You are on page 1of 9

368 HYDRAULIC GRADIENT

Bibliography studying or predicting site-specific water flow and solute


Introduction to Environmental Soil Physics (First Edition) 2003 transport processes in the subsurface. This includes using
Elsevier Inc. Daniel Hillel (ed.) http://www.sciencedirect.com/ models as tools for designing, testing, or implementing
science/book/9780123486554 soil, water, and crop management practices that optimize
water use efficiency and minimize soil and water pollution
by agricultural and other contaminants. Models are
equally needed for designing or remediating industrial
HYDRAULIC GRADIENT waste disposal sites and landfills, or assessing the for
long-term stewardship of nuclear waste repositories.
The slope of the hydraulic grade line which indicates the Predictive models for flow in variably saturated soils are
change in pressure head per unit of distance. generally based on the Richards equation, which combines
the Darcy–Buckingham equation for the fluid flux with
a mass conservation equation to give (Richards, 1931):
HYDRAULIC HEAD  
@yðhÞ @ @h
¼ KðhÞ  KðhÞ (1)
@t @z @z
The sum of the pressure head (hydrostatic pressure rela-
tive to atmospheric pressure) and the gravitational head in which y is the volumetric water content (L3 L3), h is
(elevation relative to a reference level). The gradient of the pressure head (L), t is time (T), z is soil depth (positive
the hydraulic head is the driving force for water flow in down), and K is the hydraulic conductivity (L T1).
porous media. Equation 1 holds for one-dimensional vertical flow; similar
equations can be formulated for multidimensional flow
Bibliography problems. The Richards equation contains two constitutive
Introduction to Environmental Soil Physics (First Edition) 2003 relationships, the soil water retention curve, y(h), and the
Elsevier Inc. Daniel Hillel (ed.) http://www.sciencedirect.com/ unsaturated soil hydraulic conductivity function, K(h).
science/book/9780123486554 These hydraulic functions are both strongly nonlinear func-
tions of h. They are discussed in detail below.

HYDRAULIC PROPERTIES OF UNSATURATED SOILS Water retention function


The soil water retention curve, y(h), describes the relation-
Martinus Th. van Genuchten1, Yakov A. Pachepsky2 ship between the water content, y, and the energy status of
1
Department of Mechanical Engineering, COPPE/LTTC, water at a given location in the soil. Many other names
Federal University of Rio de Janeiro, UFRJ, Rio de may be found in the literature, including soil moisture
Janeiro, Brazil characteristic curve, the capillary pressure–saturation
2
Environmental Microbial and Food Safety Laboratory, relationship, and the pF curve. The retention curve histor-
Animal and Natural Resource Institute, Beltsville ically was often given in terms of pF, which is defined as
Agricultural Research Center, USDA-ARS, Beltsville, the negative logarithm (base 10) of the absolute value of
MD, USA the pressure head measured in centimeters. In the unsatu-
rated zone, water is subject to both capillary forces in soil
pores and adsorption onto solid phase surfaces. This leads
Definition to negative values of the pressure head (or matric head)
Hydraulic Properties of Unsaturated Soils. Properties relative to free water, or a positive suction or tension.
reflecting the ability of a soil to retain or transmit water As opposed to unsaturated soils, the pressure head h is
and its dissolved constituents. positive in a saturated system. More formally, the pressure
head is defined as the difference between the pressures of
Introduction the air phase and the liquid phase. Capillary forces are the
Many agrophysical applications require knowledge of the result of a complex set of interactions between the solid
hydraulic properties of unsaturated soils. These properties and liquid phases involving the surface tension of the liq-
reflect the ability of a soil to retain or transmit water and uid phase, the contact angle between the solid and liquid
its dissolved constituents. For example, they affect the phases, and the diameter of pores.
partitioning of rainfall and irrigation water into infiltration Knowledge of y(h) is essential for the hydraulic charac-
and runoff at the soil surface, the rate and amount of redis- terization of a soil, since it relates an energy density
tribution of water in a soil profile, available water in the (potential) to a capacity (water content). Rather than using
soil root zone, and recharge to or capillary rise from the pressure head (energy per unit weight of water), many
the groundwater table, among many other processes in agrophysical applications use the pressure or matric
the unsaturated or vadose zone between the soil surface potential (energy per unit volume of water, usually mea-
and the groundwater table. The hydraulic properties are sured in Pascal, Pa), cm = rwgh, where rw is the density
also critical components of mathematical models for of water (ML3) and g the acceleration of gravity (L T2).
HYDRAULIC PROPERTIES OF UNSATURATED SOILS 369

