You are on page 1of 6

WATER RESOURCES RESEARCH, VOL. 33, NO.

3, PAGES 491-496, MARCH 1997


Exact mathematical derivation of a two-term infiltration
equation
D. Swartzendruber
Department of Agronomy, University of Nebraska-Lincoln
Abstract. Previous mathematical developments of so-called two-term infiltration
equations have been approximate instead of exact and thus provide little insight to
account for any experimental shortcomings of such equations. In contrast, it is here shown
that one form of two-term infiltration equations is obtainable by integration from an exact
solution of the one-dimensional downward Richards equation subject to the customary
initial and boundary conditions. The mathematical trial solution is a variables-separable
combination of product and additive forms. Intrinsic to the solution are two further
stipulations: (1) the unsaturated hydraulic conductivity function must be linear with the
water content, and (2) the ponded-water head on the soil surface must increase as the
square root of time from initial water application. The absence of both these stipulations
in customary ponded-infiltration measurements thus accounts for why the two-term
equation frequently fails experimentally. Nonetheless, the mathematical simplicity of the
new solution makes it useful pedagogically.
Introduction
The two-term water infiltration equation [Philip, 1957c] can
be written
I = Ss tl/2 q- Mt (1)
where I is the cumulative quantity of infiltration after time t of
ponded-water application to the soil surface, and S s is the
water sorptivity of the soil alone, with subscript s denoting zero
ponding (i.e., the applied water layer on the soil surface is
infinitesimally thin). As originally presented, parameter M,
akin to hydraulic conductivity, could be linked rigorously with
Philip's [1957a] exact, t 1/2 series solution only by constraining
t from becoming too large. This would mean that 0 < M < Ko
[Philip, 1957c, pp. 261-262; 1969b, p. 275], where Ko is the
sated (satiated) hydraulic conductivity of Miller and Bresler
[1977].
Notwithstanding the early optimistic expectations [e.g.,
Philip, 1957c; Watson, 1959] for the practical utility of (1) in
spite of its intrinsic time constraint, subsequent work did un-
cover difficulties, in the sense that the least-squares fitting of
(1) (Watsoh's [1959] procedure) to experimental field data
yielded negative values of M in half or more of the data sets
[Skaggs et al., 1969; Fahad et al., 1982]. Admittedly, such be-
havior could be rationalized or excused by appealing to the
problematic influence of the time constraint inherent in the
origins of (1). Unfolding confusingly alongside these findings
was the erroneous claim that M = Ko was rigorous [Miller and
Klute, 1967, p. 226]. There was a passing suggestion [Philip,
1969b, p. 284] that (1) fitted to the whole t range of infiltration
data would tend to give M = Ko.
In the Childs [1969, p. 279] sense of "an empirical formula to
which one is led on physical grounds," Swartzendruber and
Youngs [1974] set M = Ko in (1) and compared it with two
other infiltration equations. Swartzendruber [1987, p. 812] also
Copyright 1997 by the American Geophysical Union.
Paper number 96WR03906.
0043-1397/97/96WR-03906509.00
found that (1) with M - Ko would emerge from a quasi-
solution, multiparameter exponential form when the number
of parameters was reduced from six to just two: S and K o. This,
in turn, ultimately prompted the question, Can we pose an
exact-solution boundary value problem that will produce (1)
with M = Ko and thus bestow exactness upon the two-term
infiltration equation (in Ko) without constraint on time t? The
answer is yes, and we now turn to the details of this problem,
the exact solution of which also affords insight into the reasons
for failure in I - Ss tl/2 q- Kot when fitted by least squares to
infiltration data as customarily obtained.
Mathematical Analysis
The Problem
Consider one-dimensional downward infiltration of water
into an infinitely long, uniform, rigid soil (Figure 1). The gov-
erning partial differential equation is that of Richards [1931],
which when written in transformed manner [Philip, 1957a, p.
