Introduction - EPA · An Introduction to the Water Quality Guidelines Chapter 1 Introduction...

29
NATIONAL WATER QUALITY MANAGEMENT STRATEGY An Introduction to the Australian and New Zealand Guidelines for Fresh and Marine Water Quality 4A Australian and New Zealand Environment and Conservation Council and Agriculture and Resource Management Council of Australia and New Zealand 2000

Transcript of Introduction - EPA · An Introduction to the Water Quality Guidelines Chapter 1 Introduction...

NATIONAL WATER QUALITY MANAGEMENT STRATEGY

An Introduction tothe Australian and New Zealand Guidelines

for Fresh and Marine Water Quality

4A

Aus t ra l ian and New Zea land Env i ronment and Conserva t ion Counc i l and Agr i cu l ture and Resource Management Counc i l o f Aus t ra l ia and New Zea land

2000

Copies of this publication may be obtained from:

Australian Water AssociationPO Box 388ARTARMON NSW 2064Tel: +61 2 9413 1288Fax: +61 2 9413 1047

OR

Australian Government Info Shops in capital cities andTownsville. For locations, contact and ordering details go to

www.dofa.gov.au/infoaccess/general/purchase_info_products.html

or phone 132 447 (toll free in Australia, 24 hour service)

OR

NZ Water & Wastes AssociationPO Box 13880Onehunga, AUCKLAND 1006Tel: +64 9 636 3636Fax: +64 9 636 1234Email: [email protected]

Material included in this document may be freelyreproduced provided that due acknowledgment is givento the Australian and New Zealand Environment andConservation Council and the Agriculture and ResourceManagement Council of Australia and New Zealand.

For further information on acknowledgment, contact:

The Secretary Australian and New Zealand Environmentand Conservation CouncilGPO Box 787CANBERRA ACT 2601Tel: +61 2 6274 1428Fax: +61 2 6274 1858

OR

The SecretaryAgriculture and Resource Management Council of Australia and New ZealandGPO Box 858CANBERRA ACT 2601Tel: +61 2 6272 5216Fax: +61 2 6272 4772

Disclaimer: The contents of this document have been compiled using a range of source materials and while reasonable care has beentaken in its compilation, the member governments of ANZECC and ARMCANZ and the organisations and individuals involved withthe compilation of this document shall not be liable for any consequences which may result from using the contents of this document.

Printed in Australia on recycled paper for the Australianand New Zealand Environment and ConservationCouncil and the Agriculture and Resource ManagementCouncil of Australia and New Zealand.

Environment Australia Cataloguing-in-Publication:

An introduction to the Australian and New Zealandguidelines for fresh and marine water quality/Australianand New Zealand Environment and Conservation Counciland Agriculture and Resource Management Council ofAustralia and New Zealand.

p. cm.

(National water quality management strategy ; no. 4a)

ISBN 0 9578245 2 1ISSN 1038 7072

1. Water quality — Australia — Measurement. 2. Waterquality — New Zealand — Measurement. 3. Water —Pollution — Environmental aspects — Australia. 4. Water— Pollution — Environmental aspects — New Zealand. 5.Water quality management — Australia. 6. Water qualitymanagement — New Zealand. I. Australian and NewZealand Environment and Conservation Council. II.Agriculture and Resource Management Council of Australiaand New Zealand. III. Series

628.161’0994–dc21

Text: Julie Freeman

Design and layout: Clockwork Communicators

Photography: Front cover (from top left) — child drinking,Ministry for the Environment, New Zealand; Logan River(Qld), Qld EPA; Maori war boat, Photosource NewZealand Limited Image Library; Hereford, Qld EPA;irrigation channels, Bruce Cooper, NSW DLWC;monitoring, Bruce Cooper; Bondi Beach (NSW), BrianRobson; Rakaia River, South Island, New Zealand, ClintMcCullough, ERISS; aquaculture, Qld EPA; Beenleigh Rumdistillery (Qld), Qld EPA; Cultural ceremony on upperKatherine River (NT), Diane Lucas [‘Because our greatgrandmothers and grandfathers been here before, their spirits arestill here. Now the spirits smell your sweat and it goes down tothe deep water and makes it alright for you to be here, withoutany harm.’ — Margaret Oenpelli and Penny Long, Barunga,NT. The washing of people by spraying water on their headis an Arnhem Land ceremony for newcomers to country tokeep away bad health for people and water.]

Back cover — Copeton Dam (NSW), Bruce Cooper

page 6 — red lotus lilies, billabong on West Alligator River(NT), Ben Bayliss, ERISS; page 10 — Australian Daphniacarinata, used for toxicity testing, showing eggs, MorenoJulli, Centre for Ecotoxicology, NSW EPA; page 13 —estuary, Qld EPA; page 16 — irrigation, Heather Hunter,Qld DNR; page 18 — cow and calf, Lesley Johnson, QldDNR; oyster beds, Agriculture, Fisheries & Forestry,Australia; page 20 — divers, Qld EPA; page 22 — samplingmacroinvertebrate communities in Micalong Creek (NSW),Philip Sloane, CRCFE ACT.

NATIONAL WATER QUALITY MANAGEMENT STRATEGY

An Introduction tothe Australian and New Zealand Guidelines

for Fresh and Marine Water Quality

Aus t ra l ian and New Zea land Env i ronment and Conserva t ion Counc i l and Agr i cu l ture and Resource Management Counc i l o f Aus t ra l ia and New Zea land

October 2000

AUSTRALIAAustralian Capital TerritoryBob NeilEnvironment ACTPO Box 144, Lyneham, ACT 2602Tel: +61 2 6207 2581 Fax: +61 2 6207 6084Email: [email protected]

New South WalesPollution LineNSW Environment Protection AuthorityPO Box A290, Sydney South, NSW 1232Tel: 131 555 Fax: +61 2 9995 5911Email: [email protected]

Northern TerritoryDirector Resource ManagementNatural Resources DivisionDepartment of Lands Planning & EnvironmentPO Box 30, Palmerston, NT 0831Tel: +61 8 8999 4455 Fax: +61 8 8999 4403Email: [email protected]

QueenslandAll enquiries for Chapter 4 (Volume 1) and Volume 3 should be addressed to:Heather HunterDepartment of Natural ResourcesBlock B, 80 Meiers RoadIndooroopilly, QLD 4068 Tel: +61 7 3896 9637 Fax: +61 7 3896 9591Email: [email protected]

All other Queensland enquiries to:Andrew MossEnvironmental Protection AgencyPO Box 155Brisbane Albert Street, QLD 4002Tel: +61 7 3896 9245 Fax: +61 7 3896 9232Email: [email protected]

South AustraliaJohn CugleySouth Australian Environment Protection AgencyGPO Box 2607, Adelaide, SA 5001Tel: +61 8 8204 2055 Fax: +61 8 8204 2107Email: [email protected]

TasmaniaDirector of Environmental ManagementEnvironmental Planning and Scientific ServicesScientific and Technical Branch, Water SectionDepartment of Primary Industries, Water and EnvironmentGPO Box 44A, Hobart, TAS 7001Tel: +61 3 6233 6518 Fax: +61 3 6233 3800Email: [email protected]

VictoriaLisa DixonManager Freshwater SciencesEnvironment Protection AuthorityGPO Box 4395QQ, Melbourne, VIC 3001Tel: +61 3 9616 2361 Fax: +61 3 9614 3575Email: [email protected]

Western AustraliaVictor TalbotDepartment of Environmental ProtectionPO Box K8222, Perth, WA 6842Tel: +61 8 9222 8655 Fax: +61 8 9322 1598Email: [email protected]

NEW ZEALANDNigel BradlyLand and Water Group, Ministry for the EnvironmentPO Box 10362Wellington, New ZealandTel: +64 4 917 7489 Fax: +64 4 917 7523Mobile: NZ 025 379 391Email: [email protected]

For information and advice about the Water Quality Guidelines and to adviseof possible errors, omissions and changes required for future revisions, pleasecontact the designated agency for your state or territory in Australia or for NewZealand. The agency contacts are listed below:

The Water Quality Guidelines at a glance 2

1 Introduction 4

Philosophical basis of the Guidelines 4

Water quality management framework 6

Tailoring guidelines for local conditions 8

Benefits of site-specific guidelines 9

Case study: Cost savings to industry through use of a decision tree 9

2 Guidelines for aquatic ecosystems 10

The indicators 10

Integrated assessment 11

Site-specific guidelines 11

Applying the guidelines for aquatic ecosystems 11

Case study: Applying the guidelines to a lowland river 14

3 Guidelines for primary industries 16

Irrigation and general water use 16

Livestock drinking water 17

Aquaculture and human consumers of aquatic foods 18

4 Guidelines for human health values 20

Recreational water quality and aesthetics 20

Drinking water 21

5 Monitoring and assessment 22

Key tables in the Water Quality Guidelines 23

References 24

Contents

2

The Water Quality Guidelines…

An Introduction tothe Water Quality

Guidelines

Chapter 1Introduction

Introduction The Water Quality GuidelinesVolume 1

Briefly describes the scope of the Water Quality Guidelines in terms of:• why and how the Guidelines were revised, and• the philosophical basis governing the Guidelines.

