exploring alternative futures of the world water system. a ... · conservancy, 2010). marginal...

39
Exploring alternative futures of the World Water System. Building a second generation of World Water Scenarios Driving force: Water Resources and Ecosystems Martina Bertsch 2010 Prepared for the United Nations World Water Assessment Programme UN WWAP

Upload: others

Post on 19-May-2020

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

  

    

 Exploring alternative futures of the World Water System.  

 Building a second generation of World Water Scenarios 

 Driving force: Water Resources and Ecosystems 

     

  

Martina Bertsch 2010 

      

    

Prepared for the United Nations World Water Assessment Programme UN WWAP  

  

Page 2: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

2

    

Table of contents      1. Introduction2. Current assessment of the driver 2.1. Ecosystem services 2.2. Water resources 2.3. Floods and droughts 2.4. Biodiversity and invasive species  2.5. Deforestation 2.6. Water availability, use and productivity 2.7. Water quality  3. Possible future developments in the domain of the driver 3.1. Ecosystem services 3.2. Water resources 3.3. Floods and droughts 3.4. Biodiversity and invasive species  3.5. Deforestation 3.6. Water availability, use and productivity 3.7. Water quality  Acknowledgements Bibliography 

               

Page 3: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

3

      

1. Introduction This report examines global ecosystems as a driving force  impacting the water resources. 

The  report  begins  with  a  description  of  the  current  state  of  global  ecosystems  and  water resources,  depicted  through  a  series  of  key  elements  and  their  complex  interactions  and interdependencies.  It  identifies major  impacts on the current state.  It then extracts major trends and  projections  for  future  states  from  available  global  scenarios.  The  table  of  possible developments  is provided, detailing events, time factors, positive and negative  impacts on water resources,  low  probability  developments  and  principal  sources  of  information.  Causal  links  are established between this driver and the other nine water drivers. Lastly, critical uncertainties and future research needs are  identified. Driver elements analyzed  in this report  include: 1) Value of ecosystem  services;  2)  Freshwater  resources  (e.g., wetlands);  3)  Extreme  climate/water‐related events  (floods  and droughts); 4) Biodiversity  and  invasive  alien  species; 5) Changes  in  land use (deforestation);  6)  Changes  in  water  availability,  use  and  productivity,  including  groundwater overdraft; 7) Changes in water quality (e.g., eutrophication).  

2. Current assessment of the driver  

2.1. ECOSYSTEM SERVICES An ecosystem is a dynamic complex of plant, animal and microorganism communities and 

the  nonliving  environment  interacting  as  a  functional  unit.  Humans  are  an  integral  part  of ecosystems.  The Millennium  Ecosystem  Assessment  (MA,  2005)  defines  ecosystem  services  as benefits provided to people, comprising: 1) Provisioning services, i.e., the products obtained from ecosystems,  including  fresh  water,  food,  fuel,  fiber,  biochemical  and  genetic  resources.  2) Regulating  services,  i.e.,  the  benefits  obtained  from  the  regulation  of  ecosystem  processes, including  climate  regulation,  air  quality  maintenance,  regulation  of  hydrological  flows,  water purification,  human  disease  and  pest  control,  pollination,  coastal  protection,  flood  and  erosion control.  3)  Cultural  services,  i.e.,  the  nonmaterial  benefits  obtained  from  ecosystems  through spiritual enrichment, cognitive development,  reflection,  recreation and aesthetic experiences. 4) Supporting services, necessary  for the production of other services, e.g., oxygen production, soil formation, nutrient storage, recycling, processing and acquisition.1 

 

1 Provision and regulation of water by aquatic ecosystems is essential for numerous biochemical and

geochemical processes. Another important service attributed to aquatic ecosystems is provisioning of fish and seafood. Fish are an important source of proteins, micronutrients, minerals and essential fatty acids, and significantly contribute to the diets of many communities. Globally, 1 billion people rely on fish as their main source of animal proteins, and some small island nations depend on fish almost exclusively. According to FAO (1999), 79% of fish products are harvested from marine sources.

Page 4: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

4

The ability of any particular aquatic ecosystem to supply the range of services listed above depends  upon  a  variety  of  factors,  e.g.,  the  type  of  ecosystem,  the  presence  of  key  species, management  interventions,  the  location  of  human  communities,  the  surrounding  climate  and topography. Few sites have the capacity to provide all of the above services. Generally, the more biologically diverse an ecosystem  is, the greater the range of services that can be derived from it (MA, 2005). 

Current decision‐making processes often  ignore or underestimate the value of ecosystem services. There are two paradigms of value: 

1) The anthropocentric (utilitarian) concept is based on the principle of humans’ preference satisfaction  (welfare).  Ecosystem  services  have  value  to  societies  because  people  directly  or indirectly  derive  utility  from  their  use  (use  values).  Value  is  often  ascribed  to  knowing  that  a resource exists even if people never use that resource directly (non‐use or existence values), e.g., the deeply held historical, national, ethical, religious and spiritual values ascribed to ecosystems. 

2)  The  non‐utilitarian  paradigm  holds  that  an  ecosystem  can  have  intrinsic  value irrespective  of  its  contribution  to  human  well‐being.  Intrinsic  value  may  complement  or counterbalance considerations of utilitarian value (Alcamo et al., 2005).  

Well‐documented impacts of human activities on ecosystem services at a variety of scales include changes in the number and distribution of species (Chapin et al., 2000; Higgins et al., 2002; Sala et al., 2000), and the quality and quantity of fresh water (Brinson and Malvarez, 2002).  

 For  certain  ecosystem  services, monetary  values  are  estimated.2  Conversely,  ecological 

degradation results in economic losses. Coastal erosion from altered currents and sediment loads can be caused by changes in coastal and upstream land use. The beaches largely disappear in areas where new ports are built, resulting in substantial tourism income losses. Societal consequences of altered biodiversity are also measured.3  

 MA (2005) reports that the degradation of lakes, rivers, marshes and groundwater systems 

is more  rapid  than  that  of  other  ecosystems.  The  status  of  freshwater  species  is  deteriorating faster  than  those  of  other  ecosystems.  Direct  drivers  of  degradation  of  freshwater  ecosystem services  include:  infrastructure development,  land conversion, water withdrawal, eutrophication, pollution,  overexploitation  and  the  introduction  of  exotic  species  (Mayers  et  al.,  2009).  UNEP (2009a)  provides  a  review  of water  security  and  ecosystem  services  case  studies  and  lessons learned about habitat rehabilitation, pollution control and integrated watershed management. 

2 Direct use values of water buffering (e.g., flood prevention) are estimated at US $350 billion at 1994 prices, and recreational values at US $304 billion (Costanza et al., 1997). It is estimated that reef habitats provide human beings with living resources (e.g., fish), and services (e.g., tourism returns and coastal protection), worth about US $375 billion each year (Costanza et al., 1997). 3 In the U.S., total annual damage associated with aquatic invasive species is estimated at $73 billion (the Nature Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are valued at US $16-56 million per year. In South Africa's Cape region, the presence of rapidly transpiring exotic pines raises the unit cost of water procurement by nearly 30%. Increased evapotranspiration due to the invasion of Tamarix in the U.S. costs an estimated US $65-180 million per year in reduced municipal and agricultural water supplies. The presence of sediment-trapping Tamarix also narrows river channels and obstructed over-bank flows, increasing annual flood damages by $50 million (as cited by Chapin et al., 2000).

Page 5: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

5

 2.2. WATER RESOURCES 

Less than 3% of the global total water resources is represented by freshwater and less than 1% of  that  (less  than 0.01% of  total water) occurs  in  the Earth’s  liquid  surface  freshwater. The remainder  represents  ice‐caps  or  groundwater.  The  small  fraction  of  liquid  surface  freshwater hosts an extraordinary  level of biodiversity  supported  through a  range of  freshwater ecosystem types:  1)  running  waters,  2)  standing  waters  and  3)  areas  of  transient  water  availability. Freshwater ecosystems  include: permanent and temporary rivers and streams; permanent  lakes, reservoirs; seasonal lakes, marshes and swamps, including floodplains; forested, alpine and tundra wetlands; springs and oases; groundwater systems and geothermal wetlands (Mayers et al., 2009).  

The  volume  of water  in  rivers  and  streams  is only  a  fraction  of  the water  in  the  entire hydrosphere, but often this water constitutes the most accessible and  important water resource (WWDR3, 2003). Uneven global distribution of precipitation and runoff  influence the distribution of  river  networks.  Asia  and  Latin  America  each  contribute  approximately  30%  of  the  world freshwater  discharged  into  the  ocean, North  America  contributes  17, Africa  10,  Europe  7,  and Australasia 2%  (Fekete et al., 1999). There are several published  inventories of rivers,  listing  the major river systems with their drainage area, length and average runoff (e.g., Shiklomanov, 1997). The variability between estimates can be explained by different definitions regarding the extent of a river system and different time periods or locations for the measurement. 

 There are 5‐15 million  lakes across  the globe  (WWDR3, 2003) and approximately 10,000 

lakes have a surface area over 1 km2. Only about 15‐20 existing lakes in the world are older than 1 million years (LakeNet, 2003). A disproportional share of large lakes, with a surface area over 500 km2,  is  found  in North America, where glacial scouring created many depressions  in which  lakes have formed. The International Lake Environment Committee (ILEC) maintains a database of over 500 lakes worldwide, with some physiographic, biological and socioeconomic information. 

 The definition of  inland wetlands adopted by the Convention on Wetlands  (Ramsar,  Iran, 

1971)  covers  all wetland  types,  including  artificial wetlands,  and  encompasses  permanent  and temporary  riverine,  lacustrine  and  palustrine  systems,  above  ground  and  underground  systems (karst  and  cave),  freshwater,  brackish  and  saline  systems  (Ramsar  Classification  System  for Wetland  Type  in  the  Annex  to  Resolution  VII.11).  A  Global  Review  of Wetland  Resources  and Priorities for Wetland Inventory (GRoWI, Finlayson & Spiers, 1999) estimated the global extent of wetlands  including  coastal  wetlands  in  some  countries,  at  1,276‐1,279 million  hectares  (ha).4 Revenga and Kura (2003) offer some reasons for the incompleteness of global wetland inventories, notably inconsistent interpretations of definitions, difficulty defining the boundaries, limitation of maps and remote sensing products. 

 

4 Compared with previous estimates of inland waters of 530-970 million ha. These estimates include 530 million ha for natural freshwater wetlands. A substantial portion of the inland water resource is known to be peatland. The Global Peatland Initiative estimates coverage of 324 million ha, or 25-46% of the total area of inland waters from different sources. Cultivated rice paddies also form a significant area of the inland resource, covering at least 130 million ha (Aselmann and Crutzen, 1989).

Page 6: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

6

During the past hundred years, the groundwater portion of the water cycle was subjected to massive changes, as humans learned to dig or drill wells and abstract groundwater using pumps. Groundwater  serves  as  a  water  shortage  buffer  during  short‐term  climate  variations  and  is important  to  adaptation  strategies.  Groundwater mining  from  fossil  aquifers  is  often  the  only reliable means of obtaining water.  

 These  groundwater  resources  are  increasingly being  used  for  agricultural,  industrial  and 

domestic water supplies, although they are almost never recharged.  Changing land use and water infrastructure  also  greatly modify  groundwater  regimes, with  groundwater pumping  from deep aquifers  now  a  worldwide  phenomenon.  In many  instances,  groundwater  is  pumped  with  no understanding  of  its  source  or  its  annual  recharge,  and  therefore  of  how much may  be  used sustainably. Heavy groundwater pumping has led to unsustainable conditions. Stresses result in: 1) Falling water levels; 2) Degraded groundwater aquifers; 3) Increased salinization; 4) Chemical and microbiological  pollution;  5)  Desiccated  wetlands;  6)  Dewatered  rock  sequences;  7)  Land subsidence.  Use  of  groundwater  for  irrigated  agriculture  increased  enormously  in  the  past  50 years.  Some  70%  of  global  groundwater  abstraction  is  now  estimated  to  be  used  in  irrigation. Pollution of  shallow  aquifers became widespread  four or  five decades  ago  and  triggered water quality protection measures in many countries. Groundwater abstractions also contributed to the development of rural economies.   

 2.3. FLOODS AND DROUGHTS 

Extreme  climate/water‐related  events  can  have  positive  and  negative  impacts  on ecosystems and their services. They recharge natural ecosystems, providing more abundant water for  food production, health and  sanitation. But extreme water‐related events also destroy  lives and property. Adikari and Yoshitani  (2009) analyze water‐related disaster events between 1980 and 2006.5 The  factors  that  lead  to  increased water‐related disasters  include: natural pressures, e.g.,  climate variability; management pressures, e.g.,  lack of appropriate organizational  systems and  inappropriate  land management;  and  social  pressures,  e.g.,  escalation  of  population  and settlements in high‐risk areas (particularly for poor people).   

Flooding  is  broadly  defined  to  include  both  excess  water  caused  by  rainstorms  that subsequently  leads rivers  to  flood  (spill over  their banks), and severe coastal storms or cyclones that can  lead  to excess water  through  largely  tidal surges  (Cosgrove and Rijsberman, 2000). The moisture‐holding capacity of the atmosphere has been increasing at a rate of about 7% per 1°C of warming, creating the potential for heavier precipitation. There have  likely been  increases  in the number of heavy precipitation events in many land regions, consistent with a warming climate and 

5 Numbers of floods and windstorms increase drastically from 1997 to 2006 (factor of 2 and 1.5, respectively), but other types of disasters do not increase significantly in this period. Drought is severe at the beginning of the 1980’s and gains momentum again during the late 1990’s and afterwards. The numbers of landslides and waterborne epidemics are highest during 1998–2000 and then decrease. Waves and surges increase between 1980 and 2006. In general, water-related disaster fatalities follow a decreasing trend, but the fatalities record has occasional peaks: droughts, during the period 1983–1985; windstorms, 1989–1991; waves and surges, 2004–2006. Global estimates of water-related economic losses show an increasing trend, the main contributors in descending order being windstorms, floods and droughts; the rest of the water-related disasters are insignificant, but underestimated (Adikari and Yoshitani, 2009).

Page 7: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

7

the observed  increase  in atmospheric water vapor, even where  total precipitation has declined. Some people  rely on  floods  for  irrigation and natural  fertilization, but disruption, devastation of human lives and infrastructure are common. Among the regions devastated by floods in the period 2001‐2010 were Southern China, Haiti,  India,  Indonesia, U.S., Mexico, South Eastern Europe and most recently Portugal. In developing countries extreme floods can result in many deaths, while in developed countries extreme floods cause material damage  in the billions and tens of billions of dollars.   

Van Lanen et al. (2007) define drought as a sustained and regionally extensive occurrence of  below  average  natural  water  availability.  The  main  causes  are  low  precipitation  and  high evaporation rates. In regions with a cold climate, temperatures below zero can also give rise to a winter drought. Drought is characterized as a deviation from normal climate and hydrology, which is reflected in precipitation, soil water, groundwater and streamflow. Droughts have a substantial impact on  the ecosystem and agriculture of  the affected  region. Climate  change  is expected  to influence precipitation, temperature and potential evapotranspiration and to influence occurrence and severity of droughts. But it is difficult to disentangle the impacts of climate change from those of other human  influences, e.g., engineered effects,  land use  changes and multidecadal climate variability. More  intense droughts, affecting more people and  linked to higher temperatures and decreased precipitation have been observed  in the 21st century (Zhang et al., 2007). Burke et al. (2006) assess meteorological drought in the Hadley Centre global climate model using the Palmer Drought  Severity  Index  (PDSI).  On  average,  the  limiting  PDSI  values  for  extreme,  severe  and moderate drought are −4.3, −3.3, and −2.0, respectively. The land surface in drought increased by the beginning of  the 21st century  for all  three types of drought,  from 1%  to 3%  for  the extreme droughts,  from  5%  to  10%  for  the  severe  droughts,  and  from  20%  to  28%  for  the moderate droughts.  

