Mechanical equation-of-state data of initially liquid and solid CO2 shock-compressed to terapascal conditions are reported. Diamond-sapphire anvil cells were used to vary the initial density and state of CO2 samples that were then further compressed with laser-driven shock waves, resulting in a data set from which precise derivative quantities, including Grüneisen parameter and sound speed, are determined. Reshock states are measured to 800 GPa and map the same pressure-density conditions as the single shock using different thermodynamic paths. The compressibility data reported here do not support current density-functional-theory calculations, but are better represented by tabular equation-of-state models.
I. INTRODUCTION
The covalent double-bonds that bind the atoms in a CO2 molecule at ambient conditions are among the strongest of chemical bonds, but at pressures reaching tens of GPa, the compression energy (PΔV) becomes comparable to this bonding energy (hundreds of kJ/mol), and the previously stable molecule exhibits complex chemical behavior.1,2 Laser-heated diamond-anvil cell experiments have characterized the solid phase diagram of CO2 up to 120 GPa, which exhibits five molecular crystalline polymorphic phases before transforming into both crystalline and amorphous polymeric phases.1–7
The fluid phase diagram of CO2 has been experimentally explored to 1 TPa (Refs. 8–14) and is proposed to exhibit similar chemical complexity to the solid phase diagram.15 When shock compressed, molecular liquid CO2 (Fluid-I) is stable up to 40 GPa,8,10 above which it transforms into an insulating 3- and 4-coordinated polymeric fluid (Fluid-II).12,15 Above 100 GPa, CO2 transitions into the Fluid-III phase and begins to ionize.14 The present work is a study of the Fluid-III phase of CO2.
The pressure, density, temperature, and reflectivity of shocked CO2 have been measured to 1 TPa and 93 000 K in Ref. 14. Experimental evidence indicates that CO2 at the highest pressures and temperatures studied is in a complex bonded state as opposed to the previously predicted13 fully atomic C, O fluid. This work reports further details of the study presented in Ref. 14, and additionally reports the experimentally determined Grüneisen parameter and isentropic sound speed of shocked CO2, and the mechanical behavior of CO2 under reshock.
II. EXPERIMENTAL METHOD
These experiments took place on the OMEGA Laser System at the Laboratory for Laser Energetics in Rochester, NY.16 The laser-shocked diamond-sapphire anvil cell17–21 containing the precompressed CO2 sample is depicted in Fig. 1. The CO2 samples were cryogenically loaded into cells comprising 350- or 560-μm-thick diamond and 5000-μm-thick sapphire anvils. The CO2 was contained between the anvils in a stainless-steel gasket. The gasket was initially 250-μm thick; after compression, the gasket thickness was reduced to approximately 150 μm. The anvils were mounted in tungsten carbide (WC) seats, the diamond side seat had a 900-μm lateral window, and the sapphire side seat had an 800-μm window. The diamond was coated with a 15-μm plastic ablator and 3 μm of gold; the gold served to absorb x-rays produced at the ablation front and prevent photoionization of the sample. The sample-side of the drive diamond was coated with a 0.1-μm Ti flash coating. Two 25-μm-thick α-quartz pieces, referred to as the pusher and the window, were located in the sample chamber with the CO2. The quartz pusher served as an impedance matching,22 reflectivity, and temperature standard23–26 for singly shocked CO2. The quartz rear window allowed for the determination of a reshock state in the CO2.
