You are on page 1of 10

Journal of Applied Geophysics 140 (2017) 135–144

Contents lists available at ScienceDirect

Journal of Applied Geophysics

journal homepage: www.elsevier.com/locate/jappgeo

Determining the relationship of thermal conductivity and compressional


wave velocity of common rock types as a basis for
reservoir characterization
P. Mielke ⁎, K. Bär, I. Sass
Technische Universität Darmstadt, Department of Geothermal Science and Technology, Germany

a r t i c l e i n f o a b s t r a c t

Article history: A comprehensive dataset detailing thermal conductivity and acoustic (compressional) wave velocity of 1430
Received 7 August 2016 oven-dry rock samples from clastic sedimentary (sandstone, arkose, greywacke), carbonatic (limestone, marl,
Received in revised form 31 March 2017 dolomite, marble, coquina), plutonic (gabbro, gabbrodiorite, diorite, granodiorite, granite) and volcanic (basalt,
Accepted 1 April 2017
andesite, rhyolite) rock types is presented. Correlation of thermal conductivity, compressional wave velocity
Available online 6 April 2017
and porosity are discussed in detail for each tested rock type. The study confirms that thermal conductivity of
Keywords:
dry rocks can be predicted from acoustic velocity for porous rock types such as volcanites and sandstones,
Thermal conductivity prediction while non- and low-porous rocks show no to minor trends. With a prediction accuracy ±0.5 W m−1 K−1 and
Compressional wave velocity a confidence of N80% for sediments and mafic volcanites the calculated data is far more comprehensive than
Porosity data collected from literature, and is likely accurate enough for most first exploration approaches or geoscientific
Saturation models before detailed site-scale investigation or modelling is conducted.
To investigate the effect of water saturation on thermal conductivity and compressional wave velocity 118 sed-
imentary samples (arkose and fine-, medium- and coarse sandstones) were saturated in de-aired water and the
heat conduction and acoustic velocity were remeasured. The obtained data shows that both thermal conductivity
and compressional wave velocity of saturated samples markedly increase in contrast to dry samples. The extent
of the thermal conductivity and compressional wave velocity gain is mainly controlled by porosity. Thermal con-
ductivity of saturated samples increases twice as much for higher porous samples than for low porous fine and
medium sandstones. In contrast, the gain of compressional wave velocity of saturated sandstones decreases
with increasing porosity.
© 2017 Elsevier B.V. All rights reserved.

1. Introduction Individual rock types typically exhibit a great variability of thermal


conductivity due to heterogeneous mineral composition, variable tex-
Thermal conductivity characterizes the heat transfer capabilities of tures and different porosity (Schön, 2015). Collections of thermal data
materials as a result of a temperature gradient. It is a key property for of rocks (e.g. Bär et al., 2015; Cermak and Rybach, 1982; Clark, 1966;
various geoscientific applications such as geothermal modelling, sedi- Clauser and Huenges, 1995) show the variability of rock thermal con-
mentary and basin studies, but also for geotechnical and construction ductivities (Fig. 1). Most rock types typically exhibit a thermal conduc-
applications. The heat transfer through a rock formation is typically re- tivity range that spans over 3 to 4 W m−1 K−1.
alized through conduction and convection. Conduction (Fourier's law) The correlation of thermal conductivity of rocks and other parame-
dominates in dense, low-porous and impermeable rock types such as ters or properties such as mineral composition, density and porosity,
most plutonic and metamorphic rocks, but also in low permeable to im- and fluid saturation has been investigated by numerous studies
permeable sediments such as mudstones, dense carbonates or highly (Brigaud and Vasseur, 1989; Robertson, 1979; Schön, 2015; Somerton,
compacted sandstones. Under in-situ conditions convective heat trans- 1958, 1992; Zimmerman, 1989). These studies have shown that the
port dominates in permeable rock formations such as porous sediments thermal conductivity of rocks is primarily controlled by mineral compo-
or highly fractured or karstified rocks, where fluids can circulate sition and porosity. In non- to very low-porous rocks, such as plutonic
through the interconnected pores, fractures and cavities. and metamorphic rocks, thermal conductivity is mainly controlled by
the mineral composition and texture. In porous rocks, such as most sed-
⁎ Corresponding author. iments and extrusive rocks, thermal conductivity is primarily controlled
E-mail address: mielke@geo.tu-darmstadt.de (P. Mielke). by the porosity and structure of the pore space, which in turn primarily

http://dx.doi.org/10.1016/j.jappgeo.2017.04.002
0926-9851/© 2017 Elsevier B.V. All rights reserved.
136 P. Mielke et al. / Journal of Applied Geophysics 140 (2017) 135–144

