to organic forms, longitudinal analysis

indicates that NO3- is present at concentrations that markedly exceed recommended guideline

values for aquatic ecosystem protection. Moreover, the analyses highlight that the concentration

of NO3- varies considerably along the river. Therefore, depending on where samples are

collected, comparisons against guideline values may or may not reflect the magnitude of

potential impacts on ecosystem health. The data clearly demonstrate that measurements at the

river mouth do not necessarily represent the maximum levels of NO3- attained in the river

system. Hence, while end-of-river sampling can be useful for monitoring downstream impacts,

such as the potential transport of nutrients to the marine environment, longitudinal sampling is

important for river management, including identification of hotspots of high nutrient

availability. Implications of the results for nutrient monitoring and management are discussed

further in the concluding chapter of the thesis.

Page 263: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

239

Chapter 8 Research Conclusions

8.1 INTRODUCTION

The thesis commenced with a depiction of an integrated catchment management framework that

highlighted the significance of water resources as a link between land management practices

and outcomes for environmental, social and economic systems (Figure 1-1, page 2). The

management of water relates to both quality and quantity. Both of these elements can be

significantly influenced by the interactions between groundwater and surface water. This

research has focussed on the quality implications of connected water resources, motivated by

growing worldwide concern over the pollution of surface and groundwater resources by

nutrients, which can have damaging effects for both ecosystem and human health. Water is a

vehicle for mobilising dissolved constituents, including nutrients, between surface and

subsurface waters, and between terrestrial and marine systems. Therefore, an understanding of

river-aquifer connectivity is a key aspect of nutrient management in landscapes where nutrient-

bearing surface and groundwaters interact. In particular, ascertaining the significance of

groundwater fluxes for river nitrogen budgets is an important application of river-groundwater

linkages. This overarching concept has been explored and developed through the course of the

thesis.

Due to the ecological significance of the Great Barrier Reef World Heritage Area, an adjacent

coastal agricultural catchment, the Herbert River, was chosen as the target area for this research.

The selected case study catchment, located in the tropical climate zone of northeastern

Australia, also provided a unique opportunity to study groundwater contributions to the river

with minimal overland flows. Characterising the dynamics of water movement between river

and aquifer storages is a crucial step to understanding the mobility of dissolved N between

them. Accordingly, a major component of the thesis has been to investigate river-groundwater

interactions at the catchment scale, which is underpinned by characterisation of the

hydrogeological system. The role of groundwater as a vector for N is hence evaluated in light of

this conceptual understanding.

The major contributions of this thesis have been to:

(1) generate new knowledge about the Herbert River catchment, particularly in regards to

the alluvial aquifer system and surface-groundwater interactions;

Page 264: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 8

240

(2) establish the significance of groundwater as a vector for dissolved N to the river and

hence raise important implications for nutrient management, particularly in tropical

rivers; and

(3) demonstrate the value of combining analytical techniques, not provided by any one

method, to unfold different layers of a complex water resource problem involving both

surface and groundwater systems.

These research contributions are discussed in the following sections, with avenues for future

work also identified.

8.2 KEY FINDINGS

The analytical findings from this research are important because they provide a new

understanding of the sources of water that sustain flow in the lower Herbert River. The analyses

demonstrate that discharge of groundwater from the alluvial aquifers is a dominant influence on

both the flow and chemistry of the river in the dry season. In particular, groundwater is a key

vector for the delivery of nitrate (NO3-) to the river during low flow conditions. This provides a

new perspective for monitoring and management of nutrients in tropical rivers where there is

good connectivity with the underlying groundwater system.

The main analytical components of the thesis include:

(1) characterising river-aquifer interactions; and

(2) ascertaining the role of groundwater for the nitrogen budget of river systems.

A major element of (1) was also to conceptualise the hydrogeological system. The key

analytical findings from each of these components are summarised below.

8.2.1 Hydrogeological framework

The alluvial sequence of the lower Herbert River catchment is conceptualised as a two-aquifer

system comprised of a shallow unconfined aquifer (HSs) and semi-confined deep aquifer (HSd).

Although the aquifers are distinct, the extent of vertical hydraulic connection varies spatially.

The western half of the alluvial system is characterised by good/strong downwards connectivity.

In contrast, poor inter-aquifer connection is evident in the east, except towards the coast, where

strong upward heads and good connectivity results in vertical discharge from HSd to HSs.

Groundwater in each aquifer flows from the upland lateral recharge areas in the northwest and

southwest towards the lower Herbert River, and then eastward in the direction of the coastal

discharge zone. In addition to lateral sources, diffuse rainfall is an important component of

recharge to HSs.

Page 265: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Research Conclusions

241

Shallow groundwaters are dominated by Na-HCO3-Cl facies, with both Ca-Mg enriched and Na

enriched groups relating to clay content of the overlying soils. The exception is towards the

coast, where Na-Cl ± HCO3 facies groundwaters are associated with contributions from the deep

aquifer and mixing with the Herbert River estuary. The western half of the deep aquifer is

dominated by Na-HCO3 and Na-HCO3-Cl facies waters. Lateral hydrochemical evolution is

suppressed by enhanced vertical leakage from the shallow aquifer and the convergence of

flowpaths contributing groundwater of different compositions. The eastern half of the deep

aquifer is represented by Na-Cl facies waters with high salinities, influenced by seawater from

past and/or present-day intrusion via a preferential pathway in the northeast.

Despite good vertical connectivity between the aquifers in some areas, dissolved inorganic

forms of N (DIN) are very different between aquifers due to redox controls. Therefore, whilst

oxidising conditions in the main recharge area favour NO3-, the concentration of DIN in the

deep aquifer dramatically declines down the flowpath due to N reduction processes. Hence,

elevated NO3- in HSd is restricted to the high country in the northwest, the main source region

for DIN leached to the deep aquifer. In contrast, relatively high concentrations of NO3- are

found throughout the shallow aquifer, with the distribution and speciation of DIN influenced by

aquifer composition, including soil type and redox state. Oxidising Ca-Mg-enriched waters are

associated with high concentrations of NO3-, while more reduced Na-enriched waters are

dominated by NH4+. Importantly, high concentrations of NO3

- are present in shallow and deep

groundwaters adjacent to the Herbert River. Therefore, given the hydrogeological framework,

there is potential for subsurface N to contribute to the river system. Ultimate discharge of

groundwater from both aquifers to the sea indicates an additional pathway for the movement of

N offshore. A conceptual diagram summarising water and N movement in the alluvial aquifer

system was presented in Chapter 5 (Figure 5-33). This figure is repeated below as Figure 8-1.

8.2.2 River-groundwater interactions

Groundwater discharge is the major process that influences the chemistry of the lower Herbert

River during the course of the dry season, particularly during the lowest streamflow conditions.

Whilst other processes such as evaporation are not discounted, a progressive increase in

baseflow contributions is the dominant mechanism that drives an increase in river salinity as

flow decreases. Hence, there is a seasonal shift from runoff-dominated to baseflow-dominated

flow in the river. Tributaries are also an important source of water and dissolved constituents

along some reaches; however, many of these tributaries are supplied by groundwater, especially

at the end of the dry season. Therefore, groundwater discharge to the river in the dry season

represents a combination of direct and indirect (via tributary inflow) aquifer contributions.

Page 266: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 8

242

Year-round, the dominant potential direction of groundwater flow is from the aquifers to the

river; however, towards the coast, the direction of potential flux reverses during short periods of

high river stage in the wet season. Hence, the upper reaches of the river are dominantly gaining,

whilst the lower reaches are characterised as variably gaining/losing. Whilst groundwater

discharges to the river along the entire freshwater section throughout the dry season, the flux is

variable along the river. Considerably greater flux is evident between approximately 60-70 km

downstream from the gorge in October and June, attributed to a transition from HSd to HSs as

the dominant contributing aquifer. Hence, discharge from the deep aquifer is the main influence

on river chemistry in the upper reaches, while discharge from the shallow aquifer dominates the

lower reaches. Groundwater inflow to the river declines towards the estuary, but with a dramatic

increase estimated in the zone of seawater mixing. Although strong discharge of groundwater in

the tidal zone is plausible, there is lower confidence in the flux, as mixing with seawater masks

the groundwater signal.

Uncertainty analysis, using independent mass balances of radon, and solute loads, provides

reasonable confidence in the estimated local baseflow contribution along the freshwater reach

characterised by the greatest change in ionic concentrations. In addition, the estimated discharge

between the two gauges along the river is consistent with the difference in gauged discharge

during the same period. Improvements to measurements of the most sensitive parameters and

their uncertainties would provide greater confidence in the estimated groundwater fluxes

determined by each of the mass balance approaches.

