JOURNAL OF APPLIED PHYSICS 111, 112609 (2012) Thermal conductivity of argon at high pressures and high temperatures Alexander F. Goncharov,1 Michael Wong,1,2 D. Allen Dalton,1 J. G. O. Ojwang,1 Viktor V. Struzhkin,1 Zuzana Konôpková,3,4 and Peter Lazor3 1 Geophysical Laboratory, Carnegie Institution of Washington, Washington, D.C. 20015, USA Department of Earth & Planetary Science, University of California, Berkeley, Berkeley, California 94720, USA 3 Department of Earth Sciences, Uppsala University, SE-752 36 Uppsala, Sweden 4 Deutsches Elektronen Synchrotron (DESY), 22607 Hamburg, Germany 2 (Received 15 April 2011; accepted 13 November 2011; published online 15 June 2012) Knowledge of the thermal conductivity of Ar under conditions of high pressures and temperatures (P-T) is important for model calculations of heat transfer in the laser heated diamond anvil cell (DAC) as it is commonly used as a pressure transmitting medium and for thermal insulation. We used a modified transient heating technique utilizing microsecond laser pulses in a symmetric DAC to determine the P-T dependent thermal conductivity of solid Ar up to 50 GPa and 2500 K. The temperature dependent thermal conductivity of Ar was obtained by fitting the results of finite element calculations to the experimentally determined time dependent temperature of a thin Ir foil surrounded by Ar. Our data for the thermal conductivity of Ar are larger than that theoretically calculated using the Green-Kubo formalism, but they agree well with those based on kinetic theory. These results are important for ongoing studies of the thermal transport properties of C 2012 American Institute of minerals at pressures and temperatures native to the mantle and core. V Physics. [http://dx.doi.org/10.1063/1.4726207] INTRODUCTION Knowledge of the thermal conductivity of materials under extreme conditions of high pressure and high temperature is important for a number of applications, including geoand planetary sciences, technology, and industry. At ambient pressures and variable temperatures, there are well established experimental techniques that provide very accurate data on the thermal transport properties using both contact and noncontact methods.1–5 Remarkably, thermal properties of very thin (nm scale) films can be also measured using ultrafast laser pump-probe techniques.6,7 Measurements of the thermal conductivity at high P-T conditions remain a challenging task because of limitations in space and sample access imposed by high-pressure devices. Moreover, the thermochemical properties of all materials change at high P-T conditions, which cause additional complications in determining thermal transport properties. Currently, there are well established techniques for measuring thermal conductivity at moderate P (<20 GPa) and high-T in large presses (e.g., Ref. 8). The Ångstrom technique uses the standard cylindrical sample geometry and determines the thermal conductivity analytically, based on the phase and amplitude shift of the temperature wave created at the sample axis. At higher pressures, where diamond anvil cells (DACs) are commonly used, this technique is difficult to use due to the necessity of very local temperature probes. For this reason, extrapolations based on scaling relations4,9 and theoretical calculations9–12 are mainly used under P-T conditions relevant to the planetary interiors. Measurements of the thermal conductivity (more accurately, thermal diffusivity D ¼ K/(q Cp), where K is the thermal conductivity, q is the density, and Cp is the specific heat capacity) in the DAC are rare.13,14 Model calculations of the heat fluxes through the 0021-8979/2012/111(11)/112609/6/$30.00 sample cavity15–17 are the necessary part of these measurements due to the complex sample geometry and limited capabilities of temperature measurements in the sample cavity of the DAC. Measurements of thermal conductivity in the DAC are based on the determination of thermal gradients across a sample of known thickness or on the heat transfer rate from the heated material spot. In the former case, a continuous heating method is appropriate, and the