350; Swartzendruber, 1969, p. 229] is
OZ O[D] dK
Ot = O0 oZ/00 + dO (2)
wherein Z - Z(0, t) is the depth (positive downward, below
the soil surface) at which the volumetric water content is 0 at
time t. D = D (0) and K = K(0) are the soil water diffusivity
and hydraulic conductivity functions, respectively, and (2) is
taken subject to
Z>0 0= 0n t=0 (3)
Z=Z0 0= 00 t>0 (4)
Zo = roh/(dI/dt - ro) (5)
where 0n is the uniform initial water content; 0o is the sated
water content (constant) corresponding to K(0o) = Ko result-
ing from the water depth h being ponded instantaneously on
the soil surface (Z - 0) from t = 0 + onward; Z o is the depth
491
492 SWARTZENDRUBER: TECHNICAL NOTE
Soil Surface, Z = 0
time t < 0 time t = 0 time t > 0
(a) (b) (c)
Figure 1. Schematic diagram of the one-dimensional down-
ward infiltration problem showing three time stages in the
application and maintenance of ponded water.
to which 0o (at zero pressure or suction head) has penetrated
in time t (for zero air entry value); dI/dt is the volumetric
infiltration flux of the water across (into) the soil surface; and
cumulative infiltration I is the total volume of water infiltrated
(in time t) per unit of bulk soil cross-sectional area perpen-
dicular to the downward direction of water flow. Figures lb
and lc depict conditions (3) and (4), respectively.
Boundary condition (4), which limits (2) to the region Z _>
Z o with Z o specified by [5), was first given in essence by Philip
[1958a, p. 280]. He noted for the region 0 <- Z <_ Z o that the
water content is the constant 0o, so that K(0o) = Ko, and that
the pressure head is h at Z = 0 and is zero at Z = Z o. With
the total difference in hydraulic head across Z o thus being h +
Z o, he applied Darcy's equation to the Z o region and solved
for Z o to obtain (5).
We here shall also explicitly inject the possibility of a more
general, time-dependent expectation for h, namely
h =hc+p(t) (6)
where h c is constant and p(t) is a function only of time, with
p(0) = 0. This means that the initial ponded head h c is
applied instantaneously on the soil surface at t = 0. In the
special case p (t) = 0 for all t, (6) would then reduce to the
constant ponded-head case h -- h c [Philip, 1958a]. Last, note
that the instantaneous imposition of ponded-head h c at t = 0
will in turn raise the water content instantaneously from 0,, to
0o at the soil surface (Z = 0). Hence the position coordinate
Z of any 0 between 0,, and 0o will originate from zero at t =
0; analytically
Z=0 t=0 for On< 0--< 00 (7)
The Solution
As a trial solution of (2) subject to conditions (3), (4), and
(5), along with stipulation (6) and corollary (7), we introduce
Z = O(o)r(t) + U(t) (8)
where 4>(0) is a function only of 0, and T(t) and U(t) are each
functions only of t, but T and U are different functions. The
form of (8) is a variables-separable combination of product and
additive types. From partial differentiation of (8) with respect
to t and 0, respectively,
og/ot = O(o)r'(t) + U'(t) (9)
og/00 = O'(o)r(t) (t0)
where T' = dT/dt, U' = dU/dt, and 4>' = dO/dO. Using (9)
and (10) in (2)yields
or' + U' = -O(D/O'T)/00 + g' (lt)
where K' = dK/dO, but
O(D/O'T)/00 = T-O(D/O')/00 = r-d(D/O')/dO (12)
so that the use of (12) in (tt) gives
or' + U' = -r-d(D/O')/dO + g' (13)
As it stands, the ordinary differential equation (13) is not of
variables-separable form but can be made so by setting
U' =K' (14)
and then (13) becomes, after rearrangement,
rr' = -O-d(D/O' )/d O (15)
Beginning with (14), note that the left-hand side is a function
only of t while the right-hand side is a function only of 0, thus
meaning that both sides can be set equal to the separation
constant a, or
d U/dt = a (16)
dK/d 0 = a (17)
Integrating (16) and (17) yields, respectively,
U = at + C (18)
K=a0+C2 (19)
where C and C2 are constants of integration. The successive
