You are on page 1of 10

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/318702726

Water Supply Infrastructure Planning: Decision-Making Framework to Classify


Multiple Uncertainties and Evaluate Flexible Design

Article  in  Journal of Water Resources Planning and Management · October 2017


DOI: 10.1061/(ASCE)WR.1943-5452.0000823

CITATIONS READS

11 855

6 authors, including:

Sarah Marie Fletcher Magdalena Klemun


Massachusetts Institute of Technology Massachusetts Institute of Technology
13 PUBLICATIONS   38 CITATIONS    6 PUBLICATIONS   40 CITATIONS   

SEE PROFILE SEE PROFILE

Kenneth M. Strzepek Afreen Siddiqi


Massachusetts Institute of Technology Massachusetts Institute of Technology
30 PUBLICATIONS   680 CITATIONS    64 PUBLICATIONS   1,108 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

EPA - Climate Change Impacts and Risks Analysis (CIRA) project View project

All content following this page was uploaded by Sarah Marie Fletcher on 09 October 2017.

The user has requested enhancement of the downloaded file.


Water Supply Infrastructure Planning: Decision-Making
Framework to Classify Multiple Uncertainties and
Evaluate Flexible Design
Sarah M. Fletcher, S.M.ASCE 1; Marco Miotti 2; Jaichander Swaminathan 3; Magdalena M. Klemun 4;
Kenneth Strzepek, A.M.ASCE 5; and Afreen Siddiqi 6
Downloaded from ascelibrary.org by Sarah Fletcher on 07/26/17. Copyright ASCE. For personal use only; all rights reserved.

Abstract: Urban planners face challenges in water infrastructure development decisions due to short-term variation in water availability and
demand, long-term uncertainty in climate and population growth, and differing perspectives on the value of water. This paper classifies these
multiple uncertainties and develops a decision framework that combines simulation for probabilistic uncertainty, scenario analysis for deep
uncertainty, and multistage decision analysis for uncertainties reduced over time with additional information. This framework is applied to a
case from Melbourne, Australia, where a drought from 1997 to 2009 prompted investment in a $5 billion desalination plant completed in 2012
after the drought ended. The results show opportunities for significant reduction in capital investment using flexible design. Building no
infrastructure is best in most simulations. However, in 10% of simulations, building no infrastructure leads to regret of greater than $10 billion
compared with a small, flexible desalination plant. Scenario analysis for deep uncertainties underlines the significant impact of assumptions
about the future and also on value judgments about the cost of water scarcity in evaluating infrastructure performance. DOI: 10.1061/(ASCE)
WR.1943-5452.0000823. © 2017 American Society of Civil Engineers.

Introduction systems in many industrialized countries are reaching the end of


their planned lifetimes, prompting further need for infrastructure
Urban water planners face the decision of how much water infra- investment.
structure to build in order to reliably meet demand for high-quality Most urban water supply infrastructure (such as distribution
water while minimizing cost to meet budget constraints. The chal- pipelines, treatment plants, and reservoirs) in Australia, the United
lenge of balancing the trade-off between shortage risk, cost, and States, and other industrialized countries was built between the
environmental protection is compounded by several critical uncer- 1930s and 1980s, before sophisticated risk and decision analysis
tainties. Urbanization is driving population growth in many cities at methods had been developed for practical use (Stakhiv 2011;
rapid but uncertain rates, natural variation in annual runoff is high Ferguson et al. 2013; Frost et al. 2016). Traditionally, planners de-
and expected to increase due to climate change in many regions veloped a long-term supply and demand forecast, and added a
(IPCC 2014), and energy price volatility and variable maintenance safety factor to account for uncertainty (Stakhiv 2011). This ap-
requirements drive cost risk for desalination and other energy- proach can lead to overdesign when demand ends up being lower
intense infrastructure options. Additionally, numerous water supply than forecasted. If demand exceeds the forecast, society can face
economic impacts due to unserved demand, environmental degra-
1 dation, and expensive measures for building additional infrastruc-
Ph.D. Candidate, Institute for Data, Systems, and Society,
Massachusetts Institute of Technology, 77 Massachusetts Ave.,
ture in short time frames.
Bldg. E18-407A, Cambridge, MA 02139-4307 (corresponding author). Since the 1980s, reliability, or outage frequency, has been em-
ORCID: https://orcid.org/0000-0003-3289-2237. E-mail: sfletch@mit.edu phasized as a risk-based metric for assessing water supply system
2 performance (Hashimoto et al. 1982). More recently, researchers
Ph.D. Candidate, Institute for Data, Systems, and Society,
Massachusetts Institute of Technology, 77 Massachusetts Ave., and planners have developed strategies to improve resilience and
Bldg. E18-407A, Cambridge, MA 02139-4307. robustness under uncertainty. Early work focused on the use of
3
Ph.D. Candidate, Dept. of Mechanical Engineering, Massachusetts adaptive management to operate water supply systems more flexibly
Institute of Technology, Room 3-17377 Massachusetts Ave., Cambridge, to ensure resilience under extreme operating conditions (National
MA 02139-4307.
4
Ph.D. Candidate, Institute for Data, Systems, and Society,
Research Council 2004; European Commission 2013). Adaptive
Massachusetts Institute of Technology, 77 Massachusetts Ave., management requires planners to change from a predict-then-act ap-
Bldg. E18-407A, Cambridge, MA 02139-4307. proach to a learn-then-adjust approach (Pahl-Wostl 2007). A related
5 strategy is flexible infrastructure design, which allows planners to
Professor Emeritus, Dept. of Civil, Environmental and Architectural
Engineering, Univ. of Colorado Boulder, 1111 Engineering Dr., 428 respond to future uncertainties. The use of real options, in which
UCB, Boulder, CO 80309-0428. infrastructure is designed so that it can be modified or expanded
6
Research Scientist, Institute for Data, Systems, and Society, in the future depending on how supply and demand evolve, provides
Massachusetts Institute of Technology, 77 Massachusetts Ave., flexibility that can reduce reliability and cost trade-offs in water
Bldg. E18-407A, Cambridge, MA 02139-4307.
supply systems (Wang and de Neufville 2006; de Neufville and
Note. This manuscript was submitted on October 5, 2016; approved on
April 17, 2017; published online on July 26, 2017. Discussion period open Scholtes 2011). The application of a real options approach, includ-
until December 26, 2017; separate discussions must be submitted for in- ing staged deployment, has been demonstrated in several recent
dividual papers. This paper is part of the Journal of Water Resources Plan- studies on design of water distribution systems (Marques et al.
ning and Management, © ASCE, ISSN 0733-9496. 2014), urban water systems (Deng et al. 2013; Gersonius et al. 2013;

© ASCE 04017061-1 J. Water Resour. Plann. Manage.

