You are on page 1of 4

Ultrasonics 44 (2006) e1495e1498 www.elsevier.

com/locate/ultras

Hypersonic velocity measurement using Brillouin scattering technique. Application to water under high pressure and temperature
Frederic Decremps *, Frederic Datchi, Alain Polian
IMPMC Physique des Milieux Denses, University of Paris VI, casier 77 Campus Boucicaut, 140 rue de Lourmel, 75015 Paris, France Available online 9 June 2006

Abstract This paper presents recent improvement on sound velocity measurements under extreme conditions, illustrated by the hypersonic sound velocity measurements of water up to 723 K and 9 GPa using Brillouin scattering technique. Because water at high pressure and high temperature is chemically very aggressive, these experiments have been carried out using a specic experimental set-up. The present data should be useful to better constrain the water equation of state at high density. This new development brings high-quality elastic data in a large pressure/temperature domain, which may afterwards benet the understanding of many other elds as nonlinear acoustics, underwater sound, or physical acoustics from a more general point of view. 2006 Elsevier B.V. All rights reserved.
Keywords: Speed of sound; Brillouin scattering; Extreme conditions

1. Introduction For many years, inherent problems of carrying out elastic measurements in high pressurehigh temperature cells have prevented acoustics experiments under extreme thermodynamic conditions. Consequently, little is known about experimental sound velocity of liquids and solids at high density, data of major interest for physics, chemistry and geophysics. For example, understanding the Earth interior is one of the most evident application of high pressure work in the (0300 GPa) range (1 GPa = 10 kbar). Up to now, the deepest core sampling has been extracted at a depth of only few kilometers, and more than 99% of the Earth interior has to be studied by reproducing the thermodynamic conditions in the laboratory (Fig. 1). Beyond this application, the study of matter under pressure is a fascinating eld since pressure appears to be one of the most powerful thermodynamical parameter to tune properties for understanding fundamental physics and chemistry. As a matter of fact, the eect of pressure at room temperature
*

Corresponding author. Fax: +33 1 44 27 44 69. E-mail address: frederic.decremps@upmc.fr (F. Decremps).

on the energetic of a crystal is much greater than the temperature change up to the melting point at room pressure [1]. Particularly, the measurement of sound velocity versus pressure of solids and liquids enables to probe with a high sensitivity the most unknown part of the interatomic potential, say the repulsive one. Finally, the thermoelasticity of stressed material not only provides a way to study fundamental physics, chemistry and geophysics, but also gives crucial insight in the applied physics eld through, for example, the structural stability of solids, the indirect determination of the piezoelectric properties and the third-order elastic constants, or the mechanical properties of important material from a technological point of view. When pressure and temperature are increased, liquid water has a very interesting evolution, going from a highly-structured molecular liquid with strong hydrogen bonds, to an ionic conductor and eventually to a metal. The knowledge of its thermodynamic properties at extreme conditions is of great importance for a number of scientic problems in physics, chemistry, Earth and planetary physics or biology. However, there exists rather few data that bridges the gap between the extensive low pressure measurements made with pistoncylinder apparatuses in the

0041-624X/$ - see front matter 2006 Elsevier B.V. All rights reserved. doi:10.1016/j.ultras.2006.05.150

e1496

F. Decremps et al. / Ultrasonics 44 (2006) e1495e1498

Fig. 1. Typical pressure range that can be reached in the laboratory with high pressure apparatus. Some examples from the Earth atmosphere to the interior of planets give an idea of thermodynamical conditions.

