Вы находитесь на странице: 1из 22

Thermal Conductivity of Rocks and Minerals

Christoph Clauser and Ernst Huenges

1. INTRODUCTION cient only above 1200 “C. However, with single crystals
and glasses (e.g. obsidian) radiation becomes important
The interior heat of the earth is transmitted to its surface from temperatures as low as 200-400 “C. For the usual
mainly by three mechanisms: radiation, advection, and con- range of crustal temperatures and temperature gradients a
duction. In the earth’s lithosphere conduction of heat gener- linearization of the radiation law yields a “radiative thermal
ally dominates amoung these mechanisms. However, there conductivity” which can be formally added to the coeffi-
are two exceptions: cient of lattice or phonon thermal conductivity in Fourier’s
law of heat conduction. Thermal conductivities determined
(1) If the hydraulic permeability of crustal material is
at very high temperatures in the laboratory always include
sufficiently high, convection driven advection of heat can
this radiative component. Radiative thermal conductivity
be an equally or even much more efficient transfer mecha-
will therefore not be treated separately here; a review of
nism, provided sufficiently strong driving forces are sup-
heat radiation in the earth is given by Clauser (1988).
plied by forced or free convection systems. This is often
the case in sedimentary basins. However, fluid driven heat
2. BACKGROUND
advection can be important also in crystalline rocks and on
a crustal scale (Etheridge et al., 1983, Torgersen, 1990,
Clauser, 1992). 2.1 Fourier’s law
Fourier’s law of heat conduction defines heat flow density
(2) At ambient temperatures above 600 “C radiation of
heat begins to contribute sizeably to the overall heat trans- qi, the vector of specific energy flow rate, as the product of
fer in most polycrystalline materials, but it is really effi- the thermal conductivity tensor hi, and the temperature
gradient vector aT/axj:

C. Clauser, Niedersachsisches Landesamt fur Bodenforschung,


qi = ;li j . g (1)
J
Geowissenschaftliche Gemeinschaftsaufgahen (NLfB-GGA),
Postfach 510153 D-30631 Hannover, Germany Temperature measurements are usually performed along
vertical profiles in boreholes. Therefore only the vertical
Present Address: C. Clauser, NLfB-GGA (Geological Survey),
component of the temperature gradient is generally known
Stillweg 2, D-30655 Hannover, Germany
from measurements. Thermal conductivity for many rocks
E.Huenges,NiedersilchsischesLandesamtf~rBodenforschung,
is, to a good approximation, isotropic, particularly for vol-
Kontinentales Tietbohrprogramm (NLfB-KTB), Postfach canic and plutonic rocks. In these cases heat flow will be
5 10 153 D-3063 1 Hannover, Germany predominantly vertical, and it is sufficient to consider only
the vertical component of (1). In contrast to this, thermal
Rock Physics and Phase Relations conductivity of many sedimentary and metamorphic rocks
A Handbook of Physical Constants is strongly anisotropic, and lateral heat flow will be sig-
AGU Reference Shelf 3 nificant. Hence information on anisotropy is often needed,

Copyright 1995 by the American Geophysical Union. 105


106 THERMAL CONDUCTIVITY OF ROCKS AND MINERALS

requiring laboratory measurements in different directions. are purely empirical.


Anisotropy exists on several scales : (1) On the microscopic
Estimation from mineral content and saturating
scale many minerals are anisotropic (Table 3). (2) On the
fluids. Thermal conductivity of rocks may be estimated
laboratory scale, the thermal conductivity of many rocks is
from their mineral content, as minerals, due to their well
also anisotropic. However, even if rocks are composed of
defined composition, exhibit a much smaller variance in
anisotropic minerals, random orientation of the crystals
thermal conductivity than rocks. Similarly, as a porous
within the rock may make the rock’s thermal conductivity
rock’s bulk thermal conductivity varies with different
appear isotropic macroscopically. (3) Still on a larger scale,
saturants, it may be of interest to know the thermal
if rocks are exposed to folding, orogeny or other tectonic
conductivity of a rock when it is saturated with other fluids
processes, the thermal conductivity of the resulting rock
than those used in the laboratory measurement. Numerous
formation may be either isotropic as or anisotropic.
models have been proposed for this, but all have their
disadvantages: some overestimate while others under-
estimate systematically the true bulk thermal conductivity.
2.2 Measurement techniques
Most of them are valid only for a specific range of volume
Thermal conductivity can be measured in the laboratory ratios (or porosities), and yield completely unreasonable
on rock samples, i.e. cores or cuttings, or in-situ in results outside this range. Parallel and series model are easy
boreholes or with marine heat flow probes. There are to understand, but have the disadvantage of being rather
numerous steady state and transient techniques available for special cases, applicable mostly to bedded sediments. They
measuring thermal conductivity, the most prominent being lead to the well known arithmetic and harmonic means,
the “divided bar” and the “needle probe” method. As these respectively, and define upper and lower limits for all other
methods are discussed in detail in several textbook and models. Thus they constrain the maximum variance of
review articles (Beck, 1965, 1988, Davis, 1988, Desai et possible predictions. Quite successful in describing the data
al., 1974, Kappelmeyer & Hanel, 1974, Roy et al., 1981, in many cases, but unfortunately without a clearly defined
Somerton, 1992, Tye, 1969), we will neither address them physical model, the geometric mean falls in between these
here again nor comment on the many details involved in two extremes. If h, is the thermal conductivity and ni the
performing the actual measurements. volume fraction of the i-th phase relative to the total
volume, with 1=&r,, these three means are defined by:
As is the case with most other petrophysical properties,
in-situ thermal conductivity may deviate significantly from
laboratory values, even if the effect of temperature, pressure
Aari = C n;J.,
and pore-fluid is accounted for. The reason for this problem
is a certain scale dependence in which different aspects are
involved: in-situ measurements, as a rule, represent an (2)
average over a much larger rock volume than laboratory
measurements performed on small samples. On the other
hand, small-scale variations may thus be lost. Which 3Lgee= rI 3Li”i
thermal conductivity is the “correct” one will depend on the
specific question. This problem is quite similar to one
encountered in hydrology: the difficulty of defining a
In this context it must suffice to present only these three
“representative elementary volume” for which sensible
most well known models, as this subject can be addressed
averages for transport parameters like permeability and
only briefly here. Beck (1988) reviews the topic in
dispersion lengths can be defined.
considerable detail, and, in particular, presents and
discusses several other well known mixing-models.
Somerton (1992) discusses unconsolidated sands, effects of
2.3 Indirect methods
multi-fluid saturation, and illustrates the topic with many
When no data are available or no direct measurements examples from hydrocarbon reservoirs. Horai (1991) tests
can be performed, thermal conductivity can be inferred the results of predictions from several different mixing-
from a number of indirect data: mineralogical composition models on a remarkable data set in which porosity virtually
and saturating fluids, well-log correlations, and correlations varies from O-100 %. As can be expected, most of the
with other physical parameters. While some of these models tested were valid only for certain porosity ranges.
methods are based on well defined physical models, others Only two more recent two-phase models, assuming that
CLAUSER AND HUENGES 107

pores can be treated as spheroidal inclusions in a limited to unfractured rocks, since the effects of fracturing
homogeneous and isotropic material, are capable of on compressional and shear velocities lead to inaccurate
explaining the complete data set (Horai, 1991). However, results. There are indications, however, that shear-wave
additional information on the spheroids’ aspect ratio or birefringence may pose a limit to the application of this
orientation, respectively, is required by these two models. method in foliated rocks as well (Pribnow et al., 1993).
Given the typical conductivity ratios we observe in
nature, i.e. < IO, most of the conductivity models work to
2.4 Sources of data
within 10-l 5 % accuracy. For larger ratios some break
down more than others, and the geometric mean is one of For a large number of rocks thermal conductivity data are
them. The reason why it is still quite popular with many, available and classified according to rock name and origin
even in extreme cases, is that it has often been found that in several extensive compilations (Birch, 1942, Clark, 1966,
errors introduced in the inverse problem (i.e. in predicting Desai et al., 1974, Kappelmeyer & Hanel, 1974, Roy et al.,
the matrix conductivity from measurements on samples 1981, Cermak & Rybach, 1982, Robertson, 1988). How-
saturated with one fluid) are automatically compensated for ever, it is important to realize that compilations for rocks
when using this incorrect matrix value in the subsequent are inevitably comprised of data that are heterogeneous in
forward calculation (i.e. in predicting the bulk conductivity many respects, such as mineral composition, porosity, satu-
of the matrix saturated with another fluid). ration, and experimental conditions. This is responsible for
the great variability of thermal conductivity for each parti-
Well-log correlations. There are three different ways in
cular rock. Therefore, the merit of a summary purely
which well-logs can be used to infer estimates for in-situ
according to rock type is limited if users are interested
thermal conductivity:
primarily in general questions of heat transfer in the earth
(1) One approach is to establish empirical relationships rather than in data from a specific location.
between thermal conductivity and parameters derived from
well logs, such as porosity, bulk density, sonic (p-wave)
velocity, and seismic travel times. In principle, this 2.5 Outline of this compilation
approach is not limited to well logs, if petrophysical para-
In this review we therefore take a complementary
meters are known from laboratory measurements, for
approach to those previous compilations. We do not attempt
instance. A useful summary of these different approaches is
to present a complete table of all available thermal
presented by Blackwell (1989), who also illustrates their
conductivity data published to date. Instead, we build on the
application to a specific case.
data compiled previously (Birch & Clark, 194Oa,b, Clark,
(2) The second approach is, in principle, an extension of 1966, Desai et al., 1974, Kappelmeyer & Hanel, 1974, Roy
the mixing-model approach to the borehoie scale: the et al., 1981, Cermak & Rybach, 1982, Robertson, 1988)
volume fractions of the different mineral (or fluid) phases and arrange them into four basic groups: sedimentary,
are either taken directly from induced gamma ray metamorphic, volcanic, and plutonic rocks.
spectroscopy logs (Williams & Anderson, 1990) or
Data on thermal conductivity of minerals is not quite as
determined from a joint analysis of other logs such as
abundant as for rocks. Both measurements on single crys-
gamma ray, sonic traveltime, gamma density, and neutron
tals and on mineral powder are reported in the literature. In
porosity (Demongodin et al., 1991). Then an appropriate
this review we present a summary of both types of data
mixing model is applied. Both approaches apply the
from original contributions and from previous compilations
geometric mean as mixing model and test their method on
by Birch & Clark (194Oa,b), Sass (1965), Clark (1966),
detailed data from two case-study boreholes. A limitation of
Horai & Simmons (1969), Horai (1971), Dreyer (1974),
both methods is that mineralogy-based conductivity models
Robertson (1988), and Diment & Pratt (1988).
cannot account for the effect of anisotropy observed in
many sedimentary and metamorphic rocks.
(3) In a third approach, Willams & Anderson (1990) 3. THERMAL CONDUCTIVITY OF ROCKS
derive a phonon conduction model for thermal conductivity,
which utilizes temperature, acoustic velocity, and bulk
density measurements from well-logs. The method is Inspection of any of the compilations quoted above
claimed to be accurate to within + 15 %, both in isotropic reveals that thermal conductivity may vary by as much as
and anisotropic formations. Its application, however, is a factor of two to three for any given rock type. This is due
108 THERMAL CONDUCTIVITY OF ROCKS AND MINERALS