thermal conductivity can be deduced from simultaneous radiative temperature measurements from both sample sides when the sample is heated from one side.14 The results depend moderately on the thermal conductivity of the medium, so better knowledge of this key parameter would increase the accuracy of the measurements. In the latter case, time-dependent radiative temperature measurements are needed. For materials transparent to laser heating, the laser radiation should be absorbed by a material with a short skin depth (coupler). The absorptive and thermochemical properties of the coupler can also affect the measurements, so these in principle need to be known under conditions of high P-T to make the measurements more accurate. In the transient heating technique (THT),13 a 10 ns long laser pulses was employed, which was assumed to thermalize (make temperature even) the coupler much faster than the pulse duration thus making the thermal diffusivity of the coupler material unimportant for determination of the temperature decay rate. In these measurements and the following finite element (FE) model calculations, the energy of the laser pulse has not been used to constrain the temperature rise of the coupler; instead, it was used as a free parameter in the fitting of the measured temperature-time dependencies. In subsequent work using the THT on other materials,18 it was determined that the use of energetic ns pulses (which are needed to heat the sample to high temperatures, which can be measured radiometrically) 111, 112609-1 C 2012 American Institute of Physics V Downloaded 25 Jun 2012 to 206.205.250.4. Redistribution subject to AIP license or copyright; see http://jap.aip.org/about/rights_and_permissions 112609-2 Goncharov et al. creates non-equilibrium conditions on the coupler-sample (medium) interface. These conditions cannot be described by FE calculation using classical heat transfer equations, thus diminishing the accuracy of the results. It is interesting that classic heat transfer formalism works very well for much shorter (100 fs) laser pulses (albeit of much lower energy), which are used for the time-domain-thermoreflection (TDTR) measurements of the thermal conductivity, including those under high pressures.19,20 Those measurements also need an input for the thermal conductivity of the medium and the coating material (usually metal). Here, we address the issues mentioned above and the current needs to improve the accuracy of the thermal conductivity measurements in the DAC. We utilized a modified transient heating technique by using pulses of microsecond duration (which have much lower peak power) to determine the thermal conductivity of Ar surrounding the metallic coupler. We have chosen Ar because it is a common pressure transmission medium for the laser experiments (e.g., Ref. 21), and because its thermal conductivity and other thermochemical properties have been recently calculated using modern molecular dynamics simulation techniques in a wide P-T range,22 thus making an easy comparison available. Previous experimental works were all performed at ambient pressure.23–25 We determined the thermal conductivity of solid Ar up to 50 GPa and 2500 K by fitting the measured time-dependent temperature of the coupler, which in this case followed the classic heat transfer equations. Unlike the previous THT works,13,18 the value of the laser pulse energy served as an important constraint in this determination. EXPERIMENTAL PROCEDURE We used a symmetric DAC with diamond anvils possessing flat culets of 300 lm diameter up to 50 GPa and 2500 K. Rhenium foil of 300 lm initial thickness preindented to 20 GPa served as a gasket. A thin Ir foil (1 lm thick) of approximately 80 lm diameter was prepared by compressing a small amount of material between two diamond anvils to the desired thickness controlled by spectral measurements of the interference fringes. The foil was positioned in a recessed gasket hole (Fig. 1) as parallel as possible to the diamond tips and approximately axially centered between them to insure that there is no direct contact between the foil and diamond anvils and to