8.2.3 The significance of groundwater for river N budgets

This research demonstrates that groundwater is an important transport vector for DIN to the

lower Herbert River during months of the dry season. Although NH4+ and NO2

- also comprise

the inorganic component, NO3- is by far the dominant species throughout the dry season. An

increase in NO3- below the Herbert River gorge is attributed to discharge from the deep aquifer

and inflow from groundwater-fed tributaries, whilst in the middle reaches of the river, NO3- is

sustained by groundwater fluxes from the shallow aquifer. Importantly, whilst some tributaries

influence both the salinity and NO3- load of the river, others contribute to the solute load alone

with no input of NO3-. Towards the estuary, a decline in NO3

- corresponds with a decrease in

groundwater flux, coupled with losses due to in-stream processes such as denitrification. Whilst

there is an estimated increase in groundwater flux in the zone of seawater mixing, the observed

concentrations of NO3- in the river could be from groundwater or seawater sources. A

conceptual diagram summarising the movement of water and N between the aquifers and the

lower Herbert River is presented below in Figure 8-2 (equivalent to Figure 7-38, Chapter 7).

Page 267: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Figure 8-1 Conceptual diagram summarising the movement of water and N in the alluvial aquifer system and potentially to the Herbert River (HR)

Page 268: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 8

244

Figure 8-2 Conceptual diagram summarising the movement of water and N between the aquifers and the lower Herbert River at the end of the dry season

Page 269: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Research Conclusions

245

The N budget of the lower Herbert River varies both temporally and spatially. Particulate forms

dominate in the wet season, whilst dissolved forms comprise 80-90% of total N during months

of the dry season. Although the proportion of DIN in the dry season is relatively minor

compared to organic forms, the concentration of NO3-, particularly within 70 km downstream of

the gorge, markedly exceeds the recommended Wet Tropics and ANZECC guideline values for

aquatic ecosystem protection. An important result arising from this research is that the

concentration of NO3- is highly variable along the river during the dry season. Measurements

immediately below the gorge and towards the river mouth do not reflect the much higher

concentrations of NO3- present in the intervening reaches. Therefore, water quality data must be

interpreted with regard to the location of sample collection as well as the timing. This is of

particular importance when comparing water quality indicators such as NO3- against trigger

values to assess the magnitude of potential ecological impacts. Hence, whilst end-of-river

sampling can be useful for monitoring downstream impacts, such as the potential transport of

nutrients to the Great Barrier Reef, longitudinal sampling is important for river management,

including identification of hotspots of high nutrient availability.

8.3 RESEARCH APPROACH

The key findings were determined through an integrated research approach comprising a data

collection component and an assessment methodology. These elements were guided by the

fundamental research questions, together with regard to existing available information,

resources to collect additional data, and the specific attributes of the case study catchment. The

water quality sampling program took advantage of the distinct seasonal nature of tropical

catchments by sampling during the period characterised by minimal overland flows. Hence the

pure baseflow component in the river has been captured with reasonable confidence. The

analytical approach applied in this thesis clearly demonstrates that methods traditionally used by

separate disciplines, both qualitative and quantitative, can be combined to provide a powerful

toolbox of complementary techniques that enrich the conceptual process understanding of the

system.

8.3.1 Data collection

In order to assess large scale river-groundwater interactions and the application to nutrient

transport, a catchment scale investigation was undertaken. Given the broad scale of the study,

spatial relationships, rather than temporal dynamics, were considered to be of greatest interest.

Hydrological processes in the tropics are generally less dynamic during the dry season, as

overland flow is minimal and baseflow dominates streamflow. Accordingly, an extensive water

quality sampling program was instigated in the case study catchment during low flow

conditions. Grab samples of groundwater and surface waters at the beginning and end of the dry

Page 270: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 8

246

season were considered to be adequate for comparing differences in river-aquifer connectivity

relationships between the extremes of the dry season at a catchment scale. The results highlight

that the end of the dry season is particularly useful for isolating the groundwater signal in the

river.

A key element of the research approach was the use of physical datasets sourced primarily from

existing government department records. These provide a backbone for understanding the

hydrogeology and hydrology in the case study catchment. Whilst it is noted that these datasets

generally have gaps in historical records and issues with data quality, they are nonetheless

considered to be a valuable source of high temporal resolution information, at least during

particular monitoring periods. The collection of time series data was not feasible within the

scope and resources of the project. Existing datasets were therefore an important supplement to

the hydrochemical data collected specifically for this investigation. As this research

demonstrates, verification and extension of physical concepts can be accomplished through

analysis of water quality data. The extensive database of hydrochemical information is of

reasonably high spatial resolution and low temporal resolution, thus complementing the

physical datasets.

8.3.2 Characterising river-aquifer interactions

A combination of hydrogeological, hydrometric, hydrological and hydrochemical techniques

were applied in the thesis to characterise the interaction between the alluvial aquifers and the

lower Herbert River. An important aspect of establishing the relationships between surface and

subsurface waters was to build a conceptual understanding of the hydrogeological system. This

included:

(1) classifying distinct aquifers based on lithostratigraphic interpretation and assessment of

vertical hydraulic behaviour;

(2) establishing lateral groundwater flow patterns in each aquifer; and

(3) identifying aquifer recharge and discharge zones and potential interactions with the

Herbert River

Whilst the analysis of physical attributes alone enabled a reasonable framework for the

groundwater system to be developed, the complementary analysis of hydrogeochemical data

allowed verification and enhancement of the conceptual model. For example, bore hydrograph

analysis indicated the potential direction and degree of vertical flow between aquifers, while the

analysis of hydrogeochemical data provided evidence of actual exchange of water and the

relative extent of inter-aquifer mixing. Furthermore, hydrogeochemical analysis highlighted the

importance of vertical recharge processes over lateral hydrochemical evolution in each aquifer,

Page 271: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Research Conclusions

247

which could not be inferred from the available hydraulic data. Conversely, the interpretation of

hydrogeochemical information would have limited scope in the absence of a physical process

understanding of groundwater flow. The analyses presented in Chapters 4 and 5 demonstrated

that the hydrogeological conceptualisation was strengthened by the coupling of physical and

chemical datasets and the corresponding hydrogeological and hydrochemical assessment

techniques.

Similar to the approach for characterising the hydrogeological system, a combination of

physical-and chemical-based methods were utilised in Chapters 6 and 7 to examine the

connectivity between groundwater and the lower Herbert River. The potential for hydraulic

connection and the direction of flux between the aquifer system and the river were evaluated

through qualitative hydrometric approaches, including:

(1) depth relationships of the river channel with that of the underlying alluvial sediments;

(2) historical groundwater elevation-stream stage relationships; and

(3) groundwater flow patterns around the river

Hydrological techniques such as the analysis of (4) stream hydrographs and (5) flow duration

curves were also applied to determine the temporal characteristics of flow in the river, while the

groundwater flux to the river was quantified during the wet and dry seasons by (6) hydrograph

separation. Hence, the hydrometric methods provided an indication of the theoretical direction

of interaction along particular reaches of the river, whilst examination of flow characteristics

provided evidence of actual volumetric flux over a broad area.

The physical understanding of river-aquifer linkages was verified and enriched through the

assessment of surface water chemistry data, in conjunction with the conceptual hydrogeological

model incorporating the hydrochemical signature of groundwater. Surface water integrates

across the entire catchment as it is the cumulative result of hydrological processes upstream of

the sampling point. Therefore, water quality samples analysed along the river reflect upstream

inputs, such as the discharge of groundwater. The utility of different environmental tracers

(electrical conductivity, temperature, major ions, radon, stable isotopes) for distinguishing

between the key processes that influence river chemistry was illustrated, including the

application of radon as a tracer of groundwater discharge to the river. The analyses

demonstrated how the combination of tracers, at varied spatial and temporal resolution, can

provide a powerful toolbox for characterising the chemistry of the river in space and time.

Consistency between tracers provided confidence that the observed longitudinal trends were

important for identifying relative changes in water sources to the river, even though the

magnitude of hydrochemical variation was only minor for some dissolved components.

Page 272: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 8

248

Furthermore, the spectrum of tracer techniques yielded both qualitative and quantitative

information regarding the flux of groundwater along the length of the lower Herbert River.

Whilst the absolute groundwater fluxes determined have a degree of uncertainty, the various

mass balance approaches using radon and solute loads highlighted the general methodology and

efficacy of quantitative estimates in combination with visual qualitative trends to characterise

river-aquifer relationships.