Provides the water quality management framework for applying the Guidelines, including:• an outline of key steps in this framework including environmental values, water quality guide-

lines and water quality objectives, and where stakeholders are involved in decision making; and• an overview of some important issues that underpin the way Guidelines should be applied.

Recommends water and sediment quality guidelines that will sustain the ecological health ofaquatic ecosystems. Includes:• guidelines for different ecosystem types and for three different levels of protection,• default generic guidelines as well as detailed advice on deriving site-specific guidelines, and• holistic assessment of water and sediment quality, including the use of biological indicators.

Provides water quality guidelines recommended to sustain primary industries and to protecthuman consumers of food products.Guidelines are applicable to:• irrigation and general water use,• livestock drinking water, and• aquaculture and human consumers of aquatic foods.

Summarises the water quality guidelines recommended to protect designated waters forrecreational activities, such as swimming and boating, and to preserve the aesthetic appeal of these waters.

Introduces the concept of safe and aesthetically pleasing drinking water, at the point of use, andrefers the reader to the relevant Guideline documents.

Provides advice on collecting and analysing data for physical, chemical and biological indicatorsand, for a range of scenarios, makes recommendations on the number and mix of indicatortypes that should be considered when monitoring aquatic ecosystems.

Chapter 2A framework for applying the Guidelines

Chapter 3Aquatic ecosystems

Chapter 4Primary industries

Chapter 5Recreational water quality and aesthetics

Chapter 6Drinking water

Chapter 7Monitoring and assessment

KEY

Should be read by all users

Environmental values supported by the Guidelines

Support chapters/ volumes/ documents

Summarises the mainfeatures of the Australianand New ZealandGuidelines for Fresh andMarine Water Quality(the Water QualityGuidelines) to helpreaders understand anduse the documents.

3

at a glance

Volume 1 is the main part ofthe Water Quality Guidelines.It provides a framework forwater resource management,and states specific waterquality guidelines for eachenvironmental value and thecontext within which theyshould be applied.

Part 2 is a CD-ROM, whichcontains the Water QualityGuidelines support volumes,databases and software toenable the user to under-stand the rationale and cal-culate site-specific triggervalues, if needed. In particu-lar, support Volumes 2 and 3from the CD-ROM providetechnical information foraquatic ecosystems, and pri-mary industries (irrigationand general water uses, stockdrinking water, aquacultureand human consumers ofaquatic foods) respectively.

The majority of users willnot need all tables and cantailor the Water QualityGuidelines for their own useby printing relevant tables(e.g. from the key tableslisted at the back of thisIntroduction), and extra fig-ures from the electronic PDFdocuments on CD-ROM, ordownloading them from theEnvironment Australia Web-site. A divider is supplied toenable users to store thispersonalised kit in the frontof the Water Quality Guide-lines folder.

Chapter 8 (Volume 2)Aquatic ecosystems —

rationale andbackground information

Chapter 9 (Volume 3)Primary industries —

rationale andbackground information

Support volumes to the External supportWater Quality Guidelines documents

ARMCANZ & ANZECCImplementation

Guidelines (1998)

ANZECC & ARMCANZMonitoring and Reporting

Guidelines (2000)

New Zealand: 1999Ministry for the

Environment GuidelinesAustralia: Until revised,apply 1992 ANZECC

Guidelines

New Zealand: 1995a,bMinistry of Health

Guidelines and StandardsAustralia: 1996 NHMRC &ARMCANZ Guidelines

A reference volume containingsupport information for chapter 3 of the Water QualityGuidelines, Aquatic ecosystems

A reference volume containingsupport information for chapter 4 of the Water QualityGuidelines, Primary industries

Stakeholder involvement

Mixing zones (Appendix 1,Volume 2)

The main objective of the Australian and New ZealandGuidelines for Fresh and Marine Water Quality (theWater Quality Guidelines) is:

to provide an authoritative guide for setting waterquality objectives required to sustain current, orlikely future, environmental values [uses] for naturaland semi-natural water resources in Australia andNew Zealand.

The Water Quality Guidelines have been prepared aspart of Australia’s National Water Quality Manage-ment Strategy (NWQMS) and relate to NewZealand’s National Agenda for Sustainable WaterManagement. They provide government and the gen-eral community (particularly catchment/water man-agers, regulators, industry, consultants andcommunity groups) with a sound set of tools forassessing and managing ambient water quality innatural and semi-natural water resources. They arenot meant to be applied directly to recycled waterquality, contaminant levels in discharges from indus-try, mixing zones, or stormwater quality, unlessstormwater systems are regarded as having conserva-tion value.

The NWQMS provides a framework for water qualitymanagement that is based on policies and principlesthat apply nationwide. In particular, the strategy isbased on the philosophy of ecologically sustainabledevelopment (ESD). This can be defined as ‘[develop-ment] using, conserving and enhancing the commu-nity’s resources so that ecological processes, on whichlife depends, are maintained, and the total quality oflife, now and in the future, can be increased’.

The guidelines are not mandatory, nor should they beregarded as such. The vast range of environments,ecosystem types and food production systems in Aust-ralia and New Zealand require a critically discerningapproach to setting water quality objectives. TheNWQMS aims to achieve sustainable use of waterresources by protecting and improving their quality

while maintaining economic and social develop-ment. A three-tiered approach — national, State orTerritory, and regional or catchment — is required.

Ultimately, it is the responsibility of local stake-holders and State and Territory or regional govern-ments to agree on the level of protection to beapplied to water bodies. State or Territory and/orlocal jurisdictions are encouraged to use thesenational water quality guidelines to formulate theirown regional guidelines or specific water qualityobjectives. Each State or Territory uses its own waterplanning and environmental policy tools to establisha framework that is compatible and consistent withthe agreed national guidelines.

The Water Quality Guidelines provide recommenda-tions that water managers can use to guide practiceand formulate policy, taking into account local con-ditions and associated costs and benefits. The resultshould be more efficient and cost-effective environ-mental management.

A ‘roadmap’ on pages 2–3 of this Introduction de-scribes the layout of the Water Quality Guidelines andexternal support documents. It will enable users toidentify the chapters or other materials relevant tothem. Key tables are listed on page 23 of this Intro-duction, while a glossary of key terms is contained inAppendix 1, Volume 1 of the Water Quality Guidelines.Volume 1 also contains an index.

Philosophical basis of theGuidelinesIncreased scientific understanding of the complexity ofecosystems and food production systems has led tonew ways of managing water quality. Traditional scien-tific and management approaches may not deal well

4

1 Introduction

with contemporary water quality issues. In their place,holistic, best-practice approaches are being taken toensure that water resources are managed sustainably.

We now recognise that all aspects of the environmentare interdependent. Environmental values (or uses),in particular, are interdependent, and cannot be con-sidered in isolation. Nor can influences on the envi-ronment be considered in isolation. Clearing of riverbank vegetation and effects of nutrients from urbanstormwater can give rise to algal blooms. Construc-tion of dams can lead to reduction in water flow andtemperature downstream. Spills of toxic chemicals orinappropriate use of pesticides can kill fish. Types ofecosystems or food production systems, interactions,cumulative effects and modifying factors must all beexamined in studying water quality.