Figure 1 shows typical soil water retention curves for initially very dry sample to produce the main wetting
relatively coarse-textured (e.g., sand and loamy sand), curve, which is generally displaced by a factor of
medium-textured (e.g., loam and sandy loam), and fine- 1.5–2.0 toward higher pressure heads closer to saturation.
textured (e.g., clay loam, silty loam, and clay) soils. The This phenomenon of having different wetting and drying
curves in Figure 1 may be interpreted as showing the equi- curves, including primary and secondary scanning curves
librium water content distribution above a relatively deep is referred to hysteresis. Hysteresis is caused by the fact
water table where the pressure head is zero and the soil that drainage is determined mostly by the smaller pore in
fully saturated. The plots in Figure 1 show that coarse- a certain pore sequence, and wetting by the larger pores
textured soils lose their water relatively quickly (at small (this effect is often referred to as the ink bottle effect).
negative pressure heads) and abruptly above the water Other factors contributing to hysteresis are the presence
table, while fine-textured soils lose their water much more of different liquid–solid contact angles for advancing
gradually. This reflects the particle or pore-size distribu- and receding water menisci, air entrapment during wet-
tion of the medium involved. While the majority of pores ting, and possible shrink–swell phenomena of some soils.
in coarse-textured soils have larger diameters and thus
drain at relatively small negative pressures, the majority Hydraulic conductivity function
of pores in fine-textured soils do not drain until very large
tensions (negative pressures) are applied. The hydraulic conductivity characterizes the ability of
As indicated by the plots in Figure 1, the water content a soil to transmit water. Its value depends on many factors
varies between some maximum value, the saturated water such as the pore-size distribution of the medium, and the
content, ys, and some small value, often referred to as tortuosity, shape, roughness, and degree of interconnec-
the residual (or irreducible) water content, yr . As a first tedness of the pores. The hydraulic conductivity decreases
approximation and on intuitive ground, the saturated considerably as soil becomes unsaturated since less pore
water content is equal to the porosity, and yr equal to zero. space is filled with water, the flow paths become increas-
In reality, however, the saturated water content, ys, of soils ingly tortuous, and drag forces between the fluid and the
is generally smaller than the porosity because of entrapped solid phases increase.
and dissolved air. The residual water content yr is likely to The unsaturated hydraulic conductivity function gives
be larger than zero, especially for fine-textured soils with the dependency of the hydraulic conductivity on the water
their large surface areas, because of the presence of content, K(y), or pressure head, K(h). Figure 2 presents
adsorbed water. Most often ys and especially yr are treated examples of typical K(y) and K(h) functions for relatively
as fitting parameters without much physical significance. coarse-, medium-, and fine-textured soils. Notice that the
Soil water retention curves such as shown in Figure 1 hydraulic conductivity at saturation is significantly larger
are not unique but depend on the history of wetting and for coarse-textured soils than fine-textured soils. This differ-
drying. Most often, the soil water retention curve is deter- ence is often several orders of magnitude. Also notice
mined by gradually desaturating an initially saturated soil that the hydraulic conductivity decreases very significantly
by applying increasingly higher suctions, thus producing as the soil becomes unsaturated. This decrease, when
a main drying curve. One could similarly slowly wet an expressed as a function of the pressure head (Figure 2;
right), is much more dramatic for the coarse-textured soils.
The decrease for coarse-textured soils is so large that at
100,000 a certain pressure head the hydraulic conductivity becomes
smaller than the conductivity of the fine-textured soil. The
10,000 water content where the conductivity asymptotically
Pressure head, |h| (cm)

becomes zero (Figure 2; left) is often used as an alternative


working definition for the residual water content, yr.
1,000

Soil water diffusivity


100
Another hydraulic function often used in theoretical and
management application of unsaturated flow theories is
10 the soil water diffusivity, D(y), (L2 T1), which is defined as
 
dh
1 DðyÞ ¼ KðyÞ : (2)
0.0 0.1 0.2 0.3 0.4 0.5 dy
Volumetric water content [–] This function appears when Equation 1 is transformed
into a water-content-based equation in which y is now
Hydraulic Properties of Unsaturated Soils, Figure 1 Typical the dependent variable:
soil water retention curves for relatively coarse- (solid line),  
medium- (dashed line), and fine-textured (dotted line) soils. The @y @ @h
curves were obtained using Equation 6a assuming hydraulic ¼ DðyÞ  KðyÞ : (3)
parameter values as listed in Table 1. @t @z @z
370 HYDRAULIC PROPERTIES OF UNSATURATED SOILS

4 4

2 2
log (K), [cm/day]

log (K), [cm/day]


0 0

−2 −2

−4 −4

−6 −6
0.0 0.1 0.2 0.3 0.4 0.5 0 1 2 3 4 5 6
Volumetric water content, [–] log(|h|, [cm])

Hydraulic Properties of Unsaturated Soils, Figure 2 Typical curves of the hydraulic conductivity K, as a function of the pressure
head (left) and water content (right) for coarse- (solid line), medium- (dashed line), and fine-textured (dotted line) soils. The curves were
obtained using Equation 6b assuming hydraulic parameter values as listed in Table 1.

Equation 3 is very attractive for approximate analytical (using a cut-and-paste concept of a cross-section of the
modeling of unsaturated flow processes, especially for medium containing different-sized pores), and then inte-
modeling horizontal (without the K(y) gravity term) and grating over all water-filled capillaries leads to the hydrau-
vertical infiltration (e.g., Philip, 1969; Parlange, 1980). lic conductivity of the complete set of capillaries, and
However, the water-content-based equation is less attractive consequently of the soil itself. The approach allows infor-
for more comprehensive numerical modeling of flow in mation of the soil water retention curve to be translated in
layered media, flow in media that are partially saturated predictive equations for the unsaturated hydraulic conduc-
and partially unsaturated, and for highly transient flow tivity. Many theories of this type, often referred to also as
problems. statistical pore-size distribution models, have been pro-
posed in the past, including Childs and Collis-George
(1950), Burdine (1953), Millington and Quirk (1961),
Analytical representations and Mualem (1976). A review of the different approaches
To enable their use in analytical or numerical models for is given by Mualem (1992). Examples of analytical y(h)
unsaturated flow, the soil hydraulic properties are often and K(h) equations resulting from this approach are the
expressed in terms of simplified analytical expressions. hydraulic functions of Brooks and Corey (1964), based
A large number of functions have been proposed over on the approach by Burdine (1953), and equations by
the years to describe the soil water retention curve, y(h), van Genuchten (1980) and Kosugi (1996), based on the
and the hydraulic conductivity function, K(h) or K(y). theory of Mualem (1976).
A comprehensive review of the performance of some of The classical equations of Brooks and Corey (1964) for
many these models is given by Leij et al. (1997). The func- y(h), K(h), and D(y) are given by
tions range from completely empirical equations to
models based on the simplified conceptual picture that   l
soils are made up of a bundle of equivalent capillary tubes y¼ yr þ ðys  yr Þhhe  h < he (4a)
that contain and transmit water. ys h  he
While extremely simplistic as indicated by Tuller and
Or (2001) among others, conceptual models that view KðhÞ ¼ Ks Se2=lþlþ1 (4b)
a soil as a bundle of capillaries of different radii are still
useful for explaining the shape of the water retention curve Ks
for different textures, as well as to provide a means for DðyÞ ¼ S 1=lþl (4c)
predicting the hydraulic conductivity function from soil aðys  yr Þ e
water retention information. These models typically
assume that pores at a given pressure head are either where, as before, yr is the residual water content (L3 L3),
completely filled with water, or empty, depending upon ys is the saturated water content (L3 L3), he is often
the applied suction. Flow in each water-filled capillary referred to as the air-entry value (L), l is a pore-size distri-
tube is subsequently calculated using Poiseuille’s law for bution index characterizing the width of the soil pore-size
flow in cylindrical pores. By adding the contribution of distribution, Ks is the saturated hydraulic conductivity
all capillaries that are still filled with water at a particular (LT1), l a pore-connectivity parameter assumed to be
pressure head, making some assumption about how small 2.0 in the original study of Brooks and Corey (1964),
and large capillaries connect to each other in sequence and Se = Se(h) is effective saturation given by
HYDRAULIC PROPERTIES OF UNSATURATED SOILS 371