The loaded cell was mechanically compressed to pressures of 0.36–1.16 GPa at ambient temperature. These pressures correspond to densities in the CO2 ranging from 1.35 g/cm3 (Ref. 27) to 1.74 g/cm3 (Refs. 5 and 7), and densities in the quartz ranging from 2.68 g/cm3 and 2.73 g/cm3 (Ref. 25). Above 0.5 GPa, CO2 crystallizes into solid phase-I.6 The formation of crystals in the CO2 was observed in cells with the highest precompression. The precompression pressures were measured using calibrated fluorescence spectroscopy of ruby beads within the cell,28 and the density of the CO2 was then determined from a 295 K isotherm.5–7,27
The OMEGA laser irradiated the plastic ablator with intensities between 1.2 × 1014 W/cm2 and 10.0 × 1014 W/cm2; these experiments used 12 beams with up to 480 J per beam (5760 J total) in an 865-μm focal spot and a 1-ns pulse duration. The laser ablation of the CH layer drove shock waves through the diamond anvil and quartz pusher into the liquid or solid CO2 sample. The pressures in these experiments were sufficiently high to ionize the CO2 and produce an optically reflective shock front. In these experiments, the transit times through the CH ablator, Au layer, and drive diamond ranged from 9 to 22 ns, depending on the laser intensity and drive diamond thickness. The 1 ns pulse shape leads to decaying shocks in the CO2 samples, from which a range of shock states could be observed in a single experiment.
The velocity of the reflecting shock front was measured with a dual-channel line-imaging velocity interferomer system for any reflector (VISAR).29 The Ti flash coating was opaque to VISAR, so the velocity of the shock front was not measured until the shock broke out into the quartz window. This allowed for finer time resolution, as the zero-velocity phase is a necessary reference.29 The apparent velocity from VISAR is corrected for the precompressed refractive index of quartz25 and CO2.6,7 The dual-interferometer system allows one to resolve 2π fringe ambiguities and determine a unique velocity solution. Integrating the velocity as a function of time must yield the thickness of the quartz and CO2 sample chamber; this serves as another check on the velocity solution.
Impedance matching22 was performed at the quartz pusher/CO2 interface using the Rankine–Hugoniot conditions for conservation of mass, momentum, and energy to calculate the particle velocity, pressure, density, and internal energy in the shocked CO2. A Mie–Grüneisen linear release23,25 was used to model the release of the higher-impedance quartz into the lower-impedance CO2. Additionally, the intensity of the VISAR signal is used to determine the reflectivity of the shocked CO2 at 532 nm by referencing to the known quartz reflectivity as a function of shock velocity.25,26
Simultaneously with the VISAR, the self-emission (590–850 nm) from the shock front was measured using streaked optical pyrometry (SOP),30,31 from which a brightness temperature was determined. Brightness temperature is inferred from the measured emission with the assumption that the shock front emits as a gray body with reflectivity as measured with VISAR. While particle velocity, density, and pressure are determined from impedance matching only at the instant the shock wave is transmitted from the quartz pusher into the CO2, shock velocity, temperature, and reflectivity are tracked continuously through the shock transit of the entire experiment.
III. SINGLE-SHOCK COMPRESSIBILITY RESULTS AND DISCUSSION
The pressure and density results from these experiments are plotted in Figs. 2(a) and 2(b) (triangles), along with previous CO2 single-shock data.8–13 We performed a linear fit to the shock velocity () vs particle velocity () data between 189 and 995 GPa (this work and Ref. 13) with a linear term to account for variation in initial density ρ0,
The data all lie within 2-σ of the fit. Other fits, including quadratic, cubic, and exponential, were performed, but statistical analysis showed that the data did not justify a fit more complex than linear. Parameters and covariance matrix elements for Eq. (1) are given by: , and .14
This fit was converted to pressure vs density using the Rankine–Hugoniot conservation relations and plotted in Fig. 2 with solid lines, and extrapolated to lower pressure with dotted lines. Quantum mechanical calculations [density function theory, (DFT),32 dashed–dotted lines], benchmarked in Ref. 12 (squares), predict significant curvature between 50 and 500 GPa due to changes in the molecular bonding of CO2.13,15,32 Our measurements from initially 1.4 g/cm3 (green) do not support such curvature and are in better agreement with LEOS models.33 LEOS table 2274, constructed using the quotidian equation-of-state methodology, expresses the Helmholtz free energy as a function of volume and temperature, and includes a dissociation term and a non-dissociation term in the ion free energy.33 These terms are coupled by the molar mass. It is significant that the LEOS table represents the mechanical behavior of the data in the present work, but does not represent the increase in compressibility seen in Ref. 12.