Improvement in water quality resulting from improvedmanagement of all aspects of water use benefits allusers. Long-term sustainability in agriculture requiresthe adoption of management practices that maintainproductivity and minimise the off-farm movement orleaching of potential aquatic contaminants.

The Water Quality Guidelines recognise the inter-dependence of all aspects of the aquatic environmentand uses of water. Previous water quality guidelinesand water quality literature dealt with only two cate-gories of waters (freshwater and marine). Theyapplied single values to physical, chemical and bio-logical indicators or factors that can have adverseeffects on water quality, aquatic ecosystems or agri-cultural production systems, without taking intoaccount other factors that might reduce or exacerbatetheir effects.

The new Water Quality Guidelines consider the pro-tection of up to six ecosystem types and a range offactors that can influence the effects of specific con-taminants. Generally, the Water Quality Guidelinesalso apply to the quality of groundwater since theenvironmental values that they protect relate toabove-ground uses (e.g. irrigation, drinking water,animal or fish production and maintenance ofaquatic ecosystems). However, little is known ofunderground aquatic ecosystems and the fate ofmany organic chemicals in groundwater systems dif-fers from above-ground systems. Care is thereforeneeded when applying the Water Quality Guidelinesto groundwater.

Sustainable use

The fundamental objective of both the NationalWater Quality Management Strategy in Australia andthe Resource Management Act in New Zealand is thesustainable use and management of water resourcesin an environmental, economic and social context.Integrated catchment management (ICM) is essentialto achieving this objective. ICM encompasses allaspects of environmental management within acatchment, including water quality, and recognisesthe interdependence of environmental values.Within the ICM framework, all stakeholders —landowners and the community, in partnership withrelevant government agencies — identify environ-mental values to be protected and formulate specificwater quality objectives.

Cooperative best management

Environmental regulation and management in Aust-ralia and New Zealand are currently undergoing majorchange. Both countries have moved towards best prac-tice and cooperative best management. This requires ashift from control to prevention, from a focus on pre-scriptive regulation to a focus on outcomes, and anemphasis on cooperation rather than direction. TheWater Quality Guidelines encourage industry, govern-ments and communities to work cooperatively tomaintain or improve the quality of water bodies.

Cooperative best management involves a range oftools, for example:• memoranda of understanding, • impact assessment,• catchment management plans, and• monitoring.

Water or environmental quality

Before investing in local water quality managementstrategies, managers need to be sure that water qual-ity is the key issue in the water body under consider-ation. Water (and sediment) quality, whileimportant, is only one aspect of management. Inmany parts of Australia and New Zealand water qual-ity is reasonably good, but management goals formaintaining aquatic ecosystems are not met becauseof loss or degradation of habitat, particularly riparianvegetation. In such cases, it may be best for managers

5

to allocate resources to improve flow, riparian vege-tation or habitat in order to achieve their goals.

Management focus on issues

The Water Quality Guidelines aim to protect environ-mental values through management goals that focuson issues (concerns or potential problems). Forexample, they focus on toxicity of a chemical, soilsalinity or animal health rather than the concentra-tion of specific chemicals and salts. The water qualityconcern or problem (e.g. toxicity, sodicity, algalblooms, decrease in dissolved oxygen, loss of bio-diversity) should be identified, so that the environ-mental processes that contribute to it can beexamined. Appropriate water quality indicators canthen be determined and relevant guidelines selected.

Continual improvement

Continual improvement should be a fundamentalprinciple guiding water quality management. Inbadly polluted waters, managers may need to set sev-eral intermediate levels of water quality to beachieved in well-defined stages, until the requiredwater quality objective is finally met.

In waters whose quality is higher than the level speci-fied in set water quality objectives, attention should begiven to preventing contamination from all sources,particularly for highly modified water resources.Wherever possible, managers are encouraged to aim toimprove the quality of natural and semi-natural waterresources rather than allow it to degrade.

Integrated assessment

Water quality, ecosystem health and the surroundingenvironment are all intimately connected. For pro-tection of aquatic ecosystems, the Water QualityGuidelines have been broadened to include biologicalassessment of such ecosystems, which should accom-pany investigation of physical and chemical indica-tors in assessing impacts on ecosystem health. Theyinclude sediment guidelines, and provide advice onhow to determine suitable environmental flows inrivers and streams. Similarly, the Water Quality Guide-lines for irrigation consider potential impacts on soilsand the wider environment as well as on the cropsand pastures being produced.

For truly integrated water quality assessment, it isimportant that broadscale issues be consideredacross whole catchments where appropriate. Forexample, while water quality objectives might be metin riverine ecosystems, the cumulative effects of dis-charges and contaminant build-up downstream (e.g.in wetlands or estuaries) should also be consideredwhen water quality criteria are set.

Water quality managementframeworkLong-term management of any water resourcerequires:• clear definition of environmental values, or uses;• a good understanding of links between human

activity (including indigenous uses) and environ-mental quality;

• setting of unambiguous management goals;• identification of appropriate water quality objec-

tives, or targets; and• effective management frameworks, including

cooperative, regulatory, feedback and auditingmechanisms.

Water resource management is most effective whennational, state and regional powers and responsibili-ties are consistent and, wherever possible, integrated.Australia and New Zealand both have a regional orlocal government framework in place. In Australia,responsibility for the management of naturalresources mostly rests with the States and Territories.In New Zealand, primary responsibility for watermanagement rests with regional councils.

6

The Water Quality Guidelines provide a framework thatwater managers can use to implement a broadnational management strategy at a local level (figure1). Terms used in this framework are explained below.

‘Environmental values’ is the term applied to particu-lar values or uses of the environment. The WaterQuality Guidelines recognise the following environ-mental values:• aquatic ecosystems,• primary industries (irrigation and general water

uses, stock drinking water, aquaculture andhuman consumers of aquatic foods),

• recreation and aesthetics,• drinking water, • industrial water (no water quality guidelines are

provided for this environmental value), and • cultural and spiritual values (no water quality guide-

lines are provided for this environmental value).

Cultural and spiritual values can be taken into accountthrough the process of establishing the specific waterquality objectives for a particular water resource.

The roadmap on pages 2–3 indicates which chaptersof the Water Quality Guidelines deal with particularenvironmental values.

These environmental values may be important for ahealthy ecosystem or for public benefit, welfare,safety or health. They require protection from theeffects of pollution and inappropriate land manage-ment practices. Identification of community needsand wants is an essential step in defining environ-mental values for a particular water resource.

Associated with each environmental value are ‘guide-lines’ or ‘trigger values’ for substances that mightpotentially impair water quality (e.g. pesticides, met-als or nutrients). If these values are exceeded, theymay be used to trigger an investigation or initiate amanagement response. Where two or more agreedenvironmental values apply to a water body, the moreconservative, or stringent, of the associated guidelinesshould be selected as the water quality objectives.These are the specific or detailed targets that managerswill aim to meet in order to protect the agreed valueof the water body. In the absence of a clear and agreedset of environmental values for a particular waterresource, managers should take a conservativeapproach and assume that all appropriate environ-mental values apply by default. For example, drink-ing water would not apply as a default environmental

value for nearshore marine waters, but ecosystemprotection and recreation would apply.

Once the environmental values to be protected havebeen decided, the level of protection of environmentalor water quality necessary to maintain each valueshould be determined.