8.3.3 Characterising nutrient mobility

As an application of river-aquifer interactions, the potential for movement of N from the

aquifers to the river was assessed. Approaches for characterising nutrient mobility range from

qualitative field-based studies to quantitative nutrient modelling. Given the focus on processes

and the extensive data requirements which could not be met with existing data, a modelling

approach was considered beyond the scope of the thesis. However, it is noted that the

interpretation of model outputs can be enhanced with process knowledge of key components

such as physical and chemical river-aquifer linkages. The field-based hydrochemical approach

for characterising the movement of N between surface water and groundwater comprised three

key components, involving:

(1) examination of the spatial distribution and speciation of DIN in the alluvial aquifers;

(2) appraisal of river-groundwater interactions and hence whether transport of DIN to

surface waters via groundwater is a plausible mechanism; and

(3) assessment of whether there is hydrochemical evidence of N transport to the lower

Herbert River via this mechanism

Investigating the significance of groundwater N fluxes to the river was founded therefore on an

understanding of the hydrogeology and the dynamics of water movement between the river and

alluvial aquifers. Furthermore, given that NO3-, the dominant form of DIN in the river, is non-

conservative under reduced chemical conditions, the interpretation of NO3- data was

strengthened by complementary analysis of conservative tracer data such as major ions and

solute loads. In addition, comparison of NO3- data with radon was a key diagnostic for a

groundwater source of NO3-. The hydrochemical analysis thus underlines the importance of a

multi-tracer approach for verifying and enriching a conceptual process-based model.

Page 273: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Research Conclusions

249

8.4 IMPLICATIONS FOR THE USE OF NUTRIENT BUDGETS AND MODELS

This thesis illustrates a methodology for undertaking a catchment scale study concerned with

characterising river-groundwater interactions and evaluating the significance of these

interactions for river nutrient budgets. The analyses have provided an understanding of the

dynamics of water movement between the alluvial aquifers and the lower Herbert River during

the dry season and shown that groundwater is a dominant source and vector for the transport of

NO3- to the river. The results highlight that there is considerable spatial variability in the system.

For example, the sources and fluxes of groundwater vary along the river, as do the

concentrations of NO3- in the groundwater and measured in-stream. In the absence of the multi-

tracer approach presented in the thesis, this variability would be difficult to identify in a nutrient

budget or other quantitative modelling approaches. Therefore, the results demonstrate the value

of on-ground research in order to gain process understanding, which could inform the inputs

into and/or design of a nutrient model and hence be used for nutrient management purposes.

However, it is acknowledged that resource constraints may inhibit comprehensive data

collection for a particular study. Hence, an important consideration arising from the thesis is the

minimal dataset requirements to undertake a meaningful nutrient budget or modelling approach,

in a system with high river-aquifer connectivity.

Based on the range of analytical techniques applied in this thesis, a basic understanding of river-

groundwater interactions can be gained through baseflow filtering and qualitative hydrometric

methods such as assessment of bore and stream hydrographs. Coupled with these tools,

literature values can be used as an indicator of the typical concentrations of DIN expected in

surface waters that drain catchments under a particular land use. Although these desktop

analytical methods are relatively simple, they rely on reasonably high temporal resolution river

and/or groundwater level monitoring data for examining connectivity. Furthermore, surrogate

literature values of DIN fail to provide information on the spatial variability of nutrient

concentrations found in the stream. In the absence of such datasets or appropriate monitoring

networks, simple and relatively inexpensive hydrochemical data can be collected in surface

waters, such as for the analysis of radon and NO3-. Radon data, coupled with other

measurements (depth, width, discharge) can identify sites of groundwater discharge and

potential changes in the flux along the river, while NO3- samples can indicate longitudinal

variability and the actual concentrations of NO3- reaching the stream. Ideally, an understanding

of the hydrogeological system would be necessary to capture the variability and complexity of

river-groundwater interactions. Nonetheless, for the purposes of a broad scale nitrogen budget

or other nitrogen models, a combination of simple desktop tools and targeted water quality

sampling are the minimum requirements. It is noted however, that the reliability of nutrient

model outputs and their interpretation in a connected river-aquifer system can be enhanced with

Page 274: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 8

250

sound process understanding of the physical and chemical relationships between surface water

and groundwater resources.

8.5 MONITORING AND MANAGEMENT IMPLICATIONS FOR THE TROPICS

This research raises important implications for nutrient monitoring, management and policy,

particularly in tropical rivers. A defining characteristic of tropical systems is the marked

seasonality of surface water flows, driven by the distinct wet season-dry season rainfall patterns.

Accordingly, there is a shift in the dominant sources of water contributing to the lower Herbert

River. As a consequence of the shift from overland flow to baseflow dominance, the chemistry

of the river changes, including with respect to dissolved inorganic forms of N. In general, N

monitoring and management of tropical rivers has traditionally focussed on high intensity flow

events that result in high total N loads in surface waters, especially of particulate forms.

However, the results of this thesis demonstrate that significant concentrations of dissolved N are

also present in surface waters of the Herbert River throughout the dry season, due to N inputs

from groundwater sources. Even though wet season flows dissipate rapidly, longer residence

times of water during low flow conditions provide greater opportunity for nutrient contributions

to impact on river health. In addition, the analyses highlight that the concentration of DIN varies

within the alluvial aquifers contributing NO3- and that there are hotspots of high nutrient

availability in the groundwater system. The results clearly illustrate that there is considerable

variation in measured concentrations of DIN, particularly NO3-, along the entire length of the

lower Herbert River. Importantly, NO3- concentrations along particular reaches of the river are

at potentially harmful levels for the health of riverine ecosystems.

The results from this research in the lower Herbert River catchment, combined with the general

characteristics of tropical rivers, underline the following key points and recommendations in

regards to nutrient monitoring, management and policy guidelines in connected river-aquifer

systems within the tropics.

• Water quality sampling should be undertaken at recognised periods on the

stream/groundwater hydrograph, with an understanding of temporal and spatial river-

aquifer connectivity relationships. Sampling during low flow conditions, particularly at

the end of the dry season, is the most effective time for isolating the groundwater signal

and hence evaluating potential nutrient contributions in gaining stream situations.

• Surface and subsurface sources of water and dissolved nutrients must be considered.

Hence, nutrient hotpots in both surface water and groundwater systems should be

identified and targeted for land management actions.

Page 275: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Research Conclusions

251

• Sampling location, as well as timing, should be carefully chosen. End-of-river

monitoring does not necessarily capture the maximum nutrient concentrations along

upstream reaches and hence may not reflect the magnitude of potential impacts on

ecosystem health. Longitudinal sampling is considered to be of greatest relevance for

assessing river health.

• Appropriate water quality guideline values must be set that account for seasonal

changes in both the sources that deliver nutrients and the forms of N transported to

surface waters. Single water quality targets are not appropriate in distinctly seasonal

catchments.

8.6 FURTHER RESEARCH

Whilst the aims of the thesis were adequately addressed, further work would improve spatial

and temporal process understanding of the dynamics of water and N movement in the river-

aquifer system of the Herbert River catchment. Furthermore, this thesis provides a platform for

undertaking a catchment-scale nutrient budget or other modelling approach, which could

potentially be applied to other case study areas.

This research identifies a number of areas for future work:

(1) Improvements to the hydrogeological framework

This could be achieved by:

• determining aquifer hydraulic properties over a greater spatial extent, both laterally and

vertically, including measurements in multiple sandy units;

• measuring groundwater levels and groundwater quality at close proximity to the river

during the wet and dry seasons;

• analysing deep aquifer samples for isotopes such as 36Cl, SF6 and CFC’s to estimate

groundwater residence time; and

• strategic development of additional monitoring bores coupled with further enhanced

surveying techniques (e.g. ground and airborne geophysics).

Page 276: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Chapter 8

252

(2) Improvements to the spatial and temporal understanding of river-groundwater fluxes

Refinements could include:

• collection of stream hydrograph data at multiple gauges along the river;

• measurement of groundwater elevations close to the river and at a greater number of

sites adjacent to and within the river;

• installation of automatic piezometers in the river and along transects away from the

river, as well as temperature and salinity probes;

• repetition of radon sampling in the river and its tributaries at the beginning, end and part

way through the dry season, including during different tidal cycles in the estuary; and

• accurate determination of river parameters such as river depth, width and stream

discharge at all radon sampling sites, as well as increasing the spatial extent of radon

measurements in the aquifers. These improvements could provide greater confidence in

groundwater flux estimates.

