You are on page 1of 8

Relationship between thermal conductivity and water

content of soils using numerical modelling


P. CO S E N Z A , R. GU E

R I N & A. TA B B A G H
UMR 7619 Sisyphe, Universite Pierre et Marie Curie and CNRS, case 105, 4 place Jussieu, 75252 Paris Cedex 05, France
Summary
There is no simple and general relationship between the thermal conductivity of a soil, `, and its
volumetric water content, 0, because the porosity, n, and the thermal conductivity of the solid fraction,
`
s
, play a major part. Experimental data including measurements of all the variables are scarce. Using a
numerical modelling approach, we have shown that the microscopic arrangement of water influences the
relation between ` and 0. Simulated values for n ranging from 0.4 to 0.6, `
s
ranging from 2 to 5 Wm
1
K
1
and 0 from 0.1 to 0.4 can be fitted by a simple linear formula that takes into account n, `
s
and 0. The
results given by this formula and by the quadratic parallel (QP) model widely used in physical property
studies are in satisfactory agreement with published data both for saturated rocks and for unsaturated
soils. Consequently, the linear formula and the QP model can be used as practical and efficient tools to
investigate the effects of water content and porosity on the thermal conductivity of the soil and hence to
optimize the design of thermal in situ techniques for monitoring water content.
Introduction
The determination and the monitoring of water content in
soils and unsaturated superficial layers are among the funda-
mental requirements in both agronomy and water manage-
ment. One needs to know, on the one hand, the water reserve
and, on the other, the rate of water flow through the unsat-
urated layers. Many methods and techniques have been tested
to determine the water content of a given soil volume, in
particular those measuring electrical properties, conductivity
or permittivity that are quick and cheap. Thermal properties
also merit attention: they depend significantly on water con-
tent, and measuring temperature is easy and much used in soil
studies. Two independent thermal properties are involved in
the dominant transfer process, namely conduction. The first,
the volumetric heat capacity, C
v
, has been recognized as lin-
early linked to the volumetric water content, 0 (de Vries, 1963),
and is now in common use both in the laboratory and in the
field to determine 0 with so-called heat-pulse probes
(Campbell et al., 1991). For the second, the thermal conduct-
ivity `, it has been recognized that the relationship, though
monotonically increasing, can be more complicated. Despite
continuous progress in the non-stationary step-pulse technique
with needle probes (de Vries, 1952; Blackwell, 1954; Tabbagh &
Jolivet, 1974; Tabbagh, 1985b; Larson, 1988), usable experi-
mental data on the relation between ` and 0 are scarce
(Ochsner et al., 2001), perhaps because the mineralogy of the
solid fraction and the air content have more pronounced
influence than 0 (Smith & Byers, 1938; Smith, 1939; Tabbagh,
1976). This is in contrast to the case of dielectric permittivity
(Topp et al., 1980). An increase in the conductivity with
increasing water content can be the consequence of the
decrease of the air content. An experimental study of the
conductivity against water content would have thus to take
into account air content and solid mineralogy, which would
lead to much experimentation with at least three variables.
This explains why recent studies with active thermal tech-
niques use C
v
rather than `. However, if one attempts to
deduce 0 from natural variation in temperature in a soil then
both properties have to be considered, and one needs a simple
and reliable expression of the relationship between ` and 0 or
at least an unequivocal expression of 0`/00 to allow water
content to be monitored.
To reach this objective, we chose to use numerical simula-
tions based on a physical analogy: if heat moves purely by
conduction then our problem is mathematically the same as
that in static electricity for electrical conductivity or dielectric
permittivity, so the same modelling techniques can be applied.
Consequently, we can ignore heat transfer by vapour flow
(Hiraiwa & Kasubuchi, 2000) or water flow, i.e. by convection.
As in the electrical analogue, we chose to apply the method
of moments (MoM) (Tabbagh et al., 2000, 2002) without
physical assumptions (except the unavoidable assumptions
bearing on the discretization of the geometries of solid, water
Correspondence: P. Cosenza. E-mail: cosenza@ccr.jussieu.fr
Received 10 September 2002; revised version accepted 30 January 2003
European Journal of Soil Science, September 2003, 54, 581587
# 2003 Blackwell Publishing Ltd 581
or air). Following this approach, we have considered a hetero-
geneous volume of soil where different cells corresponding to
the different phases that can be in any type of geometrical
arrangement. From these numerical calculations we propose
simple approximate formulae, which will be compared to the
few existing experimental results. First, we review the empirical
formulae commonly used and the principle of the MoM with
its links with previous attempts to take into account the shapes
and interactions of the different constituent elementary
volumes (de Vries, 1963).
Simple expressions or processes used as empirical
formulae
The behaviour of multiphase media can be complex, and the
first approach to determine any bulk physical property is
experiment. Given a series of experimental data, one wants
to express the results in a manner that is easy to use and links
the bulk property with that of each constituent and, thus, one
attempts to fit simple mathematical expressions to the results.
The choice of the expressions is wide open, but several are in
common use for properties of soils and rocks.
The first is the simple mixing law, which would be written as
`
X
N
i1
`
i
x
i
. 1
where `
i
is the conductivity, N is the total number of compon-
ents in the mixture, and x
i
is the volumetric content of the ith
component. This formula is exact for the density and the
volumetric heat capacity of porous media. By analogy with
an electrical lattice it is often called parallel. Following the
electrical analogy a series formula could be proposed:
`
X
N
i1
`
1
i
x
i
!
1
. 2
However, neither of themhave been recognized as convenient for
thermal conductivity. A third expression is the geometrical law:
`
Y
N
i1
`
xi
i
. 3
Woodside & Messmer (1961) proposed this expression for
thermal conductivity of two-phase saturated media. A fourth
expression is
`
X
N
i1
x
i

