Você está na página 1de 20

Special Section: VZJ

Anniversary Issue

Shmuel Assouline*
Dani Or

The water retention curve (WRC)


and the hydraulic conductivity function (HCF) are key ingredients in most
analytical and numerical models for
flow and transport in unsaturated
porous media. We review some of
the primary models and highlight
their physical basis, assumptions,
advantages and limitations.
S. Assouline, The Dep. of Environmental
Physics, Institute of Soil, Water and Environmental Sciences, A.R.O.- Volcani Center,
Bet Dagan 50250, Israel; D. Or, Dep. of Environmental Sciences (D-UWIS), Institute of
Terrestrial Ecosystems (ITES), Soil and Terrestrial Environmental Physics (STEP), Swiss
Federal Institute of Technology (ETH), Zurich,
Switzerland. Contribution of the Agricultural Research Organization, Institute of
Soil, Water and Environmental Sciences, Bet
Dagan, Israel, no. 606/12. *Corresponding
author (vwshmuel@agri.gov.il).
Vadose Zone J.
doi:10.2136/vzj2013.07.0121
Received 8 July 2013.

Soil Science Society of America


5585 Guilford Rd., Madison, WI 53711 USA.

All rights reserved. No part of this periodical may


be reproduced or transmitted in any form or by any
means, electronic or mechanical, including photocopying, recording, or any information storage
and retrieval system, without permission in writing
from the publisher.

Conceptual and Parametric


Representation of Soil Hydraulic
Properties: A Review
The water retention curve (WRC) and the hydraulic conductivity function (HCF) are key ingredients in most analytical and numerical models for flow and transport in unsaturated porous
media. Despite their formal derivation for a representative elementary volume (REV) of soil
complex pore spaces, these two hydraulic functions are rooted in pore-scale capillarity and viscous flows that, in turn, are invoked to provide interpretation of measurements and processes,
such as linking WRC with the more difficult to measure HCF. Numerous conceptual and parametric models were proposed for the representation of processes within soil pore spaces and
inferences concerning the two hydraulic functions (WRC and HCF) from surrogate variables. We
review some of the primary models and highlight their physical basis, assumptions, advantages,
and limitations. The first part focuses on the representation and modeling of WRC, including
recent advances such as capillarity in angular pores and film adsorption and present empirical
models based on easy to measure surrogate properties (pedotransfer functions). In the second
part, we review the HCF and focus on widely used models that use WRC information to predict the saturated and unsaturated hydraulic conductivity. In the third part, we briefly review
issues related to parameter equivalence between models, hysteresis in WRC, and effects of
structural changes on hydraulic functions. Recent technological advances and monitoring networks offer opportunities for extensive hydrological information of high quality. The increase
in measurement capabilities highlights the urgent need for building a hierarchy of parameters
and model structures suitable for different modeling objectives and predictions across spatial
scales. Additionally, the commonly assumed links between WRC and HCF must be reevaluated
and involve more direct measurements of HCF. The modeling of flow and transport through
structured and special porous media may require special functions and reflecting modifications in the governing equations. Finally, the impact of dynamics and transient processes at
fluid interfaces on flow regimes and hydraulic properties necessitate different modeling and
representation strategies beyond the present REV-based framework.
Abbreviations: BCC, bundle of cylindrical capillaries; HCF, hydraulic conductivity function; PSD, particle size
distribution; PTF, pedotransfer function; PVD, pore volume distribution; REV, representative elementary
volume; WRC, water retention curve.

Most operational models for flow and transport in porous media rely on the
concept of representative elementary volume (REV) (Bear, 1972; Scheidegger, 1974) to
facilitate continuum description of water retention and flow in complex pore spaces. Such
representation invokes various simplifying assumptions concerning capillarity and viscous
flow within soil pores. The outcome of such volume averaged pore-scale processes yields the
familiar macroscopic hydraulic functions such as the soil water retention curve (WRC)
and the hydraulic conductivity function (HCF). The WRC describes the relationship
between soil matric potential (or capillary head), y, and the soil water content, q (expressed
volumetrically or gravimetrically), under equilibrium conditions. The pioneering work
of Buckingham (1907) established the basis for expressing the hydraulic conductivity as
a function of the hydrologic state variables q and y. The HCF takes on a form similar to
Darcys saturated hydraulic conductivity Ks parameter (a coefficient of proportionality
between flux and hydraulic gradient; Darcy, 1856). The main difference between Darcys
constant, Ks , and Buckinghams (1907) HCF stems from unsaturated conditions giving
rise to a highly nonlinear HCF that varies with q or y. The HCF is often scaled by Ks ,
and expressed in a dimensionless form as Kr(q) = K(q)/Ks or Kr(y) = K(y)/Ks. The soil
water content may also be scaled and represented as dimensionless effective saturation, Se,
defined by Se(q) = [(q q r)/(q s q r)], with q s and q r being the soil saturated and residual
volumetric water contents, respectively.
Experimental determination of WRC by means of traditional methods (e.g., hanging
water column, pressure membrane apparatus) is laborious and time-consuming, typically

www.VadoseZoneJournal.org

yielding only a few pairs of (y, q) values for a soil sample. These
limitations motivated efforts to relate WRC to other simpler
and easy-to-measure surrogate soil properties. Recent measuring
devices based on the evaporation method (Gardner and Miklich,
1962; Wind, 1968) give water retention data under laboratory conditions almost as a continuous function (Schindler and Mller,
2006; Peters and Durner, 2008a; Schelle et al., 2010). However,
the need for continuous representation of the WRC for numerical
modeling provided the motivation for the development of numerous mathematical expressions fitting the measured (y, q) data.
The experimental determination of Kr is considerably more complicated than that of WRC. Hence, for many practical applications,
the HCF is deduced from WRC information (soil pore sizes), often
supplemented by direct measurements of soil Ks. Most models
that link information deduced from WRC to soil HCF estimation invariably assume simple pore geometry to enable capillary
and viscous considerations for obtaining mathematically tractable
HCF expressions. Among the conceptual models for porous media
proposed over the years, the simplest and most widely used considers soil pores as an assembly of parallel cylindrical capillaries (Fatt
and Dykstra, 1951). An increment of complexity was added by considering a statistical distribution of capillary radii for the assumed
bundle of capillaries (Purcell, 1949; Burdine, 1953). Additionally,
Childs and Collis-George (1950) considered interconnections
along the bundle of capillaries that were conceptually cut and
randomly rejoined. More recently, authors have reformulated
the concepts of Childs and Collis-George (1950) by considering
bundles of capillaries with random pore size distributions that may
vary along each capillary (Dullien et al., 1977).
In the following we review conceptual approaches for modeling and representing soil WRC and HCF starting with models
for soil WRC in the first part, followed by conceptual and
empirical models for soil Ks and Kr (HCF) in the second part.
We will address herein only the hydraulic properties at equilibrium. Dynamic conditions could affect the expression of these
hydraulic functions (Diamantopoulos and Durner, 2012). In
the third part of the review, we consider topics related to model
parameter equivalence, representation of WRC hysteresis, effects
of structural changes on soil hydraulic functions, and properties of coarse porous media. Finally, we close with a look at key
challenges and opportunities for improved representation of soil
hydraulic functions.

66Functions for the Soil Water

Retention Curve

The most widely used approaches for representing soil WRC could
be grouped into: (i) empirical or parametric functions that fit a
wide range of experimental data, (ii) expressions based on soil particle size distribution, (iii) expressions on fractal representation of
soil pore spaces, and (iv) pedotransfer functions based on simple

www.VadoseZoneJournal.org

soil surrogate variables. A related class of WRC expressions was


derived from pore-scale water retention in different geometries
(Tuller et al., 1999; Lebeau and Konrad, 2010; Likos and Jafaar,
2013) that were upscaled to represent sample scale WRC. However,
these geometry-explicit models have not been widely used in practice and were implemented primarily for the estimation of the soil
hydraulic conductivity function.

Parametric and Empirical Expressions for WRC

We refer to a definitive review of parametric models by Kosugi et


al. (2002) and present in the following only the most widely used
expressions representing WRC. One of the simplest parametric
models for WRC was proposed by Tani (1982), it is an exponential
function with a single fitting parameter:
y

y Se (y) = 1 + e yo
y o

[1.1]

where y o is the capillary head at the WRCs inflection point.


Equation [1.1] was extended by Russo (1988):
2

y 2+ k

y -

Se (y) = 1 + e yo

y o

[1.2]

where k is an empirical constant (related to the power l in Eq. [2.17]


below). Interestingly, Russo (1988) derived Eq. [1.2] starting from
an exponential HCF as discussed in the next section.
Brooks and Corey (1964) proposed a widely used power function
of y to represent Se(y):
y -l

Se (y) =
; y < yc

y c
S e ( y) = 1
; y yc

[1.3]

The Brooks and Corey (1964) model contains a discontinuity


at the pressure value of y c that represents the air entry value, or
the capillary pressure at which the largest connected pores in
the soil sample are invaded by air. The dimensionless parameter
l reflects the pore size distribution index (large l values correspond to sandy soils). Campbell (1974) proposed an expression
similar to Eq. [1.3].
With the advent on numerical solutions for unsaturated flow, the
discontinuity in the WRC derivative at y c introduced numerical
difficulties. Additionally, evidence suggests that the discontinuity
is not present in field-measured WRC data (Milly, 1987). To overcome this limitation, Clapp and Hornberger (1978) and Hutson
and Cass (1987) suggested replacing the sharp discontinuity at y c
with a parabolic curve joining the two-part function.

p. 2 of 20

Brutsaert (1966) proposed an expression that provides for a


continuous description of WRC for entire range from Se = 0 to
Se = 1:

WRC representing the related pore size distribution. The proposed


models have used similarity in the distributions of particle to pore
sizes, or invoked packing arguments giving rise to the WRC.

-1
Se (y) = 1 + ( a y )n

Arya and Paris (1981) developed a model to predict the WRC of a


soil from its PSD, bulk density, and particle density, based on the
assumption that the mean pore radius is proportional to the mean
particle radius. The pore radii are then converted in volumetric
water content and in equivalent soil water capillary head using the
equation of capillarity (y = 14.9 r1 for r and y in mm). Model
predictions for seven soil materials show close agreement with the
experimental data. This model was further developed in Arya et
al. (1999).

[1.4]

where a and n are empirical parameters resulting from best-fit


procedures to data. Brutsaerts (1966) expression did not receive
much attention until van Genuchten (1980) introduced a similar
expression with a third parameter, m:
-m
Se (y) = 1 + (a y )n

[1.5]

Compared to the Brooks and Coreys model (Eq. [1.3]), the expressions in Eq. [1.4] and [1.5] do not account directly for an air entry
matric potential, but the parameter a is inversely proportional to
the air entry matric potential. The expression by van Genuchten
(1980) provided good fit to WRC data from many soils, particularly for data near saturation (van Genuchten and Nielsen, 1985),
and thus became very popular among modelers. Nimmo (1991)
and Ross et al. (1991) found that the van Genuchten model is successful at high and medium water contents, but often give poor
results at low water contents. Campbell and Shiozawa (1992) proposed a modification of the van Genuchten model for improving
model fit to dry range data.
Kosugi (1994, 1996) proposed a WRC model that results from
representing pore radii distribution based on a three-parameter
lognormal distribution and applying the capillary rise equation to
deduce the corresponding capillary head distribution. The resulting expression for the WRC is:
y -y 2
-s
ln c

y c -y o
1
; y < y
Se (y) = erfc


2
2s

S e ( y) = 1
; y yc

[1.6]

This model presents four fitting parameters: q r, y c, y o, and a


dimensionless parameter, s, which is related to the width of the
pore radius distribution. The model was tested for fifty soils and
produced consistent and reasonable results.

Linking Soil Particle Size Distribution


to Pore Sizes and WRC

The WRC is often interpreted as representative of distribution of


pore sizes of a prescribed simple geometry (e.g., cylinder, sphere).
These textural pores are the voids resulting from packing of solid
particles. It is thus attractive to seek quantitative links between
easy-to-measure particle size distribution (PSD) and the desired

www.VadoseZoneJournal.org

Haverkamp and Parlange (1986) also assumed a linear relationship between soil particle diameter and equivalent pore radius to
derive an analytical expression for the WRC. The application of
the model to sandy soils (with no organic matter) yielded reasonable agreement.
Nimmo et al. (2007) relied on the Arya and Paris model to develop
a property transfer model that translates PSD into WRC. Fredlund
et al. (2002) suggested a model of the WRC based on the PSD,
assuming a constant packing factor independent of grain size to
convert PSD into pore volume distribution (PVD). Mohammadi
and Vanclooster (2011) proposed a model that related PSD to
WRC assuming a packing state parameter that characterized each
size fraction in the PSD.
Detailed experimental results (Crisp and Williams, 1971; Glover
and Walker, 2009) paint a different picture than the constant proportionality between PSD and PVD. Results suggest that such
proportionality holds only for mean values of particle and pore
sizes, but not for the entire distribution. Hence, it is not surprising
that the coefficient of proportionality between particle and pore
sizes is expected to vary with soil type, mean particle size, and
packing. Even simple geometrical considerations for the packing of
monodisperse spheres show that the ratio of inscribed pore diameter, d, to particle diameter, D, varies with packing angle:
d (1- cos J)
=
[1.7]
D
cos J
where J is half-packing angle between centers of neighboring
spheres (with d/D = 0.41 for J = 45 corresponding to cubic
packing; and d/D = 0.15 for J = 30 corresponding to tetrahedral packing). Analyses of Glover and Walker (2009) show a value
of d/D = 0.29 for packs of spherical particles but a wide range
of proportionality between particles and pore sizes for different
natural media. Rouault and Assouline (1998) have shown that for
polydisperse dense packs, linear relationship between particle and
pore sizes is inadequate, and, instead, a power function describes
better that relationship. In short, while particle and pore size
p. 3 of 20

distributions are linked, the relationships are not straightforward,


and often the resulting pore size distribution is broader than the
particle size distribution even for monodisperse packs (see Fig. 1
reproduced from Crisp and Williams, 1971; and results of Zhang
et al., 1997; Assouline and Rouault, 1997).
Chan and Govindaraju (2003, 2004) conceptualized a porous
medium composed of polydisperse and lognormally distributed
spheres (similar to the assumptions made by Shirazi and Boersma,
1984; Buchan, 1989; Shiozawa and Campbell, 1991). They applied
concepts from Torquato (2002) to statistically link PSD for polydisperse spherical particles and the resulting WRC with reasonable
results for sand and loamy sand soils.
An interesting aspect concerns the statistical distribution of particles or aggregate sizes in relation to the underlying fragmentation
process in their formation (Bittelli et al., 1999). The lognormal distribution of particle sizes results from breakup process where the
fragmentation probability is independent of particle size (Tenchov
and Yanev, 1986). Evidence suggests that in soil, the fragmentation

probability of soil particles or aggregates is proportional to their


sizes (Hadas and Wolf, 1984). For a uniform and random fragmentation process, the probability distribution function of particle
sizes is an exponential function (Tenchov and Yanev, 1986):
F (va ) = 1- exp(-eva ) ; va =

va - va min
va max - va min

[1.8]

where va is the soil particle/aggregate size constrained between a


minimal, va min, and a maximal value, va max . Invoking a power
function to link between particles and pores size (Assouline and
Rouault, 1997; Rouault and Assouline, 1998), such as vp = gvau,
the resulting probability distribution function of the pore sizes is
a Weibull distribution:
F (vp ) = 1- exp[-w(va )u ] ; vp =

vp - vp min

[1.9]

vp max - vp min

where vp is the soil pore size varying between a minimal value,


vp min, and a maximal one, vp max. Assouline et al. (1998, 2000)
applied such approach to model the related WRC according to:
Se(y) = 1 exp[x(|y|-1 |yL|-1)m]; 0 |y| |yL|