2.4. BIODIVERSITY AND INVASIVE SPECIES Biodiversity is the variability among living organisms from all sources, including terrestrial, 

marine  and  freshwater  ecosystems.  It  is  a  composite measure  of  the  number  of  species,  i.e., species richness, and the number of individuals of different species, i.e., relative abundance (MA, 2005).  It  includes diversity within and between species and diversity of ecosystems.   Freshwater biodiversity  is high  relative  to  the  limited portion of  the Earth’s  surface  covered by  freshwater. Freshwater  fish make up 40% of all  fish, and  freshwater molluscs make up 25% of all molluscs. Freshwater biodiversity tends to be greatest in tropical regions, with a large number of species in northern South America, central Africa, and Southeast Asia. Worldwide, the number of freshwater species  is estimated to be between 9,000 and 25,000 (Cosgrove and Rijsberman, 2000). Revenga and  Kura  (2003)  review  inland  water  species  richness,  distribution  and  conservation  status. Ecological  indicators are  catalogued by Olli and Tamminen  (2006):  inherently univariate  indices; complex indices; marine benthic indices; entropy indices; river indices; and physiological indices.6 

6 1) Inherently univariate indicators are simple to measure and interpret, e.g., sea surface temperature, secchi depth, concentration and percentage of dissolved oxygen, biological and chemical oxygen demand, bulk measurements of dissolved and particulate nutrients. They are included in routine monitoring programs and long time series are available; 2) Complex indices are various arithmetic combinations of univariate parameters; 3) Marine benthic indices rely on the paradigm stating that benthic communities respond to improvements in habitat quality in three steps: the abundance

Page 8: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

8

Changes in ecosystem biodiversity can influence the services they provide. In addition, the diversity of  living  species has  intrinsic value  independent of any human  concern  (Alcamo et al., 2005). A growing number of species and habitats are disappearing. According  to  the 2004  IUCN Red  List of Threatened Species, a  total of 15,589  species of plants and animals are  threatened, facing a high risk of extinction in the near future, in almost all cases as a result of human activities. Since  the beginning of  the  century, 50% of wetlands have disappeared.   The  loss of  freshwater biodiversity is poorly monitored except for some larger, commercial species.7 The most important causes of biodiversity  loss  in  freshwater  systems are: 1) Habitat  loss and degradation; 2) Water quality, not discussed in detail because of the lack of comprehensive data at global scale (Revenga and Kura, 2003); 3)  Intentional and unintentional  introduction of exotic or  invasive alien species, IAS (Hill et al., 1997).  

In  some  lake  ecosystems,  IAS  are  considered  by  some  to  be  the  primary  cause  of biodiversity  loss  (Ciruna et al., 2004).  IAS can become established  in an ecosystem as a result of political,  demographic,  cultural,  socio‐economic  or  ecological  factors.  The  rich  endemic ichthyofauna of Lake Victoria in Africa have been reduced by predatory Nile perch, overfishing and eutrophication  (McAllister  et  al.,  1997).  Today’s  global  spread  of  IAS  is  mainly  driven  by aquaculture,  shipping  and  global  commerce  (Welcomme,  1988).  Intentional  introductions comprise exotic  flora and  fauna  in gardens or waterways, or government‐sanctioned releases of organisms for propagation or harvest. Unintentional introductions occur as a result of the escape of aquaculture operations or the accidental transport of organisms attached to boats, structures, garbage or  in ballast water. There  is  some  correlation between  levels of human  activity,  trade, ecological  integrity and the resistance of ecosystems to  invasion from  introduced species (Ciruna et al., 2004). Degradation of ecosystem  functions generally  results  in a greater  susceptibility  to invasions.  In  modern  environmental  management,  a  two‐phase  solution  is  recommended  to address aquatic  IAS: 1) Prevention of  their spread  through education, partnership and policy; 2) Management of invasions and restoration of freshwater habitats.  

IAS  affect  native  species  through  predation,  competition,  disruption  of  food  webs, alteration  of water  quality  and  the  introduction  of  diseases.  The  species  introduced  include  a 

increases; species diversity increases; and dominant species change from pollution-tolerant to pollution-sensitive species; 4) Entropy indices are based on information theory; 5) River indices are incorporated into governmental monitoring programs and specific to one region or country. Many use specific indicator species, applicable to a restricted region or context. An important concept is the integration time of the environmental signal by the organisms. Unicellular phytoplankton or benthic diatoms respond rapidly to fluctuations in the environment, while sessile benthic fauna integrates over several years; 6) Physiological indices are usually expressed as ratios of physiological rates, e.g., the ratio of primary production to community respiration indicates the degree of heterotrophy of the system. Some physiological indices are very specialized and often use modern molecular techniques, e.g., evaluation of the total antioxidant status in algal tissue extracts to assess the stress response of algae to environmental impacts in urban aquatic habitats. 7 Over 37% of wetland dependant mammals, nearly one third of amphibian species and 50% of turtles are threatened with extinction; 41% of waterbird populations are in decline. Data suggest that 20–35% of freshwater fish are vulnerable or endangered, mostly because of habitat alteration. Other factors include pollution, non-native species and overharvesting.

Page 9: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

9

variety of taxa  from  fish  (e.g., Asian carp8), higher plants  (e.g., water hyacinth9) to  invertebrates and microscopic plants  (e.g., dinoflagellates10). Worldwide,  two  thirds of  the  freshwater  species introduced into the tropics and more than 50% of those introduced to temperate regions become established (Welcomme, 1988). In aquaculture, exotic fish are usually introduced to enhance food production  and  recreational  fisheries,  or  to  control  pests,  e.g., mosquitoes  and  aquatic weeds. Introduced  fish  account  for  96.2%  of  fish  production  in  South  America  and  84.7%  in  Oceania (Garibaldi  and Bartley, 1998).  In  Europe, North America, Australia  and New  Zealand,  in  77% of cases native fish populations are reduced or eliminated following the introduction of exotic fish.  

 2.5. DEFORESTATION 

Deforestation  is  the  clearance  of  naturally  occurring  forests  by  logging  and  burning.11 Disregard  or  ignorance  of  intrinsic  value,  lack  of  ascribed  value,  lax  forest  management  and deficient environmental  law are some of the factors that allow deforestation to occur on a  large scale.  In many countries, deforestation  is an ongoing  issue  that  is causing extinction, changes to climatic  conditions,  desertification  and  displacement  of  indigenous  people  (FAO,  2005). Deforestation also significantly impacts the water cycle: 

 1)  It  reduces  the  content  of water  in  the  soil,  groundwater  and  atmospheric moisture. Trees extract groundwater through their roots and release it into the atmosphere. When part of a forest  is  removed,  the  trees  no  longer  evaporate  away  this  water,  resulting  in  a much  drier climate.  

2)  It  reduces  soil  cohesion, which  results  in erosion,  flooding  and  landslides. Vegetation litter,  stems  and  trunks  slow  down  runoff.  Roots  create macropores  in  the  soil  that  increase infiltration of water.  8 Asian carp have been found in the Illinois River, which connects the Mississippi River to Lake Michigan (USA). Two species of Asian carp (the bighead and silver carp) were imported by catfish farmers in the 1970’s to remove algae and suspended matter out of their ponds. During large floods in the early 1990’s, catfish farm ponds overflowed their banks, and the Asian carp were released into local waterways in the Mississippi River basin. They can weigh up to 100 pounds, and can grow to a length of more than four feet. They are well-suited to the climate of the Great Lakes region, which is similar to their native Asian habitats. Due to their large size and rapid rate of reproduction, these fish can disrupt the food chain that supports the native fish and pose a significant risk to the Great Lakes Ecosystem. Additionally, silver carp can leap several feet out of the water when disturbed by boat propellers, possibly causing damage to boats and injuries to boat operators and fishermen. To prevent the carp from entering the Great Lakes, USEPA and other authorities are working to install and maintain a permanent electric barrier between the fish and Lake Michigan (USEPA, 2004a). 9 Water hyacinth (Eichhornia crassipes) became invasive in many tropical and sub-tropical inland waters. Indigenous to the upper Amazon basin, it was spread throughout much of the planet for use as an ornamental plant beginning in the mid-19th century. By 1900, it spread to every continent except Europe. Today, the plant has a pan-tropical distribution. It quickly extends its range throughout rivers and lakes in the tropics, clogging waterways and infrastructure, reducing light and oxygen in freshwater systems, and causing changes in water chemistry and species assemblages (Hill et al., 1997). 10 Dinoflagellates are protists, mostly marine plankton, which sometimes bloom in concentrations of more than a million cells per milliliter. Certain species produce neurotoxins (e.g., saxitoxin, a powerful paralytic), which kill fish and accumulate in filter feeders, e.g., shellfish, which in turn cause poisoning in people. Their rapid and copious reproduction due to the abundant nutrients in the water can result in the red tide phenomenon. 11 Trees or derived charcoal are used or sold for fuel or as a commodity. Cleared land is used as pasture for livestock, plantations of commodities and settlements. The removal of trees without sufficient reforestation results in damage to habitat, biodiversity loss and aridity. Deforestation has adverse impacts on biosequestration of atmospheric carbon dioxide.

Page 10: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

10

3)  It  lessens  the  landscape’s  capacity  to  intercept,  retain  and  transpire  precipitation. Instead of trapping precipitation, which then percolates to groundwater systems, deforested areas become  sources of  surface water  runoff, which moves much  faster  than  subsurface  flows. That quicker transport of surface water can translate into flash flooding and more localized floods than would occur with the forest cover.  

4)  It  contributes  to  decreased  evapotranspiration,  which  lessens  atmospheric moisture which can affect precipitation levels downwind from the deforested area, as water is not recycled to downwind forests, but is lost in runoff and returns directly to the oceans.   

According to the FAO Global Forest Resource Assessment (FRA 2005), forests cover 30% of the  total  land area. The  total  forest area  in 2005  is  just under 4 billion ha, corresponding  to an average of 0.62 ha per capita. But the area of forest  is unevenly distributed: 64 countries with a combined population of 2 billion have  less than 0.1 ha of forest per capita. The ten most forest‐rich countries account for two‐thirds of the total forest area. Seven countries or territories have no forest at all, and an additional 57 have forest on less than 10% of their total land area.  

2.6. WATER AVAILABILITY, USE AND PRODUCTIVITY Water availability  is  critical  for  the maintenance of  flora,  fauna,  freshwater  systems and 

human activities. Water is needed for forests and wetlands, which in turn recharge aquifers, store runoff  and  regulate  pollution  by  processing  organic waste  (Johnson  et  al.,  2001).  Sufficient  in‐stream water availability can help temper water pollution through the dilution of contaminants in the watercourse. Extreme changes  in ecosystems may occur  if water available for environmental uses  falls below a certain  threshold. There  is a growing  interest  in understanding environmental demands for water. Environmental flows refer to the quality, quantity and timing of water flows required  to  protect  an  ecosystem,  enable  ecologically  sustainable  development  and  water resource utilization. The Nature Conservancy developed the following tools for environmental flow management: 1) Ecological Limits of Hydrologic Alteration12  (ELOHA); 2) Ecologically Sustainable Water Management13  (ESWM);  and  3)  Indicators  of  Hydrologic  Alteration14  (IHA). The  possible trade off between water for nature and water for food production can be a contentious issue, as demonstrated during the 2nd World Water Forum in The Hague (2000). Participants in the “Water for Food” theme stressed the need for slow growth in water consumption in agriculture, while the “Water  for  Nature”  theme  called  for  significant  reallocation  of  water  from  agriculture  to environmental uses.  

12 ELOHA is a scientifically robust and flexible framework for assessing environmental flow needs across large regions (states, provinces, major river basins or entire countries), aimed at understanding the ecological ramifications of human-induced alterations in river flows. Information gathered through ELOHA allows water managers and stakeholders to develop environmental flow targets for rivers without requiring detailed, site-specific hydrologic or biological information for each river. 13 ESWM is a six step framework that determines how to meet human water needs by storing, diverting and releasing water in a manner that either sustains or restores a river’s ecological integrity. 14 IHA is a free software program developed by Nature Conservancy, which provides information for developing environmental flow recommendations and understanding the impacts of human activities on water flows. IHA displays statistical information about water flow changes and trends in multiple formats, e.g., comparing the flow in a particular river between pre-dam and post-dam periods.

Page 11: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

11

   The total rainfall on the Earth’s  land surfaces amounts to 110,000 cubic kilometers (km3). Rain  replenishes blue water  sources,  i.e.,  rivers,  lakes,  etc,  and  green water,  i.e.,  soil moisture. About  39%  of  rain  contributes  to  blue  water,  supporting  biodiversity,  fisheries  and  aquatic ecosystems;  56%  of  green water  is  evapotranspired  by  landscape  uses  that  support  bioenergy, forest  products,  livestock  grazing  lands  and  biodiversity;  4.5%  is  evapotranspired  by  rainfed agriculture supporting crops and  livestock; total evapotranspiration by  irrigated agriculture  is 2% of rain, of which 30%  is directly  from rain  (green water) and the remainder  from  irrigation  (blue water). Blue water withdrawals  for  irrigation,  industry  and municipal use  represent 9% of  total blue water. Cities and industries return more than 90% of their withdrawals to blue water, but the return flows are often of lower quality (Molden, 2007).    Only a small  fraction of the total available water  is accessible by humans. Therefore,  it  is more  useful  to  consider  total  water  withdrawn  for  human  use.  Groundwater/Surface‐Water Maximum Allowable Water Withdrawals  (G/S MAWW) are actual water withdrawals, which are constrained  by  allowable  water  withdrawal  for  surface  water  and  groundwater.  MAWW  are limited by  infrastructure  constraints,  e.g., physical diversion  structures  and pumping  capacities, and policy constraints, e.g., the amount of water which must be  left  in‐stream for environmental purposes. Of  the  total water withdrawn  for human uses, withdrawals  for  agriculture  represent 70%, those for industry 20%, and for municipal use about 10%. Shiklomanov (1999) presents tables of the dynamics of water use by continents and by economic sectors.      Between  1950  and  1995,  agricultural  uses  for water more  than  doubled  (Cosgrove  and Rijsberman,  2000;  Shiklomanov,  1999).  Sixty  years  ago,  the world  population was  smaller,  less wealthy, consumed  fewer calories, ate  less meat and required  less water  to produce  their  food. When agricultural activities change the quality, quantity and timing of water flows to support food provision,  the ecosystems’ capacity  to provide other services  is  reduced. Regulating services are primarily  affected:  pollination,  biological  pest  control,  flood  retention  capacity,  changes  in microclimate regulation, the loss of biodiversity and habitats (Molden, 2007). The development of dams and irrigation systems has significant environmental and human resettlement costs, e.g., 48 million  people  were  displaced  by  dam  projects.  In  addition,  agricultural  expansion  caused  an estimated 56‐65% of the available wetlands  in Europe and North America to disappear. Globally, wetland loss is 26%, and still intensifying in many regions.     Agricultural water withdrawal  is  largely  attributed  to  irrigation.  The  key  environmental issues for rivers related to unsustainable agriculture are: excessive water depletion, water quality reduction,  waterlogging  (60–80 million  ha  of  irrigated  lands  are  affected),  salinization  (20–30 million  ha  irrigated  lands  severely  affected)  and  loss  of  ecosystem  services  (FAO,  1996).  River downstream effects are temperature changes, depleted fish stock, effects on wetland, reduction in silt and water quality. The upstream effects are siltation, salinization and deforestation.      Many  countries  and  basins  currently  exploit  their  groundwater  reserves  at  a  rate substantially exceeding recharge. Overdraft occurs when pumping rates exceed the rate of natural recharge. Unsustainable groundwater use is often associated with irrigation, and can lead to both 

Page 12: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

12

water  scarcity  and water quality problems. Postel  (1999) draws on  several  sources  to estimate total  annual  global  groundwater  overdraft  at  163  km3.  The  threshold  at  which  localized groundwater overdraft occurs at the whole‐basin  level can be set at 0.55. The key impacts which result from overdraft are:  increasing  lifts and costs from the  lowered water table, subsiding  land (sometimes  irreversibly  damaging  the  aquifer)  and  degrading  water  quality  (including  saline intrusion and arsenic).    