(3) Improvements to inputs of a N mass balance

Not all forms of N were measured in surface water and groundwater in this study. Therefore, to

complement the existing data, additional samples could be collected at the beginning and end of

the dry season for analysis of:

• gaseous forms of N such as nitrous oxide and nitrogen gas and dissolved organic N in

the river and in groundwater proximal to the river;

• particulate N in surface waters; and

• dissolved and particulate N in adjacent seawater.

In addition, further work could include:

• examination of N transformations in groundwater through N isotopes and other redox-

sensitive tracers;

• geochemical modelling to identify where changes in groundwater chemistry can

account for changes in N; and

• improvement to the quantification of water movement between the aquifers and the

river in order to quantify the flux of N between them.

Page 277: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

Research Conclusions

253

8.7 CONCLUDING STATEMENT

The results from this thesis, based on a case study catchment in the tropics of northeastern

Australia, demonstrate that groundwater can be an important vector for the delivery of dissolved

forms of nitrogen to surface waters during low flow conditions. Therefore, characterising the

interaction between surface water and groundwater resources, through a combination of

complementary analytical techniques, is crucial for effective management of water quality. It is

hoped that this research also provides a new perspective for monitoring and management of

nutrients, particularly in tropical river systems.

Page 278: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to
Page 279: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

255

References

Ahern KS, Udy JW and Pointon SM (2006) Investigating the potential for groundwater from

different vegetation, soil and landuses to stimulate blooms of the cyanobacterium, Lyngbya

majuscula, in coastal waters. Marine and Freshwater Research, 57(2):177-186

ANZECC and ARMCANZ (1994) 'Water quality management - an outline of the policies',

National Water Quality Management Strategy, Australian and New Zealand Environment and

Conservation Council and Agriculture and Resource Management Council of Australia and

New Zealand

ANZECC and ARMCANZ (2000) 'Australian and New Zealand Guidelines for Fresh and

Marine Water Quality', Australian and New Zealand Environment and Conservation Council

and Agriculture and Resource Management Council of Australia and New Zealand, National

Water Quality Management Strategy, http://www.mincos.gov.au/pub_anzwq.html

Appelo CAJ (1994) Cation and proton exchange, pH variations, and carbonate reactions in a

freshening aquifer. Water Resources Research, 30(10):2793-2806

Appelo CAJ and Postma D (1994) 'Geochemistry, groundwater and pollution.' A.A. Balkema:

Rotterdam, 536 pp.

ARMCANZ (1996) 'Allocation and use of groundwater: a national framework for improved

groundwater management in Australia', Policy position paper for advice to States and Territories

Occasional paper number 2, Agriculture and Resource Management Council of Australia and

New Zealand

ARMCANZ (1998) 'Special Water Meeting: Resolution No.6 Groundwater Management',

Agriculture and Resource Management Council of Australia and New Zealand,

www.affa.gov.au, accessed 19/09/2003

ARMCANZ and ANZECC (1994) 'Policies and principles - a reference document', National

Water Quality Management Strategy, Agriculture and Resource Management Council of

Australia and New Zealand and Australian and New Zealand Environment and Conservation

Council

Page 280: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

References

256

ARMCANZ and ANZECC (1995) 'Guidelines for groundwater protection in Australia',

National Water Quality Management Strategy, Agriculture and Resource Management Council

of Australia and New Zealand and Australian and New Zealand Environment and Conservation

Council

Armstrong D and Narayan K (1998) Using groundwater responses to infer recharge. In 'The

basics of recharge and discharge, Part 5'. (Eds L Zhang and G Walker), CSIRO Australia

Back W (1960) Origin of hydrochemical facies of ground water in the Atlantic Coastal Plain. In

'Benchmark papers in Geology, v. 73: Chemical hydrogeology'. (Eds W Back and RA Freeze),

pp. 79-95, Hutchinson Ross Publishing Company: Stroudsberg

Bartley R, Henderson A, Prosser IP, Hughes AO, McKergow L, Lu H, Brodie J, Bainbridge Z

and Roth CH (2003) 'Patterns of erosion and sediment and nutrient transport in the Herbert

River Catchment, Queensland', Consultancy Report, CSIRO Land and Water

Bohl HP, Mitchell D, Fanning D and Roth CH (2000a) Water and nitrogen balance for a heavy

floodplain soil under sugarcane in the Australian Wet Tropics. Proc. NZSSS/ASSSI Soil 2000

Conference, Canterbury NZ

Bohl HP, Mitchell DC, Penny RS and Roth CH (2000b) Nitrogen losses via subsurface flow

from sugar cane on floodplain soils in the Australian Wet Tropics. Proc. Aust. Soc. Sugar Cane

Technol., 22:302-307

Bohl HP, Roth CH, Tetzlaff D and Timmer J (2001) Estimation of groundwater recharge and

nitrogen leaching under sugarcane in the Ripple Creek catchment, lower Herbert. Proc. Aust.

Soc. Sugar Cane Technol., 23:84-89

Bohlke JK and Denver JM (1995) Combined use of groundwater dating, chemical, and isotopic

analyses to resolve the history and fate of nitrate contamination in two agricultural watersheds,

Atlantic coastal plain, Maryland. Water Resources Research, 31(9):2319-2340

Bonell M (2005) Runoff generation in tropical forests. In 'Forests, water, and people in the

humid tropics: past, present, and future hydrological research for integrated land and water

management'. (Eds M Bonell and LA Bruijnzeel), pp. 314-406, UNESCO, International

hydrology series, Cambridge University Press: Cambridge

Bonell M and Balek J (1993) Recent scientific developments and research needs in hydrological

processes of the humid tropics. In 'Hydrology and water management in the humid tropics -

hydrological research issues and strategies for water management'. (Eds M Bonell, MM

Hufschmidt and JS Gladwell), pp. 167-260, UNESCO - Cambridge University Press:

Cambridge

Page 281: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

References

257

Bonell M, Barnes CJ, Grant CR, Howard A and Burns J (1998) High rainfall response-

dominated catchments: a comparative study of experiments in tropical north-east Queensland

with temperate New Zealand. In 'Isotope tracers in catchment hydrology'. (Eds C Kendall and JJ

McDonell), pp. 347-390, Elsevier

Bonell M and Bruijnzeel LA (Eds) (2005) 'Forests, water, and people in the humid tropics: past,

present, and future hydrological research for integrated land and water management', UNESCO,

International hydrology series, Cambridge University Press: Cambridge

Bonell M, Hufschmidt MM and Gladwell JS (Eds) (1993) 'Hydrology and water management in

the humid tropics - hydrological research issues and strategies for water management',

UNESCO - Cambridge University Press: Cambridge

Boulton AJ, Findlay S, Marmonier P, Stanley EH and Valett HM (1998) The functional

significance of the hyporheic zone in streams and rivers. Annu. Rev. Ecol. Syst., 29:59–81

Bouwer H and Maddock T (1997) Making sense of the interactions between groundwater and

streamflow: lessons for water masters and adjudicators. Rivers, 6(1):19-31

Bramley RGV and Johnson AKL (1996) Land use impacts on nutrient loading in the Herbert

River. Proceedings from the conference of the downstream effects of land use, Brisbane, (Eds

HM Hunter, AG Eyles and GE Rayment), pp. 93-96, Department of Natural Resources

Bramley RGV and Muller DE (1999) 'Water quality in the lower Herbert River - the CSIRO

dataset', CSIRO Land and Water

Bramley RGV and Roth CH (2002) Land-use effects on water quality in an intensively managed

catchment in the Australian humid tropics. Marine and Freshwater Research, 53(5):931-940

Bratten R and Gates G (2003) Groundwater-surface water interaction in inland New South

Wales: a scoping study. Water science and technology, 48(7):215-224

Bristow KL, Thorburn PJ, Sweeney CA and Bohl HP (1998) 'Water and nitrogen balance in

natural and agricultural systems in the wet tropics of north Queensland: a review', LWRRDC

Occasional Paper RAPPS03/98

Brodie J, Christie C, Devlin M, Haynes D, Morris S, Ramsay M, Waterhouse J and Yorkston H

(2001) Catchment management and the Great Barrier Reef. Water Science and Technology,

43(9):203-211

Page 282: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

References

258

Brodie J and Mitchell A (2005) Nutrients in Australian tropical rivers: changes with agricultural

development and implications for receiving environments. Marine and Freshwater Research,