use in (19) ofK = K(On) = Kn at 0,, and ofK = K(0o) =
K o at 0o enables the evaluation of a by subtraction and rear-
rangement to be written
a = (go- gn)/(0o- On) (20)
Turning next to (15), the variables again are separable with
separation constant b, namely,
rr' = b (21)
d(D/O')/dO = -bO (22)
Solution of (21) and (22) is akin to that of Swartzendruber
[1966] but with some change in details. Integrating (21) and
solving for T yields
T = (2bt + C3) /2 (23)
where C3 arises from the constant of integration. Putting (23)
and (18) back into the trial solution (8) gives
Z = O(O)(2bt + C3) 1/2 --[- at + C (24)
wherein a is retained for economy in writing symbols until the
right-hand side of (20) is expressly needed. Using corollary (7)
in (24) provides, after rearrangement,
C 1 = -C31/20(0 ) (25)
SWARTZENDRUBER: TECHNICAL NOTE 493
in which finite nonzero C and C3 are not permissible, since
that would force qb(0) to be constant over the whole range of 0
( 0, to 0o). Neither could qb(0) be chosen as zero for the general
nontrivial case. It might seem consistent with (25) to choose C
and C3 to be large without bound (i.e., c), but such choices
would not be useful in (24). Hence we select C3 = 0 for the
general finite qb(0) > 0, which from (25) makes C = 0, so that
(24) becomes
g = &(O)(2bt) /2 + at (26)
Although (2b)/2&(O) in (26) could be defined as a new
function of only 0, the same end is achieved without loss of
generality by choosing b = 1/2 and continuing the use of qb(0),
which makes (26) and (22) become, respectively,
Z = c(O)t /2 + at (27)
d(D/&')/dO = -qb/2 (28)
Note that the consequence embodied in (14), arising out of
the variables-separation process, has forced the K( 0 ) relation-
ship to be the linear (19). Such linearity is indeed a noteworthy
limitation, but has nonetheless been employed theoretically
before by Parlange and Fleming [1984, equation (6)], taking
0, = 0, D(0 ) as a power function, and allowing 0 o to be time
dependent. Although they hypothesized a similarity expression
(their equation (7)) reminiscent of (27) but with t /2 replaced
by t -m, they did not pursue our problem here in which 0o
(condition (4)) is constant rather than time dependent.
To determine whether the initial condition (3) is satisfied,
rewrite (27) as
git /2= 49(0) + at /2 (29)
Putting condition (3) into (29) by letting 0 0, as t - 0 for
nonzero finite Z yields
4'(0) (30)
which thus supplies the first condition to which the ordinary
differential equation (28) becomes subject. Next, putting
boundary condition (4) into (27) gives
Z o = Co t/2 + at (31)
where qb o = qb(0o). To investigate whether (31) can be matched
with (5) will require I for use in (5), which as a general for-
mulation [Philip, 1957a, p. 352] for fixed t:
I
I = rnt + Z dO (32)
If (27) is the solution to the problem, the Z from it can be
substituted into (32) to obtain
i00
I = Knt + t /2 c(O) dO + at dO (33)
where t /2 and t are taken through their respective integral
signs because t is fixed. In (33), carry out the simple evaluation
of the second integral and substitute for a from (20), to yield
ultimately
I
I = t /2 &(O) dO + rot (34)
In (34), since 4> is a function only of 0 and the limits of
integration are fixed, the integral is fixed. For an infinitesimally
thin layer of ponded water (h = 0), Philip [1957c, p. 257;
1969b, p. 237] defined a similar fixed integral as the soil sorp-
tivity, Ss, in (1). Philip [1958a] extended the sorptivity concept
to include the effect of nonzero ponding, which is here ex-
pressed explicitly as total sorptivity S, namely
s = 0(o) ao (35)
although Philip [1958a, b] did not write (35) explicitly in the
sense of total sorptivity. Nonetheless, it is to be emphasized in
the present context that the S of (35) is conceived to include
the condition of nonzero water ponding on the soil surface.