Water  productivity  is  largely  determined  by  crop  yield  in  the  agricultural  sector, which dominates water use. Relatively small gains can be achieved by trending towards more beneficial water  consumption.  The  greatest  benefit  comes  from  having  a  stable  climate  with  adequate rainfall. Water productivity  increased by  at  least 100% between 1961  and 2001, mainly due  to increases  in  crop  yields.  Irrigated  rice  yields  doubled  and  rainfed wheat  yields  rose  by  160%. Globally, FAO  (2003) estimated  that water needs  for  food per  capita halved between 1961 and 2001. A 10% increase in water productivity equals current domestic water consumption, therefore investing  in agricultural water management  is an attractive  strategy  for  freeing water  for other purposes. The water productivity of rice  is significantly  lower than that of other cereals, ranging from 0.15‐0.60 kilograms per cubic meter  (kg/m3), while that of other cereals ranges  from 0.20‐2.40 kg/m3. The average water productivity of other cereals  in the developed world  is 1.0 kg/m3, while in the developing world it is 0.56 kg/m3 (Rosegrant et al., 2002).  

In many developing countries,  industrial production and hence  the  sectoral use of water grow  fast,  putting  increasing  pressure  on  scarce  water  resources.  The  relationship  between industrial water withdrawal and  industrial growth  is not  linear, as technological advances  lead to water savings and water reuse in industry. Hence industrial water withdrawals in many developed countries have  flattened off, while  industrial water consumption  (which  is only a  fraction of  the total water withdrawal) continues to grow.     In  terms  of  domestic  water  use,  delivery  of  safe  water  and  sanitation  are  critical. Withdrawals  for domestic and  industrial uses grew  four‐fold between 1950 and 1995  (Cosgrove and Rijsberman, 2000; Shiklomanov, 1999). There  is much promise  in the ongoing developments for desalination technology, but today  it contributes only 0.02% of global water withdrawals and perhaps 1% of drinking water (Martindale and Gleick, 2001). It is mostly limited to coastal regions. Three  key  issues  are: high  energy  costs,  concentrate disposal  and high  capital  costs. Poverty  is directly associated with lack of access to safe drinking water, sanitation and water‐related disease. Global sanitation coverage rose from 49% in 1990 to 58% in 2002 (WHO, 2004).15 

15 Many people still access water and sanitation the same way their ancestors did thousands of years ago: 1 billion people lack access to improved water supply and 2.6 billion lack access to improved sanitation, resulting in water-related disease. There are four classes of water-related disease: 1) Waterborne, for which water is the agent of transmission, caused by pathogens transmitted from excreta to water to humans, e.g., cholera, hepatitis A and E; 2) Water-based, from hosts that either live in water or require water for part of their life cycle, e.g., schistosomiasis spreads in regions where irrigation and dam projects produce habitats that favor the host snails; 3) Water-washed, caused by water scarcity where people cannot wash themselves, their clothes or homes regularly, e.g., trachoma; and 4) Water-related insect vectors, e.g., malaria, onchocerciasis (UNEP, 2010). Water, sanitation and hygiene risk factors contribute annually to over 2 million deaths from diarrhea (cholera, typhoid) and over 1 million deaths from malaria.

Page 13: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

13

2.7. WATER QUALITY   Water quality is affected by chemical, microbiological and thermal pollution (Mayers et al., 2009;  UNEP,  2010):  1)  Chemical  contamination  can  occur  as  a  result  of  excess  nutrients, acidification16, salinity17, heavy metals and other trace elements18, persistent organic pollutants19 and  changes  in  sediment  loads. Anthropogenic  substances  can  enter  the  environment  through intentional,  measured  releases;  as  regulated  or  unregulated  industrial  and  agricultural  by‐products;  through  accidental  spills  or  leaks  during  the  manufacturing  and  storage  of  these chemicals; or as household waste. 2) Microbiological contaminants, bacteria, viruses and protozoa in water  pose  one  of  the  leading  global  human  health  hazards.  The  greatest  risk  of microbial contamination comes from consuming water contaminated with pathogens from human or animal 

16 Acidity of aquatic ecosystems (expressed as the pH value) determines their health and biological characteristics, disproportionately affecting young organisms. Most surface waters have a pH between 6 and 8.5; values below 6 can be hazardous to aquatic life. Lower pH can also mobilize metals from natural soils, e.g., aluminum, leading to additional stresses. Extensive acidification occurs downwind of power plants emitting nitrogen and sulfur dioxides, or downstream of mines releasing contaminated groundwater. Mining and power production can cause localized acidification of freshwater systems. Acid rain, a product of the interaction of emissions from fossil fuel combustion and atmospheric processes, can affect large regions (UNEP, 2010). 17 Freshwater plant and animal species typically do not tolerate high salinity, because it negatively impacts their metabolic function and oxygen saturation levels. Rising salinity can also alter riparian and emergent vegetation, affect the characteristics of natural wetlands and marshes, decrease habitat for some aquatic species, and reduce agricultural productivity and crop yields (Carr and Neary, 2008). Actions that can cause salinization include agricultural drainage from high-salt soils, groundwater discharge from oil and gas drilling or other pumping operations, industrial activities and certain municipal water-treatment operations. The chemical nature of the salts introduced by human activities may differ from those occurring naturally. 18 Trace elements, e.g., zinc, copper, arsenic and selenium, are naturally found in many waters. High natural concentrations of arsenic in groundwater are found in parts of Bangladesh and adjacent parts of India. Mining, industry and agriculture can lead to an increase in the mobilization of the trace elements out of soils or waste products into fresh waters. Even at extremely low concentrations, these materials can be toxic to aquatic organisms, impair growth, reproductive and other functions (UNEP, 2010). Heavy metals can accumulate in the tissues of humans and other organisms. Mercury and lead from industrial activities, landfills, commercial and artisanal mining also threaten human and ecosystem health in some areas. Emissions from coal-fired power plants are a major source of the mercury accumulating in the tissues of fish at the top of fish trophic levels (UNEP, 2008). Heavy metals also affect food quality, e.g., the preferential accumulation of cadmium in rice grain when effluent from zinc mines is used for irrigation. The United Nations Economic Commission for Europe finds that mining activities have severe impacts on water and the environment in eastern and South Eastern Europe, the Caucasus and central Asia. In some river basins, tailing dams for mining wastes also present substantial risks. 19 Diverse organic substances enter surface and groundwater through human activities, including pesticide use, industrial processes, and as degradation products of other chemicals (Carr and Neary, 2008). Numerous organic contaminants are persistent, bioaccumulative and toxic (PBT), are used globally and can be transported long ranges to regions where they have never been produced (UNEP, 2009b). For some compounds, non-lethal doses may be ingested by invertebrates and stored in their tissues. As larger organisms consume these prey species, PBT substances bioaccumulate, eventually to toxic levels. Certain pollutants break down over time, but degradation products may also be toxic. Pesticides leach into groundwater from agricultural runoff and urban landscapes. Insecticide DDT is banned in many countries, but still used for malaria control in Africa, Asia and Latin America (Jaga and Dharmani, 2003). It persists in the environment and resists complete microbial degradation. Even in countries where DDT is banned, it is still found in sediments, waterways and groundwater. Other PBT compounds are byproducts of industrial processes, e.g., dioxins, furans, and polychlorinated biphenyls (PCBs), which enter the environment through their use and disposal (UNEP, 1998). Oil and grease cause contact poisoning in aquatic organisms, coating and asphyxiation, behavioral changes, carcinogenic and mutagenic effects, e.g., petroleum products. Incomplete combustion of materials during manufacturing, incineration of waste or accidental fires produce polycyclic aromatic hydrocarbons and other mutagens and carcinogens which can leach into water (UNEP, 2008).

Page 14: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

14

feces.20  3)  Altering  natural  water  temperature  cycles  can  impair  biological  functions,  e.g., spawning, growth patterns, migration, and affect metabolic rates in aquatic organisms, leading to long‐term population declines. Warmer water dissolves less oxygen, impairing metabolic function and  reducing  fitness.  Thermal  impacts  are  severe  downstream  of  thermal  or  nuclear  power generation facilities or industrial activities (Carr and Neary, 2008).     Multiple  contaminants  often  combine  synergistically  to  cause  amplified,  or  different, impacts than the cumulative effects of pollutants considered separately (UNEP, 2010). Continued input of contaminants can ultimately exceed an ecosystem’s  resilience,  leading  to dramatic and irreversible losses. The level of anthropogenic pollution is a function of the structure of a country’s economy  and  its  institutional  and  legal  capacity  to  address  it.  Groundwater  systems  are particularly vulnerable  freshwater  resources: once contaminated,  they are difficult and costly  to clean. Based on data availability and relevance to trends in biodiversity and human well‐being, the following  three  parameters  of  water  quality  are  considered  to  be  particularly  useful:  nitrate concentrations,  biological  oxygen  demand  and  sediment  loads  in  rivers/turbidity  (UNEP,  2005). These three parameters are discussed in more detail below.  

Eutrophication  is a result of high‐nutrient  loads  (mainly phosphorus and nitrogen), which substantially  impairs beneficial uses of water. Major nutrient sources  include agricultural runoff, domestic  sewage,  industrial effluents and atmospheric  inputs  from  fossil  fuel burning and bush fires. Lakes and  reservoirs are particularly  susceptible  to  the negative  impacts of eutrophication because of  their complex dynamics,  relatively  longer water  residence  times and  their  role as an integrating sink for pollutants from their drainage basins (ILEC, 2005; Lääne et al., 2005). Eutrophic lakes  frequently  experience  noxious  algal  blooms,  increased  aquatic  plant  growth  and  oxygen depletion,  leading to degradation of their ecological, economic and aesthetic value by restricting use  for  fisheries,  drinking  water,  industry  and  recreation  (National  Research  Council,  1992). Nitrogen concentrations exceeding 5mg/L of water may indicate pollution from human and animal waste or fertilizer runoff from agricultural areas.    Three  macronutrients  (nitrogen,  phosphorus  and  potassium,  NPK)  and  numerous micronutrients  are  required  for  plant  growth.  Crop  production  often  requires  NPK supplementation.  N  and  P  can  travel  beyond  the  field  to  which  they  are  applied,  potentially affecting  ecosystems  offsite.  In  addition,  P  in  detergents  and  output  from  sewer  systems contributes to aquatic plant growth near global population centers. Here, the focus is on N and P as drivers of ecosystem changes; K  is not discussed, because  it  is relatively  immobile and benign offsite.  Consequences  of  increased  emissions  of NOx, NH3  and  dissolved  organic N  compounds include: decreased drinking water quality (Spalding and Exner, 1993), eutrophication of freshwater ecosystems  (McIsaac  et  al.,  2001),  hypoxia  in  coastal marine  ecosystems  (Rabalais,  2002).  The potential  health  effects  of  nitrates  in  humans  are  numerous: methemoglobinemia  (blue  baby 

20 Moreover, a number of pathogens are free-living in certain areas or represent IAS capable of colonizing a new environment, e.g., Vibrio bacteria and a few types of amoebas, which can cause major health problems: intestinal infections, amoebic encephalitis, amoebic meningitis (WHO, 2008); viruses and protozoa, including Cryptosporidium and Giardia, Guinea worm, etc (UNEP, 2010).

Page 15: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

15

syndrome), cancers,  thyroid disruptions and birth defects  (Carr and Neary, 2008). P compounds are  the  primary  cause  of  most  freshwater  eutrophication  and  a  component  of  estuarine eutrophication.  In  marine  systems,  excess  P  input  causes  harmful  algal  blooms21;  growth  of attached  algae  and  macrophytes  that  damage  benthic  habitats,  including  coral  reefs; deoxygenation and foul odors; fish mortality; and economic losses associated with increased costs of  water  purification  and  impairment  of  water  supply  for  agriculture,  industry  and municipal consumption. Global warming may further exacerbate toxic blooms of cyanobacteria, because at higher temperatures they have a competitive advantage over other algae (Bates et al., 2008).  

Organic materials from domestic wastewater treatment plants, food processing discharges and  algal  blooms,  are  decomposed  by  oxygen‐consuming microorganisms  in  water  bodies,  as measured by biochemical oxygen demand (BOD). Thermal stratification in nutrient‐enriched lakes with high BOD levels can produce conditions that allow nutrients and heavy metals in lake bottom sediments to re‐enter the water column. The eastern and southern coasts of North America, the southern  coasts  of  China  and  Japan  and  large  areas  around  Europe  have  undergone  oxygen depletion. One of  the world’s  largest hypoxic or dead zones has appeared off  the mouth of  the Mississippi River in the Gulf of Mexico, attributed to excessive N loads from the river, with harmful impacts on biodiversity and fisheries (Rabalais et al., 2002). Projected food production needs and increasing wastewater  effluents  associated with  an  increasing  population  over  the  next  three decades  suggest  a  10‐15%  increase  in  the  river  input  of  N  loads  into  coastal  ecosystems, continuing  the  trend  observed  during  1970‐95  (MA,  2005).  Human  sewage  and  wastewater treatment are also significant sources of methane and N2O, potent greenhouse gases.  

   There  are  two  primary  drivers  causing  changes  in  the  sediment  loads  of  the  rivers:  1) Catchment disturbance, which  results  in  increased  sediment  loads.  It  is  linked  to unsustainable land use practices, e.g., deforestation, land clearance for agriculture, mining, mineral exploitation, construction  and  infrastructural  development;  2)  Dam  construction, which  results  in  sediment trapping and  reduced  sediment  loads  (Walling, 2009). Climate change may also affect  sediment loads, but  the distinction between  climatic  and other  anthropogenic  contributors  is not  always straightforward. Excessive sediment loads in rivers disturb the life‐cycle of fish, through interfering with  breathing  and  impacting  spawning  areas.  Deposited  sediments  can  smother  benthic organisms. Sediments also disturb nutrient cycles in wetlands, especially estuaries. Fine sediments adsorb nutrients and toxins, e.g., P, lead and pesticides, altering water chemistry (Carr and Neary, 2008). Significant reductions  in natural  loads are equally harmful by reducing nutrient availability downstream (UNEP, 2005). From the standpoint of ecosystem services,  increased sediment loads also have implications for navigation and flood control. Reduced sediment input to a delta results in shoreline retreat, loss of land and increased coastal flooding (Walling, 2009).  