56(3):279-302

Brodie R, Sundaram B, Tottenham R, Hostetler S and Ransley T (2007) 'An overview of tools

for assessing groundwater-surface water connectivity', Bureau of Rural Sciences

Brunke M and Gonser T (1997) The ecological significance of exchange processes between

rivers and ground-water. Freshwater Biology, 37:1-33

Brunskill G (2000) Summary of continental shelf research for the Herbert River Sector of the

Great Barrier Reef Lagoon, North Queensland. In 'Herbert River Integration Study 1st HRIS

technical workshop - Report', CSIRO: Townsville

Bureau of Mineral Resources (1965) 1:250,000 geological series sheet SE 55-10 (Ingham)

Buttle JM (1998) Fundamentals of small catchment hydrology. In ' Isotope tracers in catchment

hydrology'. (Eds C Kendall and JJ McDonell), pp. 1-49, Elsevier

Callaghan J and Bonell M (2005) An overview of the meteorology and climatology of the

humid tropics. In 'Forests, water, and people in the humid Tropics: past, present, and future

hydrological research for integrated land and water management'. (Eds M Bonell and LA

Bruijnzeel), pp. 158-193, International hydrology series, Cambridge University Press:

Cambridge

Calvert FJ (1959) 'Notes on groundwater investigations Herbert River Delta', Unpublished

report, Queensland Water Resources Commission

Canter LW (1997) 'Nitrates in groundwater.' CRC Press, Inc.: Boca Raton, 263 pp.

Chang J-H and Lau LS (1983) Definition of the humid tropics. Paper presented at the IAHS

Symposium, Hydrology of humid tropical regions, Hamburg, Germany

Chapman T (1999) A comparison of algorithms for stream flow recession and baseflow

separation. Hydrological Processes, 13(5):701-714

Chebotarev II (1955) Metamorphism of natural waters in the crust of weathering. Geochimica

Et Cosmochimica Acta, 8:22-48, 137-170, 198-212

Cirmo CP and McDonnell JJ (1997) Linking the hydrologic and biogeochemical controls of

nitrogen transport in near-stream zones of temperate-forested catchments: a review. Journal of

Hydrology, 199(1-2):88-120

Page 283: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

References

259

Clark I and Fritz P (1997) 'Environmental isoptopes in hydrogeology.' CRC Press LLC: Boca

Raton, New York, 328 pp.

COAG (2004) 'Intergovernmental agreement on a national water initiative',

http://www.coag.gov.au/meetings/250604/#water_initiative

Cook PG (2005) Groundwater chemistry and environmental isotopes as tracers. Workshop on

groundwater surface water interaction in the tropics, Darwin, International Association of

Hydrogeologists

Cook PG, Bohlke JK and Solomon DK (2002) Measuring groundwater recharge and discharge

using environmental tracers. In 'Proceedings of the international groundwater conference,

balancing the groundwater budget', International Association of Hydrogeologists: Darwin

Cook PG, Favreau G, Dighton JC and Tickell S (2003) Determining natural groundwater influx

to a tropical river using radon, chlorofluorocarbons and ionic environmental tracers. Journal of

Hydrology, 277(1-2):74-88

Cook PG and Herczeg A (1998) Groundwater chemical methods for recharge studies. In 'The

basics of recharge and discharge; Part II'. (Ed. L Zhang), CSIRO Australia

Cook PG, Stieglitz T and Clark J (2004) 'Groundwater discharge from the Burdekin floodplain

aquifer, North Queensland', Technical Report 26/04, CSIRO Land & Water

Council Directive 91/676/EEC of 12 December 1991 concerning the protection of waters

against pollution caused by nitrates from agricultural sources, Official Journal L 375 of

31.12.1991, p. 1

Cox RB (1979) 'Herbert River groundwater investigation report on hydrogeology of the Herbert

Delta, Ingham, vol 1-2', Queensland Water Resources Commission

Craig H (1961) Isotopic variations in meteoric waters. Science, 133:1833-1834

Cresswell RG and Herczeg AL (2004) 'Groundwater recharge, mixing and salinity across the

Angas-Bremer Plains, South Australia: geochemical and isotopic constraints', Technical report

29/04, CSIRO Land & Water

Croke BFW (in prep.) The use of data analysis techniques in hydrology.

Dahm CN, Grimm NB, Marmonier P, Valett MH and Vervier P (1998) Nutrient dynamics at the

interface between surface waters and groundwaters. Freshwater Biology, 40:427-451

Page 284: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

References

260

De Keyser F (1963) The Palmerville Fault - a fundamental structure in north Queensland.

Journal of the Geological Society of Australia, 10:273-278

Dixon-Jain P, Croke BFW, Letcher R and Sims JP (2005) Managing water quality in river and

groundwater systems: a case study from the Herbert River catchment, North Queensland. Water

Capital, 29th Hydrology and Water Resources Symposium, 21-23 February 2005, Canberra

ACT

Domenico P and Schwartz FW (1990) 'Physical and chemical hydrogeology.' John Wiley &

Sons, 506 pp.

Douglas I and Guyot JL (2005) Erosion and sediment yield in the humid tropics. In 'Forests,

water, and people in the humid Tropics: past, present, and future hydrological research for

integrated land and water management'. (Eds M Bonell and LA Bruijnzeel), pp. 407-421,

UNESCO, International hydrology series, Cambridge University Press: Cambridge

EPA (2006) 'Queensland water quality guidelines, version 2', The State of Queensland,

Environmental Protection Agency,

http://www.epa.qld.gov.au/publications/p01414ak.pdf/Queensland_water_quality_guidelines_2

006.pdf

Evans R (2005) Keynote address. Groundwater surface water interaction in the Tropics, IAH

Northern Territory Branch, Darwin

Eyre BD and Pont D (2003) Intra- and inter-annual variability in the different forms of diffuse

nitrogen and phosphorus delivered to seven sub-tropical east Australian estuaries. Estuarine,

Coastal and Shelf Science, 57(1-2):137-148

Fabricius KE (2005) Effects of terrestrial runoff on the ecology of corals and coral reefs: review

and synthesis. Marine Pollution Bulletin, 50:125-146

Fetter CW (1988) 'Applied hydrogeology.' Merril publishing company: Ohio, 691 pp.

Fleming PM (1993) The impact of land-use change on water resources in the tropics: an

Australian view of the scientific issues. In 'Hydrology and water management in the humid

tropics - hydrological research issues and strategies for water management'. (Eds M Bonell, MM

Hufschmidt and JS Gladwell), pp. 405-413, UNESCO, Cambridge University Press: Cambridge

Foster SSD and Chilton PJ (1993) Groundwater systems in the humid tropics. In 'Hydrology and

water management in the humid tropics - hydrological research issues and strategies for water

management'. (Eds M Bonell, MM Hufschmidt and JS Gladwell), pp. 261-272, UNESCO,

Cambridge University Press: Cambridge

Page 285: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

References

261

Freeze RA and Cherry JA (1979) 'Groundwater.' Prentice-Hall, Inc: New Jersey, 604 pp.

Freyney JR, Denmead OT, Wood AW and Saffigna PG (1994) Ammonia loss following urea

addition to sugar cane trash blankets. Proceedings of the Australian Society of Sugar Cane

Technologists:114-121

Furey PR and Gupta VK (2001) A physically based filter for separating base flow from

streamflow time series. Water Resources Research, 37(11):2709-2722

Furnas M (2003) 'Catchments and corals: terrestrial runoff to the Great Barrier Reef.' Australian

Institute of Marine Science and CRC Reef Research Centre: Townsville, 334 pp.

Furnas M and Mitchell A (2000) 'Sediment and nutrient exports from the Herbert River

catchment', Herbert River Integration Study 1st HRIS technical workshop - Report, CSIRO

Furnas M and Mitchell A (2001) Runoff of terrestrial sediment and nutrients into the Great

Barrier Reef World Heritage Area. In 'Oceanographic processes of coral reefs: physical and

biological links in the Great Barrier Reef'. (Ed. E Wolanski), pp. 37-51, CRC Press: Boca Raton

Furnas M, Mitchell AW and Skuza M (1995) 'Nitrogen and phosphorus budgets for the Great

Barrier Reef', Research Publication No. 36, Great Barrier Reef Marine Park Authority

Gagan MK, Ayliffe LK, Opdyke BN, Hopley D, Scott-Gagan H and Cowley J (2002) Coral

oxygen isotope evidence for recent groundwater fluxes to the Australian Great Barrier Reef.