to the natural variation of a rock’s mineral content as well Plutonic and metamorphic rocks display a much smaller
as to several physical and diagenetic factors. All rocks are porosity. Here the dominant mineral phase controls different
therefore arranged into the four basic groups characterizing conductivity distributions. For plutonic rocks the feldspar
the special conditions prevailing at their formation, content determines the nature of the histogram (Figure lc):
deposition, or metamorphism: sediments, volcanics, pluton- while rocks with a low feldspar content (i.e. less than about
its, and metamorphics. In each group we study statistical 60 %) seem to define a nearly symmetrical histogram, a
quantities (such as histograms, median, mean, and standard high content in feldspar biases the distribution towards low
deviation) and investigate the variation of thermal conductivities. Interestingly enough, means and medians for
conductivity with those factors that have the most both distributions are nearly identical within the given
pronounced effect on this group of rocks. These are standard deviation.
petrological aspects or petrophysical influences such as Metamorphic rocks may be classified according to their
porosity (in sediments and volcanic rocks), the dominant quartz content. Figure Id displays the resulting bimodal
mineral phase (in metamorphic and plutonic rocks), and distribution. While the low conductivity part is made up of
anisotropy (in sediments and metamorphic rocks). More rocks with low quartz-content, the high-conductivity portion
recent data was included, mainly (but not only) from the consists of quartzite only.
German continental deep drilling project KTB, when the
Influence of ambient temperature. Thermal
existing data base seemed insufficient for our statistical
conductivity is a function of temperature. Lattice (or pho-
approach. Where additional specific data is available, the
non) thermal conductivity varies inversely with temperature.
effect of temperature, pressure, saturation, and saturant is
As thermal expansions increases with temperature, but dif-
demonstrated.
ferently for all minerals, “thermal cracking” by differential
expansion may create contact resistances between mineral
3.1 Thermal conductivity of sedimentary, volcanic, grains, thus contributing to the observed decrease of con-
plutonic, and metamorphic rocks ductivity with temperature. This effect is probably not as
severe in water-saturated rocks as it is in dry rocks, the
Before any details are discussed we first provide an condition in which most laboratory experiments are
overview on the distribution of thermal conductivity in conducted. Conductivity-temperature determinations of
general as well as on the variation of thermal conductivity crystalline water-saturated rocks are now under way in
with ambient temperature for the four basic rock-types. some laboratories. The “radiative thermal conductivity”, in
Influence of porosity and the dominant mineral phase. contrast, follows a T3-law (see e.g. Clauser 1988). Thus
Figure 1 shows histograms for thermal conductivity measurements on thermal conductivity as function of
according to rock type. For sedimentary rocks (Figure la) increasing temperature generally show initially a decrease
the controlling factors on thermal conductivity are porosity with temperature, until around 1000-1200 “C the radiative
and origin of a particular sediment. It appears as if component balances and sometimes even inverts this
chemical sediments, mainly formed by precipitation of decreasing trend.
dissolved minerals or by compaction of organic material, Figure 2a shows this effect for sediments. Up to 300 ‘C
and low porosity (c about 30 %) physical sediments, there is a reduction by nearly a factor of two, both for
formed by the compaction and cementation of elastic elastic and carbonaceous sediments. Above 300 “C the de-
material, have nearly identical frequency distributions, crease in thermal conductivity comes nearly to an end, with
means, and medians. In contrast, high porosity (> about 80 carbonates decreasing still a little more than elastic sedi-
%), mainly marine physical sediments display a distribution ments. However, as there are very few data for this temper-
which is biased towards low conductivities, with mean and ature range, this last observation is not very sound
median about half the size of the former two. This, of statistically.
course, is due to the low-conductivity fill of the void space,
Volcanic rocks (Figure 2b) display quite a different
which can be either air or water.
behaviour, depending on their opacity, i.e. on how well
For volcanic rocks (Figure lb), spanning nearly the total they transmit thermal energy by radiation. Due to this
possible range of porosity from O-1, porosity is again the additional “radiative thermal conductivity”, volcanic glasses
controlling factor on thermal conductivity: mean and and rocks with a small iron content experience an increase
median of the high- and low-porosity histograms differ by in thermal conductivity for temperatures above 800-l 000 ‘C
nearly a factor of two, and the high porosity distribution is (see e.g. Clauser, 1988). In contrast, conduction dominated
clearly skewed towards low conductivities. rocks show a much more pronounced decrease in thermal
Fig. 1. Histograms of thermal conductivity for sedimentary,
im n
volcanic, plutonic, and metamorphic rocks. All data, taken
12 10 04
from [ 141, were measured at room temperature and atmo-
. . . . . . . .
. . . . . . . .
spheric pressure. n is the number of data, m the median, p
the mean, and o the standard deviation for all three data
sets. Please note that superposition of different domains
results in new hatchure styles in some diagrams.
(a). Thermal conductivity of sedimentary rocks, subdivid-
ed according to chemical or physical sedimentation proces-
0 5 I 5 25 3.5 4.5 5.5 ses. Histogram for chemical sediments inc!udes data for
h (\\ IN I<-')
limestone, coal, dolomite, hematite, chert, anhydrite, gyp-
sum, rock salt, and sylvinite. For physical sediments the
influence of porosity is considered additionally: low porosi-
FZBB highporosrly 92
Esssxss low porosity 234
ty sediments include data from shale (including dolomitic,
29 32 07 pyritic, and carbonaceous shale), marl, clayey marl, marl-
stone, conglomerate, tuff-conglomerate, tuffite, breccia,
quartz breccia, and sandstone (including limy and quartz
sandstone), while high-porosity sediments are ocean- and
lake-bottom sediments.
(b). Thermal conductivity of volcanic rocks, subdivided
0 according to porosity. The high porosity histogram repre-
05 15 3.5 4.5 :5.:
B h (W m-’ K-‘) sents data fom lava, tuff, tuff breccia, and mid-ocean ridge
basalt (MORB). Low porosity data are from rhyolite, lipa-
rite, trachodolerite, andesite, and basalt (excluding MORB).
-- (c). Thermal conductivity of plutonic rocks, subdivided
s 70 according to feldspar content. Histogram for high feldspar
2 60 content (i.e. more than about 60 %) is made up of data
F 50 from syenite (including alkali and nepheline syenite), grano-
5 10 syenite, syenite porphyry, and anorthosite. Data for vari-
: 30
able, but generally lower feldspar content (i.e. less than
20
about 60 %) are from granite (including alkali granite,
10
plagiogranite, granodiorite, tonalite, quartz monzonite),
0
05 10 15 20 25 30 3.5 4.0 4.5 5.0
quartz- and quartz-feldspar-porphyry, diorite (including
C X (W m‘l K-‘)
monzonite), gabbro (including quartz and olivine gabbro),
100
porphyrite dykes (lamporphyre, diabase, quartz dolerite),
and ultramafic rocks (pyroxenite, peridotite, lherzolite,
hypersthenite, bronzitite, dunite, olivinite, homblendite,
cumberlandite).
(d). Thermal conductivity of metamorphic rocks, sub-
divided according to quartz content. Histogram for high
quartz content is made up of data fom quartzite. Data for
low quartz content are from quartz-mica schist, gneisses,
marble, serpentinite, talc, serpentinized peridotite, homfels,
0 eclogite, albitite, leptite, schist, slate, phyllite, amphibolite,
1 2 3 4 5 6 7
D h (W m-l K-l) mylonite and greenstone.
110 THERMAL CONDUCTIVITY OF ROCKS AND MINERALS