Using (35) in (34) hence yields
I: St /2 + Kot (36)
To determine whether (36) will satisfy (5) and (6), we find by
differentiation of (36) that dI/dt = K o + S/2t /2, which when
substituted into (5) along with the h of (6) yields
Zo = 2(Kohct/2/S) + 2Kop(t)t/2/S (37)
The Z o of (31) matches exactly the Z o of (37) if we make the
identifications
Oo: 2gohc/S (38)
p(t) = Sat/2/2Ko (39)
where (39) provides the particular form of p(t) in (6), con-
forming as well to p(0) - 0 as mentioned. Also, (38), the
second condition to which (28) is subject (along with (30)), is
the same as that found by Philip [1958a, p. 280, second of
conditions (15)] for constant-head nonzero ponding in hori-
zontal water adsorption.
Integration of (28) from 0, to 0 [Philip, 1955] leads to
o dO
ck dO = -D(O) dck (40)
In principle, this integration between limits also yields the term
D(On)[dO/dck]o:o, added to the right-hand side of (40), but
this term vanishes because condition (30) implies d O/d ck -- 0
at 0 - 0,. The ordinary integrodifferential equation (40)
remains subject only to 4> = qbo at 0 = 0o, with qb o given by (38).
For D(0), 0,, 0o, and h c supplied as inputs, (40) is solvable
(for qb(0)) by the numerical method of Philip [1955] as modified
[Philip, 1958a, p. 280] for the nonzero qb o (for h c > 0) of (38).
If, however, h c = 0, meaning that initial ponding (at t = 0) is
infinitesimally thin because h (0) = 0 in (6) by virtue of p (0)
= 0, then qb o of (38) is also zero, and the numerical method for
determining qb(0) reverts back to the original one of Philip
[1955]. Also, of course, with qb o = 0 (from h c = 0), S would
then be replaced by S s in (35) through (39), thus reducing (36)
to (1).
With the mathematical proof thus completed, the two-term
infiltration equation (36) is therefore seen to arise exactly from
the two-term solution embodied in (27), namely, the solution
of (2) subject to conditions (3) and (4) as supplemented by (5),
(6), and (7). Included in (36) are both zero and nonzero initial
surface ponding, for hc = 0 and hc > 0, respectively, although
even the case of zero initial ponding becomes both ponded and
494 SWARTZENDRUBER: TECHNICAL NOTE
progressively more deeply ponded for t > 0, by virtue ofp (t)
from (39) appearing in (6).
Results and Discussion
From the two-term (27), it is clear that the new exact solu-
tion is simply the additive combination of b() t/:, the clas-
sical horizontal solution [Philip, 1955], plus the linear term at.
This same additive linear term also appears in the classical,
constant-ponding infiltration solution for asymptotically large
times [Philip, 1957b].
Graphical illustration of the new two-term solution requires
numerical values of b(O) and a in (27). Such values were given
by Philip [1958a] for the now classical Yolo soil material under
nonzero ponding (h = 0.25 m of water) and also for zero
ponding (h = 0) [Philip, 1957a, b], along with K o = 123.00
rim/s, K = 0.12 rim/s, o = 0.4950, and = 0.2376, to yild
a = 477.39 nm/s from (20). We recognize from the outset,
however, that these numerical values of Philip's were deter-
mined from a nonlinear K() rather than from a linear K(O),
as in (19). Nonetheless, it is still deemed of interest to utilize
these values for illustrating profiles in this first-time explor-
atory sense.