3. Possible future developments in the domain of the driver 

21 Cyanobacteria, also known as blue-green algae, have increased in freshwater and coastal systems, e.g., the East China Sea. The toxins produced by the excessive algal blooms are concentrated by filter-feeding bivalves, fish and other marine organisms. In humans, they can cause acute poisoning, skin irritation and gastrointestinal illnesses (e.g., shellfish poisoning).

Page 16: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

16

 3.1. ECOSYSTEM SERVICES 

Simulations by Alcamo et al.  (2005) utilize  the  four MA  scenarios  (Global Orchestration, Order  from  Strength,  Adapting  Mosaic  and  TechnoGarden22)  to  estimate  future  worldwide ecosystem services. Results  indicate that  if agricultural practices are  intensified, the side effects, e.g.,  contamination  of  groundwater  or  N  transport  to  coastal  zones,  can  limit  the  ecosystem services provided by groundwater, surface waters and the coastal zone, e.g.,  fishery, recreation. The scenarios project a much higher delivery of  freshwater services to households and  industry, especially in developing countries. The development of water infrastructure and delivery of clean water  to  households  in  developing  countries  can  significantly  reduce  water‐related  diseases. However,  increased withdrawals  can  lead  to  large  increases  in wastewater,  especially  in  Latin America (factor of 2 to 4) and Sub‐Saharan Africa (factor of 3.6 to 5.6) between 1995 and 2050. If wastewater  is discharged untreated  into  surface waters  and  groundwater,  it will pose  a  risk of contamination for society and reduce the habitat suitable for freshwater fish and other organisms.  

 Higher  income makes  it possible to deliver freshwater services to many more households 

and  industries  in  the  future, but a  consequence of  increasing water withdrawals  is a  significant intensification of water stress  throughout  the developing world. Authors also point out  that  the global demand for fish grows, but it is questionable whether marine fisheries can keep up with this demand, even with an increase in aquaculture‐produced fish. Ecological limits make an increase in production unlikely in two out of the three important marine fisheries. Overall, these simulations raise  a  question  about  the  reliability  of  future  ecosystem  services.  They  do  not  indicate  the likelihood that thresholds or breaking points will be reached.  

Within  the  MA  scenarios,  the  highest  per  capita  demand  levels  for  food  fish  (17.3 kilograms) and production of 128 million tons by 2020 occur under Global Orchestration, due to rapid  increases  in urbanization and  income growth. Under Order from Strength, rapid population growth combined with slower economic progress result in the lowest production increases, at 117 millions tons in 2020, corresponding to depressed per capita demand of 14.8 kilograms. The values for the other MA scenarios fall between these two extremes.  

An  improved  understanding  of  the  relationships  between  ecosystems  and  human well‐being  is needed  to  facilitate  sustainability. Further  research  is needed  to better understand  the connections among the production of multiple ecosystem services, the local and global impact of 

22 Millennium Assessment considers the following four scenarios: 1) Global Orchestration, describing a globally connected society that focuses on global trade and economic liberalization. This society takes a reactive approach to ecosystem problems, but also takes strong steps to reduce poverty and inequality and to invest in public goods, e.g., infrastructure and education. 2) Order from Strength depicts a regionalized and fragmented world, concerned with security and protection, which emphasizes regional markets, paying little attention to public goods, and takes a reactive approach to ecosystem problems. 3) Adapting Mosaic chooses regional watershed-scale ecosystems as the focus of political and economic activity. Local institutions are strengthened and local ecosystem management strategies are common. Societies develop a strongly proactive approach to the management of ecosystems. 4) TechnoGarden is a globally connected world relying strongly on environmentally sound technology. This world uses highly managed, often engineered ecosystems to deliver ecosystem services, and takes a proactive approach to the management of ecosystems in an effort to avoid problems (Alcamo et al., 2005).

Page 17: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

17

ecological  processes,  and  the  determinants  of  ecological  resilience  (Alcamo  et  al.,  2005).  Our current  lack  of  knowledge  concerning  the  resilience  of  ecosystem  services makes  it  difficult  to assess the degree of concern. If ecosystems are relatively robust, it is possible that current trends may not greatly alter the provision of the more vital ecosystem services. By contrast, if ecosystems are relatively fragile and if the relationship between ecological impacts and ecosystem services is nonlinear,  cumulative  human  impacts  will  some  day  push  ecosystems  over  one  or  more thresholds, resulting in the collapse of a bundle of ecosystem services (Peterson et al., 2003). The true  state  probably  lies  between  these  two  extremes  and  will  differ  for  different  ecosystem services. There are high levels of uncertainty concerning the relative magnitude of human impacts on ecosystems. Rates of habitat destruction and species extinctions are higher than they have ever been  in  the history of humanity  (McNeill, 2000). Ecosystem services may be  intricately  linked  to one another in surprising or unforeseen ways, e.g., deforestation can lead to the unexpected rise of a saline water table, severely affecting food production (Keating et al., 2002). 

 The  type  of  ecosystem  change  that  is  of  greatest  concern  involves  regime  shifts, which 

occur when an entire  system  flips  into an alternative  stable  state  (Scheffer et al., 2001). These changes  are  a  strongly  nonlinear  (accelerating,  abrupt  and/or  irreversible)  response  to  gradual changes  in  the  variables  that define  a  system’s  stable  state. Regime  shifts  in ecosystems  cause rapid,  substantial  changes  in  ecosystem  services.  Ecological  surprises  are  unexpected  regime shifts, e.g. degradation of  fisheries, coral reefs, quality of  freshwater, emergence of disease and spread of nuisance species. Carpenter (2006) offers a reading list of ecological and socio‐ecological case  studies of eco‐surprises.  Increasing pressure on ecosystems will  increase  the  frequency of regime  shifts  that  affect  ecosystem  services  and  human  well‐being.  Ecological  feedbacks may accentuate  anthropogenic  modifications  of  ecosystems.  Changes  in  ecological  functioning produced by unintended ecological feedbacks from human actions appear likely to amplify climate change, decrease agricultural productivity, reduce human health, and increase the vulnerability of ecosystems to IAS (MA, 2005).  Research needs in the domain of ecological surprises include: data‐based risk analysis for thresholds that can be quantified, e.g., freshwater eutrophication, drought persistence  in  drylands;  assessment  of  leading  indicators  that  are  grounded  in  empirical observation; practical, empirically confirmed guidelines  for building  resilience  in sensible day‐to‐day decision‐making (Carpenter, 2006). 

 3.2. WATER RESOURCES 

The primary direct drivers of degradation and loss of rivers, lakes, freshwater marshes and other  inland  wetlands  include  infrastructure  development  (e.g.,  dams),  land  conversion  (e.g., deforestation),  water  withdrawal  (e.g.,  irrigation),  pollution  (e.g.,  excess  nutrients), overexploitation  (e.g., overfishing, overgrazing) and the  introduction of  IAS. The primary  indirect drivers of degradation and loss are population growth and increasing economic development. The primary direct driver of the loss and degradation of coastal wetlands, including saltwater marshes, mangroves,  seagrass meadows  and  coral  reefs  is  conversion  to  other  land  uses.  Other  direct drivers affecting coastal wetlands include diversion of freshwater flows, N loading, overharvesting, siltation, changes  in water  temperature and  IAS. The primary  indirect drivers of change are  the population growth in coastal areas coupled with growing economic activity (MA, 2005).  

Page 18: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

18

Most river‐related scenarios and projections for water use and water scarcity are obtained from global models that utilize a combination of climate and elevation data. These models do not supply  the  level of detail  required  to establish management options, assess  threats or evaluate tradeoffs. More  reliable  information  on  actual  stream  and  river  discharge,  and  the  amount  of water withdrawn  and  consumed  at  the  river  basin  level, would  increase  our  ability  to manage inland  water  ecosystems more  efficiently,  evaluate  tradeoffs  between  different  uses,  and  set conservation measures.  

Lakes undergo seasonal changes and are continually modified through  tectonic events or volcanic activity, landslides, erosion or wind. Understanding of lake water movement and potential productivity requires calculations of the  lake depth and water volume. Estimates are often used, because the calculations require detailed knowledge of the lake bottom. Revenga and Kura (2003) specify the most common threats affecting  lake ecosystems: toxic pollution, eutrophication,  IAS, water level change and siltation. 

 Wetland  ecosystem  processes  are  driven  by  the  complex  coupling  of  climatic,  soil  and 

biotic  factors, which change quickly over  space and  time.  Interactions between  these processes promote  the  persistence  of  rare  indigenous  species  and  the maintenance  of  critical  ecosystem functions  that would disappear with  the  loss of  the ephemeral and seasonal wetlands  (Revenga and Kura, 2003). Sea  level rise may affect a range of freshwater wetlands  in  low‐lying regions.  In tropical regions, low lying floodplains and swamps may be displaced by salt water habitats due to the combined actions of sea level rise, more intense monsoonal rains and larger tidal/storm surges (Bayliss et al., 1997). This may dislocate or displace wetland flora and fauna. Flora not tolerant to increased salinity or  inundation may be eliminated. Salt‐tolerant mangrove species could expand from nearby coastal habitats. Migratory and resident fauna, e.g., birds and fish, may lose resting, feeding  and  breeding  grounds.  Fish migrations may  be  affected  by  both  temperature  and  flow patterns. The most apparent faunal changes may occur with migratory and nomadic bird species that use wetland habitats across or within continents (Walther et al., 2002). Changed rainfall and flooding  patterns  across  arid  land may  adversely  affect  bird  species  that  rely  on  a  network  of wetlands and lakes that are alternately or episodically wet and fresh or dry and saline (Roshier et al., 2001). The wetland carbon sink may also be affected by climate change, although the direction of the effect is uncertain due to the number of climate factors and possible responses. 

 Under  the  MA  scenarios  Order  from  Strength  and  Adapting  Mosaic,  wetland  loss  is 

relatively slow in the short term, for different reasons. Reduction in international trade and direct investments  limits  new  large‐scale  water  schemes.  Within  Adapting  Mosaic,  small‐scale  and integrated water management projects are developed, which releases some pressure on wetlands. This  is  not  the  case  in  Global  Orchestration.  In  TechnoGarden,  technical  improvements  are expected. Some wetlands are allowed to decline, because their ecosystem services are provided by  technological alternatives.   For  the  second half of  the period until 2050,  in both Order  from Strength and Global Orchestration, there is a long‐term increase of conversion to agricultural land use. For TechnoGarden and Adapting Mosaic,  the  technology and ecosystem management  skills induce  a  restoration  of wetlands.  Climate  change may  put  further  pressure  on wetlands  in  all scenarios. Sea level rise is expected to lead to loss of coastal wetlands, e.g., estuaries and deltas. 

Page 19: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

19

This  effect  is  most  pronounced  in  Global  Orchestration,  Order  from  Strength,  and  Adapting Mosaic, where it may overcompensate for the effects of learning.  

 Based on  the  IPCC scenarios modified  in Ramsar COP8‐Doc.11, by 2025 coastal wetlands 

will experience loss due to sea level rise. Erosion of shorelines increases. The loss of wetlands and shore erosion is more extensive by 2100. 

 3.3. FLOODS AND DROUGHTS 

Examination  of  evidence  for  the  intensification  of  the  global water  cycle  by Huntington (2006) reveals upward trends for the following variables: precipitation, runoff, tropospheric water vapor,  soil  moisture,  glacier  mass  balance,  evapotranspiration,  growing  season  length  and droughts, but no change in cloudiness, floods, tropical storm frequency and intensity. Substantial uncertainty  regarding  trends  in  hydro‐climatic  variables  remains  because  of  differences  in responses among variables and regions and major spatial and temporal limitations  in data. There are large gaps in spatial coverage for some hydrologic indicators; therefore, trends in these areas are unknown. 

 Van Lanen et al.  (2009)  stress  the  importance of processes underlying  the generation of 

hydrological extremes for computations of flood and drought characteristics and trend prediction. Floods can be generated through  intense and  long‐lasting rainfall, snowmelt or flow obstruction, e.g., ice jam or landslide. Droughts can be caused either by low precipitation often combined with high evaporation losses, or result from precipitation being stored as snow. The authors also review the  physical  indices  for  large  scale  floods  and  for  soil moisture,  groundwater  and  streamflow droughts  (Van  Lanen  et al., 2009). Climate  change  is expected  to primarily  affect precipitation, temperature and potential evapotranspiration, and  is  likely to affect the occurrence and severity of meteorological  droughts.  Changes  in meteorological  drought  affect  soil  water  drought  and hydrological drought, i.e., groundwater and streamflow drought. Soil water drought is relevant for agriculture, terrestrial ecosystems, and health through the occurrence of heat waves. Hydrological drought  impacts water resources, namely agriculture, domestic and  industrial water use, aquatic ecosystems, power generation and navigation (Van Lanen et al., 2007).  

Documented trends in floods show no evidence for a globally widespread change, despite increased media coverage, flood costs and flood risks. One study identified an apparent increase in the  frequency of  large  floods  (exceeding 100‐year return period  levels)  in 16  large basins across much of  the globe during  the 20th  century  (Milly et al., 2002). A global  change detection  study does not support the hypothesis of a global increase of annual maximum river flows23 (Kundzewicz et al., 2005).  

 

23 The study finds increases in 27 cases, decreases in 31 cases and no trend in the remaining 137 cases of the

195 catchments examined worldwide. However, the overall maxima for the period (1961-2000) occur more frequently (46 times) in the later subperiod, 1981-2000, than in the earlier subperiod (24 times), 1961-80. Evidence is stronger for changes in the timing of floods, with increasing late fall and winter floods.

Page 20: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

20

Summaries of observed and predicted  impacts of climate change on hydrologic droughts are  available  (Huntington,  2006; Van  Lanen  et  al.,  2007;  Zhang  et  al.,  2007).  In  the  past  three decades,  droughts  have  become more widespread, more  intense  and more  persistent  due  to decreased  precipitation  over  land  and  rising  temperatures,  resulting  in  enhanced evapotranspiration and drying. The occurrence of droughts is determined largely by changes in sea surface  temperatures, especially  in  the  tropics,  through  changes  in  atmospheric  circulation  and precipitation. In the western United States, diminishing snow pack and resultant reductions in soil moisture also appear to be factors. In Australia and Europe, the extremely high temperatures and heat waves accompanying recent droughts imply direct links to global warming (Bates et al., 2008). 

 The  IPCC scenarios modified  in Ramsar COP8‐Doc.11 predict an  increase  in  flood damage 

by 2025 due  to more  intense precipitation events. By 2100,  the  flood damage becomes  several times higher than in a “no climate change” scenario. Increased drought frequency is predicted for the period up to 2025, followed by further increase in drought events and their impacts by 2100. 

 Future  projections  of  drought  in  the  21st  century  made  using  the  Special  Report  on 

Emissions Scenarios (SRES, Arnell, 2004) A2 emission scenario show regions of strong wetting and drying with a net overall global drying trend. This  increase continues throughout the 21st century and by the 2090s the percentage of the  land area  in drought  increases to 30%, 40% and 50% for extreme,  severe  and moderate  drought,  respectively.  In  all  cases,  as  the  severity  of  drought increases,  the number of drought events and  the duration of each decreases.  In  the  future,  the number of extreme and severe drought events is projected to double; for moderate drought, the number of events remains stable. There is a significant increase in the mean event duration for all forms of drought. The  spatial distribution of changes  in  the PDSI over  the 21st century was also explored. The model predicts drying over Amazonia, the United States, northern Africa, southern Europe  and western  Eurasia,  and wetting  over  central  Africa,  eastern  Asia  and  high  northern latitudes (Burke et al., 2006). 