J. Water Resour. Plann. Manage., 2017, 143(10): 04017061


Urich and Rauch 2014), water treatment investments (Zhang and require minimum environmental flows to be met (Melbourne Water
Babovic 2012), and dam investments (Jeuland and Whittington 2015b); environmental flows ranged from 100 to 410 MCM be-
2014). Real options can be compared and evaluated using multiob- tween 1995 and 2011 (Low et al. 2015). Minimal groundwater
jective decision analysis to incorporate multiple planning goals and is used (Low et al. 2015). A single bulk wholesaler, Melbourne
evaluate flexible design options as uncertainties unfold over multi- Water, is responsible for harvesting, storage, and treatment.
ple planning stages (de Neufville and Scholtes 2011). After more than a decade of below average rainfall, reservoir
Recent methods have also been developed for new infrastructure storage for the city reached a record low of 25.6% of capacity in
planning that is robust to deep uncertainties such as climate change. 2009 (Low et al. 2015). This led to a range of demand management
The likelihoods of deep uncertainties cannot be accurately quanti- efforts including efficient appliance installations, reduced environ-
fied using historical data, rendering probability-based risk assess- mental flows, water restrictions for outdoor uses, a domestic rain-
ment challenging or inappropriate (Groves and Lempert 2007; water tank installation program, and treated wastewater recycling.
Milly et al. 2008). The goal is to generate and evaluate planning There were also two large infrastructure investments: the 150 MCM
strategies that are robust to a range of future outcomes. For exam- per year Wonthaggi reverse osmosis (RO) desalination plant and the
ple, robust decision making (RDM) uses Latin hypercube sam- 100 MCM=year maximum capacity Sugarloaf pipeline at capital
pling to generate many possible future scenarios across multiple costs of $5 billion and $550 million, respectively, able to provide
Downloaded from ascelibrary.org by Sarah Fletcher on 07/26/17. Copyright ASCE. For personal use only; all rights reserved.

uncertainties, assumes equal likelihood of each scenario, and se- up to approximately 40% of the city’s demand (Grant et al. 2013).
lects strategies that meet threshold performance criteria across a However, the drought ended before the desalination plant was
large percentage of scenarios (Lempert et al. 2006; Lempert and completed, leaving the plant unused for years. Several studies dis-
Groves 2010). Info-gap theory, in contrast, develops increasingly cuss the political pressure on Melbourne Water and the Victorian
large multidimensional uncertainty sets and identifies the solutions government, a detailed timeline of actions taken, and the institu-
that meet threshold performance criteria for each uncertainty set tional decision-making process in responding to the drought
(Ben-Haim 2001). Decision scaling links decision analysis with (Ferguson et al. 2013; Grant et al. 2013; Porter 2013; Low et al.
bottom-up climate vulnerability analysis, identifying climate-driven 2015).
action thresholds without relying on general circulation models This paper addresses the following question: given similarly un-
to generate climate scenarios (Brown et al. 2012). Such approaches certain and dire situations like those Melbourne faced in 2007 and
have been widely applied to problems in water resources planning in the inherent trade-off between cost, supply risk, and environmental
various countries (Pallottino et al. 2005; Dessai and Hulme 2007; protection, what approach can water planners use in evaluating
Groves et al. 2008; Kasprzyk et al. 2012; Moody and Brown 2012). infrastructure investments when facing multiple uncertainties of
Climate change and other deep uncertainties should be integrated different natures? A decision framework using classification and
into a water resources modeling framework that accounts for the incorporation of multiple uncertainties is applied in order to
full range of uncertainty planners face (Ray and Brown 2015). (1) evaluate and communicate the cost and water supply risk of
While scenario analysis has been demonstrated to be a powerful proposed infrastructure alternatives, and (2) identify the best infra-
planning tool for urban water managers facing deep uncertainties structure alternative based on a planner’s risk preferences across
and varying stakeholder concerns (Lienert et al. 2006), this does simulated water supply futures. A single best solution is not pre-
not preclude the use of probabilistic approaches for different, more sented because that would require value judgments and preferences
easily modeled uncertainties. Population growth, which is more ap- from stakeholders and planners. Rather, results are presented for a
propriate for statistical treatment, is expected to have a larger im- range of sample preferences to demonstrate the usefulness of the
pact on water resource systems to midcentury than climate change framework. The remainder of the paper is organized as follows.
(Vörösmarty et al. 2000; Schlosser et al. 2014; Sapkota et al. 2015; “Methods” describes the methods used: the uncertainty framework
Fant et al. 2016). The stationary statistical variation in rainfall, is presented first, followed by a description of its application to
which can be modeled stochastically, often drives uncertainty in the Melbourne. Key results and conclusions are presented in “Results”
short to medium term more than climate change (Lins and Stakhiv and “Conclusions”, respectively. More detailed descriptions of the
1998; Doczi 2014). methods and additional results are provided in the Supplemental
This paper builds on traditional decision analysis and deep Data.
uncertainty methods for water supply planning by developing a
framework to classify and model multiple uncertainties of different
natures. It enables water supply planners to quantify statistical un- Methods
certainties when appropriate, while the impact of different, deep
uncertainties on a risk profile can be assessed using scenario analy-
Decision Framework for Multiple Uncertainties
sis. It also identifies which uncertainties enable learning over time
and incorporates them into a multistage decision analysis model. Many uncertainties simultaneously impact water supply infra-
This model allows the evaluation of real options such as staged structure planning. These include annual reservoir inflows, popu-
deployment where planners learn as uncertainties unfold and utilize lation growth, energy prices, the cost of water supply shortages, and
flexible options if and when they are needed. climate change. A framework for decision making under multiple
The usefulness of the method to water planners is demonstrated uncertainties was developed, illustrated in Fig. 1.
through an application in Melbourne, Australia. The Millennium The first step in Fig. 1 is to classify uncertainties along two
Drought in Southeast Australia from 1997 to 2009 motivated this dimensions: whether they are deep or statistical, and whether there
work. Melbourne’s water supply system comprises a network of 10 is high or low potential for learning over time. Deep uncertainties
storage reservoirs with a total of 1,812 million m3 (MCM) of stor- are those for which the likelihood of the outcomes cannot be de-
age. Net average annual inflows (evaporation and losses subtracted) termined; they are modeled using scenario analysis in order to
were 571 MCM between 1926 and 2014 and demand was avoid estimating likelihood. In this context, scenario analysis refers
401 MCM in 2014–2015 (Melbourne Water 2015b). This results to varying the value of uncertain parameters in order to understand
in a storage to annual runoff ratio of 3.2, which indicates a highly the impact of their uncertainty on the results. While scenarios are
managed system. Catchment-level streamflow management plans not explicitly probabilistic, the choice of values does assign some

© ASCE 04017061-2 J. Water Resour. Plann. Manage.

