Jump to content

Groundwater: Difference between revisions

From Wikipedia, the free encyclopedia
[accepted revision][accepted revision]
Content deleted Content added
→‎Climate change: more info added
Line 144: Line 144:


[[Groundwater pollution]] may also increase indirectly due to climate change: More frequent and intense storms can pollute groundwater by mobilizing contaminants, for example fertilizers, wastewater or human excreta from pit latrines.<ref name=":82" />{{rp|611}} Droughts reduce river dilution capacities and groundwater levels, increasing the risk of groundwater contamination.
[[Groundwater pollution]] may also increase indirectly due to climate change: More frequent and intense storms can pollute groundwater by mobilizing contaminants, for example fertilizers, wastewater or human excreta from pit latrines.<ref name=":82" />{{rp|611}} Droughts reduce river dilution capacities and groundwater levels, increasing the risk of groundwater contamination.

[[Effects of climate change]] could have different impacts on groundwater storage: The expected more intense (but fewer) major rainfall events could lead to ''increased'' groundwater recharge. But more intense drought periods could result in soil drying-out and compaction which would ''reduce'' infiltration to groundwater.<ref>IAH (2019) [https://iah.org/wp-content/uploads/2019/07/IAH_Climate-ChangeAdaptationGdwtr.pdf CLIMATE-CHANGE ADAPTATION & GROUNDWATER], Strategic Overview Series</ref>


Using more groundwater, particularly in Sub-Saharan Africa, is seen as a method for [[climate change adaptation]] in the case that climate change causes more intense or frequent droughts.<ref>WaterAid and BGS (2022) [https://washmatters.wateraid.org/publications/groundwater-the-worlds-neglected-defence-against-climate-change Groundwater: The world’s neglected defence against climate change]</ref>
Using more groundwater, particularly in Sub-Saharan Africa, is seen as a method for [[climate change adaptation]] in the case that climate change causes more intense or frequent droughts.<ref>WaterAid and BGS (2022) [https://washmatters.wateraid.org/publications/groundwater-the-worlds-neglected-defence-against-climate-change Groundwater: The world’s neglected defence against climate change]</ref>

Revision as of 14:07, 24 January 2023

An illustration showing groundwater in aquifers (in blue) (1, 5 and 6) below the water table (4), and three different wells (7, 8 and 9) dug to reach it.

Groundwater is the water present beneath Earth's surface in rock and soil pore spaces and in the fractures of rock formations. About 30 percent of all readily available freshwater in the world is groundwater.[1] A unit of rock or an unconsolidated deposit is called an aquifer when it can yield a usable quantity of water. The depth at which soil pore spaces or fractures and voids in rock become completely saturated with water is called the water table. Groundwater is recharged from the surface; it may discharge from the surface naturally at springs and seeps, and can form oases or wetlands. Groundwater is also often withdrawn for agricultural, municipal, and industrial use by constructing and operating extraction wells. The study of the distribution and movement of groundwater is hydrogeology, also called groundwater hydrology.

Typically, groundwater is thought of as water flowing through shallow aquifers, but, in the technical sense, it can also contain soil moisture, permafrost (frozen soil), immobile water in very low permeability bedrock, and deep geothermal or oil formation water. Groundwater is hypothesized to provide lubrication that can possibly influence the movement of faults. It is likely that much of Earth's subsurface contains some water, which may be mixed with other fluids in some instances.

Groundwater is often cheaper, more convenient and less vulnerable to pollution than surface water. Therefore, it is commonly used for public water supplies. For example, groundwater provides the largest source of usable water storage in the United States, and California annually withdraws the largest amount of groundwater of all the states.[2] Underground reservoirs contain far more water than the capacity of all surface reservoirs and lakes in the US, including the Great Lakes. Many municipal water supplies are derived solely from groundwater.[3] Over 2 billion people rely on it as their primary water source worldwide.[4]

Use of groundwater has related environmental issues. For example, polluted groundwater is less visible and more difficult to clean up than pollution in rivers and lakes. Groundwater pollution most often results from improper disposal of wastes on land. Major sources include industrial and household chemicals and garbage landfills, excessive fertilizers and pesticides used in agriculture, industrial waste lagoons, tailings and process wastewater from mines, industrial fracking, oil field brine pits, leaking underground oil storage tanks and pipelines, sewage sludge and septic systems. Additionally, groundwater is susceptible to saltwater intrusion in coastal areas and can cause land subsidence when extracted unsustainably, leading to sinking cities (like Bangkok)) and loss in elevation (such as the multiple meters lost in the Central Valley of California). These issues are made more complicated by sea level rise and other changes caused by climate changes which will affect the water cycle.

Definition

Groundwater is fresh water located in the subsurface pore space of soil and rocks. It is also water that is flowing within aquifers below the water table. Sometimes it is useful to make a distinction between groundwater that is closely associated with surface water and deep groundwater in an aquifer (sometimes called "fossil water").

Role in the water cycle

Water balance
Dzherelo, a common source of drinking water in a Ukrainian village

Groundwater can be thought of in the same terms as surface water: inputs, outputs and storage. The natural input to groundwater is seepage from surface water. The natural outputs from groundwater are springs and seepage to the oceans. Due to its slow rate of turnover, groundwater storage is generally much larger (in volume) compared to inputs than it is for surface water. This difference makes it easy for humans to use groundwater unsustainable for a long time without severe consequences. Nevertheless, over the long term the average rate of seepage above a groundwater source is the upper bound for average consumption of water from that source.

Groundwater is naturally replenished by surface water from precipitation, streams, and rivers when this recharge reaches the water table.[5]

Groundwater can be a long-term 'reservoir' of the natural water cycle (with residence times from days to millennia),[6][7] as opposed to short-term water reservoirs like the atmosphere and fresh surface water (which have residence times from minutes to years). Deep groundwater (which is quite distant from the surface recharge) can take a very long time to complete its natural cycle.

The Great Artesian Basin in central and eastern Australia is one of the largest confined aquifer systems in the world, extending for almost 2 million km2. By analysing the trace elements in water sourced from deep underground, hydrogeologists have been able to determine that water extracted from these aquifers can be more than 1 million years old.

