CP620, Shock Compression of Condensed Matter - 2001 edited by M. D. Furnish, N. N. Thadhani, and Y. Horie © 2002 American Institute of Physics 0-7354-0068-7/02/$ 19.00 SHOCK TEMPERATURE OF NaCl MEASURED WITH WIDE-BAND OPTICAL RADIOMETRY Toshiyuki Ogura1, Kazutaka G. Nakamura1, Hisataka Takenaka2, and Ken-ichi Kondo1 Materials and Structures Laboratory, Tokyo Institute of Technology, 4259 Nagatsuta, Midori, Yokohama 2268503, Japan 2 NTT Advanced Technology Corporation, 3-9-11, Midori, Musashino, Tokyo 180-8585, Japan Abstract. Shock temperature of NaCl was measured with time-resolved (3-nanosecond resolution) wide-band optical radiometry observing the radiation ranging from 0.6 to 13 jim in the pressure range between 17 and 43 GPa. In case of samples above 2000 K, the emitted radiation was measured with a visible to near-IR radiometer (0.6-1.6 jum). This radiation associated with a phase transition which arises at the defect sites such as dislocations. An IR radiometer sensitive to 9-13 urn was used to measure the bulk temperature below 1000 K. In the mixed-phase region between 23 and 33 GPa, thermal heterogeneity was observed in shock-loaded NaCl. In the low-pressure phase (Z?l) between 17 and 21 GPa, the shock temperature obtained with IR radiometer was almost 400 K lower than the values obtained by Fritz et al.1 problem for measuring low shock temperature below 2000 K because of an insufficient emission in a visible wavelength range. For NaCl in the pressure range below 45 GPa, while the radiometric technique revealed the nature of local hot spot,3 it was impossible to determine the bulk temperature. The thermal information of the bulk is indispensable not only for constructing EOS but also for understanding the mechanism of the phase transition under dynamic high pressure loading. In a previous study, the infrared radiometry was developed for the purpose of observing low shock temperature.4 In this study, we measured the shock temperature of NaCl by observing shock-induced emission at the wavelength between 0.6 and 13 |um. An isothermal compression curve was derived from experimentally determined shock temperature data and compared with the available NaCl pressure standards. INTRODUCTION Mechanical properties of sodium chloride (NaCl) under high pressure have been pursued by various methods, x-ray analysis, shock and particle velocity measurements, and ultrasonic measurements. These experimental devotions were succeeded in the construction of high pressure EOS of NaCl by Decker,2 which is utilized as a pressure standard below 30 GPa. The shock wave data by Fritz et al.l play an important role on this pressure standard. In their treatment, the shock temperature, however, was derived from Hugoniot with assumptions on thermodynamic parameters. It is desirable to determine the shock temperature experimentally without any assumption. Optical radiometry has been usually used for shock temperature measurement. It has, however, a 1215 0.10 (a) £0.05 0.00 4-COLOR RADIOMETER - IR RADIOMETER 1 0 1 Time [>s] 2 3 FIGURE 1. The schematic drawing of the experimental setup for shock temperature measurement. EXPERIMENT AND DATA ANALYSIS Shock wave was generated by a plate-impact method with a flyer accelerated with 20mm-bore, double-stage light-gas gun at Tokyo Institute of Technology. The copper impactor 1.7-3.7 km/s brought NaCl single crystal ([100] or [111] perpendicular to the shock wave front) to the pressure 17-43 GPa. Emission between 0.6 and 13 urn from shockloaded NaCl was observed using wide-band optical radiometric system shown in Fig. 1. The system consisted of the 4-channel visible - near IR radiometer (3-ns temporal resolution, silicon and InGaAs optical device) sensitive to 0.6-1.6 um emission and the 1-channel IR radiometer (HgCdTe device cooled to 76 K) sensitive to 9-13 jam. The temporal change of the IR (9-13 jam) emission was analyzed in order to derive the temperature with the equation; 0.2 0.4 Time [us] FIGURE 2. The temporal change of the emission from shockloaded NaCl. (a) infrared (9-13 jam) at 21 GPa, (b) near IR (1.1 \im) at 24 and 43 GPa. (2) where A, /?, T are absorption, reflection, and transmission of the sample, respectively. The temporal profile of 0.6-1.6 um emission changed with the shock pressure as shown in Fig. 2 (b). While the emission obeyed Eq. (1) for the pressure above 40 GPa, it peaked during the shock duration for the pressure below 35 GPa. The 4-color 0.6-1.6 um spectrum was fitted to the graybody spectrum; ', CD where & is the emissivity and /^ is the corresponding blackbody radiation. as and av are absorption coefficients of the shocked and the unshocked NaCl, respectively, d is the initial thickness of the sample. The experimental record is shown in Fig. 2 (a) along with the least square fit to Eq. (1). In order to determine the value of £, the reflectivity of the shock wave front was measured and Kirchhoff s relation was applied; s^e*'"-\ (3) to determine the temperature, T, and the emissivity, £, where A is wavelength and Cl and C2 are constants (Q= 1.191xlO-16 W-m'-sr-1, C2= 1.439xlO-2m-K). 1216 5000 Hugoniot temperature below the phase transition pressure 23 GPa. This is explained as follows; the expected temperature is obtained by separating the increase in internal energy induced by shock wave into lattice compression part and thermal part, o c 4000 £ 3000 8 shock ~ ^^compression "^ ^tL thermal • ! &2000 —V— IR A NIRflOO] T IRflll] O Schmitt A Kondo D Kormer O Ahrens ———Fritz •••••Al'tshuler ---Isentrope 1000 0 10 20 30 40 50 60 70 (4) On the other hand, it is reported that a large number of defects such as slip surfaces, dislocations and point defects are generated in the shock-compressed solids. In considering dislocation, formation of dislocations will consume the shock energy as the form of strain energy. The new term should be added toEq.(4), 80 Pressure [GPa] FIGURE 3. The shock temperatures of NaCl below 80 GPa determined with the wide-band optical radiometry along with the data in the literatures. shock ~ * compression + A/i thermal defect . (5) Comparing with Eq. (4), the thermal part of the internal energy can decrease the shock temperature. Assuming the formation of the screw dislocation as an example, the elastic strain energy per unit length is written as; RESULTS The experimental results are shown in Fig. 3. Above 35 GPa, the temperatures measured by the IR (9-13 urn) observations are comparable with those from the near-IR (0.6-1.6 um). These agree well with the results by Kormer.5 Between 23 and 33 GPa, where the crystal is in the mixed phase state of Bl (NaCl) and B2 (CsCl) structures, the results obtained by the two methods differ from each other. The IR temperatures keep constant value around 800 K. In contrast, the nearIR temperatures are significantly higher values from 2000 to 4000 K and the brightness of the near-IR emission is small (the emissivity is in the order of 0.01). All the values of temperatures differ from the expected Hugoniot temperature obtained by Fritz et al. (solid line in Fig. 3). The IR temperatures become lower and approach to the temperatures for the isentropic compression (dashed line) below 23 GPa. The 0.6-1.6 um emission becomes weaker abruptly and is below the sensitivity of the radiometer. The reflectivity of the shock wave front is 44% at maximum at 19 GPa for the IR emission range. ,,5 Gb ,( R E =——ln| — 4n (6) where G, b, R, r0 are shear modulus, Burgers vector, grain size including dislocation considered, and radius of dislocation core (almost equal to 5&), respectively. Using Eq. (6) and the deviation of the temperature from the expected Hugoniot temperature, the dislocation density in shocked NaCl is estimated to be 1017 