Vous êtes sur la page 1sur 5

The Water Cycle

SJ Marshall, University of Calgary, Calgary, AB, Canada


ã 2014 Elsevier Inc. All rights reserved.

Introduction 1
Evaporation and Transpiration 1
Precipitation 2
Runoff 3
Watershed Processes 3
Hydrological Models 4
References 4

Introduction

The hydrological cycle describes exchanges of water between the oceans, atmosphere, land surface, biosphere, soils, groundwater
systems, and the solid Earth. The overview article on Hydrology considers aspects of the global water cycle and the amount of water
stored in these different reservoirs. Figure 1 indicates the global-scale fluxes of water between these reservoirs, expressed in
1012 m3 yr1 (trillions of m3 yr1). Evaporation and precipitation over the world’s oceans dominate the fluxes within the global
water cycle, and are truly prolific. Ocean evaporation rates, estimated at 424 trillion m3 yr1, equate to 1.2 m of water skimmed off
the surface of the global ocean each year.
There is tremendous turnover within the global water cycle; the average residence time of a water molecule in the atmosphere is
about 9.2 days. Roughly 90% of water molecules cycle quickly through the atmosphere, evaporating and precipitating over the
world’s oceans (Figure 1). Water vapor that is blown inland and precipitates over the continents has a more circuitous path back to
the oceans, and can get diverted and stored in snowpacks, vegetation, soils, wetlands, lakes, or groundwater systems. Water
molecules can be ‘recycled’ many times on the continents, through cycles of evapotranspiration and precipitation. Eventually, rivers
and subsurface drainage systems return this water to the ocean, and over a period of a year or more this discharge equals the net
evaporation minus precipitation from the oceans.
This must be true or global sea level would be highly unstable (i.e., dropping by 1.2 m yr1 if rivers did not return water to the
ocean). There are exceptions to this rule, particularly over longer timescales of Earth history. Interannual variability in precipitation,
continental storage, and runoff rates can give small-scale (mm yr1) fluctuations in sea level (e.g., Boening et al., 2012). Growth
and decline of glaciers and ice sheets involves large quantities of water storage on land. During the Quaternary glaciations, for
instance, ice-sheet buildup over many thousand years led to a sea-level drawdown of more than 120 m.
In practice, hydrologists typically focus on terrestrial environments and examine the water balance for a particular location or
region, considering the inputs, outputs, exchanges, and storage of water within that region of interest. There are many scales of
interest with respect to water balance. Examples include agricultural fields, lakes or reservoirs, river basins, political jurisdictions
(e.g., a state or nation), or larger regions which are vulnerable to hydrological extremes such as drought (e.g., East Africa, the
western U.S.). Water management practices depend critically on accurate accounting of water balance on seasonal to multiannual
time scales, while shorter-term water fluctuations are important to many other considerations such as crop growth, river flows,
wildfires, and flood events.
Inputs to a region come in the form of precipitation and water flow (e.g., rivers, springs), while outputs include evaporation,
transpiration, biological uptake, and surface or groundwater flows that transport water out of the region of interest.
Within a region, there are exchanges between the surface environment (lakes, rivers, wetlands), soils, vegetation, the ground-
water system, and human uptake in service of agricultural, industrial, and municipal water demands. Water can also be stored for
extended periods, amounting to many years in the case of some large lakes and groundwater systems. This generalized water
balance for a region (or a specific lake or catchment) can be expressed as:

dV
¼ S ¼ QIN  QOUT [1]
dt
where V is the volume of water, t is time, S is the rate of storage (or release) of water, and Q refers to the various volume fluxes of
water into or out of the region.

Evaporation and Transpiration

Evaporation is the phase change of water from liquid to vapor, driven primarily by energy from solar radiation. Water in the
atmosphere has in fact been referred to as ‘liquid sunlight,’ as the latent energy stored in water vapor and released during

Reference Module in Earth Systems and Environmental Sciences http://dx.doi.org/10.1016/B978-0-12-409548-9.09091-6 1


2 The Water Cycle

Figure 1 The global water cycle, with fluxes in 1012 m3 yr1 after the U.S. University Corporation for Atmospheric Research, https://spark.ucar.edu/
longcontent/water-cycle, with updates from Durack et al. (2012). Graphic adapted from NOAA National Weather Service, http://www.srh.noaa.gov/
jetstream/index.htm.