By comparing the age of groundwater obtained from different parts of the Great Artesian Basin, hydrogeologists have found it increases in age across the basin. Where water recharges the aquifers along the Eastern Divide, ages are young. As groundwater flows westward across the continent, it increases in age, with the oldest groundwater occurring in the western parts. This means that in order to have travelled almost 1000 km from the source of recharge in 1 million years, the groundwater flowing through the Great Artesian Basin travels at an average rate of about 1 metre per year.

Groundwater recharge

Groundwater recharge or deep drainage or deep percolation is a hydrologic process, where water moves downward from surface water to groundwater. Recharge is the primary method through which water enters an aquifer. This process usually occurs in the vadose zone below plant roots and is often expressed as a flux to the water table surface. Groundwater recharge also encompasses water moving away from the water table farther into the saturated zone.[8] Recharge occurs both naturally (through the water cycle) and through anthropogenic processes (i.e., "artificial groundwater recharge"), where rainwater and/or reclaimed water is routed to the subsurface.

The most common methods to estimate recharge rates are: chloride mass balance (CMB); soil physics methods; environmental and isotopic tracers; groundwater-level fluctuation methods; water balance (WB) methods (including groundwater models (GMs)); and the estimation of baseflow (BF) to rivers.[9]

Location in aquifers

Schematic of an aquifer showing confined zones, groundwater travel times, a spring and a well
An aquifer is an underground layer of water-bearing material, consisting of permeable or fractured rock, or of unconsolidated materials (gravel, sand, or silt). Aquifers vary greatly in their characteristics. The study of water flow in aquifers and the characterization of aquifers is called hydrogeology. Related terms include aquitard, which is a bed of low permeability along an aquifer, and aquiclude (or aquifuge), which is a solid, impermeable area underlying or overlying an aquifer, the pressure of which could lead to the formation of a confined aquifer. The classification of aquifers is as follows: Saturated versus unsaturated; aquifers versus aquitards; confined versus unconfined; isotropic versus anisotropic; porous, karst, or fractured; transboundary aquifer.

Characteristics

Temperature

The high specific heat capacity of water and the insulating effect of soil and rock can mitigate the effects of climate and maintain groundwater at a relatively steady temperature. In some places where groundwater temperatures are maintained by this effect at about 10 °C (50 °F), groundwater can be used for controlling the temperature inside structures at the surface. For example, during hot weather relatively cool groundwater can be pumped through radiators in a home and then returned to the ground in another well. During cold seasons, because it is relatively warm, the water can be used in the same way as a source of heat for heat pumps that is much more efficient than using air.

The entire surface water flow of the Alapaha River near Jennings, Florida, going into a sinkhole leading to the Floridan Aquifer groundwater

Quantities

Groundwater makes up about thirty percent of the world's fresh water supply, which is about 0.76% of the entire world's water, including oceans and permanent ice.[10][11] About 99% of the world's liquid fresh water is groundwater.[12] Global groundwater storage is roughly equal to the total amount of freshwater stored in the snow and ice pack, including the north and south poles. This makes it an important resource that can act as a natural storage that can buffer against shortages of surface water, as in during times of drought.[13]

The volume of groundwater in an aquifer can be estimated by measuring water levels in local wells and by examining geologic records from well-drilling to determine the extent, depth and thickness of water-bearing sediments and rocks. Before an investment is made in production wells, test wells may be drilled to measure the depths at which water is encountered and collect samples of soils, rock and water for laboratory analyses. Pumping tests can be performed in test wells to determine flow characteristics of the aquifer.[3]

The characteristics of aquifers vary with the geology and structure of the substrate and topography in which they occur. In general, the more productive aquifers occur in sedimentary geologic formations. By comparison, weathered and fractured crystalline rocks yield smaller quantities of groundwater in many environments. Unconsolidated to poorly cemented alluvial materials that have accumulated as valley-filling sediments in major river valleys and geologically subsiding structural basins are included among the most productive sources of groundwater.

Fluid flows can be altered in different lithological settings by brittle deformation of rocks in fault zones; the mechanisms by which this occurs are the subject of fault zone hydrogeology.[14]

Uses by humans

Groundwater may be extracted through a water well

Groundwater is the most accessed source of freshwater around the world, including as drinking water, irrigation, and manufacturing. Groundwater accounts for about half of the world's drinking water, 40% of its irrigation water, and a third of water for industrial purposes.[12]

Another estimate stated that globally groundwater accounts for about one third of all water withdrawals, and surface water for the other two thirds.[15]: 21 

A similar estimate was published in 2021 which stated that "groundwater is estimated to supply between a quarter and a third of the world’s annual freshwater withdrawals to meet agricultural, industrial and domestic demands."[16]: 1091 

Global groundwater depletion has been calculated to be between 100 and 300 km3 per year. This depletion is mainly caused by "expansion of irrigated agriculture in drylands".[16]: 1091 

As of 2010, the top five countries by volume of groundwater extraction were India, China, the US, Pakistan, and Iran.

Aquifers that provide sustainable fresh groundwater to urban areas and for agricultural irrigation are typically close to the ground surface (within a couple of hundred metres) and have some recharge by fresh water. This recharge is typically from rivers or meteoric water (precipitation) that percolates into the aquifer through overlying unsaturated materials.

Water supply for municipal and industrial uses

Municipal and industrial water supplies are provided through large wells. Multiple wells for one water supply source are termed "wellfields", which may withdraw water from confined or unconfined aquifers. Using groundwater from deep, confined aquifers provides more protection from surface water contamination. Some wells, termed "collector wells", are specifically designed to induce infiltration of surface (usually river) water.

Irrigation

Center-pivot irrigated fields in Kansas covering hundreds of square miles watered by the Ogallala Aquifer

In general, the irrigation of 20% of farming land (with various types of water sources) accounts for the production of 40% of food production.[17][18] Irrigation techniques across the globe includes canals redirecting surface water,[19][20] groundwater pumping, and diverting water from dams. Aquifers are critically important in agriculture. Deep aquifers in arid areas have long been water sources for irrigation. A majority of extracted groundwater, 70%, is used for agricultural purposes.[21]

In India, 65% of the irrigation is from groundwater[22] and about 90% of extracted groundwater is used for irrigation.[23]

Occasionally, sedimentary or "fossil" aquifers are used to provide irrigation and drinking water to urban areas. In Libya, for example, Muammar Gaddafi's Great Manmade River project has pumped large amounts of groundwater from aquifers beneath the Sahara to populous areas near the coast.[24] Though this has saved Libya money over the alternative, desalination, the aquifers are likely to run dry in 60 to 100 years.[24] Aquifer depletion has been cited as one of the causes of the food price rises of 2011.[25]

Families collecting water from a water well in Niger.