Geophysical Research Letters, 29(20)

Gat JR (1980) The isotopes of hydrogen and oxygen in precipitation. In 'Handbook of

environmental isotope geochemistry, Vol 1, The terrestrial environment'. (Eds P Fritz and J-C

Fontes), pp. 21-48, A. Elsevier: Amsterdam

Germanov AI, Volkov GA, Lisitsin AK and Serebennikov VS (1958) Investigation of the

oxidation reduction potential of ground waters. Geokhimiya:322-329

Gibbs RJ (1970) Mechanisms controlling world water chemistry. Science, 170:1088-90

Gonfiantini R, Klaus F, Araguas LA and Rozanski K (1998) Isotopes in groundwater

hydrology. In 'Isotope tracers in catchment hydrology'. (Eds C Kendall and JJ McDonnel), pp.

203-246, Elsevier

Grayson RB, Argent RM, Nathon RJ, McMahon TA and Main RG (1996) 'Hydrological

recipes. Estimation techniques in Australian hydrology.' CRC for Catchment Hydrology, Dept

of Civil Engineering, Monash University, pp.

Page 286: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

References

262

Grimaldi C, Viaud V, Massa F, Carteaux L, Derosch S, Regeard A, Fauvel Y, Gilliet N and

Rouault F (2004) Stream nitrate variations explained by ground water head fluctuations in a

pyrite-bearing aquifer. Journal of Environmental Quality, 33(3):994-1001

Hamilton SK and Gehrke PC (2005) Australia's tropical river systems: current scientific

understanding and critical knowledge gaps for sustainable management. Marine and Freshwater

Research, 56(3):243-252

Harris GP (2001) Biogeochemistry of nitrogen and phosphorus in Australian catchments, rivers

and estuaries: effects of land use and flow regulation and comparisons with global patterns.

Marine and Freshwater Research, 52:139-149

Hatton TJ (Ed.) (1998) 'Catchment scale recharge modelling', The basics of recharge and

discharge, Part 4, CSIRO Australia

Heath RC (1987) 'Basic ground-water hydrology', Geological Survey water-supply paper 2220,

North Carolina Department of Natural Resources and Community Development

Heathwaite AL, Burt TP and Trudgill ST (1989) 'Runoff, sediment and solute delivery in

agricultural drainage basins - a scale dependent approach', International Association of

Hydrological Sciences Publication 182:175-191, IAHS Press

Heathwaite L, Sharpley A and Gburek W (2000) A conceptual approach for integrating

phosphorus and nitrogen management at watershed scales. Journal of environmental quality,

29(1):158-166

Hefting M, Beltman B, Karssenberg D, Rebel K, van Riessen M and Spijker M (2006) Water

quality dynamics and hydrology in nitrate loaded riparian zones in the Netherlands.

Environmental Pollution, 139(1):143-156

Hem JD (1985) 'Study and interpretation of the chemical characteristics of natural water', Water

Supply Paper 2254, U.S Geological Survey

Herbert G (1994) 'Coastal groundwater hydrology - Herbert delta', office memo., Queensland

Department of Primary Industries

Herczeg A, Dighton J, Easterbrook M and Salomans E (1994) Radon-222 and Ra-226

measurements in Australian groundwaters using liquid scintillation counting. Proceedings of the

workshop on radon and radon progeny measurements in environmental samples, Canberra, pp.

53-57

Page 287: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

References

263

Herczeg A and Edmunds W (2000) Inorganic ions as tracers. In 'Environmental tracers in

subsurface hydrology'. (Eds PG Cook and AL Herczeg), pp. 35-76, Kluwer Academic: Boston

Herczeg A, Lamontagne S, Pritchard J, Leaney F, Dighton J, Jiwan J and Smith P (2001)

Groundwater – surface water interactions: testing conceptual models with environmental

tracers. Proceedings of the 8th Murray-Darling Basin groundwater workshop 2001, Victor

Harbour, South Australia.

Hill AR (1996) Nitrate removal in stream riparian zones. Journal of environmental quality,

25:743-755

Holmes KH, Stephens AW, Jones MR and Searle DE (1991) 'Quaternary geology of the coast

between Cape Pallarenda and Cape Grafton - notes accompanying 1:50 000 map sheets.'

Queensland resource industry record, 1991/21

Hooper RP and Shoemaker CA (1986) A comparision of chemical and isotopic hydrograph

separation. Water Resources Research, 22:1444-1454

Hunter HM and Walton RS (1997) 'From land to lagoon: land use impacts on water quality in

the Johnstone River catchment', Department of Natural Resources

IAEA (2006) Oxygen-18 and deuterium stable isotope data. Isotope hydrology section database,

International Atomic Energy Agency, http://isohis.iaea.org

Ingraham NL (1998) Isotopic variations in precipitation. In 'Isotope tracers in catchment

hydrology'. (Eds C Kendall and JJ McDonnel), pp. 87-118, Elsevier

Johnson AKL, Ebert SP and Murray AE (1999) Distribution of coastal freshwater wetlands and

riparian forests in the Herbert River catchment and implications for management of catchments

adjacent the Great Barrier Reef Marine Park. Environmental Conservation, 26(3):229-235

Johnson AKL, Ebert SP and Murray AE (2000) Land cover change and its environmental

significance in the Herbert River catchment, north-east Queensland. Australian Geographer,

31(1):75-86

Johnson AKL and Murray AE (1997) 'Herbert River catchment atlas.' CSIRO Tropical

Agriculture: Townsville, 48 pp.

Kendall C (1998) Fundamentals of small catchment hydrology. In 'Isotope tracers in catchment

hydrology'. (Eds C Kendall and JJ McDonell), pp. 519-576, Elsevier

Korom SF (1992) Natural denitrification in the saturated zone: a review. Water Resources

Research, 28(6):1657-1668

Page 288: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

References

264

Kuhanesan S, Scriven DC, Huxley WJ, Keating BA and Verburg K (1998) Modelling nitrate

movement in the Gooburrum aquifer complex, Bundaberg, Queensland. Proceedings of the

International Groundwater Conference, Groundwater Sustainable Solutions, Melbourne, pp.

491-496

Lamontagne S, Herczeg A, Leaney F, Dighton J, Pritchard J, Ullman W and Jiwan J (2001)

Nitrogen attenuation by stream riparian zones: prospects for Australian landscapes. Proceedings

of the 8th Murray-Darling Basin groundwater workshop 2001, Victor Harbour, South Australia

Lamontagne S, Herczeg AL, Dighton JC, Jiwan JS and Pritchard JL (2005) Patterns in

groundwater nitrogen concentration in the floodplain of a subtropical stream (Wollombi Brook,

New South Wales). Biogeochemistry, 72(2):169 - 190

Lamontagne S, Herczeg AL, Dighton JC, Pritchard JL, Jiwan JS and Ullman WJ (2003)

'Groundwater–surface water interactions between streams and alluvial aquifers: Results from

the Wollombi Brook (NSW) study (Part II – Biogeochemical processes)', Technical report

42/03, CSIRO Land and Water

Lamontagne S, Leaney F and Herczeg A (2002) 'Streamwater-groundwater interaction: the river

Murray at Hattah-Kulkyne Park, Victoria', Technical report 27/02, CSIRO Land and Water

Langmuir D (1997) 'Aqueous environmental geochemistry.' Prentice Hall: New Jersey, 600 pp.

Latrubesse EM, Stevaux JC and Sinha R (2005) Tropical rivers. Geomorphology, 70(3-4):187-

206

Laudon H and Slaymaker O (1997) Hydrograph separation using stable isotopes, silica and

electrical conductivity: an Alpine example. Journal of Hydrology, 201:82-101

Lawrence CR (1983) 'Nitrate-rich groundwaters of Australia', Australian Water Resources

Council, technical paper no. 79

Letcher RA, Jakeman AJ and Ekasingh B (2005) Principles of integrated assessment, Chapter 3.

In 'Integrating Knowledge for River Basin Management. Progress in Thailand'. (Ed. AJ

Jakeman, Letcher, R.A., Rojanasoonthon, S., Cuddy, S.M.), pp. 55-66, Monograph No.118,

ACIAR

Linderfelt WR and Turner JV (2001) Interaction between shallow groundwater, saline surface

water and nutrient discharge in a seasonal estuary: the Swan-Canning system. Hydrological

Processes, 15(13):2631-2653

Page 289: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

References

265

Linsley RK, Kohler MA, Paulhus JLH and Wallace JS (1958) 'Hydrology for engineers.'

McGraw Hill: New York, 508 pp.