Management goals that describe precisely and in detailwhat is to be protected can then be formulated. Amanagement goal could be to eliminate or reducethe occurrence of algal blooms, to improve livestock

7

DefinePRIMARY MANAGEMENT AIMS

(including environmental values, management goals andlevel of protection)

Determine appropriateWATER QUALITY GUIDELINES

(tailored to local environmental conditions)

DefineWATER QUALITY OBJECTIVES(specific water quality to be achieved)

taking account of social, cultural, political and economic concerns where necessary

EstablishMONITORING AND ASSESSMENT PROGRAM

(focused on water quality objectives)after defining acceptable performance or

decision criteria

Initiate appropriate MANAGEMENT RESPONSE

(based on attaining or maintaining water quality objectives)

Figure 1 The framework for applying the guidelines

health and productivity, to minimise the occurrenceof fish kills, or to increase biodiversity and ecosystemhealth. Management goals should be definedaccording to community needs and desires and afterconsultation with stakeholders. They should beachievable and measurable and should be realisedthrough clear management plans.

A water quality guideline is a recommended numeri-cal concentration level (e.g. of a contaminant) or adescriptive statement (e.g. visual appearance of awater body) that will support and maintain the des-ignated use of a particular water. Water quality guide-lines are provided for chemical and physicalparameters of water and sediment, as well as biolog-ical indicators. They form the basis for determiningwater quality objectives.

A water quality objective goes a step further than a waterquality guideline. It is a numerical concentration levelor descriptive statement used by managers to measureand report on performance. Water quality objectivesare targets agreed between stakeholders, or set by localauthorities. These then become the indicators, or meas-ures, of success in meeting agreed goals.

Although it is based on scientific water quality guide-lines, a water quality objective may be modified byother inputs such as social, cultural, economic orpolitical constraints. The process of modifying guide-lines to establish water quality objectives would nor-mally involve considering costs and benefits. Thecommunity might decide, after such consideration,to allow a longer period to achieve the desired waterquality or even accept a lower water quality.

Tailoring guidelines for localconditionsIt is not possible to develop a universal set of specificguidelines that apply equally to the very wide rangeof ecosystem types or production systems, in varyingdegrees of health, in Australia and New Zealand.Environmental factors can reduce or increase theeffects of physical and chemical parameters at a siteand these factors can vary considerably across thetwo countries. A framework is provided that allowsthe user to move beyond single-number, necessarily

conservative values, to guidelines that can be refinedaccording to local environmental conditions — thatis, to developing site-specific guidelines. This is a keymessage of the Water Quality Guidelines.

Numerical and descriptive guidelines provided in theWater Quality Guidelines are intended to help man-agers establish water quality objectives that willmaintain ecosystems and meet the needs of peoplewho use a water resource. The Water Quality Guide-lines provide risk-based decision frameworks wher-ever possible, simply to help the user refine guidelinetrigger values for local and/or regional use. (Theseframeworks are most explicitly developed for themanagement of aquatic ecosystems.)

The frameworks hinge on the use of guideline ‘triggervalues’ (equivalent to the 1992 guideline defaultvalues). Guideline trigger values are concentrationsof a chemical or nutrient that, if exceeded, have thepotential to cause a problem and so trigger a man-agement response. (This could include further inves-tigation or management action.)

If the level of a particular chemical or nutrientexceeds the trigger value specified in the WaterQuality Guidelines, managers may obtain additionalinformation to determine whether the default triggervalue is appropriate. For toxicants in water, the WaterQuality Guidelines include an extensive database andsoftware package, which managers can use to calcu-late trigger values relevant to local conditions. Forexample, a given trigger value may be too stringent ina water body where high turbidity reduces theamount of the chemical available to biota.

Should managers decide that the cost of developingtrigger values tailored to local conditions is too high,they have the option of applying the default triggervalues, which are necessarily conservative by nature.Should stakeholders all agree that the economic,social, cultural or political concerns outweigh theenvironmental benefit achieved by meeting the rec-ommended guidelines, they may decide to accept alower level of protection or consider alternative man-agement, including land use practices. Conversely,should stakeholders agree that the environmentalbenefits are of overriding importance, then therecommended guidelines should prevail and man-agement strategies should be implemented to ensurethat they are achieved.

8

Benefits of site-specificguidelinesIt is not mandatory to use decision frameworks, butthey can reduce the amount of conservatism incor-porated in the guideline trigger values. This can pro-duce values more appropriate to a particular water

resource. Although tailoring guidelines to local con-ditions requires more work in some cases, it resultsin much more realistic management goals. It there-fore has the potential to reduce costs for industry.The case study below illustrates the benefits of apply-ing a decision tree.

9

The Lower Molonglo Water Quality Control Centre(LMWQCC) treats all of Canberra’s sewage, which isdischarged to the Molonglo and Murrumbidgee Riversystems. In 1995, the LMWQCC was asked by ACTElectricity and Water (ACTEW) to establish that theconcentrations of copper being discharged were notdetrimental to the receiving waters. Most samplestaken from just downstream of the discharge pointexceeded the guideline values for copper current atthe time (ANZECC 1992), and so a decision treeapproach was applied, similar to that advocated in therevised Water Quality Guidelines.

The steps taken included:1. considering whether water hardness should

modify the guideline, i.e. make it less stringent;2. considering the form and type of copper, includ-

ing the way in which it could be made lessbioavailable by complexation to dissolvedorganic matter and adsorption to sediments(through speciation modelling and complexationstudies); and

3. assessing the resulting concentration of copperagainst available (published) toxicity data, andeffluent toxicity assessment.

Waters downstream of the discharge point are natu-rally soft, and after application of hardness correc-tion factors copper concentrations in most samplesstill exceeded the guideline value.

Results of speciation modelling and complexationcapacity studies showed that free copper that wasnot bound up by dissolved organic matter and sedi-ments (i.e. the more toxic form of copper) was belowthe guideline value. Published toxicity data also indi-cated that aquatic organisms would be unlikely to beaffected by these concentrations of free copper.Also, toxicity testing of effluent carried out with acopper-sensitive freshwater alga showed no evidenceof toxicity.

Outcomes

(i) Copper discharged with sewage effluent did notcontain free copper at concentrations likely tobe a hazard to aquatic ecosystems.

(ii) At the time the study was conducted, the stepsinvolved in applying the decision tree cost theLower Molonglo treatment plant somewhere inthe order of $l0 000.

(iii) The study saved the plant tens of millions of dol-lars that it might otherwise have spent had thewater needed to be treated to remove copper.

(iv) If the same study were conducted today, the esti-mated cost to industry would be of the order of$500 for modelling and complexation studies. Inthis study, free copper was below the guidelinevalue and so toxicity testing would be unneces-sary; costs to conduct a freshwater alga testtoday are of the order of $2000.

Case study: Cost savings to industry through use of adecision tree

Aquatic ecosystems comprise the animals, plants andmicro-organisms that live in water, and the physicaland chemical environment and climatic conditionswith which they interact. Aquatic ecosystems varyconsiderably within and between tropical and tem-perate zones. The Water Quality Guidelines provideguidelines for biological and physico-chemical indi-cators of water and sediment quality that will protectthe ecological health of aquatic ecosystems, bothfreshwater and marine.

The physical components (e.g. light, temperature,mixing, flow, habitat) and chemical components(e.g. organic and inorganic carbon, oxygen, nutri-ents) of an ecosystem largely dictate what lives inthat ecosystem, and the structure of the food web.Natural physical and chemical water quality varia-tions can have important consequences for aquaticecosystems. Human activities can cause variations inthe living and non-living components of a systemthat can lead to biological changes more dramaticthan those that occur naturally.

The physical, chemical and biological aspects ofwater and sediment provide valuable indications ofthe overall health of the ecosystem.

The indicatorsBiological indicators such as algae, macrophytes,macroinvertebrates and fish are continuous monitorsof water quality, integrating the effects of past andpresent exposure to contaminants or pressures. Bio-logical assessment provides information on biologi-cal or ecological changes that may result fromchanges in water quality but may also result fromchanges in physical habitat (e.g. increased tempera-ture) or biological interactions (e.g. introduction ofexotic species or diseases). Biological assessment is avital part of assessing changes in aquatic ecosystems.It is the key tool used to assess achievement of man-agement goals and water quality objectives. To thisend, responses of biological indicators can beregarded as the end point in the decision trees usedfor physical and chemical indicators.