J. Water Resour. Plann. Manage., 2017, 143(10): 04017061


uncertainties. Statistical uncertainties with high learning potential
are incorporated into the decision analysis model explicitly so that
flexible options can be exercised in response to new information.
The decision model ranks the infrastructure alternatives according
to an objective function for each simulation run, resulting in a dis-
tribution of best infrastructure alternatives across simulated varia-
bles. Scenario analysis is then used to repeat the simulation process
for different sets of input variables representing deep uncertainties
to assess the impact of scenarios on the risk profile.
In the third step shown in Fig. 1, these results are visualized.
Several visualization methods are used including bagplots, which
show a two-dimensional (2D) representation of the risk profile
Fig. 1. Method for decision making under multiple uncertainties; against two key performance metrics, cost and water shortages.
Step 1: uncertainties are classified according to whether they are deep Risk ranking is also used. The goal is not to recommend a single
or statistical and whether they have high or low potential for learning as best solution as in traditional decision analysis, but rather to show a
Downloaded from ascelibrary.org by Sarah Fletcher on 07/26/17. Copyright ASCE. For personal use only; all rights reserved.

uncertainties are realized over time; Step 2: water system and decision risk profile for each alternative based on how it performs under
models evaluate infrastructure alternatives under these uncertainties; many different conditions drawn from simulation and scenarios.
Step 3: performance indicators are displayed in a risk profile for de- Because stakeholder values are not elicited and incorporated in this
cision support study, the identification of a single solution is not feasible. This is
consistent with decision support tools that offer technical analysis
and identify the effects of value judgments on the results but do not
implicit probability to those outcomes; this is important to recog- recommend a single choice to planners (Benayoun et al. 1971;
nize in all deep uncertainty methods. Statistical uncertainties are Loucks 2000; Brugnach et al. 2007).
those that can be estimated using data-driven probability distribu-
tions. Statistical uncertainties are modeled using Monte Carlo sim- Application to Urban Water Supply Planning in
ulation in order to obtain the most precise risk profile. The high Melbourne, Australia
versus low learning potential indicates whether, as the uncertainty
is realized over time, additional observations meaningfully update This analysis framework is applied to an illustrative example from
the planner’s expectations of the future. Based on updated expect- Melbourne, Australia, where planners in 2007 decide what, if any,
ations, planners decide to exercise or ignore flexible options. These additional supply investments should be made over a 30-year plan-
uncertainties are therefore incorporated directly into a multistage ning period (the approximate lifetime of a RO plant). Key uncer-
decision analysis model (de Neufville and Scholtes 2011), in which tainties are identified and classified. Infrastructure alternatives are
flexible options are optimally exercised or ignored. This allows the chosen based on those considered by Melbourne’s water planners
full value of flexible options to be modeled. Uncertainties that do in 2007, with additional flexible alternatives designed for com-
not yield meaningful updates on expected payoffs as observations parison. A model of Melbourne’s water system is developed using
are made over time can be analyzed more efficiently by varying the a simple hydrological model and demand forecasts. A decision
inputs across model runs. analysis model is then used to evaluate the infrastructure alterna-
This uncertainty classification explicitly incorporates different tives, including some with flexible expansion options, based on an
objective function that considers lifetime costs and water shortages.
levels of uncertainty (Walker et al. 2003) and varying potentials
for learning over time. The level dimension can also incorporate
ambiguity, or uncertainties that arise through differences in stake- Uncertainty Classification
holder perspectives (Brugnach et al. 2008; Kwakkel et al. 2010), Five uncertainties, listed in Table 1, are included in this analysis.
such as value judgments about the cost of water shortages. Uncer- They were chosen because of their high degree of uncertainty and
tainties arising from ambiguity are similar to deep uncertainties in potential for impact on Melbourne’s planning decisions; however,
that they are not appropriate for probabilistic analysis, so scenario they are not comprehensive. Certain uncertainties were excluded,
analysis is used instead. Different objective functions can be used to such as the potential for desalination technology costs to decline over
represent different preferences or risk profiles. For instance, a risk- time, because initial analysis of historical data suggested they were
averse planner may prefer to invest in water supply infrastructure unlikely to impact planning decisions on a 30-year timeline. Other
that often sits idle as an insurance policy against drought that min- uncertainties, such as climate change, future agricultural production,
imizes shortage risk. While there is a large amount of literature on policy changes, and structural and observation uncertainty, could be
eliciting and incorporating stakeholder values into decision analysis included in future work. The included uncertainties are classified as
(Renn et al. 1993; Gregory 2000; Hämäläinen et al. 2001), such indicated in Table 1 with brief justifications. Decisions to classify
analysis is not included in this paper. Rather, a sample objective uncertainties using the framework presented in Fig. 1 ultimately rely
function and a few scenarios comparing alternative objective on analyst judgment but are informed by analyses of available histori-
weights are used to demonstrate the method. Additionally, other di- cal data and forecasts. The analysis was based on historical inflow data
mensions of uncertainty, such as the location of uncertainty within a (Melbourne Water 2015b), historical and forecasted population
decision-making process, structural uncertainty, and observational growth (Australian Bureau of Statistics 2014), and historical and fore-
uncertainty, are not addressed here. casted electricity price (ACILTasman 2011). Further details are avail-
In the second step in Fig. 1, a model of the water system is com- able in methodology sections A–C of the Supplemental Data.
bined with the decision model to evaluate and compare infrastruc-
ture alternatives. These infrastructure alternatives can be generated
Infrastructure Alternatives
through a screening model or planner input. The water system
model is simulated to generate the distribution of cost and water Six infrastructure alternatives comprised of combinations of three
supply risk for each infrastructure alterative over all statistical projects—a 150 MCM=year RO plant based on the Wonthaggi

© ASCE 04017061-3 J. Water Resour. Plann. Manage.

J. Water Resour. Plann. Manage., 2017, 143(10): 04017061


Table 1. Methods Used for Key Uncertainties in Analysis Based on Uncertainty Classification Framework
Uncertainty Category Justification
Reservoir inflows Statistical; low learning Well characterized by 80 years of historical data; relatively stationary over 30-year planning
period
Population growth rate Statistical; high learning Well characterized by existing demographic models and forecasts in Melbourne; early
changes in growth rate predictive of future growth
Electricity price Deep uncertainty Influenced by deeply uncertain factors such as national and international policy and markets
Water shortage penalty value Deep uncertainty This is an ambiguity, or uncertainty arising from differences in stakeholder perspective
Demand per capita Deep uncertainty Influenced by deeply uncertain factors such as policy, citizen behavior, and technology
adoption

Table 2. Definitions for the Six Infrastructure Alternatives Evaluated and Compared in the Decision-Modeling Framework
Capital expenditure
Infrastructure alternatives (millions of dollars) Capacity (MCM=year)
Downloaded from ascelibrary.org by Sarah Fletcher on 07/26/17. Copyright ASCE. For personal use only; all rights reserved.