velocity as a proxy to predict thermal conductivity of rocks or entire


rock formations has the advantage of relying on a standard method
that has been successfully applied in reservoir and basin exploration
for decades, and can be applied in the laboratory as well as in boreholes
and large-scale field surveys. The relationship between thermal conduc-
tivity and acoustic velocity is based on the phonon conduction theory,
that assumes thermal energy transfer occurring through the propaga-
tion of acoustic wave packets (phonons) along a thermal gradient
(Horai and Simmons, 1969; Pribnow et al., 1993; Williams and
Anderson, 1990). Both thermal conductivity and acoustic velocity are
primarily controlled by the heat and acoustic wave conduction of the
rock-forming minerals and the type and amount of cementation,
which connects the individual grains. In contrast, “defects” such as
pores, microfractures and associated grain-to-grain boundary effects in-
terfere with the phonon flow through the rock. Consequently, both
thermal conductivity and acoustic velocity decrease with increasing
porosity (hence increase with bulk density).
By contrast, the mineral composition can influence thermal conduc-
tivity and acoustic velocity in opposing directions. Generally, mafic min-
Fig. 1. Thermal conductivity ranges of common rocks. erals exhibit higher densities than felsic minerals, hence both rock
After Cermak and Rybach, 1982. properties should increase for mafic rocks. However, quartz exhibits
low density but high thermal conductivity as compared to other rock-
depend on grain size distribution and sorting, and secondarily by miner- forming mafic and felsic minerals (mostly feldspar and other silicate
al composition and cement type. minerals). As quartz is the main constituent of felsic rock types, its rela-
However, as the range of reported values is commonly quite large tively high thermal conductivity causes an increase of thermal conduc-
and do not reflect the local geological circumstances, the validity of lit- tivity from basic/mafic (e.g. gabbro, basalt) to acid/felsic (e.g. granite,
erature values for a specific problem is often considered questionable rhyolite) rock types, while acoustic wave velocity decreases due to the
(e.g. Fuchs and Balling, 2016a, 2016b; Schintgen et al., 2015). To obtain lower density of felsic rocks as compared to mafic rocks (Schön, 2015).
reliable thermal conductivity values it is essential to conduct large
numbers of measurements that take the lithological and structural 1.3. Existing data
variability of the local strata into consideration.
Published studies that substantiate or disprove the applicability of
1.1. Thermal conductivity measurement methods predicting thermal conductivity from acoustic velocity are rare and
show inconsistent results. Esteban et al. (2015) predicted the thermal
Common direct measurement methods in laboratory are the divided conductivity of 179 dry and wet sandstones from the Perth basin
bar method (Bullard, 1939; Sass et al., 1971), the line source method (Australia) and Soultz-sous-Forêts (France) from acoustic velocity, po-
(Jaeger, 1958, and recent developments of this method e.g. Abid et al., rosity and simplified mineralogy using a model from Pimienta et al.
2014; Hammerschmidt and Meier, 2006), the laser flash method (2014), and reported a good match with measured thermal conductiv-
(Parker et al., 1961), and the optical scanning method (Popov et al., ity data from the same samples. Kukkonen and Peltoniemi (1998) mea-
1999). Unfortunately, laboratory measurements are often constrained sured petro-physical data including compressional wave velocity and
by sample accessibility. In deep boreholes thermal conductivity can thermal conductivity of more than 700 crystalline core samples from
be measured by a range of downhole tools (e.g. Beck et al., 1971; Finland, but found no significant correlation. Popov et al. (2003) pub-
Burkhardt et al., 1995; Hyndman et al., 1979; Kuriyagawa et al., 1983), lished data from more than 800 core samples of sedimentary rocks
but these methods typically work discontinuously and are uneconomic. from different Russian hydrocarbon deposits and impact rocks from
In shallow boreholes the thermal response test (e.g. Gehlin, 2002) has the well “Nördlingen 1973” drilled in the Ries impact structure
emerged to a common tool. Deeper boreholes can be investigated di- (Germany), and concluded that correlations exist between thermal
rectly applying the optical frequency domain reflectometry method conductivity and acoustic velocity, but strongly depend on the local con-
(Lehr and Sass, 2014). Unconsolidated rocks need to be measured ditions. A study by Hartmann et al. (2005) investigated the correlation
with a changeable water content and under increasing or decreasing of thermal conductivity and compressional wave velocity for shaly
compaction pressure (Sass and Stegner, 2012). sandstones and marls, and suggested that good correlations exist, but
Consequently, indirect methods to determine thermal conductivity strongly depend on the local conditions and diagenesis of the rock.
of rocks from other rock properties such as density, porosity, and miner- Gegenhuber and Schön (2012) measured both thermal conductivity
al composition have been explored in many studies (e.g. Fuchs et al., and compressional wave velocity on a total of 35 samples consisting of
2013; Goutorbe et al., 2006; Gross and Combs, 1976; Hartmann et al., mainly granite, basalt and sandstone and concluded that good correla-
2005). An extensive literature compilation of various empirical relation- tions exist. A good correlation of both properties was also confirmed
ships and proposed equations for the indirect determination of the ther- by Özkahraman et al. (2004) who tested a small sample set of
mal conductivity of a rock formation is given by Fuchs et al. (2013), and limestones and andesite.
an overview of most common model concepts is given by Schön (2015). Large quantities of laboratory measurements of thermal conductivi-
In conclusion, these regression-based empirical equations are typically ty and acoustic velocity are typically done on dry samples at ambient
limited to the specific rocks on which they were established for and conditions. The standardized measurement conditions allow direct
are therefore not universally applicable. comparison of the different rock types, but do not reflect the in-situ con-
ditions at depth. Porosity commonly decreases with depth due to in-
1.2. Thermal conductivity and acoustic velocity creasing overburden stress, which in turn facilitates improved heat
and acoustic wave conduction by up to 30% (e.g. Birch, 1960; Clauser
Another indirect method to predict thermal conductivity is its rela- and Huenges, 1995; Horai and Susaki, 1989; Schön, 2015; Walsh and
tionship to acoustic (compressional) wave velocity. Applying acoustic Decker, 1966). In contrast, elevated temperatures at depth can cause
P. Mielke et al. / Journal of Applied Geophysics 140 (2017) 135–144 137

the formation of micro cracks (“thermal cracking”) due the differential 2. Materials and methods
thermal expansion of the rock forming minerals. This can increase po-
rosity (e.g. Bauer and Handin, 1983; Chaki et al., 2008; David et al., A total of 1430 core and plug samples from 18 rock types were se-
1999), which in turn reduces both heat and acoustic wave conduction lected from the rock archive of the Department of Geothermal Science
by up to 50% due to increased contact resistance (e.g. Clauser and and Technology at the Technische Universität Darmstadt. The collection
Huenges, 1995; Fortin et al., 2011; Nara et al., 2011; Schön, 2015; consists of rock samples taken from various projects that have been
Vinciguerra et al., 2005; Vosteen and Schellschmidt, 2003). Rocks at carried out in the past decade. Selected samples comprise of plutonic,
depth are typically saturated, which increases thermal conductivity volcanic, clastic sedimentary and carbonatic rock types, and originate
and compressional wave velocity (as compared to dry conditions), as from various outcrops, quarries and cored boreholes. While the place
fluids conduct heat and compressional waves much better than air. of origin was documented for most samples, it could not be precisely de-
This study reports an extensive assessment of the relationship be- termined for a small subset of samples. A comprehensive list of the sam-
tween thermal conductivity and compressional velocity. Additionally, pled rock types, their place of origin and stratigraphic classification
the effect of water saturation on both properties are investigated. (where available) is given in Table 1. The collection of plutonic and

Table 1
Rock type, place of origin, stratigraphic classification, number of specimen tested and the original literature reference for all tested samples.

Rock type n Country and region Lithostratigraphy n Ref

Clastic sediments
Fine sandstone 91 DE Black Forest Triassic (Buntsandstein) 2 1
DE Odenwald Triassic (Buntsandstein) 8 1
DE Saar Nahe Basin Permo-Carboniferous (Rotliegend) 61 2
DE South German Scarplands Triassic (Keuper) 12 1
DE Spessart Triassic (Buntsandstein) 3 1
DE Wetterau Permo-Carboniferous (Rotliegend) 5 1
Medium sandstone 349 DE Black Forest Triassic (Buntsandstein) 13 1
DE Odenwald Triassic (Buntsandstein) 46 1
DE Pfalz Triassic (Buntsandstein) 21 1
DE Saar Nahe Basin Permo-Carboniferous (Rotliegend) 222 2
DE Spessart Triassic (Buntsandstein) 34 1
DE Wetterau Permo-Carboniferous (Rotliegend) 13 1
Coarse sandstone 114 DE Black Forest Triassic (Buntsandstein) 19 –
DE Saar Nahe Basin Permo-Carboniferous (Rotliegend) 48 2
DE Spessart Triassic (Buntsandstein) 7 –
DE Sprendlinger Horst Permo-Carboniferous (Rotliegend) 34 1
DE Wetterau Permo-Carboniferous (Rotliegend) 6 1
Arkose 29 DE Saar Nahe Basin Permo-Carboniferous (Rotliegend) 29 2
Greywacke 36 DE Kellerwald Carboniferous 11 1
DE n/a n/a 5 –
NZ Central Volcanic Zone Mesozoic 20 3