conductivity. An inversion of this decrease cannot be seen


with statistical confidence. However, above around 900 “C *CT) = A + 35oB, T ’
this trend seems to come to a halt, with thermal conductivi-
ty on a level of about 50 % of the room-temperature value. where h is given in W rn-’ K-‘, T in “C, and the empirical
constants A and B are determined from a least-squares fit
Again, there are quite few data points above 500 “C.
to measured data for different rock types (Table 1). Linear
There does not seem to be a very significant radiative relationships between temperature and the inverse of h, the
contribution in plutonic rocks (Figure 2~). However the thermal resistivity, discriminate between temperature-
decrease of thermal conductivity with temperature is quite dependent contributions and other factors, which are
different, depending on the feldspar content: while there is independent of temperature (such as micro-cracks, grain
hardly any significant decrease (- 10 %) in conductivity up boundaries, pore volume, as well as mineralogical composi-
to 300 “C for rocks that are rich in feldspar, rocks that are tion, shape and orientation of crystals and their fragments):
poor in feldspar decrease by more than 40 % over this
range. Above this temperature the decrease is more gentle,
1
spreading an additional 20 % over the next 700 K. -=D+E.T,
Interestingly, there is a large amount of data available for KQ
this high-temperature range. The different behaviour of
where h is again in W rn-’ K“ and T is in K. By measuring
rocks with a high feldspar content is due to the increase in
h and plotting the thermal resistivity versus temperature D
thermal conductivity with temperature of some plagioclase
and E may be determined as intercept and slope of a linear
feldspars (Birch & Clark, 1940a) which compensates the
regression. Buntebarth (1991) determined D and E from
decrease in thermal conductivity with temperature observed
measurements on 113 samples of metamorphic rocks from
for most other minerals and rocks (other notable exceptions
the KTB borehole (mostly gneisses and metabasites) in the
are fused silica as well as volcanic and silica glasses; see
temperature range 50-200 “C. The arithmetic means of 66
also discussion of empirical relationships below).
individual values for D and E determined for gneiss are D
For metamorphic rocks, the decrease of thermal conducti- = 0.16 _+0.03 m K W-’ and E = 0.37 _+0.14 . 10-j m W-‘.
vity with temperature depends on the content in a dominant The corresponding means of D- and E-values determined
mineral phase, similar to plutonic rocks. Quartzites decrease on 36 metabasite samples are D = 0.33 + 0.03 m K W“ and
rapidly, by nearly a factor of three up to a temperature of E = 0.22 _+0.14 . 10.’ m W-‘. Sass et al. (1992) likewise
about 500 “C. Above this, there is only a very mild further distinguish between the effects of composition and
decrease. For rocks that are poor in quartz the decrease in temperature on thermal conductivity. They propose a quite
conductivity is not quite as dramatic, amounting to about general empirical relation for h(T), the thermal conductivity
one third of the room-temperature value up to 200 “C. Then in W rn-’ K-’ at temperature T in “C as a function of h(25),
it remains roughly constant up to 500 “C. Above this, up to the measured room-temperature thermal conductivity:
750 “C, it decreases again to about one third of the room-
temperature value. There are again many data available for
this high-temperature range, at least for mafic rocks. a(T) = a(0)

Often data on thermal conductivity is available for room- 1.007 + T . (0.0036 - o.0072 )
w
temperature conditions only, even though it is required at
elevated temperatures. For this purpose we will discuss
some empirical relationships that have been proposed for where PC)
extrapolation on the basis of data measured at elevated tem-
peratures. It is emphasized, however, that there is no real
substitute for individual measurements.
a(0) = a(25) * [ 1.007 + 25 * (0.0037 - ygf)l.
It has been long recognized that for moderate tempera-
tures h varies inversely with temperature (Birch & Clark, Equation (3~) is derived from the classical experimental
194Ob). For this temperature range several approaches have data set of Birch and Clark (194Oa,b), who measured
been suggested as how to infer thermal conductivity at ele- thermal conductivity as function of temperature in the range
vated temperatures. Based on the analysis of available tabu- O-200 “C and higher on 38 samples from a large suite of
lated data of thermal conductivity as function of tempera- materials including volcanic, metamorphic, plutonic and
ture Zoth & Hanel(1988) suggest a relationship of the form sedimentary rocks. Their results for granites clearly show
0 100 200 300 400 500 600 700 800

J
0 100 200 300 400 500 600 700 800
‘I’ (“C) T (T)

Fig. 2. Variation of thermal conductivity with temperature for various rocks. Two groups of data are
considered in each plot, and for various temperature ranges (half-way up and down to the next reference
temperature) median, mean, and standard deviation are computed. Full line, big symbols, and shading
correspond to means and standard deviations, broken line and small symbols to medians. The inset
illustrates the number of data available in different temperature ranges. Measurements were performed
either with a divided bar or a line-source apparatus in dry condition and at atmospheric pressure. Please
note that superposition of different domains results in new hatchure styles in some diagrams.
(a). Sedimentary rocks. Two curves are shown for carbonates (limestone and dolomite) and elastic
sediments, i.e. (quartz) sandstone and shale. Data taken from [23] and [ 141.
.(b). Volcanic rocks. Two curves are shown for rocks with weak (basalt, rhyolite (also altered or
porphyritic), dacite, tuff) and strong radiative component (basalt glass, obsidian, diabase dolerite). Data
taken from [23] and [ 141.
(c). Plutonic rocks. Two curves are shown for rocks which are rich in feldspar (syenite, anorthosite,
hypersthenite) and poor in feldspar (granite, alkali granite, diorite, quartz diorite, monzonite, tonalite,
gabbro, hornblende gabbro, peridotite, lherzolite, bronzitite, dunite, olivinite, granodiorite). Data taken
from [23], [ 141, and [46].
(d). Metamorphic rocks. Two curves are shown for quartzites and for rocks which are poor in quartz
(marble, serpentinite, eclogite, albitite, slate, amphibolite). Data taken from [23], [14], and [46].

the coupled effect of composition and temperature, as the ture range of O-250 “C for rocks ranging in composition
normalized thermal resistivity h(O)/h(T) is a linear function from felsic gneiss to amphibolite: in spite of some slight
of temperature whose slope increases with h(O), the con- systematic differences, the deviations between measured
ductivity at 0 “C. Sass et al. (1992) report a successful test values and predictions made on the basis of equation (3~)
of equation (3~) on an independent data set over a tempera- were well within the range attributable to experimental
112 THERMAL CONDUCTIVITY OF ROCKS AND MINERALS

TABLE 1. Constants A and B from equation (3a) for conductivities of about 0.6,0.12-0.17, and 0.025 W rn-’ K-‘,
different rock types.” respectively (Grigull & Sandner, 1990). For plutonic rocks,
figure 4a demonstrates this effect for air and water on a
rock type -I- (“C) A B remarkable data set of Hawaiian marine basalt, that nearly
spans the total possible range of porosity from O-l. For
(1) rock salt -2o- 40 -2.11 2960
sedimentary rocks Figure 4b illustrates this for quartz
(2) limestones o- 500 0.13 1073 sandstones saturated with air, oil, and water. In both cases
the resulting bulk conductivity behaves according to the
(3) metamorphic rocks o- 1200 0.75 705 saturant’s thermal conductivity. Additionally, contact
(4) acid rocks O-1400 0.64 807 resistances during measurements on dry rock samples will
also reduce thermal conductivity.
(5) basic rocks 50-l 100 1.18 474
Partial saturation. The effect of partial saturation varies
(6) ultra-basic rocks 20- 1400 0.73 1293 depending whether the rock is porous or fractured. Porosity
(7) rock types (2)(5) 0- 800 0.70 770 in porous rocks consists of “bottlenecks” formed at the
contact between individual grains and the bulk pore space.
“after [57]
Dry bottlenecks act as thermal contact resistances between
grains, while the bulk pore volume contributes according to
error. This suggests that equation (3~) yields useful its size to the effective thermal conductivity. Figures 5a-5c
estimates of the temperature dependence of thermal conduc- illustrate how both types of pore space influence thermal
tivity for crystalline rocks, independent of mineralogy. conductivity under partially saturated conditions. Figure 5a
shows how thermal conductivity varies in sandstones of
low- to medium-porosity with the degree of oil saturation.
3.2 Influence of various factors for selected rock types Initially, there is a rapid increase in conductivity with
saturation: starting from completely unsaturated conditions,
Apart from temperature, thermal conductivity also varies where conductivity reaches only about 80 % of the
with pressure, degree of saturation, pore fluid, dominant saturated value, 90 % is reached at a saturation level of
mineral phase, and anisotropy. about 10 %. The remaining 10 % deficit in conductivity is
Pressure. The effect of overburden pressure is twofold, spread rather evenly over the remaining 90 % of saturation.
different for two distinct pressure ranges. First, fractures Fig. 5b illustrates these two effects for water-saturation in
and microcracks developed during stress release, when the a medium-porosity sandstone. The behavior is quite similar
sample was brought to the surface, begin to close again to the preceeding case: starting from a completely
with increasing pressure. This reduces thermal contact unsaturated conductivity of only about 60 % of the
resistance as well as porosity, which is usually filled with saturated value, 80 % is reached again at a saturation level
a low conductivity fluid. When an overburden pressure of of only about 10 %. The remaining 20 % deficit in
about 15 MPa is reached, this process comes to an end. A conductivity is made up for during the remaining 90 % of
further pressure increase to 40 MPa does not affect thermal saturation. Physically this observation indicates that the
conductivity significantly (Figure 3a). If pressure is still filling of intergranular bottlenecks, which accounts for only
further increased the second effect becomes apparent, the about 10 % or 20 % of the total porosity, respectively,
reduction of the rock’s intrinsic porosity, i.e. that which is significantly reduces contact resistance between individual
not artificially created by stress release. Figure 3b illustrates grains. Replacing low conductivity air by the higher
this effect for granite and for metamorphic rocks. While conductivity fluid in the major part of the pore volume
both curves indicate about a 10 % increase over the total accounts for the second effect.
range of pressures from O-500 MPa, the increase is stronger
If only fractures contribute to the total porosity, such as
over the first 50 MPa due to the first effect discussed
in crystalline rock, the pore space consists of bottlenecks
previously.
only, and we observe the first effect alone. This is
Porosity and saturating fluid. If porosity is important illustrated in Figure 5c for a granite of 1 % porosity.
(i.e. >> 1 %) the saturating fluid’s thermal conductivity may Starting from completely unsaturated conditions, with only
significantly affect the bulk thermal conductivity of the about 80 % of the saturated conductivity, there is a more or
saturated rock. Results are shown for three low conductivity less linear increase until 100 % is reached for complete
saturants, water, oil, and air with room-temperature saturation. In contrast, porous rocks with a considerable
CLAUSER AND HUENGES 113

