Reuse of Urban Runoff in Australia: A Review of Recent Advances and Remaining Challenges

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

SPECIAL SUBMISSIONS

TECHNICAL REPORTS

Reuse of Urban Runoff in Australia: A Review of Recent Advances and Remaining Challenges
Tim D. Fletcher,* Ana Deletic, V. Grace Mitchell, and Belinda E. Hatt Monash University

The degradation of aquatic ecosystems due to hydrologic


and water quality impacts of urbanization, combined with
increasing water scarcity, has generated increasing interest in
I n recent years, the environmental consequences of uncontrolled
and untreated discharge of urban storm water runoff have
been well documented. Changes to hydrology (Boyd et al., 1993;
the harvesting of urban storm water. This paper reviews the
rationale for integrated storm water treatment and harvesting Carlson and Arthur, 2000; Leopold, 1968; Wong et al., 2000) and
and synthesizes recent advances and trends and knowledge to pollutant concentrations and loads (Hatt et al., 2004b) have
gaps that limit its application. Storm water harvesting is been widely observed. Aquatic ecosystems in catchments with even
shown to be a viable alternative water supply and to provide small proportions of urbanization have been shown to be degraded
a potential solution to the increases in runoff frequency and
in terms of stream functioning (Groffman et al., 2002; Harbott
peak flows that occur as a result of catchment urbanization.
In general, treatment technologies for storm water harvesting and Grace, 2005; Meyer et al., 2005) and species composition
have been adapted from existing “water-sensitive urban design” (Morgan and Cushman, 2005; Roy et al., 2005; Serena and
approaches, with limited use of traditional water supply and Pettigrove, 2005). Recognition of these consequences has led to
wastewater technologies. Risk management is often lacking, in the development of a range of storm water treatment technologies,
part due to a lack of relevant guidance. Reported performance
such as sediment basins, ponds (Farm, 2002; Lloyd et al., 2002b;
shows variable levels of potable water savings, with cases of up
to 100% substitution recorded. Costs of storm water harvesting Marsalek and Marsalek, 1997), storm water wetlands (Bourguès
systems are shown to be inversely related to their scale. The and Hart, 2007; Carleton et al., 2000; Farm, 2002; House et al.,
limited cost data show the importance of context, with the 1994), and the more recent developments in biofiltration, using
harvested water costing more or less than alternative supplies, vegetated soil filters (Davis et al., 2001; Denman et al., 2007; Hatt
depending on the cost of the alternative. Limited data exist on
et al., 2006b, 2008; Henderson et al., 2007).
environmental benefits, such as reductions in pollutant loads
and flow peaks. Implementation of storm water harvesting At the same time, pressure on water resources in urban areas is
systems is impeded by inadequate data on risk, lifecycle costs, increasing, with growing demand and limited water sources. This,
externalities, and water–energy tradeoffs. Furthermore, retrofit combined with an increasing awareness of environmental issues
of storm water harvesting into existing urban areas is proving to (Harremoes, 1997), has resulted in an increased attention to the
be a challenge, creating an urgent need for specific technologies
potential for harvesting urban runoff as an alternative water re-
for use in retrofit situations.
source, primarily for nonpotable purposes (Hatt et al., 2006a).
This paper reviews recent advances in the harvesting of urban
storm water runoff. First, we briefly explore the rationale for storm
water harvesting by reviewing recent insights into the mechanisms of
environmental degradation by urban storm water and outlining the
context of increasingly limited urban water resources. We explore the
recent advances and trends in storm water harvesting and treatment
systems, documenting the evolution from “receiving-water-focused”
treatment to the new technologies aimed specifically at providing
storm water for harvesting. We also attempt to assess the performance
of storm water runoff harvesting systems in terms of water savings and
environmental, economic, and social consequences. Last, we identify
areas that we believe remain as impediments to the adoption of effec-
tive use of storm water runoff as an alternative water supply.
Copyright © 2008 by the American Society of Agronomy, Crop Science
Society of America, and Soil Science Society of America. All rights
reserved. No part of this periodical may be reproduced or transmitted Rationale for Harvesting of Storm Water Runoff
in any form or by any means, electronic or mechanical, including pho- Urbanization represents the fastest growing land use in the
tocopying, recording, or any information storage and retrieval system,
without permission in writing from the publisher. world (United Nations, 2008); for the first time, more than 50%
of the world’s populations currently live in urban areas. As a con-
Published in J. Environ. Qual. 37:S-116–S-127 (2008).
doi:10.2134/jeq2007.0411
Dep. of Civil Engineering (Institute for Sustainable Water Resources), Bldg. 60, Monash
Received 3 Aug. 2007.
Univ., Victoria, Australia, 3800.
*Corresponding author (tim.fletcher@eng.monash.edu).
© ASA, CSSA, SSSA Abbreviations: BNR, biological nutrient removal; SMURRF, Santa Monica Urban Runoff
677 S. Segoe Rd., Madison, WI 53711 USA Reclamation Facility; UV, ultraviolet; WSUD, water sensitive urban design.

S-116
sequence, demand for water in many urban areas is approach- channels with that of lining some or all drainage channels
ing, and in some cases exceeding, the available supply (Hatt et within the catchment. They concluded that 80 to 90% of
al., 2006a). For example, in 2006, six of the seven Australian the increase in peak flows is explained by the nature of the
state capital cities faced water restrictions, affecting an esti- drainage connection, rather than simply the proportion of
mated 70% of the Australian population of 21 million people the catchment that is impervious. More recently, Hatt et al.’s
(Australian Bureau of Statistics, 2006). Similar shortages face (2004b) investigation of water quality in relation to catch-
parts of Africa, China, Singapore, the USA, Israel, and Europe ment characteristics also showed that the nature of connection
(Niemczynowicz, 1996; Vlachos and Braga, 2001). of impervious areas to receiving waters was a strong indepen-
Although there are several potential alternative water sup- dent predictor of pollutant concentrations and loads.
plies that may help to address this shortage, such as recycling One aspect of the hydrologic changes brought about by urban-
of wastewater or desalination of sea water, the harvesting and ization that has been less studied is the change in evapotranspira-
treatment of urban runoff has the advantage of higher public tion. Generally, it is expected that the amount of evapotranspira-
acceptance than wastewater recycling (Brown and Davies, tion will be less after urbanization due to the loss of vegetation
2007; Coombes and Mitchell, 2006). Mitchell et al. (2003) (Oke, 1987). However, this effect may be altered by the impact of
showed that the annual storm water discharged (from separate garden watering and open-space irrigation (Grimmond and Oke,
storm water sewer systems) from three major Australian cit- 1991), particularly in areas with lower density and large gardens.
ies—Brisbane, Melbourne, and Sydney—was similar to their Although it has long been known that urbanization and the
total water demand (50% of which is for nonpotable uses). documented changes in hydrology and water quality lead to deg-
The other key driver for harvesting of storm water is an increas- radation of aquatic ecosystems (Aalderink et al., 1990; Arthing-
ing recognition of the impacts of unmitigated urban runoff. Per- ton et al., 1983; Pratt et al., 1981), the recent insights into the
haps more importantly, there has been substantial recent research importance of drainage connection has significantly informed
into the mechanisms that contribute to this degradation. Broadly recent ecological studies of the “urban stream syndrome” (Walsh
speaking, urbanization causes significant changes to a catchment’s et al., 2005b). For example, Taylor et al. (2004) showed that ben-
hydrology (e.g., Leopold, 1968; Rose and Peters, 2001) and water thic algal biomass was directly related to the degree of connec-
quality (Hatt et al., 2004b; Soranno et al., 1996). tion of impervious areas to receiving waters. Harbott and Grace
Changes to hydrology from urbanization are well docu- (2005) showed similar relationships for extracellular enzyme re-
mented and are related to two factors: (i) the proportion of the sponse, as did Walsh (2004) for macroinvertebrates. Walsh et al.
catchment, which is made up of impervious surfaces, and (ii) the (2005a, 2004) observed that the frequency of runoff into receiv-
hydraulic efficiency of flow paths from these impervious areas ing waters is a critical indicator of the likely degradation to receiv-
to receiving waters (Booth and Jackson, 1997; Leopold, 1968; ing waters. Consequently, they suggest that it is necessary to in-
Walsh et al., 2005b; Wong et al., 2000). For example, Leopold tercept and retain any rainfall up to the amount that would have
(1968) showed that urbanization leads to decreases in the lag caused surface runoff under predeveloped conditions. Ladson et
time between rainfall and runoff and increased peak flow, with al. (2007) further explore this, demonstrating that installation of
the greatest proportional increase occurring for the smaller, more rain water tanks on each property and the installation of storm
frequent events (because for larger events, pervious areas become water infiltration and biofiltration systems would be sufficient to
saturated and begin to produce runoff anyway). Leopold suggest- achieve the required predeveloped runoff frequency.
ed that 20% imperviousness would result in a doubling of the It is also well established that the increased peak flows that
frequency of overbank flows, with 50% resulting in four times occur as a result of urbanization can lead to geomorphologic
the predeveloped frequency. Similar observations have been made changes, such as bed and bank erosion and channel enlarge-
by several studies (Codner et al., 1988; Driver and Troutman, ment (Booth, 1990; Booth and Jackson, 1997; Konrad et al.,
1989; Olivera and DeFee, 2007; Wong et al., 2000). Wong et al. 2005) and the destruction of benthic habitat due to sediment
(2000) showed that the 20% annual exceedence probability flow smothering (Lenat and Crawford, 1994; Olthof, unpublished
(i.e., the flow with an average recurrence interval of 5 yr) under data, 1994; Wang et al., 1997).
rural conditions would occur on average 2 to 3 times per year, The important implication of these recent research find-
with only 20% catchment imperviousness. ings is that retention and “loss” of storm water is necessary to
The proportion of a catchment that is impervious is not restore catchment hydrology toward predevelopment levels.
the only predictor of the degree of hydrologic disturbance Storm water reuse provides one potential tool to assist in this
that derives from urbanization. Rather, a long series of studies aim. By detaining and using urban runoff, there may be po-
have examined the extent to which hydraulic efficiency of the tential reductions in peak flow, annual runoff volume, and the
drainage network controls catchment response to urbaniza- frequency of runoff (Walsh et al., 2005a).
tion. Leopold (1968) showed that the post:predevelopment Despite this apparently strong rationale for the harvesting of
peak flow ratio depends on the proportion of impervious ar- urban runoff as a means for simultaneously tackling water supply
eas that are directly connected (via a pipe or constructed lined shortages and degradation of aquatic ecosystems, urban storm
channel, for example) to the receiving water. water remains a relatively neglected water resource (Thomas et al.,
Wong et al. (2000) continued this line of enquiry, com- 1997). One reason for this may be the relative lack of technology
paring the consequences of maintaining natural drainage and facilities already in place, relative to other components of

