1: Introductory Hydrology

Back to Home

Outline

1. Definition of Hydrological Science

2. Why Study Surface Hydrology?

3. Hydrological Cycle


1. What is hydrological science?

From COHS, 1991.

A Distinct Geoscience

Over the past 60 years, the evolution of hydrologic science has been in the direction of ever-increasing space and time scales, from small catchment to large river basin to earth system, and from storm event to seasonal cycle to climatic trend. Hydrologic science should be viewed as a geoscience interactive on a wide range of space and time scales with the ocean, atmospheric, and solid earth sciences as well as with plant and animal sciences. To establish and retain the individuality of hydrologic science as a distinct geoscience, its domain is defined as follows:

Continental water processes-- the physical and chemical processes characterizing or driven by the cycling of continental water (solid, liquid and vapor) at all scales (from the micropores of soil water to the global processes of hydroclimatology) as well as those biological processes that interact significantly with the water cycle.

(This restrictive treatment of biological processes is meant to include those that are an active part of the water cycle, such as vegetal transpiration and many human activities, but to exclude those that merely respond to water, such as the life cycle of aquatic organisms.)

Global Water Balance-- the spatial and temporal characteristics of the water balance (solid, liquid and vapor) in all compartments of the global system: atmosphere, oceans and continents.

(This includes water masses, residence times, interfacial fluxes, and pathways between compartments. It does not include those physical, chemical or biological processes internal to the atmospheric and ocean compartments.)

 

The scope of this class (i.e. surface water hydrology) deals with those continental hydrological processes occurring at or near the land surface

Hydrologic science deals with the waters of the earth, including their

The science of hydrology embraces the full life history of water on the earth.

Hydrology is closely related to other natural sciences:

It is a multi-disciplinary science since water interacts (i.e. it is important to and affected by)

within all compartments of the earth system, i.e. the

as shown in COHS (1991) Figure 2.4.

 

2. Why do we need to study surface hydrology?

Since water pervades the earth system and interacts with its various components on a range of spatial and temporal scales, land surface hydrology plays a critical, if not central role, in the interactive functioning of the earth system. Therefore, it is tightly coupled to numerous earth system processes ( e.g. climate, weather, biogeochemistry, ecosystem dynamics).

For example, at large (e.g. continental) scales, soil water stored in soils near the land surface exerts a major control on climatic and weather processes by affecting land surface temperatures and by supplying moisture to the atmosphere via evapotranspiration. We are now suspecting that runoff from the continents affects ocean circulation, which via their control of sea surface temperatures, that have a major impact on climate.

At regional scales, land use practices such as irrigation, grazing, and deforestation have feedbacks to regional hydrology, climate, and weather.

At smaller scales, practical issues of water resources management, flood control

flood forecasting, water supply forecasting and groundwater quality require an intimate knowledge of surface hydrologic processes to really grasp the big picture and to function effectively in water resources careers in engineering/consulting/planning, government or academia.

Given the above background, a primary goal of this class is to educate you with the perspective that the hydrologic cycle is a fundamental earth system cycle, interactive with other earth system components across a broad range of spatial-temporal scales. We’ll see that this cross-disciplinary theme will be echoed throughout the semester.

 

3. Hydrological Cycle

The pathway of water as it moves in its various phases, through the atmosphere, over and through the land, to the ocean, and back to the atmosphere, is known as the hydrological cycle. It is the most fundamental principle of hydrologic science. These pathways are shown graphically in Figure 1.4 of Fetter, 1994.

Precipitation falling on land takes various pathways through the hydrological cycle

 

Some of the rain will drain across the landscape toward stream channels as overland flow (runoff).

Some of the rain will seep into porous soils through infiltration.

Below the land surface, the region where soil pores contain both air and water is called the unsaturated zone (zone of aeration, vadose zone).

Water stored in the unsaturated zone is called soil moisture (soil water). A measure of the amount of water in the soil is called the moisture content, or soil water content.

Near surface soil moisture is extracted by roots, and is then transpired. Water in the unsaturated zone can also migrate back to the land surface to evaporate.

Under some conditions, water can flow laterally in the unsaturated zone, e.g. through macropores, a process known as interflow.

Some water in the unsaturated zone is pulled downward by gravity drainage (percolation).

At some depth, the pores of the soil are saturated with water. The top of the saturated zone is called the water table

Water stored in the saturated zone is called groundwater. It moves through the soil layer as saturated subsurface flow (baseflow, groundwater flow) until it discharges as a spring, seeps into a pond or stream, or the ocean.

Water flowing in a stream derives from overland flow, from groundwater that has seeped into the stream channel, and perhaps from interflow.

Many of the pathways that water takes once on land have specific direction. For example, baseflow, overland flow, streamflow are topographically-driven downslope. Others are more ubiquitous, provided there is a moisture source (e.g. infiltration occurs over large spatial areas after rain events; evaporation occurs from soil, plants, from water intercepted by plant canopies, and from any open water bodies (oceans, lakes, streams)). See Fetter, (1994) Figure 1.5.


References

Committee on Opportunities in the Hydrologic Sciences (COHS), Water Science and Technology Board, Commission on Geosciences, Environment and Resources, National Research Council, 1991. Opportunities in the Hydrologic Sciences, , National Academy Press, Washington, DC.

Fetter, C. W., 1994. Applied Hydrogeology, Macmillan College Publishing Co., New York.


Back to top of page

 

 

 

 

2: Hydrology Basics I

 

Outline

1. Conservation Equations

2. Water Balance Examples


1. The Conservation Equations

A system is any conceptually-defined region of space capable of

 

A system is sometimes called a control volume. See Dingman, 1994, Figure 2-1 for a conceptual diagram of a system.

A conservative quantity is one that cannot be created or destroyed within the system, for example

The basic conservation equation states:

The amount of a conservative quantity entering a control volume during a defined time period, minus the amount leaving the volume during the time period, equals the change in the amount of the quantity

This basic conservation equation is in fact a generalization of the conservation of mass, Newton's first law (conservation of momentum) and the first law of thermodynamics (conservation of energy)

Stated more simply

Amount in - Amount out = Change in storage
 
or
 
I - Q = DS

remembering of course that this is only true for

Other useful forms can be obtained by dividing by DT as shown in the text, yielding

dS/dt = i - q

where i and q are the instantaneous rates of inflow and outflow and dS/dt is the instantaneous rate of change of storage.

Note now that there are no restrictions on the size of the control volume or the length of the time period so that the conservation equations are scale and time independent

When applied to problems of the mass of water moving through the hydrological cycle, they are called water balance equations. Since the conservation equations are space-time independent, we can apply them from the point, or infinitesimally small spatial scales, up to the global scale, and on anything from instantaneous to very long time scales. When the conservation equations are applied to energy they are called the energy balance equations.

 

2. Water balance examples

Since the conservation equation is scale independent, we could apply it at

Examples from class: the root zone of a soil column, a water balance on a lake, and a watershed on a drainage basin.

With respect to the drainage basin example, we first need to define a drainage basin (watershed, catchment, river basin) as the area which drains through a particular point in a stream network, i.e. it is the area that topographically appears to contribute all the water that passes through a given cross section of stream. Drainage basins are outlined by topographic divides. These are the boundaries which delimit watersheds. See Dingman, 1994, Figure 2-5 for an example of a watershed delineated on a topographic map.