S1: No build 0 0
S2: Pipeline and irrigation upgrade 1,002 Variable: maximum 100
S3: Small RO plant with expansion option 2,045 [þ1,095] Firm: 75 [þ75]
S4: Large RO plant 2,900 Firm: 150
S5: Small RO plant with expansion option; pipeline and irrigation upgrade 3,047 [þ1,095] Firm: 75 [þ75] variable: maximum 80
S6: Large RO plant; pipeline and irrigation upgrade 3,902 Firm: 150 variable: maximum 80
Note: S3 and S5 were developed to compare a staged deployment approach to the full, upfront deployment approach used in S4 and S6.

plant, a 100 MCM=year capacity pipeline and accompanying irri- both the existing reservoir system and a new supply infrastructure.
gation system upgrade based on the Sugarloaf pipeline and accom- The model estimates the annual cost and water shortages (i.e., un-
panying upgrades, and a 75 MCM=year RO plant designed with a served demand) over a 30-year period starting in 2007 for each in-
flexible option to expand to 150 MCM=year if desired—are de- frastructure alternative. Monte Carlo simulation is used to develop
signed and evaluated. The smaller (75 MCM=year) desalination distributions for these estimates based on uncertainty in inflows and
plant was not considered by Melbourne Water in 2007; it was in- population growth.
cluded here as an alternative to assess the value of staged deploy- The main components of the water balance are net reservoir in-
ment. The flexible design requires the plant site to be sized to fit flows (inflows minus evaporation), demand from end users, water
twice as many membrane modules as needed before the expansion, imported from new infrastructure, and environmental outflows. To
incurring additional capital costs upfront in exchange for the option model future inflows, 100,000 synthetic annual 30-year inflows are
to expand cheaply and quickly later. The no-build alternative (S1 in generated using an autoregressive moving average (ARMA) time-
Table 2) is included as a baseline option where no new infrastruc- series model fit to historical inflow data dating back to 1926. This
ture is developed. approach captures the autocorrelation observed in runoff. It as-
The desalination plants incur both high capital costs and high sumes inflows to be a stationary stochastic process over the 30-year
operating costs. The operating costs are high due to the energy planning period. Scenario analysis is used to vary the mean and
intensity of seawater desalination and membrane replacement variance of this process based on estimates of climate change im-
requirements. In addition to substantially lower capital costs, the pacts on runoff in Australia by Strzepek et al. (2013). This is a sim-
pipeline and irrigation upgrade also incurs much lower operating ple approach; climate change is not a focus of this paper. Future
costs; pumping is the largest component. The desalination plants, work evaluating longer planning periods could use more sophisti-
however, provide firm capacity; they can reliably provide the maxi- cated statistical approaches such as downscaling from general cir-
mum design capacity during a dry year. The pipeline system is culation models. Annual water demand is modeled as the product
market-based: farmers whose irrigation systems have been up- of population and demand per capita. Population projections from
graded to be more efficient sell excess water to the utility at their the Australian Bureau of Statistics for the city of Melbourne are
discretion. Farmers are less likely to provide water during dry used, ranging from 50,000 to 150,000 people per year (Australian
years, so it is unlikely that the full 100 MCM=year is available each Bureau of Statistics 2014). The base case demand per capita of
year. This dynamic is approximated by assuming a correlation be- 100 kL=person=year is based on historical demand data and is var-
tween inflows to Melbourne’s reservoir system and water available ied using scenario analysis to assess the effect of demand-reduction
from the pipeline, with 80 MCM available during average wet years measures; future work could assess and compare demand-side al-
and no water available during dry years. Cost estimates (capital ternatives directly. Water is imported from any new infrastructure
expenditures, fixed operating costs, and variable operating costs) capacity if reservoir storage levels go below one of two thresh-
are developed for each infrastructure alternative using cost data olds. Pipeline water is imported below an intervention threshold of
from the Sugarloaf pipeline project, cost information on the Won- 980 MCM and desalination water is imported below an emergency
thaggi plant, cost data from comparable RO plants, and input from threshold of 580 MCM, the maximum allowable drawdown. These
desalination experts. Details are available in “Infrastructure Cost thresholds are set by Melbourne Water (2015c) and used in other
Estimates” of the Supplemental Data. studies of Melbourne’s water system (Turner et al. 2014). In this
paper, water shortage is defined as water demand that cannot be
met without reducing demand or withdrawing below the emergency
Water System Model
threshold. The operational rules used to model the import and out-
A water system model is developed using a water balance approach flow decisions are illustrated in Fig. S2. The water system model
to model reservoir storage and water supplied to end users from aggregates the individual reservoirs and uses an annual time step.

© ASCE 04017061-4 J. Water Resour. Plann. Manage.

J. Water Resour. Plann. Manage., 2017, 143(10): 04017061


This approach is appropriate to demonstrate the new method devel- across many realizations of the simulated uncertain variables, res-
oped in this paper and given the capacity-expansion focus of the ervoir inflows and population growth, are presented. Further details
decision model (Baker et al. 2014) and the high intra-annual stor- on the decision analysis model are available in “Decision Model”
age in the system. More details on the water system model are avail- of the Supplemental Data.
able in “Infrastructure Cost Estimates” of the Supplemental Data.

Results
Decision Analysis Model
The decision model uses the estimates for cost and water shortages First results showing the impact of the statistical uncertainties, res-
from the water system model to evaluate the six infrastructure al- ervoir inflows and population growth, are presented in “Statistical
ternatives (Table 1) over a 30-year planning period. For each syn- Uncertainties.” This includes results from the water system model
thetic 30-year inflow series, the model ranks the six infrastructure showing the cost and water shortage risk of each infrastructure op-
alternatives using multistage decision analysis, which is frequently tion (Fig. 2) as well as results of the decision model (Fig. 3). These
used to evaluate real options (de Neufville and Scholtes 2011). The results are shown as distributions because Monte Carlo simulation
model can be understood as a decision tree, in which the population is used for statistical uncertainties. Then, results showing the effect
of deep uncertainties using scenario analysis are presented in “Deep
Downloaded from ascelibrary.org by Sarah Fletcher on 07/26/17. Copyright ASCE. For personal use only; all rights reserved.

growth rate has a probability of going up or down in each 10-year


planning stage. In the two infrastructure alternatives with flexible Uncertainties.” Fig. 4 shows how the distribution-based results ob-
desalination design, S3 and S5 (Table 2), the planner can react to tained from addressing the statistical uncertainties are shifted when
change in population growth by deciding to exercise the real option scenario-based deep uncertainties are also addressed.
of expanding a small desalination plant. Fig. S4 gives a tree sche-
matic of the decision model. The model chooses the infrastructure Statistical Uncertainties
alternative and the timing of the desalination plant expansion, Fig. 2 summarizes the model results for cumulative water shor-
where applicable, to minimize the following objective function: tages and infrastructure cost for each of the infrastructure alterna-
E½Total Costi  ¼ ENPV½Costi þ Penalty × Water Shortagei  tives across 100,000 water supply simulations using bagplots
(Rousseeuw et al. 1999). The bagplot is an analog to the boxplot
where E = expected value; i = choice of infrastructure alternative for bivariate data (Rousseeuw et al. 1999). For each infrastructure
and expansion strategy; Cost = total of capital costs and operational alternative except no-build, the center point marked with an asterisk
expenditures incurred from the infrastructure alternative over the is the Tukey median (Donoho and Gasko 1992); the inner, dark-
planning period; Water Shortage = total volume of water shortages colored shape is the bag, which contains 50% of the data, similar
from years where water demand without conservation or environ-
mental flow reductions exceeds available water supply, also inter-
preted as water supply vulnerability (Hashimoto et al. 1982);
ENPV = expected net present value; and Penalty = cost incurred
by the decision maker for each MCM of water shortage during
the year of the shortage. A monetary penalty for shortages does
not exist in reality in Melbourne or most municipal water utilities.
Instead, the cost of water shortages is borne by society as economic
damages. The penalty value can therefore be interpreted as the
planner’s degree of risk aversion to water supply vulnerability.
A base case value of $25 million=MCM is chosen to be approxi-
mately 50 times higher than the bulk usage price of desalinated
water ($0.55=m3 ) in Melbourne (Melbourne Water 2015a) and
consistent with previous work (Jenkins et al. 2003; Rawls and
Turnquist 2012). Using scenario analysis, different penalty values
up to $250 million=MCM are evaluated, consistent with prices in
previous work (Liu et al. 2016). The ENPV is calculated using a
discount rate of 7%, which is in the middle of the range typically
used by government agencies to evaluate projects (Harrison 2010).
This representative objective function is intended to demonstrate
the method and does not reflect real stakeholder preferences. Use
of this method in future planning could incorporate stakeholder in-
put on the value of the penalty through a collaborative stakeholder
engagement process.
Finally, after the decision model ranks the infrastructure alter-
natives based on the expected value of total cost, Monte Carlo sim-
Fig. 2. Bagplots visualize the distribution of infrastructure cost and
ulation is used to randomly choose a single population growth path
total water shortage across a random sample of 1,000 total water supply
for each model run. This shows, across many synthetic inflow and
simulations for each infrastructure alternative; because the no-build
population growth time series, how the infrastructure alternative
alternative always incurs zero cost and is therefore univariate, the plot
chosen based on expected value performs against simulated actual
shows the median shortage and a line that spans the range of shortage
conditions. For example, the decision model might choose to build
between the 0th and 99th percentiles, which is analogous to the convex
no additional infrastructure because the expected shortages are low;
hull in the bagplots; building desalination capacity increases cost while
however, there is a small probability that high population growth
mitigating risk of high-magnitude shortages; flexible design can miti-
could lead the actual shortages to be high. Using this approach,
gate shortage risk at lower cost
results that show the performance of each infrastructure alternative