[1.10]

where x and m are two fitting parameters, and y L is the capillary


head corresponding to a very low water content, q L , which
represents the limit of interest for a particular WRC application.
The expression in Eq. [1.10] can improve in some cases the representation of the WRC at the transition between saturation and
unsaturated conditions. To illustrate the main trends of the expressions presented above, the two-parameter models of Brooks and
Corey (1964) (Eq. [1.3]) and Assouline et al. (1998) (Eq. [1.10])
are fitted to the experimental data of the WRC of a sand (sable
de riviere) and a loam (Pachappa loam) as they were reported in
Mualem (1974a). The results are depicted in Fig. 2. For the sand,
the power function in Eq. [1.3] seems to fit best the WRC, while,
in the case of the loam, an expression that present an inflection
point like Eq. [1.4-1.6] and Eq. [1.10] is in better agreement with
the experimental data.
Hwang (2004) compared the performances of nine PSD models
and concluded that the model of Fredlund et al. (2000), the model
proposed by Skaggs et al. (2001), and the Weibull distribution
based on the model of Assouline et al. (1998) provided the best
fits to data corresponding to 1385 Korean soils. Cornelis et al.
(2005) compared 10 WRC models and concluded that the models
of van Genuchten (1980), Kosugi (1994, 1996), and Assouline et
al. (1998) resulted in the best overall fit for data from 48 horizons
of 24 soils from Belgium.
Fig. 1. Links between particle and pore size distributions of sand
(modified from Crisp and Williams, 1971).

www.VadoseZoneJournal.org

p. 4 of 20

In all the expressions above, one has to determine the


WRC residual water content, q r. The defi nition of this
variable remains ambiguous; some consider q r as a fitting
parameter without ascribing much physical significance to
its value; others have considered this point on the WRC
as the marker for loss of phase continuity that, in turn,
affects a wide range of processes from termination of stage
1 evaporation (Lehmann et al., 2008; Or et al., 2013) to
the onset of fi lm flow for HCF (Tuller and Or, 2001).
Regardless of the interpretation of this WRC parameter,
evidence suggests that the value of q r is closely linked
with soil specific surface area as demonstrated by Tuller
and Or (2005). In their analyses, Tuller and Or (2005)
presented evidence for the near-universal behavior of the
WRC dry-end due to dominance of fi lm adsorption (by
van der Walls forces) and thus predictable dependency
of q r on soil surface area tested for a wide range of soils.
Fig. 2. The water retention curve for a sand and a loam soil. The experimental data
Wang et al. (2008) extended this relationship and derived
(circles) were reported by Mualem (1974a). The curves correspond to Eq. [1.3]
a WRC model based on soil specific surface area and on
(dashed line) and [1.10] (solid line).
the physical and chemical behavior of the water and air
phases in the unsaturated porous medium. The model
performance was good for the soils tested and improved the representation of the dry end of the WRC.

Multi-Modal Water Retention Curve

Porous media containing large contrasts in pore sizesfor example,


pores forming between and within soil aggregates, soils with significant fractions of large biopores within a fi ner matrix (roots,
earthworms), or in fractured porous media with significant fracture void fractioncould be characterized by a bimodal WRC.
Zhang and van Genuchten (1994) proposed a model of the WRC
that represents a sigmoidal curve when its four fitting parameters
version is used and describes bimodal curves when the five fitting
parameters version is used.
However, the practical representation of the bimodal WRC could
be achieved by superposition of two WRC fitted to each of the pore
domains. The challenge of fitting several WRC to such dual porosity
media or multi-modal pore size distributions is relatively minor and
often one ends up with an expression such as (Durner, 1994):
k

-mi
Se (y) = wi 1 + (a i y )ni

[1.11]

i =1

where k is the number of the domains (each with an index i), and wi
represents the weighting factor for the ith sub-WRC corresponding to each domain. This procedure is illustrated in Fig. 3 for the
bimodal case (k = 2). Note, however, that even when the value of
the weighting parameter (wi) for a pore subdomain is relatively
small with minor effect on the apparent multimodal WRC (e.g.,
for fractured rock), the impact on the HCF could be significant,
and the treatment of the resulting HCF should be considered
carefully. For certain flow conditions, large voids (however small

www.VadoseZoneJournal.org

Fig. 3. A porous medium characterized by a bimodal particle size


distribution (upper plot) and the corresponding water retention curve
(lower plot) resulting from Eq. [1.11] with w1 = 0.6 and w2 = 0.4.

p. 5 of 20

their fraction might be) may dominate the flow regime entirely
as in the case of fractured rock or macropores in soil near saturation. Schaap and van Genuchten (2006) and Jarvis (2008) have
suggested models that improve the description of the hydraulic
functions near saturation to account for macroporosity.

of the method are presented in Or and Tuller (1999). More recently,


Likos and Jafaar (2013) extended the treatment to pores forming between sand grains demonstrating reasonable representation
for WRC for coarse-textured media based on their geometrically
explicit pore scale model.

Pore Scale Mechanistic WRCAngular Pores


and Surface Area

Fractals and Percolation Theory

In addition to general dependency on pore size, liquid retention


and flow dynamics in porous media are greatly influenced by pore
shape and angularity. Even cursory inspection shows that soil pore
spaces do not resemble cylindrical capillaries (assumed in idealized representation of capillary phenomenon), and most pores are
angular with rough surfaces (Li and Wardlaw, 1986; Mason and
Morrow, 1991; Tuller et al., 1999). Angular pore shapes are not
only a more realistic representation of natural pore spaces, but they
capture a richer range of capillary liquid behavior than possible
in the classical bundle of cylindrical capillaries (BCC) model. A
key aspect in angular pores is the possibility of dual occupancy
of liquid and gaseous phases within the same invaded pores that
provide a richer depiction of water retention and hydraulic continuity in unsaturated soils (Dullien et al., 1986; Or and Tuller,
2000; Tuller and Or, 2002). These elements were incorporated in
several pore scale models, such as proposed by Tuller et al. (1999),
that provided description of water retention within angular porous
medium that accommodate both capillary and adsorptive forces
affecting water retention. The basic unit pore is comprised of a
polygon-shaped pore (e.g., triangle, square) linked with internal surface area for adsorption (Fig. 4). The unit pore is defined
by three parameters, the polygonal pore size, L (for capillary
processes), and two dimensionless values, b and d (that scale slitspacing bL, and slit-length dL), for fi lm adsorption surface area.
The simple geometry allows accounting for different soil textural
and structural classes by adjusting pore width (L) and the proportions of exposed surfaces. The relative saturation of a pore, Sw, with
area Apore for a chemical potential (P) (or capillary pressure, y) is:
Sw =

r(P)2 F ( g)
s2
=
F ( g)
Apore
Apore P2

Since the early 1980s with the studies of Burrough (1981), and later
on with the work of Tyler and Wheatcraft (1989) and subsequently
Rieu and Sposito (1991), many researchers have attempted to harness the utility and generality of fractal theory (Mandelbrot, 1967)
to represent complex sizes and roughness properties of soil particles
(Borkovec et al., 1993) in a compact manner (typically as power
law). Such representation was not only motivated by experimental
observations of particle and aggregate size distributions (reflecting
weathering and fragmentation by soil forming processes), but often
serving as a stepping stone for compact and general description of
soil pore spaces (Tyler and Wheatcraft, 1989; Young and Crawford,
1991; Bartoli et al., 1991; Perrier et al., 1996; Gimnez et al., 1997;
Xu and P. Dong, 2004) and scaling in soils and heterogeneous
porous media (Guadagnini et al., 2013).
The criticism of Baveye and Boast (1998) notwithstanding, a key
step remains as to how to establish similarity between assumed

[1.12]

where s is the surface tension of the liquid, and r is the radius of


curvature of the liquidvapor interface that is dependent on the
chemical potential P = s/rr, or the capillary pressure y = s/r, and
F(g) a shape or angularity factor dependent on angularity of the
pore cross section only.
The pore-scale model enabled derivation of hydraulic functions
from measurable soil properties (e.g., specific surface area). The
geometry of the proposed unit pore is simple and tractable and
enables upscaling to sample scale using a statistical framework
similar to that used for representation of pore radii distribution
in the BCC (Laroussi and de Backer, 1979; Kosugi, 1994). Details

www.VadoseZoneJournal.org

Fig. 4. Calculated and measured liquid saturation as a function


of chemical potential (matric potential) for Hygiene sandstone
employing a single unit cell with triangular pore cross section and
negligibly small surface area (see Tuller and Or, 2001).

p. 6 of 20

pore space fractal dimension, D, and a practical exponent for the


power law representing the pore size distribution and the WRC.

inserting a penetration resistance parameter into PTFs improved


the estimate of WRC based on soil texture and bulk density.

Tyler and Wheatcraft (1990) obtained a power law expression


for the WRC similar to the Brooks and Corey (1964) (and the
Campbell, 1974) model, where the power (2D) is equivalent to l in
Eq. [1.3] for a porous medium similar to a Sierpinski carpet. Rieu
and Sposito (1991) have shown that starting from a fractal representation of aggregation of structured soils yields fractal voids and
thus (with appropriate scaling of solid mass and pore spaces; Perrier
et al., 1996) recovers the familiar power functions for the WRC.
Note, however, that a fractal model of soil aggregate sizes may
not necessarily lead to a power function for the WRC. Moreover,
studies have shown that a power function of soil WRC does not
guarantee a fractal geometry for soil pore spaces (Crawford et al.,
1995; Bird et al., 1996; Assouline and Rouault, 1997; Rouault and
Assouline, 1998).

The empiricalcorrelative approach at the basis of the PTF


offers reasonable initial estimates for certain large-scale analyses
(Romano, 2004). However, the limited physical basis for the estimates of WRC and applicability within the range of values used for
the regression analysis, necessitate extra caution for their general
application (Chirico et al., 2007). Often, data used for the calibration of the PTFs are from specific locations (regions or countries)
and therefore, direct transfer to elsewhere may lead to significant
errors. Gijsman et al. (2002) compared the performances of eight
different PTFs and observed significant discrepancy among the
results due to the local nature of the data basis. For the combination between the compared PTFs and the database used in that
study, the PTF proposed in Saxton et al. (1986) appeared to perform the best. However, considering different PTFs and using
different databases, the conclusions might differ. In the studies of
Tietje and Tapkenhinrichs (1993) and Kern (1995), best performances were achieved by the PTFs in Vereecken et al. (1989) and
Rawls and Brakensiek (1989), respectively.

Pachepsky et al. (1995a) showed that within the fractal framework


for soil WRC, deviation from power law representation could be
attributed to a multifractal structure of soil pores. This implies a
dependence of the fractal dimension on pore radii. Pachepsky et
al. (1995b) have developed a fractal-based WRC for self-similar
pore sizes with a correction factor, f(r) that accounts for dependence of the fractal dimension on pore radii. Representing f(r) by
a lognormal probability distribution function of the pore radii, the
resulting q(y) function resembles the expression in Eq. [1.6].
More recently, Hunt and coworkers (Hunt and Gee, 2002; Hunt,
2005) expanded the application and interpretation of such fractal
pore space models in the context of percolation theory to capitalize on universal traits in the theory for inferences of critical
volumes for connectivity and critical path to provide insights on
key transport properties (HCF, Diffusion, etc.) as affected by water
retention in such spaces.