Watersheds can be delineated for any point on the stream network, so there are an infinite number of watersheds that can be drawn for any stream. Delineation is most often done manually, but can now be done automatically using digital elevation models (DEMs). DEMs are simply computer data files that give land-surface elevations at grid points.

Regional water balance example

The regional water balance is the application of the water balance equation (conservation equation) to a watershed (or to any land area such as a state or continent).

The watershed area delimited by its divide is the upper surface of the control volume; the sides of the volume extend vertically downward from the divide some indefinite distance assumed to be below the level of significant ground-water movement. (See Dingman, 1994, Figure 2-6)

From Dingman (1994) Figure 2-6 we can write the water balance equation for watershed

DS = I - Q

Inputs Outputs

Precipitation(P) Streamflow(Q)

Groundwaterout(GWo)

DS = P + GWi - (Q + ET + GWo)

Choice of Dt. Choose Dt for your problem of interest -- instantanoeous, daily, monthly, seasonal, annual, etc. Since many of these variables are hard to quantify over large areas (e.g. ET, GWi,o), one trick is to eliminate some of the unknowns. For example, if Dt is taken over a long enough period of time (e.g. annual) then DS can be taken as zero.

 

Estimation of regional ET. One of the most common applications of the regional water balance approach is in estimating regional ET. Additional assumptions are required to simplify the draingage basin water balance equation: assume that surface water divides are equal to groundwater divides, and take GWi as zero. Then we can rewrite water balance as

P - Q - ET - GWo= 0

Assume GWo is negligible (since much of the groundwater leaving the basin may in reality leave as baseflow). Then

P - Q - ET = 0

(A note on units: Water balance variables such as P, or ET are frequently given as depths (L), i.e. volumes (L3) divided by the area (L2) over which they fell. This depth quantity could be further divided by the duration of Dt to get units of L/T)

 

Sources of error in the regional estimation of ET: model and measurement error introduce uncertainty into the water balance which propagates into, e.g. estimated ET.

Model error: the omission of terms in the basic water balance equation:

Measurement error: uncertainties due to measurement error are always present in hydrology. Briefly, with respect to regional ET estimation, errors in P and Q will propagate into errors in ET

 


Back to top of page

 

 

 

3: Hydrology Basics II

 

Outline

1. Energy Balance


1. Energy Balance

The conservation equation as applied to energy, or conservation of energy, is known as the energy balance. In general, just as we looked at how precipitation is partitioned into infiltration, runoff, evapotranspriation, etc., we can look at how incoming radiation from the sun and from the atmosphere is partitioned into different energy fluxes (where the term flux denotes a rate of tranfer (e.g. of mass, energy or momentum) per unit area).

 

We’ll see later (as we move through the course) that the water and energy balance are intimately linked:

 

Just as changes in water balance were reflected in changes in storage in water amounts (soil moisture in a root zone; level of a lake) changes in energy balance are reflected in temperature changes. Just as we wrote water balances for a number of different control volumes in the last class, we could write energy balances for the same control volumes. However, we won’t really do that because it’s beyond the scope of what we’re trying to do in this class.

Instead, we’ll look at a simplified energy budget for an ideal surface, where we define ideal as

Later, in the chapter on snow, we’ll look at the energy balance in a little more detail and add heat capacity and storage of heat in a snow layer.

The simplified energy budget for such an ideal surface is given by

Rn = H + LE + G

where

Rn is defined as net radiation at the surface;
H is defined as sensible heat exchanged with the atmosphere;
LE is defined as latent heat exchanged with the atmosphere; and
G is defined as heat exchanged with the ground.

See Arya, Figure 2.1 for a schematic representation of the simplified energy budget.

Net Radiation.

Net radiation is really composed of shortwave radiation, K, from the sun, and longwave radiation, L from the atmosphere and from the ground, so that

Rn = K + L

The radiation from the sun (solar radiation) is often referred to as shortwave radiation, and the radiation from the atmosphere and the ground (i.e. atmospheric and terrestrial radiation) as longwave radiation, since the wavelength of the electromagnetic emitted by these bodies is inversely proportional to their temperatures (See Figure 3.1 from Arya (1988)).

Shortwave radiation input. First, consider what happens to incoming solar radiation as it enters the earths atmosphere on the way to the surface. Figure 2.8.4 from Chow et al. (1988) shows that it may be:

The net flux of solar energy entering the land surface is therefore given as

K = Kin - Kout = Kin (1-a)

where

K in is the incident solar energy on the surface, and it includes direct solar radiation (i.e. that which makes it through the atmosphere unscathed) and diffuse (due to scattering by aerosols and gases);
Kout is the reflected flux;
a is the albedo

Solar radiation is measured in specialized meteorological stations with radiometers.

However, these require careful calibration and maintenance, and consequently they are not usually available at standard weather stations. Dingman (1994), Figure E1, shows the weather stations across the US at which solar radiation data are routinely collected. A more common practice is to estimate K by modifying the clear-sky solar for the geographic location and time of year by functions of slope orientation, fraction of clouds, and fraction of vegetation (See Dingman (1994), Appendix E).

Longwave radiation input. All matter at a temperature above absolute zero radiates energy in the form of electromagnetic radiation which travels at the speed of light.

The rate at which this energy is emitted is given by the Stefan-Boltzman law

Qr = esT4

where

Qr is the rate of energy emission per unit surface area per unit time
T is the absolute temperature of the surface
s is a universal constant called the Stefan-Boltzman constant
e is a dimensionless quantity called the emissivity

The emissivity ranges from 0 to 1, depending on the material and surface texture. A surface with e equal to 1 is called a blackbody. Most earth materials have emissivities near 1.

Long-wave radiation is emitted by bodies at near earth surface temperatures (the land surface and the lower atmosphere). The net input of longwave radiation, L, is the difference between the incident flux, Lin, which is emitted by the atmosphere, clouds

and overlying vegetation canopy, and the outgoing radiation emitted from the land surface:

L = Lin - Lout

Longwave radiation is measured using radiometers. As in the case of shortwave, the instruments are rare except at intense research sites. So, it is usually estimated from more readily available meteorological information. These estimates are based on the following physics.

 

The flux or radiation emitted by the atmosphere is

Lin = eatsTat4

and at denotes atmosphere . Outgoing radiation is the sum of the radiation emitted by the surface and that fraction of the incoming longwave that is reflected

Lout = essTs4+(1-es)Lin

The subscript s denotes land surface. For the case of gray bodies (e<1) reflectivity equals 1 - emissivity.

 

Sensible Heat

H = cara/rah (Ta-Ts)

Note that this equation is essentially a conductivity times a gradient, corrected for the properties of the fluid.

ca is the heat capacity of air
ra is the density of air
rah is the aerodynamic resistance to heat transport and is given by
 

rah = 1 / k2 u (z(m)) {ln{zm/zo}}2

k is von Karmann's constant (0.4)
u (z(m)) is wind speed at measurement height zm
zo is known as the roughness height of the surface and depends of the irregularity of the surface

 

Latent Heat

LE = cara/ grav (es-ea)

where

L is the latent heat of vaporization
E is the rate of evaporation
rav = rah
g is the psychrometric constant, and is a function of atmospheric pressure, density of air, etc.
es,ea are the vapor pressures measured at the surface and in the lower atmosphere

 

Ground Heat Flux

G = kG dT/dz

where

kG is the thermal conductivity of the soil
dT/dz is the vertical temperature gradient

Thermal conductivities of soils depend on soil texture, soil density, and moisture content, and vary widely in space. Owing to this variability, and the fact that dT/dz is tough to measure, G is often neglected or estimated in energy balance computations.