`
i
p
!
2
. 4
Note first that for dielectric permittivity, which determines the
velocity in electromagnetic wave propagation, this expression
corresponds to the summation of slowness and is thus equiva-
lent to the expression proposed by Wyllie et al. (1956) for
elastic P waves, and second that for electrical conductivity it
would correspond to Fn
2
, i.e. Archies law (F being the
formation factor) if the solid and air fraction conductivities are
null. As it successfully fitted the variation of both relative
dielectric permittivity and electrical conductivity of the soil
with water content, we could expect it to be equally relevant
for the mathematically analogous thermal conductivity,
even if we cannot give an underlying physical meaning. This
expression, i.e. Equation (4), has received several names:
quadratic parallel (QP), time-propagation (TP) and
complex refractive index (CRIM).
Combinations of the preceding expressions and simple poly-
nomials of order two or three could also be proposed.
Besides simple mathematical expressions, simple iterative
processes easy to implement on computers can model the
physical properties of porous media. The differential effective
medium (DEM) scheme is such a process (Sheng, 1991). One
begins by considering one phase only; at each step a small
fraction of other constituent(s) is added, it replaces an equiva-
lent volume of the existing material and is diluted in it to
constitute a new material to which the next step is applied.
The process is stopped when the expected volumetric propor-
tions are reached.
Method of moments in heat conduction
The method of moments (MoM) was proposed (Harrington,
1968) to solve electrical or electromagnetic problems (Raiche,
1974; Tabbagh, 1985a). It establishes the equivalence between
(a) the presence inside a given volume of an inclusion with
differing properties and (b) the presence of secondary dipolar
sources which are proportional to the total field and to the
local conductivity contrast.
Consider an inhomogeneity of electrical conductivity o
1
in a
uniform space of conductivity o
0
. The total electrical field E
obeys the following integral equation with the appropriate
Greens function G(r,r
0
) (e.g. Raiche, 1974; Tabbagh, 1985a):
Er E
p
r o
1
o
0


V
0
Gr.r
0
Er
0
dt
0
. 5
where V
0
is the volume of the inhomogeneity, E
p
is the
primary field, which would exist in absence of inhomogeneity,
r denotes the current field point and r
0
denotes the position
of the elementary volume dt
0
.
The mathematical analogy leads us to consider an equivalent
thermal field, which is the opposite of the temperature gradient
(rT). If we consider M inhomogeneities of constant thermal
conductivity `
i
in a uniform space of thermal conductivity `
0
,
then the equivalent thermal field satisfies the following integral
equation:
rT rT
p