yðhÞ  yr for l have been suggested in various studies. Based on an


Se ðhÞ ¼ (5) analysis of a large data set from the UNSODA database,
ys  yr Schaap and Leij (2000) recommended using l equal
to 1 as a more appropriate value for most soil textures.
For completeness we have given here also the expres- Equations 6a, 6b, and 6c assume the restrictive relation-
sion for the soil water diffusivity, D(y). Note that Equa- ship m = 1  1/n, which simplifies the predictive K(h)
tions 4b and 4c contain parameters that are also present expression compared to leaving m and n as independent
in Equation 4a, in particular yr and ys through Equation 5, parameters in Equation 6b. In particular, the convex and
as well as he and l. The value of l in Equation 4a reflects concave curvatures at the high and low pressure heads in
the steepness of the retention function and is relatively large Figure 1 have then a particular relationship with each
for soils with a relatively uniform pore-size distribution (gen- other. Other restrictions on Equation 6a have been used
erally coarse-textured soils such as those shown in Figures 1 also. For example, Haverkamp et al. (2005) used the
and 2), but small for soils having a wide range of pore sizes. restriction m = 1  2/n in connection with Equation 6a
One property of Equation 4a is the presence of a sharp and Burdine’s (1953) model to produce a different expres-
break in the retention curve at the air-entry value, he. This sion for K(h). The restrictions are not formally needed,
break (or discontinuity in the slope of the function) is since they limit the flexibility of Equation 6a in describing
often visible in retention data for coarse-textured soils, experimental data. However, the predicted K(h) function
but may not be realistic for fine-textured soils and soils obtained with the theories of Burdine or Mualem becomes
having a relatively broad pore- or particle-size distribu- then extremely complicated by containing incomplete
tion. A sharp break is similarly present in the hydraulic beta or hypergeometric functions, thus limiting the practi-
conductivity function when plotted as a function of the cality of the analytical functions.
pressure head, but not versus the water content. As an Rawls et al. (1982) provided average values of the
alternative, van Genuchten (1980) proposed a set of equa- parameters in the Brooks and Corey (1964) soil hydraulic
tions that exhibit a more smooth sigmoidal shape. The van parameters for 11 soil textural classes of the U.S. Depart-
Genuchten equations for y(h), K(h), and D(y) are given by: ment of Agriculture (USDA) textural triangle. Carsel and
ys  yr Parrish (1988) gave similar values for the van Genuchten
yðhÞ ¼ yr þ m ðm ¼ 1  1=n; n > 1Þ (1980) parameters for 12 USDA soil textural classes. In
ð1 þ jahjn Þ Table 1, we list typical van Genuchten hydraulic parameter
(6a) values for relative coarse-, medium-, and fine-textured soils.
h  m i2 The data in this table were actually used to calculate the
KðhÞ ¼ Ks Sel 1  1  Se1=m (6b) water retention and hydraulic conductivity functions,
shown in Figures 1 and 2, respectively, with Equations 6a,
ð1  mÞKs l1=m h m b. Average values such as those given in Table 1, or pro-
DðyÞ ¼ Se 1  Se1=m vided in more detail by Rawls et al. (1982) and Carsel and
amðys  yr Þ (6c) Parrish (1988), are often referred to as textural class aver-
 m i aged pedotransfer functions. Pedotransfer functions are
þ 1  Se1=m  2 relationships that use more easily measured of readily avail-
able soil data to estimate the unsaturated soil hydraulic
respectively, where a (L1), n (), and m (= 1–1/n) () parameters or properties (Bouma and van Lanen, 1987;
are shape parameters, and l is the pore-connectivity Leij et al., 2002; Pachepsky and Rawls, 2004).
parameter (). The parameter n in Equation 6 tends to We note that Equations 4 and 6 provide only two exam-
be large for soils with a relatively uniform pore-size distri- ples in which the hydraulic properties are described
bution and small for soils having a wide range of pore analytically. Many other combinations (Leij et al., 1997;
sizes. The pore-connectivity parameter l in Equation 6b Kosugi et al., 2002) are possible and have been used.
was estimated by Mualem (1976) to be about 0.5 as an For example, the combination of Equation 6a for y(h) with
average for many soils. However, many other values a simple expression like