References

 
Arya, S. Pal, Introduction to Micrometeorology, 1988, Academic Press, New York.
 
Chow, V. T., D. R. Maidment and L. W. Mays, 1988. Applied Hydrology, McGraw-Hill, Inc.
 
Dingman, S. L., 1994. Physical Hydrology, MacMillan Publishing Company, New York.

Back to top of page

 

 

 

4: Water in the Atmosphere

Outline

1. Composition of the Atmosphere

2. Atmospheric Circulation

3. Quantification of amount of water in atmosphere


1. Composition of the atmosphere

The atmosphere is a mixture of gases in which liquid and solid particles are suspended.

Table D-3 (Dingman, 1994) lists these various components. The major constituents, and many of the minor ones, are effectively constant in time. The variable constituents include those that are most significant hydrologically (e.g. liquid water, ice, water vapor, dusts, and CO2 ) since they effect the water and energy balance of the atmosphere and thus the formation of clouds and precipiation. The amount of water vapor in the atmosphere amounts to less than 1 part in 100,000 (0.001 %) of all the waters of the earth (0.04% of the fresh water), yet it plays a vital role in the hydrological cycle and the earth's climate system. Most of the water vapor in the atmosphere is concentrated in the lower 2 km of the atmosphere

2. Atmospheric circulation

The earth constantly receives heat from the sun as solar radiation, and emits heat back into space by back radiation. This heating of the earth is uneven, since, as seen in Dingman (1994) Figure 3-4, incoming radiation near the equator is almost perpendicular to the land surface (averaging 270 W/m2), while near the poles it strikes at a more oblique angle(averaging 90 W/m2).

Since the back radiation is proportional to the temperature of the earth's surface, which doesn't vary that much from the equator to the poles, the emitted radiation is more uniform than the incoming solar, so that in response to the imbalance, the atmosphere functions as a vast heat engine, transferring energy from the equator to the poles.

If the earth were a non-rotating sphere, the atmospheric circulation would appear as in Chow, Maidment, and Mays, 1988, Figure 3.1.1. Air would rise near the equator, travel towards the poles in the upper atmosphere, cool, and decend into the lower atmosphere. This one-cell circulation is called the Hadley circulation.

However, the earth does rotate, and the combined effects of the rotation of the earth and the poleward transport of heat energy are the driving forces behind the general circulation of the atmosphere which is shown in Chow, Maidment, and Mays, 1988, Figure 3.1.2.

The actual pattern of the atmospheric circulation has 3 cells: a tropical cell, a polar cell, and a middle cell, as well as jet streams and prevailing surface wind directions (See Dingman, 1994, Figure 3-8).

Hartmann (1994), Figure 1.2, shows that the atmosphere is divided vertically into various zones. The circulation described above occurs in the troposphere, which ranges in height from 8 km at the poles to 16 km at the equator. The troposphere accounts for more than 80% of the mass and virtually all of the water vapor, clouds, and precipitation in the earth's atmosphere. It is characterized by strong vertical mixing, e.g. in clear air it is not unusual for a parcel, or small control volume of air (containing water vapor) to travel the height of the troposphere in a few days (or minutes during a thunder storm).

In the troposphere, temperature decreases with altitude at a rate varying with its moisture content. The rate of decrease of tempereature with altitude in dry air is called the dry adiabatic lapse rate, and is roughly 9.8 deg C/km. Adiabatic implies no heat exchanged with surroundings. In moist air, the moist adiabatic lapse rate is typically half of the dry rate, or 5 deg C/km -- since the atmosphere is heated somewhat by the formation of precipitation. The average lapse rate in the troposphere is a weighted average of the dry and moist rates, and is roughly 6.5 deg C/km.

The top of the troposphere is called the tropopause. It is overlain by the stratosphere, stratopause, mesosphere, mesopause, etc.

 

3. Quantification of the amount of water vapor in the atmosphere

The atmosphere is a mixture of gases, each exerting a partial pressure proportional to its concentration, owing to its molecular motion and collision with other gases.

The vapor pressure is computed from the ideal gas law as

e = rvRvTa

where

rv is the water vapor density (absolute humidity), i.e. the the mass concentration of water vapor in a volume of air;

Rv is the gas constant for water vapor; and

Ta is the air temperature

It can also be written in terms of atmospheric pressure, P, and the specific humidity, qv, (the mass of water vapor per unit mass of moist air rv/ra) as

e = qvP/.622

The maximum vapor pressure that is thermdynamically stable is called the saturation vapor pressure. It is a function of temperature only ( Dingman, 1994, Figure D-3) and can be computed as

esat(T) = 6.11 exp ((17.3 T)/(T+237.3))

where esat(T) is in mb and T in deg C.

A parcel of air at T and e below the esat(T) curve is unsaturated. The relative humidity, Wa of this parcel is the ratio of its actual vapor pressure to its saturation vapor pressure

Wa = e/esat

If the parcel is now cooled, as in Dingman (1994) Figure D-3, its saturation vapor pressure drops, so that its relative humidity increases. When the parcel is cooled to the point on the esat(T) curve it has reached its saturation vapor pressure, and its relative humidity of 100%.

The temperature to which a parcel with a given vapor pressure must be cooled in order to reach saturation is called the dew point


References

Dingman, S. L., 1994. Physical Hydrology, MacMillan Publishing Company, New York.

Hartmann, D. L. 1994. Global Physical Climatology, Academic Press, New York.


Back to top of page

 

 

5: Precipitation I

Outline

1. Requirements for precipitation formation
2. Mechanisms of air mass lifting
3. Moisture sources and transport

Precipitation is defined as the condensation of water to liquid or solid form, and subsequent falling to earth. It includes:

1. Requirements for formation of precipitation

In order for hydrologically significant rates of precipitation to occur, a sequence of several conditions must be met. These include

  1. air mass cooling
  2. condensation or freezing nuclei must be present
  3. droplets must form and grow
  4. a source of water vapor must be maintained

Air mass cooling

The formation of precipitation requires that an air mass be cooled to its dew point and some of its moisture condenses

An air mass in the general circulation is a large body of air that is fairly uniform horizontally in properties such as temperature and moisture content. Air masses are sub-continental in scale

Air masses can be cooled in a number of ways:

  1. radiation to cooler surroundings
  2. mixing with cooler air
  3. conduction to a cool surface
  4. adiabatic cooling by horizontal movement to a region of lower pressure
  5. adiabatic cooling by vertical uplift

Cooling by these first four may result in fog or drizzle. Only vertical lifting can cause cooling high enough to produce significant amounts of rainfall.

From the ideal gas law we know that

P = rRT

and that pressure decreases with height into the atmosphere.

Therefore,

This is shown schematically in Dingman (1994) Figure D-6.

Condensation or freezing nuclei must be present

Condensation requires a seed called a condensation nucleus (cloud condensation nucleus or CCN) e.g. dust, particles containing ions like sea salt, sulfur and nitrogen from combustion) around which water molecules can attach themselves.