For h = 0.25 m of water, a water-content profile, or set of
paired values (, Z), was determined from (27) at each se-
lected time of 0.01, 0.04, 0.10, and 0.20 Ms. A second sequence
of profiles for the same times was determined from Philip's
[1958a] solution under constant pondcd head of 0.25 m (i.e.,
p(t) = 0 in (6) for all ! > 0), using the first three terms of his
exact t / series solution. A third sequence of profiles at these
same times was determined from just the first term, b(O), on
the right-hand side of (27), which thus constitutes the profiles
for horizontal water absorption (gravity free).
Graphic profiles for these three cases are shown in Figure 2.
At very small times, all three proilia cases merge as they
should, because b(O) predominates. As time increases, the
gravity-free profiles lag progressively behind their gravity-
aided infiltration counterparts, again, as would be expected.
Also, the forward portions of the constant-ponding (p (!) = 0)
profiles become steeper (as viewed from the Z axis) in com-
parison with their gravity-free and new-solution counterparts,
once more as expected. The new-solution profiles clearly dem-
onstrate their progressive translational advancement beyond
their respective gravity-free counterparts, by the distance +at
in (27). Also, these new-solution profiles arc progressively
ahead of even their respective constant-ponding profiles (p(t)
= 0), in response to the time buildup of pondcd-watcr head
expressed by p(t) 0 of (39) in (6).
Profiles for zero initial ponding (h - 0) on the Yolo soil
material [Philip, 1957a, b], using the scaled data of Swartzen-
dmber [1987, p. 813], are shown in Figure 3. To avoid confusing
overlappings, the zero-gravity profiles are omitted. Obviously,
the behavior and trends of Figure 2 are not only sustained but
even accentuated in Figure 3, especially the substantially
deeper penetration of the new-solution profiles as compared
with their respective constant-ponding profiles. For example,
at 3.5 Ms (17.5 times the maximum 0.2 Ms in Figure 2), the
constant-ponding profile is everywhere more shallow than even
the 3.0-Ms new-solution profile. Also, as time increases over
the range shown in Figure 3, the new-solution profiles become
progressively less steep than their respective constant-ponding
profiles, as viewed from the Z axis. This means that the new-
solution profiles do not approach the constant-shape, linearly
0.1
0.4
0.5 /
WATER CONTENT 0
h c =0.25m
0.01
0o0'

p(f) > O, new solution.
/ p(t) = O, Philip [1958a]
J
...... zero gravity.
Figure 2. Three types of water-content profiles at different
times of downward infiltration into the Yolo soil material of
Philip [1958a], for an initial ponded-water head of h c - 0.25
m of water on the soil surface.
translating, traveling-wave profiles of the constant-ponding
case [Philip, 1957b, pp. 447-448], even though the infiltration
flux dI/dt from (36) does approach K o as t - . This behavior
is qualitatively akin to the linearized, constant-diffusivity solu-
tion of Philip [1969a], which also does not produce traveling-
wave profiles, although the profiles of that linearized solution
WATER CONTENT 0
0.2Ms
N
0.5Ms
"' 1.0Ms
1.5Ms
2.5Ms / I
3.0Ms . I
3.5Ms
p(f) > O, new solution
p(t) = O, Philip [1957a].
Figure 3. Two types of water-content profiles at different
times of downward infiltration into the Yolo soil material of
Philip [1957a], for an initial ponded-water head of h c - 0 on
the soil surface. For times of 1.0 Ms and larger, the two types
of profiles at a given same time are connected by a vertical line.
SWARTZENDRUBER: TECHNICAL NOTE 495
K 0
K(e) /
i
K n
0 '
e On eo
Figure 4. Schematic curve (solid) of K(0) for a coarse-
textured soil. Also shown for the high-0 range (0n -< 0 -< 0o)
is a least-squares linear approximation (dashed line) of the
true curve.
are far less satisfactory in shape than even those of the new
solution in Figures 2 and 3.