X
M
i1
`
i
`
0


Vi
0
Gr.r
0
rTr
0
dt
0
.
6
582 P. Cosenza et al.
# 2003 Blackwell Publishing Ltd, European Journal of Soil Science, 54, 581587
where the V
i
0
are the volumes of the inhomogeneities asso-
ciated with the constant conductivity `
i
. In the MoM, one
converts the integral Equation (6) into a set of algebraic equa-
tions (see for instance Harrington, 1968). We divide the
volume into N cubic cells and assume that the equivalent
thermal field is constant in each cell. Then, the integral in
Equation (6) can be approximated by a finite summation:
rT
m
rT
m
p

X
M
i1
`
i
`
0

X
N
n1
rT
n

V
0
n
i
Gr.r
0
dt
0
. 7
where the vectors (rT)
m
and (rT)
m
p
are, respectively, the
unknown equivalent thermal field at the centre of cell m (the
centre of which is at r) and the primary equivalent thermal
field at the centre of cell m. Rearranging Equation (7), we
obtain in more concise notation:
rT
m
p
A
mn
rT
n
. 8
where
A
mn
c
mn

X
M
i1
`
i
`
0

X
N
i1

in
and

in


V
0
n
i
Gr.r
0
dt
0
.
The integral G
in
is numerically evaluated and c
mn
is the unity
tensor (c
mn
1 if mn, c
mn
0 if m6n).
Therefore, the vector (rT) is the solution of Equation (8),
each element of the matrix A
mn
representing the coupling of
one component of the temperature gradient in one cell to one
component of the temperature gradient in another (or the
same) cell. For N cells, the system has 3N linear equations
with 3N unknowns, which are the three spatial components of
the vector (rT) in the centre of each elementary cell.
This method corresponds to a general approach from which the
different assumptions can be clearly defined. By adopting
(rT) (rT)
p
, i.e. the Born approximation, one gets a simple
summation of the effect of the different cells which corresponds in
fact to the simple mixing law, Equation (1). By neglecting the
coupling between cells and considering only the effect of each cell
on itself, Maxwell-Garnett (1904) proposed an approach to the
calculation of optical properties of complex media, which has con-
stituted the basis of inclusion-based models for electrical properties
and of the later de Vries (1963) work on soil thermal conductivity.
To represent a three-phase medium one considers the first
phase as the reference medium, here the solid fraction, in
which heterogeneities corresponding to the two other phases
are located, here water and air fractions. The heterogeneities
can represent the major part of the volume. Any type of
arrangement of the different cells can be chosen: for example,
to represent elongated volumes one can consider either an
in-line series of cubic cells or parallelepiped ones. One can
also consider big cells for one phase and small cells for another
phase, and fracture-like cells as well. In a soil volume the pore
number is very large, and this would lead to a huge matrix, but
it was established in previous studies of dielectric permittivity
(Tabbagh et al., 2000) and electrical conductivity (Tabbagh
et al., 2002) that by considering successive trials of randomly
located cells the results are stable even for a number of cells as
small as 300. In the calculations we thus consider the median
of 27 different trials of 1000 cell positions.
Saturated media
As the thermal conductivity of air can be only one hundredth that
of solid grains, its presence in a medium corresponds to large
contrasts between the different elementary components, which
makes difficult the assessment of assumptions supporting simple
formulae. It is thus advisable to begin by studying saturated media
for which numerous references exist in the literature. In the earth
sciences, indeed, both oil-bearing sedimentary rocks (Zierfuss &
van der Vliet, 1956) and marine sediments (Von Herzen &
Maxwell, 1959; Ratcliffe, 1960) have been studied to determine