Loaiciga HA and Zekster IS (2002) Estimation of submarine groundwater discharge. Water

Resources, 30(5):473-479

Loucks D and Gladwell J (1999) 'Sustainability criteria for water resource systems.' Cambridge

University Press: Cambridge, 139 pp.

Lyne V and Hollick M (1979) Stochastic time-variable rainfall-runoff modeling. Institute of

Engineers Australia National Conference Publication, Canberra, pp. 89-93, Institute of

Engineers Australia

Malard F, Tockner K, Dole-Olivier M and Ward JV (2002) A landscape perspective of surface-

subsurface hydrological exchanges in river corridors. Freshwater Biology, 47(4):621-640

Mazor E (1991) 'Applied chemical and isotopic groundwater hydrology.' Open University Press:

Buckingham, 413 pp.

McKergow LA, Prosser IP, Grayson RB and Heiner D (2004) Performance of grass and

rainforest riparian buffers in the wet tropics, Far North Queensland. 1. Riparian hydrology.

Australian Journal of Soil Research, 42(473-484)

McMahon TA, Finlayson BL, Srikanthan R and Haines AT (1992) 'Global runoff: continental

comparisons of annual flows and peak discharges.' Catena Verlag: Cremlingen-Destedt, West

Germany, 166 pp.

McNamara JP, Kane DL and Hinzman LD (1997) Hydrograph separations in an Arctic

watershed using mixing model and graphical techniques. Water Resources Research, 33:1707-

1720

McNeil VH and Cox ME (2000) Relationship between conductivity and analysed composition

in a large set of natural surface-water samples, Queensland, Australia. Environmental Geology,

V39(12):1325-1333

McNeil VH, Cox ME and Preda M (2005) Assessment of chemical water types and their spatial

variation using multi-stage cluster analysis, Queensland, Australia. Journal of Hydrology,

310(1-4):181-200

MDBC (1997) 'Murray-Darling Basin groundwater quality sampling guidelines', Technical

Report no. 3, Groundwater Working Group

Page 290: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

References

266

Meier EA, Thorburn PJ and Probert ME (2006) Occurrence and simulation of nitrification in

two contrasting sugarcane soils from the Australian wet tropics. Australian Journal of Soil

Research, 44(1):1-9

Merrill AG and Benning TL (2006) Ecosystem type differences in nitrogen process rates and

controls in the riparian zone of a montane landscape. Forest Ecology and Management, 222(1-

3):145-161

Merritt WS, Letcher RA and Jakeman AJ (2003) A review of erosion and sediment transport

models. Environmental Modelling and Software, 18:761-799

Meybeck J (1987) Global chemical weathering of surficial rocks estimated from river dissolved

loads. American Journal of Science, 287:401-428

Meybeck M (2004) Global occurrence of major elements in rivers. In 'Surface and ground

water, weathering and soils (ed. JI Drever) Vol 9 Treatise on geochemistry'. (Eds HD Holland

and KK Turekian), pp. 207-224, Elsevier-Pergamon: Oxford

Mitchell AW, Bramley RGV and Johnson AKL (1997) Export of nutrients and suspended

sediment during a cyclone-mediated flood event in the Herbert River catchment, Australia.

Marine and Freshwater Research, 48(1):79-88

Mitchell AW, Reghenzani JR and Furnas MJ (2001) Nitrogen levels in the Tully River - a long-

term view. Water Science and Technology, 43(9):99-105

Nathan RJ and McMahon TA (1990) Evaluation of automated techniques for base flow and

recession analyses. Water Resources Research, 26(7):1465-1473

Neal C and Heathwaite AL (2005) Nutrient mobility within river basins: a European

perspective. Journal of Hydrology, 304(1-4):477-490

Nevill J, Maher M and Nichols P (2001) Water law, COAG, and the environment. Third

Australasian Natural Resources Law and Policy Conference, Adelaide, South Australia, pp. 93-

107

Oenema O, Kros H and de Vries W (2003) Approaches and uncertainties in nutrient budgets:

implications for nutrient management and environmental policies. European Journal of

Agronomy, 20(1-2):3-16

Parkhurst DL (1995) 'Users guide to PHREEQC - a computer program for speciation, reaction-

path, advective-transport, and inverse geochemical calculations', U.S. Geological Survey, Water

Resources Investigations Report 95-4227

Page 291: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

References

267

Pidwirny M (2005) Introduction to the biosphere: the nitrogen cycle. In 'Fundamentals of

physical geography (2nd edition)',

http://www.physicalgeography.net/fundamentals/chapter9.html, accessed 29/6/2005

Piper AM (1944) A graphic procedure in the geochemical interpretation of water-analyses. Am.

Geophys. Union Trans, 25:914-923

Productivity Commission (2003) 'Industries, land use and water quality in the Great Barrier

Reef Catchment', Research Report

Prosser IP, Rustomji P, Young WJ, Moran CJ and Hughes A (2001) 'Constructing river basin

sediment budgets for the National Land and Water Resources Audit', CSIRO Land and Water

Technical Report 15/01, CSIRO

Prove BG, Moody PW and Reghenzani JR (1997) 'Nutrient balance and transport from

agricultural and rainforest lands: a case study of the Johnstone River Catchment', DAQ3S Final

report, Queensland Department of Natural Resources and BSES

Rasiah V and Armour JD (2001) Nitrate accumulation under cropping in the ferrosols of far

north Queensland Wet Tropics. Australian Journal of Soil Research, 39:329-341

Rasiah V, Armour JD, Yamamoto T, Mahendrarajah S and Heiner DH (2003) Nitrate dynamics

in shallow groundwater and the potential for transport to off-site water bodies. Water, Air, &

Soil Pollution, (147):183-202

Rassam DW, Fellows CS, Hayr RD, Hunter H and Bloesch P (2006) The hydrology of riparian

buffer zones; two case studies in an ephemeral and a perennial stream. Journal of Hydrology,

325(1-4):308-324

Reghenzani JR, Armour JD, Prove BG, Moddy PW and McShane TJ (1996) Nutrient balances

for sugarcane plant and first ratoon crops in the Wet Tropics. Sugarcane: research towards

efficient and sustainable production. Sugar 2000 Symposium, Brisbane, (Eds JR Wilson, DM

Hogart, JA Campbell and AL Garside), pp. 275-277, CSIRO Division of Tropical Crops and

Pastures

REM (2002) 'WaterMark: sustainable groundwater use within irrigated regions. Project 2:

Conjunctive resource management, milestone 2 final report. A review of stream-aquifer

interaction assessment methods.' Prepared for the Murray-Darling Basin Commission

Rienks IP, Gunther MC, Bultitude RJ, Morwood DA, Dash PH, Withnall IW and Fanning CM

(2000) 'Ingham, second edition. Queensland 1:250 000 Geological Series Explanatory Notes',

Geological Survey of Queensland, Queensland Department of Mines and Energy

Page 292: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

References

268

Roche MA (1993) Water quality issues in the humid tropics. In 'Hydrology and water

management in the humid tropics - hydrological research issues and strategies for water

management'. (Eds M Bonell, MM Hufschmidt and JS Gladwell), pp. 344-366, UNESCO -

Cambridge University Press: Cambridge

Rose CW (1993) Erosion and sedimentation. In 'Hydrology and water management in the humid

tropics - hydrological research issues and strategies for water management'. (Eds M Bonell, MM

Hufschmidt and JS Gladwell), pp. 167-260, UNESCO - Cambridge University Press:

Cambridge

Schoeller H (1955) Geochemie des eaux souterraines. Rev Inst Fr. Petrol., 10:230-244

Silver WL, Herman DJ and Firestone MK (2001) Dissimilatory nitrate reduction to ammonium

in upland tropical forest soils. Ecology, 82(9):2410-2416

SKM (2001) 'Survey of baseflows in unregulated catchments of the Murray Darling Basin',

Murray-Darling Basin Commission Project R2005

Smakhtin VU (2001) Low flow hydrology: a review. Journal of Hydrology, 240(3-4):147-186

Smith AJ and Turner JV (2001) Density-dependent surface water-groundwater interaction and

nutrient discharge in the Swan-Canning Estuary. Hydrological Processes, 15(13):2595-2616

Sophocleous M (2002) Interactions between groundwater and surface water: the state of the

science. Hydrogeology Journal, 10(1):52-67

Sponberg ME (2000) Spectral analysis of base flow separation with digital filters. Water

Resources Research, 36(3):745-752

Statutory Instrument No. 1289 (2006) The protection of water against agricultural nitrate

pollution (England and Wales) (Amendment), Regulations 2006

Stewart BJ (1993) The hydrology and water resources of humid northern Australia and Papua

New Guinea. In 'Hydrology and water management in the humid tropics - hydrological research

issues and strategies for water management'. (Eds M Bonell, MM Hufschmidt and JS Gladwell),

pp. 67-83, UNESCO, Cambridge University Press: Cambridge

Stewart LK, Charlesworth PB, Bristow KL and Thorburn PJ (2005) Estimating deep drainage

and nitrate leaching from the root zone under sugarcane using APSIM-SWIM. Agricultural

Water Management, 81(3):315-334

Stieglitz T (2005) Submarine groundwater discharge into the near-shore zone of the Great

Barrier Reef, Australia. Marine Pollution Bulletin, 51(1-4):51-59

Page 293: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

References

269

Stieglitz TC and Ridd PV (2000) Submarine groundwater discharge from paleochannels?