Fletcher et al.: Reuse of Urban Runoff in Australia S-117


Fig. 1. Main components of urban storm water harvesting systems.

the urban water management system. For example, recycling of Pitt, 2001). Therefore, storm water treatment commonly uses
wastewater is facilitated by the fact that treatment infrastructure natural purification processes that also allow for temporary stor-
for wastewater treatment is well established, meaning that adding age and attenuation of flows, such as sedimentation in large open
infrastructure for additional treatment, along with storage and water bodies, filtration through soils (or any other filtration me-
distribution, is relatively straightforward (Thomas et al., 1997). dia), or biological uptake by plants and microorganisms. These
systems are usually energy efficient and highly diverse (Table 1).
Recent Advances and Trends These systems were originally known as Best Management Prac-
tices for storm water management (WEF and ASCE, 1998). They
Evolution of Technologies: From Storm water Treatment are now becoming known as Low Impact Development systems in
to Storm Water Harvesting the USA (Department of Environmental Resources, 1999), Water
Urban storm water harvesting systems are made up of four Sensitive Urban Design (WSUD) in Australia (Wong, 2006), and
main components: collection, treatment, storage, and distribu- Sustainable Urban Drainage Systems in the UK (CIRIA, 2000b).
tion, which may also play a role in flood protection (Fig. 1). Ur- There are numerous design guidelines for their construction, with
ban storm water is generated in large quantities within short peri- almost every developed country, state or, region issuing their own
ods, and the nature of this pollutant source is diffuse (Burton and manual (e.g., Auckland Regional Council, 2002; CIRIA, 2001;

Table 1. Storm water treatment measures and their pollutant removal effectiveness for pollution control and storm water harvesting applications
(Source: CIRIA, 2000a, 2000b; WEF and ASCE, 1998; Wong, 2006).
Effectiveness of treatment measure
Storm water Pollution control Storm water harvesting
treatment measure Pollutant removal mechanisms Sediment Nutrients Heavy metals Sediment Nutrients Heavy metals Pathogens
Swales and filter strips sedimentation 99 9 99 9 9 9 ?
coarse filtration
infiltration
Ponds and wetlands sedimentation 999 99 999 99 9 99 99
filtration
biological processes
infiltration
evapotranspiration
Biofiltration systems sedimentation 999 99 999 999 9 999 ?
filtration
biological processes
infiltration
evapotranspiration
Infiltration systems sedimentation 999 99 999 99 9 99 99
filtration
biological processes
infiltration
evapotranspiration

S-118 Journal of Environmental Quality • Volume 37 • September–October [Supplement] 2008


Wong, 2006). Regulations in many countries also require the treat- Selected Promising Storm Water Harvesting Technologies
ment of urban storm water to mitigate pollution effects.
These systems have only recently been considered for use
Porous Pavement Technologies
in storm water harvesting, motivated primarily by water re- Due to the growing needs for storm water harvesting, some
source shortages. Traditional storm water treatment systems novel concepts are emerging that may enhance the potential for
have been retrofitted, or new systems have been built for har- application of storm water harvesting systems across a greater range
vesting. Australia is a leading example due primarily to its se- of situations. One such technology is based on the proven concept
vere long-term drought. In the last few years, the practice has of porous pavement, which has been used for decades in storm
increasingly been supported by the authorities as a valuable water flow and pollution management. Traditional design of a
alternative water resource (Brown and Clarke, 2007; Com- porous pavement system consists of a porous surface overlaying a
monwealth of Australia, 2002; Radcliffe, 2004). filter layer (a bedding material) that is placed on top of a sub-base
A review of storm water harvesting systems in Australia was (usually divided by geotextile). The porous surface can be modular
recently conducted (Hatt et al., 2006a) with respect to sizing, (unbound individual and nonporous blocks, laid down with gaps
performance, operation, and integration into the total water in between) or monolithic (asphalt or concrete without fine aggre-
cycle. Key observations of that review are reported below. gate; i.e., the entire surface is porous). The sub-base may contain a
collection pipe for drainage. The systems are usually installed in car
1. System Components. Storm water harvesting is largely parks and sections of streets with low traffic volume.
restricted to smaller scale sites, and the water has been In some countries (e.g., the UK, Sweden, Japan, and the
used for nonpotable purposes, overwhelmingly for USA), porous pavements have been widely used for control
urban irrigation. Collection and storage were based of storm water (Newton et al., 2003). They have infiltration
on conventional methods, such as gutters, pits, pipes, capacities usually upward of 4500 mm h−1 when new. Reduc-
and channels for collection and large centralized stores. tions in annual runoff coefficients from around 0.95 for normal
Treatment was based mainly on WSUD techniques; pavement to around 0.4 for porous pavement are typical, with
however, advanced techniques (outlined in the many authors reporting even greater reductions (Pratt et al.,
following section) and disinfection were occasionally 1995). Consistent water quality performance has been observed
used if there were perceptions of health risk. from porous pavers, with reductions in total suspended solids,
2. Design. The same design methods were used for reuse total P, and total N of around 80, 65, and 60%, respectively,
schemes as for storm water pollution control alone. This and hydrocarbon and metal reductions commonly around 85
may cause some problems because such systems were never and 75%, respectively (Berbee et al., 1999; Pratt et al., 1995;
intended to be used to deliver reliably high water quality Pagotto et al., 2000). Early perceptions of clogging and struc-
(in particular the variability of treatment may be an issue). tural problems have hindered the adoption of porous pave-
Design of surface stores was not well defined; sometimes ment. It has been shown that most systems are not very prone
the storage was retrofitted into existing ponds and urban to clogging, with localized problems being relatively easy to fix.
lakes rather than sized to meet required demand. Studies show after 15 to 20 yr of operation, they provide high
3. Construction, Maintenance, and Operation. Construction infiltration rates (100–1000 mm h−1) (Pratt et al., 1995; Davies
tolerance in these systems was generally finer than et al., 2002; Pratt et al., 1981). These systems are rather inex-
in conventional systems. Despite the importance of pensive and suitable for densely populated areas where other
maintenance, a number of studied sites seem to have storm water treatment systems are usually not applicable.
been neglected since completion of construction due There are two cases in Australia were porous pavement was
to a lack of defined maintenance regimes. Therefore, it used for collection, treatment, and storage. In storm water reuse
was likely that the integrity of a number of the systems schemes in Manly, Sydney (McRae, 2002), modular systems were
studied has been at risk, potentially affecting water used to collect water from a 3-ha, densely populated residential
quality, public health, and amenity. catchment. Treatment was provided by filtration through engi-
4. Implementation Issues. Although there was significant public neered material that comprised of expanded polymer made from
acceptance of storm water harvesting, one of the most recycled plastics with sufficient strength to carry normal pave-
frequently encountered issues was lengthy negotiation, ment loadings and a high cation exchange capacity and biofilm
assessment, and approval processes. This was largely due to density to remove bacteria. Treated water was stored in an under-
a lack of experience and policies on the part of the water ground (400 m3) modular store. There was no proper monitoring
industry and relevant authorities. However, it was found program, and only sporadic data have been collected. These data
that a partnership approach, often between the various showed relatively high removal of heavy metals and Escherichia
levels of government and a private developer as well as coli (Department of Environment and Conservation, 2006).
design teams and researchers, was highly successful because A similar system was built at Homebush Bay in Sydney
it enabled all parties’ skills, roles, and experience to be (Hatt et al., 2004a), where porous pavement and ecosoils
combined. Ongoing monitoring of system performance (soils amended to have specific properties that are conducive
was shown to be critical. to water quality improvement) were used for collection and
treatment, respectively. The harvested water was directly used