geothermal heat flowinthe ocean. Inthe first context, Woodside &
Messmer (1961) have proposed the geometrical law model:
` `
1n
s
`
n
w
. 9
where n is the porosity, `
s
the conductivity of the solid grains
and `
w
that of water (0.60 Wm
1
K
1
). This model fits well
with marine sediments (Lovell, 1984, 1985).
In Figure 1 we present the curves for thermal conductivity
against water content for four different models with a 3 Wm
1
K
1
solid conductivity:
(a) the MoM numerical calculation with a random arrange-
ment of isotropic (cubic) water cells inside the solid;
(b) a DEM scheme;
(c) the geometrical law model, and
(d) the QP model.
The agreement is perfect between (a), (b) and (d). The decrease in
conductivitywiththe increase of water content is more pronounced
in (c) than for the three others, but this effect is clearly apparent
onlybecause inthe present calculationthe conductivityvalue of the
solid fraction is fixed. When experimental data are considered a
little change can easily be introduced in the value of `
s
to enhance
the fit. If one considers a multiphase solid fraction the results
remain similar to those of Figure 1.
Unsaturated media
For the thermal conductivity in the air one adopts
`
a
0.024 Wm
1
K
1
and for water `
w
0.60 Wm
1
K
1
and in both media one considers only conductive transfers.
Thermal conductivity and water content of soils 583
#
2003 Blackwell Publishing Ltd, European Journal of Soil Science, 54, 581587
The study of the relation between dielectric permittivity and
water content of unsaturated soils has established that the
microscopic scale arrangement of the elementary water
volumes has an important influence (Tabbagh et al., 2000).
Thus Figure 2 compares the case of a random distribution of
water and air isotropic (cubic) volumes and the case of elong-
ated randomly orientated water and air volumes. Here also a
difference exists between the two arrangements: for n 0.6 and
`
s
3 Wm
1
K
1
, one has a variation from 0.83 Wm
1
K
1
for 0 0 to 1.17 Wm
1
K
1
for 0 0.4 in the first arrange-
ment and from 0.74 to 1.12 Wm
1
K
1
in the second; the
slopes of the curves are not different, but the second curve is
systematically lower. Even if this difference is more limited
than for permittivity, we adopt in the following calculations
elongated volumes for water and air since it is for this type of
arrangement that a good agreement exists with experiment for
dielectric permittivity.
When varying n, `
s
and 0, one observes that the depend-
ences on water content are linear (Figures 3 and 4 and also
Figure 2). The slopes and the origins, however, depend on
both n and `
s
. As expected from the already published experi-
mental data, there is, thus, no possibility to have a unique
curve enabling determination of water content from the ther-
mal conductivity of a soil. The agreement with the QP formula
Figure 1 Comparison between the different models of the `(0)
relationship for a two-phase saturated medium with 3 Wm
1
K
1
solid thermal conductivity and 0.6 Wm
1
K
1
water thermal
conductivity.
Figure 2 `f(0), results of numerical simulation, comparison between
cubic and elongated microscopic elementary volumes of air and water
(`
s
3 Wm
1
K
1
and n 0.6).
Figure 3 `f(0), results of numerical simulation using elongated
elementary volumes for water, influence of the porosity n
(`
s
3 Wm
1
K
1
).
584 P. Cosenza et al.
# 2003 Blackwell Publishing Ltd, European Journal of Soil Science, 54, 581587
is good for the largest values of the water content, better than
1% for 0 0.4, but worse for smaller 0. If 0 0.15, commonly
the case in temperate climates, one can consider this formula
always as efficient. In those cases the corresponding value of
the derivative can be adopted when monitoring the water
content:
0`
0 0
1.24