 EU  Water  and  Global  Change  (WATCH)  Integrated  Project  Work  Block  titled  “Likely 

Frequency, Severity and Scale of Future Extremes” will provide an analysis of droughts and large‐scale floods based on the global and regional scenarios for the 21st century as compared with the simulations for the 20th century. The Work Block report was not available by the deadline for this report. 

 The author of this report has found no global scenarios specifically addressing  impacts of 

floods and droughts on aquatic ecosystems and  their services.  It  is possible  to develop  them by considering  flood and drought characteristics  relevant  for  the provision of  services. Elements of flood  and  drought  profiles  that  affect  ecosystem  services  include:  season/period;  frequency; duration; flood depth and flood water quality. Ecosystem services that are likely to be impacted by flooding  and  droughts  are:  biodiversity  maintenance;  hydrological  functioning;  carbon sequestration  and  nutrient  cycling.  For  drought  conditions,  Huxman  et  al.  (2004)  propose  an evaluation of the ecosystem water use efficiency. Surprisingly, the low‐precipitation regions, e.g., deserts,  are much more  sensitive  to  changes  in  precipitation  than  expected. Moreover,  during drought, total plant growth per unit of precipitation for tropical ecosystems approaches that of the 

Page 21: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

21

most  arid  areas.  For  any  ecosystem,  the  authors’  model  predicts  substantial  reductions  in productivity when rain falls below historical minimums.   

3.4. BIODIVERSITY AND INVASIVE SPECIES Future extinction rates are projected  to be  five  times higher  for  freshwater animals  than 

for  terrestrial  species  (Ricciardi and Rasmussen, 1999).  In addition  to habitat  loss, water quality and IAS, the following major threats to the biodiversity of freshwater ecosystems are identified by Revenga and Kura (2003): modification of river systems24, water scarcity25, the condition of inland water fisheries26 and climate change.27 

MA  (2005)  offers  scorecards  for  biodiversity  developments  until  2050,  considering  N‐S deposition, climate,  IAS and  land use. TechnoGarden and Adapting Mosaic scenarios are  likely to see  a  decrease  in  the  rate  of  biodiversity  loss.  In  TechnoGarden,  this  is  due  to  a  significant reduction in land use change. In Adapting Mosaic, it is due to a reduction in IAS. Movements of IAS between countries also are reduced  in Order from Strength, but the number of outbreaks within poorer countries  is higher and  the overall chance of  IAS  reaching  richer countries  is high.  In  the longer  term,  TechnoGarden  bears  the  risk  of  significant  technological  failures  that  can  lead  to outbreaks  of  new  pests  and  diseases.  The worst‐case  curve  shows  a  peak  in  biodiversity  loss midway  through  the period, which  flattens out and declines after appropriate countermeasures are developed. In Adapting Mosaic, failed experiments and climate change might also increase the 

24 Modifications of river systems aim to improve transportation, provide flood control, hydropower, more land and irrigation water, thereby boosting agriculture. Modifications include river embankments to improve navigation, drainage of wetlands for flood control and agriculture, construction of dams and irrigation channels, and the establishment of inter-basin connections and water transfers. The physical changes in the hydrological cycle disconnect rivers from their floodplains and wetlands and slow water velocity in riverine systems, converting them to a chain of connected reservoirs. This may impact the migratory patterns of fish species, alter the composition of riparian habitats, open up paths for exotic species, change coastal ecosystems and contribute to an overall loss of freshwater biodiversity (Revenga et al., 2000). Dams also affect the seasonal flow and sediment transport of rivers for an average of 160 kilometers downstream (Revenga and Kura, 2003). There are more than 45,000 large dams and an exponentially larger number of smaller dams on the world’s rivers. 25 Widespread depletion and pollution of surface and groundwater sources is expected to contribute to the global biodiversity loss. Diversions of river flow for irrigation and agricultural runoff cause a decline in species richness and a disappearance of valuable wetland habitats (Revenga and Kura, 2003). 26 Inland fisheries from rivers, lakes and other wetlands are a major source of protein for a large part of the world’s population, particularly the poor. Globally, inland fisheries landings increased at 2% per year from 1984 to 1997. In Asia, the rate is higher: 7% per year since 1992. This increase is a result of: 1) efforts to raise production above natural levels through fisheries enhancements (e.g., stocking and aquaculture) and 2) eutrophication of inland waters (FAO, 1999). Aquaculture practices may contribute to habitat degradation, pollution, introduction of IAS and the spread of pathogens (Naylor et al., 2000). 27 Warming of rivers will affect chemical and biological processes, reduce the amount of ice cover, reduce the amount of dissolved oxygen in deep waters, alter the mixing regimes and affect the growth rates, reproduction and distribution of organisms and species (Gitay et al., 2001). Species in small rivers and lakes are thought to be more susceptible to changes in temperature and precipitation than those in large rivers and lakes. Fish species distribution is expected to move towards the poles, with cold water fish being further restricted in their range, and cool and warm water fish expanding in range. Aquatic insects are less likely to be restricted given their aerial life stages. Less mobile fish and molluscs may be at higher risk due to their presumed inability to keep pace with the rate of change in freshwater habitats (Gitay et al., 2001). Warmer weather conditions are expected to facilitate the establishment of IAS. At high latitudes, warming is expected to increase biological productivity, whereas at low latitudes the boundaries between cold and cool-water species may change and possibly lead to extinctions (IPCC, 1996).

Page 22: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

22

rate of biodiversity loss in the longer term, albeit not as abruptly. The other two scenarios present a possibility of  reducing  the  rate of  loss  on  longer  terms, which  is more pronounced  in Global Orchestration than  in Order from Strength, and mainly  in richer countries. Climate change might counteract the optimistic outlook for both Order from Strength and Global Orchestration. 

 Xenopoulos et al. (2005) build global scenarios of  losses of riverine fish biodiversity. They 

consider the known relationships between fish species and river discharge and combine them with a  global  hydrological model  and  two  scenarios  from  the  Intergovernmental  Panel  for  Climate Change.  In  rivers with  reduced discharge, up  to 75% of  local  fish biodiversity would be headed toward extinction by 2070 because of combined changes in climate and water consumption. If fish are adapted to particular combinations of discharge and other habitat features, e.g., floodplains, stream gradient,  substrate,  riparian  vegetation, which become decoupled under various human influences, then many more species will tend toward extinction. Other anthropogenic  influences are likely to increase species loss above those reflected in the species‐discharge model: industrial pollution, eutrophication, physical modifications of rivers. Freshwater taxa are especially sensitive to  the  hydrologic  changes  indirectly  related  to  increased  CO2  emissions.  Fish  loss  falls disproportionately on poor countries. Authors predict that reductions  in water consumption can prevent many of the extinctions in these scenarios. 

 Sala et al.  (2000) suggest  that  relative  to  terrestrial biomes,  freshwater  taxa are  likely  to 

suffer more  from  land use  changes  and  IAS  than  from  climate,  increased CO2  emissions  and N deposition.  Land  use  is  expected  to  have  especially  large  effects  because  humans  live disproportionately  near  waterways  and  extensively modify  riparian  zones,  inputting  nutrients, sediments and contaminants. Biotic exchange  is  important because of extensive  intentional and unintentional  releases  of  organisms.  The  authors  build  three  scenarios  to  estimate  the  total change  in biodiversity  to 2100, assuming no  interactions, antagonistic or synergistic  interactions among the drivers of change. Scenarios are not applied to freshwater biomes. 

 3.5. DEFORESTATION 

Total  global  forest  area  continues  to  decrease,  but  the  rate  of  net  loss  is  slowing. Deforestation continues at a rate of 13 million ha per year. Forest planting, landscape restoration and natural expansion of  forests have  significantly  reduced  the net  loss of  forest area. The net change in forest area in the period 2000‐2005 is estimated at ‐7.3 million ha per year, down from ‐8.9 million ha per year  in  the period 1990‐2000. Africa and South America continue  to have  the largest net loss of forests. Oceania, North and Central America also have a net loss of forests. The forest area  in Europe continues  to expand, although at a slower  rate. Asia  reports a net gain of forests in the period 2000‐2005 after a loss during 1990‐2000 (FAO, 2005).  

The four MA scenarios indicate that a total of 10‐20% of current grassland and forest area (warm mixed forests, tropical woodlands and tropical forests) may be lost by 2050, mostly because of  the expansion of agricultural  land  (Alcamo et al., 2005). The global  trends are a net  result of slow  reforestation  in  the  temperate zones and strong deforestation  in  the  tropical zones. Order from Strength exhibits  the  fastest  rate of deforestation at  the beginning of  the scenario period, increasing  from  the historic  annual  rate of 0.4%  to 0.6%.  In Global Orchestration  and Adapting 

Page 23: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

23

Mosaic, the rate of loss of undisturbed forests is at the historic rate, and slightly below in Techno Garden.  In  industrial  regions,  forested  areas  generally  expand  from  2000‐2050, with  the most growth expected under Global Orchestration and Techno Garden. Forests  in developing  regions decline, with the largest loss observed under Order from Strength.28  

 According to Bates et al. (2008), reduced deforestation conserves water resources, reduces 

runoff, controls flooding, erosion, salinity and siltation, protects fisheries and hydroelectric power facilities. Currently,  there are no global scenarios exploring  the relationship between  the  forests and water resources. However, regional scenarios are available.29 A set of scenarios for Melbourne catchments considers bushfires along with a variety of clearing and thinning schemes to 2030. The Macaque model30 expects  the water yield  to  increase due  to  the natural aging of  the  forests as they become  re‐established after bushfires, even  if  there are no changes  to  the current  timber‐harvesting regime (Battad et al., 2007; DSE, 2008). Impacts of forest disturbances on water quality are examined by Feikema et al. (2005) and Poynter and Featherston (2008).31 

 

28 Overall, rapid depletion of forest areas continues under the Order from Strength scenario, while TechnoGarden expects an increase in net forest cover. Production of biofuel, particularly under the TechnoGarden scenario, is an important category of land use, especially in former Soviet countries, OECD and Latin America. The coverage of energy crops remains small, but it has a large influence on the trends in land use. Order from Strength continues to increase agricultural areas in sub-Saharan Africa and Latin America, due to a fast population growth and a limited potential to import food. The depletion of forests continues at a rate near the historic average, and slows down after 2050, because of slowing population growth. By 2050, 67% of the central African forest present in 1995 disappear. For Asia and Latin America, these numbers are 40% and 25%, respectively. In other regions, the rate of forest loss declines. 29 Odada et al. (2009) present plausible trends in land use change and resulting changes in lake levels and contribution of fisheries in the Lake Victoria basin across four scenarios to 2025. Scenarios with a low level of regional integration are the No Development Scenario (NDS) and Worst Case Scenario (WCS). The Current Practices Scenario (CPS) and Best Case Scenario (BCS) assume a high level of regional integration, including specific social, economic and environmental policies to nurture it. Increases in agricultural and urban land uses are observed in all scenarios, accompanied by a decline in forest land in all but the BCS. Deforestation is heaviest in WCS and lowest in BCS, occurring mainly in the frontier highlands of the basin, while extensive degradation under WCS and CPS centers around marginal lands closer to the lake. Water level variation shows a negative height variation trend. The key drivers propelling the water level trends are precipitation and other climate elements, agricultural land use, industrialization along the lake and energy demands. The challenges associated with WCS, which are countered in BCS with sustainable transboundary ecosystem management include: rapid population explosion; vulnerability to local and external environmental and economic shocks; unprecedented spread of AlDS, malaria; emergence of new diseases and re-emergence of old ones; and escalating poverty. 30 The Macaque model simulates physical processes within the catchment by examining variables associated with topography, vegetation, precipitation and temperature. It describes the relationship between water yield and time since disturbance, while considering variations in forest type and forest age. 31 The best management practices to address the potential impacts of timber harvesting on water quality include measures that minimize sediment generation and flow from unsealed roads and tracks. By comparison, sediment generation and flow from the harvested area are less significant and easier to manage (Poynter and Featherston, 2008). The impact of bushfires on the water quality in rivers is examined by Feikema et al. (2005). The study constructs and parameterizes models using the E2 catchment modeling framework to represent the flow, sediment and nutrient loads for the water storages and river systems in fire-affected catchments in eastern Victoria, Australia. The fire-affected catchments are predicted to deliver on average 27 times greater total suspended sediment loads, 4.9 times greater total nitrogen, and 7.9 times the amount of total phosphorus compared to pre-fire conditions. Sediment storage, adsorbed nutrients and recovery of the fire-affected areas are not represented in the model; therefore, observed increases in loads are likely to be lower.

Page 24: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

24

3.6. WATER AVAILABILITY, USE AND PRODUCTIVITY   The focus here is on water availability, use, efficiency and productivity development guided by  the  following  indicators.  Water  availability  indicates  the  volume  of  water  theoretically exploitable.  Water  withdrawals  indicate  the  volume  used  by  society  to  fulfill  its  domestic, industrial  and  agricultural  needs. Water  productivity  indicates water  availability  for  uses  other than agriculture. The criticality ratio  is an  indicator of water scarcity stress at the basin  level and the  intensity  of  human  water  use.  Water  stress  is  defined  as  the  long‐term  average  annual withdrawals  to availability ratio  (Alcamo and Henrichs, 2002).32 The change  in water stress  is an indirect  indicator of  the ability of a  freshwater ecosystem  to deliver ecosystem  services. Return flows roughly indicate the magnitude of wastewater discharged into the receiving water in a river basin. 

 Rosegrant  et  al.  (2002)  discuss  three  scenarios  projecting  the  likely  water  and  food 

outcomes  to  2025:  the business‐as‐usual  (BAU),  crisis  (CRI)  and  sustainability  (SUS)  scenarios.33 Under BAU, total global water withdrawal in 2025 is projected to increase 22% above 1995 levels to  4,772  km3.  This  is  consistent with  other  recent  projections  to  2025  including: Alcamo  et  al. (1998) medium  scenario  of  4,580  cubic  km3,  Seckler  et al.  (1998)  business‐as‐usual  scenario  of 4,569 km3 and Shiklomanov  (1999)  forecast of 4,966 km3  (excluding  reservoir evaporation). The Global  Orchestration  scenario  shows  strong  economic  growth  coupled  with  an  increase  in population, which leads to a worldwide increase in withdrawals of around 40% by 2050. For OECD, MENA  and  FSU  only  slight  increases  in  withdrawals  are  predicted,  because  economic  and population  growth  are  compensated  by  improving  water  efficiency.  By  2050,  TechnoGarden experiences strong structural changes in the domestic and industrial sectors and improvements in the  efficiency  of water  use  in  all  sectors, which  lead  to  a  net  decrease  in water withdrawals. Adapting Mosaic and Order from Strength do not have the largest economic growth, but by 2050, 

32 Water stress is severe if the ratio is greater than 0.4; medium for ratios between 0.2 and 0.4; and low for ratios below 0.2. Alcamo and Henrichs (2002) identify critical regions in which water resources have higher sensitivity to global change than other regions, using an increase in water stress as a measure of watershed sensitivity. Stress increases when water withdrawals increase or water availability decreases. The extent of computed critical regions depends upon the scenario and the set of criteria for determining critical regions. Scenarios considered are: Markets First, Policy First, Security First and Sustainability First. Certain regions appear to be critical under all scenarios, i.e., central Mexico, the Middle East, large parts of the Indian subcontinent, and the northern coast of Africa. Oki et al. (2001) derive annual water availability from annual runoff estimated by land surface models using total runoff integrating pathways. The total population under water stress estimated for 1995 is consistent with earlier estimates. This number is highly dependent upon assumptions about the volume of water from upstream of a region, which is considered “available” water within the region. Therefore, it is necessary to evaluate the upstream regional water quality and the real consumption of water resources, as well as the accessibility of water. 33 The business-as-usual scenario (BAU) projects the likely water and food outcomes for a future trajectory based on the recent past, whereby current trends for water investments, water prices, and management are broadly maintained. The water crisis scenario (CRI) projects deterioration of current trends and policies in the water sector. The sustainable water use scenario (SUS) projects improvements in a wide range of water sector policies and trends (Rosegrant et al., 2002). An analogous set of three global scenarios has been prepared for World Water Vision (Gallopín and Rijsberman, 2000). The alternative scenarios are: the Business-as-Usual (BAU), representing the future trajectory if no major policy or lifestyle changes take place; the Economics, Technology and Private Sector (TEC), which relies on the market, the involvement of the private sector and technological solutions, and predominantly national/local or basin-level action; and the Values and Lifestyles scenario (VAL), that materializes through a revival of human values, strengthened international cooperation, heavy emphasis on education, increased solidarity and changes in lifestyles and behavior.