© ASCE 04017061-5 J. Water Resour. Plann. Manage.

J. Water Resour. Plann. Manage., 2017, 143(10): 04017061


Ranks of Simulated Payoffs by Infrastructure Alternative to the inner section of a boxplot; and the outer, light-colored shape
50% is the convex hull that encompasses the rest of the data, excluding
outliers. These plots show the distribution of cost and water supply
risk faced as a result of the statistical uncertainties addressed, res-
40% ervoir inflows and population growth. The no-build alternative
costs nothing but incurs the greatest shortage risk, with the median
Frequency of Ranking

Worst Payoff
total shortage spanning the 30-year period at 260 MCM. Cumula-
30% 5th Best tive shortages greater than 2,000 MCM (over the 30-year period)
4th Best occur in 20% of simulation runs (Fig. S6). The pipeline alternative
3rd Best (S2) consistently imposes a cost of close to $1 billion and decreases
20% 2nd Best the shortage risk. The four infrastructure alternatives that include a
Best Payoff desalination plant all increase the average cost and cost variability
while decreasing the shortage risk further. The median water short-
10% age for all four alternatives with desalination is 0, with 90th per-
centile shortages all on the order of 1,000 MCM over the 30-year
Downloaded from ascelibrary.org by Sarah Fletcher on 07/26/17. Copyright ASCE. For personal use only; all rights reserved.

analysis period. Within these four alternatives, those that have more
0 built infrastructure capacity see modest reductions in water shortage
risk but significant increases in cost. Most notably, the alternative of
nt rig g

w h O on
n
Irr

t
Irr
p l la n
n

io
hi

&

t&
w at

pt

the small plant with expansion option (S3) has a very similar water
ot

an
rg e
N

io

La arg
r

pt

shortage risk profile to the large plant (S4), as shown by the width of
I
o

it
D

/O

the inner bag, with median cost approximately $1 billion lower,


al Pla
nt

again demonstrating the value of staged deployment under uncertain


la
Sm all
lP
Sm

demand. The same relationship is true of the infrastructure alterna-


tives with desalination and pipeline. This value is also demonstrated
Fig. 3. For each of 100,000 simulations, the simulated total payoff for
by the variation in utilization of the expansion option: the option is
each of the six infrastructure alternatives is calculated and used to rank
exercised in approximately 45% of simulations in the small plant
the alternatives from best to worst; doing nothing is best in more than
without pipeline alternative (S3) and approximately 40% of simu-
half the simulations but also worst in more than 30% of simulations lations in the small plant with pipeline alternative (S5).
Across all of the infrastructure alternatives, more than 80% of
the years simulated incurred no water shortage at all. Large annual
water shortages of >100 MCM were more common than smaller
shortages. These large annual shortages are concentrated in the sec-
ond half of the 30-year period because population growth leads to
higher demand in later years. This has important implications for
the capacity factor of the built infrastructure as well as for timing
the decision to build new infrastructure. The median number of
years that the built infrastructure is used is between 4 and 6 years
across all the infrastructure alternatives. This means that any infra-
structure that is built is likely to be used for 20% or less of its
lifetime. This is because the total water shortage incurred over
the 30-year planning period is typically concentrated in a small
number of years that incurred extreme lows in runoff coupled with
high demand.
The decision model evaluates the six infrastructure alternatives
based on the overall cost, which includes both infrastructure cost
and penalties for water shortages over a 30-year period. In addition
to choosing a best alternative based on expected values, it also cal-
culates the simulated payoff for each alternative in each model run,
and ranks the alternatives according to this simulated payoff. A
(a) (b) (c) simple way to communicate and compare the infrastructure alter-
natives is to compare the rankings, shown in Fig. 3 for 100,000
simulations of the base case. The no-build alternative (S1) is the
best-performing infrastructure alternative in more than 50% of
cases; however, it is the worst-performing alternative in more than
30% of cases. This underlines the risk of planning based on a single
Fig. 4. Scenario analysis on the distribution of the best infrastructure forecast. At the other extreme, the large plant plus pipeline alter-
alternative chosen by the decision model across 100,000 model runs; in
native (S6) performs the worst in more than 60% of cases and rarely
each model run, the best alternative is chosen based on the expected
performs within the top half of rankings. An interesting result is the
payoff; base cases are indicated with an asterisk; more details are avail-
flexible small desalination plant (S3): despite performing the best in
able in “Decision Model” of the Supplemental Data; in the base case,
only approximately 20% of simulations, this alternative performs in
no-build (S1) and small plant with option (S3) are each chosen in ap-
the top half of rankings in nearly 90% of simulations and never
proximately 40% of runs; results are highly sensitive to demand per
ends up as worst or second to last. This demonstrates the ability
capita and the penalty value and relatively insensitive to electricity
of paying a small premium for flexibility to mitigate downside risk
growth rate
significantly.

© ASCE 04017061-6 J. Water Resour. Plann. Manage.