Pedotransfer Functions

Experimental challenges in determining WRC limit its


availability in soil databases. On the other hand, such databases
generally provide information on other (easier-to-measure)
soil physical and mechanical properties. Hence, empirical and
statistical functions, called pedotransfer functions (PTFs), a
term coined by Bouma (1989), were proposed to relate simple
and handy soil properties to WRC (see reviews in Rawls et al.,
1991; Timlin et al., 1996; Pachepsky and Rawls, 2004; Saxton
and Rawls, 2006; Vereecken et al., 2011). Most of the variables
found in PTFs are (i) oven-dry bulk density or porosity, (ii)
organic carbon content; (iii) and soil texturesand, silt, and
clay content (Gupta and Larson, 1979; Rawls, 1983; Saxton
et al., 1986; Ritchie et al., 1987; Vereecken et al., 1989; Rawls
and Brakensiek, 1989; Pachepsky et al., 1999; Wsten et al.,
1999; Weynants et al., 2009). Pachepsky et al. (1998) found that

www.VadoseZoneJournal.org

66Modeling the Soil Hydraulic

Conductivity Function

The Saturated Hydraulic Conductivity

For viscous and relatively slow flows characteristic of porous media


(often with Reynolds number Re < 1), the water flux, q, according
to Darcys law is proportional to the gradient of total hydraulic
head, DF/L, expressed as:
q =-K s

DF

L

[2.1]

where F = P/rg + z represents the total hydraulic head driving the


flow across distance L, and Ks [L T1] is a proportionality constant
known as the saturated hydraulic conductivity, also related to the
permeability k [L2] of the porous medium (a property of pore space
geometry) and m [M L1 T1], the dynamic viscosity of the fluid.
The relationship between the saturated hydraulic conductivity Ks
and the permeability, k, is given by:
rg
K s = k
m

[2.2]

where r [M L3] is the fluid density, and g [L T2] is acceleration


due to gravity. In groundwater applications, the permeability is
often expressed in units of Darcy (D), leading to the following
unit equivalence (for water): 1 D = 1 1012 m2 ? 105 m s1.
Often, the permeability is expressed as a function of a pore space
characteristic length, l, (Scheidegger, 1960):

p. 7 of 20

k=c

l

f (n)

[2.3]

where f(n) is related to porosity, pore shape factor, and tortuosity


of the porous medium of interest, and c a dimensionless constant.
Several characteristic lengths were deduced from different pore
related attributes (Bernab and Bruderer, 1998), including the
hydraulic radius (Kozeny, 1927; Scheidegger, 1960; Johnson and
Schwartz, 1989), critical pore radius (Katz and Thompson, 1986;
Arns et al., 2005), air entry matric potential (Assouline and Or,
2008), and mean grain diameter, d (Revil and Cathles, 1999).
An interesting expression was proposed by Revil and Cathles
(1999) in which a cementation exponent, m that varies with pore
space characteristics was introduced:
k=

d 2 n3 m
24

[2.4]

This and similar expressions establish links between electrical conductivity measurements and permeability. Based on the pioneering
work of Archie (1942), the values of the cementation exponent
were near 2.0 for consolidated sandstone and 1.3 for loose sand.
Lesmes and Friedman (2005) presented a table for coarse media
reporting a similar range (m = 1.3 to 2.0).
Considering the simplest capillary model, that is, a bundle of parallel cylindrical capillaries of constant radius r, the permeability
according to Eq. [2.3] is expressed as a function of porosity, n,
according to:
2
n r
k =
2 2

[2.5]

A widely used generalization to porous media is known as the


Kozeny equation (Scheidegger, 1960):
k=

c n3
t S2

[2.6]

where t is the soil tortuosity, S is the surface exposed to the fluid


per unit volume, and c is a constant related to the pore shape that
varies between 1/2 for a circle to 2/3 for a strip. A modification
of Eq. [2.6] leads to the well-known KozenyCarman expression
(Carman, 1937):
k=

1
n3

5So2 (1- n)2

[2.7]

where S o is surface exposed to fluid per unit volume of solid


material.

www.VadoseZoneJournal.org

The concept of permeability of a porous medium can be considered


also in terms of a flow around submerged bodies, therefore defining
related drag or friction factors. The friction factor around a submerged sphere is inversely proportional to the Reynolds number
of the flow, Re (Ergun, 1952). Considering steady flow through
porous packs of granular media characterized by Re = {(dVs r)/
[m(1 n)]}, with Vs being the superficial velocity, and the friction
factor for flow in tubes, a KozenyCarman type of relationship
is obtained:
mV
n3
DF
= 180 2s

L
d (1- n)2

[2.8]

Often Kozeny-type expressions tend to overestimate k at low porosities (Bernab et al., 1982); hence, Bourbi et al. (1987) and Zhu et
al. (2007) suggested variable power of n in Eq. [2.6] and [2.7], with
a value of 3 for large porosity, and a value of 7 to 9 for low porosities.
The corrections above are linked with the concept of critical or
connected porosity (nc) that was introduced to improve the performances of Kozeny-type equations (Mavko and Nur, 1997; Quispe
et al., 2005):

k=

c (n - nc )3

t S2

[2.9]

Marshall (1958) derived an expression for soil permeability that


relies on HagenPoiseuilles equation and accounts for the distribution of pore radii in an isotropic material. This equation
resembles the Kozeny equation in the particular case of a porous
medium with an effective pore radius.
Expressions for soil saturated hydraulic conductivity, Ks, were
developed based on the KozenyCarman model, involving a power
relationship between Ks and effective soil porosity (Green et al.,
2003; Ahuja et al., 1984, 1989a,b):
Ks = aneb

[2.10]

where a and b are empirical coefficients and ne is the effective


porosity, calculated as the saturated water content q s minus the
water content at a matric tension of 33 kPa, a value usually used
to characterize field capacity. Rawls et al. (1998) found that the
power value, b, can vary between 1.59 and 3.98, and that it could
be related to the pore size distribution index, l, of the Brooks and
Corey (1964) model for the WRC (Eq. [1.3]), according to the
relationship b = 3 l. Consequently, Eq. [2.10] becomes:
Ks = ane (3 l)

[2.11]

p. 8 of 20

Assouline and Or (2008) proposed to consider the air entry matric


potential, y c, as a natural characteristic length for the porous
medium, which integrates aspects of pore size and connectivity
in a simple fashion (Katz and Thompson, 1986; Arns et al., 2005).
They followed the analysis of Sisavath et al. (2000) by capitalizing
on the Aissen formula (Aissen, 1951) that represents complex pore
shapes in a relatively simple way (see Fig. 5) to derive the following
general expression for k:
k=

c
1
1
t (n2 +1) y2c

[2.12]

To account for effects of a broad pore size distribution on estimates


of k, an expression was proposed by Wyllie and Sprangler (1952):
c (n -q r ) l
k=

t yc 2 l + 2

[2.13]

where q r is the residual water content and l is the pore size


distribution index characterizing the WRC in the Brooks and
Corey model (Eq. [1.3]).

The Unsaturated Hydraulic


Conductivity Function

Three major approaches can be identified in models developed to


predict the HCF, namely, the empirical, the macroscopic and the
statistical approaches. A comprehensive review of these different
approaches can be found in Mualem (1986) and Brutsaert (2005).
The empirical approach consists of best-fitting simple mathematical
formulas to available hydraulic conductivity data and functions
that are integrable and could yield analytical solutions for problems involving Buckingham-Darcy fluxes. These functions can be
represented by Gardners type HCFs (Gardner, 1958):
K r ( y) =

1
yp

+1

K r (y) = e(-a G y)

[2.14]

[2.15]

where a G, c, and p are fitting parameters. A power function type


was suggested by Wind (1955) and adapted to the air entry value
y c by Brooks and Corey (1964), leading to an equation of Kr(y)
similar to the power type form of Eq. [1.3]. The major limitation of
this approach is the specificity of equations to a range of potentials
(Or et al., 2000b) or to specific soil type. The lack of systematic link
with WRC limits incorporation of hysteresis into these functions
(especially in their K(y) form). Note that the WRC expression
in Eq. [1.2] was derived by Russo (1988) starting from Gardners
(1958) Eq. [2.15] to establish links between parameters of this

www.VadoseZoneJournal.org

Fig. 5. A vertical cross section through a pack of modeling clay spheres


(5 mm in diameter) arranged in octahedral packing with an inscribed
and circumscribed circles used in the Aissen formulation (Eq. [2.12])
are depicted for a typical pore in the cross section (see Assouline and
Or, 2008).

particular HCF (the parameters are relatively easy to measure)


and a compatible WRC.
Approaches similar to the statistical representation of soil PSD and
the WRC (as discussed above) were also applied to estimation of
HCF (Kosugi, 1999; Arya et al., 1999; Kosugi et al., 2002; Blank
et al., 2008; Arya and Heitman, 2010; Hwang and Hong, 2006;
Nasta et al., 2013a). Some of the simplest models for HCF yield
expressions for Kr(Se) relationship in the following general form:
Kr(Se) = Sem

[2.16]

Averjanov (1950) suggested that m = 3.5, very close to the theoretical value of m = 3.0 derived by Irmay (1954). Th is approach
considers basically uniform and parallel capillaries and thus
neglects the effect of pore-size distribution on hydraulic conductivity (Childs and Collis-George, 1950). Bresler et al. (1978)
found that m = 7.2 represent relatively well measured data for
12 soils reported by Mualem (1974a). Using experimental WRC
data of 50 soils, Mualem (1978) found that m varies between 2.5
and 24.5. A good correlation was found between the m values of
the 50 soils and the corresponding energy per unit volume of soil,
w, required to drain the saturated soil down to the wilting point.
Consequently, the derived empirical linear formula relating m to
w incorporates to some extent the effect of pore size distribution
in the macroscopic approach. Brooks and Corey (1964) suggested
accounting for the effect of pore size distribution in the power
value of Eq. [2.16] using l of Eq. [1.3] by means of the relationship
m = (3 + 2/l). Considering the data of Mualem (1978), Brutsaert
(2000) found that m = (2.18 + 2.51/l) produced a good agreement.
Assouline (2005) has proposed that:
m = a[lb + l(b 1)]

[2.17]

where the constants a and b can be considered as empirical constants. For the data set used in Assouline (2005), a = 1.40 and b
= 0.717.
p. 9 of 20

The foundations for the statistical approach that uses WRC information to predict the HCF were apparently established by Purcell
(1949), and developed through significant contributions by Childs
and Collis-George (1950), Fatt and Dykstra (1951), Burdine
(1953), Wyllie and Gardner (1958), Brutsaert (1967), Mualem
(1976), and Mualem and Dagan (1978). The outcome of the statistical approach depends on the conceptual-geometrical model
of the porous medium. For example, the power function in Eq.
[2.16] stems from considering the porous medium as an assembly
of parallel capillaries tubes having a circular cross section and a
prescribed distribution of radii.
The statistical approach assumes that the HagenPoiseuille equation is valid at the single pore level. A general form of the relative
hydraulic conductivity, Kr(Se), can be derived as a function of the
effective saturation, Se (Mualem and Dagan, 1978; Mualem, 1986;
Kosugi et al., 2002):

Se

0
l
K r ( Se ) = Se
1

dSe

y(Se )

d
S
e
n

y (Se )

[2.18]

The models of Burdine (1953) and Mualem (1976) lead to closedform analytical expressions when some models of the WRC
presented in the first section are used to express dSe/y in Eq. [2.18].
The expression resulting from the application of Mualems model
to the WRC expression of Brooks and Corey (1964) [Eq. 1.3] is:
Kr(Se) = Se(2 + 2.5l)/l

[2.19]

When Mualems model is applied to the van Genuchten (1980)


expression for Se(y) (Eq. [1.5]) imposing m = (1 1/n) with n >
1, Kr(Se) becomes:

Applying Mualems model to Assouline et al. (1998) expression of


the WRC (Eq. [1.10]) leads to:
2
-1/m 1 1
1
1
x
g , ux- e-ux +

m m
y L
y
K r (Se ) = Se

1
-1/m 1
1

x
G +

m m y L

[2.21]

where g(b,u) and G(u) are the incomplete and the complete
Gamma functions, respectively, and u = (|y|1 |y L|1) m . A
detailed description of its derivation and performances can be
found in Assouline and Tartakovsky (2001).
It is interesting to notice that Brutsaert (2000) suggested an
expression that resembles the one in Eq. [2.19]:

The factor Sel in Eq. [2.18] is a correction that accounts for the
tortuosity of the flow path (departure from straight capillaries). For
n = 2, z = 1, and l = 2, one can recognize the model of Burdine
(1953), while n = 1, z = 2 and l = 0.5 correspond to the model of
Mualem (1976). In fact, the value of the power l depends on the
specific soil-fluid properties and varies considerably for different
soils. The value of l = 0.5 suggested by Mualem (1976) was the
optimal value with regards to measured RHC data of the 45 soils
that were considered in that study.

m
K r (Se ) = Se0.5 1- 1- Se(1/ m)

among others). As a result, Kr(Se) expressed by means of Eq. [2.20]


could exhibit a high sensitivity just below Se = 1.0, which might
cause numerical instabilities in simulations of near-saturated infiltration, especially for 1 < n < 1.3 (Durner, 1994; Schaap and van
Genuchten, 2006). Forcing Kr(Se) to remain constant for a small
range of y values (few centimeters) below y = 0 could be enough
to obtain stable numerical simulations (Vogel et al., 2000; Schaap
and van Genuchten, 2006).

[2.20]

Kr(Se) = Se(2.5 + 2l)/l

[2.22]

In terms of Mualems model, the expression of Brutsaert (2000) is:

Se dSe 2.5

0
y
-0.5
K r (Se ) = Se

1 dSe
0 y

[2.23]

meaning that l = (-0.5). This is in agreement with the results of


Leij et al. (1997) who have obtained a mean value of l = (-0.72)
when 401 pairs of water retention and unsaturated hydraulic conductivity data were considered. The model in Eq. [2.17], with n =
1, z = 2 and l = 0.5, that corresponds to Mualems model (1976),
performs the best when compared to the models of Averjanov
(1950), Wyllie and Gardner (1958) and Millington and Quirk
(1961). However, some remarks can be made: (i) the value z = 2
stems from the assumption that the pore configuration is conceptually described by a pair of capillary elements whose lengths are
proportional to their radii, which is a very strong constraint and a
restrictive assumption considering pore configuration in soils; (ii)
the assumed constant value of l = 0.5 is in disagreement with the
fact that the tortuosity factor strongly depends on soil structure
and texture in addition to its dependence on soil moisture content
(Pachepsky, 1990).

The expression in Eq. [1.5] lacks second-order continuity at saturation when 1 < n < 2 (Luckner et al., 1989; Vogel et al., 2000;

www.VadoseZoneJournal.org

p. 10 of 20

Assouline (2001, 2004a) has suggested a different model for Kr(Se),


which requires less restrictive assumptions and improves the prediction of Kr(q) when compared with measured data:
S e dS e

0 y
K r ( Se ) =
1 dS e
0
y

[2.24]

Following the notation in Eq. [2.18], this model corresponds to n


= 1, z = h and l = 0. The quadratures in Eq. [2.24] can be evaluated
numerically, using WRC data. For some analytical y(q) functions,
closed-form analytical expressions for Kr(Se) in Eq. [2.24] do exist.
The resulting expression of the HCF when Brooks and Corey
(1964) model (Eq. [1.3]) is used in Eq. [2.24] is:

K r (Se ) = Se(h+hl )/l

A dependency between h and s was also observed by Nasta et al.


(2013b) for optimized h values from 20 soil types showing a linear
correlation between h and s.
The performance of the statistical approach is illustrated in Fig.
6, depicting the fit of Eq. [2.19] and [2.27] to experimental data
reported by Mualem (1974a) for the sand and the loam soils that
were already presented in Fig. 2. For the sand, Eq. [2.27] improves
the agreement between the estimated Kr(Se) and the measured
data, while it seems that this is the opposite for the loam soil. It
is interesting to note that at the WRC level, Eq. [1.3] performed
better than Eq. [1.10] for the sand than for the loam. These results
emphasize the complexity of the relationship between the WRC
and the HCF, and the limits of the statistical approach inherent
to the related conceptualization of the pore space.