1.3

12

;L
q 11
r:

1 0

09
0 10 20 30 40 200 300
A 1’ (MPa) B I-’ (MPa)

Fig. 3. Variation of thermal conductivity with uniaxial pressure for various rocks. For several temperature
ranges (half-way up and down to the next reference temperature) median, mean, and standard deviation
are computed. Full line, big symbols, and shading correspond to means and standard deviations, broken
line and small symbols to medians. The inset illustrates the number of data available in different
temperature ranges.
(a). Anhydrite, sandstone, dolomite, limestone, porphyry, diabasic basalt, basaltic lava, and granite. In
order to make results for such a diverse group of rocks comparable, thermal conductivity values are
normalized relative to the thermal conductivity measured at the lowest reported pressure level for each
specimen (O-4 Mpa). All data, taken from [23], [14], and [46], were measured either with a divided bar
or a line-source apparatus in dry condition.
(b). Two groups of crustal rocks. Two curves are shown, the upper curve for granite and the lower
curve for predominantly metamorphic rocks (amphibolite, serpentinite as well as gabbro). All data (taken
from [46] and Seipold: written personal communication 1993) were measured in dry condition with a
line-source apparatus.

. .
00
OL / I
0 10 20 30 40 50 60 70 80 90 100 0 10 20 30 40 50 60
A porosity (%) B porosity (%)

Fig. 4. Variation of thermal conductivity with porosity.


(a). Hawaiian basalt. All data, taken from [40], were measured with a divided-bar apparatus at 35 “C
and under 3 MPa of uniaxial pressure, both in dry and in fully saturated condition.
(b). Quartz sandstone. All data ([36], taken from [23]) were measured with a line-source apparatus at
room temperature and atmospheric pressure in air-, oil-, and water-saturated condition.
114 THERMAL CONDUCTIVITY OF ROCKS AND MINERALS

Fig. 5. Variation of thermal conductivity with p‘artial satura-


tion.
(a.) Sandstone (3-30 % porosity). Thermal conductivity
values are normalized relative to the thermal conductivity
x measured at 100 % saturation. The inset illustrates the
number of data available at different levels of saturation.
For each saturation range (half-way up and down to the
next reference point) median, mean, and standard deviation
are computed. Full line, big symbols, and shading corres-
07 I
0 10 20 30 40 50 60 70 80 90 100 pond to means and standard deviations, broken line and
A oil saturation (%) small symbols to medians. All data ([36], taken from [23])
were measured with a line-source apparatus at room tem-
perature and atmospheric pressure, both dry and partially
saturated with Soltrol “C”, an oil from Phillips Petroleum
Co. (Bartlesville, Ok.).
(b). Sandstone (18 % porosity). All data, taken from [39],
were measured with a half-space line-source apparatus at
room temperature and atmospheric pressure, both dry and
T<
- 0.7
partially saturated with water. The degree of saturation was
30 4 determined by weighing. Squares are means and vertical
- 0.6 lines are standard deviations determined from multiple
0 10 20 30 40 50 60 70 60 90 100
B water saturation (X) measurements.

32 (c). Granite (1 % porosity). All data, taken from [39],


r I 1.0 were measured with a half-space line-source apparatus at
3 1
room temperature and atmospheric pressure, both dry and
7 30 partially saturated with water. The degree of saturation was
x In determined by weighing. Squares are means and vertical
7 29 .+
c lines are standard deviations determined from multiple
; 28 3.g 5 measurements.
x 27

26

0 10 20 30 40 50 60 70 80 90 1 9"
C wdtcr saluration (%)

amount of bulk pore volume, display this linear between the two histograms which is due to the different
conductivity increase within the first 10 % of saturation. mineral content, Figure 6a illustrates the effects of
Dominant mineral phase and anisotropy. The variation anisotropy: measurements on both rock types were
of thermal conductivity with the dominant mineral phase performed parallel and perpendicular to the apparent
was previously discussed in the context of plutonic and direction of foliation. While for amphibolite the means,
metamorphic rocks in general (Figures lc and Id). Figure medians and histograms are nearly identical for either
6a demonstrates this dependence for two particular direction, this is not the case for gneiss. Here the means
metamorphic rock types, a low-conductivity amphibolite, and medians differ by about 20 %, and the histograms are
and a high conductivity gneiss. Apart from the obvious shift skewed towards lower values for measurements
CLAUSER AND HUENGES 115

nh
v. 1 3 4 5 6 -60 -:!I0 0 30 60 90
A h (W me’ K-‘) B rixi rnul 11 (dr,gree)

Fig. 6. Thermal conductivity of two metamorphic rocks from the KTB-VB pilot hole, an amphibolite,
consisting mostly of low-conductivity, basic minerals and a felsic gneiss, made up mostly of high-
conductivity quartz. n is the number of data, m the median, p the mean, and o the standard deviation for
both data sets. All samples were measured with a half-space line-source apparatus at room temperature
and atmospheric pressure, in nearly fully water-saturated condition, i.e. 70-90 % [31]. Please note that
cross-hatchure results from superposition of different domains.
(a) The effect of anisotropy and of the dominant mineral phase on thermal conductivity. Components
parallel and perpendicular to foliation are determined from knowledge of dip and azimuth of foliation.
(b). Variation of thermal conductivity with direction of heat flow relative to foliation for a dry (broken
line) and an almost fully water-saturated (70-90 %) sample (full line). The gneiss sample (squares) is
from a depth of 1908.7 m, the amphibolite sample (circles) from 3839.5 m.

perpendicular and towards higher values for measurements interpolation formula that accounts for the effects of both
parallel to foliation. water- or air-filled porosity and variable mineral content:
This directional dependence is further illustrated for two
particular rock samples of amphibolite and gneiss in Figure
h = 3Lf+y2q(3LS+p*S)-a.J, (4)
6b. It is quite apparent that thermal conductivity for the
same rock sample may vary from 100 % (parallel) to about
60 % (perpendicular), depending on the azimuth of the where h, is the pore fluid’s thermal conductivity, h, the
measurement relative to the foliation. In contrast to this and solid rock intercept at $ = 1 for zero per cent specific
in spite of an existing slight anisotropy, the variation of
mineral content, p the actual percentage of the specific
thermal conductivity in the amphibolite sample is less than mineral, and S a slope constant equal to the change of h
about f 5 %. It is interesting to note that the amount of
with specific mineral content, determined from intercept
anisotropy is identical irrespective of the state of saturation
values obtained from experimental data at ‘$ = 1. Table 2
of the sample. In contrast to seismic velocity this indicates
lists some of the data reported by Robertson (1988) which
that anisotropy of thermal conductivity does not seem to be
may be inserted into equation (4) to obtain estimates of
influenced by the pore-space or fracture geometry and the
thermal conductivity as function of porosity, pore-fluid, and
saturation of this rock. This is valid as long as the fluid’s
mineral content for mafic and felsic igneousrocks, and for
thermal conductivity is less than the rock’s.
sandstones.
Robertson (1988) discusses an empirical approach which
permits to account for the combined effects of porosity,
saturating fluid and dominant mineral phase. Plotting meas- 4. THERMAL CONDUCTIVITY OF MINERALS
ured thermal conductivities of various rocks versus the
square of solidity y (where y is l-porosity), he finds linear
relationships whose slopes vary with the per cent content in Thermal conductivity of minerals is much better con-
a specific mineral (e.g. quartz, olivine, etc). He proposes an strained than that of rocks, as there is a well defined
116 THERMAL CONDUCTIVITY OF ROCKS AND MINERALS