reduce the axial temperature gradients. Pressure was determined by the ruby J. Appl. Phys. 111, 112609 (2012) fluorescence technique at room temperature and no correction for the thermal pressure has been made. The sample cavity was filled with Ar (loaded at room temperature at 0.2 GPa), and the iridium foil served as a laser absorber (coupler) to pump thermal energy into the sample (surrounding Ar). To determine the temperature history of the sample, we measured the iridium coupler’s time resolved thermal emission spectra as a function of time. We used a custom optical microscope system similar to that described in Ref. 13, except that achromatic glass optics were used. The fiber laser power was finely controlled using a k/2 waveplate and polarizing cube beamsplitter. The laser power at the sample position was calibrated in a separate experiment using a power meter positioned at the sample location. The laser power reaching the sample within the DAC was corrected for transmission of the DAC using the Fresnel formula. The laser beam profile in the focal spot was determined by a laser beam profiler. The determined beam spot profile was approximated to a 2D round Gaussian function of 22 lm full width (at 1/e height). We used 6 ls width pulses from electronically modulated Yb-based fiber laser21 operating at 10 kHz repetition rate. This allowed the coupler temperature to return to essentially room temperature between laser pulses. This is verified by the FE calculations, which reproduces the temperature history of the coupler in the available measurement range. The temporal pulse profile (measured by a photodiode and oscilloscope) was kept constant for the whole set of measurements and used in FE calculations (approximated by a sum of the Gaussians). The thermal emission spectra were accumulated using a 300 mm focal length spectrograph with 300 gr/mm grating equipped with a gated CCD detector.13 The detector gate was synchronized with the laser pulses. The spectra were measured with a 500 ns gate time by averaging the laser heating events for a total accumulation time between 0.1 to 120 s. The emission spectra measured within the spectral window of 550-710 nm were energy calibrated using a standard lamp with the NIST calibrated spectral irradiance. An automation program controlling the delay generator electronically adjusted the temporal delay of the detector gate relative to the heating laser pulse. The temperature was determined synchronously with the variation of the time delay using a Wien’s fitting procedure (Fig. 2).26 By performing this fit as a function of electronic delay of the spectrometer gate, we are able to collect a full history of the thermal evolution of the sample. Our system allows measurement of the radiative temperatures as low as 1300 K. The temperature determination uncertainty varies with temperature: It is approximately 50 K above 1900 K and up to 150 K approaching the low temperature limit of the measurements. This has been estimated based on the scatter between different consecutive measurements. COMPUTATIONAL METHODS FIG. 1. The DAC sample schematic. Ar sample fills the volume above and below the Ir foil (coupler). Since the coupler has a rectangular shape, Ar can freely move between the upper and the lower pockets. We used a commercial FE solver code (FlexPDE6 3D professional, PDE Solutions, Inc.) for modeling the timedependent heat fluxes in the DAC cavity (e.g., Ref. 27). The Downloaded 25 Jun 2012 to 206.205.250.4. Redistribution subject to AIP license or copyright; see http://jap.aip.org/about/rights_and_permissions 112609-3 Goncharov et al. J. Appl. Phys. 111, 112609 (2012) FIG. 3. The distances between the Ir foil and diamond culets measured using the spectral distance between the interference fringes. The fits applied to the data are used for the sample cavity dimensions in the FE calculations. FIG. 2. Representative radiometric time-resolved temperature measurements. The temperature is determined as an inverse slope of the linear lines fitted to the data. The data represent the thermal emission spectra (Ik) transformed as shown and plotted as a function