Carbonates
Limestone 108 DE Franconian Jura Jurassic (Malm) 31 4
DE Swabian Jura Jurassic (Malm) 64 4
DE n/a n/a 13 –
Marl 43 DE Swabian Jura Jurassic (Malm) 43 4
Dolomite 24 DE Kellerwald Permian (Zechstein) 2 1
DE Swabian Jura Jurassic (Malm) 22 4
Marble 38 GR n/a n/a 24 –
IT Piemont n/a 14 –
Coquina 48 DE n/a n/a 48 –

Plutonites
Gabbro 84 DE Odenwald Paleozoic 65 –
SA Karoo n/a 19 –
Gabbrodiorite 17 DE Odenwald Paleozoic 17 5
Diorite 24 DE Sprendlinger Horst Paleozoic 24 5
Granodiorite 72 DE Odenwald Paleozoic 64 1
DE Sprendlinger Horst Paleozoic 8 5
Granite 153 CN Gonghe Basin Triassic 93 –
DE Odenwald Paleozoic 26 1
DE Sprendlinger Horst Paleozoic 3 5
n/a n/a n/a 31 –

Volcanites
Basalt 75 DE Rhön n/a 2 1
DE Odenwald n/a 10 5
DE Vogelsberg Tertiary 4 1
DE n/a n/a 16 –
IS Reykjavik Quaternary 16 –
IS n/a Quaternary, Tertiary 15 –
JO Jordanian Harrat Tertiary 12 6
Andesite 46 NZ Central Volcanic Zone Quaternary 46 3
Rhyolite 79 DE Odenwald Permian 63 –
NZ Central Volcanic Zone Quaternary 16 3

References: 1 — Bär (2012); 2 — Aretz et al. (2015); 3 — Mielke et al. (2015); Mielke et al. (2016); 4 — Homuth et al. (2015); 5 — Sass et al. (2015); 6 — Al-Zyoud et al. (2014).
138 P. Mielke et al. / Journal of Applied Geophysics 140 (2017) 135–144

volcanic rocks reflects the most common rock types of each group, and same direction as the compressional wave velocity measurements.
comprise of rocks types ranging from felsic (granite and rhyolite) to Boundary effects on the sample edges were removed during the post-
mafic (diorite, gabbro, basalt, and andesite). processing phase. The device has an accuracy of 3% in the supported
measuring range from 0.5 to 25 W m−1 K−1 as specified by the manu-
2.1. Sample description facturer Lippmann & Rauen.

Gabbro, gabbrodiorite, diorite, granodiorite and parts of the 2.2.2. Compressional wave velocity
granite mainly derive from the variscan basement of the Odenwald Acoustic velocity (or compressional wave velocity, vp) was deter-
and Sprendlinger Horst (both located in Hesse, Germany). The mined using a commercial ultrasound generator (Geotron USG 40)
majority of the granite originates from the uplifted basement of the with mounted emitter and receiver probes. Measurements were con-
Gonghe Basin, Eastern Tibet Plateau (China) and the Odenwald. ducted along the axis of the core. The specimens were mounted on a
Most specimens come from active open cast mines and are generally framework, and the emitter and receiver were pressed against the cen-
fresh and unaltered, and exhibit no visible porosity. Diorites and tre of the plan parallel grinded end surfaces with a contact pressure of
some granodiorites and granites exhibit minor microcracks along 2000 kPa. To facilitate a better connection between probes and sample
grain boundaries. a shear gel (Magnaflux 54-T04) was applied. Continuous measurements
Volcanites mainly derive from active volcanic zones in Iceland and were performed with a frequency of 80 Hz for samples with an axial
New Zealand, but also from extinct regional volcanic zones in Central length N 12 cm, and with 250 MHz for samples with an axial
Germany and Jordan. Specimens exhibit pore space ranging from mas- length b 12 cm length. Measured data was averaged out of 8 single mea-
sive, non-porous (mainly basalts and andesite) to well-developed con- surements on each specimen.
nected or isolated pore structures (some andesite). Rhyolite samples Both thermal conductivity and acoustic velocity for all samples were
exhibit either abundant or minor porosity and were therefore separated determined on oven-dry rocks and at ambient conditions (0.1 MPa at-
in two subgroups. mospheric pressure and 25 °C). To be able to investigate the effect of
Clastic sedimentary rocks mainly derive from reservoir potential fluid saturation on thermal conductivity and compressional wave veloc-
studies of the Permian Rotliegend- and Triassic Buntsandstein- ity also, selected sets of clastic sandstones and arkose with known po-
formations in Hesse, Germany (Aretz et al., 2015; Bär, 2012). Within rosities were remeasured under saturated conditions. Sets of samples
the scope of both studies rocks were sampled in large number from from the same rock type and place of origin were selected and measur-
quarries, outcrops and boreholes. Both studies generated large amounts ing results were averaged to compensate for possible heterogeneities of
of data, and existing thermal conductivity and porosity measurements an individual specimen. For each rock type, series exhibiting differing
of the regional sedimentary Rotliegend survey were integrated in this porosities were selected. A total of 118 samples were selected,
study. Sedimentary rocks were classified depending on the modal consisting of 3 series of fine sandstone, 6 series of medium sandstone,
composition and matrix content into sandstone, arkose, and greywacke. 3 series of coarse sandstone and 3 series of arkose. The specimens
The dominant grain size of the sandstones was defined macroscopi- were evacuated and subsequently immersed in de-aired water for
cally, and the samples were grouped by a simplified grain size 48 h inside an evacuated container.
classification into a) fine sandstone (N 0.0625 mm to b0.25 mm);
b) medium sandstone (N0.25 mm to b0.5 mm); and c) coarse sand- 2.2.3. Porosity
stone (N 0.5 mm to b2 mm). Sampled arkoses were primarily coarse- Porosity (ϕ) data was available from previous studies for about 50%
grained, while greywacke was fine-grained. of all tested specimens. In all those studies (published and unpublished)
Limestone, marl and dolomite originate from quarries in the Swabi- porosity was measured using the same commercial gas expansion pyc-
an and Franconian Jura in Southern Germany. The specimens are nometer (AccuPyc II 1340) in combination with an envelope density
micritic, massive and exhibit no visible porosity (Homuth et al., 2015). analyser (GeoPyc 1360).
Marble and coquina derives from unknown locations in Greece and
Germany, respectively. 3. Results and discussion