`
p
. 10
with
`

`
s
p
1 n

`
w
p
0

`
a
p
n 0

g
2
. 11
The relation in Equation (10) can be used to study the sensi-
tivity of thermal conductivity to changes in water content. To
provide an easy to use result of the numerical calculations, we
propose a polynomial formula fitting to the whole set of
simulated data for n ranging from 0.4 to 0.6, `
s
ranging from
2 to 5 Wm
1
K
1
and 0 from 0.1 to 0.4:
` 0.8908 1.0959n`
s
1.2236 0.3485n0. 12
This formula is called hereafter the numerical simulation (NS)
formula. The agreement between the whole set of numerically
simulated data and this formula is better than 1%, while the
standard deviation remains at 7% for the QP formula as a
result of the discrepancy for small water contents.
Comparison with published experimental data
The first published data acquired with a transient method
(Van Duin & de Vries, 1954) exhibit a linear variation of `
with 0. The slope of the curve linking ` with 0 depends on
porosity, which is qualitatively in complete agreement with the
present simulation results. However, as noted above, experi-
mental series comprising thermal conductivity of solid and
porosity together with a variation of the water content are
few. Moreover, the first studies used a stationary method
where the too high temperature gradients resulted in thermo-
migration, i.e. thermally driven transport of water (Smith &
Byers, 1938; Smith, 1939), which renders the data unusable for
a comparison.
In order to fix the parameters of his model, de Vries (1963)
used thermal conductivity data for Fairbanks sand and Healy
clay (Kersten, 1949), and it is interesting to compare these data
with the models. Unfortunately no data were given for the
thermal conductivities of the solid, and the only thing that
can be done is to calculate the values of this parameter that
give the best fit with the above formulae and to assess the
likelihood of these values. By considering data for 0 0.1, one
obtains for Fairbanks sand `
s
3.72 Wm
1
K
1
for the NS
formula and 3.84 Wm
1
K
1
for the QP formula; both values
are likely for a sand with quartz content around 60%. For the
Healy clay, one obtains 2.04 Wm
1
K
1
for the NS formula
and 2.08 Wm
1
K
1
for the QP formula. A comparison
between experimental data and the calculated values from
the QP formula and the NS formula is given in Figure 5. A
good agreement is obtained.
Two soils among the series considered in the study by
Ghuman & Lal (1985) present an extended range of water
content (from 0.06 to 0.24), a loam (numbered 3), and a
sandy clay loam (numbered 17). For both one can determine
the porosity from the bulk density, the dry density and the
water content. Comparison with the models leads to small
values of `
s
: 1.45 Wm
1
K
1
with NS and 1.0 Wm
1
K
1
with QP for the loam and 0.67 Wm
1
K
1
with NS and
0.73 Wm
1
K
1
with QP for the sandy clay loam. These values
are small and provide theoretical curves with a good trend but
a poorer agreement with experimental data than for other soils
(Figure 6). This suggests that NS and QP formulae would not
be valid for soils with unusually small values of `
s
, typically
less than 1.0. Note that the value of 1.0 is much smaller than
the minimum of the `
s
range used in our numerical simulations
to establish the NS model.
Hopmans & Dane (1986) published data on `(0) for
Norfolk sandy loam. The porosity equals 0.412 (solid mineral
volumetric content 0.584 and organic matter volumetric con-
tent 0.004). At 22