[2.25]

The resulting expression of the HCF when van Genuchten


(1980) expression (Eq. [1.5] with m = 1 1/n) is used in Eq.
(2.24) is:

m
K r (Se ) = 1- 1- Se(1/ m)

[2.26]

The resulting expression of the HCF when Assouline et al.


(1998) model (Eq. [1.10]) is used in Eq. [2.24] is:
h
-1/m 1 1
1
1
x
g , xa - e-xa +

m m
y L
y
K r (Se ) =

1
-1/m 1

+ 1
x
G

m m y L

[2.27]

The power h in Eq. [2.24] depends on the soil structure and


texture, which both shape the WRC. Therefore, the power
h was found to be highly correlated with l of Brooks and
Corey (1964) (Eq. [1.3]), and to the coefficient of variation,
e, of the WRC expressed according to Eq. [1.10] (Assouline,
2005):
h = 1.40l0.717 (r2 = 0.84)

[2.28]

h = 1.10e-0.624 (r2 = 0.88)

[2.29]

Based on a restricted data set of 10 soils, from sand to sandy


clay loam, it can be shown that h is correlated with s of the
model of Kosugi (1994) (Eq. [1.6]:
h = 0.99s-0.82 (r2 = 0.92)

www.VadoseZoneJournal.org

[2.30]

Fig. 6. The hydraulic conductivity function of a sand and a loam soil expressed in
terms of Se (upper plots) or y (lower plots). The experimental data (black dots)
were reported by Mualem (1974a). The curves correspond to Eq. [2.19] (dashed
line) and [2.27] (solid line).

p. 11 of 20

The PTF approach was also applied to correlate between soil structure and texture to soil hydraulic conductivity, both saturated and
unsaturated (Campbell, 1985; Vereecken et al., 1990; Cosby et al.,
1984; Brakensiek et al., 1984; Wsten, 1997; Wsten et al., 1999;
Saxton et al., 1986; Rawls et al., 1998). Wagner et al. (2001) compared the performances of some of these expressions and concluded
that the PTF proposed by Wsten (1997) provided the best estimate of K(y). However, their conclusion was that the discrepancy
between the estimated soil hydraulic conductivity and experimental data indicated that the application of PTFs must be performed
with great caution.

Hydraulic Conductivity Function Models Based


on Corner and Film Flows

The detailed picture of liquid-vapor interface configuration in


angular pores and in liquid fi lms (discussed above) and assuming
these liquidvapor interfaces remain relatively stable under slow
laminar flow conditions enable introduction of hydrodynamic
considerations for flow in unsaturated pore spaces. Studies by
Ransohoff and Radke (1988), Blunt and Scher (1995), and others,
have used the equilibrium interfaces as boundaries for flowing
liquid. Th is is a reasonable assumption considering the large viscous forces required for breakup or deformation of liquidvapor
interfaces (Romero and Yost, 1996), especially at low matric potentials. Tuller and Or (2001) considered viscous flow in a direction
perpendicular to the unit cell cross section (similar to derivations
for the BCC model invoked by Fatt and Dykstra, 1951; Millington
and Quirk, 1961; and Mualem, 1976), and ignored network effects
to obtain estimates of saturated and unsaturated hydraulic conductivity for a unit pore. For different fi lling stages of a unit pore
(determined by matric potential and geometrical attributes) one
may identify three primary flow regimes to derive HCF: (i) flow
in completely fi lled pore spaces, (ii) corner flow in partially fi lled
pores and grooves, and (iii) fi lm flow on solid surfaces. Specific
solutions of the NavierStokes equations for the geometry and
interfacial conditions could be averaged to yield average cross-sectional velocities for each of these regimes. Two key assumptions are
(i) that equilibrium liquidvapor interfaces remain stable and (ii)
that flow pathways are parallel. Liquidvapor configurations for
different matric potentials are estimated and statistically upscaled
to obtain unsaturated hydraulic conductivity from velocity expressions weighted by the appropriate liquid-occupied cross-sectional
areas. An example of such calculations for Hygiene sandstone is
shown in Fig. 7 for the same unit cell presented in Fig. 4.
Application of the procedure for a wide range of soils by Tuller and
Or (2001) highlighted a key feature of this approach, namely, the
consideration of water flow in thin fi lms that is likely to become
important for dry range of WRC and for fine-textured soils with
considerable internal surface area. An example of the HCF tail is
shown for Gilat loam (Fig. 8). It is considered an important transport pathway for arid regions and biological activity in dry soils.

www.VadoseZoneJournal.org

Fig. 7. Measurements (Mualem, 1976) and estimated hydraulic


conductivity function for Hygiene sandstone deduced from a single
unit cell with triangular central pore (see Tuller and Or, 2001).

Fig. 8. Measured (Mualem, 1976) and calculated hydraulic


conductivity function for Gilat loam using the van Genuchten
Mualem model (Eq. [2.20]) and using the corner-film flows model of
Tuller and Or (2001).

More recently, Peters and Durner (2008b) and Lebeau and Konrad
(2010) devised more practical approaches for incorporation of the
water flow in fi lms using simplified representation of the WRC
required as input for HCF estimation.

p. 12 of 20

6 Special Considerations
Parameter Equivalence between
Different WRC and HCF Models

Table 1. Parameters equivalence for some WRC models based on the studies of van
Genuchten (1980), Morel-Seytoux et al. (1996) and Assouline (2005).
Models

Brooks and Corey (1964)


Model
parameters (Eq. [1.3])

van Genuchten (1980)


(Eq. [1.5])

The most widely used empirical models for WRC are


n
l
a
yc
the Brooks and Corey (1964) (Eq. [1.3]) (favored for
a = y c1
analytical solutions of flow problems) and the van van Genuchten (1980) a
(Eq. [1.5])
n
Genuchten (1980) (Eq. [1.5]) model (often favored
n = l+1
for numerical schemes). Ability to compare parameters and results based on use of these two models Assouline et al. (1998) m
m = 1.49 l1.24
m = 0.51n1.1
1 m
x = 1.21
motivated several studies that proposed correspon- (Eq. [1.10])
a
dence between the model parameters (van Genuchten,
x
x = 1.57 (y cl) -1.21
1980; Morel-Seytoux et al., 1996; Lenhard et al., 1989;
Stankovich and Lockington, 1995; Sommer and
Stckle, 2010). For example, assuming m = 1 1/n
A porous medium can be represented by a bivariate statistical distriin Eq. [1.5], the comparison between Eq. [1.3] and [1.5] for large
bution function for the capillary head for water filling or emptying,
-1
y values leads to a = y c and n = l + 1 (van Genuchten, 1980).
f(y w, y d), where y w and y d are the capillary head values of fi llMorel-Seytoux et al. (1996) suggested parameter equivalence for
ing or emptying, respectively, of a representative volume of the
Eq. [1.3] and [1.5] based on the preservation of the capillary drive
porous medium. The resulting hysteretic WRC is composed of
and the asymptotic behavior of y(q) at low q values. Under such
main drying and wetting curves that define the limits of an infinite
-1
-1
conditions, a is proportional to y c with the product (ay c )
number of interior scanning curves of several orders. The main and
for a given soil being dependent on l. Assouline (2005) suggested
primary scanning loops are illustrated in Fig. 9, using data reported
equivalences between the parameters in Eq. [1.10] and those in
by Huang et al. (2005). The different approaches that have been
Eq. [1.3] and [1.5]. Table 1 provides some of the main parameter
applied to the modeling of hysteresis in WRC include empirical
equivalences that were proposed.
expressions, invasionpercolation theory, and independent and
dependent domains theory, as reviewed by Pham and Barbour
Similarly, parameters equivalences have been proposed in terms of
(2005) and Mualem and Beriozkin (2009).
the HCF models. Rucker et al. (2005) and Ghezzehei et al. (2007)
introduced a correspondence between the parameter a G of the
Several empirical models have been proposed (Hanks et al., 1969;
model of Gardner (1958) (Eq. [2.15]) and the parameters a and
Haverkamp et al., 1977; Scott et al., 1983; Kool and Parker, 1987;
n of the Mualemvan Genuchten model (Eq. [2.20]). A concise
Hogarth et al., 1988; Nimmo, 1992; Huang et al., 2005). Some
relationship resulted from the study of Ghezzehei et al. (2007) in
of these models are based on a scaling approach assuming shape
the form of a G = 1.3an.
similarity between main and primary curves. If such similarity
The suggested equivalences enable transfer of parameters across
models. It increases also the opportunities for applications, reuse
of previously reported data, and comparison of approaches and
results from previous studies.

Hysteresis in the Water Retention Curve

Pore space morphology and interconnectedness coupled with


boundary conditions (rate of drainage or imbibition) may result
in different sequences of pore fi lling or emptying and affect the
resulting phase distribution (Haines, 1930; Poulovassilis, 1962;
Vachaud and Thony, 1971; Smiles et al., 1971). Consequently, for
the same porous medium, different WRCs characterize drying or
wetting processes and even for different rates of wetting and drying,
as Davidson et al. (1966) observed: The size of the pressure increment or redistribution rate will control not only the number of
pore sequences which fi ll and conduct water, but will result in a
different water content distribution within the pore sequences.

www.VadoseZoneJournal.org

Fig. 9. Hysteresis in the water retention curve: (a) drying and wetting
events, departing from the main wetting curve (black triangles). (b)
Wetting and drying events, departing from the main drying curve
(black dots). Data are from Huang et al. (2005).

p. 13 of 20

exists between wetting curves, it does not exist between drying


ones, thus inducing inconsistencies when applied (Mualem and
Beriozkin, 2009). These models failed to reproduce accurately
the experimental data, especially in the cases where a large portion of the hysteretic domain is within the range of the air entry
matric potential.
The domains-based approach pioneered by Nel (1942, 1943)
and Everett (1955) offered a quantitative basis for description of
hysteresis. First, the independent domain approach assumes that
domains in the pore space (pore clusters) are filled or emptied independently (Poulovassilis, 1962; Parlange, 1976). Topp and Miller
(1966) and Everett (1967) introduced interconnections between
the independent elements and opened the way to the dependent
domain approach (Topp, 1971; Mualem, 1974b; Mualem and
Dagan, 1975; Mualem and Miller, 1979; Mualem, 1984).
Comparison between the performances of the different approaches
and models can be found in Viaene et al. (1994), Pham and Barbour
(2005), and Mualem and Beriozkin (2009). The test performed by
Viaene et al. (1994) using 10 soils indicated that models calibrated
by both branches of the main hysteresis loop performed better than
those calibrated by a single main branch. Mualem and Beriozkin
(2009) demonstrated that primary wetting curves of all orders
could be predicted based on scaling approaches, while accurate
estimates of the primary drying curves of all orders has to rely on
the dependent domain theory. The hysteresis in the WRC reflects
also in terms of the HCF (Mualem, 1986). The stronger expression
of the hysteresis in the HCF is when it is expressed as K(y). It is
much less noticeable when it is expressed as K(q).

Effects of Soil Structural Changes


on the WRC and HCF

The review above emphasizes the crucial role of soil structure


in shaping the hydraulic functions. Soil structure can undergo
dynamic changes at different time scales both in natural environments and in agricultural fields due to, among other factors,
sedimentation, illuviation, consolidation, cycles of wetting and
drying or freezing and thawing, tillage, swelling, and mechanical compaction. These changes will induce variations in the soil
hydraulic properties, thus affecting many aspects of the soilwater
plantatmosphere system. Appropriate quantitative tools with
predictive ability are therefore essential to account for these effects
on agricultural, hydrological and environmental aspects.
The main approach applied relied on empirical or semi-empirical
models considering that changes in soil structure could be well
represented by changes in porosity or bulk density (Assouline et
al., 1997; Ahuja et al., 1998; Stange and Horn, 2005; Assouline,
2006a,b). This approach was the premise of a conceptual model
aiming to derive the hydraulic properties of a seal layer developing on the soil surface following the impact of high-kinetic energy
raindrops (Mualem and Assouline, 1989; Assouline, 2004b). Or et

www.VadoseZoneJournal.org

al. (2000a) used the FokkerPlanck equation to model soil deformation and coupled it to the probabilistic nature of the PSD to
develop a stochastic model for post-tillage dynamic changes in soil
structure and consequently on soil hydraulic properties.
A large body of studies has investigated the relationships between Ks
and r b (Carman, 1937; Laliberte et al., 1966; Reicovsky et al., 1980;
Young and Voorhees, 1982; Mualem and Assouline, 1989; Or et al.,
2000a,b; Or and Ghezzehei, 2002; Green et al., 2003; Assouline,
2006b; Berli et al., 2008). A concise review of the different
approaches developed to relate the permeability of porous media to
porosity can be found in Assouline and Or (2008). An interesting
result of that study was that the characteristic length inherent to the
relationship between soil structure and permeability could be well
represented by the air entry matric potential, y c.
Many soil types, including some of the most fertile soils, contain
significant amounts of active clay minerals that exhibit shrink
swell behavior in response to changes in soil water content and
chemical composition of the soil solution (Warkentin et al., 1957;
Giraldez and Sposito, 1983; Giraldez et al., 1983; Quirk, 1986;
Revil and Cathles, 1999). The theory for crystalline and osmotic
swelling of clay minerals at the scale of individual clay lamellae
is well established (Derjaguin and Landau, 1941; Quirk, 1986;
Derjaguin et al., 1987; Israelachvili, 1991; Ohshima, 1995;
McBride and Baveye, 2002). However, translation of this behavior
to the prediction of hydraulic properties of swelling soils remains
limited. Changes in pore space attributes induced by the shrink
swell behavior of clay minerals is still presenting a challenge to
predictive modeling of hydraulic properties of clayey soils and of
flow and transport processes in soils. Tuller and Or (2003) capitalized on modeling liquid distribution in rigid angular pore spaces
(Tuller and Or, 2001) to provide the framework for pore-scale
geometrical changes in swelling porous media. Other modeling
and measurement frameworks have been proposed for quantifying volume changes and soil hydraulic properties (Chertkov
and Ravina, 2002; Chertkov, 2012; Garnier et al., 1997; White
et al., 2003; among others). A key feature in these models is the
need to combine physicochemical processes with mechanical and
hydrodynamic considerations towards predicting mechanical state,
volume changes, and constitutive hydraulic relationships for swelling porous media.