J. Water Resour. Plann. Manage., 2017, 143(10): 04017061


The total costs including shortage penalties incurred by each including electricity price growth and climate change, is available
infrastructure alternative across the 100,000 inflow simulations in “Additional Results” in the Supplemental Data.
were also analyzed. The no-build alternative is the lowest cost in
50% of runs, with more than 40% of runs incurring zero total cost,
indicating no water shortages over 30 years. However, the down- Conclusions
side risk is significant; in 10% of runs the total cost of the no-build
alternative is more than $10 billion greater than the cheapest alter- The primary contribution of this paper is the development of a
native. The pipeline alternative (S2) performs similarly to the no- method that can be used to inform water supply planning decisions
build alternative, with approximately $1 billion higher cost in the under multiple uncertainties. A classification framework was devel-
low-cost half of simulations and lower cost in the high-cost half of oped for uncertainties that enables the use of probability-based risk
simulations. This is due to the effect of the market-based system for quantification and information gathering where appropriate, while
water sales through the pipeline, demonstrating the limits of water leaving deep uncertainties to be addressed using scenario analysis.
markets to provide firm capacity when there is high correlation be- This approach allows evaluation and comparison of infrastructure
tween drought in the two interconnected regions. The four alterna- alternatives, including those with flexible design, as well as
tives that include desalination have high capital costs, but in more demand-reduction strategies in a way that takes into account the
Downloaded from ascelibrary.org by Sarah Fletcher on 07/26/17. Copyright ASCE. For personal use only; all rights reserved.

than 30% of simulations they incur cheaper overall costs than the diverse and multifaceted uncertainties water planners face. The
no-build or pipeline alternatives. The small plant (S3) provides a use of simple yet information-dense visualizations of risk profiles
midrange capital cost alternative that significantly mitigates down- can provide decision support for policy makers.
side risk through the use of a flexible option. Fig. S5 gives more The results demonstrate that moderate investment increases, to-
details. gether with flexible infrastructure design, can mitigate water short-
age risk significantly. Using a staged deployment strategy to expand
if necessary when more information is available reduces shortage
Deep Uncertainties risk at lower cost compared with other, less-flexible infrastructure
So far, results have incorporated uncertainties arising from inflow options. The value of this flexibility is dependent on the nature and
variability and population growth, the two statistical uncertainties magnitude of uncertainties in the system. Furthermore, demand-
addressed. Next, scenario analysis is used to show the effects of the reduction strategies may reduce the need for additional capacity.
deep uncertainties addressed: demand per capita, the shortage pen- These results also suggest that supply infrastructure can be regarded
alty value, and electricity price growth (Table 1). Each bar in Fig. 4 as an insurance policy against drought. Although the investments
shows how often the decision model selected each infrastructure may be used infrequently over their lifetime, they mitigate the risk
alternative as the best across a set of 100,000 simulations. In of severe economic consequences due to water shortages. This view
the base case, indicated using an asterisk, the model chooses both can allow for improved risk assessment along with appropriate val-
the no-build alternative and the small plant (S3) approximately 40% uation and expectations of its utility. However, there are trade-offs
of the time each. The results are highly sensitive to the penalty fac- between infrastructure cost and water shortage risk.
tor, shown in Fig. 4(a). The no-build alternative performs best The Victorian government’s decision to build a 150 MCM=year
in more than 95% of simulations when the penalty is reduced to desalination plant, one of the world’s largest, has been the subject
$5 million=MCM. Likewise, increasing the penalty makes alterna- of heated debate and political backlash given the wealth of reservoir
tives with greater capacity more favorable. Increasing the penalty to water available for years after the plant came online in 2012 (Porter
$50 million=MCM makes the no-build alternative best in approx- 2013; Ferguson et al. 2014). However, as drought impacts the re-
imately 25% of simulations, down from 40%, and the small plant gion again and the population continues to grow rapidly, the plant
with option plus pipeline best in 20% of simulations, up from ap- delivered its first water order in March 2017 (Gordon 2017), dem-
proximately 5%. Increasing the penalty by a full order of magni- onstrating that desalination capacity can play a role in mitigating
tude to $250 million=MCM, consistent with the values in Liu et al. supply and demand uncertainty even if it is used infrequently.
(2016), increases the magnitude of this shift. The no-build alterna- These insights are relevant to other places facing infrastruc-
tive is chosen as best in less than 5% of simulations and the small ture decisions driven by multiyear droughts. For example, several
plant with option plus pipeline (S5) is chosen in approximately 60% municipalities in California are considering desalination invest-
of simulations. The small plant with option plus pipeline dominates ments (Gillis 2015). Although drought in California is often framed
the distribution of best options over the large plant (S4). Both alter- as the new normal because of climate change, it is important for
natives have 150 MCM of capacity without the expansion in S5; both planners and the public to remember that high variability and
however, half that capacity in S5 is variable rather than firm. This extended periods of low rainfall are normal even in the absence
suggests that the value of flexibility from the expansion option out- of climate change. Using a learning-driven staged deployment ap-
weighs the variability in some of the capacity in S5. proach or increasing demand-reduction efforts may reduce the need
The model results are also highly sensitive to the demand per for capital investment. Alternatively, framing any investments as
capita, shown in Fig. 4(b). Lowering the demand to 80 kL per capita drought insurance that may be used infrequently may increase pub-
per year substantially reduces the need for additional infrastructure, lic acceptance. This strategy was successfully employed in the
with the no-build alternative performing best in more than 95% of United Kingdom, where Thames Water built a large desalination
simulations. The high sensitivity to demand per capita suggests that plant in east London in 2010 citing risks of severe water shortages
demand-side conservation measures may be able to significantly re- (Jowit 2010).
duce the need for supply infrastructure investments, especially if The analysis of the Melbourne case shows that much of the
demand reductions are firmly available during dry years. This dem- choice of the best answer is predicated not only on assumptions
onstrates the profound potential of demand management and points about the city’s future but also value judgments about the value of
to the need for more rigorous analysis of infrastructure expansion water during times of scarcity and society’s appetite for risk. The
decisions in conjunction with demand management options. Results sensitivity to the penalty value suggests that working with stake-
are relatively insensitive to the electricity price growth rate, shown holders to choose a value that reflects society’s tolerance to risk
in Fig. 4(c). Additional scenario analysis on model parameters, will be important in applying this method to prospective planning

© ASCE 04017061-7 J. Water Resour. Plann. Manage.

J. Water Resour. Plann. Manage., 2017, 143(10): 04017061


decisions. Future work could incorporate economic methods to Ben-Haim, Y. (2001). Info-gap decision theory: Decisions under severe
value the effects of water shortages on society as a way to inform uncertainty, Elsevier, Oxford, U.K.
the choice of penalty. Modeling separate end uses—agriculture, Bhaduri, A., et al. (2016). “Achieving sustainable development goals from a
municipal, industrial, and environmental—could be used to assess water perspective.” Front. Environ. Sci., 4, 64.
the impact of risk sharing across multiple end-use sectors. Alter- Brown, C., Ghile, Y., Laverty, M., and Li, K. (2012). “Decision scaling:
Linking bottom-up vulnerability analysis with climate projections in
native planning strategies can also be analyzed. For example,
the water sector.” Water Resour. Res., 48(9), 1–12.
the water provider may invest in a water fund (at a much lower
Brugnach, M., et al. (2007). “Uncertainty matters: Computer models at the
cost than committing capital for new infrastructure development) science-policy interface.” Water Resour. Manage., 21(7), 1075–1090.
that allows the provider to subsidize costs incurred by users in years Brugnach, M., Dewulf, A., Pahl-Wostl, C., and Taillieu, T. (2008). “Toward
of drought. Additionally, the use of financial risk mitigation op- a relational concept of uncertainty: About knowing too little, knowing
tions, such as the private-public partnership that Melbourne Water too differently, and accepting not to know.” Ecol. Soc., 13(2), 30.
used when contracting the Wonthaggi plant, may further reduce the de Neufville, R., and Scholtes, S. (2011). Flexibility in engineering design,
need for capital-intensive projects. MIT Press, Cambridge, MA.
Future work can extend this study by incorporating addi- Deng, Y., Cardin, M.-A., Babovic, V., Santhanakrishnan, D., Schmitter, P.,
tional uncertainties and comparing the nature of key uncertainties and Meshgi, A. (2013). “Valuing flexibilities in the design of urban
Downloaded from ascelibrary.org by Sarah Fletcher on 07/26/17. Copyright ASCE. For personal use only; all rights reserved.