2.2. Methodology The obtained thermal conductivity and compressional wave velocity
data (Fig. 2) correspond to ranges reported by other studies (e.g. Bär
The specimens selected for this study were either present in the et al., 2015; Birch, 1960; Cermak and Rybach, 1982; Clark, 1966;
sample archive as drill cores or plugs. Drill cores had a common diame- Clauser and Huenges, 1995; Popov et al., 2011; Schön, 2015). A compre-
ter of 64 mm and axial lengths of about 80 mm to 150 mm. Core end hensive overview of the measured thermal conductivity and acoustic
faces were often rough and fractured. Where necessary, both end velocity results, as well as corresponding porosity data is given in
faces of the samples were cut to be parallel. Plugs had a diameter of Table 2. A detailed list of all thermal conductivity, compressional wave
about 40 mm and axial lengths of about 20 mm to 40 mm with parallel velocity and associated porosity values is provided in Appendix A as
cut end faces. The samples were then oven-dried at 65 °C (carbonates) supplementary material.
or 105 °C (all other rock types) for at least 48 h. About 24 h before the In porous rocks, both thermal conductivity and compressional wave
actual measurements samples were removed from the oven and stored velocity are mainly controlled by porosity (Fig. 3). Both thermal conduc-
in a desiccator for cooling down to ambient temperature. tivity and compressional wave velocity of most porous rock types, such
as sediments and volcanites, generally decreases with increasing poros-
2.2.1. Thermal conductivity ity. Some coarse sandstones and (coarse) arkoses exhibit increased
Bulk thermal conductivity (λ) was measured using the optical scan- thermal conductivity and deviate from the general regression trend of
ning method after Popov et al. (1999). The method is based on heating similar material. These deviations are likely to be caused by different
the sample with a contactless heat emitter, and measuring the subse- types and amounts of cementation in the pore space (e.g. Aretz et al.,
quent cooling rate with contactless temperature sensors. By comparison 2015). By contrast, low-porous rocks such as plutonites, greywacke
with a set of standards with known thermal conductivity the thermal and carbonates exhibit no or only weak correlation of thermal conduc-
conductivity of the sample can be calculated. The thermal conductivity tivity and porosity, or acoustic velocity and porosity. Fig. 3 shows that
was scanned in 2 mm intervals along a measuring line on the plan par- the influence of the rock type is less striking for the compressional
allel end faces of the cores. Thus, the thermal vector detected by the de- wave velocity than for thermal conductivity. For example, volcanites and
vice is parallel to the core or plug axis and therefore oriented in the plutonites form clearly delimitable point clusters in the porosity vs.
P. Mielke et al. / Journal of Applied Geophysics 140 (2017) 135–144 139

Fig. 2. Box-Whisker plots of thermal conductivity (λ) and compressional wave velocity (vp) at dry condition for all tested rock types.

thermal conductivity plot (Fig. 3a), but cannot be distinguished in the po- velocities than the sandstones. Greywacke has commonly about 10%
rosity vs. compressional wave velocity plot (Fig. 3b). The reason for this higher thermal conductivities than the other tested sediments, while
are the differing heat conduction properties of the rock forming minerals, the compressional wave velocity is about twice as high as compared
which cause significantly differing thermal conductivities of the whole to the sandstones due to the significantly lower porosity (b2%). Carbon-
rock, but have little influence on the acoustic velocity. ates exhibit thermal conductivities between 2 W m−1 K−1 and
Fine, medium and coarse sandstones exhibit nearly equal mean 3 W m−1 K−1, where limestone has the lowest and dolomite has the
thermal conductivities, while the range is relatively wide. Porosity of highest values, which corresponds to reported values (Clauser and
clastic sandstones commonly ranges from about 5% up to about 22%, Huenges, 1995). Although porosity is generally low (b 5%) and homoge-
and increases with grain size. As the compressional wave velocity is neous, carbonates exhibit a wide range of acoustic velocities, ranging
mainly controlled by porosity, it decreases with increasing grain size. from 2.2 km s−1 up to 7 km s−1.
Arkoses have porosities comparable to medium and coarse sandstones, Thermal conductivity of plutonites is lowest for mafic rocks (gabbro,
but marginally higher mean thermal conductivity and mean acoustic gabbrodiorite) and highest for felsic rocks (granodiorite, granite) due to
the relatively high thermal conductivity of quartz. Compressional wave
velocity decreases with the increasing felsic character of the rock, which
Table 2
corresponds to previous studies (Schön, 2015). Porosity of plutonites is
Results of the thermal conductivity (λ) and compressional wave velocity (vp) measure-
ments at dry condition and corresponding porosity values (Φ) from published and unpub-
generally b2%.
lished internal data. Volcanites have about 50% lower thermal conductivity than
plutonites but span over a wider range. Within the here presented sam-
λ vp Φ
ple set thermal conductivity of porous felsic volcanites (rhyolite) is
W m−1 K−1 km s−1 % lower than that of the porous mafic volcanites (andesite, basalt),
Mean Std. Mean Std. n Mean Std. n Ref. which contradicts the trend from the plutonites and is most likely
Clastic sediments strongly related to porosity of the volcanic sample sets. Volcanites ex-
Fine sandstone 2.49 0.43 3.31 0.72 91 11.1 5.3 65 1, 2 hibit a wide range of porosities, which greatly influences both thermal
Medium sandstone 2.5 0.37 2.93 0.57 349 15 4.5 219 1, 2 conductivity and acoustic velocity.
Coarse sandstone 2.47 0.63 2.64 0.55 114 17.7 3.7 51 1, 2
Arkose 2.64 0.18 2.81 0.32 29 19.1 2.6 29 2
Greywacke 2.86 0.45 5.05 0.79 36 1.7 1.4 29 3
3.1. Correlation of thermal conductivity and acoustic velocity