Hydraulic Properties of Unsaturated Soils, Table 1 Typical values of the soil hydraulic parameters in the analytical functions
of van Genuchten (1980) for relatively coarse-, medium-, and fine-textured soils. The parameters were used to calculate the
hydraulic properties plotted in Figures 1 and 2 using Equations 6a and 6b, respectively

yr ys a n Ks
Soil texture (cm day1) (cm3 cm3) (cm3 cm3) (cm1) ()

Coarse 0.045 0.430 0.145 2.68 712.8


Medium 0.057 0.410 0.124 2.28 350.2
Fine 0.020 0.540 0.0010 1.2 45.0
372 HYDRAULIC PROPERTIES OF UNSATURATED SOILS

KðhÞ ¼ Ks Seb (7) capacity or flow attributes (e.g., water contents, pressure
heads, boundary fluxes), which are then used in combina-
which is essentially identical to Equation 4b, for K(h) is tion with a mathematical solution (generally numerical) to
also very realistic. Another attractive alternative equation obtain estimates of the hydraulic parameters such as
for K(h) is of the form (e.g., Vereecken et al., 1989) those that appear in Equations 4 and 6, or other functions.
Popular methods include one-step and multi-step outflow
Ks methods (Kool et al., 1987; van Dam et al., 1994), tension
KðhÞ ¼ (8)
1 þ jahjb infiltrometers methods (Šimůnek et al., 1998a), and evap-
oration methods (Šimůnek et al., 1998b), although many
Many alternative expressions have been used also for other laboratory and field methods also exist or can be
the soil water diffusivity function, D(y), mostly to facili- similarly employed (Hopmans et al., 2002). This also per-
tate simplified analytical analyses of unsaturated flow tains to different approaches for minimizing the objective
problems (e.g., Parlange, 1980). function, including quantification of parameter uncer-
tainty (Abbaspour et al., 2001; Vrugt and Robinson,
Experimental procedures 2007). Very attractive now also is the use of combined
A large number of experimental techniques can be used to hard (e.g., directly measured) and soft (e.g., indirectly esti-
estimate the hydraulic properties of unsaturated soils. mated) data, including hydrogeophysical measurements
A direct approach for the water retention function would and information derived from pedotransfer functions, to
be to measure a number of water content (y) and pressure extract the most out of available information (e.g.,
head (h) pairs, and then to fit a particular retention func- Kowalski et al., 2004; Segal et al., 2008).
tion to the data. Direct measurement techniques include
methods using a hanging water column, pressure cells,
pressure plate extractors, suction tables, soil freezing, Hydraulic properties of structured soils
and many other approaches. Comprehensive reviews of The Richards equation 1 typically predicts a uniform flow
various methods are given by Gee and Ward (1999) and process in the vadose zone. Unfortunately, the vadose
Dane and Hopmans (2002). Once the pairs of y and h zone can be extremely heterogeneous at a range of scales,
data are obtained, the data may be analyzed in terms of from the microscopic (e.g., pore scale) to the macroscopic
specific analytical water retention and conductivity func- (e.g., field or larger scale). Some of these heterogeneities
tions such as those discussed earlier. Several convenient can lead to a preferential (or bypass) flow process that
software packages are available for this purpose (van macroscopically is very difficult to capture with the stan-
Genuchten et al., 1991; Wraith and Or, 1998). Alterna- dard Richards equation. One obvious example of prefer-
tively, the data can be analyzed without assuming specific ential flow is the rapid movement of water and dissolved
analytical functions for y(h) and K(h) or K(y). This could solutes through soil macropores (e.g., between soil aggre-
be done using linear, cubic spline, or other interpolation gates, or created by earthworms or decayed root channels
techniques (Kastanek and Nielsen, 2001; Bitterlich et al., or rock fractures), with much of the water bypassing
2004). (short-circuiting) the soil or rock matrix. However, many
Similar direct measurement approaches involving pairs other causes of preferential flow exist, such as flow
of conductivity (or diffusivity) and pressure head (or water instabilities caused by soil textural changes or water repel-
content) data are also possible for the K(h) and D(y) lency (Hendrickx and Flury, 2001; Šimůnek et al., 2003;
functions, at least in principle (Dane and Topp, 2002), Ritsema and Dekker, 2005), and lateral funneling of water
including for the saturated hydraulic conductivity, Ks. along sloping soil layers (e.g., Kung, 1990).
The saturated hydraulic conductivity can be measured in While uniform flow in granular soils is traditionally
the laboratory using a variety of constant or falling head described with a single-porosity model such as the
methods, and in the field using single or double ring Richards equation given by Equation 1, flow in structured
infiltrometers, constant head permeameters, and various media can be described using a variety of dual-
auger-hole and piezometer methods (Dane and Topp, porosity, dual-permeability, multi-porosity, and/or multi-
2002). Unfortunately, because of the strongly nonlinear permeability models (Šimůnek and van Genuchten,
nature of the soil hydraulic properties, pairs for the K(h) 2008; Köhne et al., 2009). While single-porosity models
and D(y) data are not easily measured directly, especially assume that a single pore system exists that is fully acces-
at relatively low (negative) pressure heads, unless more sible to both water and solute, dual-porosity and dual-
specialized techniques are used such as centrifuge permeability models both assume that the porous medium
methods (Nimmo et al., 2002). Even then, the data are consists of two interacting pore regions, one associated
generally not distributed evenly over the entire water con- with the inter-aggregate, macropore, or fracture system,
tent range of interest. Consequently, unsaturated hydraulic and one comprising the micropores (or intra-aggregate
conductivity properties are most often estimated using pores) inside soil aggregates or the rock matrix. Whereas
inverse or parameter estimation procedures. dual-porosity models assume that water in the matrix is
Parameter estimation methods generally involve the stagnant, dual-permeability models allow also for water
measurement during some experiment of one or several flow within the soil or rock matrix.
HYDRAULIC PROPERTIES OF UNSATURATED SOILS 373