CCN are typically dust or particles containing ions. Major natural sources include

Natural concentrations of CCN are

Human activities can raise this number over land by a factor of 10 (e.g. combustion products including sulfur and nitrogen products that lead to acid rain)

 

Droplet formation and growth

After condensing on CCN, drop size initially increases by continued condensation. Cloud droplets have sizes that range from 0.001 to 0.2 mm, and have fall velocities between 0.01 and 70 cm/s. For precipitation to fall from clouds, droplets must continue to grow to a size such that they are large enough so their terminal velocity is greater than the velocity of the updrafts that lifted them to the LCL, keeping in mind that there is evaporation as the drops start to fall. Therefore, the droplets must grow a few orders of magnitude to sizes on the order of 0.4 to 4 mm in diameter, or to snowflakes or even larger sizes.

Continued growth occurs by

• collision

• ice crystal growth

Collision: at temperatures above 0 deg C, produce droplets of varying sizes.

See Chow, Maidment, Mays (1988) Figure 3.3.1.

Ice crystal growth. In saturated air less than -40 deg C, vibrational energy of H20 molecules is low enough so that clusters of the molecules can spontaneously form ice crystals. When saturated air is between -40 and 0, a CCN with a molecular structure similar to that of ice is required to nucleate ice crystal growth. Given that some nucleation has occurred, since the saturation vapor pressure of ice is always lower than that of water for a given temperature below zero, water will evaporate from the liquid water and condense on the ice crystals. Growth continues in this fashion until the ice crystals are large enough to fall from the cloud.

Importation of Water Vapor

 

The concentration of liquid water and/or ice in most clouds is 0.1 to 1 g / m3. Consider a 10,000 m thick cloud, so that the total cloud volume over a m2 of ground is 10,000 m3.

If the concentration of water in the cloud is 0.5 g/m3 , the total volume of cloud water above each m2 is 5000 cm3. If all this water fell as precipitation, its depth would be 0.5cm, which is not much rain. Since most rain-producing clouds are less than 10,000 m in thickness, and water concentrations are less than 0.5 g/m3, a final requirement for the occurrence of significant amounts of precipitation is that a continual water vapor supply be imported to clouds to replace the rain that falls. This inflow of moisture is produced by winds that converge on precipitation producing clouds.

2. Mechanisms of air mass lifting

The three major mechanisms of air mass lifting include uplift due to

1. convergence

2. convection

3. orography

Convergence:

Airflow is induced by pressure gradients and movement is generally towards regions of low pressure. Looking back at the Dingman's Figure 3-8 we can imagine that high pressure regions are located where air descends from atmospheric cells, and low pressures where air is heated and rises. Thus low pressure areas are regions of convergence

• air converging from several directions is forced to rise

• adiabatic cooling results

• frontal convergence is characteristic of mid latitudes

• non-frontal convergence is a tropical phenomenon

Frontal convergence:. Outside of the tropics, in the mid latitudes, much of the precipitation results from frontal convergence (cyclonic convergence)

A cyclone is the convergence of air masses around a low pressure system.

• Air masses with contrasting temperatures, moisture contents and densities converge around the low

• The boundaries between air masses are called fronts

• In the northern hemisphere, the circulation is counterclockwise

• The diameter of a fully developed extratropical cyclone is about 1500 km

• Because of the temperature differences between the air masses, they tend not to mix, rather warm air is lifted over cold

Cold front: a cold air mass moves into a warm one characterized by

  • steep slopes 1:30 1:40
  • rain is more intense and restricted to a narrow region around the surface front
  • moves faster than warm front
  • lifting faster than warm front
  • Warm front: a warm air mass displaces a cold one characterized by

  • gentle slopes 1:60 1:120
  • precipitation lighter and spread over a larger area ahead of the front
  • moves slower than cold front
  • lifting slower than cold front
  • Nonfrontal convergence. Convergence at the ITCZ: the low latitude Hadley cells create a tendency for convergence that circles the globe in tropical regions. This leads to adiabatic lifting and a band of heavy rain that circles the equator.

    Tropical cyclones (Hurricanes) Tropical cyclones are non-frontal features and can develop into severe storms (hurricanes, typhoons, cyclonesm etc)

    Uplift due to convection:

    Convective precipitation results from adiabatic cooling when parcels of air at the surface are heated and rise because they are less dense. Look at Dingman (1994) Figure 4-7

    Uplift due to orography

    In most parts of the world, long-term mean precipitation increases with elevation (orographic effect)

    Critical temperature for rain/snow transition.

    Much of the precip that falls outside the tropics originates as snow or hail in supercooled clouds. The formation of precipitation reaching the surface is determined largely by the height of the 0 deg C surface. Rain occurs if the 0 deg surface is high enough so that the precipitation can melt before it reaches the ground

    3. Moisture sources and transport

    In the first half of the century, it was widely believed that most of the moisture falling as precipitation on a given land area entered the atmosphere as evaporation from the same or nearby areas.

    However, starting around 1940, studies began to show that

    1. most of the precipitation falling on land came from maritime air masses which obtained their moisture from the sea; and
    2. most of the evapotranspiration from the land moved off the continent without yielding much precipitation

    More generally, precipitation over land is derived from two sources:

    Recycled precipitation is defined as water that evaporates from land within a specified control volume and falls again as precipitation within the same control volume. We'll consider the remainder of the precipitation falling inside the control volume to be of advective origin (regardless of whether it was recently evaporated from an ocean or from the land surface outside the control volume).

    Why is knowledge of vapor transport and recycling important?

    For recent estimates of continental-scale precipitation recycling, see the papers by Brubaker et al (1993) including figures 1-6; and the review paper by Eltahir and Bras (1996).


    References

    Brubaker, K. L., D. Entekhabi and P. S. Eagleson, 1993. Estimation of Continental Precipitation Recycling, J. Clim., 1077-1089.

    Dingman, S. L., 1994. Physical Hydrology, MacMillan Publishing Company, New York.

    Eltahir, E. A. B. and R. L. Bras, 1996. Precipitation Recycling, Rev. Geophys., 367-368.


    Back to top of page

     

     

     

    6: Precipitation II

    Outline

    1. Measurement at a point
    2. Checking the consistency of point measurements

    Measurement at a point

    As an input to the land phase of the hydrological cycle, it is critical that we quantify the space-time variability of the precipitation falling to the earth's surface. Whether these problems are of local, small watershed, regional, continental, or global scale, we need to understand the amount of rainfall falling to the ground, and how that amount varies in space and time.

    This problem has 2 components:

    Types of precipitation gages

    The principal behind measuring (gaging) precipitation is conceptually simple:

    Recording rain gages

    weighing type --

    Water collected is funneled to a vessel on a scale and the weight is recorded.

    float and siphon type

    A float rises as water level rises in the gage; height of the float is recorded.

    tipping-bucket type

    Water in the gage is funnelled to a pair of vessels with a known small capacity. When one vessel fills, it tips and drains, (bringing the other vessel into position) and the time of the tipping is recorded.

    Non-recording rain gages

    Measurement of snow

    Factors affecting measurement accuracy

    To what degree is the gauge catch an accurate representation of the actual amount of precipitation that fell?

    A number of issues need to be addressed:

    How big should the gage be?
    What should its shape be?
    Should it be placed level or parallel to the ground?
    Should it be shielded to minimize wind effects?
    Should it be fenced off to minimize disturbance by wildlife?
    How can evaporation be minimized?

    There are no universal standards, and in fact, standards vary between countries. Dingman summarizes a number of studies which provide guidance on these issues.