Carbonates The strength and direction of the correlation between thermal


Limestone 2.45 0.22 5.03 0.73 108 3 1.3 45 4
Marl 2.59 0.12 5.01 0.94 43 1.3 0.7 40 4
conductivity and acoustic velocity differ for the rock groups (Fig. 4).
Dolomite 2.68 0.1 5.14 1.12 24 2.4 1.6 22 1, 4 While most porous rocks exhibit minor to good correlation of both
Marble 2.84 0.17 3.18 0.99 38 n/a n/a n/a – properties, no or only weak correlations are present for non-porous or
Coquina 1.86 0.17 3.85 0.34 48 n/a n/a n/a – low-porous rock types.
Plutonites In detail, fine, medium and coarse sandstones exhibit moderate to
Gabbro 2.42 0.19 6.25 0.48 84 2.6 1.9 3 – good correlations of thermal conductivity and acoustic velocity with re-
Gabbrodiorite 2.55 0.13 5.85 0.26 17 n/a n/a n/a 5 gression gradients gradually increasing with grain size (and consequently
Diorite 2.48 0.2 5.73 0.84 24 1 0.6 20 5
porosity), while the intercept is gradually decreasing (Fig. 4a–c). Arkose
Granodiorite 2.7 0.34 4.69 1.04 72 1.6 2.2 46 1, 5
Granite 2.79 0.29 5.14 0.96 153 0.2 0.8 46 1, 5 (Fig. 4d) and greywacke (Fig. 4e) exhibit no correlation, which could be
due to a mixture of samples of a different origin and thus different
Volcanites
modal compositions, matrix contents and cementation types. Thus,
Basalt 1.7 0.47 4.73 1.16 75 11.8 9.6 15 1, 5, 6
Andesite 1.68 0.37 4.2 0.82 49 4.7 5.2 49 3 these influences might mask correlations, which might well exist for
Rhyolite (massive) 2.84 0.16 4.22 0.47 63 n/a n/a n/a - samples with homogeneous composition from one geological setting
Rhyolite (porous) 1.04 0.12 2.02 0.54 16 19.1 9 16 3 but differences in grain size and porosity.
References: 1 — Bär, 2012; 2 — Aretz et al., 2015; 3 — Mielke et al., 2015, 2016; 4 — Homuth Carbonates exhibit no correlation between thermal conductivity and
et al., 2015; 5 — Sass et al., 2015; 6 — Al-Zyoud et al., 2014. compressional wave velocity. The massive, low-porous limestones
140 P. Mielke et al. / Journal of Applied Geophysics 140 (2017) 135–144

Fig. 3. Cross plots of porosity and compressional wave velocity (vp), and porosity and thermal conductivity (λ) at dry condition for all samples grouped according to rock type.

Fig. 4. Cross plots of thermal conductivity and compressional wave velocity of all tested rock types at dry condition.
P. Mielke et al. / Journal of Applied Geophysics 140 (2017) 135–144 141

(Fig. 4f), marls (Fig. 4g), dolomites (Fig. 4h) and marbles (Fig. 4i) exhibit properties markedly increase (Fig. 6a). Thermal conductivity of saturat-
relatively equal results, which span over a wide range of compressional ed samples increases notably stronger for higher porous samples
wave velocities, but exhibit little deviations in the thermal conductivity. (Fig. 6b). For example, the thermal conductivity of low porous fine
Coquina, which exhibits a considerable amount of visible pore space, and medium sandstones (sample series FS 1 and MS 1 in Fig. 5) in-
was included in this study to check whether porosity in carbonates creases by 0.68 W m−1 K−1 and 0.76 W m−1 K−1, respectively, while
causes a correlation between thermal conductivity and compressional the thermal conductivity of higher porous sandstones such as sample
wave velocity comparable to clastic sandstones. As shown in Fig. 4j, co- series FS 3 and MS 6 increases twice as high. In fact, the resulting ther-
quina plots within a dense point cloud with a weak negative correlation, mal conductivity of saturated highly porous sandstones exceeds the
which may be attributed to the heterogeneous and chaotic distribution thermal conductivity of saturated lower porous samples, although the
of the pore space in the rock forming shells. initial dry thermal conductivity of the latter was markedly higher. In
The tested plutonic rock types mainly exhibit weak correlations be- contrast, the gain of the compressional wave velocity of saturated sand-
tween thermal conductivity and compressional wave velocity that grad- stones decreases with increasing porosity (Fig. 6c).
ually change from a negative to a positive regression with increasing The effective gain of thermal conductivity and compressional wave
felsic character of the rock (Fig. 4k–o). A detailed explanation on this ob- velocity with saturation, expressed as the difference of λsat. − λdry and
servation is given in the Introduction. Altogether, the results match to vp,sat. − vp,dry, respectively, in relation to porosity is shown in Fig. 6d
the point clusters published by Kukkonen and Peltoniemi (1998), and and e, which illustrate the contrasting effects of saturation on both
widely confirm their assumption that thermal conductivity of plutonites properties as a function of porosity.
does not correlate well with acoustic velocity.
Volcanites exhibit the best correlation of thermal conductivity and
compressional wave velocity among all tested rock types (Fig. 4p–r). 3.3. Predicting thermal conductivity from acoustic velocity
Both rock properties are primarily controlled by porosity, while mineral
composition and texture has a subordinate impact under the given The data suggests that the thermal conductivity of dry porous
porosity range of the sample set (Fig. 3). This effect is particularly well rocks such as sandstone, basalt and andesite can be predicted from com-
noticeable for the rhyolites, which were subdivided in a massive (i.e. pressional wave velocities without exact knowledge of porosity and
no visible pore space) and a porous (i.e. visible pore space) group mineralogy. In contrast, the correlation of thermal conductivity and
(Fig. 4r). While the regression of both the massive and porous rhyolite compressional wave velocity of carbonates and plutonites (except for
is about the same, the intercept of the porous specimens is much diorite and granodiorite) is only poorly established and therefore ren-
lower than that of the massive specimens. ders a prediction of thermal conductivity from acoustic data widely im-
The data suggests that the thermal conductivity of dry porous possible. Given these conditions an approach for the prediction of
rocks such as sandstone, basalt and andesite can be predicted thermal conductivity for sandstones and volcanites can be proposed,
from compressional wave velocities with an accuracy of about which bases on the linear increase of thermal conductivity with acoustic
± 0.5 W m− 1 K− 1 without exact knowledge of porosity and miner- velocity. In our case, where both parameters are primarily controlled by
alogy. In contrast, for carbonates and plutonites (except for diorite the porosity, this linear increase is independent from origin and strati-
and granodiorite) the correlation of thermal conductivity and com- graphic association of the respective rock types. As the effect of the
pressional wave velocity is only poorly established. Therefore, for grain size of sandstones is negligible, all sandstone types and arkose
these rock types a prediction of thermal conductivity from acoustic were merged and one linear regression was calculated. The regression
data seems widely impossible. equation facilitates the calculation of thermal conductivity from
acoustic data with an accuracy of ±0.5 W m−1 K−1 and a confidence
3.2. Effect of water saturation of N 80% for both rock types (Fig. 7).
Table 3 lists calculated thermal conductivities for volcanites and
The effect of saturation on the thermal conductivity and compres- sandstones in 1000 m s−1 intervals. The thermal conductivity of
sional wave velocity of sandstone and arkose is shown in Fig. 5. Gener- volcanites increases by 0.27 W m−1 K−1 for each 1000 m s−1 acoustic
ally, both thermal conductivity and compressional wave velocity of wave velocity interval, while sandstones exhibit an increase of
dry samples decrease with increasing porosity. Once saturated, both 0.43 W m−1 K−1.