Page 25: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

25

they  do  have  larger  withdrawals  than  TechnoGarden.  They  show  slow  improvement  in  the efficiency of water use and fast population growth. Under CRI, unregulated and illegal withdrawals increase. After 2010, key aquifers begin to  fail; declining water tables make extraction costs too high for continued pumping, causing a big drop in groundwater extraction, further reducing water availability  for  all  uses.  Under  SUS,  withdrawals  decrease,  leaving  more  water  in‐stream  for environmental purposes.  Investments  result  in  improved water use efficiency. Under CRI, water withdrawals  result  in sharply  reduced environmental committed  flow  in 2025, 460 km3 globally. SUS results in higher environmental committed flows, with a global increase of 1,030 km3 (Alcamo et al., 2003; Rosegrant et al., 2002). 

 Global Orchestration predicts 18% of  the world’s  river basins are  in  severe water  stress, 

with  60%  of  the  global  population  residing  in  these  regions  by  2050,  or  4.9  billion  people compared to current 2.3 billion. Under Adapting Mosaic and Order from Strength, the population increases to 5.3 and 5.5 billion, respectively. Under BAU, the criticality ratio increases globally from 0.08 in 1995 to 0.10 in 2025. Under SUS, allowable groundwater pumping in 2025 falls to 594 km3 representing a decline of 11% from the value in 1995 and a drop from the 2025 BAU value of 23%. Criticality  ratios  under  CRI  result  in  significantly  higher  ratios  in  2025  than  1995,  while  SUS maintains 1995 levels. The ratio is relatively low globally and for large aggregated regions because abundant water in some countries masks scarcity in others, but the ratio is far higher for individual dry regions. It is in these dry regions the impact of SUS is most beneficial (Rosegrant et al., 2002). 

 By 2025,  total worldwide water consumption under CRI  is 13%  (or 261 km3) higher  than 

that under BAU, while under SUS it is 20% (or 408 km3) lower. This reduction in consumption frees water for environmental uses. Under SUS, industrial water demand declines compared with BAU. Technological improvements in water use and recycling and increased water prices, which induce reductions  in demand, are  the cause. Under CRI, with weakened  incentives and  regulations and lower investment in technology, industrial water demand increases compared with BAU, as more water  is needed  to produce a unit of output.  In 2025,  total worldwide  industrial water demand under CRI is 80 km3 (or 33%) higher than under BAU, while it is 85 km3 (or 35%) lower under SUS. Under CRI, global  industrial water use  intensity  is 1.2 m3 per thousand U.S. dollars higher under CRI, and 1.3 m3 per thousand U.S. dollars lower under SUS (Rosegrant et al., 2002). By 2050, most scenarios  predict  an  increase  in  efficiency  with  a  net  water  availability  increase  in  nearly  all regions. Notable exceptions  are MENA under BAU  and CRI.  From  the perspective of ecosystem services, these results imply an increase in services delivered by freshwater systems.   

A year‐to‐year variation  in agricultural water productivity from 1995–2025  is  likely caused by climate variability, which affects water availability and thus water productivity. The change of water consumption per hectare  is small compared with the change of crop yield, which  indicates the  increase of water productivity results mainly from  increases  in crop yield. Water productivity of  irrigated crops  is higher than for rainfed crops  in developing countries. Water productivity for rainfed  crops  is  higher  than  for  irrigated  crops  in  developed  countries  for  both  rice  and  other cereals, which indicates favorable rainfall conditions for crop growth (Rosegrant et al., 2002).  

Page 26: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

26

By 2025, water consumption under CRI is 13% higher than under BAU, while under SUS it is 20%  lower.  This  reduction  in  consumption  frees water  for environmental uses. Virtually  all  the difference  in water consumption between  the CRI and BAU  scenarios  is  in  the  irrigation  sector. Under  CRI,  declining  water  use  efficiency  causes  higher  losses  through  non‐beneficial  water consumption  and  greater  water  withdrawals.  Compared  with  BAU,  the  2025  global  water withdrawal  is  10%  higher  under  CRI,  while  under  SUS,  it  is  22%  lower.  Under  CRI,  beneficial irrigation water consumption is 8% lower than BAU in the developed world and 16% lower in the developing world. Under  SUS,  total  irrigation  consumption declined, but higher basin efficiency brought irrigation consumption close to BAU levels. Beneficial irrigation consumption under SUS is 2% lower than BAU in the developed world and 9% lower in developing countries (Rosegrant et al., 2002). 

Low  Groundwater  pumping  scenario  (LGW)  assumes  that  groundwater  overdraft  in  all countries and regions using water unsustainably is phased out over 25 years beginning in 2000 by reducing annual groundwater pumping‐to‐recharge  ratios  to below 0.55 at  the basin or country level. Under  LGW, pumping  in  the overdrafting  countries and  regions declines by 163 km3. The projected increase in pumping for areas with more plentiful groundwater resources remains about the  same  as  under  BAU.  Total  global  groundwater  pumping  falls  to  753  km3  in  2021–2025,  a decline from the 1995 value of 817 km3 and from the 2021–2025 BAU value of 922 km3 (Rosegrant et al., 2002). 

   Molden  (2007)  summarizes  emerging  trends  in  the water  sector. Global  food  trade  and consequent flows of virtual water (embodied  in food exports) offer prospects for better national food  security  and  relieving  water  stress.  The  changing  climate  affects  temperatures  and precipitation patterns, severely  impacting tropical areas of  intense poverty.  Irrigators dependent on snow melt are even more vulnerable to changes in river flows. Urbanization increases demand for  water,  generates  more  wastewater  and  alters  demands  for  agricultural  products.  Higher energy prices increase the costs of pumping water, applying fertilizers and transporting products, with  implications  for  access  to  water  and  irrigation.  Hydropower  generation  and  agriculture compete for water resources. Greater reliance on bioenergy affects the production and prices of food  crops  and  increases  the  amount  of  water  used  by  agriculture.  The  Comprehensive Assessment estimates that  in 2050 evapotranspiration necessary to support  increased bioenergy use will approach total current depletion by agriculture. Perceptions about water are changing, as water professionals and policymakers realize the need to optimize the use of both blue and green water.  More  attention  is  paid  to  environmental  flows  and  integrated  approaches  to  water management.    Available  land and water  resources  can  satisfy  future  food and  fiber demands  in  several ways: 1) Rainfed  scenario:  investing  to  increase production  in  rainfed agriculture. An optimistic scenario assuming significant upgrades results in an increase of yields from 2.7 metric tons per ha in 2000 to 4.5 in 2050. Risks include low adoption rates and low yield improvements; 2) Irrigation scenario: investing in irrigation. Under optimistic assumptions, 75% of the additional food demand can be met by  improving water productivity on existing  irrigated  lands. Alternatively, expansion requires 40% more withdrawals of water  for agriculture.  Irrigation  could  contribute 55% of  the total value of food supply by 2050. Risks  include degradation of aquatic ecosystems and capture 

Page 27: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

27

fisheries;  3)  Trade  scenario:  conducting  agricultural  trade  within  and  between  countries  can mitigate  scarcity,  but  poor  countries  depend  on  their  national  agriculture  sector,  and  their purchasing  power  is  low;  4)  Reducing  gross  food  demand  by  influencing  diets,  reducing  post‐harvest  losses  (industrial  and  household  waste);  5)  The  Comprehensive  Assessment  scenario combines  elements of different  approaches.  Through  appropriate  investments  in both  irrigated and rainfed agriculture, by 2050 the cropped area will  increase by 9% and water withdrawals for agriculture will  increase by 13%. One challenge  is to minimize the adverse  impacts on ecosystem services (Molden, 2007). 

 Among  the  more  comprehensive  regional  tools,  a  new  computational  model  of  the 

Paraguay‐Paraná river basin (South America) is being designed to help local governments, farmers and ranchers: understand the factors that lead to water scarcity and impurity; make conservation‐friendly decisions about  future  land‐use projects; assess how  landscape planning, water and soil conservation can improve water quality and sustain biodiversity downstream.34  

3.7. WATER QUALITY With the exception of the TechnoGarden scenario, the MA scenarios show an  increase  in 

global  river N  loading  to coastal ecosystems  ranging  from 10%  to 30% between 1995 and 2030. The TechoGarden scenario shows a 10% decrease because a high level of N removal from sewage is assumed and a lower level of NOx emissions and N deposition are computed for this scenario. All scenarios experience  increase  in the use of N and P fertilizers due to growing human population and increased food demand. Changes in nutrient use vary among the scenarios, depending on the key indirect drivers of food demand and technological change. 

 The four MA scenarios (2005) address eutrophication trends up to 2050. The technologies 

developed  in TechnoGarden are expected to reduce the risks of eutrophication, despite a rise  in areas  under  agriculture,  by  reducing  the  impacts  of  agriculture,  sewage  and  energy  use.  In Adapting Mosaic, better  local management of agricultural runoff,  fossil  fuel burning and sewage will reduce the chances of eutrophication. These benefits may be offset by increasing use of fossil fuels  and  expansion  in  area  under  agricultural  areas. Water  quality  in  inland  waterways may improve, but coastal eutrophication in densely populated areas will increase. Global Orchestration expects  increases  in  eutrophication.  Minimal  population  growth  combined  with  technological advances  increase energy and agriculture efficiency. However,  improvements are moderate and do  not  address  the  nutrient  runoff,  resulting  in  an  increase  in  nutrient  pollution,  especially  in 

34 The new tool is being developed by The Nature Conservancy in partnership with IBM, the University of Wisconsin’s Center for Sustainability and the Global Environment (SAGE), and Brazil's Water Agency and local institutions. Development around the Paraguay-Paraná river system is accelerating, with major agriculture and infrastructure projects planned for the near future. More than half of Brazil's extensive Cerrado grasslands have already been converted for ranching and agriculture, causing soil erosion and pollution. South America's largest city, São Paulo, is faced with the difficult task of supporting a population of 18 million with water resources threatened by upstream deforestation, sedimentation and pollution. The need to chemically treat water for São Paulo's residents rose 51% in the past five years. The modeling tool will explore questions about the relationship between changes in native vegetation and water quality; targeted native vegetation restoration to maximize improvement in water quality and flow; selection of national parks and protected areas in order to benefit river system integrity; and effects of different land uses on the integrity of each river. (The Nature Conservancy, 2010).

Page 28: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

28

poorer countries. The reactive management is too slow to address changes in soil and sediment P, setting  up  for  eutrophication  that  is  very  difficult  to  reverse.  Order  from  Strength  is  very vulnerable to increases in nutrient pollution, and areas of eutrophication greatly expand. There is little  investment  in environmental  technology or management  to offset  substantial  increases  in agricultural  area  and  fossil  fuel  use.  In  all  scenarios,  people  have  to  cope with  eutrophication caused by storage of P in soils and recycling of P from lake sediments. 

The  IPCC  scenarios modified  in Ramsar COP8‐Doc.11 predict  lower water quality due  to higher temperature by 2025, which  is further altered by changes  in water flow volume. By 2100, water quality effects are amplified.  

   Presently,  it  is not possible  to  compute global  changes  in water quality  for  the different scenarios. MA uses  a  concept of  return  flows  as  a  surrogate  variable  for water quality. Return flows are the difference between withdrawals and consumption, and represent a rough estimate of the magnitude of wastewater discharged into the receiving water in a watershed. The type and concentration of water pollutants  is different  from different sources. For certain types of return flows, high  rates of  these  flows may  correspond  to  low water quality  and high  levels of water contamination  and  pressure  on  freshwater  ecosystems.  The  adverse  impact  of  return  flows  is greater for untreated flows. In OECD, the partial treatment of wastewater, i.e., removal of organic wastes and pathogens,  is common.  It  is more uncertain whether agricultural return flows will be collected  and  treated  because  of  the  high  costs  of  large  volume  collection  and  treatment. Wastewater treatment in poorer countries is now quite uncommon, except in cities.35     Water scarcity in rapidly expanding cities is projected to increase (Vörösmarty et al., 2000). Between  1995  and  2025,  urban  population  doubles  to  5  billion,  and  faces  escalating  water pollution and waterborne disease. Water supply and demand are driven by population growth and economic development, and to a lesser degree by climate. 

 An important source of uncertainty in estimating return flows is the uncertainty of the ratio 

of consumption to withdrawals in different water use sectors. Also, the tools used to estimate the 

35 By 2050, Global Orchestration experiences a 42% increase in global return flows, which follows an increase in water withdrawals. In OECD and the former Soviet countries population levels off and efficiency of water use improves, resulting in lower withdrawals and return flows. Furthermore, the richer societies maintain their environmental management efforts. Conversely, withdrawals and return flows substantially increase between 2000 and 2050 in sub Saharan Africa (factor of 3.7), Latin America (factor of 2), Asia (49%), and the MENA countries (24%). Under TechnoGarden, the trends up to 2050 are in the same direction as Global Orchestration. Stronger emphasis on improving water efficiency and somewhat lower economic growth rates lead to a stronger decrease in return flows up to 2050 in OECD (18%) and the former Soviet countries (43%), and to slower growth of return flows in sub-Saharan Africa (factor of 3.5), MENA (17%), and Asia (nearly 20%). The increase in Latin America is the same as for Global Orchestration (factor of two). In Adapting Mosaic, return flows decrease in the former Soviet Union because withdrawals decrease. In all other regions, return flows increase compared to Global Orchestration or TechnoGarden because of lower water use efficiency and larger population: sub-Saharan Africa (factor of 5.5), Latin America (factor of 2.6), Asia (76%), MENA (56%) and OECD (4%). Order from Strength has the largest withdrawals because of its slower improvement of the efficiency of water use and faster population growth. Return flows double worldwide between now and 2050. Increases by region are: former Soviet countries (9%), OECD (40%), Asia and MENA countries (approx. 100%), Latin America (factor of 3), and sub-Saharan Africa (factor of 4.7).