To avoid over-parameterization of the governing equa- provide more realistic simulations of field data than the
tions, one useful simplifying approach is to assume instan- standard approach using unimodal hydraulic properties
taneous hydraulic equilibration between the fracture and of the type shown in Figures 1 and 2. In soils, the two parts
matrix regions. In that case, the Richards equation can still of the conductivity curves may be associated with soil
be used, but now with composite hydraulic properties of structure (near saturation) and soil texture (at lower nega-
the form (e.g., Peters and Klavetter, 1988) tive pressure heads).
The use of composite hydraulic functions such as those
yðhÞ ¼ wf yf ðhÞ þ wm ym ðhÞ (9a) shown in Figure 2 is consistent with field measurements
suggesting that the macropore conductivity of soils at
KðhÞ ¼ wf Kf ðhÞ þ wm Km ðhÞ (9b) saturation is generally about one to two orders of magni-
where the subscripts f and m refer to the fracture tude larger than the matrix conductivity at saturation,
(macropore) and matrix (micropore) regions, respectively, depending upon texture. These findings were confirmed
and where wi are volumetric weighting factors for the two by Schaap and van Genuchten (2006) using a detailed
overlapping regions such that wf + wm = 1. Rather than neural network analysis of the UNSODA unsaturated soil
using Equations 6a,b directly in Equations 9a and 9b, hydraulic database (Leij et al., 1996). The analysis
Durner (1994) proposed a slightly different set of equa- revealed a relatively sharp decrease in the conductivity
tions for the composite functions as follows away from saturation and a slower decrease afterward.
Schaap and van Genuchten (2006) suggested an improved
yðhÞ  yr wf wm composite function for K(h) to account for the effects of
Se ðhÞ ¼ ¼ m þ m
ys  yr ½1 þ jaf hjnf  f ½1 þ jam hjnm  m macropores near saturation as follows:
(10a) RðhÞ
Ks
KðhÞ ¼ Km ðhÞ (11a)
KðSe Þ ¼ Km ðhÞ
 l n 1=m mf
o
mm 2
wf Sef þ wm Sem wf af ½1  ð1  Sef f Þ  þwm am ½1  ð1  Se1=m
m
m
Þ  where
Ks  2 8
wf af þ wm am
< 0 h < 40 cm:
(10b) RðhÞ ¼ 0:2778 þ 0:00694h 40  h < 4 cm
:
where ai, ni, and mi (=1  1/ni) are empirical parameters of 1 þ 0:1875h 4  h  0 cm
the separate hydraulic functions (i = f,m). An example of (11b)
composite retention and hydraulic conductivity functions
based on Equations 10a and 10b is shown in Figure 2 for and where Km(h) is the traditional hydraulic conductiv-
the following set of parameters: yr = 0.00, ys = 0.50, ity function for the matrix as given by Equation 6b.
l = 0.5, Ks = 1 cm d1, am = 0.01 cm1, nm = 1.50, wm = Equations 11a and 11b were found to produce very small
0.975, wf = 0.025, af = 1.00 cm1, and nf = 5.00. The systematic errors between the observed (UNSODA) and
fracture domain in this case represents only 2.5% of the calculated hydraulic conductivities across a wide range
entire pore space, but accounts for almost 90% of the of pressure heads between saturation and 150 m.
hydraulic conductivity close to saturation (Figure 3). While the macropore contribution was most significant
While still leading to uniform flow, models using such between pressure heads 0 and 4 cm, its influence on
composite media properties do allow for faster flow and the conductivity function extended to pressure heads as
transport during conditions near saturation, and as such low as 40 cm (Equation 11b).

0.5
Volumetric water content, [–]

0
0.4
−2
log(K), [cm/day])

0.3
−4

0.2 −6

0.1 −8

0.0 −10
−1 0 1 2 3 4 −1 0 1 2 3 4
log(|h|, [cm]) log(|h|, [cm])

Hydraulic Properties of Unsaturated Soils, Figure 3 Bimodal water retention (left) and hydraulic conductivity (right) functions as
described with the composite soil hydraulic model of Durner (1994).
374 HYDRAULIC PROPERTIES OF UNSATURATED SOILS