Developing countries

Groundwater provides critical freshwater supply, particularly in dry regions where surface water availability is limited.[26] Globally, more than one-third of the water used originates from underground. In the mid-latitude arid and semi-arid regions lacking sufficient surface water supply from rivers and reservoirs, groundwater is critical for sustaining global ecology and meeting societal needs of drinking water and food production. The demand for groundwater is rapidly increasing with population growth, while climate change is imposing additional stress on water resources and raising the probability of severe drought occurrence.[26]

The anthropogenic effects on groundwater resources are mainly due to groundwater pumping and the indirect effects of irrigation and land use changes.[26]

Groundwater plays a central role in sustaining water supplies and livelihoods in sub-Saharan Africa.[27] In some cases, groundwater is an additional water source that was not used previously.[28]

Reliance on groundwater is increasing in Sub-Saharan Africa as development programs work towards improving water access and strengthening resilience to climate change.[29] In lower-income areas, groundwater supplies are typically installed without water quality treatment infrastructure or services. This practice is underpinned by an assumption that untreated groundwater is typically suitable for drinking due to the relative microbiological safety of groundwater compared to surface water; however, chemistry risks are largely disregarded.[29] Chemical contaminants occur widely in groundwaters that are used for drinking but are not regularly monitored. Example priority parameters are fluoride, arsenic, nitrate, or salinity.[29]

Challenges

First, flood mitigation schemes, intended to protect infrastructure built on floodplains, have had the unintended consequence of reducing aquifer recharge associated with natural flooding. Second, prolonged depletion of groundwater in extensive aquifers can result in land subsidence, with associated infrastructure damage – as well as, third, saline intrusion.[30] Fourth, draining acid sulphate soils, often found in low-lying coastal plains, can result in acidification and pollution of formerly freshwater and estuarine streams.[31]

Overdraft

Diagram of a water balance of the aquifer

Groundwater is a highly useful and often abundant resource. Most land areas on Earth have some form of aquifer underlying them, sometimes at significant depths. In some cases, these aquifers are rapidly being depleted by the human population. Such over-use, over-abstraction or overdraft can cause major problems to human users and to the environment. The most evident problem (as far as human groundwater use is concerned) is a lowering of the water table beyond the reach of existing wells. As a consequence, wells must be drilled deeper to reach the groundwater; in some places (e.g., California, Texas, and India) the water table has dropped hundreds of feet because of extensive well pumping.[32] The GRACE satellites have collected data that demonstrates 21 of Earth's 37 major aquifers are undergoing depletion.[12] In the Punjab region of India, for example, groundwater levels have dropped 10 meters since 1979, and the rate of depletion is accelerating.[33] A lowered water table may, in turn, cause other problems such as groundwater-related subsidence and saltwater intrusion.[34]

Another cause for concern is that groundwater drawdown from over-allocated aquifers has the potential to cause severe damage to both terrestrial and aquatic ecosystems – in some cases very conspicuously but in others quite imperceptibly because of the extended period over which the damage occurs.[30] The importance of groundwater to ecosystems is often overlooked, even by freshwater biologists and ecologists. Groundwaters sustain rivers, wetlands, and lakes, as well as subterranean ecosystems within karst or alluvial aquifers.

Not all ecosystems need groundwater, of course. Some terrestrial ecosystems – for example, those of the open deserts and similar arid environments – exist on irregular rainfall and the moisture it delivers to the soil, supplemented by moisture in the air. While there are other terrestrial ecosystems in more hospitable environments where groundwater plays no central role, groundwater is in fact fundamental to many of the world's major ecosystems. Water flows between groundwaters and surface waters. Most rivers, lakes, and wetlands are fed by, and (at other places or times) feed groundwater, to varying degrees. Groundwater feeds soil moisture through percolation, and many terrestrial vegetation communities depend directly on either groundwater or the percolated soil moisture above the aquifer for at least part of each year. Hyporheic zones (the mixing zone of streamwater and groundwater) and riparian zones are examples of ecotones largely or totally dependent on groundwater.

A 2021 study found that of ~39 million investigated[how?] groundwater wells 6-20% are at high risk of running dry if local groundwater levels decline by a few meters, or – as with many areas and possibly more than half of major aquifers[35] – continue to decline.[36][37]

Fresh-water aquifers, especially those with limited recharge by snow or rain, also known as meteoric water, can be over-exploited and depending on the local hydrogeology, may draw in non-potable water or saltwater intrusion from hydraulically connected aquifers or surface water bodies. This can be a serious problem, especially in coastal areas and other areas where aquifer pumping is excessive.

Subsidence

Subsidence occurs when too much water is pumped out from underground, deflating the space below the above-surface, and thus causing the ground to collapse. The result can look like craters on plots of land. This occurs because, in its natural equilibrium state, the hydraulic pressure of groundwater in the pore spaces of the aquifer and the aquitard supports some of the weight of the overlying sediments. When groundwater is removed from aquifers by excessive pumping, pore pressures in the aquifer drop and compression of the aquifer may occur. This compression may be partially recoverable if pressures rebound, but much of it is not. When the aquifer gets compressed, it may cause land subsidence, a drop in the ground surface.[38]

In unconsolidated aquifers, groundwater is produced from pore spaces between particles of gravel, sand, and silt. If the aquifer is confined by low-permeability layers, the reduced water pressure in the sand and gravel causes slow drainage of water from the adjoining confining layers. If these confining layers are composed of compressible silt or clay, the loss of water to the aquifer reduces the water pressure in the confining layer, causing it to compress from the weight of overlying geologic materials. In severe cases, this compression can be observed on the ground surface as subsidence. Unfortunately, much of the subsidence from groundwater extraction is permanent (elastic rebound is small). Thus, the subsidence is not only permanent, but the compressed aquifer has a permanently reduced capacity to hold water.