Physical and chemical stressors include the naturalwater quality parameters, nutrients, biodegradableorganic matter, dissolved oxygen, turbidity, suspendedparticulate matter, temperature, salinity, pH andchanges in flow regime. Physical and chemical stres-sors are major contributors to changes in aquaticecosystems, such as nuisance growth of aquatic plants,smothering of organisms living in aquatic environ-ments, and stress to or death of native freshwater fish.They may also modify the effects of toxicants. TheWater Quality Guidelines classify physical and chemicalstressors into two broad types, those having direct andthose having indirect effects on ecosystems.

‘Toxicants’ is the term given to chemical contami-nants such as metals, aromatic hydrocarbons, pesti-cides and herbicides that can potentially have toxiceffects at concentrations that might be encounteredin the environment.

Sediments are a sink for many contaminants. Theyare also habitats, food sources and refuges for manybiological communities. Sediments influence surface

10

2 Guidelines for aquaticecosystems

water quality, and can act as a source of contami-nants to benthic organisms and, hence, potentially tothe aquatic food chain. Sediments are assessed todefine contaminant concentrations above which alikely risk is posed to ecosystem health. Such assess-ments can be used by managers to decide whetherremediation or restoration is necessary.

Integrated assessmentNatural aquatic ecosystems should not be consideredas static environments. Their fauna and flora popula-tions fluctuate in accordance with seasonal and long-term climatic influences, as well as the effects ofchanges to land use and vegetation cover within theircatchments. The Water Quality Guidelines promoteassessment that integrates biological and chemicalmonitoring of surface waters and sediments to assessprogress towards achieving the goals of ecosystemprotection. The holistic approach advocated in theWater Quality Guidelines manages issues rather thansingle stressors. The aquatic environment must bemanaged to consider changes to all aspects — pol-luted sediments, reduction in stream flow, removal ofhabitat (e.g. by draining of wetlands) and changes incatchment land use.

Site-specific guidelinesThe concept of tailoring guidelines for local condi-tions was discussed on page 8 of this Introduction tothe Water Quality Guidelines. This developmenttowards site-specific guidelines is most advanced foraquatic ecosystems. Through a number ofapproaches, expanded upon below, the application ofguidelines to all Australian and New Zealand aquaticenvironments is improved. Overall, it should result inmore relevant guidelines and thereby provide betterand more cost-effective environmental protection.

In the new Water Quality Guidelines, a greater empha-sis is placed on the use of reference sites. Measuresand values for biological and many chemical indica-tors (particularly physical and chemical stressors)from suitable local reference waters are the bench-marks for assessing and maintaining biologicaldiversity at a site.

The most relevant guidelines are those tailored fordifferent ecosystem types. Default guideline valuesfor up to six ecosystem types (upland rivers andstreams, lowland rivers, freshwater lakes andreservoirs, wetlands, estuaries, and coastal andmarine) are provided.

The Water Quality Guidelines adopt an innovative risk-based approach (not full quantitative risk assess-ment). In this approach, the old single-numberguidelines (see ANZECC 1992) are regarded as guide-line trigger values that can be modified into regional,local or site-specific guidelines. Decision frameworkshelp users tailor the default guidelines to localenvironmental conditions. This can be done bytaking into account factors such as variability of theparticular ecosystem or environment, soil type, rain-fall and the level of exposure to contaminants. Theavailability of data, expertise, resources and time willdetermine which steps in the framework are used.

Ecosystems may have been modified to variousdegrees and it is unrealistic to impose a uniformmanagement goal or regulatory framework acrossdifferent locations. In reality, there is a ‘spectrum’ ofecosystem conditions. However, to assist users,guidelines are provided for three ecosystem con-ditions, with a different level of protectionrecommended for each:• high conservation/ecological value systems;• slightly to moderately disturbed systems (where

the guidelines will mostly be applied); and• highly disturbed systems.

Applying the guidelines foraquatic ecosystemsCommon to the application of all indicator types(biological, physical and chemical, toxicant and sedi-ment), the primary management aims, includingmanagement goals and level of protection, are firstdetermined (figure 1, p. 7). Preliminary steps indetermining appropriate guideline trigger levels arealso common to all indicator types, including deter-mining a balance of indicator types and selectingrelevant and specific indicators.

From this point, however, it is convenient to considerguidelines application separately for biologicalindicators and non-biological indicators (physical and

11

12

Step 1: Derive trigger values using methods in preferred order shown

A: Local or site-specific information available

yes

yes

B: No local data available (default)

1. Use local biological effects data (e.g. ecotoxicity, including multiple species toxicity tests, mesocosms)

2. Use local reference data (mainly physical and chemical stressors; for toxicants and sediments, applies only for the casewhere background data exceed default values from the box immediately below)

Use regional reference data (physical and chemical stressors)

Use generic effects-based guidelines (toxicants and sediments)

Step 2 (optional): Using decision-tree framework to assess test site and possibly refine trigger values

Test against trigger values(Compare test site data with guideline ‘trigger’ values)

If low risk If potential risk

No special actionrequiredc

Consider further site-specific investigations (Consider effects of ecosystem-specific modifying

factors)

If low risk If high risk

Initiate remedialactions

WATER QUALITY GUIDELINES

Figure 2 Procedures for deriving and refining water quality guidelines, and assessing test sites, for physicaland chemical stressors and toxicants in water and sediment. Step 1B (dark shading) indicates the most likelypoint of entry for users requiring guideline trigger values.

yesa

yes, if possible guidelines exceedance is to be assessed or if thetrigger value can be refined, using site-specific informationa, b

c possible refinement of trigger value

a further/site-specific investigations are optionalb local biological effects data and data from localreference site(s) that closely match test site generally not required in Step 2

no

no

No special actionrequiredc

chemical stressors and toxi-cants in water and sediment).For non-biological indicators,decision trees provide optionsto refine guidelines, bringingvalues closer to those that mayelicit a biological response at a particular site. Conversely,for biological indicators, themeasured responses are themanagement targets or endpoints in the decision treesthat apply for otherindicators. A different frame-work, therefore, is providedfor the application of bio-logical indicators.

Biological indicators

Comparison of biological indicators at the site(s) ofinterest with the same indicators from relatively nat-ural or unimpacted sites is the basis for detecting andassessing important changes in ecological health.The steps involved in biological assessment follow ageneral framework. After management goals havebeen determined, specific indicators are selectedthrough appropriate assessment objectives. There arethree such objectives: • broad-scale assessment — quickly determining

the extent of a problem;• early detection — pre-empting/preventing irre-

versible or important damage, habitat loss etc.; and• biodiversity or ecosystem-level response — assess-

ing the ecological importance of impact.

The Water Quality Guidelines provide advice on whichindicators to select for specific water quality issues(e.g. nutrients, metals) and suitable methods ofassessment (protocols). They outline procedures forformulating management decision criteria, and inter-preting results to assess whether water quality objec-tives are being achieved.

Physical and chemical stressors andtoxicants in water and sediment

Step 1 of figure 2 describes procedures for derivingwater quality guidelines for physical and chemicalstressors and toxicants in water and sediment. The

preferred approach to deriving guideline trigger val-ues is to use local information: ecological effects data,if available, or local reference data (mainly physicaland chemical stressors). However, in many casesdefault values will be used for decision makingbecause local information is not yet available or thereare not likely to be benefits in further refining the val-ues beyond the default values provided in the WaterQuality Guidelines.

Default guidelines for physical and chemical stres-sors are based on regional reference data for fiveclimatic/geographical regions across Australia andNew Zealand: south-east Australia, south-west Aust-ralia, tropical Australia, south central Australia —low rainfall area, and New Zealand. Default guide-lines for toxicants in water and sediment are derivedfrom biological effects (i.e. toxicological) databasesfrom Australia/New Zealand and overseas; the datahave been divided into statistical bands so that moreor less stringent guideline trigger values may beselected, depending upon stakeholders’ decision onthe level of protection to be afforded to the particu-lar ecosystem. An example of this is for most toxi-cants in slightly to moderately disturbed ecosystems,where trigger values are recommended that protect95% of species with 50% confidence.