Fig. 5. Boxplots of porosity, thermal conductivity and compressional wave velocity of tested sandstones and arkose at dry and saturated conditions.
142 P. Mielke et al. / Journal of Applied Geophysics 140 (2017) 135–144

Fig. 6. Cross plots of averaged thermal conductivity and compressional wave velocity (a), thermal conductivity and porosity (b) and compressional wave velocity and porosity (c) for dry
and saturated sandstone and arkose. Plots d and e show the correlation of λsat. − λdry and vp,sat. − vp,dry, respectively, with porosity.

4. Conclusions literature, and is likely accurate enough for most first exploration ap-
proaches or geoscientific models before detailed site-scale investigation
In this study the data of the measurement of thermal conductivity or modelling is conducted.
and acoustic wave velocity of 1430 rock samples from 18 clastic sedi- For a large scale in-situ application the effects of saturation, stress
mentary, carbonatic, plutonic and volcanic rock types on oven-dry and temperature of the rocks need to be considered. Most effects can
cores and plugs was presented. Additionally, 118 samples were selected be calculated by generic equations to transfer data from measurements
for measurements of both properties under water saturated conditions. at lab conditions on dry samples to in-situ conditions considering the
The study confirms that thermal conductivity of dry rocks can be temperature, pressure degree of saturation and the saturating fluid
predicted from the acoustic velocity for porous rock types such as (see references given above). The comparative study of dry and saturat-
volcanites and sandstones, while non- and low-porous rocks show no ed sandstones showed contrasting effects of saturation on thermal
to minor trends. conductivity and compressional wave velocity, which render the pre-
Acoustic velocity data is often routinely collected within the scope of diction very complex and clearly illustrate, that rock type specific em-
various exploration projects (2D or 3D seismic investigations or vertical piric relations need to be defined. To date, few studies have been
seismic profiles in boreholes), and can therefore easily be used for conducted that included all factors. New methods are necessary to val-
predicting thermal conductivity (e.g. Gu et al., 2017). With a prediction idate the results obtained for oven-dry rocks at reservoir conditions.
accuracy ±0.5 W m−1 K−1 for sediments and mafic volcanites the cal- Promising approaches include the application of high temperature tri-
culated data is far more comprehensive than any data collected from axial cells to measure rock properties at elevated temperature and

Table 3
Acoustic wave velocity intervals of volcanites and sandstones, and the predicted thermal
conductivity for each interval.

vp λ
−1
km s W m−1 K−1

±0.5

Volcanites Sandstones

1 0.68 1.68
2 0.96 2.11
3 1.24 2.53
4 1.53 2.96
5 1.81 3.39
6 2.09
Fig. 7. Cross plot with confidence intervals of compressional wave velocity (vp) and
7 2.37
thermal conductivity (λ) for sediments and mafic volcanites.
P. Mielke et al. / Journal of Applied Geophysics 140 (2017) 135–144 143