condensation accounts for a large transfer of energy from tropical and subtropical regions (where most evaporation occurs) to
higher latitudes. Transpiration is the release of water vapor to the atmosphere during photosynthetic activity in plants. Water is
drawn into vegetation from the soil and water vapor is diffused to the atmosphere through plant stomata. This is costly for the
plant; roughly 98% of water transpires to the atmosphere rather than being used for photosynthesis or plant cellular growth (e.g.,
Baldocchi et al., 1987), but it is an unavoidable consequence of opening up stomata to take in CO2 during photosynthesis. Solar
radiation again provides the main source of energy driving the phase change in transpiration.
Together, evaporation and transpiration release about 495  1012 m3 to the atmosphere each year, with 86% of this derived
from ocean evaporation (Figure 1). Of the 71  1012 m3 derived from land, roughly 30% is due to transpiration and 70% is from
direct evaporation of lakes, wetlands, and soil water. Changing water levels in evaporation pans provide direct measurements of
evaporation rates, or natural water bodies can provide similar inferences. It is generally difficult to measure transpiration or larger-
scale evaporation rates, however (e.g., over the ocean), so these tend to be inferred from water balance (for instance, changes in soil
moisture or atmospheric humidity). Empirical formulae have also been developed to estimate potential evapotranspiration as a
function of vegetation type and atmospheric conditions, but these are approximate as actual evapotranspiration rates depend on
the availability of soil moisture and the life cycle of the plant.

Precipitation

About 78% of global precipitation falls over the oceans, with the remaining 22% distributed unevenly over the landscape. Water
vapor is transported in the atmosphere, with roughly 39  1012 m3 advected inland from the oceans each year. The additional
71  1012 m3 from evapotranspiration over land gives a total atmospheric water volume of 110  1012 m3 that falls as precipitation
over the continents each year. When averaged over Earth’s land mass, this equates to an average terrestrial precipitation rate of
740 mm yr1.
Condensation of water vapor occurs when air masses are cooled enough to become saturated, inducing cloud development.
Precipitation can follow where there is enough moisture and cloud droplets grow to a large enough mass to overcome updrafts and
air resistance and fall to the ground as hydrometeors. In-cloud processes that produce precipitation can be very complex, and are
described in detail in the Atmospheric Sciences section of the ESES module. Large-scale precipitation processes and precipitation
patterns are discussed in the Water Cycle section. The main ways to cool an air mass and wring out moisture involve four different
mechanisms: (i) orographic uplift (forced uplift over mountain ranges), (ii) frontal uplift where air masses collide, particularly
along the polar front and in mid-latitude cyclones, (iii) isobaric cooling as air masses travel inland or to higher latitudes, and
(iv) adiabatic cooling during convection (buoyancy-driven uplift). The windward sides of mountain ranges receive high precip-
itation totals, particularly where mountains are parallel to coastlines and intersect incoming maritime air masses (e.g., coastal
Norway, New Zealand, western Canada). Mid-latitude ‘storm tracks’ as well as stationary low pressure systems along the polar front,
The Water Cycle 3

such as the Aleutian and Icelandic lows, see frequent frontal precipitation, while hot, humid, and unstable air in the tropics gives
rise to high amounts of convective precipitation.
Regions of the world with high annual precipitation totals typically experience one or more of these conditions. Conversely,
stable air masses in polar regions and in the subtropical desert belts are associated with low precipitation totals, particularly in
interior continental regions where moisture sources are limited.

Runoff

Surface water collects in lakes, wetlands, and soils, or it flows downhill under the force of gravity where topographic gradients and
river channels permit. River discharge is the dominant form of surface runoff, and effectively drains most of the land mass on the
planet. Water that drains into the subsurface becomes part of the groundwater system, where it may be stored in aquifers or it may
drain under the control of hydraulic gradients, the combination of topographic and pressure gradients. Subsurface drainage does
not always follow surface slopes, as it depends on the water levels (pressure head) in the local and regional groundwater system, but
in general groundwater pathways also help to drain water from high to low elevations, discharging in river valleys and estuaries and
contributing to continental runoff to the oceans. Articles in the Runoff section examine different aspects of hydrological drainage,
overland vs. channel flow, and river hydraulics.