of an energetic variable. C1 and C2 are first and second radiation constant with values of C1 ¼ 119.1044 (W nm2), C2 ¼ 1.4388 107 (nm K), respectively. based on the high melting temperature at ambient pressure (2683 K) and the results of our experiments (e.g., Ref. 31). Using FE calculation methods, we simulated the heat flux transfer in the DAC cavity using the experimentally geometrical parameters used for calculations were determined experimentally; the thermochemical parameters were chosen as described below. The thermal conductivity of Ar was approximated as suggested in the theoretical calculations22 and then fit to the experimental data (2-parameters fit) by forward modeling. The results of calculations obtained using PDE solver were confirmed using the software package COMSOL Multiphysics,14 which gave nearly identical results. RESULTS AND DISCUSSION Two experimental runs were performed ranging from 10 to 50 GPa. At each pressure point, we determined the sample geometry by measuring the cavity thickness and the distances between the Ir foil and diamond culets using the optical interferometry technique (Fig. 3). The pressure dependences of these parameters have been fit and thus averaged values have been used in the FE calculations (see below). The data show some scatter related to the foil deformation under pressure, resulting in spatially variable foil–diamond culet distances. We neglect these effects in our simplified FE model calculations. The refractive index of Ar, which is required for calculation, was taken from the Brillouin data of Ref. 28 and was extrapolated to higher pressure as described in Ref. 29. The laser heating experiments were performed to variable maximum temperatures depending on the nominal pressure. In the experimental runs used for comparison with FE calculations (Fig. 4), we did not want to exceed the melting temperature of Ar30 to avoid probing the fluid state, which can have substantially different thermochemical parameters. This would make our model calculations less reliable. We assumed that Ir has a higher melting temperature than Ar FIG. 4. Temperature history of the DAC for pulse laser heating at 43 GPa. Many such plots were constructed for pressures up to 50 GPa. Radiometric data from the Wien’s fits (the error bars are the temperature determination uncertainties) illustrate an increase in the sample temperature corresponding to the front edge of the laser pulse and then plateaus before decaying below the detection limit. The thick solid line is the results of the FE calculations, which represent the best fit to these data yielding the following parameters for the temperature dependent thermal conductivity of Ar: K300 ¼ 72 W/(K m), m ¼ 1.35—see Table I for the description of parameters. The best fit to the data calculated in the assumption that the emissivity of Ir decreased by 10% at 43 GPa essentially coincides with this curve (not shown); the parameters yielded are K300 ¼ 79 W/(K m), m ¼ 1.7. The thermal history, calculated taking into account the isobaric changes in density of Ar and Ir and the isobaric change in the thermal heat capacity of Ir with temperature, is shown by a thin dashed line. The temporal profile (a.u. of intensity) of the incident laser pulse (measured via a photodiode) is also included. Downloaded 25 Jun 2012 to 206.205.250.4. Redistribution subject to AIP license or copyright; see http://jap.aip.org/about/rights_and_permissions 112609-4 Goncharov et al. J. Appl. Phys. 111, 112609 (2012) determined geometric and laser heating parameters (Fig. 3). This also includes the laser power needed to reach the measured temperature-time profiles (Fig. 4). The thermochemical parameters and optical emittance data of Ir at ambient pressure and high temperature are available in the literature.32 The thermal conductivity of Ir at high pressures was determined by scaling these data to high pressures using the available equation of state31 and scaling relations.33 We assumed that the temperature dependence of the thermal conductivity of Ir (taken from Ref. 32) changes congruently with pressure (Table I). As there are no data, we