Wonky holes in the inner shelf of the Great Barrier Reef. HYDRO 2000 - 26th National and 3rd

International Hydrology and Water Resources Symposium of the Institute of Engineers,

Australia, Perth, pp. 1265-1271

Taylor JR (1982) 'An introduction to error analysis: the study of uncertainties in physical

measurements.' University Science Books: Mill Valley, California, 270 pp.

Thayalakumaran T, Charlesworth P and Bristow KL (2004) 'Assessment of the geochemical

environment in the lower Burdekin aquifer: implications from the removal of nitrate through

denitrification', Technical report 32/04, CSIRO Land & Water

Thorburn PJ, Biggs JS, Weier KL and Keating BA (2003) Nitrate in groundwaters of intensive

agricultural areas in coastal Northeastern Australia. Agriculture, Ecosystems and Environment,

94:49-58

Thorburn PJ, Meier EA and Probert ME (2005) Modelling nitrogen dynamics in sugarcane

systems: recent advances and applications. Field Crops Research, 92(2-3):337-351

Thorburn PJ, Probert, M.E.,Lisson, S.N., Wood, A.W. and Keating, B.A. (1999) Impacts of

trash retention on soil nitrogen and water: an example from the Australian sugarcane industry.

Proc. S. Afr. Sug. Technol. Ass., 73:75-79

Ticehurst JL, Newham LTH, Rissik D, Letcher RA and Jakeman AJ (2007) A Bayesian network

approach for assessing the sustainability of coastal lakes in New South Wales, Australia.

Environmental Modelling & Software, 22(8):1129-1139

Toth J (1963) A theoretical analysis of groundwater flow in small drainage basins. Journal of

Geophysical Research, 68:4795-4812

Vengosh A (2004) Salinization and saline environments. In 'Environmental geochemistry (ed.

BS Lollar) Vol 9 Treatise on geochemistry'. (Eds HD Holland and KK Turekian), pp. 333-365,

Elsevier-Pergamon: Oxford

Verburg K, Keating BA, Probert ME, Bristow KL and Huth NI (1998) Nitrate leaching under

sugarcane: interactions between crop yield, soil type and management strategies. Agronomy-

growing greener future. Proceedings 9th Australian Society Agronomy Conference, Wagga

Wagga, (Eds DL Nichalk and JE Pratley), pp. 717-720

Wade AJ, Neal C, Whitehead PG and Flynn NJ (2005) Modelling nitrogen fluxes from the land

to the coastal zone in European systems: a perspective from the INCA project. Journal of

Hydrology, 304(1-4):413-429

Page 294: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

References

270

Ward RC (1967) 'Principles of hydrology.' McGraw Hill Publishing Company Limited: London,

402 pp.

Weier K (1999) The quality of groundwater beneath Australian sugarcane fields. Australian

sugarcane, 3(2):26-27

Whitehead PG, Wilson EJ and Butterfield D (1998) A semi-distributed Integrated Nitrogen

model for multiple source assessment in Catchments (INCA): Part I - model structure and

process equations. The Science of the Total Environment, 210-211:547-558

Wilson PR and Baker DE (1990) 'Soils and agricultural land suitability of the wet tropical coast

north Qld: Ingham area', Queensland Department of Natural Resources

Winter TC (1999) Relation of streams, lakes, and wetlands to groundwater flow systems.

Hydrogeology Journal, 7(1):28-45

Winter TC, Harvey JW, Franke OL and Alley WM (1998) 'Ground water and surface water - a

single resource', Circular 1139, U.S. Geological Survey, Denver, Colorado

Woessner WW (2000) Stream and fluvial plain groundwater interactions: rescaling

hydrogeologic thought. Ground Water, 38(3):423-429

Wolf J, Rotter R and Oenema O (2005) Nutrient emission models in environmental policy

evaluation at different scales - experience from the Netherlands. Agriculture, Ecosystems &

Environment, 105(1-2):291-306

Wood A, Schroeder B and Stewart B (2003) 'Soil specific management guidelines for sugarcane

production.' CRC for Sustainable Sugar Production: Townsville, 92 pp.

Young EO and Briggs RD (2005) Shallow ground water nitrate-N and ammonium-N in

cropland and riparian buffers. Agriculture, Ecosystems & Environment, 109(3/4):297-309

Young WJ, Prosser IP and Hughes AO (2001) 'Modelling nutrient loads in large-scale river

networks for the National Land and Water Resources Audit', Technical Report 12/01, CSIRO

Land and Water

Zaporozec A (1972) Graphical interpretation of water-quality data. Ground Water, 10(2):32-43

Page 295: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

271

Appendix A Laboratory analyses and field data

Raw data collected for this research is provided electronically in a spreadsheet labelled

Herbert_River.xls (CD in back pocket of the thesis). GPS location data for surface water and

groundwater samples are in map datum GDA94. Other units are as specified in the spreadsheet.

Page 296: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to
Page 297: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

273

Appendix B Radon sampling procedure

The procedure used for sampling radon in this study is a modification of the method described

in Cook et al. (2004). Outlined below are the key steps (adapted from the method of F. Leaney

2004, pers. comm.).

(1) Fill a 1250 mL plastic bottle with the required river/groundwater sample, ensuring minimal

air bubbles. To minimise agitation, attach a short length of clear tubing to the submersible

pump; insert the tubing into the collection bottle and allow it to gradually fill (at a low flow

rate) and overflow prior to capping. Record the time and location of sampling; for river samples

also record the width and depth of the stream.

(2) To extract radon from the river/groundwater sample, remove the lid of the bottle and syringe

out 50 ml of water to discard. Pour approximately 20 mL of scintillant9 from a pre-weighed vial

into the bottle and recap.

(3) Mix the water and scintillant by inverting the bottle every 2 seconds for a period of 4

minutes. Allow the bottle to stand for one minute to allow the scintillant to move to the top.

(4) Remove the lid from the bottle and insert a glass nozzle onto the top of the bottle.

(5) Remove the lid from the empty scintillant vial (from step 2) and hold at the outlet of the

glass nozzle. Slowly squeeze the bottle to push the scintillant out of the bottle into the vial. Stop

squeezing when the scintillant/water interface reaches the start of the capillary tubing at the start

of the nozzle, avoiding water from being transferred to the vial.

(6) Put the cap firmly back on the scintillant vial and tighten using pliers. Record the sampling

location, time and corresponding scintillant vial number for sending to the laboratory.

(7) Rinse the flask and nozzle with methylated spirits and dry in preparation for the next sample.

(8) Courier bottles to the laboratory within 2-3 days after sampling for radon counting.

9 Scintillant vial contains Packard NEN mineral oil cocktail

Page 298: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to
Page 299: Groundwater-Surface Water Interactions: Implications for ... › ... › 2 › 02Whole_Dixon-ja… · Groundwater-Surface Water Interactions: Implications for Nutrient Transport to

275

Appendix C Uncertainty analysis

In equation 7-13, the uncertainty in xbest was expressed as:

2/1

1

⎟⎠

⎞⎜⎝

⎛=Δ ∑

N

ibest wx

The derivation of this is as follows:

∑∑=Δ

i

iibest w

xwx (7-12)

i.e. }{ 2211 )(.....)(1

NNi

best xwxww

x Δ++Δ=Δ∑

i.e. ⎪⎩

⎪⎨⎧

⎭⎬⎫

Δ++

Δ=Δ∑

22

1

)1(.....)1(1

Nibest xxw

x where 2

1

ii x

=

Therefore,

}{ ∑∑=Δ i

ibest w

wx 1

or 2/1

1

⎟⎠

⎞⎜⎝

⎛=Δ ∑

N

ibest wx as in equation 7-13.