    Orifice size and orientation

    greater than 30 mm
    orifice should be level

    Orifice height and wind shielding

    Distance to obstructions

    Few studies have been done, but some general guidelines exist:

    Splash and evaporation

    If the surface of the water in the gage is too close to the gage opening, the impact of falling drops can cause collected water to spash out. Prevent by

    Limiting evaporation

    Instrument errors

    Systematic (mechanical) errors are common with recording rain gages

    Observer errors

    Very common but difficult to quantify

    Can identify by periodically comparing record from a number of gages in the same region and noting outliers

    was there a thunderstorm to explain it?
    were there high streamflows locally to support it?

    If an error, can correct objectively (statisically or otherwise)

    Errors due to differences in observation time

    e.g. in non-recording gages, observations are often daily

    Errors due to occult precipitation

    Occult precipitation: precipitation induced by cloud contact with trees or other vegetation.

    Fog drip is the liquid form and it occurs when clouds move through forests

    Rime is formed when supercooled clouds encounter exposed objects, like trees, and provide nucleation sites for ice-crystal formation and the buidup of ice, which ultimately falls to the ground as a solid or liquid

    Errors due to low-intensity rains

    Rain may enter gage, but not enough to measure accurately

     

    2. Checking the consistency of point measurements

    Changes in type, location, or environment of a gage are common

    • Buildings may be constructed nearby
    • location may be changed
    • urbanization can have important effects

    These changes can significantly affect the gage catch

    The hydrologist must determine whether the precipitation record is affected by such artificial changes and correct for them so they do not influence analyses. The most common technique for detecting and correcting for inconsistent precipitation data is by a double mass curve analysis. A double mass curve is a plot, on arithmetic graph paper, of the successive cumulative annual precipitation collected at a gage where measurement conditions may have changed, versus the successive cumulative annual precipitation for the same period of years collected at several gages in the same region (or the average of them).

    A change in the proportionality between the measurements at a suspect station and those in the region is reflected as a change in slope. Dingman, Figure 4-20 displays double mass curve examples. The second graph in this figure indicates some inconsistency.

    How are corrections made? General guidelines: the break in slope is only significant if it persists for 5 or more years, and then only if it is clearly associated with a change in measurement and it is determined to be statistically significant. Under these conditions, the annual values of the earlier part of the curve should be adjusted to be consistent with the later portions of the curve. Simply multiply the data for the period before the slope change by a factor K, where

    K = slope for the period after the change/slope for the period before the change

    Slope breaks in double mass curves can occur due to climatic shifts, so that adjustments should only be made if there is reason to believe that the break is due to a change in measurement conditions


    References

    Dingman, S. L., 1994. Physical Hydrology, MacMillan Publishing Company, New York.


    Back to top of page

     

     

     

    7: Precipitation III

    Outline

    1. Areal Estimation
    2. Precipitation Gage Networks
    3. Radar Measurement of Radar

    Areal Estimation

    Estimates of areally averaged precipitation are required at scales from the hillslope up to global.

    If the true areal average is P, where

    P = 1/A p(x,y) dx dy

    and A is the area of the region and p is the precipitation falling at the coordinates x and y, then what we're after is , the estimated value of P.

    Approaches to the estimation of include

    Direct weighted averages

    Compute directly as a weighted average of measured values,

    where

    G
    = S vgpg,
    g=1

    and

    G
    S vg = 1
    g=1

    pg are the precipitation values measured at each gage

    Estimating the weights, vg:

    Arithmetic average: Weights are equal to 1/G so that

    G
    = 1/G S pg
    g=1

    Thiessen polygon

    G
    = 1/A S agpg,
    g=1

    Construction of polygons

    See Figure 2.7, Fetter, 1994

    Surface-Fitting methods

    Identify a surface which represents the spatial variability in precipitation which can be depicted on a map

    General features:

    i = 1/2 (pi- + pi+)

    where i is the average precipitation value in the ith subregion

    I
    = 1/A S aii
    i=1

    and ai is the area between the two contours

    Contours can be drawn by hand (the eyeball isohyetal method) or automatically using specialized software. The software estimates the surface of precipitation given the gage values, and draws the contours on that surface.

    These algorithms follow the following procedure

     

    G
    j = S wjgpg
    g=1

    where j is the estimated precipitation at the jth point, pg is the gage precipitation at station g, and wjg is the weight assigned to a station g for point j.

  • • Most contouring routines use only the pg within a radius around point j to estimate pj

    • The difference between the various surface fitting methods is the manner by which the wjg are computed.

    • Once the j 's are computed, another algorithm is used to compute the contours

  • Important points regarding surface fitting:

    • Contouring algorithms involve arbitrary methods for choosing weights
    • Different algorithms will produce different patterns
    • Point values may contain error, as discussed previously--don't use method that is more elaborate than warranted by quality of the data
    • Objective methods based on mathematical or statistical criteria may have little relation to actual spatial variability, i.e. it may be unwise to draw contours based only on the pg values without consideration of physiography

    Hypsometric method : good in regions where orographic effect is important

    See Dingman (1994) figure 4-27

    1. Select an interval DZ and divide elevation range into H increments of DZ
    2. Determine the fraction of total area, ah within each increment of elevation
    3. Use the orographic equation to get (z)h where (z)h is taken at the midpoint of each elevation increment
    4. Compute as
    H
    = S (z)h ah
    h = 1

    2. Precipitation gage networks

    Recommended gage density: see Dingman Table 4-5

    In the US

    278 primary stations

    staffed 24 hours by paid technicians
    located mainly at airports

    8000 cooperative stations

    operated by unpaid volunteers

    Precipitation at these stations is recorded on at least a daily basis. Of these, hourly totals are recorded at 241 first order stations and 2600 cooperative stations. Many of these stations have been in operation for more than 100 years.

    3. Radar Measurement of Radar

    By 1996, a network of more than 120 high-quality radars, the Next Generation Weather Radar system, (NEXRAD) will be deployed in the U.S. This network will provide near complete coverage in the continental US.

    Most important advantage of using radar for precipitation measurement

    The NEXRAD systems are termed WSR-88D for weather surveillance radar, D is for Doppler. The radar unit transmits electromagnetic radiation (waves) of 10cm wavelength, and receives the returned signal, which is measured as

    backscattered power, P,

    P= C L Z / r2

    which is related to characteristics of the radar and characteristics of the precipitation target.

    P is related to

    Measure P, know C, L, r, solve for Z. Determine rainfall rate from the a power law model

    R = aZb

    where the parameters a and b are determined using paired samples of reflectivity (from samples with known drop size distribution) and rainfall rate.


    References

    Dingman, S. L., 1994. Physical Hydrology, MacMillan Publishing Company, New York.
     
    Fetter, C. W., 1994. Applied Hydrogeology, Macmillan College Publishing Co., New York.
     
    Maidment, D. R., 1993. Handbook of Hydrology, McGraw-Hill Publishing Co., New York.

     


    Back to top of page

     

     

    8: Snow and Snowmelt I

    Outline

    1. Introduction
    2. Formation and Interception
    3. Material Characteristics

    1. Introduction

    The objective of this chapter is to become familiar with the physical characteristics of seasonal snowcover and the processes affecting them. Volumetrically, seasonal snowcover forms only a small fraction of the world's fresh water. Hydrologically, though , its importance is immense. Furthermore, it seasonally affects large portions of the landscape in North America.

    In high and midlatitudes, and where precipitation is slight, melt of seasonal snow cover is the most significant hydrologic event of the year. In these regions, runoff from shallow snowcover can provide more than 80% of the annual surface runoff, as well as augment soil water reserves and recharge groundwater supplies

    Even at lower latitudes, particularly in alpine regions, snowmelt is a primary source of water. For example, in California, as much as 85% to 90% of annual streamflow comes from spring snowmelt.