assumed that the temperature dependent optical properties of Ir32 are pressure independent. A possible departure from this assumption would affect the final results for the thermal conductivity of Ar as the absorbed laser pulse energy used in our calculations is directly proportional to the optical emissivity of Ir. The uncertainty in the value of the thermal conductivity, which results from the unknown emissivity of Ir at high pressures, has been estimated by performing calculations with the emissivity varied by 10% of the nominal value (maximum estimated pressure induced change) (Fig. 4); the results obtained are within the error bars shown in the Figs. 5 and 6. To evaluate whether the thermal expansion of Ar and Ir would affect our results intrinsically through the change in density with temperature, we performed finite element calculations using the thermal equations of state, which were determined experimentally for Ir31 and theoretically for Ar.22 We found that the calculated changes in time profiles are small (within the experimental uncertainties in measured temperatures) (Fig. 4), so they can be neglected. We did not attempt calculating the extrinsic effects related to the change in the interface distances as such calculations are beyond our current capabilities. We also neglected the temperature dependences of the specific heat capacities (Table I) as these are expected to be small for Ar in the temperature range studied,22 while for Ir the use of the temperature dependent specific heat capacity32 affects the calculated time series very little (Fig. 4). Finally, we neglected the stress and temperature dependences of the thermochemical parameters of diamond (it remains at almost 300 K in all calculations). An initial estimate for the Ar thermal conductivity was applied using the theoretical work.22 The temperature dependent thermal conductivity of Ar was determined by fitting the results of FE calculations to the experimentally determined time dependent coupler temperature for each experimental pressure point (Figure 4). The uncertainty in the FIG. 5. Thermal conductivity as a function of temperature. Dashed lines show the uncertainty interval. The points indicate thermal conductivity results from MD simulations from Ref. 22 and the corresponding line is the fit to these results. thermal conductivity was determined by finding the range of parameters for which the FE model and the experimental curves agree within the experimental uncertainty. In comparing the dependence with temperature of our thermal conductivity data to that of Tretiakov and Scandolo,22 we find that (on examples of the results for 10 and 50 GPa—Fig. 5) the thermal conductivity is higher than the results of the Green-Kubo (GK) MD simulations at all temperatures. However, our data agree well with the theoretical calculations obtained using kinetic theory22 (Fig. 6). From the log(K)-log(T) plots, simulations give a slope of 1.3 while our data give a slope of 1.4. With respect to the pressure dependence of thermal conductivity for Ar at 300 K (Figure 6), our data have a similar dependence on a log(K)log(P) scale (theoretical value 1.29 vs. our data 1.24). The pressure effect on the thermal conductivity can be estimated using the Leibfried-Schlömann (LS) equation K¼A V 1=3 xD 3 ; c2 T (1) TABLE I. Thermochemical parameters of materials used in model FE calculations. Property/material Diamond Density, q0 (kg/m3) Bulk modulus, B0 (GPa) BM derivative, B00 Thermal conductivity (W/(m K)) Specific heat capacity (J/(kg K)) Emissivity 3500 2000 509 0 Sample/medium (Ar) 1809 5.513 5.20 K300 (300/T)m 568.7 0 Coupler (Ir) 22 650 390.5 3.26 (q/q0)4 (147 þ 3.2 106 (T 300) 7 1010 (T 300)2) 130 0.551 (5.661 106 K1) T þ (8.586 109 K2) T2 (5.784 1012 K3) T3 Downloaded 25 Jun 2012 to 206.205.250.4. Redistribution subject to AIP license or copyright; see http://jap.aip.org/about/rights_and_permissions 112609-5 Goncharov et al. FIG. 6. Thermal conductivity as a function of pressure at 300 K. The data points (closed circles) are those determined in this work at 300 K with the corresponding errors. The separate solid and dashed lines