Fletcher et al.: Reuse of Urban Runoff in Australia S-119


Fig. 2. Conceptual drawing of the Monash University storm water harvesting system.

for irrigation of nearby trees, and therefore the system did not acceptable. Among the studies of biofiltration system perfor-
have storage. A small system built in the UK is used to collect, mance, there is a lack of studies into the removal efficiency for
treat, and store storm water runoff from a car park (375 m2) pathogens. Given the public health implications of high levels
that is then used for flushing toilets in a youth hostel (Pratt, of bacteria, viruses, or protozoans in harvested storm water,
1999). The interesting feature here is that the sub-base of the this is an important gap that needs to be addressed.
pavement was used to treat and store the harvested water. There are several storm water harvesting systems that use
It has been concluded that porous pavements can provide the biofiltration as their treatment mechanism currently being
basis for further developments in treatment and storage of storm trialed in Australia, including one at Monash University in
water for reuse. Technologies for achieving an effective imple- Melbourne, which has been extensively monitored. This sys-
mentation of porous pavements for storm water reuse are still tem captures runoff from a carpark; the runoff passes through
to be refined, but they seem to have great potential, particularly pretreatment settling tanks and then into a fully lined biofiltra-
given their ability to provide treatment and storage within the tion system (Fig. 2). The treated water is stored in an adjacent
existing impervious surface, thus overcoming space constraints. pond before being used to irrigate a nearby sports oval. Three
Biofiltration Technologies different filter media types were trialed in separate filter strips:
sandy loam; 80% sandy loam, 10% vermiculite, 10% perlite;
Biofiltration is an emerging technology that has been used as and 80% sandy loam, 10% composted pine bark, and 10%
an effective treatment of storm water, primarily for protection hardwood mulch. The entire biofilter is densely vegetated with
of aquatic ecosystems, and more recently for nonpotable use. native species that are tolerant of variable wetting and drying.
For example, reductions in heavy metal and total suspended The hydraulic and treatment performance of the biofilter
solids concentrations through biofiltration media of more than was monitored via a fully automated monitoring system. Lev-
90% are commonly observed (Davis et al., 2003; Hatt et al., eled V-notch weirs and ultrasonic depth sensors were used to
2008). However, nutrient removal tends to be far more vari- continuously (1-min time step) measure flow rates, whereas au-
able, depending on the filter media (Hatt et al., 2006b, 2007; tosamplers collected flow-weighted water quality samples at the
Henderson et al., 2007), the presence of vegetation (Henderson inlet and three outlet points (i.e., from each filter strip; Fig. 2).
et al., 2007), plant type (Read et al., 2007), and the presence of On average, 30 to 35% of inflow volume was lost via evapo-
an anaerobic zone to promote coupled nitrification–denitrifi- transpiration. Although this is beneficial in terms of pollution con-
cation within the filter medium (Hsieh and Davis, 2005; Kim trol (reducing loads and ensuring that contaminants in the filter
et al., 2003). Most studies show good removal of total Kjeldahl are held there for longer periods, thus allowing treatment processes
nitrogen and NH3 that is often offset by leaching of NO3 to occur), this loss of water is an important consideration when
(Davis et al., 2001, 2003; Henderson et al., 2007; Hatt et al., designing a storm water harvesting system. In terms of pollutant
2008). Phosphorus removal is typically high (e.g., Davis et al., removal, removal of sediment and heavy metals was consistently
2006) and may increase with depth (Hatt et al., 2008). high, outflow concentrations of nutrients were variable, and all
Results from laboratory studies are not always repeated three biofilter strips tended to leach P and N (Table 2). In addition
by monitoring of field systems, particularly due to variations to the regular water quality monitoring, two spiking tests were per-
in soil properties (Hunt et al., 2006) and climate conditions. formed to evaluate the pathogen removal efficiency of the biofilter.
For storm water harvesting systems, this variability may be The pathogen surrogates—baker’s yeast (Saccharomyces cerevisiae),
more problematic than it is for the protection of aquatic eco- E. coli, and the coliphage MS2—were used in these tests to rep-
systems, where some periods of poor water quality may be resent the behavior of the protozoan pathogen Cryptosporidium,

S-120 Journal of Environmental Quality • Volume 37 • September–October [Supplement] 2008


bacterial pathogens, and human Table 2. Mean event-mean pollutant concentrations at the inlet and outlets of the Monash University biofilter. †
enteric viruses, respectively. Sampling point TSS TP FRP TN NH4+ NOx DON Cu Pb Zn
The mean log10 reductions for ––––––––––––––––––––––––––––––– mg L−1–––––––––––––––––––––––––––––––
each of the microbial surrogates Inlet 39 0.072 0.006 1.1 0.073 0.37 0.31 0.016 0.0052 0.25
were as follows (ranges are Outlet 1 (SL)‡ 5 0.21 0.10 1.2 0.023 0.40 0.64 0.0055 0.0017 0.029
Outlet 2 (SLVP) 3 0.16 0.10 1.2 0.016 0.53 0.57 0.0051 0.0015 0.014
shown in parentheses): E. coli,
Outlet 3 (SLCM) 4 0.16 0.10 1.0 0.020 0.15 0.64 0.0037 0.0017 0.013
1.63 (1.06– 3.30); yeast, 2.29 † DON, dissolved organic nitrogen; FRP, filterable reactive phosphorus; NOx, nitrate + nitrite;
(0.84–3.93); and bacteriophage TN, total nitrogen; TP, total phosphorus; TSS, total suspended solids.
MS2, 1.79 (0.50–3.21) (Petter- ‡ SL, sandy loam; SLCM, sandy loam/compost/mulch; SLVP, sandy loam/vermiculite/perlite.
son et al., 2007).
Like porous pavements, Post-treatment Disinfection
biofiltration systems are a promising treatment option for There is limited knowledge about the levels of pathogenic mi-
storm water harvesting systems. Although further develop- croorganisms in storm water and even less knowledge regarding
ment is required to refine their design with respect to nutrient the removal efficiency of pathogenic microorganisms achieved by
removal, they are able to reliably and efficiently remove sedi- the range of treatment and storage components that are currently
ment and heavy metals. Questions remain about their ability in use in typical storm water harvesting and treatment systems
to remove produce water that satisfies pathogen levels that (McCarthy et al., 2007; Petterson et al., 2007). There is thus un-
satisfy end-use requirements. certainty surrounding the requirement for post-treatment disin-
fection (where the water is disinfected immediately before being
Application of Traditional Physico-chemical Water distributed to its end uses). In the case of Australian systems, the
Treatment Technologies inclusion of a disinfection component has not been a regulatory
Storm water can be treated using the traditional physicochemi- requirement, and limited monitoring has been conducted on
cal processes more commonly used for water supply and wastewa- post-storage levels of microorganisms to determine the need for
ter treatment, although this is less common than the WSUD-based disinfection. In this context, when disinfection is included in
systems described. The physicochemical processes that have been the system it is largely as a precaution rather than as an essential
used to treat storm water include filter screens, sedimentation component of the treatment train. Possible disinfection methods
tanks, lagoons, trickling filters, aeration, activated sludge tanks, and include UV light, chlorination, ozonation, and microfiltration.
more advanced processes such as dissolved air flotation, membrane Emerging technologies, such as solar photocatalysis (Blangis and
filtration, high rate clarification, and biological nutrient removal Legube, 2007), may help to further improve the reliability of
(BNR). A range of proprietary products using physicochemical treatment while reducing its cost.
techniques have been developed in recent years, including filtration Ultraviolet light is the preferred disinfection technique based
systems, catch basin inserts, in-line flocculation, and vortex separa- on cost, environmental risks, ease of operation, and its appro-
tors. A high-profile example of the use of more advanced processes priateness across a wide range of scales (Mitchell et al., 2006).
as part of the treatment train is the Santa Monica SMURRF facil- However, it does not provide residual disinfection, which may be
ity in California, which uses coarse and fine screening, dissolved air required when there is an extensive distribution system (due to
flotation, microfiltration, and ultraviolet (UV) disinfection (Boyle the detention time in the distribution network). Ultraviolet treat-
Engineering Corporation, 1999). ment is used at the Santa Monica SMURRF facility, Royal Park
Where separate storm water and wastewater drainage systems Storm water Reuse System (Melbourne, Australia), and the South
are used, the use of advanced treatment techniques is largely re- Australian Museum in Adelaide, whereas filtration is used at
stricted to situations where storm water and wastewater are being Kogarah Town Square (Sydney, Australia) (Mitchell et al., 2006).
harvested for end uses where human contact is intended or pos- Based on the post-storage quality in three storm water har-
sible. Examples of these combined storm water/wastewater har- vesting systems being monitored by the authors (located in Mel-
vesting systems are Homebush Bay and Taronga Zoo in Sydney, bourne, Australia), poststorage turbidity levels are likely to limit the
Australia (Hatt et al., 2006a). These systems typically have a high effectiveness of storm water disinfection unless care is taken to en-
public profile, and the end use is widely distributed; it is therefore sure low turbidity levels are maintained. For example, weekly grab
more difficult to control public access. sample measurements of the turbidity of the storm water delivered
Physico-chemical treatment is typically more expensive than to the UV disinfection unit at the Royal Park Storm water Reuse
traditional biophysical storm water treatment, is less likely to system have ranged from 0 to 64 nephelometric turbidity units,
provide visual amenity benefits (compared, for example, with whereas the UV unit was designed assuming influent turbidity
a landscaped wetland and storage pond), and is more techni- levels of no higher than 10 nephelometric turbidity units.
cally complex to operate (Mitchell et al., 2006). However, these The disinfection performance of this UV unit is being as-
physicochemical treatment systems typically require less land sessed with a challenge test using MS2 bacteriophage, with
area and can be more intensively controlled to produce a more preliminary results of this research reported by Petterson et
constant quality of treated storm water (Hatt et al., 2006a) be- al. (2007). Given the inactivation of MS2 estimated from the
cause they do not rely on the variability of biological processes. challenge test data, the calculated dose–response relationship