There exists a clear anomaly in the low-pressure gas-gun data from Ref. 12 (blue squares) and Ref. 8 (green pentagons) seen in Fig. 2. For a given shock pressure, it is expected that CO2 of a higher initial density will have a higher final density due to reduced heating. The low-pressure data demonstrate this behavior until 42 GPa; above this pressure, the data from Ref. 12 (blue squares) show a higher final density for a given shocked pressure than the data from Ref. 8, despite having a lower initial density. The high-pressure data (this work and Ref. 13, 189–995 GPa) demonstrate the expected behavior. The anomaly in the gas-gun data has led to disagreement in the modeling of CO2, and the present measurements support LEOS33 over current DFT32 calculations in the high-pressure fluid regime.
IV. GRÜNEISEN PARAMETER AND SOUND SPEED
Variation in initial density was leveraged to measure multiple Hugoniot curves, from which derivative quantities were probed using a difference method.26 The Grüneisen parameter was determined from the mechanical equation-of-state given in Eq. (1) and a difference method between initially 1.17 and 1.4 g/cm3; 1.17 and 1.7 g/cm3; and 1.4 and 1.7 g/cm3 Hugoniots. The Grüneisen parameter shows little dependence on initial density; all three difference methods yield the same result within 10%. The averaged result is plotted in Fig. 3(a) with a 1-σ confidence interval based on the uncertainty in the – fit. Additionally plotted are predictions of γ from tabular equations-of-state SESAME 5212 and LEOS 2274.33 Theoretical γ is systematically higher than that of the experimental result, but all curves tend to the ideal gas limit of 2/3.
The Eulerian sound speed can be directly calculated from the measured Hugoniot and Grüneisen parameter γ with34
was calculated from Eqs. (1) and (2) for three different initial densities: 1.17 g/cm3, 1.4 g/cm3, and 1.7 g/cm3. The sound speed on the Hugoniot from each initial density differed by less than 3%, showing even less dependence on initial density than the Grüneisen parameter. The averaged result is plotted in Fig. 3(b) with a 1-σ confidence interval propogated from the uncertainty in the parameters in Eq. (1) and the uncertainty in γ with a 100 000 trial Monte Carlo method. SESAME 5212, a single-phase equation-of-state, underpredicts the sound speed of shocked CO2, but LEOS 227433 shows an excellent agreement with our experimental data. This is expected given the good agreement between LEOS 227433 and our Hugoniot data.
V. RESHOCK RESULTS AND DISCUSSION
When the shock wave traverses the CO2 sample and enters the higher-impedance quartz rear window, a second shock (reshock) is launched back into the CO2 sample. Impedance matching22 is performed at the CO2/window interface to determine the pressure, density, and internal energy of this reshock in CO2. The shock velocity of the CO2 (US,CO2) and the quartz window (US,Q) are measured on either of the interface with VISAR. From the known quartz Hugoniot24,25 and Eq. (1) for an initial density of , the pressure, density, and particle velocity on either side of the interface for both CO2 (P ) and quartz () are also known. By impedance matching, the pressure of the reshocked CO2 (PR) must be equal to PQ. By the Rankine–Hugoniot conservation relations, the density of the reshocked CO2 () is given by
The present work measured four reshock states in CO2, which are summarized in Table I. The reshock results from this work and from Ref. 13 are summarized in Figs. 4 and 5. The experimental observables and US,Q are plotted in Fig. 4 along with LEOS curves33 based on intersection of the modeled reshock with the experimental quartz Hugoniot.24,25 This work shows strong agreement with LEOS, as does most of the data from Ref. 13.