Watershed Processes

While decision makers in political jurisdictions need to understand their regional water balance for water resource management,
the most natural way to examine water budgets is to delineate hydrological catchments (also known as watersheds or river basins),
which commonly transcend political boundaries. Watersheds are delineated by surface topography and drainage patterns;
precipitation that falls in a given watershed eventually discharges through a common drainage channel at the lowest point in
the watershed. It is perhaps easiest to think of watersheds in the context of a major river basin such as the Amazon. Precipitation on
the western slopes of the Andes in Columbia, Ecuador, and Peru drains to the Pacific Ocean, while rainfall on the eastern slopes, in
the upper Amazon basin, drains to the Atlantic Ocean via the Amazon River (Figure 2).

Figure 2 The Amazon River Basin, indicating the dendritic network of watersheds that drain into the main trunk of the Amazon. Graphic courtesy of the
World Wildlife Fund http://wwf.panda.org/what_we_do/where_we_work/amazon/about_the_amazon/.
4 The Water Cycle

Figure 3 Example of the processes considered in a distributed hydrological model, including 3D water fluxes in the groundwater and surface systems,
a moving water table, a treatment of vegetation (canopy interception, transpiration), and exchanges with the atmosphere (precipitation,
evapotranspiration). Graphic adapted from http://www.dhigroup.com/News/2009/02/02/OpenMIForClimateModelling.aspx.

Numerous sub-basins exist within any major river like the Amazon, where smaller hydrological catchments collect water and
deliver it to the main drainage trunk through tributaries. It is convenient to examine the water balance of a watershed rather than an
arbitrary region because the upper watershed boundary is an inflection point in the terrain (a local drainage divide), and there are
no river flows into the basin. Hence the water flux Qin in eqn [1] is restricted to just the precipitation that enters the system. It is still
difficult to estimate the runoff and river discharge from the basin, as numerous other hydrological processes are at work involving
storage, delays, and exchanges of water within the basin (e.g., Figure 3).

Hydrological Models

Hydrological models are used to simulate the processes and exchanges of water within a catchment or, in the case of global climate
models, over the land surface on a continental scale. Catchment-scale models provide guidance to water resource management,
hydroelectric power production, flood forecasts, and numerous other applications.
Figure 3 depicts the typical components of a distributed hydrological model, which evaluates water fluxes, storage, and phase
changes within a three-dimensional representation of the landscape and the subsurface. The subsurface is typically divided into the
root zone, where vegetation takes up water and soil water is subject to evaporation, the underlying unsaturated zone, and the
saturated groundwater system that lies beneath. The ‘water table’ is the horizon separating the unsaturated and groundwater zones.
Exchanges of moisture and energy with the atmosphere can be prescribed (i.e., assigning a certain amount of precipitation and
sunlight according to observations, along with other meteorological conditions that matter for evapotranspiration (wind, humid-
ity, etc.)). Alternatively, hydrological models can be coupled with atmospheric models in an Earth system modeling framework,
where land surface feedbacks on the atmosphere and details of surface energy balance, moisture fluxes, etc., are explicitly included.
There are numerous approaches to hydrological modeling, described in detail in this unit of the Hydrology section.

References
Abbott MB and Refsgaard JC (eds.) (1996) Distributed hydrological modelling. Water science and technology library, 22 vols. Dordrecht, The Netherlands: Kluwer Academic
Publishers.
Baldocchi DB, Verma SB, and Anderson DE (1987) Canopy photosynthesis and water-use efficiency in a deciduous forest. Journal of Applied Ecology 24(1): 251–260.
Boening C, Willis JK, Landerer FW, Nerem RS, and Fasullo J (2012) The 2011 La Niña: So strong, the oceans fell. Geophysical Research Letters 39(19). http://dx.doi.org/10.1029/
2012GL053055.
The Water Cycle 5

Durack PJ, Wijffels SE, and Matear RJ (2012) Ocean salinities reveal strong global water cycle intensification during 1950 to 2000. Science 336: 455–458.
Shuttleworth WJ (2012) Terrestrial hydrometeorology. Oxford, U.K.: Wiley-Blackwell.
Yu L (2007) Global variations in oceanic evaporation (1958–2005): The role of the changing wind speed. Journal of Climate 20: 5376–5390.

Further Reading
Standard textbooks on hydrometeorology discuss atmospheric water vapor, evaporation, and precipitation processes in detail. Shuttleworth (2012) offers a thorough treatment. Abbot
and Refsgaard (1996) discuss numerous aspects of hydrological modeling.

Vous aimerez peut-être aussi