The city of New Orleans, Louisiana is actually below sea level today, and its subsidence is partly caused by removal of groundwater from the various aquifer/aquitard systems beneath it.[39] In the first half of the 20th century, the San Joaquin Valley experienced significant subsidence, in some places up to 8.5 metres (28 feet)[40] due to groundwater removal. Cities on river deltas, including Venice in Italy,[41] and Bangkok in Thailand,[42] have experienced surface subsidence; Mexico City, built on a former lake bed, has experienced rates of subsidence of up to 40 cm (1'3") per year.[43]

For coastal cities, subsidence can increase the risk of other environmental issues, such as sea level rise.[44] For example, Bangkok is expected to have 5.138 million people exposed to coastal flooding by 2070 because of these combining factors.[44]

Groundwater becoming saline due to evaporation

If the surface water source is also subject to substantial evaporation, a groundwater source may become saline. This situation can occur naturally under endorheic bodies of water, or artificially under irrigated farmland. In coastal areas, human use of a groundwater source may cause the direction of seepage to ocean to reverse which can also cause soil salinization.

As water moves through the landscape, it collects soluble salts, mainly sodium chloride. Where such water enters the atmosphere through evapotranspiration, these salts are left behind. In irrigation districts, poor drainage of soils and surface aquifers can result in water tables' coming to the surface in low-lying areas. Major land degradation problems of soil salinity and waterlogging result,[45] combined with increasing levels of salt in surface waters. As a consequence, major damage has occurred to local economies and environments.[46]

Aquifers in surface irrigated areas in semi-arid zones with reuse of the unavoidable irrigation water losses percolating down into the underground by supplemental irrigation from wells run the risk of salination.[47]

Surface irrigation water normally contains salts in the order of 0.5 g/L or more and the annual irrigation requirement is in the order of 10,000 m3/ha or more so the annual import of salt is in the order of 5,000 kg/ha or more.[48]

Under the influence of continuous evaporation, the salt concentration of the aquifer water may increase continually and eventually cause an environmental problem.

For salinity control in such a case, annually an amount of drainage water is to be discharged from the aquifer by means of a subsurface drainage system and disposed of through a safe outlet. The drainage system may be horizontal (i.e. using pipes, tile drains or ditches) or vertical (drainage by wells). To estimate the drainage requirement, the use of a groundwater model with an agro-hydro-salinity component may be instrumental, e.g. SahysMod.

Seawater intrusion

Aquifers near the coast have a lens of freshwater near the surface and denser seawater under freshwater. Seawater penetrates the aquifer diffusing in from the ocean and is denser than freshwater. For porous (i.e., sandy) aquifers near the coast, the thickness of freshwater atop saltwater is about 12 metres (40 ft) for every 0.3 m (1 ft) of freshwater head above sea level. This relationship is called the Ghyben-Herzberg equation. If too much groundwater is pumped near the coast, salt-water may intrude into freshwater aquifers causing contamination of potable freshwater supplies. Many coastal aquifers, such as the Biscayne Aquifer near Miami and the New Jersey Coastal Plain aquifer, have problems with saltwater intrusion as a result of overpumping and sea level rise.

Seawater intrusion is the flow or presence of seawater into coastal aquifers; it is a case of saltwater intrusion. It is a natural phenomenon but can also be caused or worsened by anthropogenic factors, such as sea level rise due to climate change.[49] In the case of homogeneous aquifers, seawater intrusion forms a saline wedge below a transition zone to fresh groundwater, flowing seaward on the top.[50][51] These changes can have other effects on the land above the groundwater. For example, coastal groundwater in California would rise in many aquifers, increasing risks of flooding and runoff challenges.[49]

Sea level rise causes the mixing of sea water into the coastal groundwater, rendering it unusable once it amounts to more than 2-3% of the reservoir. Along an estimated 15% of the US coastline, the majority of local groundwater levels are already below the sea level.[52]

Pollution

Waterborne diseases can be spread via a groundwater well which is contaminated with fecal pathogens from pit latrines
Groundwater pollution in Lusaka, Zambia, where the pit latrine in the background is polluting the shallow well in the foreground with pathogens and nitrate

Groundwater pollution (also called groundwater contamination) occurs when pollutants are released to the ground and make their way into groundwater. This type of water pollution can also occur naturally due to the presence of a minor and unwanted constituent, contaminant, or impurity in the groundwater, in which case it is more likely referred to as contamination rather than pollution. Groundwater pollution can occur from on-site sanitation systems, landfill leachate, effluent from wastewater treatment plants, leaking sewers, petrol filling stations, hydraulic fracturing (fracking) or from over application of fertilizers in agriculture. Pollution (or contamination) can also occur from naturally occurring contaminants, such as arsenic or fluoride.[53] Using polluted groundwater causes hazards to public health through poisoning or the spread of disease (water-borne diseases).

The pollutant often produces a contaminant plume within an aquifer. Movement of water and dispersion within the aquifer spreads the pollutant over a wider area. Its advancing boundary, often called a plume edge, can intersect with groundwater wells and surface water, such as seeps and springs, making the water supplies unsafe for humans and wildlife. The movement of the plume, called a plume front, may be analyzed through a hydrological transport model or groundwater model. Analysis of groundwater pollution may focus on soil characteristics and site geology, hydrogeology, hydrology, and the nature of the contaminants. Different mechanisms have influence on the transport of pollutants, e.g. diffusion, adsorption, precipitation, decay, in the groundwater.

Climate change

Climate change causes changes to the water cycle which in turn affect groundwater in several ways: There can be a decline in groundwater storage, and reduction in groundwater recharge and water quality deterioration due to extreme weather events.[54]: 558  However, the exact impacts of climate change on groundwater are still under investigation.[54]: 579  This is because scientific data derived from groundwater monitoring is still missing, such as changes in space and time, abstraction data and "numerical representations of groundwater recharge processes".[54]: 579 

There is an observed declined in groundwater storage in many parts of the world. This is due to more groundwater being used for irrigation activities in agriculture, particularly in drylands.[16]: 1091  Some of this increase in irrigation can be due to water scarcity issues made worse by effects of climate change on the water cycle.