Page 29: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

29

indicators  of  ecosystem  services  of  fresh water  are  too  aggregated  to  include  the  role  of  local water policies and management.    An  emerging  water  quality  concern  is  the  impact  of  pharmaceuticals,  e.g.,  antibiotics, painkillers, birth control pills and antidepressants, on aquatic ecosystems. Their long‐term impact on  human  or  ecosystem  health  is  poorly  understood.  Some  compounds  in  personal  care  and household  cleaning  formulations  are  thought  to  be  endocrine  disruptors:  they mimic  natural hormones  in  humans  and  other  species,  e.g.,  nonylphenol  ethoxylates36  (NPE).  Endocrine disruptors may  not  be  removed  by  existing  wastewater  treatment  operations.  Upon  entering freshwater systems, these contaminants can feminize male fish and amphibians, cause thinning of eggshells  in birds,  inadequate parental behavior, cancerous growths, etc  (Carr and Neary, 2008). Other  pollutants which  have  only  recently  received  attention  include:  plastics  debris37;  deicing fluids  in  tarmac  runoff38;  hydraulic  fracturing  (fracking)  compositions39;  aging  and  improperly maintained underground fuel tanks40; stormwater.41 

36 Used in some detergents and industrial cleansers to lift surface dirt, NPE are believed to mimic estrogen. NPE residues are linked to hormonal disruption in fish and marine life. 37 There is a growing concern over the environmental impact of plastic debris on marine ecosystems, as well as public health risks stemming from consumption of contaminated seafood (Derraik, 2002). The deleterious effects of plastic debris on marine fauna include: entanglement in and ingestion of plastic litter, the use of plastic debris by IAS and the absorption of PCBs from ingested plastics. Bisphenol A (BPA) and other plastics additives are not chemically bound to plastics and over time leach into the environment. BPA is believed to be an endocrine disruptor. BPA is detectable in most samples of bottled and piped water. 38 During winter season, airports in snow covered regions spray large volumes of deicing chemicals onto airliners and allow the runoff to seep into nearby waterways. The two most often used compounds are ethylene glycol and propylene glycol. Deicing fluids can turn streams bright orange, lower oxygen levels in waterways and create dead zones for aquatic life. The environmental risk can be minimized by collecting and recycling the used fluid, applying the maximum dilution consistent with safety and by biological digestion, or eliminated by opting for infrared deicing. 39 One of the key environmental concerns surrounding the exploitation of coalbed methane (natural gas) is the use of chemicals in hydraulic fracturing. The process forces large volumes of water, mixed with sand and small amounts of chemicals, into the earth to break coal deposits and release gas. The chemicals used in the process may be responsible for groundwater contamination incidents in drilling areas. Investigations are impeded because the chemical compositions are proprietary trade secrets (USEPA, 2004b). 40 Gasoline is a complex contamination agent, which may originate from a variety of sources, e.g., underground fuel storage tanks in gas stations and large farms. In AlQassim region of Saudi Arabia, most of the existing gas stations are more than 25 years old. The technologies used for the construction of the stations and the underground gasoline tanks are outdated and do not meet international environmental standards. A statistical study by ElKholy et al. (2009) assesses the risks associated with soil and groundwater contamination due to gasoline leakage. Two gas stations are selected for a geoenvironmental investigation. An actual case of soil contamination due to gasoline leakage is confirmed; leaching into the groundwater resources is also possible. 41 In most countries, storm-generated runoff from agricultural and urban areas is the most important non-point pollutant source, e.g., nitrates in runoff and its accumulation in rivers. Point sources are becoming more controlled, e.g., urban sewage discharge. Non-point pollutant loads are becoming an increasing concern. USEPA (2010) notes that about 13% of U.S. rivers, 18% of lakes and 32% of estuaries are impaired by stormwater, i.e., they are rendered unsafe for swimming or fishing. In a natural system, rainwater soaks into the soil and is taken up by plants. The quick infiltration prevents the water from transporting contaminants and keeps waterways from eroding. In the urban environment, instead of soaking into the ground, rain runs across impervious surfaces, picking up contaminants along the way: metals, oil, organic waste, bacteria, etc. When the traveling stormwater hits a stream, it scours sediment, dislodges benthic invertebrates and erodes banks. Regulations need to define what is expected of developers, possibly by setting limits for stormwater volume or concentrations of contaminants. The rules may include guidelines for techniques such

Page 30: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

30

 ACKNOWLEDGEMENTS 

   The author acknowledges UNESCO for supporting her work on this report. The author also wishes  to  acknowledge  the  valuable  comments  and  suggestions  made  by  William  Cosgrove, Jerome Glenn and Gilberto Gallopín. 

                                   

as infiltration basins, green roofs, rain barrels and rain gardens to capture runoff. Developers can also reduce impervious surfaces by constructing narrow roads and using porous pavement.

Page 31: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

31

      

BIBLIOGRAPHY  

Adikari, Y. and Yoshitani, J.: Global Trends in Water‐Related Disasters: an Insight for Policymakers. International Centre for Water Hazard and Risk Management (ICHARM) for UNESCO, Paris, France. 2009. http://unesdoc.unesco.org/images/0018/001817/181793E.pdf  Alcamo,  J.,  van Vuuren, D., Ringler, C., Cramer, W., Masui,  T., Alder,  J.,  Schulze, K.: Changes  in nature’s balance sheet: model‐based estimates of  future worldwide ecosystem services. Ecology and  Society,  10(2):  19.  Resilience  Alliance  Publications.  2005. http://www.ecologyandsociety.org/vol10/iss2/art19/  Alcamo,  J., Döll, P., Henrichs, T., Kaspar, F., Lehner, B., Rösch, T., Siebert, S.: Global estimates of water  withdrawals  and  availability  under  current  and  future  “business‐as‐usual”  conditions. Hydrological Sciences  Journal, 48(3): 339‐348.  International Association of Hydrological Sciences. 2003.  Alcamo, J. and Henrichs, T.: Critical regions: A model‐based estimation of world water resources sensitive to global changes. Aquatic Sciences, 64: 1‐11. Birkhäuser Basel. 2002.  Alcamo,  J., Döll, P., Kaspar, F., Siebert, S.: Global  change and global  scenarios of water use and availability: An application of water GAP 1.0. Center  for Environmental System Research  (CESR), University of Kassel, Kassel, Germany. 1998.  Arnell,  N.:  Climate  change  and  global  water  resources:  SRES  emissions  and  socio‐economic scenarios. Global Environmental Change, 14: 31‐52. Elsevier. 2004.  Aselmann, I. and Crutzen, P.J.: Global distribution of natural freshwater wetlands and rice paddies, their  net  primary  productivity,  seasonality  and  possible  methane  emissions.  Journal  of Atmospheric Chemistry, 8(4): 307‐358. Springer Netherlands. 1989.  Bates, B.C., Kundzewicz, Z.W., Wu, S., Palutikof, J.P. (Eds.): Climate Change and Water. Technical Paper  of  the  Intergovernmental  Panel  on  Climate  Change.  210  pp.  IPCC  Secretariat,  Geneva, Switzerland. 2008. http://www.ipcc.ch/pdf/technical‐papers/climate‐change‐water‐en.pdf  Battad,  D.T.,  Villanueva,  G.,  Li,  C.,  Miller,  V.,  Sutton,  M.:  Assessing  the  Impacts  of  Timber Harvesting  on Water  Yield  in  the  Central Highlands,  Victoria. Department  of  Sustainability  and Environment, Melbourne, Australia. 2007. 

Page 32: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

32

http://www.forestry.org.au/pdf/pdf‐public/conference2007/papers/battad%2020070504.pdf  Bayliss,  B.,  Brenman,  K.,  Elliot,  I.,  Finlayson, M.,  Hall,  R.,  House,  T.,  Pidgeon,  B., Walden,  D., Waterman, P.: Vulnerability Assessment of Predicted Climate Change  and  Sea  Level Rise  in  the Alligator Rivers Region, Northern Territory Australia. Supervising Scientist Report No. 123, 134 pp. Supervising Scientist, Canberra, Australia. 1997.  Brinson, M.M.  and Malvarez,  A.I.:  Temperate  freshwater wetlands:  types,  status,  and  threats. Environmental Conservation, 29(2): 115–133. Cambridge Journals. 2002.  Burke, E.J., Brown,  S.J.  and Christidis N.: Modeling  the Recent Evolution of Global Drought  and Projections  for  the  Twenty‐First  Century  with  the  Hadley  Centre  Climate  Model.  Journal  of Hydrometeorology, 7(5): 1113–1125. American Meteorological Society. 2006.  Carpenter,  S.:  Ecosystem  Surprises:  Research  Needs.  Background  for  Millennium  Ecosystem Assessment Science Follow‐Up Group. ICSU Secretariat, Paris, France. 2006. http://www.icsu‐asia‐pacific.org/resource_centre/6%20EcoSurprise_ResearchNeeds.pdf  Carr, G.M. and Neary,  J.P.: Water Quality  for Ecosystem and Human Health, 2nd Edition. United Nations  Environment  Programme, Global  Environment Monitoring  System,  Burlington,  Canada. 2008.  http://www.gemswater.org/publications/pdfs/water_quality_human_health.pdf  Chapin, F.S., Zavaleta, E.S., Eviner, V.T., Naylor, R.L., Vitousek, P.M., Reynolds, H.L., Hooper, D.U., Lavorel, S., Sala, O.E., Hobbie, S.E., Mack, M.C., Diaz, S.: Consequences of changing biodiversity. Nature, 405: 234–242. Nature Publishing Group, Macmillan Publishers Ltd. 2000.  Ciruna,  K.A., Meyerson,  L.A.  and  Gutierrez,  A.:  The  ecological  and  socio‐economic  impacts  of invasive alien species in inland water ecosystems. Report to the Convention on Biological Diversity on behalf of the Global Invasive Species Programme, Washington, DC. 2004.  Cosgrove, W. and Rijsberman, F.: World water vision: Making water everybody’s business. World Water Council. Earthscan Publications. 2000.  Costanza, R., d’Arge, R. and de Groot, R.: The value of the world’s ecosystems services and natural capital. Nature, 387: 253–260. Nature Publishing Group, Macmillan Publishers Ltd. 1997.  Derraik,  J.G.B.:  The  Pollution  of  the Marine  Environment  by  Plastic  Debris:  a  Review. Marine Pollution Bulletin, 44: 842–852. Elsevier. 2002.  DSE,  Department  of  Sustainability  and  Environment,  Victoria:  Research  Results  for Wood  and Water project: Using  the  latest  science  to develop  a  future management plan  for  State  forests supplying water to Melbourne. Melbourne, Australia. 2008. 

Page 33: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

33

http://www.ourwater.vic.gov.au/__data/assets/pdf_file/0018/12744/Summary_of_Research_Results.pdf  ElKholy, S.M., Sabry, T.I. and AlSalamah,  I.S.: Environmental Contamination Risks Due  to Leaking Underground  Fuel  Tanks  (LUFT)  of Gas  Stations  in  AlQassim  Region,  Saudi  Arabia.  pp.  3‐8,  2nd International Conference on Environmental and Computer Science, Dubai , United Arab Emirates. 2009.  FAO, Food and Agriculture Organization of the United Nations: Global Forest Resource Assessment (FRA). FAO Forestry Paper No. 147. Rome, Italy. 2005. http://www.fao.org/forestry/fra/fra2005/en/  FAO, Food and Agriculture Organization of the United Nations: Unlocking the Water Potential of Agriculture. Rome, Italy. 2003. ftp://ftp.fao.org/agl/aglw/docs/kyotofactsheet_e.pdf  FAO,  Food  and  Agriculture  Organization  of  the  United  Nations:  Review  of  the  State  of World Fishery Resources: Inland Fisheries. FAO Inland Water Resources and Aquaculture Service, Fishery Resources Division, FAO Fisheries Circular No. 942. Rome, Italy. 1999.  FAO, Food and Agriculture Organization of the United Nations: Food production: The critical role of water. World Food Summit, Rome, Italy. 1996.  Feikema,  P.M.,  Sheridan, G.J., Argent, R.M.,  Lane  P.N.J., Grayson, R.B.: Using  E2  To Model  The Impacts Of Bushfires On Water Quality  In South‐Eastern Australia.  In Zerger, A. and Argent, R.M. (Eds.)  MODSIM  2005  International  Congress  on  Modelling  and  Simulation.  pp.  1126‐1132. Modelling  and  Simulation  Society  of  Australia  and  New  Zealand.  2005. http://www.mssanz.org.au/modsim05/papers/feikema.pdf  Fekete, B., Vörösmarty, C.J.  and Grabs W.: Global Composite Runoff  Fields Based  on Observed River Discharge and  Simulated Water Balance. WMO Global Runoff Data Center Report No. 22. World Meteorological Organization, Koblenz, Germany. 1999.  Finlayson,  C.M.  and  Spiers,  A.G.  (Eds.): Global  Review  of Wetland  Resources  and  Priorities  for Inventory. Supervising Scientist Report No. 144. Supervising Scientist, Canberra, Australia. 1999.  Gallopín, G.C.  and Rijsberman,  F.: Three  global water  scenarios.  International  Journal of Water, 1(1): 16‐40. Inderscience Enterprises Ltd. 2000.  Garibaldi,  L.  and  Bartley,  D.M.:  The  database  on  introductions  of  aquatic  species  (DIAS).  FAO Aquaculture Newsletter No. 20. Food and Agriculture Organization of the United Nations, Rome, Italy. 1998.  http://www.fao.org/fi/statist/fisoft/dias/index.htm  

Page 34: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

34

Gitay, H., Brown, S., Easterling, W., Jallow, B. et al.: Chapter 5. Ecosystems and Their Goods and Services.  In:  Climate  Change  2001:  Impacts,  Adaptations,  and  Vulnerability.  Contribution  of Working Group  II  to  the Third Assessment Report of  the  International Panel on Climate Change. McCarthy, J.J., Canziani, O.F., Leary, N.A., Dokken, D.J., White, K.S. (Eds.) 235‐342. IPCC/Cambridge University Publication Press. 2001.  The Global Peatland Initiative. http://www.wetlands.org/projects/GPI/peatland.htm  Higgins,  P.A.T.,  Mastrandrea,  M.D.  and  Schneider,  S.H.:  Dynamics  of  climate  and  ecosystem coupling: abrupt changes and multiple equilibria. Philosophical Transactions of the Royal Society: Biological Sciences, 357: 647–655. Royal Society Publishing. 2002.  Hill, G., Waage,  J. and Phiri, G.: The Water Hyacinth Problem  in Tropical Africa.  In: Delfosse, E.S. and Spencer, N.R. (Eds.) Proceedings of the International Water Hyacinth Consortium. World Bank, Washington, DC. 1997.  http://www.sidney.ars.usda.gov/scientists/nspencer/water_h/appendix5.htm  Huntington,  T.G.:  Evidence  for  Intensification of  the Global Water Cycle: Review  and  Synthesis. Journal of Hydrology, 319: 83‐95. Elsevier. 2006.  Huxman T.E., Smith, M.D., Fay, P.A., Knapp, A.K., Shaw, M.R., Loik M.E., Smith S.D., Tissue, D.T., Zak, J.C., Weltzin, J.F., Pockman, W.T., Sala, O.E., Haddad, B.M., Harte, J., Koch, G.W., Schwinning, S., Small, E.E., Williams, D.G.: Convergence across biomes to a common rain‐use efficiency. Nature, 429: 651‐654. Nature Publishing Group, Macmillan Publishers Ltd. 2004.  ILEC, International Lake Environment Committee Foundation: Managing Lakes and Their Basins for Sustainable  Use:  A  Report  for  Lake  Basin  Managers  and  Stakeholders.  International  Lake Environment Committee Foundation, Kusatsu, Japan. 2005.  IPCC,  Intergovernmental Panel on Climate Change: Climate Change 1995 —  Impacts, Adaptation and Mitigation of Climate Change: Scientific technical analysis. Contribution of Working Group II to the Second Assessment Report. Cambridge University Press. 1996.  Jaga,  K.  and  Dharmani,  C.:  Global  Surveillance  of  DDT  and  DDE  Levels  in  Human  Tissues. International  Journal  of  Occupational  Medicine  &  Environmental  Health,  16(1):  7‐14.  Public Knowledge Project. 2003.  Johnson, N., Revenga, C. and Echeverria, J.: Managing water for people and nature. Science, 292: 1071–1072. The American Association for the Advancement of Science, Washington, DC. 2001  Keating,  B.A., Gaydon,  D.,  Huth, N.I.,  Probert, M.E.,  Verburg,  K.,  Smith,  C.J.,  Bond, W.: Use  of modelling  to explore  the water balance of dryland  farming systems  in the Murray‐Darling Basin, Australia. European Journal of Agronomy, 18: 159–169. Elsevier. 2002.  