Multiphase constitutive relationships media (macroporous soils and fractured rock). The
The use of Equation 1 implies that the air phase has no hydraulic properties of such media may require special
effect of water flow. This is a realistic assumption for most provisions to account for the effects of soil texture and soil
flow simulations, except near saturation in relatively structure on the shape of the hydraulic functions near sat-
closed systems where air may not move freely. The uration, thus leading to dual- or multi-porosity formula-
resulting situation may need to be described using two tions as indicated by Schaap and van Genuchten (2006)
flow equations, one for the air phase and one for the liquid and Jarvis (2008), among others. Estimation of the effective
phase. The same is true for multiphase air, oil, and water properties of heterogeneous (including layered) field soil
systems in which the fluids are not fully miscible. Flow profiles also remains an important challenge. Very promis-
in such multiphase systems generally require flow equa- ing here is the increased integration of hard (directly
tions for each fluid phase involved. Two-phase air-water measured) data and soft (indirectly estimated) information
systems hence could be modeled also using separate equa- for improved estimation of field- or larger-scale hydraulic
tions for air and water. This shows that the standard properties, including the use of noninvasive geophysical
Richards equation is a simplification of a more complete information. New noninvasive technologies with enormous
multiphase (air-water) approach in that the air phase is potential range from neutron and X-ray radiography and
assumed to have a negligible effect on variably saturated magnetic resonance imaging at relatively small (laboratory)
flow, and that the air pressure varies only little in space scales, to electrical resistivity tomography and ground
and time. This assumption appears adequate for most var- penetrating radar at intermediate (field) scales, to passive
iably saturated flow problems. Similar assumptions, how- microwave remote sensing at regional or larger scales.
ever, are generally not possible when nonaqueous phase Challenges remain on how to optimally integrate,
liquids (NAPLs) are present. Mathematical descriptions assimilate, or otherwise fuse such information with direct
of multiphase flow and transport hence in general require laboratory and field hydraulic measurements (Yeh and
flow equations for each of the fluid phases involved. Šimůnek, 2002; Kowalski et al., 2004, Looms et al.,
Assuming applicability of the van Genuchten hydraulic 2008; Ines and Mohanty, 2008), including the optimal
functions and ignoring the presence of residual air and and cost-effective use of pedotransfer function and soil
water, the hydraulic conductivity functions for the liquid texture information, and resultant quantification of uncer-
(wetting) and air phase (non-wetting) phases are given tainty (Minasny and McBratney, 2002; Wang et al., 2003).
by (e.g., Luckner et al., 1989; Lenhard et al., 2002): These various integrated technologies undoubtedly will
h  m i2 further advance in the near future, as well as the use of
Kw ðSe Þ ¼ Kw Sel 1  1  Se1=m (12a) increasingly refined pore-scale modeling approaches
(e.g., Tuller and Or, 2001) at the smaller scales for more
h i2m precise simulation of the basic physical processes
Ka ðSe Þ ¼ Ka ð1  Se Þl 1  Se1=m (12b) governing the retention and movement of water in unsatu-
rated media.
where the subscripts w and a refer to the water and air
phases, respectively, and Kw and Ka are the hydraulic con-
ductivities of the medium to water and air when filled Bibliography
completely with those fluids. A detailed overview of var- Abbaspour, K. C., Schulin, R., and Th. van Genuchten, M., 2001.
ious approaches for measuring and describing the hydrau- Estimating unsaturated soil hydraulic parameters using ant col-
ony optimization. Advances in Water Resources, 24, 827–841.
lic properties of multi-fluid systems is given by Lenhard Bitterlich, S., Durner, W., Iden, S. C., and Knabner, P., 2004. Inverse
et al. (2002). estimation of the unsaturated soil hydraulic properties from col-
umn outflow experiments using free-form parameterizations.
Vadose Zone Journal, 3, 971–981.
A look ahead Bouma, J., and van Lanen, J. A. J., 1987. Transfer functions and
The unsaturated soil hydraulic properties are key factors threshold values: from soil characteristics to land qualities. In
determining the dynamics and movement of water and Beek, K. J., et al. (eds.), Quantified Land Evaluation. Enschede:
its dissolved constituents in the subsurface. Reliable esti- Int. Inst. Aerospace Surv. Earth Sci. ITC Publ. No. 6,
mates are needed for a broad range of agrophysical appli- pp. 106–110.
Brooks, R. H., and Corey, A. T., 1964. Hydraulic properties of
cations, including for subsurface contaminant transport porous media, Hydrol. Paper No. 3, Colorado State Univ., Fort
studies. A large number of approaches are now available Collins, CO.
for describing and measuring the hydraulic properties, Burdine, N. T., 1953. Relative permeability calculations from pore-
especially for relatively homogeneous single-porosity size distribution data. Petr. Trans. Am. Inst. Mining Metall. Eng.,
soils. This includes direct measurement of discrete y(h), 198, 71–77.
and K(h) or K(y) data points and fitting appropriate analyt- Carsel, R. F., and Parrish, R. S., 1988. Developing joint probability
distributions of soil water retention characteristics. Water
ical models to the data, and the use of increasingly sophis- Resources Research, 24, 755–769.
ticated inverse methods. Childs, E. C., and Collis-George, N., 1950. The permeability of
Considerable challenges remain in the description and porous materials. Proceedings of the Royal Society of London,
measurement of the hydraulic properties of structured Series A, 201, 392–405.
HYDRAULIC PROPERTIES OF UNSATURATED SOILS 375