    2. Formation and Interception.

    Physics of formation.

    We've previously discussed ice crystal formation: Continued growth of ice crystals leads to formation of snow crystals. A snow crystal is a large particle, having a very complex shape, and of such a size that it is visible to the naked eye. Their size and shape may change as they fall owing to freezing and accretion of water droplets with which they collide. This results in rimed crystals and in the extreme, snow pellets.

    A snowflake is an aggregation of snow crystals that may also grow in size during its fall owing to the adhesion of colliding snow crystals. Whether a snowflake formed in the atmosphere arrives at the earth's surface as snow or rain depends on the height of the zero degree line above the earth's surface.

    Interception

    Interception of snowfall by vegetation canopies plays a major role in the snow hydrology of forests, including, the energy balance at the ground surface, the amount of snowfall lost to evaporation and sublimation, and the distribution of snow within a forest and forest openings

    Interception occurs in the following manner. Early in a storm snowflakes fall through branches and needles until small bridged form at narrow openings. Bridges increase the collection area, and snow is retained on the bridges by cohesion. At some point a tree reaches its maximum holding capacity, when snow retention from subsequent snowfall is roughly balanced by the loss of intercepted snow falling to ground.

    Morphological vegetation factors affecting interception include branch strength and flexure,

    needle configuration and orientation, mass and surface area of vegetation, tree age and density. Interception by coniferous forests is much greater than by deciduous, as deciduous lose their leaves. Meteorological factors like air temperature and wind speed are generally of less importance than morphological characteristics

    Wind effects interception by disrupting bridges, vibrating vegetation and thus reducing interception, erosion of intercepted snow and by retention of snow particles in suspension.

    The amount of snow stored in the canopy continuously changes in response to wind erosion, evaporation and sublimation and melting.

    3. Material characteristics of snow

    The physical properties of a snowpack most important to hydrologists are depth, density, and snow water equivalent, the equivalent depth of water of a snow cover.

    Snow is a granular porous medium consisting of ice and pore spaces. When snow is cold (i.e. when its temp is below 0°C) the pore spaces contain only air. At the melting point, the pore spaces contain liquid water as well as air (a 3-phase system)

    For a snowpack of height hs and area A

    Snow volume

    Vs = Vi + Vw+ Va = hsA

    Porosity is the ratio of pore volume to total volume

    F = (Vw+ Va)/Vs

    The liquid water content Q

    Q = Vw/Vs

    Snow density is the mass per unit volume of snow

    rs = (Mi + Mw)/Vs

    =(1-F )ri+Qrw

    The water equivalent, hm, of the snow pack is the the depth of water that would result from completely melting the snow in place

    hm = Vm/A

    and Vm is the volume of meltwater, and can be shown to be equal to

    hm = rs/rw hs

    snow water equivalent equals relative density times depth

    More on the density. Average density for new snowfall often assumed to be 100 kg/m3

    However, in actuality, it varies with environment, though it is difficult to measure. In general, density depends upon the configuration of the snowflakes (i.e. the amount of air contained within the lattice of the snow crystals). Densities in the range of 50 to 120 kg/m3 are common. Lower values are associated with dry, cold conditions ane higher values found in wet snows at warm temperatures. Density is also a function of wind speed at the deposition surface. Higher winds tend to break up snowflakes so that they pack into denser layers. Maidment (1993), Fig 7.2.2 shows that thedensity of new fallen snow decreases exponentially as air temp decreases below freezing,


    References

    Dingman, S. L., 1994. Physical Hydrology, MacMillan Publishing Company, New York.

    Maidment, D. R., 1993. Handbook of Hydrology, McGraw-Hill Publishing Co., New York.


    Back to top of page

     

     

     

    9: Infiltration and Redistribution I

    Outline

    1. Introduction to infiltration and terminology
    2. Material properties of soils

    1. Introduction to infiltration and terminology

    In this chapter we'll discuss the movement of water through the unsaturated zone. Some of the processes we'll discuss are shown in Dingman (1994) Figure 6.1. Infiltration is the movement of water from the soil surface into the soil. Redistribution is the subsequent movement of infiltrated water in the unsaturated zone, vertically, laterally, etc. Percolation is the general term for downward flow in the unsaturated zone.

    Redistribtution can also occur in an upward direction:

    • exfiltration (evaporation)

    • capillary rise (upward movement from the water table to the unsat zone under capillary, or surface tension forces)

    • Plant uptake by roots

    Lateral redistribution is called interflow

    Importance of infiltration and redistribution of soil water: globally, 3/4 of the precipitation falling on land infiltrates so it is of great importance to the global cycle, e.g.

    • via recharging soil moisture and ground water reservoirs

    • soil water stored near the land surface is available to interact with the atmosphere and climate system.

    • Provides all the water for natural and agriculture

    We also need to understand infiltration and redistribution for water resource management purposes, including irrigation strategies, understanding the geochemistry of soils, forecasting runoff and flooding.

    Let's begin our discussion of infiltration and redistribution by reviewing some of the material and hydraulic properties of soils

    2. Material properties of soils

    Distribution of pore and particle sizes

    For the purposes of this class we'll consisder soil to be a matrix of solid grains (mineral or organic) between which are interconected pore spaces. Pores contain water and air in varying proportions

    i.e. soil is a 3-phase system of solid mineral grains, water and air.

    Pore size, or the size of pores through which water flows, is roughly equal to the grain size. The distribution of pore sizes is determined largely by the grain size.

    Most soils have a mixture of grain sizes. Grain size distribution is often portrayed as a cumulative frequency plot of grain diameter vs weight fraction of grains with smaller diameter (See Dingman (1994) Figure 6-2).

    Particle size distribution is further characterized by soil texture.

    Texture is determined by the weight fraction of clay, sand, and silt

    as shown in Dingman (1994) Figure 6-3. This is the USDA classification scheme for soil texture. Note that texture is determined by the fraction of silt, sand,clay after sand and larger particles have been removed. The weight fractions of soils of various diamters are meaured by sieve analysis.

    Particle density is the mass of the mineral particles divided by the volume of the mineral particles

    rm = Mm/Vm

    Usually rm is not measured but estimated based on the composition of the soil. A value of 2.65 g/cm3 (the density of mineral quartz) is assumed for most soils.

    Bulk density is the dry density of the soil

    rb = Mm/Vs= Mm/(Va+Vw+Vm)

    Bulk density increases with depth owing to compaction by weight of overlying soil.

    Porosity, f, is the ratio of the volume of voids to the total volume

    f = (Va+Vw)/Vs

    Porosity decreases with depth owing to compaction, and owing to the presence of macropores near the surface.

    Dingman (1994) Figure 6-4 shows a range of porosity values for different soil textures, and Figure 6-5 provides the basis ofr explanation in terms of the packing arrangement of clay particles vs. sand particles for range of porosities and explanation

    Volumetric water content

    q = Vw/Vs

    Measurement in the lab consists of weighing a sample of known volume, oven drying it at 105 deg C, and re-weighing, then calculating

    q = (Mswet - Msdry)/rwVs

    where Ms stands for mass of soil

    Degree of saturation (relative saturation) is the volume of water normalized by the porosity

    S = Vw / (Va + Vw) = q / f


    References

    Dingman, S. L., 1994. Physical Hydrology, MacMillan Publishing Company, New York.
     