TABLE 2a. Constants h,, h,, and S from equation (4) for mafic igneous rocks.b

solidity y (-) pore fluid olivine content D (%) h (W rn-’ K-‘) S (W me’ Km’ %-‘)

0 air 0 &- = 0.188

1 air 0 h, = 1.51

1 air 30 h, = 1.96 0.015

0 water 0 h, = 0.75

1 water 0 h, = 1.84
1 water 30 h, = 2.60 0.025

bafter [41]; determined on tholeitic basalt samples with 0 - 40 % olivine content [40]

TABLE 2b. Constants &., h,, and S from equation (4) for felsic igneous rocks.“

solidity y (-) pore fluid quartz content p (%) h (W me’ K-‘) S (W rn-’ K“ %-I)

0 air 0 h, = 0.026
1 air 0 h, = 1.47
1 air 100 h.. = 5.23 0.038

‘after [41]; determined on samples with 0 - 45 % quartz content [5, 8, 91

TABLE 2c. Constants h,, h,, and S from equation (4) for sandstone.d

solidity y (-) pore fluid quartz content p (%) h (W rn-’ K-‘) S (W mm’ K“ %-I)

0 air 0 h, = 0.026
1 air 0 h, = 1.47

1 air 100 li, = 5.23 0.038


0 water 0 hf = 0.62

1 water 0 h, = 1.52

1 water 100 h. = 8.10 0.038

dafter [4l]; determined on samples with 0 - 100 % quartz content [I, 2, 8, 9, I I, 12, 15, 32,
35, 48, 49, 50, 54, 561
CLAUSER AND HUENGES 117

specific crystal structure and chemical formula for each in Dreyer (1974) are averages over an unspecified number
mineral. However there are two specific principal diffi- of individual measurements. The data in Horai (1971)
culties associated with the measurement of thermal conduc- consist of individual needle-probe measurements on water-
tivity on mineral samples: purity and sample size. Lattice saturated mineral powder. Ambient conditions are specified
imperfections in crystals significantly decrease the thermal as “ordinary temperature and pressure” by Horai (197 1) and
conductivity of many minerals. Correction of alien mineral as “room temperature” by Dreyer (1974). Diment & Pratt
phases in samples is possible (Horai, 1971), but requires (1988) quote specific temperature and pressure conditions
further microscopic and X-ray examination of the samples. during measurement for most of the data they report.
If measurements on single crystals or monomineralic, Comparing the data in the first two columns, there is a
polycrystalline aggregates are performed with a divided-bar good general agreement. However there are exceptions,
(or comparable) method, a minimum sample size is when lattice imperfections lead to a significant decrease in
required. Large single crystals that can be machined to the thermal conductivity, as can be seen, for instance, in arti-
desired size, however, are relatively rare. When mono- fical vs. natural periclase, halite and rock salt, as well as in
mineralic aggregates are used instead, uncertainty is chlorite, magnetite, corundum, pyrite and fluorite.
introduced by porosity. Alternatively, Horai & Simmons
Aggregate data from Diment & Pratt (1988) and Dreyer
(1969) and Horai (197 1) use a needle-probe technique to
(1974) are in generally good agreement with those from
measure the conductivity of finely ground samples of
Horai (1971). There are exceptions, though, that can be
minerals saturated with water. This way sample size poses
related to the technical details of the sample preparation
no problem, but all information on anisotropy is lost.
and the measurement technique. Horai’s (1971) value for
Moreover, the interpretation of measurements of thermal
talc, for instance, is twice as high as the two other
conductivity on fragments is not without ambiguity, as
aggregate data, but does not quite reach the parallel
pointed out by Sass et al. (197 I): their comparison between
conductivity value. This is probably due to an oriented
Horai & Simmons’ (1969) transient needle-probe method
sedimentation of the powdered talc particles in the saturated
and their steady-state divided-bar “cell” method on splits
sample compartment. Thus a needle probe measurement,
from Horai & Simmons’s (1969) original mineral samples
which produces radial heat flow, will preferentially sample
indicates that the results obtained from measurements on
the component of thermal conductivity which is parallel to
fragments depend on both the measurement technique and
the sheet-like crystal structure of the talc particles. A
the mode1 used for inferring the thermal conductivity of the
similar effect, however not quite as pronounced, can be
solid constituents of the mixture.
observed with serpentine.
Table 3 presents both kinds of data, measured both on
As with rocks, data on the temperature dependence of
single crystals as well as on natural monomineralic,
thermal conductivity for minerals not very abundant. Yang’s
polycrystalline aggregates, and on artificial monomineralic
(198 1) temperature dependent data for rock salt represent
aggregates produced from a mixture of powdered mineral
“recommended values” based on a great number of indivi-
specimens and distilled water. Data from three sources are
dual determinations and cover the temperature range 0.4-
presented: (1) Diment & Pratt (1988) who report their own
1000 K. Table 4 lists thermal conductivity and thermal
measurements as well as those performed or reported
diffusivity function of temperature for some rock-forming
previously by Birch (1942, 1954), Birch & Clark (194Oa,b),
minerals as reported by Kanamori et al. (1968). Table 5,
Clark (1966), Coster (1947), Ratcliffe (1959), Robertson
finally, provides values for the constants D and E from
(1988), and Sass (l965), (2) Dreyer (1974), a compilation
equation (3b), which may be used for inferring the
of data measured by a variety of researchers, and finally (3)
temperature dependence of h-‘, the thermal resitivity, for
Horai (1971), including Horai & Simmons (1969). Addi-
some monomineralic aggregates (Clark, 1969). Diffusivity
tionally, recommended values for rocksalt as function of
and conductivity are related by
temperature reported by Yang (198 1) are included in the
first group, as well as an aggregate value for graphite from a
K- 3
Grigull & Sandner (1990) and data on artificial periclase P *cp
measured by Kanamori et al. (1968) in the second one. The
data from Diment & Pratt (1988) represent measurements where p is density and c, specific heat capacity at constant
on individual samples; averages and standard deviations are pressure. Therefore the information on diffusivity is
computed, when several values are available for the same included here as well, although diffusivity is not further
mineral and for comparable P,T-conditions. Data reported discussed in this review. On the basis of equation (5)
118 THERMAL CONDUCTIVITY OF ROCKS AND MINERAL8

TABLE 3. Thermal conductivity h (W rn-’ Km’) of different rock-forming minerals.’

mineral T, state, 3L, (n) [24] state, 3L [25] state, 3L, (n) [28]

MISCELLANEOUS
graphite, C h, ,: 355.0, hg89.4,
a: 155.0 [27]
diamond, iso A,,: 545.3

Pyrex 774 glass T/ 1, amorphous: 011.21, 5Oj1.26,


(T in “C) lOOl1.32, 15Ol1.38, 20011.44, 25011.49,
3OOA.55, 35OA.61, 4OOA.66, 45Oi1.72,
500/1.83 [8]

ice (OOC), H,O A,,: 1.9, kx3: 2.3,


a: 2.0 (-125 “C: 4.0)

ORTHO- AND RINGSILICATES


olivine group
forsterite, MgJSiO,] 30 “C, a: 4.65 + 0.33 (3) [8] a: 5.03 f 0.18 (5)
(dunite, 97 % Fo,,Fa,) Wd%-Fo,,Fq)
fayalite, FeJSiO,] 30 “C, a: 3.85 f 0.07 (4) a: 3.16
(dunite, mostly Fa) (FOP%,,)

garnet group (Mg,Fe,Mn,Ca),(Al,Fe),[SiO,/,


almandine, iso ? “C, a: 3.56 h,,: 3.6 a: 3.31

grossularite, is0 ? “C, a: 5.32 a: 5.48 + 0.21 (3)

zircon group
zircon, ZrSiO, A,,: 3.9, is,: 4.8 a: 5.54

titanite group (sphene)


titanite CaTi[SiO,][OH] a: 2.34

AI,SiO, group

kyanite 35 “C, a: 7.15 + 0.14 (4) a: 14.16


35 “C, a: 12.45 f 0.58 (3)
(< 5 % quartz impurity)
andalusite 35 “C, a: 6.56 + 0.42 (8) a: 7.58
(<5 % quartz impurity)

sillimanite 35 “C, a: 10.73 f 0.52 (3) a: 9.10


CLAUSER AND HUENGES 119

TABLE 3. (continued)

mineral T, state, 1, (n) [24] state, 3L [25] state, A, (n) [28]

epidote group

epidote, 31 "C,I: 3.10


Ca,(A1,Fe),Si,O,,[OHl 32 "C, 11:2.93
3 I “C, a: 2.50 f 0.02 (2) a: 2.83 f 0.21 (2)

CHAIN SILICATES
pyroxene group (Na,Ca)(Mg,Fe~l)(Al,Si),O,
enstatite a: 4.47 f 0.30 (4)
diopside, augite 35 “C, a: 4.23 + 0.05 (4) a: 4.66 f 0.31 (4)

jadeite 34 “C, a: 5.59 + 0.86 (2) a: 5.64 f 1.02 (2)

amphibole group NaCa,(Mg,FePI)(Al,Si),O,,(OH),


hornblende 20 “C, a: 2.91 f 0.09 (2) [ 161 2.81 + 0.41 (2) a: 2.81 f 0.27 (2)