70s and the shock-wave experiments that explore the very high pressuretemperature portion of the phase diagram. New speeds of sound experimental data are denitively necessary to accurately extend our knowledge of uid water. We present here original experimental measurements of the sound velocity and a perspective of equation of state determination in liquid water up to 723 K and 9 GPa using Brillouin scattering in a diamond anvil cell. This study is an excellent example of what can only be achieved with the Brillouin technique [2], since, up to now, ultrasonic experiment in this (P, T) range cannot be performed on liquids. These state-of-the-art experiments will certainly be useful in several applied problems and many other elds such as nonlinear acoustics, underwater sound, or physical acoustics from a more general point of view. 2. High pressure and high temperature Brillouin scattering technique In Brillouin scattering, monochromatic light interacts inelastically with thermal acoustics phonons. These hypersonic waves produce a periodic modulation of the refractive index in the material. The incident light undergoes a Doppler shift and is inelastically scattered with a frequency modication depending on the sound wave velocity and the interaction geometry. The energy range involved in Brillouin scattering from thermal excitations is of the order of 10 GHz. The corresponding measured wave number shift Dr (in cm1) is given by: 2nv sin h 2 1 Dr kc where v is the acoustic phonon velocity, n the refractive index at k (the wavelength of the exciting light), c the speed of light in vacuum and h the angle between the incident and the scattered light. Due to the small frequency shift and the small intensity of the inelastically scattered light (Brillouin) with respect to the elastically scattered one (Rayleigh), a highly dispersive power spectrometer is used, namely the Sandercock spectrometer, made of two FabryPerot interferometers in tandem. A typical spectrum recorded on water at high pressure and high temperature is shown in Fig. 2. Combined with diamond anvil cell (DAC) [3], the Brillouin technique can be used to measure the high pressure sound velocity of small (dimensions on the order of 10 lm) and (only) transparent materials [4].

Fig. 2. Backscattering Brillouin scattering spectrum of water at 3 GPa and 653 K. The central unshifted peak comes from the elastic scattering (Rayleigh peak). The frequency of the scattered light is increased in the case of a phonon annihilation (Brillouin peak labelled B2), decreased in the case of a creation (Brillouin peak labelled B1). In this geometry, knowledge of the refractive index n is necessary to calculate the sound velocity v through the Eq. (1).

The principle of high pressure generation with DAC is shown in Fig. 3 and can be briey described as the following. A metal gasket (stainless, inconel or rhenium for example) is placed between the small at parallel faces of two opposed diamond anvils. The sample (with typically a volume of 100 100 20 lm3) is conned in a hole (about 200 lm in diameter) drilled in the pre-indented gasket

Fig. 3. Diamond anvil cell principle. A metallic gasket connes the sample with the pressure gauge and the pressure transmitting medium. Diamond is chosen for its hardness and its transparency in a wide electromagnetic radiation frequency range.

F. Decremps et al. / Ultrasonics 44 (2006) e1495e1498

e1497

together with a pressure transmitting uid (e.g., helium or nitrogen). Pressure is measured in situ with accuracy better than 3% using the uorescence emission of a ruby ball [5] and/or SrB4O7:Sm2+ powder [6] placed into the gasket hole. The high quasi-hydrostatic pressure is generated on the sample as a force pushes the two diamond anvils together. In our experiments, we used a membrane DAC [7], i.e., the force on the diamonds is applied through the displacement of the piston due to a membrane deformation (produced by a pressurized helium gas lling the membrane). The highest pressure attainable depends on the area of the diamond anvil at (with 0.03 mm anvil ats, pressure of about 300 GPa can be reached). The DAC can also be tted into a resistive cylindrical heater for high temperature studies up to 1000 K, a thermocouple being glued on one end of the diamond anvil to determine the average temperature on the sample with an accuracy of a few K.

Because water at high pressure and high temperature is chemically very aggressive, these experiments have been carried out using a specic experimental technique. The diamond anvil cell, made of high temperature steel, was heated by a resistive external heater. The temperature was controlled by a thermocouple glued onto the side of one anvil. The Brillouin scattering measurements were performed in the backscattering geometry (h = p), using the 514.5 nm line of an argon laser. Water was conned in a rhenium/gold composite gasket, rhenium being well adapted to high pressurehigh temperature studies. Gold was used to prevent chemical reactions between the water sample and the rhenium gasket on one hand, and the pressure gauge (SrB4O7:Sm2+) on the other hand, the latter being dissolved at temperatures above $500 K. A ruby ball serves as an additional in situ temperature gauge (see Fig. 4). 3. Hypersonic sound velocity of water up to 723 K and 9 GPa by Brillouin scattering experiment in DAC Brillouin spectra have been collected along several isotherms between 300 K and 723 K. The measurements extend to the solidication pressure, which usually occurred less than 1 GPa above the melting point, unless for the 723 K isotherm where one of the anvils failed before completion. The experimental data have been obtained in the backscattering geometry which means that the sound velocity can be extracted only if the refractive index is known. For liquid water, the refractive index at 514.5 nm is accurately known at densities up to 1.06 g cm3 and temperatures up to 773 K, which has been conveniently expressed in a simple formulation by Schiebener et al. [8] and Harvey et al. [9]. Direct and independent measurements of the refractive index in the visible range by Dewaele et al. [10] (using the same method as in Ref. [11]) established that the latter

Fig. 4. Picture of the rhenium/gold gasket which connes H2O samples at the liquidice VII transition (T = 572 K and P = 6.2 GPa).