Fletcher et al.: Reuse of Urban Runoff in Australia S-121


Table 3. Estimated Log10 reduction by Royal Park UV disinfection unit and water usage patterns. However, the general relationship be-
for selected reference pathogens (Source: Petterson et al, 2007). tween lifecycle costs and the number of connections is expected
Reference organism Log10 reduction at 5.217 mJ cm−2 to be generally applicable. This observation is supported by
Campylobacter 2.687 similar economies of scale trends in gray water reuse (Booker,
Salmonella 4.591
1999), with costs falling rapidly as the scale grew from 1 to 100
Adenovirus 0.1252
Rotavirus 0.5321 connections and then leveling out after 1000 connections.
Cryptosporidium 2.261 The question of scale has other elements related to respon-
sibility for operation and maintenance and the associated risks
was used to estimate the actual dose received within the Royal (Coombes and Mitchell, 2006; Cunliffe, 2004). Although the al-
Park UV unit. This dose was used to estimate the expected lotment scale may offer flexibility, it relies on the knowledge and
Log10 reductions of reference pathogens presented in Table 3. commitment of individual landholders to recognize and manage
these risks. In addition to issues of scale, there is a wide range
Integration and Scale of issues relating to the integration of storm water harvesting
The spatial scale of the storm water harvesting systems imple- systems into the landscape and into the urban water cycle. For
mented to date has generally been determined by opportunistic example, urban storm water harvesting systems may be required
drivers rather than strategic considerations such as the relation- to provide additional services, such as flood protection, amenity
ship between scale and cost (Mitchell et al., 2005). As the storm and recreational opportunities, aquatic habitats, and protection
water harvesting industry matures in the coming years and expe- of receiving waters (in terms of water quality and restoration of
rience develops, storm water harvesting may be a component of flow regimes toward their predeveloped levels) (Mitchell et al.,
long-term water management strategies, supported by govern- 2007). Rainwater tanks may be designed with a predetermined
ment policies about the integration of storm water harvested freeboard storage to reduce or delay flow peaks (Coombes et al.,
within integrated urban water management plans. 2003). Similarly, retrofit of online storm water harvesting stores
If this occurs, then strategic considerations about spatial scale into a retarding basin has been shown to be a feasible means of
should inform this policy and planning process. Initial desktop reducing peak flows from average recurrence intervals ranging
research into this topic has been conducted (Mitchell et al., from 1 to at least 50 yr (Mitchell et al., 2006).
2005), looking at the life cycle cost of 1, 100, 1000, and 10,000 Storm water harvesting systems may also need to be integrat-
connections in Melbourne, Australia (in this case study, the single ed into the landscape, and thus the design of the storage system
connection systems were rainwater tank systems). The life cycle may need to take into account such issues as maintenance of
costing accounted for the total acquisition, renewal, and decom- aquatic vegetation and the aesthetic effect of excessive drawdown
missioning costs of all the capital items and for the maintenance of the store (Mitchell et al., 2007). In many situations, it may
and operating costs over a 50-yr period, using a discount rate of be necessary to separate out the different sources of water (e.g.,
5.2% per annum. The systems contained necessary infrastructure rainwater from roof areas and storm water from other surfaces
for the collection, storage, and distribution of storm water. Mini- such as roads and carparks) in terms of the type of storage and
mal storm water treatment was included beyond that provided as the likely differences in water quality of the different sources
part of the storage components of the system (primarily due to (Coombes et al., 2003; Coombes and Mitchell, 2006).
sedimentation). A large number of scenarios were analyzed, con-
sidering two residential densities (15 and 40 households per hect- Reported Performance
are), a range of rainwater tank and storm water storage sizes, and
In this section, reported performance of storm water har-
different combinations of end use types (toilet flushing, laundry
vesting systems is reviewed, with consideration not only of the
cold water, hot water, and garden irrigation).
primary objective (providing a substitute for potable water) but
The results of this analysis provided an upper and lower
also of the integrated objectives, such as environmental impacts,
bound cost per m3 (Fig. 3). These per m3 costs will be influ-
flooding, and landscape amenity. Data on the performance of
enced by location specific factors, such as storm water runoff
storm water harvesting systems are relatively scarce, with few
systems having been monitored or even documented (Hatt et
al., 2006a). A synthesis of available data is provided here. The
following section, which reports on key research gaps, outlines
some of the most pressing needs for future monitoring.

Water Savings and Financial Benefits


Few studies are available that report on storm water harvesting,
with most reporting only the performance of rainwater harvesting
(i.e., water from rainfall onto roofs) systems. For example, Gardner
et al. (2006) monitored water use from a 22-lot subdivision with
Fig. 3. Life cycle cost of rain water and storm water harvesting for
individual household tanks coupled with water-efficient appli-
a range of residential storm water harvesting scales (adapted ances. With large allotments (average size of 1100 m2), the contri-
from Mitchell et al., 2005).

S-122 Journal of Environmental Quality • Volume 37 • September–October [Supplement] 2008


Table 4. Performance data for storm water reuse systems (adapted from Hatt et al., 2006a). The equivalent cost for TN reduction is based on the
equivalent cost of constructing a storm water treatment wetland to remove the given mass of TN.
Pollution reduction
Capital Mean annual Runoff Potable water Potable water Equivalent cost
Site cost runoff collected Storage volume use reduction cost savings TN† removed for TN reduction
$A‘000 mL % % of runoff volume % % of capital cost kg $ ha−1
Inkerman Oasis 434 7.6 20 1 30 0.1 1.1–9.1 684–2036
Figtree Place 110 6.3 83 5 65 1.4 3.7–31 2636–22,454
Kogarah 629 9.4 85 15 17 0.4 5.6–48 4400–37,700
Oaklands Park 73 75 100 65 – 30.8 53–450 237–2036
Hawkesbury 3900 800 50 11 28 3.1 280–2400 2204–18,889
Homebush Bay 15,800 1179 100 42 50 2.2 825–7074 855–7325
Parafield 4500 2210 100 29 – 14.7 1547–13,262 761–6523
† TN, total nitrogen.