After deriving guideline trigger values, users may pro-ceed to step 2 of figure 2, which provides guidanceon the use of a decision tree. If data from a test siteexceed the trigger value, the decision trees are used to

13

determine if the test values are inappropriately(unnecessarily) ‘triggering’ potential risk and hencemanagement response. For this, ecosystem-specificmodifying factors are introduced to assess test data.The decision trees also enable the guideline triggervalues to be adjusted and refined. The decision treeswill be most relevant to users who have selected thedefault guideline values. Thus, using the Water Qual-ity Guidelines, managers may assess guidelines’exceedance or derive guidelines for specific sites, tak-ing into account factors such as:• background concentrations;• analytical limits;• locally important species;• physical and chemical water quality modifiers,

such as suspended/dissolved organic matter, pH,temperature, salinity, hardness and speciation;

• interactions; and• data quantity, quality and extent in time.

While it is not mandatory to use decision frame-works, they are recommended so that the resultingguidelines are relevant to the site. Users wanting toundertake site-specific assessments to refine triggervalues to be used for regulatory purposes will need toobtain prior agreement from the relevant State orTerritory or regional authorities. The guideline triggervalues are based on ‘bioavailable’ concentrations,and hence measurements of total concentrationsmay overestimate the amount of chemical thatcauses biological effects at that site. As a con-sequence, the use of the decision frameworks will, inmost cases, increase (i.e. make less stringent) guide-line concentrations at the site. Alternatively, somesteps in the frameworks guide users towards bettermeasurements of the ‘bioavailable’ concentrations atthe site.

14

A hypothetical large river has a forested upper catch-ment, while its lowlands and flood plains have beenextensively cleared for agriculture. A medium-sizedcity is located just upstream of the tidal limit anddownstream of a series of weirs. Part of the river hasreceived waste from a smelter sited near the river.This discharge has contaminated the sediments withcadmium.

The following steps show how the guidelines may beused to define specific water quality objectives formanaging environmental problems, especially heavymetal impacts.

Define the water body — A lowland river. The uppersection is only slightly to moderately disturbed, but thesection below the weirs is highly disturbed.

Establish the environmental values — Aquaticecosystem protection, recreational water quality andaesthetics, and primary industries are established inconsultation with stakeholders.

What information is available — Physical and chem-ical monitoring data are available for the disturbed sec-tion but biological data are scant.

Determine the level of protection — The target formanagement is to return ecosystem features in lower

sections of the river from a highly disturbed conditionto slightly to moderately disturbed.

Identify key environmental concerns — An agreedconcern is possible toxicity to aquatic biota arising fromcadmium in water and particularly in sediments.

Define management goals — To maintain a quality ofwater suitable for recreation and aesthetics; enhanceaquatic ecosystem health (improve conditions fornative fish, retain sediment quality that will maintainnatural biota, improve understanding of the ecosys-tem); and protect downstream environmental valuessuch as aquatic ecosystems, aquaculture or agriculturalwater supplies for irrigation and stock.

Determine appropriate guideline trigger levels forselected indicatorsFirst, determine a balance of indicator types and otherrequirements based on level of protection and localinformation constraints. The situation involves highlydisturbed ecosystems where water and sedimentquality are of concern; little is available in the way ofpre-disturbance data and the long-term managementgoal is improvement in river health. The Water Quality Guidelines recommend water and sedimentphysico-chemistry, and rapid biological assessmentand/or quantitative biodiversity measurement.

Case study: Applying the guidelines to a lowland river

15

Next, identify all indicators relevant to the environ-mental concerns and management goals, then deter-mine guideline trigger levels and specific indicators tobe applied.

Determine guidelines,define water quality objectivesand establish a monitoring and assessment programThe steps for each of the broad indicator types aredescribed below. For this case study, we assume thatthe stakeholders (different catchment agencies, utilitiesand industries) have agreed that the water qualityobjectives are the same as the water quality guidelines.

Given the broad-scale nature of the environmentalproblems in the catchment, different catchment agen-cies, utilities and industries pooled resources for themonitoring and assessment program. Additional sav-ings in resources were made in the sampling programby sharing reference site data held by agencies con-ducting similar monitoring programs in the region.

Biological indicators The primary biological assessment objectives havebeen determined as biodiversity or ecosystem-levelresponse and catchment-wide assessment. The bio-logical indicator selected is composition and structureof macroinvertebrate communities, to be measured byrapid biological assessment (RBA) and quantitative bio-logical assessment.

RBA showed many fewer macroinvertebrate familiesthan expected in the river in the vicinity of historicalpollution sources and around the city, indicatingsubstantial impacts on water and/or habitat quality.Key point sources of contamination in the river arisingfrom urban, agricultural and industrial wastewatersduring high-flow events were shown to contribute tomuch of the deterioration in river health. This wasdemonstrated using a combination of RBA and quanti-tative assessment methods, with repetitive sampling atsites upstream and downstream of suspected contam-ination sources and from suitable reference sites.

ToxicantsCadmium has been identified as a toxicant requiringinvestigation. A monitoring program is designed andimplemented, based on the decision criterion thataction is triggered if the 95th percentile of the test

distribution exceeds the guideline value. Cadmiumtotal concentrations of 1.0 µg/L are found in the water.Water in the river has a median hardness of 80 mg/LCaCO3 (moderate) and so the relevant trigger value is0.54 µg/L. This last value is 2.7 × the low hardnesstrigger value of 0.2 µg/L. Options are to:• use the trigger value as an objective and develop

management strategies to meet this, or• conduct further studies to determine if any site-

specific factors might modify the guideline value.

The decision is made to conduct a further inex-pensive study to determine the dissolved cadmiumconcentration. The dissolved cadmium concentra-tion is 0.022 µg/L, which is below the guideline valuefor moderate hardness. The risk of adverse biologi-cal effects is therefore considered to be low.

Sediment qualityCadmium has been identified as requiring investigation.The sediment cadmium at 5 mg/kg dry weight is abovethe interim sediment quality guidelines (ISQG)-lowtrigger value but below the ISQG-high value. Back-ground cadmium measured in the surrounding soilsand upstream river sediments is below the ISQG-lowvalue, indicating that original levels in the river werealso low and that elevated levels can be attributed tothe industrial discharge. Further investigation isrequired to determine whether these levels pose ahazard to aquatic life. This will involve examining fac-tors controlling availability to biota, and then acutetoxicity and chronic toxicity testing if results indicatethese are needed. If the results indicate a high risk ofbiological impact, remedial action will be required.

Next steps in managementThe steps already taken have shown that aquatic biotaare adversely affected by various pollution sources andthat cadmium concentrations in the sediments may betoxic. Further investigation is required to derive site-specific guidelines. Also needed are managementstrategies to reduce the levels of these indicators, inaccordance with the philosophy of continual improve-ment, and continued monitoring and assessment toevaluate progress. The biological assessment programwill be an important component. Where site-specificguidelines indicate a high risk of an impact occurring,remediation programs should be developed.

16

Both the quality and the quantity of water resourcesare critical issues for agriculture and aquaculture inAustralia and New Zealand. Water quality is also ofmajor importance for the protection of human con-sumers of food products. Primary industries, alongwith other urban and industrial development, haveincreased the demand for good-quality water, but atthe same time have exerted escalating pressures onthe quality of the water resources that are available.Assessment of water quality for primary industriestherefore requires balanced consideration of produc-tivity issues and the possible adverse impacts of theseenterprises on downstream water quality.

The Water Quality Guidelines can be used at the broadlevel to assist in defining water quality objectives forcatchments in which irrigation use is prevalent. Theycan also help individual irrigators assess the suit-ability of water for their uses, and possible miti-gation measures to reduce any impacts.

In Australia, in particular, both the irrigation andlivestock industries rely heavily on groundwater inaddition to surface water. Groundwater is also animportant source of stock water in parts of NewZealand. Thus the guidelines provided in the WaterQuality Guidelines for these industries are applicable(where appropriate) to both surface water andgroundwater quality.

Irrigation and general wateruseAgricultural practice in Australia and New Zealand isoften dependent on irrigation, because of climaticconstraints. In Australia, in particular, agriculture ispredominantly based in areas of limited rainfall witha heavy reliance on the use of surface water andgroundwaters for irrigation of crops and pastures.