The Hydraulic Properties of Coarse Porous Media

Evidence suggests that soil water storage and water flow characteristics in soils containing appreciable amounts of stones or gravel
layers is significantly modified (Fis et al., 2002; Tokunaga et
al., 2002; Unger, 1971; Verbist et al., 2013). For mixed stones in
a background of a finer soil matrix, the effect on the hydraulic
behavior may involve hydraulic decoupling of the embedded stones
due to large pore size contrast, thereby reducing effective water
availability for plants (Cousin et al., 2003) or reducing average
hydraulic conductivity due to lower-conducting obstacles (Fis et
p. 14 of 20

al., 2002; Cousin et al., 2003; Coppola et al., 2013). Tokunaga et


al. (2003) devised methods and gravitational corrections to determining WRC for gravel layers found at the Hanford site. Chapuis
(2004) discusses methods for estimation of hydraulic conductivity
of gravels with emphasis on friction based empirical functions.
Dexter (1993) conducted a series of gravity-driven flows through
beds of large particles and was able to demonstrate an increase in
flux focusing with thickness of the bed (and increased in the spatial
variance of outflow). The explanation was that rivulets of water
take a random walk and progressively combine with depth (similar to a river network on surfaces), and the number of individual
rivulets decreases with depth in the gravel bed, pointing to the
enhanced potential for preferential flows. The hydraulic properties
of coarse materials may affect evaporation dynamics from soils,
and a systematic study by Unger (1971) demonstrated significant
changes in evaporative losses and water storage due to gravel layers
(similar to effects of artificial mulch).

66Summary and Outlook

Advances in observation and measurement methods provided new


insights and enhanced understanding of key flow and transport
processes ruled by soil WRC and HCF. The translation of these
insights into a broader and more routine parameterization of the
hydraulic functions characterizing natural soils and porous media
remains hindered by limited in situ measurements methods and
the omnipresent soil spatial heterogeneity. Yet, mounting pressure
from various modeling communities (hydrology, soil, climate, environmental regulators) makes the judicious selection of models and
parameters for soil WRC and HCF more critical now than before.
Researchers are often caught in the gap between the growing need
for hydrologic parameters and the chronically limited information
regarding various important quantities. It is not surprising that
since the seminal paper of van Genuchten (1980), which popularized the link between WRC with HCF, publications reporting
direct measurements of HCF (or surrogate variable such as the
soil water diffusivity) have dropped dramatically in the past few
decades (only recently, with the renewed interest by the geotechnical engineering community in unsaturated flows that new
measurements are being reported). The reliance on WRC for inference of HCF requires improved conceptual models of capillary and
viscous interactions within realistic pore spaces.
The parametric models for the WRC reviewed in this study provide
a range of complexity and reflect various underlying assumptions.
Parameter equivalence enables transfer of parameters across models
(at least in an approximate fashion) and enriches the opportunities
for applications and reuse of available data previously reported. The
growing popularity of pedotransfer models reflects the increasing
need for spatial information on hydraulic functions at resolution
that exceed present measurement capabilities. On the other hand,
with the rapid expansion of ecohydrological observatories (Tereno,

www.VadoseZoneJournal.org

CUAHSI) and Earth observing platforms, there is a concurrent


increase in flow of hydrologic information that could be used to
constrain WRC and HCF estimates.
At the conceptual level, new and more complex models for soil pore
spaces have been proposed and tested (Dullien et al., 1977; Tuller
et al., 1999). These models not only expand the range of geometries
considered in representing soil WRC, but also include additional
and underrepresented processes, such as the role of adsorption, disjoining pressures, complex topology, and more, which feed into
the understanding of flow processes, colloidal transport, aquatic
habitats for microbial life, and more.
The key remaining challenges could be categorized into conceptual,
methodological, and scale-related challenges. At the conceptual
level, there is a growing realization that the bundle of cylindrical capillaries must be replaced by more realistic (yet tractable)
description of pore spaces and processesadvances in considering retention and flow in films, in angular pores, and expansion
of universal frameworks (percolation theory) offer a promise. We
also need to expand our efforts judicially using surrogate variables,
such as creating better links between particle and pore sizes, and
expand the physical basis for the popular pedotransfer functions.
The rapid growth in measurement capabilities and deployment of
ecohydrological observatories and networks offer unprecedented
opportunities for extensive data gathering at multiple scales.
Nevertheless, without developing a clear plan and methodologies
for quality assurance of derived parameters, the establishment of
data sharing protocols, and defining structure and hierarchy for
the information derived, the picture may become even more chaotic, and the potential of these advances may not be fully realized.
The opportunity for tailoring parameters and data for different
spatial (and temporal) scales offers abilities to describe unsaturated
flow and transport processes at an ever-increasing resolution and
quality. Nevertheless, modelers and practitioners must become
more aware of the inherent limitations of the governing equations and macroscopic representation of WRC and HCF that are
derived from pore-scale phenomena, yet are applied to very large
scales, often involving far more complex flow processes than postulated in their derivation.

Acknowledgments

The authors thank Wolfgang Durner and an anonymous reviewer for their
constructive comments.

References

Ahuja, L.R., J.W. Naney, and D.R. Nielsen. 1984. Scaling soil-water properties
and infiltration modeling. Soil Sci. Soc. Am. J. 48:970973. doi:10.2136/
sssaj1984.03615995004800050003x
Ahuja, L.R., D.K. Cassel, R.R. Bruce, and B.B. Barnes. 1989a. Evaluation of spatial
distribution of hydraulic conductivity using effective porosity data. Soil Sci.
148:404411. doi:10.1097/00010694-198912000-00002
Ahuja, L.R., D.L. Nofziger, D. Swartzendruber, and J.D. Ross. 1989b. Relationship
between Green and Ampt parameters based on scaling concepts and field-

p. 15 of 20

measured hydraulic data. Water Resour. Res. 25:17661770. doi:10.1029/


WR025i007p01766
Ahuja, L.R., F. Fiedler, G.H. Dunn, J.G. Benjamin, and A. Garrison. 1998.
Changes in soil water retention curves due to tillage and natural reconsolidation. Soil Sci. Soc. Am. J. 62:12281233. doi:10.2136/
sssaj1998.03615995006200050011x
Aissen, M.I. 1951. Estimation and computation of torsional rigidity, Ph.D. diss.
Stanford University, Stanford, CA.
Archie, G.E. 1942. The electrical resistivity log as an aid in determining some
reservoir characteristics. Petrol. Trans. AIME 146:5462.
Arns, C.H., M.A. Knackstedt, and N.S. Martys. 2005. Cross-property correlations
and permeability estimation in sandstone. Phys. Rev. E Stat. Nonlin. Soft
Matter Phys. 72:046304.. doi:10.1103/PhysRevE.72.046304
Assouline, S. 2001. A model of soil relative hydraulic conductivity based on
water retention curve characteristics. Water Resour. Res. 37:265271.
doi:10.1029/2000WR900254
Assouline, S. 2004a. Correction to A model of soil relative hydraulic conductivity based on water retention curve characteristics. Water Resour. Res.
40:W02901. doi:10.1029/2004WR003025.
Assouline, S. 2004b. Rainfall-induced soil surface sealing: A critical review of
observations, conceptual models and solutions. Vadose Zone J. 3:570591.
Assouline, S. 2005. On the relationship between the pore size distribution index and characteristics of the soil hydraulic functions. Water Resour. Res.
41:W07019. doi:10.1029/2004WR003511
Assouline, S. 2006a. Modeling the relationship between soil bulk density
and the water retention curve. Vadose Zone J. 5:554563. doi:10.2136/
vzj2005.0083
Assouline, S. 2006b. Modeling the relationship between soil bulk density and
the hydraulic conductivity function. Vadose Zone J. 5:697705. doi:10.2136/
vzj2005.0084
Assouline, S., and Y. Rouault. 1997. Modeling the relationships between particle
and pore size distributions in multicomponent sphere packs: Application
to the water retention curve. Colloids Surf. A 127:201209. doi:10.1016/
S0927-7757(97)00144-1
Assouline, S., and D. Tartakovsky. 2001. Unsaturated hydraulic conductivity
function based on a fragmentation process. Water Resour. Res. 37:1309
1312. doi:10.1029/2000WR900332
Assouline, S., and D. Or. 2008. Air entry-based characteristic length for estimation of permeability of variably compacted earth materials. Water Resour.
Res. 44:W11403. doi:10.1029/2008WR006937
Assouline, S., J. Tavares-Filho, and D. Tessier. 1997. Effect of compaction on soil
physical and hydraulic properties: Experimental results and modeling. Soil Sci.
Soc. Am. J. 61:390398. doi:10.2136/sssaj1997.03615995006100020005x
Assouline, S., D. Tessier, and A. Bruand. 1998. A conceptual model of the soil water
retention curve. Water Resour. Res. 34:223231. doi:10.1029/97WR03039
Assouline, S., D. Tessier, and A. Bruand. 2000. Correction to A conceptual
model of the soil water retention curve. Water Resour. Res. 36:3769.
doi:10.1029/2000WR900249
Arya, L.M., and J.F. Paris. 1981. A physicoempirical model to predict
the soil moisture characteristic from particle-size distribution and
bulk density data. Soil Sci. Soc. Am. J. 45:10231030. doi:10.2136/
sssaj1981.03615995004500060004x
Arya, L.M., and J.L. Heitman. 2010. Hydraulic conductivity function from water
flow similarity in idealized- and natural-structure pores. Soil Sci. Soc. Am. J.
74:787796. doi:10.2136/sssaj2009.0204
Arya, L.M., F.J. Leij, P. Shouse, and M.Th. van Genuchten. 1999. Relationship
between hydraulic conductivity function and particle-size distribution. Soil
Sci. Soc. Am. J. 63:10631070. doi:10.2136/sssaj1999.6351063x
Averjanov, S.F. 1950. About permeability of subsurface soils in case of incomplete saturation. In: I.M. Roger De Wiest, transl., quoted by P. Ya Palubarinova, 1962, The theory of ground water movement. English Collection, Vol.
7. Princeton Univ. Press, Princeton, NJ. p. 1921.
Bartoli, F., R. Phillipy, M. Doirisse, S. Niquet, and M. Dubuit. 1991. Structure
and self-similarity in silty and sandy soils: The fractal approach. J. Soil Sci.
42:167185. doi:10.1111/j.1365-2389.1991.tb00399.x
Baveye, P., and C.W. Boast. 1998. Concepts of fractals in soils: Demixing
apples and oranges. Soil Sci. Soc. Am. J. 62:14691470. doi:10.2136/
sssaj1998.03615995006200050046x
Bear, J. 1972. Dynamics of fluids in porous media. American Elsevier, New York.
Berli, M., A. Carminati, T.A. Ghezzehei, and D. Or. 2008. Evolution of unsaturated hydraulic conductivity of aggregated soils due to compressive forces.
Water Resour. Res. 44:W00C09. doi:10.1029/2007WR006501
Bernab, Y., and C. Bruderer. 1998. Effect of the variance of pore size distribution on the transport properties of heterogeneous networks. J. Geophys.
Res. 103:513525. doi:10.1029/97JB02486
Bernab, Y., W.F. Brace, and B. Evans. 1982. Permeability, porosity, and pore geometry of hot-pressed calcite. Mech. Mater. 1:173183. doi:10.1016/01676636(82)90010-2
Bird, N.R.A., F. Bartoli, and A.R. Dexter. 1996. Water retention models for fractal soil structures. Eur. J. Soil Sci. 47:16. doi:10.1111/j.1365-2389.1996.
tb01365.x

www.VadoseZoneJournal.org

Bittelli, M., G.S. Campbell, and M. Flury. 1999. Characterization of particlesize distribution in soils with a fragmentation model. Soil Sci. Soc. Am. J.
63:782788. doi:10.2136/sssaj1999.634782x
Blank, L.A., A.G. Hunt, and T.E. Skinner. 2008. A numerical procedure to calculate hydraulic conductivity for an arbitrary pore size distribution. Vadose
Zone J. 7:461472. doi:10.2136/vzj2007.0037
Blunt, M.J., and H. Scher. 1995. Pore-level modeling of wetting. Phys. Rev. E Stat.
Phys. Plasmas Fluids Relat. Interdiscip. Topics 52:63876403. doi:10.1103/
PhysRevE.52.6387
Borkovec, M., Q. Wu, G. Degovics, P. Laggner, and H. Sticher. 1993. Surface area
and size distributions of soil particles. Colloids Surf. A Physicochem. Eng.
Asp. 73:6576. doi:10.1016/0927-7757(93)80007-2
Bouma, J. 1989. Using soil survey data for quantitative land evaluation. Adv.
Soil Sci. 9:177213. doi:10.1007/978-1-4612-3532-3_4
Bourbi, T., O. Coussy, and B. Zinszner. 1987. Acoustics of porous media. Gulf
Publ. Co., Houston.
Brakensiek, D.L., W.J. Rawls, and G.R. Stephenson. 1984. Determining the saturated hydraulic conductivity of a soil containing rock fragments. Soil Sci. Soc.
Am. J. 50:834835. doi:10.2136/sssaj1986.03615995005000030053x
Bresler, E., D. Russo, and R.D. Miller. 1978. Rapid estimate of unsaturated hydraulic conductivity function. Soil Sci. Soc. Am. J. 42:170172. doi:10.2136/
sssaj1978.03615995004200010038x
Brooks, R.H., and A.T. Corey. 1964. Hydraulic properties of porous media. Hydrol. Pap. 3. Colorado State Univ., Fort Collins.
Brutsaert, W. 1966. Probability law for pore-size distributions. Soil Sci. 101:85
92. doi:10.1097/00010694-196602000-00002
Brutsaert, W. 1967. Some methods of calculating unsaturated permeability.
Trans. ASAE 10:400404.
Brutsaert, W. 2000. A concise parameterization of the hydraulic conductivity
of unsaturated soils. Adv. Water Resour. 23:811815. doi:10.1016/S03091708(00)00019-1
Brutsaert, W. 2005. HydrologyAn introduction. Cambridge Univ. Press, Cambridge, UK.
Buchan, G.D. 1989. Applicability of the simple log-normal model to particle
size distribution in soils. Soil Sci. 147:155161. doi:10.1097/00010694198903000-00001
Buckingham, E. 1907. Studies on the movement of soil moisture. Bull. 38. Bureau of Soils, USDA, Washington, DC.
Burdine, N.T. 1953. Relative permeability calculation size distribution data. Pet.
Trans. Am. Inst. Min. Metall. Pet. Eng. 198:7178.
Burrough, P. 1981. Fractal dimension of landscapes and other environmental
data. Nature 294:240242. doi:10.1038/294240a0
Campbell, G.S. 1974. A simple method for determining unsaturated hydraulic conductivity from moisture retention data. Soil Sci. 117:311315.
doi:10.1097/00010694-197406000-00001
Campbell, G.S. 1985. Soil physics with BASIC: Transport models for soil-plant
systems. Elsevier, New York.
Campbell, G.S., and S. Shiozawa. 1992. Prediction of hydraulic properties of
soils using particle size distribution and bulk density data. In: International
Workshop on Indirect Methods for Estimating the Hydraulic Properties of
Unsaturated Soils. University of California Press, Berkeley.
Carman, P.C. 1937. Fluid flow through granular beds. Trans. Inst. Chem. Eng.
(London) 15:150156.
Chapuis, R.P. 2004. Predicting the saturated hydraulic conductivity of sand and
gravel using effective diameter and void ratio. Can. Geotech. J. 41:787795.
doi:10.1139/t04-022
Chan, T.P., and R.S. Govindaraju. 2003. A new model for soil hydraulic properties based on a stochastic conceptualization of porous media. Water Resour.
Res. 39:1195. doi:10.1029/2002WR001954
Chan, T.P., and R.S. Govindaraju. 2004. Estimating soil water retention curve
from particle-size distribution data based on polydisperse sphere systems.
Vadose Zone J. 3:14431454.
Chertkov, V.Y., and I. Ravina. 2002. The effect of interaggregate capillary cracks
on the hydraulic conductivity of swelling clay soils. Water Resour. Res.
37:12451256. doi:10.1029/2000WR900319
Chertkov, V.Y. 2012. Physical modeling of the soil swelling curve vs. the shrinkage
curve. Adv. Water Resour. 44:6684. doi:10.1016/j.advwatres.2012.05.003
Childs, E.C., and G.N. Collis-George. 1950. The permeability of porous materials.
R. Soc. London. Proc. A 201:392405.
Chirico, G.B., H. Medina, and N. Romano. 2007. Uncertainty in predicting soil
hydraulic properties at the hillslope scale with indirect methods. J. Hydrol.
334:405422. doi:10.1016/j.jhydrol.2006.10.024
Clapp, R.B., and G.M. Hornberger. 1978. Empirical equations for some soil
hydraulic properties. Water Resour. Res. 14:601604. doi:10.1029/
WR014i004p00601
Coppola, A., G. Dragonetti, A. Comegna, N. Lamaddalena, B. Caushi, M.A. Haikal, and A. Basile. 2013. Measuring and modeling water content in stony
soils. Soil Tillage Res. 128:922. doi:10.1016/j.still.2012.10.006
Cornelis, W.M., M. Khlosi, R. Hartmann, M. Van Meirvenne, and B. De Vos.
2005. Comparison of unimodal expressions for the soil-water retention
curve. Soil Sci. Soc. Am. J. 69:19021911. doi:10.2136/sssaj2004.0238