bution of harvested rain water was around 56% for the occupied the construction of alternative water supply infrastructure (such
lot monitored. Coombes et al. (1999) monitored a system where as a new dam). Tools such as Triple Bottom Line Assessment
rain water was collected in a centralized tank, whereas storm water (Taylor and Wong, 2002) may help to overcome the challenge
(from roads, pathways, carparks, etc.) was conveyed to a detention of accounting for all costs and benefits.
basin (but not used). Based on the use of rainwater alone, savings
in potable water demand of around 60 to 65% were obtained. Environmental Benefits
Hatt et al. (2006a) reviewed the performance data of 77 Two categories of environmental benefit are commonly
storm water harvesting systems. Of these, useable data on water demonstrated from storm water harvesting systems: (i) at-
collected and used were available for seven sites (Table 4). At tenuation of postdevelopment hydrology and (ii) reductions
these sites, the percent collected ranged from 20 to 100% of the in pollutant concentrations and/or loads. Most of the studies
mean annual runoff, and the reductions in potable water demand of these impacts rely on modeling, with little or no monitor-
ranged from 17 to 65%. Generally, the sites that collected a low ing data available (Hatt et al., 2006a).
proportion of mean annual runoff were those that were retrofit- The harvesting of storm water will affect downstream water
ted to an existing storm water treatment facility rather than pur- quality. It will reduce pollutant loads to receiving waterways as a
pose built for the harvesting of storm water. Both figures depend function of the water taken away (depending on the use and dis-
on the size of the storage relative to the mean annual runoff vol- posal of that harvested water). For example, if urban storm water
ume, which in turn depends on site constraints. has a typical concentration of around 2.5 mg L−1 total N (Duncan,
Hatt et al.’s review showed that the capital cost of storm 2006; Taylor et al., 2005; Brombach et al., 2005), each mL of wa-
water harvesting systems was related not only to catchment ter harvested will reduce the N load to the receiving water by 2.5
size; factors such as treatment method, land characteristics kg. Hatt et al. (2006a) found significant reductions in N loads to
(slope, presence of rock, etc.), and storage type also affected receiving water and quantified the equivalent cost of this reduction
the cost. The value of potable water savings (Table 4) varied using well established costs for constructing storm water treatment
greatly, from 0.1 to 30.8% of the system’s capital cost. Based wetlands. In the best case, this resulted in an equivalent reduction
on these figures, the payback period (based simplistically on in required storm water treatment of A$37,000 per hectare.
the capital cost alone) ranged from 3 to 950 yr. The influence of storm water harvesting on pollutant concen-
Hatt et al. (2006a) note that there are few data available on trations is more complex. If only roof water, with its lower concen-
the costs and benefits of such systems. Of the systems reviewed trations of particulate pollutants (Duncan, 2006), were harvested,
by Hatt et al. (2004a; 2006a), few gave indications of the unit the concentration at the receiving water may increase. This is un-
cost of water produced. Each of the three that did (Homebush likely to be the case when harvesting of overall storm water runoff.
Bay and Parafield in Australia and Santa Monica in the USA) However, no known monitoring data exist on the impact of storm
showed different outcomes. For example, the Homebush Bay water harvesting on downstream pollutant concentrations.
system cost twice as much as the retail cost of normal reticulat- Storm water harvesting systems can have a significant benefit
ed town water (but was sold at 15% less than the mains water in terms of reducing flow peaks downstream. For example, even
price to encourage uptake). The Santa Monica system provided a distributed network of rainwater tanks specially designed to
a near-neutral outcome, whereas the Parafield system produced provide permanent freeboard by having a “trickle-outlet” at a
water at 30% of the cost of the alternative (pumping from a predetermined point below the top of the tank has been shown to
nearby river). This diversity shows that local context is critical significantly reduce peak flows (Hardy et al., 2004). More detailed
in determining the financial outcomes. modeling showed reductions of around 40 to 50% in the 3-mo
Several authors have asserted that analysis of financial bene- average recurrence interval peak flow, dropping to around 5 to
fits purely in terms of potable water savings provides an incom- 10% for the 100-yr average recurrence interval event. The influ-
plete picture because it excludes the externalities (Coombes et ence of storm water harvesting on flow behavior was also examined
al., 2002), such as reductions in the size of downstream storm by Fletcher et al. (2007). They used the concept of “environmen-
water conveyance and treatment infrastructure and avoiding tal flows,” which is more typically used in the rural context, to

Fletcher et al.: Reuse of Urban Runoff in Australia S-123


determine the amount and temporal regime of flows necessary underpinning to enable these schemes to be designed, main-
to sustain healthy aquatic ecosystems (Poff and Allan, 1995). In tained, and operated to ensure adequate long-term performance
the urban context, with increased runoff from impervious areas, in terms of human health and environmental risks (Hatt et al.,
Fletcher et al. (2007) determined to what extent storm water 2004a; 2006a). Inadequate data exist on the reliability of these
harvesting could be used to restore those flows to predevelopment systems in terms of end use water quality and the supply of
levels. The feasibility of such a strategy varied with climate (harder storm water. Due to the difficulty and costs of data collection
in subtropical than temperate climates), whereas high levels of and analysis, very few storm water recycling systems have been
urban density (as measured by imperviousness), while increasing monitored. This is of serious concern for the following reasons:
the magnitude of flow reduction required, also had the advantage (i) These systems are built to provide potable substitution, but
of a higher water demand. They found that in some cases, excessive there is a dearth of quantification of the actual volumes of pota-
harvesting could result in an overextraction of flow and recom- ble substitution achieved compared with design estimates, and
mended that for each catchment an analysis of the environmental (ii) there are insufficient data to conduct robust, quantifiable
flow requirements be undertaken to set operating rules for any assessments of likely environmental and human health risks for
proposed storm water harvesting system. The aim should be to existing and future storm water harvesting projects.
emulate the predevelopment flow regime (based on indicators such As a result, the monitoring of storm water harvesting sys-
as frequency of surface runoff to streams, peak flows for a range of tems and the assessment of risks and reliability is an active
recurrence intervals, length of dry and wet spells, etc.). This should area of research. For example, a recently commenced project is
include consideration of ground water hydrology, with an aim be- monitoring three operational storm water harvesting systems
ing to emulate the predevelopment infiltration of water to ground and the performance of biofilters for storm water treatment
water on a catchment scale. No known studies exist that relate the and harvesting (Table 2). In the interim, design will continue
health of aquatic ecosystems to the construction and operation of to be based on incomplete information.
storm water harvesting systems within the catchment.
Data Requirements for Assessing Lifecycle Costs
Social Acceptance Assessing the cost of storm water harvesting systems has
Several studies have been undertaken on the social acceptance proven to be complex, and as a result there has been relatively
of storm water harvesting. Generally, there is a preference for us- little progress in capturing reliable data on capital and operat-
ing storm water over wastewater given that the “history” of water ing costs of even storm water treatment systems (Fletcher et al.,
significantly affects its acceptance by the community (McIntosh 2005; Fletcher and Taylor, 2007; Rozis and Rahman, 2002), let
and Pugh, 1991; Po et al., 2004). Generally, perceived health alone storm water treatment systems (Hatt et al., 2006a). Because
risks are lower for use of storm water than they are for reuse of storm water harvesting systems are a relatively recent area of
recycled wastewater (ACTEW, 1994; ARCWIS, 1999). For ex- endeavor, there has been inadequate time for reliable datasets to
ample, one study in Perth showed 96% support for using storm be obtained (Gardner et al., 2006; Lloyd et al., 2001a, 2002a). A
water for garden watering and 95% support for its use for toilet more fundamental limitation is the difficulty in quantifying the
flushing. Support for personal or clothes washing was lower (at externalities, such as the reduction in environmental degradation
50 and 68%, respectively), but still much higher than for waste- to downstream receiving waters by retaining storm water (Lloyd
water. Community support for the use of rain water (collected et al., 2002a; 2001b; Prato, 1999). Although the use of broader
from roofs only) is generally higher than that for general storm assessment tools (e.g., Triple Bottom Line assessment protocols
water runoff (Coombes and Mitchell, 2006). Community sup- [Taylor and Fletcher, 2006]) is a critical goal, there remains the
port has been shown to be higher for large-scale systems, where basic need to have data with which the economic performance
there is a perception that there will be a high level of regulation of storm water harvesting systems can be assessed and later pre-
and risk management (Mitchell et al., 2005). dicted. Hatt et al. (2006a) concluded that without such data on
Some impressive examples of the development of large-scale which to base investment decisions, governments and private
storm water reuse exist that demonstrate community acceptance investors will be reluctant to fund storm water harvesting. It is
of the approach. For example, de Graaf et al. (2007) show the imperative that attention be paid to the development and ap-
transformation of Rotterdam’s water supply system to one that plication of standard protocols for the collection of lifecycle cost
relies significantly on storm water harvesting and that provides a data for storm water harvesting systems.
wide range of other social benefits. Although there are few such
examples, it seems clear that public acceptance of storm water Storm Water End Use Water Quality Guidelines
harvesting is relatively high but is likely to depend on issues of Internationally, there has been a general lack of end use and
culture and local context (Brown and Davies, 2007). receiving water quality guidelines that have been specifically devel-
oped for the harvesting of storm water. In lieu of these, wastewater
Remaining Challenges and Knowledge Gaps reuse and drinking water guidelines have often been used to pro-
vide end use water quality design targets (Hatt et al., 2006a).
Data Requirements for Assessing Reliability and Risk In Australia, Phase 2 of the national guidelines for water recy-
Largely due to limited experience in storm water harvesting cling (NRMMC and EPHC, 2006), which are currently in devel-
and the lack of monitoring data, there is inadequate scientific opment, will include specific guidance for storm water harvesting.