pressure (e.g. Milsch et al., 2008; Pei et al., 2014; Siratovich et al., 2015), Fuchs, S., Förster, A., Fuchs, S., Forster, A., 2013. Well-log based prediction of thermal con-
ductivity of sedimentary successions: a case study from the North German Basin.
but the tests are complicated and time-consuming. Once the results Geophys. J. Int. 196 (1), ggt382. http://dx.doi.org/10.1093/gji/ggt382.
have been confirmed and transferred to empiric relationships for reser- Gegenhuber, N., Schön, J., 2012. New approaches for the relationship between compres-
voir conditions, the application of well-log data for predicting thermal sional wave velocity and thermal conductivity. J. Appl. Geophys. 76:50–55. http://
dx.doi.org/10.1016/j.jappgeo.2011.10.005.
conductivity at depth will take one significant step forward. Gehlin, S., 2002. Thermal Response Test Method Development and Evaluation. Luleå
University of Technology, Luleå.
Goutorbe, B., Lucazeau, F., Bonneville, A., 2006. Using neural networks to predict thermal
Appendix A. Supplementary data conductivity from geophysical well logs. Geophys. J. Int. 166 (1):115–125. http://dx.
doi.org/10.1111/j.1365-246X.2006.02924.x.
Gross, R., Combs, J., 1976. Thermal conductivity measurement and prediction from
Supplementary data associated with this article are archived under geophysical well log parameters with borehole application. NASA STI/Recon
dataset #874146 (Mielke et al., 2017) in the PANGAEA database hosted Technical Report N 77.
Gu, Y., Rühaak, W., Bär, K., Sass, I., 2017. Using seismic data to estimate the spatial
by the Alfred Wegener Institute, Helmholtz Center for Polar and Marine distribution of rock thermal conductivity at reservoir scale. Geothermics 66:61–72.
Research (AWI) and the Center for Marine Environmental Sciences, http://dx.doi.org/10.1016/j.geothermics.2016.11.007.
University of Bremen (MARUM). Hammerschmidt, U., Meier, V., 2006. New transient hot-bridge sensor to measure
thermal conductivity, thermal diffusivity, and volumetric specific heat. Int.
J. Thermophys. 27 (3):840–865. http://dx.doi.org/10.1007/s10765-006-0061-2.
Hartmann, A., Rath, V., Clauser, C., 2005. Thermal Conductivity from Core and Well log
References Data. Rock Physics and GeomechanicsRock Physics and Geomechanics 42 (7–8):
pp. 1042–1055. http://dx.doi.org/10.1016/j.ijrmms.2005.05.015.
Abid, M., Hammerschmidt, U., Köhler, J., 2014. Temperature and moisture dependent Homuth, S., Götz, A.E., Sass, I., 2015. Reservoir characterization of the Upper Jurassic
thermophysical properties of Sander sandstone. Int. J. Therm. Sci. 86:88–94. http:// geothermal target formations (Molasse Basin, Germany): role of thermofacies as
dx.doi.org/10.1016/j.ijthermalsci.2014.06.021. exploration tool. Geotherm. Energy Sci. 3 (1):41–49. http://dx.doi.org/10.5194/gtes-
Al-Zyoud, S., Rühaak, W., Sass, I., 2014. Dynamic numerical modeling of the usage of 3-41-2015.
groundwater for cooling in north east Jordan — a geothermal case study. Renew. Horai, K.-I., Simmons, G., 1969. Thermal conductivity of rock-forming minerals. Earth
Energy 62:63–72. http://dx.doi.org/10.1016/j.renene.2013.06.027. Planet. Sci. Lett. 6 (5):359–368. http://dx.doi.org/10.1016/0012-821X(69)90186-1.
Aretz, A., Bär, K., Götz, A.E., Sass, I., 2015. Outcrop analogue study of Permocarboniferous Horai, K.-I., Susaki, J.-I., 1989. The effect of pressure on the thermal conductivity of silicate
geothermal sandstone reservoir formations (northern Upper Rhine Graben, rocks up to 12 kbar. Phys. Earth Planet. Inter. 55 (3–4):292–305. http://dx.doi.org/10.
Germany): impact of mineral content, depositional environment and diagenesis on 1016/0031-9201(89)90077-0.
petrophysical properties. Int. J. Earth Sci. (Geol. Rundsch.) http://dx.doi.org/10. Hyndman, R.D., Davis, E.E., Wright, J.A., 1979. The measurement of marine geothermal
1007/s00531-015-1263-2. heat flow by a multipenetration probe with digital acoustic telemetry and in situ
Bär, K., 2012. Untersuchung der tiefengeothermischen Potenziale von Hessen thermal conductivity. Mar. Geophys. Res. 4 (2):181–205. http://dx.doi.org/10.1007/
(Dissertation, Darmstadt). BF00286404.
Bär, K., Reinsch, T., Freymark, J., Mielke, P., Strom, A., Wiesner, P., 2015. International rock Jaeger, J.C., 1958. The measurement of thermal conductivity and diffusivity with cylindri-
property database — fast access to a comprehensive dataset. Der Geothermiekongress cal probes. EOS Trans. Am. Geophys. Union 39 (4):708–710. http://dx.doi.org/10.
2015. Der Geothermiekongress, Essen. 02.-04.11.2015. 1029/TR039i004p00708.
Bauer, S.J., Handin, J., 1983. Thermal expansion and cracking of three confined water- Kukkonen, I.T., Peltoniemi, S., 1998. Relationships between thermal and other
saturated igneous rocks to 800 °C. Rock Mech. Rock. Eng. 16 (3):181–198. http:// petrophysical properties of rocks in Finland. Phys. Chem. Earth 23 (3):341–349.
dx.doi.org/10.1007/BF01033279. http://dx.doi.org/10.1016/S0079-1946(98)00035-4.
Beck, A.E., Anglin, F.M., Sass, J.H., 1971. Analysis of heat flow data — in situ thermal con- Kuriyagawa, M., Matsunaga, I., Yamaguchi, T., 1983. An In Situ Determination of the Ther-
ductivity measurements. Can. J. Earth Sci. 8 (1):1–19. http://dx.doi.org/10.1139/ mal Conductivity of Granitic Rock. International Society for Rock Mechanics.
e71-001. Lehr, C., Sass, I., 2014. Thermo-optical parameter acquisition and characterization of geo-
Birch, F., 1960. The velocity of compressional waves in rocks to 10 kilobars: 1. J. Geophys. logic properties: a 400-m deep BHE in a karstic alpine marble aquifer. Environ. Earth
Res. 65 (4):1083–1102. http://dx.doi.org/10.1029/JZ065i004p01083. Sci. 72 (5):1403–1419. http://dx.doi.org/10.1007/s12665-014-3310-x.
Brigaud, F., Vasseur, G., 1989. Mineralogy, porosity and fluid control on thermal conduc- Mielke, P., Bär, K., Sass, I., 2017. Determining the relationship of thermal conductivity and
tivity of sedimentary rocks. Geophys. J. Int. 98 (3):525–542. http://dx.doi.org/10. compressional wave velocity of common rock types. PANGAEA http://dx.doi.org/10.
1111/j.1365-246X.1989.tb02287.x. 1594/PANGAEA.874146.
Bullard, E.C., 1939. Heat flow in South Africa. Proc. R. Soc. Lond. A Math. Phys. Sci. 173 Mielke, P., Nehler, M., Bignall, G., Sass, I., 2015. Thermo-physical rock properties and the
(955):474–502. http://dx.doi.org/10.1098/rspa.1939.0159. impact of advancing hydrothermal alteration — a case study from the Tauhara geo-
Burkhardt, H., Honarmand, H., Pribnow, D., 1995. Test measurements with a new thermal thermal field, New Zealand. J. Volcanol. Geotherm. Res. 301:14–28. http://dx.doi.
conductivity borehole tool. Heat flow and thermal regimes of continental lithosphere org/10.1016/j.jvolgeores.2015.04.007.
244 (1–3):161–165. http://dx.doi.org/10.1016/0040-1951(94)00224-W. Mielke, P., Weinert, S., Bignall, G., Sass, I., 2016. Thermo-physical rock properties of
Cermak, V., Rybach, L., 1982. Thermal Properties, in: Angenheister, G., Cermák, V., greywacke basement rock and intrusive lavas from the Taupo Volcanic Zone, New
Hellwege, K.-H., Landolt, H. (Eds.), Zahlenwerte und Funktionen aus Zealand. J. Volcanol. Geotherm. Res.
Naturwissenschaft und Technik: Neue Serie. — Numerical Data and Functional Rela- Milsch, H.H., Spangenberg, E., Kulenkampff, J., Meyhöfer, S., 2008. A new apparatus for
tionships in Science and Technology: New Series c. Springer, Berlin, pp. 310–314. long-term petrophysical investigations on geothermal reservoir rocks at simulated
Chaki, S., Takarli, M., Agbodjan, W.P., 2008. Influence of thermal damage on physical in-situ conditions. Transp. Porous Media 74 (1):73–85. http://dx.doi.org/10.1007/
properties of a granite rock: porosity, permeability and ultrasonic wave evolutions. s11242-007-9186-4.
Constr. Build. Mater. 22 (7):1456–1461. http://dx.doi.org/10.1016/j.conbuildmat. Nara, Y., Meredith, P.G., Yoneda, T., Kaneko, K., 2011. Influence of macro-fractures and
2007.04.002. micro-fractures on permeability and elastic wave velocities in basalt at elevated pres-
Clark, S.P., 1966. Handbook of Physical Constants. Geological Society of America, New sure. Tectonophysics 503 (1–2):52–59. http://dx.doi.org/10.1016/j.tecto.2010.09.027.
York. Özkahraman, H.T., Selver, R., Işık, E.C., 2004. Determination of the thermal conductivity of
Clauser, C., Huenges, E., 1995. Thermal conductivity of rocks and minerals. Am. Geophys. rock from P-wave velocity. Int. J. Rock Mech. Min. Sci. 41 (4):703–708. http://dx.doi.
Union http://dx.doi.org/10.1029/RF003p0105 (http://onlinelibrary.wiley.com/doi/10. org/10.1016/j.ijrmms.2004.01.002.
1029/RF003p0105/pdf, 105 pp.). Parker, W.J., Jenkins, R.J., Butler, C.P., Abbott, G.L., 1961. Flash method of determining ther-
David, C., Menéndez, B., Darot, M., 1999. Influence of stress-induced and thermal cracking mal diffusivity, heat capacity, and thermal conductivity. J. Appl. Phys. 32 (9):1679.
on physical properties and microstructure of La Peyratte granite. Int. J. Rock Mech. http://dx.doi.org/10.1063/1.1728417.
Min. Sci. 36 (4):433–448. http://dx.doi.org/10.1016/S0148-9062(99)00010-8. Pei, L., Rühaak, W., Stegner, J., Bär, K., Homuth, S., Mielk, P., Sass, I., 2014. Thermo-Triax: an
Esteban, L., Pimienta, L., Sarout, J., Piane, C.D., Haffen, S., Geraud, Y., Timms, N.E., 2015. apparatus for testing petrophysical properties of rocks under simulated geothermal
Study cases of thermal conductivity prediction from P-wave velocity and porosity. reservoir conditions. Geotech. Test. J. 38 (1):20140056. http://dx.doi.org/10.1520/
Geothermics 53:255–269. http://dx.doi.org/10.1016/j.geothermics.2014.06.003. GTJ20140056.
Fortin, J., Stanchits, S., Vinciguerra, S., Guéguen, Y., 2011. Influence of thermal and me- Pimienta, L., Sarout, J., Esteban, L., Piane, C.D., 2014. Prediction of rocks thermal conductiv-
chanical cracks on permeability and elastic wave velocities in a basalt from Mt. ity from elastic wave velocities, mineralogy and microstructure. Geophys. J. Int. 197
Etna volcano subjected to elevated pressure. Tectonophysics 503 (1–2):60–74. (2):860–874. http://dx.doi.org/10.1093/gji/ggu034.
http://dx.doi.org/10.1016/j.tecto.2010.09.028. Popov, Y.A., Pribnow, D.F., Sass, J.H., Williams, C.F., Burkhardt, H., 1999. Characterization of
Fuchs, S., Balling, N., 2016a. Improving the temperature predictions of subsurface thermal rock thermal conductivity by high-resolution optical scanning. Geothermics 28 (2):
models by using high-quality input data. Part 1: uncertainty analysis of the thermal- 253–276. http://dx.doi.org/10.1016/S0375-6505(99)00007-3.
conductivity parameterization. Geothermics 64:42–54. http://dx.doi.org/10.1016/j. Popov, Y., Romushkevich, R., Korobkov, D., Mayr, S., Bayuk, I., Burkhardt, H., Wilhelm, H.,
geothermics.2016.04.010. 2011. Thermal properties of rocks of the borehole Yaxcopoil-1 (Impact Crater
Fuchs, S., Balling, N., 2016b. Improving the temperature predictions of subsurface Chicxulub, Mexico). Geophys. J. Int. 184 (2):729–745. http://dx.doi.org/10.1111/j.
thermal models by using high-quality input data. Part 2: a case study from the 1365-246X.2010.04839.x.
Danish-German border region. Geothermics 64:1–14. http://dx.doi.org/10.1016/j. Popov, Y., Tertychnyi, V., Romushkevich, R., Korobkov, D., Pohl, J., 2003. Interrelations be-
geothermics.2016.04.004. tween thermal conductivity and other physical properties of rocks: experimental
144 P. Mielke et al. / Journal of Applied Geophysics 140 (2017) 135–144