We now turn to the two major restrictive assumptions resi-
dent in the new two-term solution (27) and hence also in (36).
The first is the requirement in (19) of a linear K( 0 ). Notwith-
standing the employment of such linearity both here and by
?arlange and Fleming [1984], K(0) is generally strongly non-
linear. For coarse-textured soils, however, the nonlinearity
might be dealt with approximately, as portrayed in Figure 4 by
a least-squares straight line fitted to the high-0 range of the
actual curvilinear K( 0 ).
The second and perhaps even greater restrictive assumption
is that of the time-dependent buildup of ponded-water head on
the soil surface, as embodied in the functionp (t) given by (39).
Once again, using Philip's [1957a, 1958b] values of S, a, and
K o to evaluate p(t) from (39), h was determined from (6) and
plotted in Figure 5 for h c values of 0, 0.25, and 0.50 m of water.
' I ' I ' I
1.25 - -
E
1.00 - hc = 0.50m _
-r 75 --
n,' = 0.25m
.5o -
I
.25 - hc = 0 _
z
0
0
0 1.0 2.0 5.0
TIME f (Ms)
Figure 5. Time curves of ponded head h from (6) with p(t)
given by (39) for three values of initial ponded head h c = 0,
0.25, and 0.50 m of water, for the Yolo soil material of Philip
[1957a, 1958a].
Even for h c -- 0, the value of h is 71.5 mm of water after 86.4
ks (1.00 day), and 455 mm after 3.5 Ms (40.51 days). These
values of h are in substantial and changing disagreement with
a constant h of, say, 15 mm in typical infiltration experiments,
in which the time-dependent ponded-head buildup for the
validity of the two-term (36) is not being met, even apart from
any failure of K(0) to be linear as in (19).
If an experimental assessment of the new solution were
undertaken in the sense of Figure 4, a corresponding buildup
of time-dependent ponded head in accordance with (39) and
(6) should be feasible experimentally. To carry this out, how-
ever, would require prior experimentation for determining S,
a, and K o of (39) before conducting the infiltration experiment
portrayed in Figure 1. Unfortunately, though, the overall de-
finitiveness of the experimentation would still be conditioned
by the extent to which K(0) in the high-0 range were truly
linear.
Even though the assumption of a linear K(0) is an impor-
tant limitation, it is nonetheless the only constraint on the
three soil-characterizing functions K(0), D (0), and r(0),
where r is the soil-water suction head. Specifically, there is no
need for the artificiality of any of these three functions being
held constant, in contrast with the requirement D(0) = con-
stant in the solutions of Philip [1969a] and of Knight [Philip,
1974, p. 261]. Note also that these two solutions, along with
those of Philip [1957a, b, 1958a], are far more complicated
mathematically than the new solution embodied in (27) and
(36). Hence this new and relatively simple exact solution for
downward infiltration is useful pedagogically for teachers and
students of vadose-zone hydrology, a setting in which such
simple but rigorous examples are few.
In summary, the exact derivation of the two-term equation
(36) includes the two important stipulations: (1) that the hy-
draulic conductivity function K(0) be linear for 0n -< 0 -< 0o
and (2) that the ponded-water head p(t) on the soil surface
build up with time in accordance with equation (39), namely
p(t) = Sat/2/2Ko . Since neither of these stipulations is met
in the usual measurements of ponded-water infiltration, this
accounts for why the two-term infiltration equation frequently
fails experimentally. This reminds us that the mathematical
exactness of equation (27), desirable and worthy though it is,
does not of itself ensure the applicability of equation (36).
Acknowledgments. Published as paper 11476, Journal Series, Ag-
ricultural Research Division, University of Nebraska-Lincoln. Thanks
are expressed to John R. Philip, CSIRO Fellow Emeritus, Center for
Environmental Mechanics, Canberra, Australia, for kindly supplying
data used in Figure 2.