SHEET SILICATES
mica group

muscovite, 30 "C, 11:3.89 (2) a: 2.28 k 0.07 (3)


KA~,[A~~~,O,,,IPW, 32-45 "C,I: 0.62 k 0.11 (4)
biotite, K(Mg,Fe,..),Al, 33 "C, II: 3.14 a: 2.02 f 0.32 (2)
[AWJ,,IO-W, 32 "C, I: 0.52 f 0.01 (2)
talc, WGX,W~OW, 29-34 "C, II: 10.69 f 1.35 (5)
30 "C, -I-: 1.76 f 0.00 (2)
30 “C, a: 2.97 [7] a: 3.0 f 0.1 (2) a: 6.10 f 0.90 (2)
chlorite 30 “C, a: 3.06 k 1.18 a: 5.25 f 0.15 (2) a: 5.15 f 0.77 (3)
serpentine, 32 "C, I: 2.41 f 0.10 (2)
%W,%,1KW, ? "C, (I: 2.76 + 0.03 (4)
30-34 “C, a: 2.61 f 0.38 (10) a: 2.33 + 0.21 (3) a: 3.53 f 1.28 (3)

FRAMEWORK SILICATES
feldspar group

orthoclase, K[AISi,O,] 30 "C, (100): 2.34 f 0.08 (2) [43] a: 2.31


30 “C, (010): 2.68 [43]
30 “C, (001): 2.30 + 0.21 (2) [43]
microchne, K[AISi,O,] ? “C, (001): 2.04 (431 a: 2.49 f 0.08 (3)
albite, Na[AlSi,O,] 25 “C, a: 2.34 [43] a: 2.0 f 0.1 a: 2.14 + 0.19 (4)
anorthite, Ca[AlSi,O,] 25 “C, a: 2.72 [43] a: 2.1
120 THERMALCONDUCTIVITYOFROCKSANDMINERAJS

TABLE 3. (continued)

mineral T, state, A, (n) [24] state, h [25] state, A, (n) [28]

silica group, SiO,


ci quartz 30 T, I: 6.15 [38], (I: 10.17 [8] A,,: 6.5, hj3: 11.3 a: 7.69

c1 quartz I Tl h, x: 016.82, 5015.65, lOOl4.94,


(T in “C) 15014.44, 20014.06, 25013.73, 30013.52,
350/3.31 [8]
a quartz I/ T/A, x: 001.43, 50/9.38, 10017.95,
(T in “C) 15Ol7.03, 20016.32, 25015.69, 3OOl5.15,
350/4.73 [S]
silica glass 30 “C, amorphous: 1.38 [38] a: 1.36
silica glass T/ A, amorphous: 011.36, 5011.44,
(T in “C) 100/1.48, 150/1.53, 200/1.58, 250/1.64,
3OOr1.70,35011.78,400/1.85, 450/1.94,
500/2.07 [8]

NON-SILICATES
oxides

magnetite, Fe,O,, iso 22-33 “C, a: 4.61 +_0.42 (8) h,,: 9.7 a: 5.10
hematite, Fe,O, 30 "C, a: 12.42 _+1.74 (3) [7,16] A,,: 14.7, IL,,: 12.1 a: 11.28

ilmenite, FeTiO, 35 “C, a: 1.49 f 0.02 (3) a: 2.38 + 0.18 (2)


chromite, 35 "C, a: 2.19 + 0.15 (3) a: 2.52
(Fe,Mg)Cr,O,, iso
spine], MgAI,O,, iso 35-70 T, a: 12.14 f 1.23 (3) [16] A,,: 13.8 a: 9.48
rutile, TiO, 44-67 "C, I: 7.95 f 0.84 (2) IL,,: 9.3
36-67 "C, (I: 13.19 + 0.63 (2) [16] Xx3:12.9
? “C, a: 4.90 + 0.17 (3) [16] a: 5.12
corundum, A&O, 23-77 "C, I: 17.70 5~3.11 (4) A,,: 31.2
26-70 "C, I(: 18.37 + 3.46 (5) A,,: 38.9
periclase, MgO, iso A,,: 33.5
artificial periclase 400 K, h,,: 41.05 [33]

sulfies

pyrite, Fe&, iso 35 "C, a: 23.15 _+2.00 (3) A,,: 37.9 a: 19.21
pyrrhotite, FeS 35 “C, a: 3.53 k 0.05 (3) a: 4.60
galena, PbS 35 “C, a: 2.76 +_0.18 (3) a: 2.28
CLAUSER AND HUENGES 121

TABLE 3. (continued)

mineral T, state, h., (n) (241 state, h [25] state, h, (n) [28]

sulfates
baryte, BaSO, 25-100 "C, I: 2.07 k 0.02 (2)
25-100 "C, II: 2.92 f 0.07 (4)
25-35 “C, a: 1.72 f 0.04 (4) a: 1.31

anhydrite, CaSO, 25-35 “C, a: 5.36 f 0.27 (6) a: 4.76

gypsum, ? "C, :, 1.30 [20] I: 3.16, II: 3.63


CaSO, .2H,O A,,: 2.6, &: 1.6, A,,: 3.7

carbonates
calcite, CaCO, 30 "C, I: 3.16, II: 3.63 [8] A,,: 4.2, h,,: 5.0 a: 3.59
calcite I T/A, x: 013.48, 5013.00, lOOf2.72,
(T in “C) 15012.52,20012.37,25012.25,30012.16,
350/2.09, 400/2.06 [8]
calcite II T/ 1, x: O/4.00, 5013.40, 10012.99,
(T in “C) 15012.73,20012.55,25012.41,30012.29,
350/2.20, 4OOI2.13[8]
aragonite, CaCO, 25-100 “C, a: 2.37 f 0.22 (11) a: 2.24
magnesite, MgCO, 25-100 "C, I: 7.32 f 0.57 (4)
25-100 "C, II: 7.86 f 0.17 (4)
34-35 “C, a: 8.18 f 1.20 (5) a: 5.84
siderite, FeCO, 35 “C, a: 2.99 f 0.12 (3) a: 3.01
dolomite, CaMg[CO,], 25-35 “C, a: 4.78 f 0.54 (70) a: 4.9 a: 5.51

phosphates
apatite, Ca,[PO,],(F,OH) 35 “C, a: 1.27 f 0.02 (3) a: 1.38 rt 0.01 (2)

halides
halite, NaCl, iso O-35 "C, x: 5.55 IL 1,02 (8) [16] a: 6.5
halite, NaCl, iso T/A, x: 016.11,5Ol5.02,7015.44,
(T in “C) 100/4.21, 15Oi3.59, 20013.12, 250/2.76,
300/2.49, 35Oi2.30,400/2.09 [8]
rocksalt, NaCl, iso 27 “C, a: 5.94 f 0.83 (6)
122 THERMAL CONDUCTIVITY OF ROCKS AND MINERALS

TABLE 3. (continued)

mineral T, state, h , (n) 1241 state, A [25] state, h , (n) [281

rocksalt, NaCI, iso T/k: 0.4lO.95, 0.511.78, 0.6j3.13,


(T in K) 0.714.97, 0.817.40, 0.9/10.0, l/14.0,
U99.3, 31270, 41443, 51595, 61735,
71829, 81880, 91870, 101836, 151502,
201306, 25/191, 30/130, W75.0,
50154.0, 75134.9, lOO/24.3, 15OA5.0,
2OO/10.9, 250/8.24, 29316.65, 30016.57,
40014.80, 5OOf3.67, 60012.98, 70012.47,
8OO/2.08, 9OO/1.85, 1000/1.67 [55]

sylvite, KCl, iso O-12 OC, x: 6.95 + 0.21 (2) [16] A,,: 6.4

fluorite, CaF,, is0 O-36 “C, x: 8.63 + 0.58 (6) A,,: 10.1 a: 9.51

‘minerals marked “iso” are isotropic. T is ambient temperature, and (n) is number of data for mean and standard deviation. “x” denotes
measurements of unknown orientation on single crystals, “a” on monomineralic aggregates. Directions of anisotropy are specified in
one of three ways: (1) by the mineral’s optical a-, b-, or c-axes (100, 010, OOI), (2) by the diagonal elements of the thermal
conductivity tensor (A,,, h,,, A,,), where h,, is parallel to the crystal’s optical c-axis, and the optical a-axis lies within the plane defined
by h,, and hz2, (3) by the thermal conductivity components normal or parallel to the direction of maximum thermal conductivity
(19 II),

TABLE 4. Thermal conductivity h (W mm’ Km’, lower, boldface number) and thermal diffusivity K (10e6m2 se’, upper
lightface number) at different temperatures for quartz, fused silica, olivine, and synthetic periclase.’

mineral 300K 400K 500K 600 K 700K 8OOK 900K 1000 K 11OOK
quartz (001) 7.14 3.57 2.38 1.69 1.37 1.14 1.41 1.54 1.64
13.93 8.20 6.24 4.81 3.91 3.56 3.87 4.56 5.15

quartz (010) 3.33 2.00 1.45 1.15 0.96 0.89 1.00 1.14 1.28
6.49 4.60 3.83 3.29 2.90 2.79 2.75 3.39 4.03

fused silica 0.725 0.715 0.705 0.700 0.715 0.741 0.800 0.885
1.147 1.348 1.499 1.612 1.725 1.854 2.060 2.323

olivine (001) 1.85 1.49 1.22 1.08 1.03 1.04 1.09 1.2 1.35
Fo,,Fa,,) 5.07 4.73 4.23 3.89 3.86 3.98 4.23 4.77 5.44

periclase, MgO 12.5 8.70 6.67 5.56 4.65 4.00 3.57 3.23
W1) - 46.05 34.12 27.21 23.19 19.63 17.12 15.61 14.32

jadeite, a 1.54 1.28 1.11 0.97 0.88 0.84 0.83 0.89 0.96
- - - - -

garnet, x 1.10 1.oo 0.91 0.85 0.81 0.79 0.80 0.81 0.83
(mean of 2) - -
CLAUSER AND HUENGES 123