Dane, J. H., and Hopmans, J. W., 2002. Water retention and storage. Part 1, Physical Methods, 3rd edn. Madison: Soil Science Soci-
In Dane, J. H., and Topp, G. C. (eds.), Methods of Soil Analysis, ety of America, pp. 1009–1045.
Part 1, Physical Methods. Chapter 3.3, 3rd edn. Madison, WI: Lenhard, R. J., Oostrom, M., and Dane, J. H., 2002. Multi-fluid
Soil Science Society of America, pp. 671–720. flow. In Dane, J. H., and Topp, G. C. (eds.), Methods of Soil
Dane, J. H., and Topp, G. C. (eds.), 2002. Methods of Soil Analysis, Analysis, Part 4, Physical Methods, SSSA Book Series 5,
Part 1, Physical Methods, 3rd edn. Madison: Soil Science Soci- Chapter 7. Madison: Soil Science Society of America,
ety of America. pp. 1537–1619.
Durner, W., 1994. Hydraulic conductivity estimation for soils with Looms, M. C., Binley, A., Jensen, K. H., Nielsen, L., and Hansen,
heterogeneous pore structure. Water Resources Research, T. M., 2008. Identifying unsaturated hydraulic properties using
32(9), 211–223. an integrated data fusion approach on cross-borehole geophysi-
Gee, G. W., and Ward, A. L., 1999. Innovations in two-phase mea- cal data. Vadose Zone Journal, 7, 238–248.
surements of hydraulic properties. In Th. van Genuchten, M., Luckner, L., van Genuchten, M Th., and Nielsen, D. R., 1989.
Leij, F. J., and Wu, L. (eds.), Proceedings of the International A consistent set of parametric models for the two-phase flow
Workshop, Characterization and Measurement of the Hydraulic of immiscible fluids in the subsurface. Water Resources
Properties of Unsaturated Porous Media. Parts 1 and 2. River- Research, 25(10), 2187–2193.
side: University of California, pp. 241–270. Minasny, B., and McBratney, A. B., 2002. The efficiency of various
Haverkamp, R. F. J., Leij, C. F., Sciortino, A., and Ross, P. J., 2005. approaches to obtaining estimates of soil hydraulic properties.
Soil water retention: I Introduction of a shape index. Soil Science Geoderma, 107(1–2), 55–70.
Society of America Journal, 69, 1881–1890. Millington, R. J., and Quirk, J. P., 1961. Permeability of porous
Hopmans, J. W., Šimůnek, J., Romano, N., and Durner, W., 2002. materials. Transactions of the Faraday Society, 57, 1200–1206.
Inverse modeling of transient water flow. In Dane, J. H., and Mualem, Y., 1976. A new model for predicting the hydraulic con-
Topp, G. C. (eds.), Methods of Soil Analysis, Part 1, Physical ductivity of unsaturated porous media. Water Resources
Methods, 3rd edn. Madison: SSSA, pp. 963–1008. Research, 12, 513–522.
Hendrickx, J. M. H., and Flury, M., 2001. Uniform and preferential Mualem, Y., 1992. Modeling the hydraulic conductivity of unsatu-
flow, mechanisms in the vadose zone. In Conceptual models of rated porous media. In Th. van Genuchten, M., Leij, F. J., and
flow and transport in the fractured vadose zone. Washington, Lund, L. J. (eds.), Proceedings of the International Workshop,
DC: National Research Council, National Academy, Indirect Methods for Estimating the Hydraulic Properties of
pp. 149–187. Unsaturated Soils. Riverside: University of California, pp. 45–52.
Ines, A. V. M., and Mohanty, B., 2008. Near-Surface soil moisture Nimmo, J. R., Perkins, K. S., and Lewis, A. M., 2002. Steady-state
assimilation for quantifying effective soil hydraulic properties centrifuge. In Dane, J. H., and Topp, G. C. (eds.), Methods of Soil
under different hydroclimatic conditions. Vadose Zone Journal, Analysis, Part 4, Physical Methods, SSSA Book Series 5,
7, 39–52. Chapter 7. Madison: Soil Science Society of America,
Jarvis, N., 2008. Near-saturated hydraulic properties of pp. 903–916.
macroporous soils. Vadose Zone Journal, 7, 1256–1264. Pachepsky, Y. A., and Rawls, W. J., 2004. Status of pedotransfer
Kastanek, F. J., and Nielsen, D. R., 2001. Description of soil water functions. In Pachepsky, Y. A., and Rawls, W. J. (eds.), Develop-
characteristics using cubic spline interpolation. Soil Science ment of Pedotransfer Functions in Soil Hydrology. Amsterdam:
Society of America Journal, 65, 279–283. Elsevier, pp. vii–xviii.
Köhne, J. M., Köhne, S., and Šimůnek, J., 2009. A review of model Parlange, J.-Y., 1980. Water transport in soils. Annual Review of
applications for structured soils: a) Water flow and tracer trans- Fluid Mechanics, 12, 77–102.
port. Journal of Contaminant Hydrology, 104, 4–35. Peters, R. R., and Klavetter, E. A., 1988. A continuum model for
Kool, J. B., Parker, J. C., and van Genuchten, M Th, 1987. Param- water movement in an unsaturated fractured rock mass. Water
eter estimation for unsaturated flow and transport models – a Resources Research, 24, 416–430.
review. Journal of Hydrology, 91, 255–293. Philip, J. R., 1969. Theory of infiltration. In Chow, V. T. (ed.),
Kosugi, K., 1996. Lognormal distribution model for unsaturated Advances in Hydroscience, 5, 215–296; New York: Academic,
soil hydraulic properties. Water Resources Research, 32(9), pp. 215–296.
2697–2703. Rawls, W. J., Brakensiek, D. L., and Saxton, K. E., 1982. Estimation
Kosugi, K., Hopmans, J. W., and Dane, J. H., 2002. Parametric of soil water properties. Transactions on ASAE, 25(5), 1316–
models. In Dane, J. H., and Topp, G. C. (eds.), Methods of Soil 1320 and 1328.
Analysis, Part 1, Physical Methods, 3rd edn. Madison: SSSA, Richards, L. A., 1931. Capillary conduction of fluid through porous
pp. 739–757. mediums. Physics, 1, 318–333.
Kowalski, M. B., Finsterle, S., and Rubin, Y., 2004. Estimating slow Ritsema, C. J., and Dekker, L. W. (eds.), 2005. Behaviour and man-
parameter distributions using ground-penetrating radar and agement of water-repellent soils. Australian Journal of Soil
hydrological measurements during transient flow in the vadose Research, 43(3), i–ii, 1–441.
zone. Advances in Water Resources, 27, 583–599. Schaap, M. G., and Leij, F. J., 2000. Improved prediction of unsat-
Kung, K.-J. S., 1990. Preferential flow in a sandy vadose zone. 2. urated hydraulic conductivity with the Mualem-van Genuchten
Mechanism and implications. Geoderma, 46, 59–71. Model. Soil Science Society of America Journal, 64, 843–851.
Leij, F. J., Alves, W. J., van Genuchten, M. Th., and Williams, J. R., Schaap, M. G., and van Genuchten, M Th., 2006. A modified
1996. The UNSODA-Unsaturated Soil Hydraulic Database. Mualem-van Genuchten formulation for improved description
User’s manual Version 1.0. Report EPA/600/R-96/095. National of the hydraulic conductivity near saturation. Vadose Zone Jour-
Risk Management Research Laboratory, Office of Research and nal, 5, 27–34.
Development, U.S. Environmental Protection Agency, Segal, E., Bradford, S. A., Shouse, P., Lazarovitch, N., and Corwin,
Cincinnati, OH. D., 2008. Integration of hard and soft data to characterize field-
Leij, F. J., Russell, W. B., and Lesch, S. M., 1997. Closed-form scale hydraulic properties for flow and transport studies. Vadose
expressions for water retention and conductivity data. Ground Zone Journal, 7, 878–889.
Water, 35, 848–858. Šimůnek, J., and van Genuchten, M Th., 2008. Modeling
Leij, F. J., Schaap, M. G., and Arya, L. R., 2002. Indirect methods. nonequilibrium flow and transport processes using HYDRUS.
In Dane, J. H., and Topp, G. C. (eds.), Methods of Soil Analysis, Vadose Zone Journal, 7, 782–797.
376 HYDRODYNAMIC DISPERSION