data. Pure Appl. Geophys. 160 (5–6):1137–1161. http://dx.doi.org/10.1007/ stimulation of permeability in geothermal reservoirs. Int. J. Rock Mech. Min. Sci. 80:
PL00012565. 265–280. http://dx.doi.org/10.1016/j.ijrmms.2015.09.023.
Pribnow, D., Williams, C.F., Burkhardt, H., 1993. Well log-derived estimates of thermal Somerton, W.H., 1958. Some Thermal Characteristics of Porous Rocks. Society of Petro-
conductivity in crystalline rocks penetrated by the 4-km deep KTB Vorbohrung. leum Engineers.
Geophys. Res. Lett. 20 (12):1155–1158. http://dx.doi.org/10.1029/93GL00480. Somerton, W.H., 1992. Thermal properties and temperature-related behavior of rock/
Robertson, E.C., 1979. Thermal Conductivity of Rocks (Open File Report 79-356). fluid systems *rock-fluid*. Elsevier, Amsterdam u.a., VIII, 257 S.
Sass, J.H., Lachenbruch, A.H., Munroe, R.J., 1971. Thermal conductivity of rocks from mea- Vinciguerra, S., Trovato, C., Meredith, P.G., Benson, P.M., 2005. Relating seismic velocities,
surements on fragments and its application to heat-flow determinations. J. Geophys. thermal cracking and permeability in Mt. Etna and Iceland basalts. Int. J. Rock Mech.
Res. 76 (14):3391–3401. http://dx.doi.org/10.1029/JB076i014p03391. Min. Sci. 42 (7–8):900–910. http://dx.doi.org/10.1016/j.ijrmms.2005.05.022.
Sass, I., Stegner, J., 2012. Coupled measurements of thermophysical and hydraulic Vosteen, H.-D., Schellschmidt, R., 2003. Influence of temperature on thermal conductivity,
properties of unsaturated and unconsolidated rocks. Thirty-Seventh Workshop on thermal capacity and thermal diffusivity for different types of rock. Phys. Chem. Earth
Geothermal Reservoir Engineering, Stanford (30.01–01.02.2012). A/B/C 28 (9–11):499–509. http://dx.doi.org/10.1016/S1474-7065(03)00069-X.
Sass, I., Welsch, B., Bär, K., Schulte, D.O., Rühaak, W., Chauhan, S., 2015. Report 375/13-14: Walsh, J.B., Decker, E.R., 1966. Effect of pressure and saturating fluid on the thermal con-
Simulation und Evaluierung von Kopplungs- und Speicherkonzepten regenerativer ductivity of compact rock. J. Geophys. Res. 71 (12):3053–3061. http://dx.doi.org/10.
Energieformen zur Heizwärmeversorgung. 1029/JZ071i012p03053.
Schintgen, T., Förster, A., Förster, H.-J., Norden, B., 2015. Surface heat flow and lithosphere Williams, C.F., Anderson, R.N., 1990. Thermophysical properties of the Earth's crust: in situ
thermal structure of the Rhenohercynian Zone in the greater Luxembourg region. measurements from continental and ocean drilling. J. Geophys. Res. Solid Earth 95
Geothermics 56:93–109. http://dx.doi.org/10.1016/j.geothermics.2015.03.007. (B6):9209–9236. http://dx.doi.org/10.1029/JB095iB06p09209.
Schön, J. (Ed.), 2015. Physical Properties of Rocks: Fundamentals and Principles of Zimmerman, R.W., 1989. Thermal conductivity of fluid-saturated rocks. J. Pet. Sci. Eng. 3
Petrophysics. Elsevier, Amsterdam Netherlands (1 online resource). (3):219–227. http://dx.doi.org/10.1016/0920-4105(89)90019-3.
Siratovich, P.A., Villeneuve, M.C., Cole, J.W., Kennedy, B.M., Bégué, F., 2015. Saturated
heating and quenching of three crustal rocks and implications for thermal

You might also like