p. 16 of 20

Cosby, B.J., G.M. Hornberger, R.B. Clapp, and T.R. Ginn. 1984. A statistical
exploration of the relationships of soil moisture characteristics to the
physical properties of soils. Water Resour. Res. 20:682690. doi:10.1029/
WR020i006p00682
Cousin, I., B. Nicoullaud, and C. Coutadeur. 2003. Influence of rock fragments
on the water retention and water percolation in a calcareous soil. Catena
53:97114. doi:10.1016/S0341-8162(03)00037-7
Crawford, J.W., N. Matsui, and I.M. Young. 1995. The relation between the
moisture-release curve and the structure of soil. Eur. J. Soil Sci. 46:369375.
doi:10.1111/j.1365-2389.1995.tb01333.x
Crisp, D.J., and R. Williams. 1971. Direct measurement of pore-size distribution on artificial and natural deposits. Mar. Biol. 10:214226. doi:10.1007/
BF00352810
Darcy, H. 1856. Dtermination des lois dcoulement de leau travers le
sable In: Les Fontaines Publiques de la Ville de Dijon. Victor Dalmont, Paris.
p. 590594.
Davidson, J.M., D.R. Nielsen, and J.W. Biggar. 1966. The dependency of soil water uptake and release upon the applied pressure increment. Soil Sci. Soc.
Am. Proc. 30:298304. doi:10.2136/sssaj1966.03615995003000030005x
Derjaguin, B.V., and L. Landau. 1941. Acta Physicochim. URSS 14:633662.
Derjaguin, B.V., N.V. Churaev, and V.M. Muller. 1987. Surface forces. Consultants Bureau, New York.
Dexter, A.R. 1993. Heterogeneity of unsaturated, gravitational flow of water through beds of large particles. Water Resour. Res. 29:18591862.
doi:10.1029/93WR00772
Diamantopoulos, E., and W. Durner. 2012. Dynamic nonequilibrium of water flow in porous media: A review. Vadose Zone J. 11(3). doi:10.2136/
vzj2011.0197
Dullien, F.A.L., F.S.Y. Lai, and I.F. Macdonald. 1986. Hydraulic continuity of residual wetting phase in porous media. J. Colloid Interface Sci. 109:201218.
doi:10.1016/0021-9797(86)90295-X
Dullien, F.A.L., M.S. El-Sayed, and V.K. Batra. 1977. Rate of capillary rise in porous media with nonuniform pores. J. Colloid Interface Sci. 60:497506.
doi:10.1016/0021-9797(77)90314-9
Durner, W. 1994. Hydraulic conductivity estimation for soils with heterogeneous
pore structure. Water Resour. Res. 30:211223. doi:10.1029/93WR02676
Ergun, S. 1952. Fluid flow through packed columns. Chem. Eng. Prog. 48:8994.
Everett, D.H. 1955. A general approach to hysteresis. 4. An alternative formulation of the domain model. Trans. Faraday Soc. 51:15511557. doi:10.1039/
tf9555101551
Everett, D.H. 1967. Adsorption hysteresis. In: E.A. Flood, editor, The solidgas
interface. Vol. 2. Marcel Dekker, New York. p. 10551113.
Fatt, I., and H. Dykstra. 1951. Relative permeability studies. Pet. Trans. Am. Inst.
Min. Metall. Pet. Eng. 192:249255.
Fis, J.C., N. De Louvigny, and A. Chanzy. 2002. The role of stones in soil water retention. Eur. J. Soil Sci. 53:95104. doi:10.1046/j.1365-2389.2002.00431.x
Fredlund, M.D., D.G. Fredlund, and G.W. Wilson. 2000. An equation to represent grain-size distribution. Can. Geotech. J. 37:817827. doi:10.1139/t00015
Fredlund, M.D., G.W. Wilson, and D.G. Fredlund. 2002. Use of the grain-size distribution for estimation of the soil-water characteristic curve. Can. Geotech.
J. 39:11031117. doi:10.1139/t02-049
Gardner, W.R. 1958. Some steady state solutions of the unsaturated moisture
flow equation with application to evaporation from a water table. Soil Sci.
85:228232. doi:10.1097/00010694-195804000-00006
Gardner, W.R., and F.J. Miklich. 1962. Unsaturated conductivity and diffusivity measurements by a constant flux method. Soil Sci. 93:271274.
doi:10.1097/00010694-196204000-00008
Garnier, P., M. Rieu, P. Boivin, M. Vauclin, and P. Beveye. 1997. Determining the hydraulic properties of a swelling soil from a transient evaporation experiment. Soil Sci. Soc. Am. J. 61:15551563. doi:10.2136/
sssaj1997.03615995006100060003x
Ghezzehei, T.A., T.J. Kneafsey, and G.W. Su. 2007. Correspondence of the Gardner and van GenuchtenMualem relative permeability function parameters. Water Resour. Res. 43:W10417. doi:10.1029/2006WR005339
Giraldez, J.V., and G. Sposito. 1983. A general soil volume change equation.
II. Effect of load pressure. Soil Sci. Soc. Am. J. 47:422425. doi:10.2136/
sssaj1983.03615995004700030006x
Giraldez, J.V., G. Sposito, and C. Delgado. 1983. A general soil volume change
equation. I. The two-parameter model. Soil Sci. Soc. Am. J. 47:419422.
doi:10.2136/sssaj1983.03615995004700030005x
Gijsman, A.J., S.S. Jagtap, and J.W. Jones. 2002. Wading through a swamp of
complete confusion: How to choose a method for estimating soil water retention parameters for crop models. Eur. J. Agron. 18:77105. doi:10.1016/
S1161-0301(02)00098-9
Gimnez, D., E. Perfect, W.J. Rawls, and Y.A. Pachepsky. 1997. Fractal models for predicting soil hydraulic properties. Rev. Eng. Geol. 48:161183.
doi:10.1016/S0013-7952(97)00038-0
Glover, P.W.J., and E. Walker. 2009. Grain-size to effective pore-size transformation derived from electrokinetic theory. Geophysics 74:E17E29.
doi:10.1190/1.3033217

www.VadoseZoneJournal.org

Green, T.R., L.R. Ahuja, and J.G. Benjamin. 2003. Advances and challenges in
predicting agricultural management effects on soil hydraulic properties.
Geoderma 116:327. doi:10.1016/S0016-7061(03)00091-0
Guadagnini, A., A.F.S.J. Martnez, and Y.A. Pachepsky. 2013. Scaling in soil
and other complex porous media. Vadose Zone J. 12. doi:10.2136/
vzj2013.05.0092
Gupta, S.C., and W.E. Larson. 1979. Estimating soil water retention characteristics from particle size distribution, organic matter percent and bulk density.
Water Resour. Res. 15:16331635. doi:10.1029/WR015i006p01633
Hadas, A., and D. Wolf. 1984. Soil aggregates and clod strength dependence
on clod size, cultivation and stress load rates. Soil Sci. Soc. Am. J. 48:1157
1165. doi:10.2136/sssaj1984.03615995004800050041x
Haines, W.B. 1930. Studies in the physical properties of soils. V. The hysteresis
effect in capillary properties and the modes of moisture distribution associated therewith. J. Agric. Sci. 20:97116. doi:10.1017/S002185960008864X
Hanks, R.J., A. Klute, and E. Bresler. 1969. A numeric method for estimating
infiltration, redistribution, drainage, and evaporation of water from soils.
Water Resour. Res. 5:10641069. doi:10.1029/WR005i005p01064
Haverkamp, R., and J.Y. Parlange. 1986. Predicting the water retention curve
from particle size distribution: I. Sandy soils without organic matter. Soil Sci.
142:325339. doi:10.1097/00010694-198612000-00001
Haverkamp, R., M. Vauclin, J. Touma, P.J. Wierenga, and G. Vachaud.
1977. A comparison of numerical simulation models for one-dimensional infiltration. Soil Sci. Soc. Am. J. 41:285294. doi:10.2136/
sssaj1977.03615995004100020024x
Hogarth, W.L., J. Hopmans, J.-Y. Parlange, and R. Haverkamp. 1988. Application of a simple soil-water hysteresis model. J. Hydrol. 98:2129.
doi:10.1016/0022-1694(88)90203-X
Huang, H.C., Y.C. Tan, C.W. Liu, and C.H. Chen. 2005. A novel hysteresis model in
unsaturated soil. Hydrol. Processes 19:16531665. doi:10.1002/hyp.5594
Hunt, A.G. 2005. Basic transport properties in natural porous media. Complexity 10:2237. doi:10.1002/cplx.20067
Hunt, A.G., and G.W. Gee. 2002. Water-retention of fractal soil models using
continuum percolation theory: Tests of Hanford Site soils. Vadose Zone J.
1:252260.
Hutson, J.L., and A. Cass. 1987. A retentivity function for use in soil-water
simulation models. J. Soil Sci. 38:105113. doi:10.1111/j.1365-2389.1987.
tb02128.x
Hwang, S.I. 2004. Effect of texture on the performance of soil particle-size
distribution models. Geoderma 123:363371. doi:10.1016/j.geoderma.2004.03.003
Hwang, S.I., and S.P. Hong. 2006. Estimating relative hydraulic conductivity
from lognormally distributed particle-size data. Geoderma 133:421430.
doi:10.1016/j.geoderma.2005.08.006
Israelachvili, J.N. 1991. Intermolecular and Surface Forces. 2nd ed. Academic
Press, New York.
Irmay, S. 1954. On the hydraulic conductivity of unsaturated soils. Trans. Am.
Geophys. Union 35:463467. doi:10.1029/TR035i003p00463
Jarvis, N. 2008. Near-saturated hydraulic properties of macroporous soils. Vadose Zone J. 7:13021310. doi:10.2136/vzj2008.0065
Johnson, D.L., and L. Schwartz. 1989. Unified theory of geometrical effects in
transport properties of porous media, 30th Annual Logging Symposium,
SPWLA, Paper E.
Katz, A.J., and A.H. Thompson. 1986. Quantitative prediction of permeability
in porous rock. Phys. Rev. B 34:81798181. doi:10.1103/PhysRevB.34.8179
Kern, J.S. 1995. Evaluation of soil water retention models based on basic
soil physical properties. Soil Sci. Soc. Am. J. 59:11341141. doi:10.2136/
sssaj1995.03615995005900040027x
Kool, J.B., and J.C. Parker. 1987. Development and evaluation of closed-form
expressions for hysteretic properties. Water Resour. Res. 23:105114.
doi:10.1029/WR023i001p00105
Kosugi, K. 1994. Three-parameter lognormal distribution model for soil water
retention. Water Resour. Res. 30:891901. doi:10.1029/93WR02931
Kosugi, K. 1996. Lognormal distribution model for unsaturated soil hydraulic
properties. Water Resour. Res. 32:26972703. doi:10.1029/96WR01776
Kosugi, K. 1999. General model for unsaturated hydraulic conductivity for soils
with lognormal pore-size distribution. Soil Sci. Soc. Am. J. 63:270277.
doi:10.2136/sssaj1999.03615995006300020003x
Kosugi, K., J.W. Hopmans, and J.H. Dane. 2002. Parametric models. In: J.H. Dane
and G.C. Topp, editors, Methods of soil analysis, Part 4. SSSA Book Ser. 5.
SSSA, Madison, WI. p. 739757.
Kozeny, J. 1927. Uber kapillare leitung des wassers im boden. Sitzungsber. Akad.
Wiss. Wien 136:271306.
Laliberte, G.E., A.T. Corey, and R.H. Brooks. 1966. Properties of unsaturated
porous media. Hydrology Paper 17. Colorado State University, Fort Collins.
Laroussi, Ch., and L.W. de Backer. 1979. Relations between geometrical properties of glass beads media and their main y(q) hysteresis loops. Soil Sci. Soc.
Am. J. 43:646650. doi:10.2136/sssaj1979.03615995004300040004x
Lebeau, M., and J.-M. Konrad. 2010. A new capillary and thin film flow model
for predicting the hydraulic conductivity of unsaturated porous media. Water Resour. Res. 46:W12554. doi:10.1029/2010WR009092