An important goal of the Water Quality Guidelines is tomaintain the productivity of irrigated agricultural landwhile protecting associated water resources, in accor-dance with the principles of ecologically sustainabledevelopment. In terms of water quality, the focus forsustainable farming systems is on adopting manage-ment practices that maintain productivity and min-imise the off-farm movement or leaching of potentialaquatic contaminants. Key off-site issues include soilerosion, landscape salinity, fertiliser and pesticidemanagement, livestock access to streams, and safe dis-posal of effluent from intensive animal industries.

Soil, plant and water resource issues have been takeninto account in developing the water quality guide-lines for irrigation use (see table 1). The guidelinesfor irrigation water quality include trigger values for: • biological parameters; • salinity and sodicity;• inorganic constituents (specific ions, including

heavy metals and nutrients); and • pesticides and radiological contaminants.

3 Guidelines for primaryindustries

17

A new five-step approach has been developed forassessing irrigation water salinity and sodicity, withsoftware provided so users can determine the suit-ability of their water on a site-specific basis. Thisallows soil type, rainfall and other local factors to betaken into account and gives irrigators more flexibil-ity in their management.

In addition, trigger values for heavy metals and metal-loids have been developed using an internationallyrecognised approach. Two trigger values for water areincluded (for short-term and long-term use), generallyin conjunction with an associated trigger value for thesoil under irrigation. A similar approach has beenused to develop long-term and short-term trigger val-ues for nitrogen and phosphorus in irrigation water.

For the first time, the Water Quality Guidelines dealwith the corrosion and fouling potential of watersused for general on-farm water use. These character-istics are important considerations in the main-tenance of farm equipment (pumps, pipes etc.) butcan also be applied more widely where corrosionand fouling are of concern.

Livestock drinking waterGood-quality drinking water is essential for success-ful livestock production. Poor-quality drinking watermay reduce animal production or impair fertility; inextreme cases, stock may die. Water quality require-ments for livestock may differ between animalspecies, stages of growth and animal condition.

Contaminants in drinking water can produceresidues in animal products (e.g. meat, milk andeggs), adversely affecting their saleability and/orcreating human health risks. Animal industriesthemselves can impair water quality downstream(e.g. through faecal contamination). The need for anintegrated approach to land and water managementin rural catchments is clear.

The scope of the Water Quality Guidelines for livestockdrinking water includes biological, chemical andradiological constituents that can affect animalhealth. Guidelines for the commonly occurring blue-green alga, Microcystis, and its toxins have beenincluded in the Water Quality Guidelines for the first

Table 1 Key issues concerning irrigation water quality effects on soil, plants and water resources

Key issues

Soil Root zone salinitySoil structural stabilityBuild-up of contaminants in soilEffects on soil biotaRelease of contaminants from soil to crops and pastures

Plants YieldProduct qualitySalt toleranceSpecific ion toleranceFoliar injuryUptake of toxicants in produce for human consumptionContamination by pathogens

Water resources Deep drainage and leaching below root zoneMovement of salts, nutrients and contaminants to groundwaters and surface waters

Other important factors Quantity and seasonality of rainfallSoil propertiesCrop and pasture species and management options Land typeGroundwater depth and quality

18

time. Significant revisions of trigger values for sulfateand magnesium are other notable features of therevised guidelines for livestock drinking water.

Aquaculture and humanconsumers of aquatic foodsThis is the first time guidelines have been providedfor aquaculture industries in Australia and NewZealand. Recreational and commercial fisheries arebased on wild populations of fish, crustacea andshellfish species, which are supported by naturalhabitats and food webs. The guidelines for the pro-tection and maintenance of aquatic ecosystemsshould therefore be applied to protect these wild ani-mal stocks.

Aquaculture involves the production of food forhuman consumption, fry for recreational fishing andnatural fisheries, ornamental fish and plants for theaquarium trade, raw materials for energy andbiochemicals, and a number of items for the fashionindustry. The need for satisfactory water quality tomaintain viable aquaculture operations is wellaccepted. Poor water quality can result in loss of pro-duction of culture species, and can also lower thequality of the end product. Aquaculture also has thepotential to have an impact on water quality fordownstream users and this is to be addressed throughother environmental values.

Aquaculture is a burgeoning industry in both Aust-ralia and New Zealand. About sixty aquatic speciesare currently being farmed, but only limited

information is available on the water quality require-ments of many of these species and on the require-ments of larval and juvenile stages of their life cycles.Trigger values for aquaculture should therefore beconsidered as interim guidelines. In developing theseguidelines, the objective has been to:• promote the quality of water necessary to protect a

wide range of aquatic animals and plants culturedby existing and future aquaculture activities; and

• protect human consumers of harvested aquaticfood species from a variety of sources, includingaquaculture and commercial, recreational andindigenous fishing.

Guidelines for protectingaquaculture species

The Water Quality Guidelines provide a basis foraspects of aquaculture management, such as:• environmental planning and management, • environmental assessment and monitoring

requirements, • appropriate environmental zoning and legislation, • selection of appropriate species and suitable sites, • site capacity, • farm design criteria, • stocking densities and feeding regimes, and • production schedules.

The guidelines do not deal with effluent water qualityfrom aquaculture activities, however, and aqua-culturists need to manage their operations withdownstream water quality in mind.

As an initial approach, guidelines have been derivedfor the protection of freshwater and marine species.

19

Given the large number of aquaculture species inAustralia and New Zealand, and the general lack ofinformation on most of them, all finfish, molluscand crustacean species were divided into eightindicative groups. Efforts were then directed atreviewing toxicity and tolerance data for one or tworepresentative species within each of those groups,the species being chosen according to the level ofproduction and availability of scientific data. Wherediscrepancies in the data were identified, the moreconservative data were generally used.

The guidelines are provided under the following fourcategories:• physico-chemical stressors,• inorganic toxicants,• organic toxicants, and• pathogens and biological contaminants.

General guideline values for freshwater and saltwater(brackish and marine water) aquaculture uses are rec-ommended. In addition, specific guideline values havebeen provided for species groups where informationwas available on their water quality requirements.

Water quality guidelines for theprotection of human consumers ofaquatic foods

The Australia New Zealand Food Authority (ANZFA)

develops and administers uniform (statutory) stan-

dards for chemical contamination in foods (includ-

ing aquatic foods). Unlike the water quality

guidelines, the ANZFA food standards are enforce-

able through legislation. In addition to those for

chemical contaminants, guidelines are provided for:

• viral contaminants,

• bacterial contaminants,

• natural toxins,

• parasites, and

• off-flavour compounds (which cause tainting of

aquatic animal flesh).

The Water Quality Guidelines are consistent with and

do not duplicate this work.

20

Recreational water quality andaestheticsChapter 5 of the Water Quality Guidelines deals withrecreational water quality and aesthetics. Water-based recreational activities are popular with Aus-tralians and New Zealanders. Much of the coastlinein both countries is inaccessible for recreational pur-poses, resulting in high pressures on areas that areaccessible. The same situation applies to estuarineand freshwater rivers and lakes, especially those closeto urban centres. Guidelines are necessary to protectthese waters for recreational activities, such as swim-ming and boating, and to preserve their aestheticappeal. Factors considered include:

• microbiological characteristics;• nuisance organisms (macrophytes, phytoplankton

scums, filamentous algal mats, blue-green algae,sewage fungus and leeches); and

• physical and chemical characteristics (pH, temper-ature, toxic chemicals and surface films).

Recreational guidelines accommodate two categoriesof sporting activity:• sports such as swimming or surfing in which the

user comes into frequent direct contact with water,either as part of the activity or accidentally (pri-mary contact); and

• sports such as boating or fishing that generallyinvolve less-frequent body contact with the water(secondary contact).

A third recreational category concerns the passiverecreational use of water bodies, mainly as pleasantplaces to be near or to look at (no body contact).