In the tropics intense precipitation and flooding events appear to lead to more groundwater recharge.[54]: 582  However for the higher altitudes regions, the reduced duration and amount of snow may lead to reduced recharge of groundwater in spring.[54]: 582 

Sea level rise that is caused by climate change can also lead to seawater intrusion into aquifers in low-lying areas and small islands (see above in section on sewater intrusion).[54]: 611  This is expected to threaten coastal ecosystems and livelihood resilience. Bangladesh is a vulnerable country for this issue, and mangrove forest of Sundarbans is a vulnerable ecosystem.[54]: 611 

Groundwater pollution may also increase indirectly due to climate change: More frequent and intense storms can pollute groundwater by mobilizing contaminants, for example fertilizers, wastewater or human excreta from pit latrines.[54]: 611  Droughts reduce river dilution capacities and groundwater levels, increasing the risk of groundwater contamination.

Effects of climate change could have different impacts on groundwater storage: The expected more intense (but fewer) major rainfall events could lead to increased groundwater recharge. But more intense drought periods could result in soil drying-out and compaction which would reduce infiltration to groundwater.[55]

Using more groundwater, particularly in Sub-Saharan Africa, is seen as a method for climate change adaptation in the case that climate change causes more intense or frequent droughts.[56]

Regulations

Groundwater withdrawal rates from the Ogallala Aquifer in the Central United States

United States

In the United States, laws regarding ownership and use of groundwater are generally state laws. Regulation of groundwater to minimize pollution of groundwater is addressed in both state and federal law; in the latter case, through regulations issued by the U.S. Environmental Protection Agency (EPA).

  • The Rule of Capture, based on English common law, provides each landowner the ability to capture as much groundwater as they can put to a beneficial use, but they are not guaranteed any set amount of water. As a result, well-owners are not liable to other landowners for taking water from beneath their land. State laws or regulations will often define "beneficial use", and sometimes place other limits, such as disallowing groundwater extraction which causes subsidence on neighboring property.
  • Limited private ownership rights similar to riparian rights in a surface stream. The amount of groundwater right is based on the size of the surface area where each landowner gets a corresponding amount of the available water. Once adjudicated, the maximum amount of the water right is set, but the right can be decreased if the total amount of available water decreases as is likely during a drought. Landowners may sue others for encroaching upon their groundwater rights, and water pumped for use on the overlying land takes preference over water pumped for use off the land.
  • The Reasonable use rule in American drainage law does not guarantee the landowner a set amount of water, but allows unlimited extraction as long as the result does not unreasonably damage other wells or the aquifer system. Usually this rule gives great weight to historical uses and prevents new uses that interfere with the prior use.
  • EPA published its "groundwater Rule", applicable to public water systems, in 2006. The rule focuses on groundwater-supplied systems that may be subject to contamination from fecal bacteria, and requires such systems to take corrective action.[57][58]
  • In real estate property transactions both groundwater and soil are the subjects of scrutiny. For brownfields sites (formerly contaminated sites that have been remediated), EPA requires preparation of Phase I Environmental Site Assessments, to investigate and disclose potential pollution issues.[59] In the San Fernando Valley of California, real estate contracts for property transfer below the Santa Susana Field Laboratory (SSFL) and eastward have clauses releasing the seller from liability for groundwater contamination consequences from existing or future pollution of the Valley Aquifer.
A woman pumps water from a handpump in her village in Sindh, Pakistan

India

The groundwater regulation is controlled and maintained by the central government and four organizations; 1) Central Water Commission, 2) Central groundwater, 3) Central groundwater Authority, 4) Central Pollution Control Board.[60]

Laws, regulations and scheme regarding India's groundwater:

  • 2019 Atal Bhujal Yojana (Atal groundwater scheme), a 5 years (2020-21 to 2024-25) scheme costing INR 6 billion (US$854 million) for managing demand side with village panchayat level water security plans, was approved for implementation in 8,350 water-stressed villages across 7 states, including Haryana, Gujarat, Karnataka, Madhya Pradesh, Maharashtra, Rajasthan and Uttar Pradesh.[61]
  • 2013 National Water Framework Bill ensures that India's groundwater is a public resource, and is not to be exploited by companies through privatization of water. The National Water Framework Bill allows for everyone to access clean drinking water, of the right to clean drinking water under Article 21 of 'Right to Life' in India's Constitution. The bill indicates a want for the states of India to have full control of groundwater contained in aquifers. So far Andhra Pradesh, Assam, Bihar, Goa, Himachal Pradesh, Jammu & Kashmir, Karnataka, Kerala, West Bengal, Telangana, Maharashtra, Lakshadweep, Puducherry, Chandigarh, Dadra & Nagar Haveli are the only ones using this bill.[60]
  • In 2012, National Water Policy was updated, which had previously been launched in 1987 and updated in 2002 and later in 2012.[62]
  • In 2011, the Indian Government created a Model Bill for Groundwater Management; this model selects which state governments can enforce their laws on groundwater usage and regulation.
  • 1882 Easement Act gives landowners priority over surface and groundwater that is on their land and allows them to give or take as much as they want as long as the water is on their land. This act prevents the government from enforcing regulations of groundwater, allowing many landowners to privatize their groundwater instead accessing it in community areas. 1882 Easement Act's Section 7(g) states that every landowner has the right to collect within his limits, all water under the land and on its surface which does not pass in a defined channel.[60]

Canada

A significant portion of Canada’s population relies on the use of groundwater. In Canada, roughly 8.9 million people or 30% of Canada's population rely on groundwater for domestic use and approximately two thirds of these users live in rural areas.[63]

  • The Constitution Act, 1867, does not give authority over groundwater to either order of Canadian government; therefore, the matter largely falls under provincial jurisdiction
  • Federal and Provincial governments can share responsibilities when dealing with agriculture, health, inter-provincial waters and national water-related issues.
  • Federal jurisdiction in areas as boundary/trans-boundary waters, fisheries, navigation, and water on federal lands, First Nations reserves and in Territories.
  • Federal jurisdiction over groundwater when aquifers cross inter-provincial or international boundaries.