p. 17 of 20

Lehmann, P., S. Assouline, and D. Or. 2008. Characteristics lengths affecting


evaporative drying from porous media. Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 77. doi:10.1103/PhysRevE.77.056309
Leij, F.J., W.B. Russell, and S.M. Lesch. 1997. Closed form expressions for
water retention and conductivity data. Ground Water 35:848858.
doi:10.1111/j.1745-6584.1997.tb00153.x
Lenhard, R.J., J.C. Parker, and S. Mishra. 1989. On the correspondence between
Brooks-Corey and van Genuchten Models. J. Irrig. Drain. Eng. 115:744751.
doi:10.1061/(ASCE)0733-9437(1989)115:4(744)
Lesmes, D.P., and S.P. Friedman. 2005. Relationships between the electrical and
hydrogeological properties of rocks and soils. In: Y. Rubin and S.S. Hubbard,
editors, Hydrogeophysics. Springer, Dordrecht, the Netherlands. p. 87128.
Li, Y., and N.C. Wardlaw. 1986. Mechanisms of nonwetting phase trapping
during imbibition at slow rates. J. Colloid Interface Sci. 109:473486.
doi:10.1016/0021-9797(86)90325-5
Likos, W.J., and R. Jafaar. 2013. Pore-scale model for water retention and fluid
partitioning of partially saturated granular soil. J. Geotech. Geoenviron.
Eng. 139:724737. doi:10.1061/(ASCE)GT.1943-5606.0000811
Luckner, L., M.Th. van Genuchten, and D.R. Nielsen. 1989. A consistent set of
parametric models for the two-phase flow of immiscible fluids in the subsurface. Water Resour. Res. 25:21872193. doi:10.1029/WR025i010p02187
McBride, M.B., and P. Baveye. 2002. Diffuse double-layer models, long-range
forces, and ordering in clay colloids. Soil Sci. Soc. Am. J. 66:12071217.
doi:10.2136/sssaj2002.1207
Mandelbrot, B. 1967. How long is the coast of Britain? Statistical self-similarity
and fractional dimension. Science. New Series 156:636638. doi:10.1126/
science.156.3775.636.
Marshall, T.J. 1958. A relation between permeability and size distribution of
pores. J. Soil Sci. 9:18. doi:10.1111/j.1365-2389.1958.tb01892.x
Mason, G., and N. Morrow. 1991. Capillary behavior of a perfectly wetting
liquid in irregular triangular tubes. J. Colloid Interface Sci. 141:262274.
doi:10.1016/0021-9797(91)90321-X
Mavko, G., and A. Nur. 1997. The effect of percolation threshold in the Kozeny
Carman relation. Geophysics 62:14801482. doi:10.1190/1.1444251
Millington, R.J., and J.P. Quirk. 1961. Permeability of porous solids. Faraday Soc.
Trans. 57:12001206. doi:10.1039/tf9615701200
Milly, P.C.D. 1987. Estimation of BrooksCorey parameters from water retention
data. Water Resour. Res. 23:10851089. doi:10.1029/WR023i006p01085
Mohammadi, M.H., and M. Vanclooster. 2011. Predicting the soil moisture
characteristic curve from particle size distribution with a simple conceptual
model. Vadose Zone J. 10:594602. doi:10.2136/vzj2010.0080
Morel-Seytoux, H.J., P.D. Meyer, M. Nachabe, J. Touma, M.T. van Genuchten,
and R.J. Lenhard. 1996. Parameter equivalence for the Brooks-Corey and
van Genuchten soil characteristics: Preserving the effective capillary drive.
Water Resour. Res. 32:12511258. doi:10.1029/96WR00069
Mualem, Y. 1974a. A catalogue of the hydraulic properties of unsaturated soils.
Technion, Israel Inst. of Technology, Haifa, Israel.
Mualem, Y. 1974b. A conceptual model of hysteresis. Water Resour. Res.
10:514520. doi:10.1029/WR010i003p00514
Mualem, Y. 1976. A new model of predicting the hydraulic conductivity of
unsaturated porous media. Water Resour. Res. 12:513522. doi:10.1029/
WR012i003p00513
Mualem, Y. 1978. Hydraulic conductivity of unsaturated porous media: Generalized macroscopic approach. Water Resour. Res. 14:324334.
Mualem, Y. 1984. A modified dependent-domain theory of hysteresis. Soil Sci.
137:283291. doi:10.1097/00010694-198405000-00001
Mualem, Y. 1986. Hydraulic conductivity of unsaturated soils, predictions and
formulas. In: A. Klute, editor, Methods of soil analysis. Agron. Monogr. 9.
ASA and SSSA, Madison, WI. p. 799823.
Mualem, Y., and S. Assouline. 1989. Modeling soil seal as a nonuniform layer.
Water Resour. Res. 25:21012108. doi:10.1029/WR025i010p02101
Mualem, Y., and A. Beriozkin. 2009. General scaling rules of the hysteretic water retention function based on Mualems domain theory. Eur. J. Soil Sci.
60:652661. doi:10.1111/j.1365-2389.2009.01130.x
Mualem, Y., and G. Dagan. 1975. A dependent domain model of capillary hysteresis. Water Resour. Res. 11:452460. doi:10.1029/WR011i003p00452
Mualem, Y., and G. Dagan. 1978. Hydraulic conductivity of soils: Unified
approach to the statistical models. Soil Sci. Soc. Am. J. 42:392395.
doi:10.2136/sssaj1978.03615995004200030003x
Mualem, Y., and E.E. Miller. 1979. A hysteresis model based on explicit domain
dependence function. Soil Sci. Soc. Am. J. 43:10671073. doi:10.2136/
sssaj1979.03615995004300060002x
Nasta, P., N. Romano, S. Assouline, J.A. Vrugt, and J.W. Hopmans. 2013a. Prediction of spatially-variable unsaturated hydraulic conductivity using
scaled particle-size distribution functions. Water Resour. Res. doi:10.1002/
wrcr.20255.
Nasta, P., S. Assouline, J.B. Gates, J.W. Hopmans, and N. Romano. 2013b. Prediction of relative unsaturated hydraulic from Kosugis water retention
function. Procedia Environmental Sciences 19:609617. doi:10.1016/j.proenv.2013.06.069

www.VadoseZoneJournal.org

Nel, L. 1942. Thories des lois daimantation de Lord Rayleigh, 1. Cahiers de


Physique 12:120.
Nel, L. 1943. Thories des lois daimantation de Lord Rayleigh, 2. Cahiers de
Physique 13:1930.
Nimmo, J.R. 1991. Comment on the treatment of residual water content in A
consistent set of parametric models for the two-phase flow of immiscible
fluids in the subsurface by L. Luckner et al. Water Resour. Res. 27:661662.
doi:10.1029/91WR00165
Nimmo, J.R. 1992. Semiempirical model of soil water hysteresis. Soil Sci. Soc.
Am. J. 56:17231730. doi:10.2136/sssaj1992.03615995005600060011x
Nimmo, J.R., W.N. Herkelrath, and A.M. Laguna Luna. 2007. Physically based
estimation of soil water retention from textural data: General framework,
new models, and streamlined existing models. Vadose Zone J. 6:766773.
doi:10.2136/vzj2007.0019
Ohshima, H. 1995. Effective surface potential and double-layer interaction
of colloidal particles. J. Colloid Interface Sci. 174:4552. doi:10.1006/
jcis.1995.1362
Or, D., and T.A. Ghezzehei. 2002. Modeling post-tillage soil structure dynamics:
A review. Soil Tillage Res. 64:4159. doi:10.1016/S0167-1987(01)00256-2
Or, D., P. Lehmann, E. Shahraeeni, N. Shokri. 2013. Advances in soil evaporation
physicsA review. Vadose Zone J. 12. doi:10.2136/vzj2012.0163 (this issue).
Or, D., F.J. Leij, V. Snyder, and T.A. Ghezzehei. 2000a. Stochastic model of
post-tillage soil pore space evolution. Water Resour. Res. 36:16411652.
doi:10.1029/2000WR900092
Or, D., U. Shani, and A.W. Warrick. 2000b. Subsurface tension permeametry.
Water Resour. Res. 36:20432053. doi:10.1029/2000WR900121
Or, D., and M. Tuller. 1999. Liquid retention and interfacial area in variably saturated porous media: Upscaling from single-pore to sample-scale model.
Water Resour. Res. 36:11651177. doi:10.1029/2000WR900020
Or, D., and M. Tuller. 2000. Flow in unsaturated fractured porous media: Hydraulic conductivity of rough surfaces. Water Resour. Res. 36:11651177.
doi:10.1029/2000WR900020
Pachepsky, Ya. 1990. Mathematical models of physical chemistry in soil science.
(In Russian.) Nanka, Moscow.
Pachepsky, Ya.A., T.A. Polubesova, M. Hajnos, Z. Sokolowska, and G. Jozefaciuk. 1995a. Fractal parameters of pore surface area as influenced by
simulated soil degradation. Soil Sci. Soc. Am. J. 59:6875. doi:10.2136/
sssaj1995.03615995005900010010x
Pachepsky, Y.A., and W.J. Rawls. 2004. Status of pedotransfer functions. In: Y.A.
Pachepsky and W.J. Rawls, editors, Development of pedotransfer functions
in soil hydrology. Developments in Soil Science. Vol. 30. Elsevier Science,
New York. p. viixvi.
Pachepsky, Y.A., W.J. Rawls, and D.J. Timlin. 1999. The current status of pedotransfer functions: Their accuracy, reliability and utility in field and regionalscale modeling. In: D.L. Corwin, K.M. Loague, and T.R. Ellsworth, editors,
Assessment of non-point source pollution in the vadose zone. Geophys.
Monogr. 108. American Geophysical Union, Washington, DC. p. 223234.
Pachepsky, Ya.A., R.A. Shcherbakov, and L.P. Korsunskaya. 1995b. Scaling of soil water retention using a fractal model. Soil Sci. 159:99104.
doi:10.1097/00010694-199502000-00003
Pachepsky, Ya.A., W.J. Rawls, D. Gimenez, and J.P.C. Watt. 1998. Use of soil
penetration resistance and group method of data handling to improve
soil water retention estimates. Soil Tillage Res. 49:117126. doi:10.1016/
S0167-1987(98)00168-8
Parlange, J.-Y. 1976. Capillary hysteresis and the relationship between drying and wetting curves. Water Resour. Res. 12:224228. doi:10.1029/
WR012i002p00224
Perrier, E., M. Rieu, G. Sposito, and G. de Marsily. 1996. Models of the water
retention curve for soils with a fractal pore size distribution. Water Resour.
Res. 32:30253031. doi:10.1029/96WR01779
Peters, A., and W. Durner. 2008a. Simplified evaporation method for determining soil hydraulic properties. J. Hydrol. 356:147162. doi:10.1016/j.jhydrol.2008.04.016
Peters, A., and W. Durner. 2008b. A simple model for describing hydraulic conductivity in unsaturated porous media accounting for film and capillary
flow. Water Resour. Res. 44:W11417. doi:10.1029/2008WR007136
Pham, H.Q.D.G.F., and S.L. Barbour. 2005. A study of hysteresis models for soilwater characteristic curves. Can. Geotech. J. 42:15481568. doi:10.1139/
t05-071
Poulovassilis, A. 1962. Hysteresis of pore water, an application of the concept
of independent domains. Soil Sci. 93:405412. doi:10.1097/00010694196206000-00007
Purcell, W.R. 1949. Capillary pressuresTheir measurements using mercury
and the calculation of permeability therefrom. Pet. Trans. Am. Inst. Min.
Metall. Pet. Eng. 186:3948.
Quirk, J.P. 1986. Soil permeability in relation to sodicity and salinity. Philos.
Trans. R. Soc. Lond. A 316:297317. doi:10.1098/rsta.1986.0010
Quispe, J.R., R.E. Rozas, and P.G. Toledo. 2005. Permeability-porosity relationship from a geometrical model of shrinking and lattice Boltzmann and Monte Carlo simulations of flow in two-dimensional pore networks. Chem. Eng.
J. 111:225236. doi:10.1016/j.cej.2005.02.003