show the results of the MD calculations using the Green-Kubo method and kinetic theory, respectively.22 The squares and the thin dashed line show the LeibfriedSchlömann equation (1) fit to our data. where V is the volume, xD is the Debye frequency, c is the Grüneisen parameter, T is the temperature, and A is a constant that is independent of pressure.34 We estimated the values on the right side of Eq. (1) at 300 K based on the available experimental data29,35 and by fitting the constant A to our 300 K data. The results are shown in Fig. 6. The slope of the thermal conductivity dependence on pressure obtained from Eq. (1) agrees well with our experimental data. Moreover, the slope agrees well with the scaling expression obtained from kinetic theory (e.g., Refs. 15 and 22). It is remarkable that our data for the temperature dependence agree reasonably well with the results of theoretical calculations22 and ambient pressure experimental data.23 The departure of the temperature dependence of the thermal conductivity from a common T1 form has been found at ambient pressure23–25 and attributed to various mechanisms such as, for example, the effects of thermal expansion on the lattice vibrational frequencies.23 Our measurements show that this departure holds at high pressures (see also Ref. 22) and possibly even increases with pressure. Our high-pressure experiments cover a much wider temperature range than that in the ambient pressure studies,23–25 mainly probing the classical high-temperature region T HD (HD is the Debye temperature; HD ¼ 85 K at 0 GPa, increasing slightly above 300 K at 50 GPa36), thus making difficult the direct comparison. CONCLUSIONS Our flash laser heating measurements in the DAC with ls pulses determined the thermal conductivity of Ar based on good agreement between experimentally observed temperature history and the results of finite element calculations with realistic geometric and thermochemical parameters. The experiment has been designed to minimize the effect of J. Appl. Phys. 111, 112609 (2012) the unknown high-pressure optical emissivity and thermal conductivity of the Ir coupler, while other thermochemical parameters could be constrained rather accurately. Using the theoretically predicted power dependencies of the thermal conductivity of Ar on temperature, we find good agreement between experimental and FE models calculated temperature histories. Knowledge of the laser power and laser focal spot shape allowed constraining the Ar thermal conductivity values. We find that larger low-temperature values and slightly faster temperature dependencies of thermal conductivity better match our observations compared to the GK theoretical predictions, while our data nearly coincide with the theoretical results based on kinetic theory. The effect of thermal pressure increase at high temperature was found to be negligible for the current determination of thermal conductivity. The measurements presented have several sources of uncertainty, which could affect the results. We believe that the major one is related to inhomogeneity of the optical properties of Ir coupler arising from micro-defects created during previous heating cycles or because of coupler deformation. Larger than nominal Ir emissivity values would result in increased values of the thermal conductivity. Accurate thermal conductivity data of Ar at high pressures and high temperatures are essential to future measurements of material thermal conductivity under extreme P-T conditions using the modern ultrafast pump-probe TDTR technique.19,20 Ar is a commonly used and convenient pressure medium; recent measurements of the thermal conductivity of MgO show that it can be used at high pressures up to 60 GPa.37 Measurements of the thermal conductivity of Fe using the static laser heating technique14 would also benefit from these new data. Our measurements open the possibility of using Ar under extreme P-T conditions with higher confidence using the experimentally determined values of the thermal conductivity. ACKNOWLEDGMENT We thank R. S. McWilliams for important comments and suggestions on the manuscript. We acknowledge support from NSF EAR 0711358 and EAR 1015239, Carnegie Institution of Washington, Army Research Office, STINT IG2010-2 062, and DOE/NNSA (CDAC). 