Šimůnek, J., Wang, D., Shouse, P. J., and van Genuchten, M Th.,
1998a. Analysis of field tension disc infiltrometer data by param- HYDRODYNAMIC DISPERSION
eter estimation. International Agrophysics, 12, 167–180.
Šimůnek, J., Wendroth, O., and van Genuchten, M Th., 1998b. The tendency of a flowing solution in a porous medium
A parameter estimation analysis of the evaporation method for that is permeated with a solution of different composition
determining soil hydraulic properties. Soil Science Society of
America Journal, 62(4), 894–905. to disperse, due to the non-uniformity of the flow velocity
Šimůnek, J., Jarvis, N. J., Th. van Genuchten, M., and Gärdenäs, A., in the conducting pores. The process is somewhat analo-
2003. Nonequilibrium and preferential flow and transport in the gous to diffusion, though it is a consequence of
vadose zone: review and case study. Journal of Hydrology, 272, convection.
14–35.
Tuller, M., and Or, D., 2001. Hydraulic conductivity of variably sat-
urated porous media – film and corner flow in angular pore
Bibliography
space. Water Resources Research, 37, 1257–1276. Introduction to Environmental Soil Physics (First Edition) 2003
van Dam, J. C., Stricker, J. N. M., and Droogers, P., 1994. Inverse Elsevier Inc. Daniel Hillel (ed.) http://www.sciencedirect.com/
method for determining soil hydraulic functions from multi-step science/book/9780123486554
outflow experiments. Soil Science Society of America Journal,
58, 647–652.
van Genuchten, M.Th., 1980. A closed-form equation for predicting
the hydraulic conductivity of unsaturated soils. Soil Science
Society of America Journal, 44, 892–898.
van Genuchten M.Th., Leij, F. J., and Yates, S. R., 1991. The RETC HYDROPEDOLOGICAL PROCESSES IN SOILS
code for quantifying the hydraulic functions of unsaturated soils.
Report EPA/600/2–91/065. U.S. Environmental Protection Svatopluk Matula
Agency, Ada, OK. Department of Water Resources, Food and Natural
Vereecken, H., Maes, J., Feyen, J., and Darius, P., 1989. Estimating
the soil moisture retention characteristic from texture, bulk- Resources, Czech University of Life Sciences, Prague,
density, and carbon content. Soil Science, 148, 389–403. Czech Republic
Vrugt, J. A., and Robinson, B. A., 2007. Improved evolutionary
optimization from genetically adaptive multimethod search. Synonyms
Proceedings of the National Academy of Sciences of the United
States of America, 104, 708–711. Soil hydrophysical processes; Soil physical processes;
Wraith, J. M., and Or, D., 1998. Nonlinear parameter estimation Soil water processes; Soil water–soil morphology
using spreadsheet software. Journal of Natural Resources and interactions
Life Sciences Education, 27, 13–19.
Wang, W., Neuman, S. P., Yao, T., and Wierenga, P. J., 2003. Simu- Definition
lation of large-scale field infiltration experiments using a hierar-
chy of models based on public, generic, and site data. Vadose The hydropedological processes in the broader sense are all
Zone Journal, 2, 297–312. soil processes in which flowing or stagnant water acts as the
Yeh, T.-C. J., and Šimůnek, J., 2002. Stochastic fusion of informa- environment or agent or the vehicle of transport. These pro-
tion for characterizing and monitoring the vadose zone. Vadose cesses affect the visible or otherwise discernible morpholog-
Zone Journal, 1, 227–241. ical features of the soil profile and analogous features on the
pedon, polypedon, catena, and soil landscape or soil series
scales. These features can be distinguished and categorized
Cross-references
according to various pedological classification systems
Bypass Flow in Soil (Lal, 2005) and, vice versa, used to identify and semi-
Databases of Soil Physical and Hydraulic Properties
Ecohydrology quantify the soil water processes (e.g., Stewart and Howell,
Evapotranspiration 2003) that have produced or affected them. In the narrower
Field Water Capacity sense, only those processes in which water itself (its content,
Hydropedological Processes in Soils energy status, movement, and balance) is in the focus are
Hysteresis in Soil regarded as hydropedological processes.
Infiltration in Soils
Layered Soils, Water and Solute Transport Introduction
Pedotransfer Functions
Physics of Near Ground Atmosphere Pedology is the branch of soil science dealing with soil
Pore Size Distribution genesis, morphology, and classification. In some parts of
Soil Hydrophobicity and Hydraulic Fluxes the world, however, the world pedology has been or still
Soil Water Flow is used to denote the whole of soil science. Under these
Soil Water Management conditions, it was quite natural to name that branch of soil
Sorptivity of Soils science that deals with soil water (e.g., Stewart and
Surface and Subsurface Waters
Tensiometry Howell, 2003) and is otherwise referred to, for example,
Water Balance in Terrestrial Ecosystems as soil physics, soil water physics, physics of soil water,
Water Budget in Soil or soil hydrology as hydropedology, notwithstanding its
Wetting and Drying, Effect on Soil Physical Properties relations (or rather the absence of such relations) to soil

You might also like