Page 35: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

35

Kundzewicz, Z.W., Graczyk, D., Maurer, T., Pinskwar, I., Radziejewski, M., Svensson, C., Szwed, M.: Trend Detection  in  River  Flow  Series:  1.  Annual Maximum  Flow. Hydrological  Sciences  Journal, 50(5): 797‐810. International Association of Hydrological Sciences. 2005.  LakeNet: Global Lake Database. 2003. http://www.worldlakes.org/lakes.asp  Lääne,  A.,  Kraav,  E.  and  Titova,  G.:  Global  International Waters  Assessment.  Baltic  Sea,  GIWA Regional Assessment 17. Global  International Water Assessments, University of Kalmar, Sweden. 2005.  MA, Millennium Ecosystem Assessment: Ecosystems and Human Well‐being: Wetlands and Water Synthesis. World Resources Institute, Washington, DC. 2005.  http://www.maweb.org/documents/document.358.aspx.pdf  MA Scenarios. http://www.millenniumassessment.org/en/Scenarios.aspx  Martindale,  D.  and  Gleick,  P.:  Safeguarding  our water:  How we  can  do  it.  Scientific  American, February: 52–55. Nature Publishing Group. 2001.  Mayers, J., Batchelor, C., Bond, I., Hope, R.A., Morrison, E., Wheeler, B.: Water ecosystem services and poverty under climate change: Key issues and research priorities. Natural Resource Issues No. 17. International Institute for Environment and Development, London, UK. 2009.  McAllister,  D.E.,  Hamilton,  A.L.  and  Harvey,  B.:  Global  freshwater  biodiversity:  striving  for  the integrity of freshwater ecosystems. Sea Wind—Bulletin of Ocean Voice International, 11(3): 1‐140. Ocean Voice International, Ottawa, Canada. 1997.   McIsaac,  G.F.,  David,  M.B.,  Gertner,  G.Z.,  Goolsby,  D.A.  Nitrate  Flux  in  the  Mississippi  River. Nature, 414: 166‐167. Nature Publishing Group, Macmillan Publishers Ltd. 2001.  McNeill, J.R.: Something New Under the Sun: An Environmental History of the Twentieth‐century World. W.W. Norton & Co., Inc. 2000.  Milly,  P.C.D., Wetherald,  R.D., Dunne,  K.A., Delworth,  T.L.:  Increasing  Risk  of Great  Floods  in  a Changing Climate. Nature, 415: 514‐17. Nature Publishing Group, Macmillan Publishers Ltd. 2002.  Molden,  D.  (Ed.): Water  for  Food, Water  for  Life —  A  Comprehensive  Assessment  of Water Management  in  Agriculture.  International  Water  Management  Institute  and  Earthscan Publications. 2007.  National  Research  Council:  Restoration  of  aquatic  ecosystems:  science,  technology  and  public policy. National Academy Press. 1992.  The Nature  Conservancy: Helping water  users  see  the  future:  the  Paraguay‐Paraná  river  basin. 2010. http://www.nature.org/initiatives/freshwater/work/art24060.html 

Page 36: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

36

 Naylor,  R.L.,  Goldburg,  R.J.,  Primavera,  J.H.,  Kautsky,  N.,  Beveridge, M.C.M.,  Clay,  J.,  Folke,  C., Lubchenco, J., Mooney, H., Troell. M.: Effect of Aquaculture on World Fish Supplies. Nature, 405: 1017‐1024. Nature Publishing Group, Macmillan Publishers Ltd. 2000.  Odada,  E.O.,  Ochola, W.O.  and  Olago,  D.O.:  Understanding  future  ecosystem  changes  in  Lake Victoria basin using participatory  local  scenarios. African  Journal of Ecology, 47  (Suppl. 1): 147–153. Blackwell Publishing Ltd. 2009.  Oki, T., Agata, Y., Kanae, S., Saruhashi, T., Yang, D.W., Musiake, K.: Global assessment of current water resources using total runoff integrating pathways, Hydrological Sciences Journal, 46(6), 983‐995. International Association of Hydrological Sciences. 2001.  Olli, K. and Tamminen, T.: Thresholds of Environmental Sustainability. Sustainable Development, Global Change and Ecosystems. Thematic catalogue of univariate indicators. 2006. http://www.thresholds‐eu.org/public/thresholdsdeliverablespdf/D5.1.2_UT.pdf  Peterson, G.D., Carpenter, S.R. and Brock, W.A.: Uncertainty and  the management of multistate ecosystems: an apparently rational route to collapse. Ecology, 84(6): 1403–1411. Ecological Society of America, Washington, DC. 2003.  Postel, S.L.: Pillar of sand — Can the irrigation miracle last? W.W. Norton & Co., Inc. 1999.  Poynter, M. and Featherston, G.: Review of timber production on Melbourne’s Water Catchments. Victorian  Association  of  Forest  Industries,  Melbourne,  Australia.  2008. http://www.vafi.org.au/documents/WaterCatchmentReview08.pdf  Rabalais, N.N.: Nitrogen  in aquatic ecosystems. Ambio, 31: 102‐112. The Royal Swedish Academy of Sciences and Allen Press, Inc. 2002.  Rabalais,  N.N.,  Turner,  R.E.  and Wiseman,  Jr., W.J.:  Gulf  of Mexico  Hypoxia,  a.k.a.  “The  Dead Zone”. Annual Review of Ecology and Systematics, 33:235‐63. Annual Reviews. 2002.  The Ramsar Convention on Wetlands: Ramsar Classification System for Wetland Type.  http://www.ramsar.org/cda/en/ramsar‐documents‐official‐documents‐of/main/ramsar/1‐31%5E7761_4000_0__  Revenga,  C.  and  Kura,  Y.:  Status  and  Trends  of  Biodiversity  of  Inland Water  Ecosystems.  SCBD Technical Series No. 11. Secretariat of  the Convention on Biological Diversity, Montreal, Canada. 2003.  Revenga, C., Brunner, J., Henninger, N., Kassem, K., Payne, R.: Pilot Analysis of Global Ecosystems: Freshwater Systems. World Resources Institute, Washington, DC. 2000.  http://www.wri.org/wr2000/freshwater_page.html 

Page 37: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

37

 Ricciardi,  A.  and  Rasmussen,  J.B.:  Extinction  Rates  of  North  American  Freshwater  Fauna. Conservation Biology, 13(5): 1220‐1222. Wiley‐Blackwell. 1999.  Rosegrant, W.,  Cai,  X.  and  Cline,  S.A.: World Water  and  Food  to  2025:  Dealing with  Scarcity. International Food Policy Research Institute, Washington, DC. 2002.  http://www.ifpri.org/sites/default/files/publications/water2025.pdf  Roshier, D.A., Whetton, P.H., Allan, R.J., Robertson, A.I.: Distribution and persistence of temporary wetland  habitats  in  arid  Australia  in  relation  to  climate.  Austral  Ecology,  26:  371‐384. Wiley‐Blackwell. 2001.  Sala,  O.E.,  Chapin,  F.S.,  Armesto,  J.J.,  Berlow,  E.,  Bloomfield,  J.,  Dirzo,  R.,  Huber‐Sanwald,  E., Huenneke, L.F., Jackson, R.B., Kinzig, A., Leemans, R., Lodge, D.M., Mooney, H.A., Oesterheld, M., Poff,  N.L.,  Sykes, M.T., Walker,  B.H.,   Walker, M., Wall,  D.H.:  Biodiversity—Global  biodiversity scenarios  for  the  year  2100.  Science,  287:  1770–1774.  The  American  Association  for  the Advancement of Science, Washington, DC. 2000.  Scheffer, M.,  Carpenter,  S.R.,  Foley,  J.,  Folke,  C., Walker,  B.:  Catastrophic  shifts  in  ecosystems. Nature, 413: 591‐596. Nature Publishing Group, Macmillan Publishers Ltd. 2001.  Seckler, D., Amarasinghe, U., Molden, D.,  de Rhadika,  S.,  Barker,  R.: World water  demand  and supply, 1990  to 2025:  Scenarios  and  issues.  IWMI Research Report No. 19.  International Water Management Institute, Colombo, Sri Lanka. 1998.  Shiklomanov,  I.A.: World Water Resources at the Beginning of the 21st Century. Prepared for the International Hydrological Programme, UNESCO. St. Petersburg, Russian Federation. 1999.  http://webworld.unesco.org/water/ihp/db/shiklomanov/  Shiklomanov,  I.A.:  Comprehensive  assessment  of  the  freshwater  resources  of  the  world: assessment  of  water  resource  and  water  availability  in  the  world.  World  Meteorological Organization and Stockholm Environment Institute, Stockholm, Sweden. 1997.  Spalding,  R.  and  Exner,  M.:  Occurrence  of  nitrate  in  groundwater.  Journal  of  Environmental Quality,  22:392‐402.  The American  Society  of Agronomy,  Crop  Science  Society  of America,  Soil Science Society of America and HighWire Press. 1993.  UNEP, United Nations Environment Programme: Clearing the Waters: A Focus on Water Quality Solutions. Nairobi, Kenya. 2010. http://www.unep.org/PDF/Clearing_the_Waters.pdf  UNEP,  United  Nations  Environment  Programme:  Water  Security  and  Ecosystem  Services.  A Contribution  to  the  World  Water  Assessment  Programme.  Nairobi,  Kenya. 2009ahttp://www.unesco.org/water/wwap/wwdr/wwdr3/pdf/Water_Security_and_Ecosystems.pdf 

Page 38: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

38

 UNEP,  United  Nations  Environment  Programme:  Persistent  Organic  Pollutants.  Nairobi,  Kenya. 2009b. http://www.chem.unep.ch/pops/  UNEP, United Nations  Environment  Programme, Global  Environment Monitoring  System/Water Programme: Water Quality for Ecosystem and Human Health. 2nd edition. Nairobi, Kenya.  2008. http://www.gemswater.org/publications/pdfs/water_quality_human_health.pdf  UNEP, United Nations  Environment  Programme:  Indicators  for  Assessing  Progress  Towards  the 2010 Target: Water Quality  in Aquatic Ecosystems. Convention on Biological Diversity, Subsidiary Body on Scientific, Technical and Technological Advice, 10th Meeting, Bangkok, Thailand. 2005. http://www.cbd.int/doc/meetings/sbstta/sbstta‐10/information/sbstta‐10‐inf‐19‐en.pdf  UNEP, United Nations Environment Programme: Inventory of Worldwide PCB Destruction. Nairobi, Kenya. 1998. http://www.chem.unep.ch/POPs/pdf/pcbrpt.pdf   USEPA, US  Environmental  Protection Agency:  Proposed National  Rulemaking  to  Strengthen  the National Stormwater Program. 2010. http://cfpub.epa.gov/npdes/stormwater/rulemaking.cfm  USEPA,  US  Environmental  Protection  Agency:  Asian  Carp  and  the  Great  Lakes.  2004a. http://www.epa.gov/greatlakes/invasive/asiancarp/  USEPA, US  Environmental Protection Agency: Evaluation of  Impacts  to Underground  Sources of Drinking  Water  by  Hydraulic  Fracturing  of  Coalbed  Methane  Reservoirs.  National  Study  Final Report. 2004b. http://www.epa.gov/ogwdw000/uic/wells_coalbedmethanestudy.html  Van  Lanen, H.A.J., Kundzewicz, Z.W., Tallaksen,  L.M., Hisdal, H., Fendeková, M., Prudhomme C.: Indices  for different  types of droughts and  floods at different  scales. Water and Global Change (WATCH)  Integrated Project, European Commission Sixth Framework Programme, Global Change and Ecosystems Thematic Priority Area. WATCH Technical Report No.11. Centre  for Ecology and Hydrology, Oxfordshire, UK. 2009.  Van Lanen, H.A.J., Tallaksen, L.M. and Rees G.: Droughts and Climate Change. In: Commission Staff Working Document  Impact Assessment: Accompanying Document to Communication Addressing the Challenge of Water Scarcity and Droughts in the European Union. Commission of the European Communities, Brussels, Belgium. 2007.  Vörösmarty,  C.J., Green,  P.,  Salisbury,  J.,  Lammers,  R.B.: Global Water  Resources:  Vulnerability from Climate Change and Population Growth. Science, 289: 284‐288. The American Association for the Advancement of Science, Washington, DC. 2000.  

Page 39: Exploring alternative futures of the World Water System. a ... · Conservancy, 2010). Marginal water losses to the invasive star thistle in the Sacramento River valley, U.S., are

39

Walling, D.E.: The Impact of Global Change on Erosion and Sediment Transport by Rivers: Current Progress  and  Future  Challenges.  UNESCO,  International  Hydrological  Programme  Sediment Initiative, Paris, France. 2009. http://unesdoc.unesco.org/images/0018/001850/185078E.pdf  Walther, G.R.,  Post,  E.,  Convey,  P., Menzel,  A.,  Parmesan,  C.  Beebeef,  T.J.C.,  Fromentine,  J.M., Hoegh‐Guldberg, O., Bairlein, F.: Ecological responses to recent climate change. Nature, 416: 388‐395. Nature Publishing Group, Macmillan Publishers Ltd. 2002.  Welcomme, R.L.: International Introductions of Inland Aquatic Species. FAO Technical Series Paper No. 294. Food and Agriculture Organization of the United Nations, Rome, Italy. 1988.  WHO, World Health Organization: Guidelines  for Drinking Water Quality:  Incorporating  the First and Second Addenda, Volume 1, Recommendations. 3rd Edition. Geneva, Switzerland. 2008.  WHO,  World  Health  Organization:  The  World  Health  Report:  Changing  History.  Geneva, Switzerland. 2004.  WWDR3, United Nations World Water Development Report: Water  for People — Water  for Life. 576 pp. UNESCO and Berghahn Books, Paris, France. 2003.  Xenopoulous, M.,  Lodge,  D.,  Alcamo,  J., Märker, M.,  Schulze,  K.,  van  Vuuren,  D.:  Scenarios  of freshwater  fish extinctions  from  climate  change and water withdrawals. Global Change Biology, 11: 1–8. Wiley‐Blackwell. 2005.  Zhang X., Zwiers, F.W., Hegerl, G.C., Lambert, F.H., Gillett, N.P., Solomon, S., Stott, P.A., Nozawa, T.: Detection of Human Influence on Twentieth‐Century Precipitation Trends. Nature, 448: 461‐65. Nature Publishing Group, Macmillan Publishers Ltd. 2007.