References
Childs, E. C., An Introduction to the Physical Basis of Soil Water Phe-
nomena, John Wiley, New York, 1969.
Fahad, A. A., L. N. Mielke, A.D. Flowerday, and D. Swartzendruber,
Soil physical properties as affected by soybean and other cropping
sequences, Soil Sci. Soc. Am. J., 46, 377-381, 1982.
Miller, E. E., and A. Klute, The dynamics of soil water, I, Mechanical
forces, in Irrigation of Agricultural Lands, edited by R. M. Hagan, H.
R. Haise, and T. W. Edminster, pp. 209-244, Am. Soc. of Agron.,
Madison, Wis., 1967.
Miller, R. D., and E. Bresler, A quick method for estimating soil water
diffusivity functions, Soil Sci. Soc. Am. J., 41, 1020-1022, 1977.
Parlange, J.-Y., and J. F. Fleming, First integrals of the infiltration
equation, 1, Theory, Soil Sci., 137, 391-394, 1984.
Philip, J. R., Numerical solution of equations of the diffusion type with
496 SWARTZENDRUBER: TECHNICAL NOTE
diffusivity concentration dependent, Trans. Faraday Soc., 51, 885-
892, 1955.
Philip, J. R., The theory of infiltration, 1, The infiltration equation and
its solution, Soil Sci., 83, 345-357, 1957a.
Philip, J. R., The theory of infiltration, 2, The profile of infinity, Soil
Sci., 83, 435-448, 1957b.
Philip, J. R., The theory of infiltration, 4, Sorptivity and algebraic
infiltration equations, Soil Sci., 84, 257-264, 1957c.
Philip, J. R., The theory of infiltration, 6, Effect of water depth over
soil, Soil Sci., 85, 278-286, 1958a.
Philip, J. R., The theory of infiltration, 7, Soil Sci., 87, 333-337, 1958b.
Philip, J. R., A linearization technique for the study of infiltration, in
Proceedings of the Symposium on Water in the Unsaturated Zone, vol.
1, pp. 471-478, Int. Assoc. of Hydrol. Sci. and UNESCO, Gent-
brugge, Belgium, 1969a.
Philip, J. R., Theory of infiltration, Adv. Hydrosci., 5, 215-296, 1969b.
Philip, J. R., Recent progress in the solution of nonlinear diffusion
equations, Soil Sci., 117, 257-264, 1974.
Richards, L. A., Capillary conduction of liquids through porous medi-
ums, Physics, 1,318-333, 1931.
Skaggs, R. W., L. E. Huggins, E. J. Monke, and G. R. Foster, Exper-
imental evaluation of infiltration equations, Trans. ASAE, 12, 822-
828, 1969.
Swartzendruber, D., Variables-separable solution of the horizontal
flow equation with nonconstant diffusivity, Soil Sci. Soc. Am. Proc.,
30, 7-11, 1966.
Swartzendruber, D., The flow of water in unsaturated soils, in Flow
Through Porous Media, edited by R. J. M. De Wiest, pp. 215-292,
Academic, San Diego, Calif., 1969.
Swartzendruber, D., A quasi-solution of Richards' equation for the
downward infiltration of water into soil, Water Resour. Res., 23,
809-817, 1987.
Swartzendruber, D., and E.G. Youngs, A comparison of physically
based infiltration equations, Soil Sci., 117, 165-167, 1974.
Watson, K. K., A note on the field use of a theoretically derived
infiltration equation, J. Geophys. Res., 64, 1611-1615, 1959.
D. Swartzendruber, Department of Agronomy, University of Ne-
braska-Lincoln, 133 Keim Hall, East Campus, Lincoln, NE 68583.
(Received April 8, 1996; revised December 9, 1996;
accepted December 16, 1996.)

You might also like