in different regions. For example, while population growth pro- water management systems.” Water Res., 47(20), 7162–7174.
vides significant learning opportunities for planners in Melbourne, Dessai, S., and Hulme, M. (2007). “Assessing the robustness of adaptation
Australia, there may be different uncertainties that drive staged decisions to climate change uncertainties: A case study on water resour-
deployment decisions in other regions. Future work could also ex- ces management in the East of England.” Global Environ. Change,
17(1), 59–72.
plore the effect of different water conservation strategies on the ne-
Doczi, J. (2014). “Managing climate risk for the water sector with tools and
cessity for water supply capacity expansion or sustainability goals
decision support.” Climate change and water resources, S. Shrestha,
and objectives (Bhaduri et al. 2016). Some demand-reduction strat- M. S. Babel, and V. P. Pandey, eds., CRC Press, Boca Raton, FL,
egies such as fines or public education may yield uncertain and 239–290.
variable reductions rather than the firm capacity provided by desali- Donoho, D., and Gasko, M. (1992). “Breakdown properties of location
nation. Future work could also include more detailed models of estimates based on halfspace depth and projected outlyingness.”
operational reservoir management to assess the impact of improved Ann. Stat., 20(4), 1803–1827.
operations as an alternative to or in conjunction with infrastructure European Commission. (2013). “Adapting infrastructure to climate
additions. Additionally, new strategies such as fit-for-purpose water change.” Brussels, Belgium, 1–37.
supply, in which water of different quality is used for applications Fant, C., Schlosser, C. A., Gao, X., Strzepek, K., and Reilly, J. (2016).
for which the quality level is adequate (Nair et al. 2014), can be “Projections of water stress based on an ensemble of socioeconomic
evaluated for effects on infrastructure scale and design decisions. growth and climate change scenarios: A case study in Asia.” PLoS
One, 11(3), e0150633.
Ferguson, B. C., Brown, R. R., Frantzeskaki, N., de Haan, F. J., and Deletic,
Acknowledgments A. (2013). “The enabling institutional context for integrated water man-
agement: Lessons from Melbourne.” Water Res., 47(20), 7300–7314.
We thank Professor Hector Malano and Dr. Meenakshi Arora at the Ferguson, B. C., Brown, R. R., Frantzeskaki, N., de Haan, F. J., and Deletic,
University of Melbourne for discussions and assistance in identify- A. (2014). “The enabling institutional context for integrated water man-
ing initial data sources for this work. We are grateful to Professor agement: Lessons from Melbourne.” Water Res., 47(20), 7300–7314.
Richard Larson at Massachusetts Institute of Technology (MIT) for Frost, L., et al. (2016). “Water, history and the Australian city: Urbanism,
suburbanism and watering a dry continent, 1788-2015.” Cooperative
insights that shaped the early stages of this work and to three anon-
Research Centre for Water Sensitive Cities, Melbourne, Australia.
ymous reviewers whose thoughtful comments greatly improved this
Gersonius, B., Ashley, R., Pathirana, A., and Zevenbergen, C. (2013). “Cli-
paper. mate change uncertainty: Building flexibility into water and flood risk
infrastructure.” Clim. Change, 116(2), 411–423.
Gillis, J. (2015). “For drinking water in drought, California looks warily to
Supplemental Data
sea.” The New York Times, Apr. 11, A1.
Gordon, J. (2017). “Desal water will be needed every year, says Victorian
Further description of methods used, Figs. S1–S13, and Tables S1
minister Lisa Neville.” The Age, Mar. 19.
and S2 are available online in the ASCE Library (www
Grant, S. B., et al. (2013). “Adapting urban water systems to a changing
.ascelibrary.org). climate: Lessons from the millennium drought in southeast Australia.”
Environ. Sci. Technol., 47(19), 10727–10734.
Gregory, R. (2000). “Using stakeholder values to make smarter environ-
References
mental decisions.” Environ. Sci. Policy Sustainable Dev., 42(5), 34–44.
ACIL Tasman. (2011). “Wholesale energy cost forecast for serving residen- Groves, D. G., and Lempert, R. J. (2007). “A new analytic method for find-
tial users.” Australian Energy Market Commission, Melbourne, ing policy-relevant scenarios.” Global Environ. Change, 17(1), 73–85.
VIC, Australia. Groves, D. G., Yates, D., and Tebaldi, C. (2008). “Developing and applying
Australian Bureau of Statistics. (2014). “Regional population growth.” uncertain global climate change projections for regional water manage-
〈http://www.abs.gov.au/ausstats/abs@.nsf/products/AC53A071B4B231 ment planning.” Water Resour. Res., 44(12), 1–16.
A6CA257CAE000ECCE5?OpenDocument#PARALINK1〉 (May 1, Hämäläinen, R. P., Kettunen, E., and Ehtamo, H. (2001). “Evaluating a
2016). framework for multi-stakeholder decision support in water resources
Baker, J., Block, P., Strzepek, K., and de Neufville, R. (2014). “Power management.” Group Decis. Negotiation, 10(4), 331–353.
of screening models for developing flexible design strategies in hy- Harrison, M. (2010). “Valuing the future: The social discount rate in cost-
dropower projects: Case study of Ethiopia.” J. Water Resour. Plann. benefit analysis.” Australian Government Productivity Commission,
Manage., 10.1061/(ASCE)WR.1943-5452.0000417, 4014038. Canberra, Australia.
Benayoun, R., de Montgolfier, J., Tergny, J., and Laritchev, O. (1971). Hashimoto, T., Stedinger, J. R., and Loucks, D. P. (1982). “Reliability, resil-
“Linear programming with multiple objective functions: Step method iency, and vulnerability criteria for water resource system performance
(stem).” Math. Program., 1(1), 366–375. evaluation.” Water Resour. Res., 18(1), 14–20.

© ASCE 04017061-8 J. Water Resour. Plann. Manage.