A large federal government groundwater initiative is the development of the multi-barrier approach. The multi-barrier approach is a system of processes to prevent the deterioration of drinking water from the source. The multi-barrier consists of three key elements: Source water protection, drinking water treatment, and drinking water distribution systems.[64]

By country

Groundwater is an important water resource for the supply of drinking water, especially in arid countries.

Extraterrestrial groundwater

Groundwater may not be confined only to Earth. The formation of some of the landforms observed on Mars may have been influenced by groundwater. There is also evidence that liquid water may also exist in the subsurface of Jupiter's moon Europa.[65]

See also

References

  1. ^ "What is Groundwater? | International Groundwater Resources Assessment Centre". www.un-igrac.org. Retrieved 2022-03-14.
  2. ^ National Geographic Almanac of Geography, 2005, ISBN 0-7922-3877-X, p. 148.
  3. ^ a b "What is hydrology and what do hydrologists do?". The USGS Water Science School. United States Geological Survey. 23 May 2013. Retrieved 21 Jan 2014.
  4. ^ Famiglietti, J. S. (November 2014). "The global groundwater crisis". Nature Climate Change. 4 (11): 945–948. Bibcode:2014NatCC...4..945F. doi:10.1038/nclimate2425. ISSN 1758-6798. Retrieved 2 March 2022.
  5. ^ United States Department of the Interior (1977). Ground Water Manual (First ed.). United States Government Printing Office. p. 4.
  6. ^ Bethke, Craig M.; Johnson, Thomas M. (May 2008). "Groundwater Age and Groundwater Age Dating". Annual Review of Earth and Planetary Sciences. 36 (1): 121–152. Bibcode:2008AREPS..36..121B. doi:10.1146/annurev.earth.36.031207.124210. ISSN 0084-6597.
  7. ^ Gleeson, Tom; Befus, Kevin M.; Jasechko, Scott; Luijendijk, Elco; Cardenas, M. Bayani (February 2016). "The global volume and distribution of modern groundwater". Nature Geoscience. 9 (2): 161–167. Bibcode:2016NatGe...9..161G. doi:10.1038/ngeo2590. ISSN 1752-0894.
  8. ^ Freeze, R.A.; Cherry, J.A. (1979). Groundwater. Prentice-Hall. ISBN 978-0-13-365312-0. OCLC 643719314. Accessed from: http://hydrogeologistswithoutborders.org/wordpress/1979-english/ Archived 2020-04-06 at the Wayback Machine
  9. ^ MacDonald, Alan M; Lark, R Murray; Taylor, Richard G; Abiye, Tamiru; Fallas, Helen C; Favreau, Guillaume; Goni, Ibrahim B; Kebede, Seifu; Scanlon, Bridget; Sorensen, James P R; Tijani, Moshood; Upton, Kirsty A; West, Charles (2021-03-01). "Mapping groundwater recharge in Africa from ground observations and implications for water security". Environmental Research Letters. 16 (3): 034012. Bibcode:2021ERL....16c4012M. doi:10.1088/1748-9326/abd661. ISSN 1748-9326. S2CID 233941479. Text was copied from this source, which is available under a Creative Commons Attribution 4.0 International License
  10. ^ "Where is Earth's Water?". www.usgs.gov. Retrieved 2020-03-18.
  11. ^ Gleick, P. H. (1993). Water in crisis. Pacific Institute for Studies in Dev., Environment & Security. Stockholm Env. Institute, Oxford Univ. Press. 473p, 9.
  12. ^ a b c Lall, Upmanu; Josset, Laureline; Russo, Tess (2020-10-17). "A Snapshot of the World's Groundwater Challenges". Annual Review of Environment and Resources. 45 (1): 171–194. doi:10.1146/annurev-environ-102017-025800. ISSN 1543-5938.
  13. ^ "Learn More: Groundwater". Columbia Water Center. Retrieved 15 September 2009.
  14. ^ Bense, V.F.; Gleeson, T.; Loveless, S.E.; Bour, O.; Scibek, J. (2013). "Fault zone hydrogeology". Earth-Science Reviews. 127: 171–192. Bibcode:2013ESRv..127..171B. doi:10.1016/j.earscirev.2013.09.008.
  15. ^ Alix, Alexandre; Bellet, Laurent; Trommsdorff, Corinne; Audureau, Iris, eds. (2022). Reducing the Greenhouse Gas Emissions of Water and Sanitation Services: Overview of emissions and their potential reduction illustrated by utility know-how. IWA Publishing. doi:10.2166/9781789063172. ISBN 978-1-78906-317-2. S2CID 250128707.
  16. ^ a b c Douville, H., K. Raghavan, J. Renwick, R.P. Allan, P.A. Arias, M. Barlow, R. Cerezo-Mota, A. Cherchi, T.Y. Gan, J. Gergis, D.  Jiang, A.  Khan, W.  Pokam Mba, D.  Rosenfeld, J. Tierney, and O.  Zolina, 2021: Chapter 8: Water Cycle Changes. In Climate Change 2021: The Physical Science Basis. Contribution of Working Group I  to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change [Masson-Delmotte, V., P. Zhai, A. Pirani, S.L. Connors, C. Péan, S. Berger, N. Caud, Y. Chen, L. Goldfarb, M.I. Gomis, M. Huang, K. Leitzell, E. Lonnoy, J.B.R. Matthews, T.K. Maycock, T. Waterfield, O. Yelekçi, R. Yu, and B. Zhou (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 1055–1210, doi:10.1017/9781009157896.010.
  17. ^ "On Water". European Investment Bank. Retrieved 2020-12-07.
  18. ^ "Water in Agriculture". World Bank. Retrieved 2020-12-07.
  19. ^ McNeill 2000 pp.174.
  20. ^ Peterson 2016
  21. ^ "Facts About Global Groundwater Usage". National Ground Water Association. Retrieved 29 March 2021.
  22. ^ PM Launches Rs 6,000 Crore Groundwater Management Plan, NDTV, 25 December 2019.