S-124 Journal of Environmental Quality • Volume 37 • September–October [Supplement] 2008


The risk management taken in these new guidelines represents a they need to become able to be readily deployed in an “off-
departure from more prescriptive guideline approaches previously the-shelf ” manner by reducing complexities in the design and
used in the field of water reuse. However, there is considerable ex- construction process.
perience in the use of risk management frameworks; for example,
the food industry has used the HACCP (hazard analysis and criti- References
cal control point) system for some time, and the HACCP is seen Aalderink, R.H., L. Lijklema, and J.B. Ellis (ed.) 1990. Urban storm water
internationally as best practice for ensuring food safety (Codex quality and ecological effects upon receiving waters. Vol. 22 (10/11).
International Assoc. of Water Pollution Research and Control,
Alimentarius Commission, 1997; NRMMC and EPHC, 2006). Pergamon Press, Oxford, UK.
The same technique has recently been successfully applied to ACTEW. 1994. Australian Capital Territory future water supply strategy: our
storm water harvesting, using data on the variability of inflow and future water supply. ACT Electricity and Water, Canberra, Australia.
outflow pollutant concentrations (Page et al., 2006). Quantitative ARCWIS. 1999. The social basis for urban water provision in the 21st century.
Australian Research Centre for Water in Society, CSIRO, Perth, Australia.
Risk Assessment and Quantitative Microbial Risk Assessment are Arthington, A.H., D.A. Milton, and R.J. McKay. 1983. Effects of urban
other tools that have been widely used in the area of water reuse development and habitat alterations on the distribution and abundance
(NHMRC and NRMMC, 2004; NRMMC and EPHC, 2006). It of native and exotic freshwater fish in the Brisbane region, Queensland.
Aust. J. Ecol. 8:87–101.
is likely that in the short term, the limiting factor in the application
Auckland Regional Council. 2002. Storm water management devices: Design
of these new guidelines for storm water harvesting will be the lack guidelines manual. Revision to Technical Publication 10. Auckland
of data that can be used for robust quantitative risk assessment. Regional Council, Auckland, New Zealand.
Australian Bureau of Statistics. 2006. Yearbook 2006. Australian Bureau of
Water–Energy Tradeoffs Statistics, Canberra, Australia.
Berbee, R., G. Rijs, R. de Brouwer, and L. van Velzen. 1999. Characterisation
There is growing recognition of the need to assess the energy and treatment of runoff from highways in the Netherlands paved with
requirements of water management and to consider potential impervious and pervious asphalt. Water Environ. Res. 71:183–190.
Blangis, D., and B. Legube. 2007. Traitement des eaux de pluie par
water and energy tradeoffs in planning and design, aiming to photocatalyse solaire. J. Eur. Hydrol. 38:121–130.
maximize the water yield while minimizing the embodied energy Booker, N. 1999. Estimating the economics of scale of greywater reuse systems.
and the operational energy of a system (Gardner et al., 2006). Program Rep. FE-88. CSIRO Molecular Science, Melbourne, Australia.
Based on current practice, storm water management rep- Booth, D.B. 1990. Stream channel incision in response following drainage
basin urbanization. Water Resour. Bull. 26:407–417.
resents a modest proportion of the total operational energy Booth, D.B., and C.R. Jackson. 1997. Urbanization of aquatic systems:
of an urban water system as gravity drainage and biophysical Degradation thresholds, storm water detention, and the limits of
storm water treatment (when treatment occurs) dominate. In mitigation. J. Am. Water Resour. Assoc. 33:1077–1090.
contrast, a storm water harvesting system may involve consider- Bourguès, S., and B.T. Hart. 2007. Nitrogen removal capacity of wetlands:
Sediment versus epiphytic biofilm. Water Sci. Technol. 55:175–182.
able pumping, and treatment may use more or less operational Boyd, M.J., M.C. Bufill, and R.M. Kness. 1993. Pervious and impervious
energy depending on the processes used (Grant et al., 2006). runoff in urban catchments. Hydrol. Sci. J. 38:463–478.
However, the embodied energy and operational energy intensi- Boyle Engineering Corporation. 1999. Engineering report for dry-weather
runoff, reclamation, storage, pumping, distribution, and non-potable water
ty (defined here as the amount of operational energy consumed use area facilities: Report No. FR-C77-100-02. City of Santa Monica, CA.
per unit of water supplied to the end uses) of a storm water Brombach, H., G. Weiss, and S. Fuchs. 2005. A new database on urban
harvesting scheme should be assessed relative to that of the runoff pollution: Comparison of separate and combined sewer systems.
other sources of water being considered or substituted. Water Sci. Technol. 51:119–128.
Brown, R., and J. Clarke. 2007. Transition to water sensitive urban design;
If the storm water is being used for irrigation, some of the the story of Melbourne, Australia. Melbourne: Facility for Advancing
energy requirements of the harvesting system could be offset by Water Biofiltration, and National Urban Water Governance Program,
energy savings in terms of summer cooling of buildings (Mitchell Melbourne, Australia.
Brown, R.R., and P. Davies. 2007. Understanding community receptivity to water
et al., 2007). This is because urban irrigation has the potential to re-use: Ku-ring-gai Council case study. Water Sci. Technol. 55:283–290.
passively control urban microclimate through evapotranspiration. Burton, G.A.J., and R.E. Pitt. 2001. Storm water effects handbook: A
toolbox for watershed managers, scientists, and engineers. CRC Press
Retrofit Technologies LLC, Boca Raton, FL.
Carleton, J.N., T.J. Grizzard, A.N. Godrej, H.E. Post, L. Lampe, and P.P.
If storm water harvesting is to move beyond its present Kenel. 2000. Performance of constructed wetlands in treating urban
application primarily in greenfield sites, there is a need for storm water runoff. Water Environ. Res. 72:295–304.
technologies that allow integrated storm water harvesting, Carlson, T.N., and S.T. Arthur. 2000. The impact of land use—land cover
changes due to urbanization on surface microclimate and hydrology: A
treatment, storage, and distribution systems to be retrofitted satellite perspective. Global Planet. Change 25:49–65.
into existing urban areas. For example, where storm water CIRIA. 2000a. Sustainable urban drainage systems: Design manual for
treatment systems exist, there is a need to be able to retrofit England and Wales. CIRIA Report No. C522. London, UK.
the necessary additional treatment technologies to meet local CIRIA. 2000b. Sustainable urban drainage systems: Design manual for Scotland
and Northern Ireland. CIRIA Report No. C521. Dundee, Scotland.
end-use requirements and add storage and distribution capa- CIRIA. 2001. Sustainable urban drainage systems: Best practice manual for
bility. Although there are significant advances being made in England, Scotland, Wales, and Northern Ireland. CIRIA Rep. No.
this area, such as the adaptation of porous pavements and bio- CR086A. London, UK.
filtration systems, there remains the need for a wide range of Codex Alimentarius Commission. 1997. Hazard analysis and critical control
point (HACCP) systems and guidelines for their application: Annex to
suitable technologies to overcome the constraints imposed by CAC/RCP 1-1969, Rev 3. Available at http://www.fao.org/DOCREP/005/
retrofit conditions. If these systems are to be readily adopted, Y1579E/y1579e03.htm (verified 3 Apr. 2008). FAO, Rome, Italy.