Fig. 5. Pressure dependence of the sound velocity in water at high temperature.

e1498

F. Decremps et al. / Ultrasonics 44 (2006) e1495e1498

Fig. 6. Dierence between the present sound velocity values (cexp) and those derived from the Saul and Wagner [13] equation of state (cSW).

formulation remains valid in the PT range spanned by our experiments. As can be observed in Fig. 5, the sound velocity is a smooth and monotonous function of pressure in the studied temperature range, showing that no rst-order liquidliquid phase transition occurs such as, for example, the one observed in liquid phosphorous [12]. This work shows however that the state-of-the-art equation of state (EoS) of water (Saul and Wagner [13]) poorly describes the sound velocity at high pressure and high temperature (Fig. 6). The dierence can be represented by the following polynomial equation: cexp cSW a1 P a2 P 2 a3 PT
3 5

the help of molecular dynamics computation using an empirical potential, such as the TIP4P model [14]: a trial value of the adiabatic to the isothermal bulk modulus ratio (c) is used to obtain the EoS which is then tted to an ad hoc potential. The second step is to use this potential to recalculate c which is nally used to correct the EoS. References
[1] R.J. Hemley, N.W. Ashcroft, Phys. Today 51 (1998) 26. [2] Also including the Stimulated Brillouin Scattering technique (SBS). See for example: S. Wiryana, L.J. Slutsky, J.M. Brownb, Earth Planet. Sci. Lett. 163 (1998) 123. [3] J.C. Chervin, B. Canny, J.M. Besson, P. Pruzan, Rev. Sci. Instrum. 66 (1995) 2595. [4] A. Polian, J. Raman Spectrosc. 34 (2003) 633. [5] G.J. Piermarini, S. Block, J.D. Barnett, R.A. Forman, J. Appl. Phys. 46 (1975) 2774. [6] F. Datchi, R. Le Toullec, P. Loubeyre, J. Appl. Phys. 81 (1997) 3333. [7] R. Letoullec, J.P. Pinceaux, P. Loubeyre, High Pressure Res. 1 (1988) 77. [8] P. Schiebener, J. Straub, J.L. Sengers, J. Gallagher, J. Phys. Chem. Ref. Data 19 (3) (1990) 677. [9] A. Harvey, J.S. Gallager, J.M.H. Levelt Sengers, J. Phys. Chem. Ref. Data 27 (4) (1998) 761. [10] A. Dewaele, J.H. Eggert, P. Loubeyre, R. Le Toullec, Phys. Rev. B 67 (2003) 094112. [11] A. Dewaele, J.H. Eggert, P. Loubeyre, R. Le Toullec, Phys. Rev. B 67 (2003) 094112. [12] Y. Katayama, T. Mizutani, W. Utsumi, O. Shimomura, M. Yamakata, K. Funakoshi, Nature 403 (2000) 170. [13] A. Saul, W. Wagner, J. Phys. Chem. Ref. Data 18 (1989) 1537. [14] W.L. Jorgensen, J. Chandrasekhar, J.D. Madura, R.W. Impey, M.L. Klein, J. Chem. Phys. 79 (1983) 926.

where a1 = 8.5096 10 , a2 = 9.9053 10 and a3 = 2.3143 105 (with the following unity for P, T and c: kbar, K and km/s). 4. Conclusion and perspective Recent improvements to measure accurately hypersonic sound velocities of liquids under extreme conditions are described. To illustrate the capability of this method, original results on sound velocity of water up to 9 GPa and 723 K are given with an accuracy of about 1%. To use the present data as a means to accurately determine the EoS, we believe that dierent approaches should be undertaken. This can be achieved for example by using a model equation of state such as the one of Saul and Wagner [13], and use the present data to better constrain the parameters of the EoS. A similar method to get a decent equation of state is to use an iterative calculation with

You might also like