p. 18 of 20

Ransohoff, T.C., and C.J. Radke. 1988. Laminar flow of a wetting liquid along
the corners of a predominantly gas-occupied noncircular pore. J. Colloid
Interface Sci. 121:392401. doi:10.1016/0021-9797(88)90442-0
Rawls, W.J. 1983. Estimating soil bulk density from particle size analyses and
organic matter content. Soil Sci. 135:123125. doi:10.1097/00010694198302000-00007
Rawls, W.J., and D.L. Brakensiek. 1989. Estimation of soil water retention and
hydraulic properties. In: H.J. Morel-Seytoux, editor, Unsaturated flow in
hydrologic modelingTheory and practice. Kluwer Academic Publishing,
Dordrecht, the Netherlands. p. 275300.
Rawls, W.J., T.J. Gish, and D.L. Brakensiek. 1991. Estimating soil water retention
from soil physical properties and characteristics. Adv. Agron. 16:213234.
Rawls, W.J., D. Gimenez, and R. Grossman. 1998. Use of soil texture, bulk density and slope of the water retention curve to predict saturated hydraulic
conductivity. Trans. ASAE 41:983988.
Reicovsky, D.C., W.B. Voorhees, and J.K. Radke. 1980. Unsaturated water flow
through a simulated wheel track. Soil Sci. Soc. Am. J. 45:38. doi:10.2136/
sssaj1981.03615995004500010001x
Revil, A., and L.M. Cathles III. 1999. Permeability of shaly sands. Water Resour.
Res. 35:651662. doi:10.1029/98WR02700
Rieu, M., and G. Sposito. 1991. Fractal fragmentation, soil porosity and soil
water properties: I. Theory. Soil Sci. Soc. Am. J. 55:12311238. doi:10.2136/
sssaj1991.03615995005500050006x
Ritchie, J.T., L.F. Ratliff, and D.K. Cassel. 1987. Soil laboratory data, field descriptions and field measured soil water limits for some soils of the United
States. ARS Technical Bulletin, ARS Agric. Research Service and Soil Conservation Service/State Agric. Exp. Stn., Temple, TX.
Romano, N. 2004. Spatial structure of PTF estimates. In: Y.A. Pachepsky and W.J.
Rawls, editors, Development of pedotransfer functions in soil hydrology. Elsevier Science, New York. p. 295319.
Romero, L.A., and F.G. Yost. 1996. Flow in an open channel capillary. J. Fluid
Mech. 322:109129. doi:10.1017/S0022112096002728
Ross, P.J., J. Williams, and K.L. Bristow. 1991. Equations for extending waterretention curves to dryness. Soil Sci. Soc. Am. J. 55:923927. doi:10.2136/
sssaj1991.03615995005500040004x
Rouault, Y., and S. Assouline. 1998. A probabilistic approach towards modeling
the relationship between particle and pore size distributions: The multicomponent sphere pack case. J. Powder Technol. 96:3341. doi:10.1016/
S0032-5910(97)03355-X
Rucker, D.F., A.W. Warrick, and T.P.A. Ferre. 2005. Parameter equivalence for
the Gardner and van Genuchten soil hydraulic conductivity functions
for steady vertical flow with inclusions. Adv. Water Resour. 28:689699.
doi:10.1016/j.advwatres.2005.01.004
Russo, D. 1988. Determining soil hydraulic properties by parameter estimation:
On the selection of a model for the hydraulic properties. Water Resour. Res.
24:453459. doi:10.1029/WR024i003p00453
Saxton, K.E., and W.J. Rawls. 2006. Soil water characteristic estimates by
texture and organic matter for hydrologic solutions. Soil Sci. Soc. Am. J.
70:15691578. doi:10.2136/sssaj2005.0117
Saxton, K.E., W.J. Rawls, J.S. Romberger, and R.I. Papendick. 1986. Estimating
generalized soil water characteristics from texture. Trans. ASAE 50:1031
1035.
Schaap, M.G., and M.Th. van Genuchten. 2006. A modified Mualemvan Genuchten formulation for improved description of the hydraulic conductivity
near saturation. Vadose Zone J. 5:2734. doi:10.2136/vzj2005.0005
Scheidegger, A.E. 1960. The physics of flow through porous media. Revised ed.
Univ. of Toronto Press, Toronto.
Scheidegger, A.E. 1974. The physics of flow through porous media. 3rd ed. Univ.
of Toronto Press, Toronto.
Schelle, H., S.C. Iden, A. Peters, and W. Durner. 2010. Analysis of the agreement
of soil hydraulic properties obtained from multistep-outflow and evaporation methods. Vadose Zone J. 9:10801091. doi:10.2136/vzj2010.0050
Schindler, U., and L. Mller. 2006. Simplifying the evaporation method for
quantifying soil hydraulic properties. J. Plant Nutr. Soil Sci. 169:623629.
doi:10.1002/jpln.200521895
Scott, P.S., G.J. Farquhar, and N. Kouwen. 1983. Hysteretic effects on net infiltration. In: Advances in Infiltration, Proceedings of the National Conference
on Advances in Infiltration. Publ. 1183. American Society of Agricultural
Engineers, St Joseph, MI. p. 163170.
Shiozawa, S., and G.S. Campbell. 1991. On the calculation of mean particle diameter and standard deviation from sand, silt and clay fractions. Soil Sci.
152:427431. doi:10.1097/00010694-199112000-00004
Shirazi, M.A., and L. Boersma. 1984. A unifying quantitative analysis of soil texture. Soil Sci. Soc. Am. J. 48:142147. doi:10.2136/
sssaj1984.03615995004800010026x
Sisavath, S., X.D. Jing, and R.W. Zimmerman. 2000. Effect of stress on the
hydraulic conductivity of rock pores. Phys. Chem. Earth 25:163168.
doi:10.1016/S1464-1895(00)00026-0
Skaggs, T.H., L.M. Arya, P.J. Shouse, and B.P. Mohanty. 2001. Estimating particle-size distribution from limited soil texture data. Soil Sci. Soc. Am. J.
65:10381044. doi:10.2136/sssaj2001.6541038x

www.VadoseZoneJournal.org

Smiles, D.E., G. Vachaud, and M. Vauclin. 1971. A test of the uniqueness of


the soil moisture characteristic during transient, nonhysteretic flow
of water in a rigid soil. Soil Sci. Soc. Am. J. 35:534539. doi:10.2136/
sssaj1971.03615995003500040018x
Sommer, R., and C. Stckle. 2010. Correspondence between the Campbell and
van Genuchten soil-water-retention models. J. Irrig. Drain. Eng. 136:559
562. doi:10.1061/(ASCE)IR.1943-4774.0000204
Stange, C.F., and R. Horn. 2005. Modeling the soil water retention curve for
conditions of variable porosity. Vadose Zone J. 4:602613. doi:10.2136/
vzj2004.0150
Stankovich, J.M., and D.A. Lockington. 1995. Brooks-Corey and van Genuchten
soil-water-retention models. J. Irrig. Drain. Eng. 121:17. doi:10.1061/
(ASCE)0733-9437(1995)121:1(1)
Tani, M. 1982. The properties of a water-table rise produces by a one-dimensional, vertical, unsaturated flow. (In Japanese with English summary.) J.
Jpn. For. Soc. 64:409418.
Tenchov, B.G., and T.K. Yanev. 1986. Weibull distribution of particle sizes obtained by uniform random fragmentation. J. Colloid Interface Sci. 111:17.
doi:10.1016/0021-9797(86)90002-0
Tietje, O., and M. Tapkenhinrichs. 1993. Evaluation of pedo-transfer functions. Soil Sci. Soc. Am. J. 57:10881095. doi:10.2136/
sssaj1993.03615995005700040035x
Timlin, D.J., Ya.A. Pachepsky, B. Acock, and F. Whisler. 1996. Indirect estimation
of soil hydraulic properties to predict soybean yield using GLYCIM. Agric.
Syst. 52:331353. doi:10.1016/0308-521X(96)00001-7
Tokunaga, T.K., J. Wan, and K.R. Olson. 2002. Saturation-matric potential relations in gravel. Water Resour. Res. 38:1214. doi:10.1029/2001WR001242
Tokunaga, T.K., K.R. Olson, and J.M. Wan. 2003. Moisture characteristics of Hanford gravels: Bulk, grain-surface, and intragranular components. Vadose
Zone J. 2:322329.
Topp, G.C. 1971. Soil-water hysteresis: The domain model theory extended to pore interaction conditions. Soil Sci. Soc. Am. Proc. 35:219225.
doi:10.2136/sssaj1971.03615995003500020017x
Topp, G.C., and E.E. Miller. 1966. Hysteretic moisture characteristics and hydraulic conductivities for glass-bead media. Soil Sci. Soc. Am. Proc. 30:156
162. doi:10.2136/sssaj1966.03615995003000020008x
Torquato, S. 2002. Random heterogeneous materials: Microstructure and macroscopic properties. Springer, New York.
Tuller, M., and D. Or. 2001. Hydraulic conductivity of variably saturated porous
media: Film and corner flow in angular pore space. Water Resour. Res.
37:12571276. doi:10.1029/2000WR900328
Tuller, M., and D. Or. 2002. Unsaturated Hydraulic Conductivity of Structured
Porous Media. Vadose Zone J. 1:1437.
Tuller, M., and D. Or. 2003. Hydraulic functions for swelling soils: Pore scale
considerations. J. Hydrol. 272:5071. doi:10.1016/S0022-1694(02)00254-8
Tuller, M., and D. Or. 2005. Water films and scaling of soil characteristic curves at low water contents. Water Resour. Res. 41:W09403.
doi:10.1029/2005WR004142
Tuller, M., D. Or, and L.M. Dudley. 1999. Adsorption and capillary condensation
in porous media: Liquid retention and interfacial configurations in angular
pores. Water Resour. Res. 35:19491964. doi:10.1029/1999WR900098
Tyler, S.W., and S.W. Wheatcraft. 1989. Application of fractal mathematics to soil
water retention estimation. Soil Sci. Soc. Am. J. 53:987996. doi:10.2136/
sssaj1989.03615995005300040001x
Tyler, S.W., and S.W. Wheatcraft. 1990. Fractal processes in soil water retention.
Water Resour. Res. 26:10471054. doi:10.1029/WR026i005p01047
Unger, P.W. 1971. Soil profile gravel layers: I. Effect on water storage, distribution, and evaporation. Soil Sci. Soc. Am. Proc. 35:631634. doi:10.2136/
sssaj1971.03615995003500040041x
Vachaud, G., and J.-L. Thony. 1971. Hysteresis during infiltration and redistribution in a soil column at different initial water content. Water Resour. Res.
7:111127. doi:10.1029/WR007i001p00111
van Genuchten, M.Th. 1980. A closed-form equation for predicting the hydraulic conductivity of unsaturated soils. Soil Sci. Soc. Am. J. 44:892898.
doi:10.2136/sssaj1980.03615995004400050002x
van Genuchten, M.Th., and D.R. Nielsen. 1985. On describing and predicting
the hydraulic properties of unsaturated soils. Ann. Geophys. 3:615628.
Verbist, K.M.J., W.M. Cornelis, S. Torfs, and D. Gabriels. 2013. Comparing methods to determine hydraulic conductivities on stony soils. Soil Sci. Soc. Am. J.
77:2542. doi:10.2136/sssaj2012.0025
Vereecken, H., J. Maes, J. Feyen, and P. Darius. 1989. Estimating the soil moisture retention characteristic from texture, bulk density, and carbon content.
Soil Sci. 148:389403. doi:10.1097/00010694-198912000-00001
Vereecken, H., J. Maes, and J. Feyen. 1990. Estimating unsaturated hydraulic conductivity from easily measured soil properties. Soil Sci. 149:112.
doi:10.1097/00010694-199001000-00001
Vereeckeen, H., M. Weynants, M. Javaux, Y. Pachepsky, M.G. Schaap, and M.Th.
van Genuchten. 2011. Using pedotransfer functions to estimate the van
GenuchtenMualem soil hydraulic properties: A review. Vadose Zone J.
9:795820. doi:10.2136/vzj2010.0045

p. 19 of 20

Viaene, P., H. Vereecken, J. Diels, and J. Feyen. 1994. A statistical analysis of


six hysteresis models for the moisture retention characteristic. Soil Sci.
157:345355. doi:10.1097/00010694-199406000-00003
Vogel, T., M.Th. van Genuchten, and M. Cislerova. 2000. Effect of the shape
of the soil hydraulic functions near saturation on variably saturated
flow predictions. Adv. Water Resour. 24:133144. doi:10.1016/S03091708(00)00037-3
Wagner, B., V.R. Tarnawski, V. Hennings, U. Mller, G. Wessolek, and R. Plagge.
2001. Evaluation of pedo-transfer functions for unsaturated soil hydraulic conductivity using an independent data set. Geoderma 102:275297.
doi:10.1016/S0016-7061(01)00037-4
Wang, Y., S.M. Grove, and M.G. Anderson. 2008. A physicalchemical model for
the static water retention characteristic of unsaturated porous media. Adv.
Water Resour. 31:701713. doi:10.1016/j.advwatres.2008.01.005
Warkentin, B.P., G.H. Bolt, and R.D. Miller. 1957. Swelling pressure of
montmorillonite. Soil Sci. Soc. Am. Proc. 21:495497. doi:10.2136/
sssaj1957.03615995002100050009x
Weynants, M., H. Vereecken, and M. Javaux. 2009. Revisiting Vereecken pedotransfer functions: Introducing a closed-form hydraulic model. Vadose Zone
J. 8:8695. doi:10.2136/vzj2008.0062
White, I., D.E. Smiles, S. Santomartino, P. van Oploo, B.C.T. Macdonald, and
T.D. Waite. 2003. Dewatering and the hydraulic properties of soft, sulfidic,
coastal clay soils. Water Resour. Res. 39:1295. doi:10.1029/2002WR001324
Wind, J.P. 1955. A field experiment on capillary rise in a heavy clay soil. Neth. J.
Agric. Sci. 3:6069.
Wind, G.P. 1968. Capillary conductivity data estimated by a simple method. In:
P.E. Rijtema and H. Wassink, editors, Water in the unsaturated zone, Proc.
Wageningen Symp., June 1966. Vol. 1. IASAH, Gentbrugge, Belgium. p.
181191.
Wsten, J.H.M. 1997. Pedotransfer functions to evaluate soil quality. In: E.G.
Gregorich and M.R. Carter, editors, Soil quality for crop production and
ecosystem health. Developments in Soil Science 25. Elsevier, p. 221245.

www.VadoseZoneJournal.org

Wsten, J.H.M., A. Lilly, A. Nemes, and C. Le Bas. 1999. Development and use
of a database of hydraulic properties of European soils. Geoderma 90:169
185. doi:10.1016/S0016-7061(98)00132-3
Wyllie, M.R.J., and M.B. Sprangler. 1952. Application of electrical resistivity
measurements to problems of fluid flow in porous media. Bull. Am. Assoc.
Petr. Geol. 36:359403.
Wyllie, M.R.J., and G.H.F. Gardner. 1958. The generalized KozenyCarman
equation. A novel approach to problems of fluid flow. World Oil Prod. Sect.
146:210228.
Xu, X.F., and P. Dong. 2004. Fractal approach to hydraulic properties in unsaturated porous media. Chaos Solitons Fractals 19:327337. doi:10.1016/
S0960-0779(03)00045-6
Young, R.A., and W.B. Voorhees. 1982. Soil erosion and runoff from planting
to canopy development as influenced by tractor wheel-traffic. Trans. ASAE
3:708712.
Young, I.M., and J.W. Crawford. 1991. The fractal structure of soil aggregates: Its measurement and interpretation. J. Soil Sci. 42:187192.
doi:10.1111/j.1365-2389.1991.tb00400.x
Zhang, R., and M.Th. van Genuchten. 1994. New models for unsaturated soil hydraulic properties. Soil Sci. 158:7785. doi:10.1097/00010694-19940800000001
Zhang, Z.P., A.B. Yu, and J.A. Dodds. 1997. Analysis of the pore characteristics of mixtures of disks. J. Colloid Interface Sci. 195:818. doi:10.1006/
jcis.1997.5130
Zhu, W., M.K. Tivey, H. Gittings, and P.R. Craddock. 2007. Permeability-porosity
relationships in seafloor vent deposits: Dependence on pore evolution processes. J. Geophys. Res. 112:B05208. doi:10.1029/2006JB004716

p. 20 of 20

Você também pode gostar