TABLE 4. (continued)

mineral 300K 400K 5OOK 600 K 700K 800 K 900K 1OOOK 1100 K

spinel, x 3.45 3.13 2.86 2.56 2.44 2.25 2.13


- - -

corundum, x 6.06 4.55 3.45 2.86 2.50 2.13 1.85 1.64


-

alkali feldspar, x 7.09 6.67 6.49 6.71 6.99 7.30 7.81 8.33 8.93
(moonstone1 - - - -

‘after [33]; “x” denotes measurements of unknown orientation on single crystals, “a” on monomineralic aggregates. Directions of
anisotropy are specified by the mineral’s optical a-, b-, or c-axes (100, 010, 001). Temperature conversion: T(V) = T(K) - 273. I5

Robertson (1988) converts the feldspar diffusivity data of 5. CONCLUSION


Kanamori et al. (1968) into conductivity, using a constant
density of p = 2.6 g cm-” and a temperature dependent spe-
cific heat capacity. However, a comparison of this data set This review provides information on thermal conductivity
with results from temperature dependent measurements of of crustal rocks in general. For modest temperatures and
feldspar conductivity performed by other authors shows pressures there is a great variation of thermal conductivity
somewhat ambiguous results: Some measurements contra- which decreases significantly for temperatures and pressures
dict the increase in conductivity with temperature displayed above 300 “C and 20 MPa, respectively. As thermal
by Kanamori at al.‘s (1968) converted data while those per- conductivity for any specific rock type varies according to
formed by Birch & Clark (194Oa,b), seem to confirm it, at its mineral content, porosity, pore fluid, and anisotropy, a
least in the temperature range 25-300 “C. table of thermal conductivity purely according to rock type

TABLE 5. Constants D and E from equation (3b) for


monomineralic aggregates.”

mineral T (“Cl D E
(lo*’ m K W-‘) (10-j m We’)

NaCl O-400 -52.55 0.788

MgO 100-800 -21.50 0.127

GO, 100-800 -28.66 0.155

SiO, (*) 100-400 62.10 0.387

W&O, 100-1000 19.11 0.122

ZrSiO, 100-800 131.37 0.093

Mg,ZrSiO, 100-600 85.98 0.282

(Mg.FejSiO, 100-300 200.63 0.222

gafter [l7]; *: single SiO, crystal, I to optical axis


124 THERMAL CONDUCTIVITY OF ROCKS AND MINERALS

quartz

h = 7.7 W&K-’

metamorphic rocks plutonic rocks

feldspars
plagioclase

X = 2 5 to 5 W rn-‘K-’ X = 1.5 to
kalifeldspar
2.5 W m-‘K-’
A
quartz
h = 7.7 wm-‘K-’

uolcanzc rocks sedimentary rocks

Fig. 7. Thermal conductivity of basic rock-forming minerals and compositional relationship with rocks.
(a) metamorphic and plutonic rocks, (b) volcanic and sedimentary rocks. Metamorphic and volcanic rocks
are in italics, plutonic and sedimentary rocks are not italicized. For volcanic and sedimentary rocks the
third “mineral” phase is air or water, due to the great importance of porosity for the thermal conductivity
of these rocks.

cannot provide site-specific information. Here site-specific spars are not further classified according to the IUGS
measurements are required. (International Union of Geological Sciences) system
For these reasons no table of thermal conductivity versus because of their low variability in thermal conductivity. The
rock type is given. However, in order to illustrate the position of a rock’s name in the compositional triangle
various factors that influence thermal conductivity at least indicates in a qualitative way its thermal conductivity. In
in a semi-quantitative way, we summarize the results of our principle, these two diagrams thus reflect the information
review in two ternary diagrams. These relate different types contained in Figures 1a-ld, presenting it in a somewhat
of rocks with those factors that have the most pronounced different way: metamorphic and plutonic rocks are made up
effect on their thermal conductivity. Two diagrams are of quartz, feldspars, and mafic minerals, and the content of
provided, one for metamorphic and plutonic rocks (Figure minerals from these three groups basically determines a
7a), and one for volcanic and sedimentary rocks (Figure rock’s thermal conductivity. In volcanic and sedimentary
7b). The different rocks are representative for various rocks the third mineral component is replaced by air or
classes of rocks within each group, thus representing the water, as the high variability of porosity in these rocks is a
total spectrum of thermal conductivity in each group. Feld- major factor controlling their thermal conductivity.
CLAUSER AND HUENGES 125

Acknowledgements. This work benefitted a great deal from for his contribution to Roy et al. (1981). Ulfert Seipold
the interdisciplinary approach and enthusiasm for cross- (Geoforschungszentrum Potsdam) and Meinrad Reibelt (Tech.
disciplinary discussions of members of the KTB team of Univ. Berlin) contributed their published and unpublished data,
scientists. Ulli Harms was a knowledgable partner in many Heiner Villinger (Univ. of Bremen) helped with marine data, and
discussions on the thermal consequences of petrological Lazi Rybach (ETH Zurich) provided valuable leads to recent
peculiarities and provided guidance through the system of literature. We are grateful to a number of kind and knowledgeable
petrological classifications. Alan Beck (Univ. of Western Ontario, reviewers who helped to improve this manuscript: Thomas
London, Ont.) shared with us his command of literature on the Ahrens, Alan Beck, Vladimir Cermak, Al Duba, Ralph Hanel,
subject and made available original data sets compiled by himself Daniel Pribnow, Lazi Rybach, John Sass, and Heiner Villinger.

REFERENCES

I. Anand, J., Somerton, W. H., and E. dependence upon temperature and the mantle, in The Earth’s Crust and
Gomaa, Predicting thermal conduc- composition, Part II, Am. J. Sci., Upper Mantle, edited by P. J. Hart,
tivities of formations from other 238(9), 6 13-635, I940b. pp. 622-626, Amer. Geophys. Union,
known properties, Sot. Petrol. Eng. IO. Blackwell, D. D., Thermal conduc- Geophysical Monograph 13, Wash-
Journal, 13, 267-273, 1973. tivity of sedimentary rocks: measure- ington, 1969.
2. Asaad, Y., A study on the thermal ment and significance, in Thermal 18. Clauser, C., Opacity - the concept of
conductivity of fluid-bearing rocks, History of Sedimentary Basins, edited radiative thermal conductivity, in
7 I pp., Dissertation, Univ. of Califor- by N. D. Naser and T. H. McCulloh, Handbook of Terrestrial Heat Flow
nia, Berkeley, 1955. pp. 13-36, Springer, Berlin-Heidel- Density Determination, edited by R.
3. Beck, A. E., Techniques of measur- berg, 1989. Htinel, L. Rybach and L. Stegena, pp.
ing heat flow on land, in Terrestrial I I. Bullard, E. C., Heat flow in South l43- 165, Kluwer, Dordrecht, 1988.
Heat Flow, edited by W. H. K. Lee, Africa, Proc. Roy. Sot. London, Se- 19. Clauser, C., Permeability of crystal-
pp. 24-57, Amer. Geophys. Union, ries A, 173, 474-502, 1939. line rocks, EOS Trans. Amer. Geo-
Washington, 1965. 12. Bullard, E. C., and E. R. Niblett, phys. Union, 73(21), 233,237, 1992.
4. Beck, A. E., Methods for determining Terrestrial heat flow in England, 20. Coster, H. P., Terrestrial heat flow in
thermal conductivity and thermal Monthly Notices Roy. Astr. Sot. Persia, Monthly Notices Roy. Astr.
diffusivity, in Handbook of Terrestri- London, Geophys. Suppl., 6, 222-238, Sot., Geophys. Suppl., 5(5), 131-145,
al Heat Flow Density Determination, 1951. 1947.
edited by R. Hanel, L. Rybach and L. 13. Buntebarth, G., Thermal properties of 21. Davis, E. E., Oceanic heat flow
Stegena, pp. 87-124, Kluwer, Dord- KTB-Oberpfalz VB core samples at density, in Handbook of Terrestrial
recht, 1988. elevated temperature and pressure, Heat Flow Density Determination,
5. Beck, A. E., and J. M. Beck, On the ScientiJic Drilling, 2, 73-80, 1991. edited by R. Hlnel, L. Rybach and L.
measurement of thermal conductivi- 14. Cermak, V., and L. Rybach, Thermal Stegena, pp. 223-260, Kluwer,
ties of rocks by observations on a conductivity and specific heat of Dordrecht, 1988.
divided bar apparatus, Trans. Amer. minerals and rocks, in Landolt- 22. Demongodin, L., B. Pinoteau, G.
Geophys. Union, 39, I I I l-l 123, Bornstein: Numerical Data and Func- Vasseur, and R. Gable, Thermal con-
1958. tional Relationships in Science and ductivity and well logs: a case study
6. Birch, F., Thermal conductivity and Technology, New Series, Group V in the Paris basin, Geophys. J. Int.,
diffusivity, in Handbook of Physical (Geophysics and Space Research), 105, 675-691, 1991.
Constants, edited by F. Birch, J. F. Volume la (Physical Properties of 23. Desai, P. D., R. A. Navarro, S. E.
Schairer, and H. C. Spicer, pp. 243- Rocks), edited by G. Angenheister, Hasan, C. Y. Ho, D. P. Dewitt, and
266, Geological Society of America, pp. 305343, Springer, Berlin-Heidel- T. R. West, Thermophysical Proper-
Special Paper 36, New York, 1942. berg, 1982. ties of Selected Rocks, 256 pp.,
Birch, F., Thermal conductivity, cli- 15. Clark, H., The effects of simple com- CINDAS Report 23, Center for Infor-
matic variation, and heat flow near pression and wetting on the thermal mation and Numerical Data Analysis
Calumet, Michigan, Am. J. Sci., conductivity of rocks, Trans. Amer. and Synthesis (CINDAS), Purdue
252(l), l-25, 1954. Geophys. Union, 22(11), 543-544, Univ., West Lafayette, Indiana
Birch, F., and H. Clark, The thermal 1941. (USA), 1974.
conductivity of rocks and its depen- 16. Clark, S. P. Jr., Thermal Conductivi- 24. Diment, W. H., and H. R. Pratt,
dence upon temperature and composi- ty, in Handbook of Physical Con- Thermal conductivity of some rock-
tion, Part 1, Am. J. Sci., 238(8), 529- stants, edited by S. P. Clark Jr., pp. forming minerals: a Tabulation,
558, 194Oa. 459-482, Geol. Sot. of America, U.S.G.S. Open jZe report 88-690, I5
Birch, F., and H. Clark, The thermal Memoir 97, New York, 1966. pp., U. S. Geol. Survey, Denver Co.,
conductivity of rocks and its 17. Clark, S. P. Jr., Heat conductivity in 1988.
126 THERMAL CONDUCTIVITY OF ROCKS AND MINERALS