  23. ^ Chindarkar, Namrata; Grafton, Quentin (5 January 2019). "India's depleting groundwater: When science meets policy". Asia & the Pacific Policy Studies. 6 (1): 108–124. doi:10.1002/app5.269.
  24. ^ a b Scholl, Adam. "Map Room: Hidden Waters". World Policy journal. Archived from the original on 30 October 2021. Retrieved 19 December 2012.
  25. ^ Brown, Lester (10 January 2011). "The Great Food Crisis of 2011". Foreign Policy Magazine.
  26. ^ a b c Wu WY, Lo MH, Wada Y, Famiglietti JS, Reager JT, Yeh PJ, et al. (July 2020). "Divergent effects of climate change on future groundwater availability in key mid-latitude aquifers". Nature Communications. 11 (1): 3710. Bibcode:2020NatCo..11.3710W. doi:10.1038/s41467-020-17581-y. PMC 7382464. PMID 32709871. Text was copied from this source, which is available under a Creative Commons Attribution 4.0 International License
  27. ^ Cuthbert MO, Taylor RG, Favreau G, Todd MC, Shamsudduha M, Villholth KG, et al. (August 2019). "Observed controls on resilience of groundwater to climate variability in sub-Saharan Africa" (PDF). Nature. 572 (7768): 230–234. Bibcode:2019Natur.572..230C. doi:10.1038/s41586-019-1441-7. PMID 31391559. S2CID 199491973.
  28. ^ Taye, Meron Teferi; Dyer, Ellen (22 August 2019). "Ethiopia's future is tied to water -- a vital yet threatened resource in a changing climate". The Conversation. Retrieved 4 August 2022.
  29. ^ a b c Nowicki, Saskia; Birhanu, Behailu; Tanui, Florence; Sule, May N.; Charles, Katrina; Olago, Daniel; Kebede, Seifu (2023). "Water chemistry poses health risks as reliance on groundwater increases: A systematic review of hydrogeochemistry research from Ethiopia and Kenya". Science of the Total Environment. 904: 166929. doi:10.1016/j.scitotenv.2023.166929. PMID 37689199. Text was copied from this source, which is available under a Creative Commons Attribution 4.0 International License
  30. ^ a b Zektser, S.; LoaIciga, H. A.; Wolf, J. T. (2004). "Environmental impacts of groundwater overdraft: selected case studies in the southwestern United States". Environmental Geology. 47 (3): 396–404. doi:10.1007/s00254-004-1164-3. S2CID 129514582.
  31. ^ Sommer, Bea; Horwitz, Pierre; Sommer, Bea; Horwitz, Pierre (2001). "Water quality and macroinvertebrate response to acidification following intensified summer droughts in a Western Australian wetland". Marine and Freshwater Research. 52 (7): 1015. doi:10.1071/MF00021.
  32. ^ Perrone, Debra; Jasechko, Scott (August 2019). "Deeper well drilling an unsustainable stopgap to groundwater depletion". Nature Sustainability. 2 (8): 773–782. doi:10.1038/s41893-019-0325-z. ISSN 2398-9629. S2CID 199503276.
  33. ^ Upmanu Lall (28 July 2009). "Punjab: A tale of prosperity and decline". Columbia Water Center. Retrieved 2009-09-11.
  34. ^ Bagheri, Rahim; Nosrati, Azad; Jafari, Hadi; Eggenkamp, Hermanus Gerardus M.; Mozafari, Morteza (5 May 2019). "Overexploitation hazards and salinization risks in crucial declining aquifers, chemo-isotopic approaches". Journal of Hazardous Materials. 369: 150–163. doi:10.1016/j.jhazmat.2019.02.024. ISSN 0304-3894. PMID 30776598. S2CID 73455611. Retrieved 2 March 2022.
  35. ^ Famiglietti, James S.; Ferguson, Grant (23 April 2021). "The hidden crisis beneath our feet". Science. 372 (6540): 344–345. Bibcode:2021Sci...372..344F. doi:10.1126/science.abh2867. PMID 33888627. S2CID 233353241. Retrieved 10 May 2021.
  36. ^ "The largest assessment of global groundwater wells finds many are at risk of drying up". ScienceDaily. Retrieved 10 May 2021.
  37. ^ Jasechko, Scott; Perrone, Debra (23 April 2021). "Global groundwater wells at risk of running dry". Science. 372 (6540): 418–421. Bibcode:2021Sci...372..418J. doi:10.1126/science.abc2755. ISSN 0036-8075. PMID 33888642. S2CID 233353207. Retrieved 10 May 2021.
  38. ^ Galloway, Devin L.; Burbey, Thomas J. (December 2011). "Review: Regional land subsidence accompanying groundwater extraction". Hydrogeology Journal (in English, French, Spanish, Chinese, and Portuguese). 19 (8): 1459–1486. Bibcode:2011HydJ...19.1459G. doi:10.1007/s10040-011-0775-5. S2CID 127084866. Retrieved 2 March 2022.
  39. ^ Dokka, Roy K. (2011). "The role of deep processes in late 20th century subsidence of New Orleans and coastal areas of southern Louisiana and Mississippi". Journal of Geophysical Research. 116 (B6): B06403. Bibcode:2011JGRB..116.6403D. doi:10.1029/2010JB008008. ISSN 0148-0227. S2CID 53395648.
  40. ^ Sneed, M; Brandt, J; Solt, M (2013). "Land Subsidence along the Delta-Mendota Canal in the Northern Part of the San Joaquin Valley, California, 2003–10" (PDF). USGS Scientific Investigations Report 2013-5142. Retrieved 22 June 2015.
  41. ^ Tosi, Luigi; Teatini, Pietro; Strozzi, Tazio; Da Lio, Cristina (2014). "Relative Land Subsidence of the Venice Coastland, Italy". Engineering Geology for Society and Territory – Volume 4. pp. 171–73. doi:10.1007/978-3-319-08660-6_32. ISBN 978-3-319-08659-0.
  42. ^ Aobpaet, Anuphao; Cuenca, Miguel Caro; Hooper, Andrew; Trisirisatayawong, Itthi (2013). "InSAR time-series analysis of land subsidence in Bangkok, Thailand". International Journal of Remote Sensing. 34 (8): 2969–82. Bibcode:2013IJRS...34.2969A. doi:10.1080/01431161.2012.756596. ISSN 0143-1161. S2CID 129140583.