Fletcher et al.: Reuse of Urban Runoff in Australia S-125


Codner, G.P., E.M. Laurenson, and R.G. Mein. 1988. Hydrologic effects for urban areas. Water Resour. Res. 27:1739–1755.
of urbanization: A case study. p. 201–205. In Proc. Hydrol. Water Groffman, P.M., N.J. Boulware, W.C. Zipperer, R.V. Pouyat, L.E. Band, and
Resour. Symp., Canberra, Australia. 1–3 Feb. 1988. The Institution of M.F. Colosimo. 2002. Soil nitrogen cycle processes in urban riparian
Engineers Australia, Sydney, Australia. zones. Environ. Sci. Technol. 36:4547–4552.
Commonwealth of Australia. 2002. The value of water: Inquiry into Harbott, E.L., and M.R. Grace. 2005. Extracellular enzyme response to
Australia’s urban water management. Rep. of the Senate Environment, bioavailability of dissolved organic C in streams of varying catchment
Communications, Information Technology, and the Arts Reference urbanization. J. North Am. Benthol. Soc. 24:588–601.
Committee. The Parliament of the Commonwealth of Australia, Hardy, M., P. Coombes, and G. Kuczera. 2004. An investigation of estate
Canberra, Australia. level impacts of spatially distributed rainwater tanks. p. 492–503. In
Coombes, P.J., J.R. Argue, and G. Kuczera. 1999. Figtree Place: A case study Int. Conf. WSUD, Adelaide, Australia. 21–25 Nov. 2004.
in water sensitive urban development. Urban Water 1:335–343. Harremoes, P. 1997. Integrated water and waste management. Water Sci.
Coombes, P.J., A. Frost, G. Kuczera, G. O’Loughlin, and S. Lees. 2003. The Technol. 35(9):11–20.
impact of rainwater tanks in the Upper Parramatta River catchment. Hatt, B.E., A. Deletic, and T.D. Fletcher. 2004a. Integrated storm water
Aust. J. Water Resour. 7:121–129. treatment and re-use systems– an inventory of Australian practice.
Coombes, P.J., G. Kuczera, and J.D. Kalma. 2002. An evaluation of the benefits CRCCH Report No.04/1. Available at http://www.catchment.crc.
of source control measures at the regional scale. Urban Water 4:307–320. org.au/archive/pubs/prog4.html (verified 3 Apr. 2008). Cooperative
Coombes, P.J., and V.G. Mitchell. 2006. Roofwater, storm water, and Research Centre for Catchment Hydrology, Melbourne.
wastewater reuse. In T.H.F. Wong (ed.) Australian runoff quality. The Hatt, B.E., A. Deletic, and T.D. Fletcher. 2006a. A review of integrated storm
Institution of Engineers Australia, Sydney, Australia. water treatment and re-use in Australia. J. Environ. Manage. 76:102–113.
Cunliffe, D.A. 2004. Guidance on the use of rainwater tanks. EnHealth Hatt, B.E., A. Deletic, and T.D. Fletcher. 2007. Storm water reuse:
(Australian government, National Public Health Partnership), Canberra. Designing biofiltration systems for reliable treatment. Water Sci.
Davies, J.W., C.J. Pratt, and M.A. Scott. 2002. Laboratory study of Technol. 55:201–209.
permeable pavement systems to support hydraulic modelling. 10 p. In Hatt, B.E., T.D. Fletcher, and A. Deletic. 2008. Hydraulic and treatment
Proc. 9ICUD. 8–13 Sept. 2002. Portland, OR. performance of fine media storm water filters. Environ. Sci. Technol.
Davis, A.P., M. Shokouhian, H. Sharma, and C. Minami. 2001. Laboratory 42:2535–2541.
study of biological retention for urban storm water management. Water Hatt, B.E., T.D. Fletcher, C.J. Walsh, and S.L. Taylor. 2004b. The influence
Environ. Res. 73:13–26. of urban density and drainage infrastructure on the concentrations and
Davis, A.P., M. Shokouhian, H. Sharma, and C. Minami. 2006. Water loads of pollutants in small streams. Environ. Manage. 34:112–124.
quality improvement through bioretention media: Nitrogen and Hatt, B.E., M. Siriwardene, A. Deletic, and T.D. Fletcher. 2006b. Filter media
phosphorus removal. Water Environ. Res. 78:284–293. for storm water treatment and recycling: The influence of hydraulic
Davis, A.P., M. Shokouhian, H. Sharma, C. Minami, and D. Winogradoff. properties of flow on pollutant removal. Water Sci. Technol. 54:263–271.
2003. Water quality improvement through bioretention: Lead, copper, Henderson, C., M. Greenway, and I. Phillips. 2007. Removal of dissolved
and zinc removal. Water Environ. Res. 75:73–82. nitrogen, phosphorus, and carbon from storm water by biofiltration
de Graaf, R., R. van der Brugge, J. Lankester, W. van der Vliet, and L. mesocosms. Water Sci. Technol. 55:183–191.
Valkenburg. 2007. Local water resources and urban renewal: A Rotterdam House, C.H., S.W. Broome, and M.T. Hoover. 1994. Treatment of nitrogen
case study. p. 2189–2196. In Novatech 2007; Sustainable techniques and and phosphorus by a constructed upland-wetland wastewater treatment
strategies in urban water management. 25–28 June 2007. Lyon, France. system. Water Sci. Technol. 29:177–184.
Denman, L., P. May, and P.F. Breen. 2007. An investigation of the potential Hsieh, C.H., and A.P. Davis. 2005. Evaluation and optimization of
to use street trees and their root zone soils to remove nitrogen from bioretention media for treatment of urban storm water runoff. J.
urban storm water. Aust. J. Water Resour. 10:303–311. Environ. Eng. 131:1521–1531.
Department of Environment and Conservation. 2006. Managing urban Hunt, W.F., A.R. Jarrett, J.T. Smith, and L.J. Sharkey. 2006. Evaluating
storm water: Harvesting and reuse. Dep. of Environment and bioretention hydrology and nutrient removal at three field sites in
Conservation, NSW, Sydney, Australia. North Carolina. J. Irrig. Drain. Eng. 132:600–608.
Department of Environmental Resources. 1999. Low-impact development: Kim, H.H., E.A. Seagren, and A.P. Davis. 2003. Engineered bioretention
An integrated design approach. Dep. of Environmental Resources, for removal of nitrate from storm water runoff. Water Environ. Res.
Prince George’s County, MD. 75:355–367.
Driver, N.E., and B.M. Troutman. 1989. Regression models for estimating Konrad, C.P., D.B. Booth, and S.J. Burges. 2005. Effects of urban
urban storm-runoff quality and quantity in the United States. J. development in the Puget Lowland, Washington, on interannual
Hydrol. 109:221–236. streamflow patterns: Consequences for channel form and
Duncan, H.P. 2006. Urban storm water quality. In T.H.F. Wong (ed.) streambed disturbance. Water Resour. Res. 41:W07009
Australian runoff quality. Available at http://www.arq.org.au) (verified 3 doi:10.1029/2005WR004097.
Apr. 2008). The Institution of Engineers Australia, Sydney, Australia. Ladson, A.R., S.D. Lloyd, C.J. Walsh, and T.D. Fletcher. 2007. Scenarios
Farm, C. 2002. Evaluation of the accumulation of sediment and heavy metals for redesigning an urban drainage system to reduce runoff frequency
in a storm-water detention pond. Water Sci. Technol. 45:105–112. and restore stream ecological condition. Water Pract. Technol. 2(2):
Fletcher, T.D., H.P. Duncan, P. Poelsma, and S.D. Lloyd. 2005. Storm water doi:10.2166/wpt.2007.0053.
flow and quality, and the effectiveness of non-proprietary storm water Lenat, D.R., and J.K. Crawford. 1994. Effects of land use on water
treatment measures- a review and gap analysis. CRCCH Rep. 04/08. quality and aquatic biota of three North Carolina Piedmont streams.
Cooperative Research Centre for Catchment Hydrology, Melbourne. Hydrobiologia 294:185–199.
Fletcher, T.D., G. Mitchell, A. Deletic, A.R. Ladson, and A. Séven. 2007. Leopold, L.B. 1968. Hydrology for urban land planning: A guidebook on the
Is storm water harvesting beneficial to urban waterway environmental hydrological effects of urban land use. Circ. No. 554. U.S. Geological
flows? Water Sci. Technol. 55:265–272. Survey, Washington, DC.
Fletcher, T.D., and A.C. Taylor. 2007. Estimating life-cycle costs of storm Lloyd, S.D., T.H.F. Wong, and C.J. Chesterfield. 2001a. Opportunities and
water treatment measures. Aust. J. Water Resour. 11:79–92. impediments to water sensitive urban design. p. 302–309. In Proc.
Gardner, E., G. Millar, C. Christiansen, A. Vieritz, and H. Chapman. 2006. 2nd. South Pacific Storm water Conf.: Rain the forgotten resource.
Energy and water use at a WSUD subdivision in Brisbane, Australia. Auckland, New Zealand.
Aust. J. Water Resour. 10:283–291. Lloyd, S.D., T.H.F. Wong, and C.J. Chesterfield. 2002a. Water sensitive urban
Grant, T., L. Ospray, A. Grant, A. Sharma, V.G. Mitchell, and F. Pamminger. design- A storm water management perspective. Industry Rep. No. 02/10.
2006. Life cycle aspects of alternative water and sewage servicing. p. Cooperative Research Centre for Catchment Hydrology, Melbourne.
121–128. In Proc. 7th Urban Drainage Modelling and 4th Water Lloyd, S.D., T.H.F. Wong, T. Liebig, and M. Becker. 2002b. Sediment
Sensitive Urban Design Conf. (Vol. 2), Melbourne, Australia. 3–7 Apr. characteristics in storm water pollution control ponds. Aust. J. Water
2006. International Water Association, London, UK. Resour. 5:137–145.
Grimmond, G.S.B., and T.R. Oke. 1991. An evaporation-interception model Lloyd, S.D., T.H.F. Wong, and B. Porter. 2001b. Implementing an