25. Dreyer, W., Properties of Anisotropic Burkhardt, Well log-derived estimates 1990.
Solid-State Materials: Thermal and of thermal conductivity in crystalline 47. Somerton, W. H., Thermal Properties
Electric Properties (Materialverhal- rocks penetrated by the 4-km deep and Temperature-Related Behaviour
ten anisotroper Festkijrper: Thermi- KTB Vorbohrung, Geophys. Res. of RocklFluid Systems, 257 pp.,
sche und elektrische Eigenschafien), Lett., 20(12), 1155-I 158, 1993. Elsevier, Amsterdam, 1992.
295 pp., Springer, Wien, 1974 (in 38. Ratcliffe, E. H., Thermal conductivity 48. Sugarawa, A., and Y. Yoshizawa, An
German). of fused and crystalline quartz, Brit. investigation on the thermal conduc-
26. Etheridge, M. A., V. J. Wall, and R. J. Appl. Phys., IO, 22-25, 1959. tivity of porous materials and its
H. Vernon, The role of the fluid 39. Reibelt, M., Study of the Influence of application to porous rock, Austral. J.
phase during regional metamorphism Surface Structure and Fluid-Satura- Phys., 14(4), 469480, 196 I.
and deformation, J. Metamorph. tion of Rocks on the Determination of 49. Sugarawa, A., and Y. Yoshizawa, An
Geol., I, 205-226, 1983. Thermal Conductivity using a Half- experimental investigation on the
27. Grigull, U., and H. Sandner, Heat Space Line Source (Untersuchung des thermal conductivity of porous mate-
Conduction (Wtirmeleitung), 2nd. Einjlusses der Oberfliichenbeschaf- rials, J. Appl. Phys., 33, 3135-3138,
edition, 163 pp., Springer, Berlin- fenheit und der Fluidsiittigung von 1962.
Heidelberg, 1990 (in German). Gesteinen auf die Messung der Whir- 50. Sukharev, G. M., and Z. V. Sterlen-
28. Horai, K., Thermal conductivity of meleitfiihigkeit mit einer Halbraum- ko, Thermal properties of sandstone
rock-forming minerals, J. Geophys. linienquelle), 1 I I pp., Diploma saturated with water and oil (Teplov-
Res., 76(5), 1278-1308, 1971. Thesis (unpublished), Inst. f. Angew. ye svotsva peshankov, nasyshennych
29. Horai, K., Thermal conductivity of Geophysik, Tech. Univ. Berlin, presnoj vody i nefti), Doklady Akade-
Hawaiian basalt: a new interpretation Berlin, 1991 (in German). mija Nauk, SSSR, 194,683-685, 1970
of Robertson and Peck’s data, J. Geo- 40. Robertson, E. C., and D. L. Peck, (in Russian).
phys. Res., 96(B3), 41254132, 1991. Thermal conductivity of vesicular 51. Torgersen, T., Crustal-scale fluid
30. Horai, K., and G. Simmons, Thermal basalt from Hawaii, J. Geophys. Res:, transport: magnitude and mecha-
conductivity of rock-forming miner- 79(32), 4875-4888, 1974. nisms, EOS Trans. Amer. Geophys.
als, Earth and Planet. Sci. Lett., 6, 41. Robertson, E. C., Thermal properties Union, 7/(l), pp. 1,4,13, 1990.
359-368, 1969. of rocks, U.S.G.S. Open jile report 52. Tye, R. P. (Ed.), Thermal Conductivi-
31. Huenges, E., H. Burkhardt, and K. 88-441, I06 pp., U. S. Geol. Survey, ty, ~01s. I and 2, 422 and 353 pp.,
ErbaS, Thermal conductivity profile Reston, Va., 1988. Academic Press, London, 1969.
of the KTB pilot corehole, Scientific 42. Roy, R. F., A. E. Beck, and Y. S. 53. Williams, C. F., and R. A. Anderson,
Drilling, I, 224-230, 1990. Touloukian, ‘Ihermophysical proper- Thermophysical properties of the
32. Hutt, J. R., and J. Berg Jr., Thermal ties of rocks, in Physical Properties earth’s crust: in situ measurements
and electrical conductivities of sand- of Rocks and Minerals, edited by Y. from continental and oceanic drilling,
stone rocks and ocean sediment, Geo- S. Touloukian, W. R. Judd, and R. F. J. Geophys. Res., 95(B6), 9209-9236,
physics, 33, 489-500, 1968. Roy, pp. 409-502, McGraw-Hill/ 1990.
33. Kanamori, H., N. Fujii, and H. Mizu- CINDAS Data Series on Material 54. Woodside, W., and J. H. Messmer,
tani, Thermal diffusivity measurement properties, Volume H-2, McGraw- Thermal conductivity of porous
of rock-forming minerals from 300 ’ Hill, New York 198 I. media, I: Unconsolidated sands, II.
to II00 “K, J. Geophys. Res., 73(2), 43. Sass, J. H., The thermal conductivity Consolidated rocks, J. Appl. Phys.,
595-605, 1968. of fifteen feldspar specimens, J. Geo- 32, 1688-1706, 1961.
34. Kappelmeyer, O., and R. Hsnel, Geo- phys. Res., 70(16), 40644065, 1965. 55. Yang, J. M., Thermophysical proper-
thermics with Special Reference to 44. Sass, J. H., A. H. Lachenbruch, and ties, in Physical Properties Data for
Application, 238 pp., Gebtider Bom- R. J. Monroe, Thermal conductivity Rock Salt, edited by L. H. Gevant-
trlger, Berlin-Stuttgart, 1974. of rocks from measurements on frag- man, pp. 205-221, Monograph 167,
35. Kunii, D., and J. M. Smith, Thermal ments and its application to heat-flow National Bureau of Standards,
conductivities of porous rocks filled determinations, J. Geophys. Res., Washington, I98 I.
with stagnant fluid, Sot. Petrol. Eng. 76(14), 3391-3401, 1971. 56. Zierfuss, H., and G. van der Vliet,
Journal,‘l(l), 37-42, 1961. 45. Sass, J. H., A. H. Lachenbruch, and Measurement of heat conductivity of
36. Messmer, J. H., The thermal conduc- T. H. Moses Jr., Heat flow from a sedimentary rocks, Bull. Am. Assoc.
tivity of porous media. IV. Sand- scientific research well at Cajon Pass, Petrol. Geol., 40, 2475-2488, 1956.
stones. The effect of temperature and California, J. Geophys. Res., 97(B4), 57. Zoth, G., and Htinel, R., Appendix, in
saturation, in Proceedings of the Fifh 5017-5030, 1992. Handbook of Terrestrial Heat Flow
Conference on Thermal Conductivity, 46. Seipold, U., Pressure and temperature Dens@ Determination, edited by R.
Vol. I, pp. l-29, Univ. of Denver, dependence of thermal transport pro- H%nel, L. Rybach and L. Stegena, pp.
Denver, Co., 1965. perties of granites, High Tempera- 449-466, Kluwer, Dordrecht, 1988.
37. Pribnow, D., C. F. Williams, and H. tures - High Pressures, 22, 541-548,

Вам также может понравиться