J. Water Resour. Plann. Manage., 2017, 143(10): 04017061


IPCC (Intergovernmental Panel on Climate Change). (2014). “Climate of concepts, state-of-art and methods.” Resour. Conserv. Recycl., 89, 1–
change 2014: Synthesis report.” Contribution of Working Groups I, 10.
II and III to the Fifth Assessment Rep. of the Intergovernmental Panel National Research Council. (2004). Adaptive management for water re-
on Climate Change, Geneva. sources project planning, National Academies Press, Washington, DC.
Jenkins, M., Lund, J., and Howitt, R. (2003). “Economic losses for urban Pahl-Wostl, C. (2007). “Transitions towards adaptive management of water
water scarcity in California.” J. Am. Water, 95(2), 1–20. facing climate and global change.” Water Resour. Manage., 21(1),
Jeuland, M., and Whittington, D. (2014). “Water resources planning under 49–62.
climate change: Assessing the robustness of real options for the Blue Pallottino, S., Sechi, G., and Zuddas, P. (2005). “A DSS for water resources
Nile.” Water Resour. Res., 50(3), 2086–2107. management under uncertainty by scenario analysis.” Environ. Modell.
Jowit, J. (2010). “Thames Water opens first large-scale desalination plant in Software, 20(8), 1031–1042.
UK.” The Guardian, Jun. 2. Porter, M. G. (2013). “A tale of two cities: Desalination and drought in
Kasprzyk, J. R., Reed, P. M., Characklis, G. W., and Kirsch, B. R. (2012). Perth and Melbourne.” Rep. prepared for National Centre of Excellence
“Many-objective de novo water supply portfolio planning under deep
in Desalination (NCEDA) under ‘Desalination for Australian Eco-
uncertainty.” Environ. Modell. Software, 34, 1–116.
nomic Development,’ Deakin Univ., Melbourne, VIC, Australia, 1–43.
Kwakkel, J. H., Walker, W. E., and Marchau, V. A. W. J. (2010).
Rawls, C. G., and Turnquist, M. A. (2012). “Pre-positioning and dynamic
“Classifying and communicating uncertainties in model-based policy
delivery planning for short-term response following a natural disaster.”
Downloaded from ascelibrary.org by Sarah Fletcher on 07/26/17. Copyright ASCE. For personal use only; all rights reserved.

analysis.” Int. J. Technol. Policy Manage., 10(4), 299.


Lempert, R. J., and Groves, D. G. (2010). “Identifying and evaluating ro- Socio-Econ. Plann. Sci., 46(1), 46–54.
bust adaptive policy responses to climate change for water management Ray, P. A., and Brown, C. M. (2015). Confronting climate uncertainty in
agencies in the American West.” Technol. Forecasting Soc. Change, water resources planning and project design: The decision tree frame-
77(6), 960–974. work, World Bank Publications, Washington, DC.
Lempert, R. J., Groves, D. G., Popper, S. W., and Bankes, S. C. (2006). “A Renn, O., Webler, T., Rakel, H., Dienel, P., and Johnson, B. (1993). “Public
general, analytic method for generating robust strategies and narrative participation in decision making: A three-step procedure.” Policy Sci.,
scenarios.” Manage. Sci., 52(4), 514–528. 26(3), 189–214.
Lienert, J., Monstadt, J., and Truffer, B. (2006). “Future scenarios for a Rousseeuw, P. J., Ruts, I., and Tukey, J. W. (1999). “The bagplot: A bivari-
sustainable water sector: A case study from Switzerland.” Environ. ate boxplot.” Am. Statistician, 53(4), 382–387.
Sci. Technol., 40(2), 436–442. Sapkota, M., et al. (2015). “An integrated framework for assessment of
Lins, H. F., and Stakhiv, E. Z. (1998). “Managing the nation’s water in a hybrid water supply systems.” Water, 8(1), 153–174.
changing climate.” J. Am. Water Resour. Assoc., 34(6), 1255–1264. Schlosser, C., Strzepek, K., and Gao, X. (2014). “The future of global water
Liu, X. M., Huang, G. H., Wang, S., and Fan, Y. R. (2016). “Water resour- stress: An integrated assessment.” Earth’s Future, 2(8), 341–361.
ces management under uncertainty: Factorial multi-stage stochastic pro- Stakhiv, E. Z. (2011). “Pragmatic approaches for water management
gram with chance constraints.” Stochastic Environ. Res. Risk Assess., under climate change uncertainty.” J. Am. Water Resour. Assoc., 47(6),
30(3), 945–957. 1183–1196.
Loucks, D. P. (2000). “Sustainable water resources management.” Water Strzepek, K., Jacobsen, M., Boehlert, B., and Neumann, J. (2013). “Toward
Int., 25(1), 3–10. evaluating the effect of climate change on investments in the water re-
Low, K. G., et al. (2015). “Fighting drought with innovation: Melbourne’s sources sector: Insights from the forecast and analysis of hydrological
response to the millennium drought in Southeast Australia.” Wiley indicators in developing countries.” Environ. Res. Lett., 8(4), 44014.
Interdiscip. Rev. Water., 2(4), 315–328. Turner, S. W. D., Marlow, D., Ekström, M., Rhodes, B. G., Kularathna, U.,
Marques, J., Cunha, M., and Savić, D. (2014). “Using real options in the and Jeffrey, P. J. (2014). “Linking climate projections to performance: A
optimal design of water distribution networks.” J. Water Resour. Plann.
yield-based decision scaling assessment of a large urban water resour-
Manage., 10.1061/(ASCE)WR.1943-5452.0000448, 4014052.
ces system.” Water Resour. Res., 50(4), 3553–3567.
Melbourne Water. (2015a). “Customers and prices.” 〈http://www.melbourne
Urich, C., and Rauch, W. (2014). “Exploring critical pathways for urban
water.com.au/aboutus/customersandprices/Pages/Bulk-water.aspx〉 (Jan.
water management to identify robust strategies under deep uncertain-
5, 2015).
ties.” Water Res., 66, 374–389.
Melbourne Water. (2015b). “Water data.” 〈http://www.melbournewater.com
.au/waterdata/waterstorages/Pages/Inflow-over-the-years.aspx〉 (May 1, Vörösmarty, C. J., Green, P., Salisbury, J., Lammers, R. B., and Vorosmarty,
2015). C. J. (2000). “Global water resources: Vulnerability from climate
Melbourne Water. (2015c). “Water outlook for Melbourne.” 〈https://www change and population growth.” 289(5477), 284–288.
.melbournewater.com.au/getinvolved/saveandreusewater/Documents/ Walker, W. E., et al. (2003). “A conceptual basis for uncertainty manage-
Water%20Outlook%20December%202015.pdf〉. ment.” Integr. Assess., 4(1), 5–17.
Milly, P. C. D., et al. (2008). “Climate change. Stationarity is dead: Whither Wang, T., and de Neufville, R. (2006). “Real options ‘in’ projects.” 8th Real
water management?” Science, 319(5863), 573–574. Options Annual Int. Conf., International Council on Systems Engineer-
Moody, P., and Brown, C. (2012). “Modeling stakeholder-defined climate ing, Orlando, FL, 1–1834.
risk on the Upper Great Lakes.” Water Resour. Res., 48(10), 1–15. Zhang, S. X., and Babovic, V. (2012). “A real options approach to the
Nair, S., George, B., Malano, H. M., Arora, M., and Nawarathna, B. (2014). design and architecture of water supply systems using innovative water
“Water-energy–greenhouse gas nexus of urban water systems: Review technologies under uncertainty.” J. Hydroinf., 14(1), 13–29.

© ASCE 04017061-9 J. Water Resour. Plann. Manage.

View publication stats J. Water Resour. Plann. Manage., 2017, 143(10): 04017061

You might also like