    Back to top of page

     

     

    10: Infiltration and Redistribution II

    Outline

    1. Other soil moisture measurement techniques
    2. Hydraulic properties of soils
    3. Pressure-Water Content relations
    4. Water Conditions in Natural Soils

    1. Other soil moisture measurement techniques

    2. Hydraulic Properties of Soils

    Flows in unsaturated porous media are described by Darcy's Law

    Vx = -Kh d(z+ p/gw)/dx

    where

    so that

    Vx = -Kh dh/dx

    In unsaturated flows, both the pressure head and the hydraulic conductivities are functions of the soil water content, Q, so that

    Vx = -Kh(q) d(z + y(q))/dx

    Let's look at K-q and y-q relations in the unsaturated zone in a little more detail. The simplest hydrologic configuration of saturated and unsaturated conditions is that of an unsaturated zone overlying a capillary fringe and a saturated zone. The water table is defined as the surface on which the fluid pressures p in the pores of a porous medium is exactly atmospheric. If p is measured in gage pressure, then on the water table, p = 0 which implies that y = 0. From your groundwater classes, in the saturated zone we know that both y and p are > 0. In the unsaturated zone, both y and p are < 0 which reflects the fact that water in the unsaturated zone is held in the soil pores under surface tension forces

    A microscopic inspection would reveal a concave meniscus extending from grain to grain across each pore channel. The radius of curvature on each mensicus reflects the surface tension on that individual, microscopic, air-water interface. In reference to the physical mechanism of water retention, soil physicists often call the negative pressure the tension head, or head

    Again, from your groundwater class, you've probably discussed that piezometers are used to measure head in the saturated zone. However,above the water table, where y < 0 , we use tensiometers to measure y in the unsaturated zone.

    A tensiometer consists of a porous cup attached to an airtight, water-filled tube. The porous cup is inserted into the soil at the desired depth, and allowed to reach a hydraulic equilibrium. The equilibration involves the passage if water through the porous cup from the tube into the soil. The vacuum created at the top of the airtight tube is a measure of the pressure head in the soil. It is usually measured by a vacuum gage attached to the tube above the ground surface. To obtain total head, h, the negative y is added to the elevation head z at the point of measurement.

    3. Pressure-Water Content relations

    Pressure head

    The relationship between pressure head, y, and moisture content, q,

    for a given soil is called the moisture characteristic curve. The relationship is highly nonlinear, and typically of the form shown in Dingman, Figure 6-7. Note that pressure head is zero when water content equals the porosity.

    Water content changes little as tension increases up to a point of inflection. This rather distinct point represents the tension at which significant volumes of air begin to appear in soil pores, and is called the air entry tension, or air entry suction head, yae. The absolute value of the air-entry suction head equals the height of the capillary fringe, or the height of the tension saturated zone. As tension increases to very high levels, the curve becomes nearly vertical reflecting a residual water content that is very tightly held in soil pores by capillary forces.

    In real soils, the value of tension at a given water content is not unique and depends on the soil's wetting a drying history. Hysteresis can have a significant influence on soil moisture movement but it is difficult to model mathematically, and is often left out of hydrological models. See Figure 6-9 from Dingman (1994)

    Hydraulic Conductivity

    Hydraulic conductivity is the rate at which water moves through a porous medium under a unit hydraulic gradient. Under saturated conditions

    K = krg/m

    Under unsaturated conditions K is also a function of the degree of saturation of the soil -- the wetter the soil, the more interconnected flow paths. See figure 6-7 of Dingman (1994).

    The relationships between y-q and K-q can be expressed parametrically

    |y(S)| = |ys|S-b

     

    Kh(S) =KhsatSc

     

    where

    Values of these parameters are functions of soil texture and are tabulated as averages over many soils in the literature.

    Summary of saturated, unsaturated and tension saturated zones

    Saturated

    Unsaturated

    Capillary fringe

    4. Water Conditions in Natural Soils

    Soil water status

    If a soil is saturated, then allowed to drain without being subject to evaporation, plant uptake and capillary rise, its water content will decrease exponentially. Within a few days, the soil will reach a value beyond which a further decrease in moisture content (drainage) occurs very slowly. This value is known as the field capacity qfc. It represents the water that can be held against gravity. The pressure head at field capacity is close to - 340 cm for most soils.

    After a soil has reached field capacity, water can only be removed by direct evaporation or by plant uptake. However, plants can't exert suctions stronger than -15,000 cm. When water content is reduced to this point on the moisture characteristic curve, transpiration ceases and plants wilt. This point is known as the permanent wilting point qpwp. It ranges from 0.05 for sands to 0.25 for clays.

    The difference between the field capacity and pwp is the water available for plant use (available water content, qa)

    These concepts are shown schematically in Dingman (1994) Figure 6-12.


    References

    Dingman, S. L., 1994. Physical Hydrology, MacMillan Publishing Company, New York.
    Freeze, R. A. and J. A. Cherry, 1979. Groundwater, Prentice-Hall, New Jersey.

    Back to top of page

     

     

    11: Infiltration and Redistribution III

    Outline

    1. Pressure-Water Content Relations
    2. Water Conditions in Natural Soils

     

    1. Pressure-Water Content Relations

    Pressure head

    The relationship between pressure head, y, and moisture content, q,

    for a given soil is called the moisture characteristic curve. The relationship is highly nonlinear, and typically of the form shown in Dingman, Figure 6-7. Note that pressure head is zero when water content equals the porosity.

    Water content changes little as tension increases up to a point of inflection. This rather distinct point represents the tension at which significant volumes of air begin to appear in soil pores, and is called the air entry tension, or air entry suction head, yae. The absolute value of the air-entry suction head equals the height of the capillary fringe, or the height of the tension saturated zone. As tension increases to very high levels, the curve becomes nearly vertical reflecting a residual water content that is very tightly held in soil pores by capillary forces.

    In real soils, the value of tension at a given water content is not unique and depends on the soil's wetting a drying history. Hysteresis can have a significant influence on soil moisture movement but it is difficult to model mathematically, and is often left out of hydrological models. See Figure 6-9 from Dingman (1994)

    Hydraulic Conductivity

    Hydraulic conductivity is the rate at which water moves through a porous medium under a unit hydrualic gradient. Under saturated conditions

    K = krg/΅

    Under unsaturated conditions K is also a function of the degree of saturation of the soil -- the wetter the soil, the more interconnected flow paths. See figure 6-7 of Dingman (1994). Freeze and Cherry (1979) Figure 2-14 shows effects of different soil textures

    The relationships between y-q and K-q can be expressed parametrically

    |y(S)| = |ys|S-b

    Kh(S) =KhsatSc

    where

    S is relative saturation
    Khsat is the saturated hydraulic conductivity
    ys is essentially the air entry suction head
    b is called the pore size distribution index, and is a fitting parameter
    c = 2b + 3

    Values of these parameters are functions of soil texture and are tabulated as averages over many soils in the literature.

    Summary of saturated, unsaturated and tension saturated zones

    Saturated

    Unsaturated

    2. Water Conditions in Natural Soils

    Soil water status

    If a soil is saturated, then allowed to drain without being subject to evaporation, plant uptake and capillary rise, its water content will decrease exponentially. Within a few days, the soil will reach a value beyond which a further decrease in moisture content (drainage) occurs very slowly. This value is known as the field capacity qfc. It represents the water that can be held against gravity. The pressure head at field capacity is close to - 340 cm for most soils.