m"2. This value is larger than the typical density of 109 - 1016 in 2 . The point defect-dislocation and the dislocation-dislocation interactions may decrease the density. In the mixed phase region between 23 and 33 GPa, the measured temperatures, both the IR and the near-IR measurements, abruptly increase. This will be explained by the disappearance of dislocations and the release of the strain energy. It was reported that the dislocation of the order of 1016 m"2 existed in the shock-loaded steel recovered from 10.4 GPa and that the density decreased when the shock pressure exceeded the transition pressure because of the DISCUSSIONS The temperature obtained with the IR radiometer is significantly lower than the expected continuum 1217 It is revealed that the EOS obtained from shock temperature measurements is above the others by 5.9% in the pressure range 16-20 GPa when the reflectivity of the shock front is negligible. Even if the reflectivity of the shock front is assumed to be 44%, which is the maximum value measured at 19 GPa, our EOS is still higher by 4.9%. This value is larger than the experimental error. It is, therefore, concluded that the isotherm may be shifted to higher pressure in the pressure range of 16-20 GPa. By fitting the Birch-Murnaghan equation;8 30 20 O 10 ———Decker (1971) ——— ftitz(1971) — •••Brown (1999) • This work ——— RtofB-Meq. 0.6 0.8 V/Vo 0.7 0.9 1.0 (8) FIGURE 4. The isothermal compression curve at room temperature obtained from Eq. (8) along with the available NaCl pressure standards. V0 it is possible to estimate the pressure derivative of bulk modulus, K0". Using K0= 23.84 GPa and K0' = 5.352 (obtained from ultrasonic measurement below 3 GPa9), K0" becomes -0.30 GPa'1. This value is smaller than that obtained from ultrasonic measurements, -0.68 GPa"1. Using the EOS by Fritz et al., K0" becomes -0.46 GPa"1. rearrangement of atoms.6 The energy released from dislocations will be converted to the thermal energy and increase the temperature. Therefore, the disappearance of dislocations and the release of the elastic strain energy can generate the high temperature regions heterogeneously scattered in shocked solid. The extremely high temperature ranging 2000-4000 K observed with the near-IR (0.6-1.6 jLim) radiometer is considered to be due to these hot spots. ACKNOWLEDGEMENT This work was supported by Core Research for Evolutional Science and Technology (CREST) program of Japan Science and Technology Corporation (1ST). DERIVATION OF THE 293 K ISOTHERMAL COMPRESSION CURVE FOR NaCl Using the temperature measurement data of NaCl below the transition pressure, 293 K isothermal compression curve was derived and compared with the available NaCl pressure scale, i.e., the calculation from atomic model by Decker2 and Brown,7 and the shock wave data by Fritz et al.1 The measured shock temperature was utilized to convert the shock wave EOS to the isotherm under room temperature using the equation, dP_ dE , REFERENCES 1. Fritz, J. NL, Marsh, S. P., Carter, W. J., and McQueen, R. G., in Accurate Characterization of the High Pressure Environment, edited by E. C. Lloyd, National Bureau of Standards, Washington DC, 1972, Vol. 326, pp. 201208. 2. Decker, D. L., J. Appl. Phys. 42, 3239 (1971). 3. Schmitt, D. R., Ahrens, T. J., and Svendsen, B., J. Appl. Phys. 63,99(1988). 4. Fat'yanov, O. V., Ogura, T, Nicol, M. R, Nakamura, K. G., and Kondo, K., Appl. Phys. Letters 77, 960 (2000). 5. Kormer, S. B., Sov. Phys. Uspekhi 11, 229 (1968). 6. Huo, D. T. C., and Ma, C. H., /. Appl. Phys. 46, 699 (1975). 7. Brown, J. M., J. Appl. Phys. 86, 5801 (1999). 8. Birch, F., J. Geophys. Res. 83, 1257 (1978). 9. Spetzler, H., Sammis, C. G., and O'Connell, R. J., J. Phys. Chem. Solids 33, 1727 (1972). (7) with the Debye model of heat capacity and the constant y/y . The result is shown in Fig. 4. 1218
© Copyright 2025 Paperzz