C the value of `
s
that provides the best fits
with the NS formula is 2.74 Wm
1
K
1
and 2.87 Wm
1
K
1
Figure 4 `f(0), results of numerical simulation using elongated
elementary volumes for water, influence of the solid fraction thermal
conductivity (n 0.5).
Thermal conductivity and water content of soils 585
#
2003 Blackwell Publishing Ltd, European Journal of Soil Science, 54, 581587
with QP. The theoretical curves associated with these values of
`
s
are in a satisfactory agreement with the experimental data
(Figure 6).
In their paper presenting the tests of a newly developed
TDR and Dual-Probe Heat Pulse (DPHP) probe, Ren et al.
(1999) presented laboratory results for a silica sand. From
their heat capacity curve one can, by adopting a value of
1.9 MJ m
3
K
1
for the solid fraction, determine the porosity
n 0.385. By extracting the numerical values from their Figure
8 showing the `(0) curve, one can compare them with the
models. The best fitting value of `
s
is 4.65 Wm
1
K
1
for
NS and 4.94 Wm
1
K
1
for QP, likely values for a silica
sand. A good agreement is obtained between the experimental
data and the theoretical curves calculated with the best fitting
values of `
s
(Figure 6).
Singh & Devid (2000) considered laboratory samples of clay,
sand and mixtures. Their results (Table 6 in the reference) on
single solid phase silty sand exhibit sufficiently large values of 0
and ` to be within the domain of application of the NS formula.
By calculating the porosity from dry density (and adopting a
2.65g cm
3
solid fraction density) and 0 from the moisture
(weight) and the dry density, one obtains `
s
2.13 Wm
1
K
1
for NS and 2.26 Wm
1
K
1
for QP. Figure 6 shows the thermal
conductivity calculated with the NS and QP formulae for a clay
and a silty clay. A satisfactory agreement with the correspond-
ing experimental data is observed.
Abu-Hamdeh et al. (2001) compared heating and cooling
hot wire methods for the determination of `, with 0 0.06.
They considered three different soils, clay loam, loam and
sandy loam. By calculating n and 0, one obtains, using, respect-
ively, the NS and the QP formulae, `
s
1.20 and 1.38 Wm
1
K
1
for the clay loam, 1.37 and 1.55 Wm
1
K
1
for the loam,
and 1.89 and 2.13 Wm
1
K
1
for the sandy loam. The increase
of `
s
with the sand content is a consistent result.
All these comparisons are for laboratory samples, the struc-
ture of which has been destroyed, and not for in situ measure-
ments. Also, experimentally determined thermal conductivities
of the solid fraction are not given. Nonetheless, both the NS
formula and the QP formula give consistent and likely results
which can be compared with the experimental data with a
satisfactory agreement (Figures 5 and 6).
Conclusions
As with other thermal properties, conductivity ` depends on
water content, but the porosity and the thermal conductivity
of the solid fraction are also strong determinants. Hence deter-
mination of water content from ` cannot be as simple as for
the dielectric permittivity . To compensate for the lack of
complete experimental data sets over a sufficiently wide
range of soil texture, we use numerical modelling by the
method of moments to simulate the influence of the different
Figure 6 Measured and calculated conductivities as a function of
volumetric water content.
Figure 5 Comparison between experimental data (de Vries, 1963) and
theoretical values calculated using the NS and QP formulae.
586 P. Cosenza et al.
# 2003 Blackwell Publishing Ltd, European Journal of Soil Science, 54, 581587
parameters. We propose a simple polynomial expression for
use as an easy to handle tool. We observe also that the
classical quadratic parallel (QP) expression gives a fair agree-
ment with the numerical results, both fitting well with the few
available sets of experimental data. If one considers the ana-
logy with , for which this expression fits well with experi-
ments, then this agreement is not surprising. In the case of ,
one component, the water, has a permittivity very different
from that of the others. In the case of `, one component, the
air, has a conductivity very different from that of the other
constituents.
References
Abu-Hamdeh, N.H., Khdair, A.I. & Reeder, R.C. 2001. A comparison
of two methods used to evaluate thermal conductivity of some soils.
International Journal of Heat and Mass Transfer, 44, 10731078.
Blackwell, J.H. 1954. A transient-flow method for determination of
thermal constants of insulating material in bulk, Part I. Theory.
Journal of Applied Physics, 25, 137144.
Campbell, G.S., Calissendorff, K. & Williams, J.H. 1991. Probe for
measuring soil specific heat using a heat-pulse method. Soil Science
Society of America Journal, 55, 291293.
De Vries, D.A. 1952. A nonstationary method for determining thermal
conductivity of soil in situ. Soil Science, 73, 8389.
De Vries, D.A. 1963. Thermal properties of soils. In: Physics of Plant
Environment (ed. W.R. van Wijk), pp. 210235. North-Holland,
Amsterdam.
Ghuman, B.S. & Lal, R. 1985. Thermal conductivity, thermal diffu-
sivity and thermal capacity of some Nigerian soils. Soil Science, 139,
7480.
Harrington, R.F. 1968. Field Computation by Moment Methods.
Macmillan, New York.
Hiraiwa, Y. & Kasubuchi, T. 2000. Temperature dependence of
thermal conductivity of soil over a wide range of temperature
(575

C). European Journal of Soil Science, 51, 211218.