    After a soil has reached field capacity, water can only be removed by direct evaporation or by plant uptake. However, plants can't exert suctions stronger than -15,000 cm. When water content is reduced to this point on the moisture characteristic curve, transpiration ceases and plants wilt. This point is known as the permanent wilting point qpwp . It ranges from 0.05 for sands to 0.25 for clays.

    The difference between the field capacity and pwp is the water available for plant use (available water content, qa)

    These concepts are shown schematically in Dingman (1994) Figure 6-12, and in a vertical soil column in Figure 6-13.


    References

    Dingman, S. L., 1994. Physical Hydrology, MacMillan Publishing Company, New York.
    Freeze, R. A. and J. A. Cherry, 1979. Groundwater, Prentice-Hall, New Jersey.

    Back to top of page

     

     

    12: Infiltration and Redistribution IV

    Outline

    1. Qualitative description at a point
    2. Measurement

     

    1. Qualitative description at a point

    Infiltration is the process by which water arriving at the soil surface enters the soil. Let's review some terminology found in the text and look at things qualitatively before moving on to quantitative description.

    Dingman (1994) defines the beginning of a water-input event begins a t time t = 0 and ending at t = tw. The infiltration rate f(t), or(actual infiltration rate) is the rate at which water enters the soil from the surface. The water input rate w(t) is the rate at which water arrives at the surface due to rain or snowmelt. The infiltration capacity fc(t) is the maximum rate at which infiltration can occur. The depth of ponding Y(t), the depth of water standing on the surface.

    Then 3 conditions are recognized which control the actual infiltration rate:

    1.) No ponding. The water input rate is less than the infiltration capacity, so that all water infiltrates and none ponds on the surface. The actual infiltration rate is equal to the water input rate.

    Y(t) = 0         f(t) = w(t)

    2.) Saturation fom above. Ponding has occurred because the water input rate is greater thatn the infiltration capacity. The actual infiltration rate proceeds at the infiltration capacity.

    Y(t) > 0         f(t) = fc(t)

    3.) Saturation from below. The water table has risen to or above the surface and the entire soil is saturated. Soil pores are full and there is no room for infiltrated water.

    Y(t) 0            f(t) = fc(t) = 0

    Based the above we can state that the actual infiltration rate is equal to the minimum of the water input rate or the infiltration capacity.

    f(t) = min [ w(t), fc(t)]

    2. Measurement

    Ring infiltrometer.

    The ring infiltrometer is a cylindrical ring,which is driven into the ground, flooded, and the level of water in the ring ismonitored to determine the infiltration rate.

    The ring is first flooded, then the infiltration rate is obtained by measuring the rate at which the level of ponded water decreases, or by measuring the rate at which water must be added to maintain constant head.

    In general, during a water input event, infiltration rates start out high and decrease exponentially with time into a storm, ultimately approaching a constant rate which is equal to Ksat.

    Because infiltrating water encounters both capillary and gravitational forces, the water beneath the ring moves laterally as well as vertically, so that the measured infiltration rate often exceeds the rate if the entire surface were ponded. A double-ring infiltrometer can be used to set up a buffer zone and obtain a more realistic estimate of infiltration. Alternatively, a capillary correction term can be applied.

    Sprinkler plot

    A sprinkler is applied to a plot of land surface at a known rate. Runoff is measured. The difference between the water applied and the runoff is infiltrated water.

    Observation of soil water

    Install a nest of tensiometers which monitor pressure head at different levels. Pressures can be converted to moisture content using the moisture characteristic equations (remembering problems with hysteresis) and the depth and rate of infiltration can be determined from analysis of the profiles.


    References

    Dingman, S. L., 1994. Physical Hydrology, MacMillan Publishing Company, New York.
    Freeze, R. A. and J. A. Cherry, 1979. Groundwater, Prentice-Hall, New Jersey.

    Back to top of page

     

     

    13: Infiltration and Redistribution V

    Outline

    1. Basic characteristics of infiltration
    2. Factors affecting infiltration rates
    3. Spatial variability of infiltration

    1. Basic characterisitics of infiltration

    In general, the infiltration capacity decreases with time into an event, and ultimately approaches the saturated hydraulic conductivity. At early times into an event, capillary forces dominate infiltration. However, these forces decrease with time as the soil wets, until eventually, gravitational forces dominate.

    Figure xx shows another characteristic of an infiltration event, the formation of a wetting front near the soil surface...

    2. Factors affecting infiltration rate

    f(t) is determined by affects of gravity and pressure forces arriving at the surface, which in turn are determined by the

    1. Rate of arrival of water at the surface, or depth of ponding
    2. Khsat at the surface
    3. Initial soil wetness
    4. Inclination and surface roughness
    5. Chemical characteristics of the soil surface
    6. Physical and chemical properties of water

    The first of these has already been discussed.

    Khsat at the surface

    organic surface layers: Litter, humus and other organic material where has a loose, open structure with a high hydraulic conductivity regardless of texture of the underlying mineral soil. Root growth, insects and burrowing mammals contribute to the surface porosity.

    frost: High water content at the surface can render soil impermeable upon freezing. Frost associated with lower water contents can increase conductivity by producing crack networks, e.g. polygonal cracks.

    swelling and drying: Clay minerals can swell and shrink during rain events. Swelling can reduce permeabiliy, porosity, and limit infiltration during dry periods. Crack networks can develop that can result in high infiltration rates.

    inwashing of fine sediments: Fine sediments may wash into soil pores reducing effective pore size and permeability.

    human impact. Depends on the nature of the activity. Plowing may increase but grazing, other heavy equipment may compact soil.

    Initial soil wetness

    As mentioned previously, saturation from below eliminates infiltration. This occurs under rising water table conditions, or also from decreased hydraulic conductivity at depth which encourages a build-up of water from below.

    In unsaturated soils, high water contents tend to increase Khsat and thus infiltration rates. However, higher water contents reduce the effects of surface tension/capillary forces.

    The net effect of antecedent water thus depends on the water input rate and duration, the distribution of soil hydraulic conductivity with depth, depth of the local water table and initial water contnent

    Surface slope and roughness

    Steeper, smoother slopes limit depth of ponding and thus infiltration rates.

    Chemical characteristics of the soil surface Some vegetation produces waxy organic substances which renders soils hydrophobic, i.e. water beads up, and thus infiltration is limited. This is a minor effect under undisturbed conditions, but more significant when, e.g. organic layers are burned off in forest fires and the waxy substances condenses on bare soil.

    Physical/chemical properties of water For example, viscosity is very temperature sensitive: higher viscosity occur at lower temperatures, and thus infiltration rates are lower.

    3. Spatial variability of infiltration

    Infilitration at a point is controlled by: soil type and properties (porosity, Khsat, air entry suction head, pore size distribution), initial moisture content and rainfall rates. As we move from the point to the hillslope to the catchment and greater scales, we encounter spatial variability in soils, moisture content and rainfall patterns, and consequently spatial variability in infiltration rates. The degree of the variability is simply a function of the amount of variability of the individual patterns listed above and how they interact. The book describes various attempts at modeling but in fact, at the catchment scale, the most direct approach is spatially-distributed modeling. At larger scales, statistical methods may be more appropriate.


    References

    Dingman, S. L., 1994. Physical Hydrology, MacMillan Publishing Company, New York.
    Freeze, R. A. and J. A. Cherry, 1979. Groundwater, Prentice-Hall, New Jersey.

    Back to top of page