S-126 Journal of Environmental Quality • Volume 37 • September–October [Supplement] 2008


ecologically sustainable storm water drainage system in a residential River, Massachusetts. Hydrobiologia 83:29–42.
development. Water Sci. Technol. 45:1–7. Pratt, C.J., J.D.G. Mantle, and P.A. Schofield. 1995. UK research into the
Marsalek, J., and P.M. Marsalek. 1997. Characteristics of sediments from a performance of permeable pavement, reservoir structures in controlling
storm water management pond. Water Sci. Technol. 36:117–122. storm water discharge quantity and quality. Water Sci. Technol. 32:63–69.
McCarthy, D., V.G. Mitchell, A. Deletic, and C. Diaper. 2007. Urban storm Radcliffe, J. 2004. Water recycling in Australia; a review undertaken by
water Escherichia coli levels: Factors that influence them. Water Sci. the Australian Academy of Technological Sciences and Engineering.
Technol. 56:27–34. Australian Academy of Technological Sciences and Engineering, Parkville.
McIntosh, G.F., and S.J. Pugh. 1991. A preliminary assessment of the Read, J., T. Wevill, T.D. Fletcher, and A. Deletic. 2007. Variation among
reuse potential of effluent and urban storm water in South Australia. plant species in pollutant removal from storm water in biofiltration
Engineering and Water Supply Dep., Adelaide, South Australia. systems. Water Res. 42:893–902.
McRae, B. 2002. Managing Manly Beach from source to sea. Water, March Rose, S., and N.E. Peters. 2001. Effects of urbanization on streamflow in the
2002:99–101. Atlanta area (Georgia, USA): A comparative hydrological approach.
Meyer, J.L., M.J. Paul, and W.K. Taulbee. 2005. Stream ecosystem function Hydrol. Processes 15:1441–1457.
in urban landscapes. J. North Am. Benthol. Soc. 24:601–612. Roy, A.H., M.C. Freeman, B.J. Freeman, S.J. Wenger, W.E. Ensign, and J.L. Meyer.
Mitchell, V.G., H. Cleugh, G.S.B. Grimmond, and J. Xu. 2007. Linking 2005. Investigating hydrological alteration as a mechanism of fish assemblage
urban water balance and energy balance models to analyse urban design shifts in urbanizing streams. J. North Am. Benthol. Soc. 24:656–678.
options. Hydrol. Processes doi:10.1002/hyp.6868. Rozis, N., and A. Rahman. 2002. A simple method for life cycle cost
Mitchell, V.G., A. Deletic, T.D. Fletcher, and B.E. Hatt. 2007. Achieving assessment of water sensitive urban design. In Proc. 9ICUD. 8–13 Sept.
multiple benefits from storm water reuse. Water Sci. Technol. 55:135–144. 2002, Portland, OR.
Mitchell, V.G., B.E. Hatt, A. Deletic, T.D. Fletcher, D.T. McCarthy, and Serena, M., and V. Pettigrove. 2005. Relationship of sediment toxicants
M. Maygar. 2006. Integrated storm water treatment and harvesting: and water quality to the distribution of platypus populations in urban
Technical guidance report. ISWR Rep. 06/05. Inst. for Sustainable streams. J. North Am. Benthol. Soc. 24:679–689.
Water Resources, Monash Univ., Melbourne. Soranno, P.A., S.L. Hubler, S.R. Carpenter, and R.C. Lathrop. 1996.
Mitchell, V.G., T.A. McMahon, and R.G. Mein. 2003. Components of the total Phosphorus loads to surface waters: A simple model to account for
water balance of an urban catchment. Environ. Manage. 32:735–746. spatial pattern of land use. Ecol. Appl. 6:865–878.
Mitchell, V.G., A. Taylor, T.D. Fletcher, and A. Deletic. 2005. Storm Taylor, A.C., and T.D. Fletcher. 2006. ‘Triple-bottom-line’ assessment of
water reuse- Potable water substitution for Melbourne. ISWR Rep. urban storm water projects. Water Sci. Technol. 54:459–466.
No. 05/12. Inst. for Sustainable Water Resources, Monash Univ., Taylor, G.D., T.D. Fletcher, T.H.F. Wong, and P.F. Breen. 2005. Nitrogen
Melbourne, Australia. composition in urban runoff- Implications for storm water
Morgan, R.P., and S.F. Cushman. 2005. Urbanization effects on Maryland management. Water Res. 39:1982–1989.
fishes. J. North Am. Benthol. Soc. 24:643–655. Taylor, S.L., S.C. Roberts, C.J. Walsh, and B.E. Hatt. 2004. Catchment
Newton, D., G.A. Jenkins, and I. Philips. 2003. The potential of porous urbanisation and increased benthic algal biomass in streams: Linking
pavement for urban storm water management. In 28th Int. Hydrol. mechanisms to management. Freshwater Biol. 49:835–851.
Water Resour. Symp., Wollongong, Australia. 10–14 Nov. 2003. Taylor, A., and T.H.F. Wong. 2002. Non-structural storm water quality best
NHMRC and NRMMC. 2004. Australian drinking water guidelines. National management practices- Guidelines. CRCCH Report No. 03/14. Cooperative
Health and Medical Research Council and Natural Resource Management Research Centre for Catchment Hydrology, Melbourne, Australia.
Ministerial Council, Government of Australia, Canberra, Australia. Thomas, J.F., J. Gomboso, J.E. Oliver, and V.A. Ritchie. 1997. Wastewater
Niemczynowicz, J. 1996. Megacities from a water perspective. Water Int. re-use, storm water management, and the national water reform
21:198–205. agenda. Research Position Paper 1. CSIRO Land and Water, Canberra.
NRMMC and EPHC. 2006. Australian guidelines for water recycling: United Nations. 2008. State of the world populations 2007; unleashing
Managing health and environmental risks (Phase 1). Environmental the potential of urban growth. Available at http://www.unfpa.org/
Protection and Heritage Council and Natural Resources Management swp/2007/english/introduction.html (verified 8 May 2008).
Ministerial Council, Canberra, Australia. Vlachos, E., and B. Braga. 2001. The challenge of urban water management.
Oke, T.R. 1987. Boundary layer climates. Taylor and Francis, London, UK. In C. Maksimovic and J.A. Tejada-Guibert (ed.) Frontiers in urban
Olivera, F., and B.B. DeFee. 2007. Urbanization and its effect on runoff in the water management. UNESCO-IHP and IWA Publishing, London.
Whiteoak Bayou watershed, Texas. J. Am. Water Resour. Assoc. 43:170–182. Walsh, C.J. 2004. Protection of instream biodiversity from urban impacts: To
Page, D., P. Dillon, M. Purdie, and S. Rinck-Pfeiffer. 2006. A risk management minimize urban density or to improve drainage design? Mar. Freshw.
method for storm water reuse. p. 65–72. In Proc. 7th Urban Drainage Res. 55:317–326.
Modelling and 4th Water Sensitive Urban Design Conf. (Vol. 1), Melbourne, Walsh, C.J., T.D. Fletcher, and A.R. Ladson. 2005a. Stream restoration in urban
Australia. 3–7 Apr. 2006. International Water Assoc., London, UK. catchments through redesigning storm water systems: Looking to the
Pagotto, C., M. Legret, and P. Le Cloirec. 2000. Comparison of the hydraulic catchment to save the stream. J. North Am. Benthol. Soc. 24:690–705.
behaviour and quality of highway runoff water according to the type of Walsh, C.J., A.W. Leonard, A.R. Ladson, and T.D. Fletcher. 2004. Urban storm
pavement. Water Res. 34:4446–4454. water and the ecology of streams. Monash Univ., Melbourne, Australia.
Petterson, S.R., C.M. Davies, N.J. Ashbolt, V.G. Mitchell, G.D. Taylor, and Walsh, C.J., A.H. Roy, J.W. Feminella, P.D. Cottingham, P.M. Groffman, and
J. Lewis. 2007. Quantifying microbial health risks for non-potable R.P. Morgan. 2005b. The urban stream syndrome: Current knowledge
reuse of storm water. In REUSE07 AWA Nation. Water Reuse and and the search for a cure. J. North Am. Benthol. Soc. 24:706–723.
Recycling Conf., 16–18 July 2007, Sydney, Australia. Wang, L., J. Lyons, P. Kanehi, and R. Gatti. 1997. Influences of watershed
Po, M., J. Kaercher, and B.E. Nancarrow. 2004. Literature review of the factors land use on habitat quality and biotic integrity in Wisconsin Streams.
influencing public perceptions of water reuse. Australian Water Conserv. Fisheries 22:6–12.
and Reuse Res. Program, CSIRO and Australian Water Assoc., Canberra. WEF and ASCE. 1998. Urban runoff quality management New York, USA: Water
Poff, N., and J. Allan. 1995. Functional organisation of stream fish Environment Federation Manual of Practice No. 23 and American Society of
assemblages in relation to hydrological variability. Ecology 76:606–627. Civil Engineers Manual and Report on Engineering Practice No. 87.
Prato, T. 1999. Multiple attribute decision analysis for ecosystem Wong, T.H.F. (ed.). 2006. Australian runoff quality. Inst. of Engineers,
management. Ecol. Econ. 30:207–222. Australia, Sydney, Australia.
Pratt, C.J. 1999. Use of permeable, reservoir pavement constructions for storm Wong, T.H.F., S.D. Lloyd, and P.F. Breen. 2000. Water sensitive road
water treatment and storage for re-use. Water Sci. Technol. 39:145–151. design- Design options for improving storm water quality of road
Pratt, J.M., R.A. Coler, and P.J. Godfrey. 1981. Ecological effects of urban runoff. CRCCH Report No. 00/1. Cooperative Research Centre for
storm water runoff on benthic macroinvertebrates inhabiting the Green Catchment Hydrology, Melbourne.

Fletcher et al.: Reuse of Urban Runoff in Australia S-127

You might also like