  43. ^ Arroyo, Danny; Ordaz, Mario; Ovando-Shelley, Efrain; Guasch, Juan C.; Lermo, Javier; Perez, Citlali; Alcantara, Leonardo; Ramírez-Centeno, Mario S. (2013). "Evaluation of the change in dominant periods in the lake-bed zone of Mexico City produced by ground subsidence through the use of site amplification factors". Soil Dynamics and Earthquake Engineering. 44: 54–66. doi:10.1016/j.soildyn.2012.08.009. ISSN 0267-7261.
  44. ^ a b Nicholls, R. J.; Hanson, S.; Herweijer, C.; Patmore, N.; Hallegatte, S.; CorfeeMorlot, J.; Chateau, Jean; Muir-Wood, Robert (2008). "Ranking Port Cities with High Exposure and Vulnerability to Climate Extremes: Exposure Estimates" (PDF). OECD Environment Working Papers (1). doi:10.1787/011766488208. Retrieved 22 May 2014.
  45. ^ "Free articles and software on drainage of waterlogged land and soil salinity control". Retrieved 2010-07-28.
  46. ^ Ludwig, D.; Hilborn, R.; Walters, C. (1993). "Uncertainty, Resource Exploitation, and Conservation: Lessons from History" (PDF). Science. 260 (5104): 17–36. Bibcode:1993Sci...260...17L. doi:10.1126/science.260.5104.17. JSTOR 1942074. PMID 17793516. Archived from the original (PDF) on 2013-08-26. Retrieved 2011-06-09.
  47. ^ ILRI (1989), Effectiveness and Social/Environmental Impacts of Irrigation Projects: a Review (PDF), In: Annual Report 1988 of the International Institute for Land Reclamation and Improvement (ILRI), Wageningen, The Netherlands, pp. 18–34{{citation}}: CS1 maint: location missing publisher (link)
  48. ^ ILRI (2003), Drainage for Agriculture: Drainage and hydrology/salinity - water and salt balances. Lecture notes International Course on Land Drainage, International Institute for Land Reclamation and Improvement (ILRI), Wageningen, The Netherlands. Download from : [1], or directly as PDF : [2]
  49. ^ a b Befus, K. M.; Barnard, P. L.; Hoover, D. J.; Finzi Hart, J. A.; Voss, C. I. (October 2020). "Increasing threat of coastal groundwater hazards from sea-level rise in California". Nature Climate Change. 10 (10): 946–952. Bibcode:2020NatCC..10..946B. doi:10.1038/s41558-020-0874-1. ISSN 1758-6798. S2CID 221146885.
  50. ^ Polemio, M.; Dragone, V.; Limoni, P.P. (2009). "Monitoring and methods to analyse the groundwater quality degradation risk in coastal karstic aquifers (Apulia, Southern Italy)". Environmental Geology. 58 (2): 299–312. Bibcode:2009EnGeo..58..299P. doi:10.1007/s00254-008-1582-8. S2CID 54203532.
  51. ^ Fleury, P.; Bakalowicz, M.; De Marsily, G. (2007). "Submarine springs and coastal karst aquifers: a review". Journal of Hydrology. 339 (1–2): 79–92. Bibcode:2007JHyd..339...79F. doi:10.1016/j.jhydrol.2007.03.009.
  52. ^ Jasechko, Scott J.; Perrone, Debra; Seybold, Hansjörg; Fan, Ying; Kirchner, James W. (26 June 2020). "Groundwater level observations in 250,000 coastal US wells reveal scope of potential seawater intrusion". Nature Communications. 11 (1): 3229. Bibcode:2020NatCo..11.3229J. doi:10.1038/s41467-020-17038-2. PMC 7319989. PMID 32591535.
  53. ^ Adelana, Segun Michael (2014). Groundwater: Hydrogeochemistry, Environmental Impacts and Management Practices. Nova Science Publishers, Inc. ISBN 978-1-63321-791-1. OCLC 915416488.
  54. ^ a b c d e f g h Caretta, M.A., A. Mukherji, M. Arfanuzzaman, R.A. Betts, A. Gelfan, Y. Hirabayashi, T.K. Lissner, J. Liu, E. Lopez Gunn, R. Morgan, S. Mwanga, and S. Supratid, 2022: Chapter 4: Water. In: Climate Change 2022: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change [H.-O. Pörtner, D.C. Roberts, M. Tignor, E.S. Poloczanska, K. Mintenbeck, A. Alegría, M. Craig, S. Langsdorf, S. Löschke, V. Möller, A. Okem, B. Rama (eds.)]. Cambridge University Press, Cambridge, UK and New York, NY, USA, pp. 551–712, doi:10.1017/9781009325844.006.
  55. ^ IAH (2019) CLIMATE-CHANGE ADAPTATION & GROUNDWATER, Strategic Overview Series
  56. ^ WaterAid and BGS (2022) Groundwater: The world’s neglected defence against climate change
  57. ^ U.S. Environmental Protection Agency (EPA). 2006-11-08. "National Primary Drinking Water Regulations: Ground Water Rule; Final Rule." Federal Register, 71 FR 65574
  58. ^ "Ground Water Rule". Drinking Water Requirements for States and Public Water Systems. EPA. 2018-12-18.
  59. ^ "Brownfields All Appropriate Inquiries". EPA. 2019-12-19.
  60. ^ a b c Suhag, Roopal (February 2016). "Overview of Groundwater in India" (PDF). PRS India.org. Retrieved 9 April 2018.
  61. ^ Centre approves Rs 6,000 crore scheme to manage groundwater, Times of India, 24 December 2019.
  62. ^ "National Water Policy 2002" (PDF). Ministry of Water Resources (GOI). 1 April 2002. p. 2. Archived from the original (PDF) on 18 January 2012. Retrieved 15 August 2012.
  63. ^ Rutherford, Susan (2004). Groundwater Use in Canada (PDF).
  64. ^ Côté, Francois (6 February 2006). "Freshwater Management in Canada: IV. Groundwater" (PDF). Library of Parliament.
  65. ^ Richard Greenburg (2005). The Ocean Moon: Search for an Alien Biosphere. Springer Praxis Books.