Guidelines for users in New Zealand

In New Zealand, water managers should refer toRecreational Water Quality Guidelines (Ministry for theEnvironment 1999). This document and the draftsupporting manual can be downloaded from:http://www.mfe.govt.nz/about/publications/water_quality/beaches-guidelines.htm.

The revised New Zealand guidelines were trialledover the 1999–2000 bathing season. This trial periodwill be followed by consultations. Any recommend-ation to the Minister for the Environment regardinga National Environmental Standard will be madeafter these consultations.

Guidelines for users in Australia

Guidelines for recreational water quality and aesthet-ics are currently being prepared for Australian users.When completed, they will replace the current sectionof the Water Quality Guidelines, in accordance withNWQMS requirements and National Health and Med-ical Research Council (NHMRC) statutory procedures.It is intended that the new guidelines should be basedlargely on recommendations from the World Health

4 Guidelines for humanhealth values

21

Organization (WHO). Until these guidelines arerevised and endorsed, users should apply the guide-lines from the Australian Water Quality Guidelines forFresh and Marine Waters (ANZECC 1992). These guide-lines are reproduced in the Water Quality Guidelines.

Drinking waterChapter 6 of the Water Quality Guidelines deals withdrinking water. Drinking water should be safe to useand aesthetically pleasing. Guidance on what consti-tutes good-quality drinking water is provided forNew Zealand by Drinking-water Standards for NewZealand (New Zealand Ministry of Health 1995a)and the Guidelines for Drinking-water Quality Manage-ment (New Zealand Ministry of Health 1995b). InAustralia, similar guidance is provided by theAustralian Drinking Water Guidelines (NHMRC andARMCANZ 1996), a companion document of theNational Water Quality Management Strategy.

For more information on the Australian DrinkingWater Guidelines, readers may wish to refer to:http://www.health.gov.au:/nhmrc/publicat/synopses/eh19syn.htm

The Drinking Water Guidelines are intended to meetthe needs of consumers and apply at the point of use,for example at the tap. They are applicable to anywater intended for drinking, irrespective of its source(municipal supplies, rainwater tanks, bores, or point-of-use treatment devices) or where that wateris to be used (in homes, restaurants etc.). The WaterQuality Guidelines summarise the key issues con-tained in the Australian Drinking Water Guidelines.

The Drinking Water Guidelines devote a chapter to themicrobiological quality of drinking water, since themost common and widespread health risk associatedwith drinking water is contamination, either directlyor indirectly, by human or animal excreta and micro-organisms contained in faeces. Micro-organisms,including pathogenic organisms, may enter watersupplies at every stage of the collection anddistribution cycle. Emphasis is placed on the need foran active watershed protection program, including anemergency plan for responding to major pollutionevents such as spills or contamination. Detailed adviceis given on the problems of surface and groundwatersupplies, and approaches to their management.

The Drinking Water Guidelines also discuss the prob-lems of small water supplies, regarded as thoseserving fewer than 1000 people, and individualhousehold supplies.

Guideline values

The individual guidelines in the Drinking WaterGuidelines cover a wide range of measurable charac-teristics, compounds or constituents that can poten-tially be found in water and affect its quality. They fallinto the following categories:• micro-organisms (including bacteria, protozoa,

toxic algae and viruses);• physical characteristics;• radionuclides; and• chemicals (including inorganic chemicals, organic

compounds, organic disinfection by-products,and pesticides).

The Water Quality Guidelines both contain, and referto, information on the practicalities of collecting andanalysing data for the measurement of water quality.This information is based on the national frameworkfor monitoring water quality and reporting on theoutcomes. The national framework aims to raise thestandard of water monitoring in Australia andshould help ensure that programs are more consis-tent. This will allow data to be better comparedacross regions and over time.

For basic, general details of how to plan a monitoringprogram, the Water Quality Guidelines (through chap-ter 7) refer the reader to the Australian Guidelines forWater Quality Monitoring and Reporting (the Monitor-ing Guidelines). Both the Monitoring Guidelinesand accompanying summary document set out aframework for developing a monitoring program.The framework outlines in broad terms:• how to define monitoring program objectives, and

the principles for both designing monitoring studiesand implementing an effective sampling program;

• laboratory analyses, choosing suitable techniquesfor data analyses and data interpretation; and

• mechanisms for reporting of the results and reach-ing conclusions.

Chapter 7 of the Water Quality Guidelines comple-ments the Monitoring Guidelines in that it containsinformation that is both very specific to issues raisedin earlier chapters of the Guidelines and, for selectedtopics, of a more detailed nature for the experienced

reader. Among the specific and/or detailed topicsthat are covered in chapter 7 and that are relevantmainly to aquatic ecosystems are:• recommendations on a balance of indicator types

to apply under different situations;• extensive advice on biological assessment; and• for physico-chemical indicators of water and sedi-

ment, information on how to compare test datawith guideline trigger values.

22

5 Monitoring and assessment

Important tables from the Water Quality Guidelines, including tables of default guideline trigger values, thatusers may need to refer to are listed below.

Environmental value and table description Reference

Aquatic ecosystems

Water quality issues and recommended biological indicators for different ecosystem types Table 3.2.2

Regional values for physical and chemical stressors Tables 3.3.2–3.3.11

Values for toxicants Table 3.4.1

Values for sediments Table 3.5.1

Primary industries: irrigation and general water use

Values for coliforms, salinity and other major ions, nutrients, general toxicants, natural physical and chemical indicators, radiological contaminants Tables 4.2.2–4.2.15

Primary industries: livestock drinking water quality

Values for salinity, metals and radiological contaminants Tables 4.3.1, 4.3.2 and 4.3.3 respectively

Primary industries: aquaculture and human consumers of aquatic foods

Values for physico-chemical stressors and toxicants Tables 4.4.2 and 4.4.3respectively

Recreational water quality and aesthetics

Values for recreational waters Table 5.2.2

Values for recreational purposes: general chemicals and pesticides Tables 5.2.3 and 5.2.4 respectively

23

Key tables in the Water Quality Guidelines

ANZECC 1992. Australian water quality guidelines for fresh

and marine waters. National Water Quality Management

Strategy Paper No 4, Australian and New Zealand Envi-

ronment and Conservation Council, Canberra.

ANZECC & ARMCANZ 1994. Policies and principles: A refer-

ence document. National Water Quality Management

Strategy Paper No 2, Australian and New Zealand Envi-

ronment and Conservation Council & Agriculture and

Resource Management Council of Australia and New

Zealand, Canberra.

ANZECC & ARMCANZ 2000a. Australian and New Zealand

guidelines for fresh and marine water quality. National

Water Quality Management Strategy Paper No 4, Aus-

tralian and New Zealand Environment and Conserva-

tion Council & Agriculture and Resource Management

Council of Australia and New Zealand, Canberra.

ANZECC & ARMCANZ 2000b. Australian guidelines for

water quality monitoring and reporting. National Water

Quality Management Strategy Paper No 7, Australian

and New Zealand Environment and Conservation

Council & Agriculture and Resource Management

Council of Australia and New Zealand, Canberra.

ANZFA 1996. Food standards code (including amendments

to June 1996). Australia New Zealand Food Authority,

Australian Government Publishing Service, Canberra.

ARMCANZ & ANZECC 1998. Implementation guidelines.

National Water Quality Management Strategy Paper No 3,

Agriculture and Resource Management Council of Aust-

ralia and New Zealand & Australian and New Zealand

Environment and Conservation Council, Canberra.

New Zealand Ministry for the Environment 1999. Recre-

ational water quality guidelines. New Zealand Ministry for

the Environment, Wellington.

New Zealand Ministry of Health 1995a. Drinking-water

standards for New Zealand. New Zealand Ministry of

Health, Wellington.

New Zealand Ministry of Health 1995b. Guidelines for

drinking-water quality management. New Zealand Min-

istry of Health, Wellington.

NHMRC & ARMCANZ 1996. Australian drinking water

guidelines. National Water Quality Management Strategy

Paper No 6, National Health and Medical Research

Council & Agricultural and Resource Management

Council of Australia and New Zealand, Australian Gov-

ernment Publishing Service, Canberra.

24

References