1 W. J. Parker, R. J. Jenkins, C. P. Butler, and G. L. Abbott, J. Appl. Phys. 32(9), 1679–1684 (1961). 2 S. Andersson and G. Backstrom, Rev. Sci. Instrum. 57(8), 1633–1639 (1986). 3 M. Pertermann and A. M. Hofmeister, Am. Mineral. 91(11–12), 1747–1760 (2006). 4 A. M. Hofmeister, Proc. Natl. Acad. Sci. U.S.A. 104(22), 9192–9197 (2007). 5 D. G. Cahill and R. O. Pohl, Phys. Rev. B 35(8), 4067 (1987). 6 Y. K. Koh, S. L. Singer, W. Kim, J. M. O. Zide, H. Lu, D. G. Cahill, A. Majumdar, and A. C. Gossard, J. Appl. Phys. 105(5), 054303 (2009). 7 K. Kang, Y. K. Koh, C. Chiritescu, X. Zheng, and D. G. Cahill, Rev. Sci. Instrum. 79(11), 114901 (2008). 8 Y. Xu, T. J. Shankland, S. Linhardt, D. C. Rubie, F. Langenhorst, and K. Klasinski, Phys. Earth Planet. Inter. 143–144, 321–336 (2004). 9 N. de Koker, Earth Planet. Sci. Lett. 292(3–4), 392–398 (2010). 10 N. de Koker, Phys. Rev. Lett. 103(12), 125902 (2009). 11 X. Tang and J. Dong, Phys. Earth Planet. Inter. 174(1–4), 33–38 (2009). Downloaded 25 Jun 2012 to 206.205.250.4. Redistribution subject to AIP license or copyright; see http://jap.aip.org/about/rights_and_permissions 112609-6 12 Goncharov et al. X. L. Tang and J. J. Dong, Proc. Natl. Acad. Sci. U.S.A. 107(10), 4539–4543 (2010). 13 P. Beck, A. F. Goncharov, V. V. Struzhkin, B. Militzer, H.-K. Mao, and R. J. Hemley, Appl. Phys. Lett. 91(18), 181914 (2007). 14 Z. Konôpková, P. Lazor, A. F. Goncharov, and V. V. Struzhkin, High Press. Res. 31(1), 228–236 (2011). 15 M. Manga and R. Jeanloz, J. Geophys. Res. 102(B2), 2999–3008, doi: 10.1029/96JB02696 (1997). 16 A. Dewaele, G. Fiquet, and P. Gillet, Rev. Sci. Instrum. 69(6), 2421–2426 (1998). 17 B. Kiefer and T. S. Duffy, J. Appl. Phys. 97(11), 114902 (2005). 18 A. F. Goncharov, V. V. Struzhkin, J. A. Montoya, S. Kharlamova, R. Kundargi, J. Siebert, J. Badro, D. Antonangeli, F. J. Ryerson, and W. Mao, Phys. Earth Planet. Inter. 180(3–4), 148–153 (2010). 19 W.-P. Hsieh, B. Chen, J. Li, P. Keblinski, and D. G. Cahill, Phys. Rev. B 80(18), 180302 (2009). 20 B. Chen, W.-P. Hsieh, D. G. Cahill, D. Trinkle, and J. Li, Phys. Rev. B 83, 132301 (2011). 21 A. F. Goncharov, V. B. Prakapenka, V. V. Struzhkin, I. Kantor, M. L. Rivers, and D. A. Dalton, Rev. Sci. Instrum. 81(11), 113902 (2010). 22 K. V. Tretiakov and S. Scandolo, J. Chem. Phys. 121(22), 11177–11182 (2004). 23 D. K. Christen and G. L. Pollack, Phys. Rev. B 12(8), 3380–3391 (1975). 24 I. N. Krupskii and V. G. Manzhelii, Sov. Phys. JETP 28, 1097 (1969). 25 F. Clayton and D. N. Batchelder, J. Phys. C 6, 1213–1228 (1973). J. Appl. Phys. 111, 112609 (2012) 26 C.-S. Zha and W. A. Bassett, Rev. Sci. Instrum. 74(3), 1255–1262 (2003). 27 A. F. Goncharov, J. A. Montoya, N. Subramanian, V. V. Struzhkin, A. Kolesnikov, M. Somayazulu, and R. J. Hemley, J. Synchrotron Radiat. 16, 769–772 (2009). 28 M. Grimsditch, R. Letoullec, A. Polian, and M. Gauthier, J. Appl. Phys. 60(10), 3479–3481 (1986). 29 M. Grimsditch, P. Loubeyre, and A. Polian, Phys. Rev. B 33(10), 7192 (1986). 30 R. Boehler, M. Ross, P. Soderlind, and D. B. Boercker, Phys. Rev. Lett. 86(25), 5731–5734 (2001). 31 A. F. Goncharov, J. C. Crowhurst, J. K. Dewhurst, S. Sharma, C. Sanloup, E. Gregoryanz, N. Guignot, and M. Mezouar, Phys. Rev. B 75(22), 224114 (2007). 32 C. Cagran and G. Pottlacher, Int. J. Thermophys. 28(2), 697–710 (2007). 33 R. G. Ross, P. Andersson, B. Sundqvist, and G. Backstrom, Rep. Prog. Phys. 47(10), 1347–1402 (1984). 34 M. Roufosse and P. G. Klemens, Phys. Rev. B 7(12), 5379–5386 (1973). 35 M. Ross, H. K. Mao, P. M. Bell, and J. A. Xu, J. Chem. Phys. 85(2), 1028–1033 (1986). 36 A. I. Karasevskii and W. B. Holzapfel, Phys. Rev. B 67(22), 224301 (2003). 37 D. A. Dalton, A. F. Goncharov, W. Hsieh, and D. Cahill, in Abstract MR13A-1899 presented at 2010 Fall Meeting, AGU, San Francisco, CA, 13–17 December 2010. Downloaded 25 Jun 2012 to 206.205.250.4. Redistribution subject to AIP license or copyright; see http://jap.aip.org/about/rights_and_permissions
© Copyright 2025 Paperzz