Hopmans, J.W. & Dane, J.H. 1986. Thermal conductivity of two
porous media as a function of water content, temperature and
density. Soil Science, 142, 187195.
Kersten, M.S. 1949. Thermal properties of soils. Bulletin of the Uni-
versity of Minnesota Institute of Technology, 52, 1225.
Larson, T.H. 1988. Thermal measurement of soils using a multineedle
probe with a pulsed-point source. Geophysics, 53, 266270.
Lovell, M.A. 1984. Thermal conductivity and permeability assessment
by electrical resistivity measurements in marine sediments. Marine
Geotechnology, 6, 205240.
Lovell, M.A. 1985. Thermal conductivities of marine sediments.
Quarterly Journal of Engineering Geology, 18, 437441.
Maxwell-Garnett, J.C. 1904. Colours in metal glasses and metal films.
Philosophical Transactions of the Royal Society of London, 203,
385420.
Ochsner, T.E., Horton, R. & Ren, T. 2001. A new perspective on soil
thermal properties. Soil Science Society of America Journal, 65,
16411647.
Raiche, A. 1974. An integral equation approach to three-dimensional
modelling. Geophysical Journal of the Royal Astronomical Society,
36, 363376.
Ratcliffe, E.H. 1960. The thermal conductivities of ocean sediments.
Journal of Geophysical Research, 65, 15351541.
Ren, T., Noborio, K. & Horton, R. 1999. Measuring soil water con-
tent, electrical conductivity and thermal properties with a thermo-
time domain reflectometry probe. Soil Science Society of America
Journal, 63, 450457.
Sheng, P. 1991. Consistent modelling of the electrical and elastic
properties of sedimentary rocks. Geophysics, 56, 12361243.
Singh, D.N. & Devid, K. 2000. Generalized relationships for estimat-
ing soil thermal resistivity. Experimental Thermal and Fluid Science,
22, 133143.
Smith, W.O. 1939. Thermal conductivities in moist soils. Soil Science
Society of America Proceedings, 4, 3240.
Smith, W.O. & Byers, H.G. 1938. The thermal conductivity of dry soils
of certain of the great soil groups. Soil Science Society of America
Proceedings, 3, 1319.
Tabbagh, A. 1976. Les proprie te s thermiques des sols, premiers re sul-
tats utilisables en prospection arche ologique. Archaeo-Physika, 6,
128148.
Tabbagh, A. 1985a. The response of a three-dimensional magnetic and
conductive body in shallow depth electromagnetic prospecting. Geo-
physical Journal of the Royal Astronomical Society, 81, 215230.
Tabbagh, A. 1985b. A new apparatus for measuring thermal proper-
ties of soils and rocks in-situ. IEEE Transactions on Geosciences and
Remote Sensing, 23, 896900.
Tabbagh, A. & Jolivet, A. 1974. Proce de de mesure in-situ des
proprie te s thermiques des sols. Science du Sol, 4, 269279.
Tabbagh, A., Camerlynck, C. & Cosenza, P. 2000. Numerical model-
ling for investigating the physical meaning of the relationship
between relative dielectric permittivity and water content of soils.
Water Resources Research, 36, 27712776.
Tabbagh, A., Panissod, C., Gue rin, R. & Cosenza, P. 2002. Numerical
modelling of the role of water and clay content in soils and rocks
bulk electrical conductivity. Journal of Geophysical Research,
107(B11) art. no 2318. DOI: 10.1029/2000JB000025.
Topp, G.C., Davis, J.L. & Annan, A.P. 1980. Electromagnetic deter-
mination of soil water content: measurement in coaxial transmission
lines. Water Resources Research, 16, 574582.
Van Duin, R.H.A. & de Vries, D.A. 1954. A recording apparatus for
measuring thermal conductivity, and some results obtained with it
in soil. Netherlands Journal of Agricultural Science, 2, 168175.
Von Herzen, R.P. & Maxwell, A.E. 1959. The measurement of thermal
conductivity of deep-sea sediments by a needle probe method.
Journal of Geophysical Research, 64, 15571563.
Woodside, W. & Messmer, J.H. 1961. Thermal conductivity of porous
media. I. Unconsolidated sands. Journal of Applied Physics, 32,
16881699.
Wyllie, M.J.R., Gregory, A.R. & Gardner, L.W. 1956. Elastic wave
velocities in heterogeneous and porous media. Geophysics, 21, 4170.
Zierfuss, H. & van der Vliet, G. 1956. Laboratory measurements of
heat conductivity of sedimentary rocks. Bulletin of the American
Association of Petroleum Geologists, 40, 24752488.
Thermal conductivity and water content of soils 587
#
2003 Blackwell Publishing Ltd, European Journal of Soil Science, 54, 581587

You might also like