Shot . | (km/s) . | (km/s) . | (g/cm3) . | P1 (GPa) . | (g/cm3) . | PR (GPa) . | (g/cm3) . |
---|---|---|---|---|---|---|---|
58917 | 15.74 (0.14) | 16.42 (0.14) | 1.64 (0.01) | 267 (6) | 4.14 (0.05) | 364 (7) | 4.46 (0.11) |
57510 | 16.87 (0.14) | 17.69 (0.14) | 1.36 (0.01) | 280 (6) | 3.97 (0.03) | 426 (8) | 4.56 (0.11) |
58920 | 19.32 (0.14) | 20.41 (0.14) | 1.39 (0.01) | 388 (6) | 4.20 (0.03) | 583 (10) | 4.78 (0.11) |
58922 | 21.94 (0.14) | 23.20 (0.14) | 1.67 (0.01) | 585 (9) | 4.78 (0.05) | 784 (12) | 5.19 (0.11) |
Shot . | (km/s) . | (km/s) . | (g/cm3) . | P1 (GPa) . | (g/cm3) . | PR (GPa) . | (g/cm3) . |
---|---|---|---|---|---|---|---|
58917 | 15.74 (0.14) | 16.42 (0.14) | 1.64 (0.01) | 267 (6) | 4.14 (0.05) | 364 (7) | 4.46 (0.11) |
57510 | 16.87 (0.14) | 17.69 (0.14) | 1.36 (0.01) | 280 (6) | 3.97 (0.03) | 426 (8) | 4.56 (0.11) |
58920 | 19.32 (0.14) | 20.41 (0.14) | 1.39 (0.01) | 388 (6) | 4.20 (0.03) | 583 (10) | 4.78 (0.11) |
58922 | 21.94 (0.14) | 23.20 (0.14) | 1.67 (0.01) | 585 (9) | 4.78 (0.05) | 784 (12) | 5.19 (0.11) |
Figure 5(a) represents the reshock data in the pressure-density plane. The pressure along a reshock curve is related to the pressure on the Hugoniot at the same specific volume V (inverse of density, ρ) by assuming a constant γ equation of state35
where V0 is the initial specific volume (1/ρ0), PR is the pressure on the reshock curve at volume V from initial state P1,V1 on the principle Hugoniot, PH is the pressure on the principle Hugoniot at volume V, and γ is the Grüneisen parameter. Because the single-shock and reshocked state of CO2 were determined with impedance matching, Eq. (4) serves as an independent way to determine γ, plotted in Fig. 3(a) (points). The uncertainty in γ represents how much γ can vary and still yield the measured reshocked state in the CO2 within the error bar. The results are consistent with those obtained from a difference method applied to the CO2 single-shock equation-of-state [Fig. 3(a) solid line], but PR in Eq. (4) does not provide strong constraints on γ given present uncertainties in the measured density of reshocked CO2. Reshock curves for the best value of γ are plotted in Fig. 5 (dotted).
There are three pairs of data points in Fig. 5(a) that reach the same pressure-density state following different thermodynamic paths. As shown in Fig. 5(b), when plotted as internal energy vs density, those points no longer overlap, implying that the states are at different temperatures. Internal energy was determined from the Rankine–Hugoniot condition for conservation of energy. To account for the different initial energy arising from different initial conditions, the initial energy for initially liquid points [blue and green in Fig. 5(b)] was taken from Ref. 36 [–421 J/g for initially 220 K and 1.167 g/cm3 (blue) and –393 J/g for initially 295 K and 1.39 g/cm3 (green)]. The initial energy in the initially solid (orange) points was then found by shifting from the initially liquid (green) points on a 295 K isotherm by integrating the pressure–volume curves from Refs. 5 and 6, and adding the latent heat of fusion from Ref. 37, to yield –552 J/g for initially 295 K and 1.67 g/cm3. In all cases, the initial internal energy of the CO2 was approximately 50% of the error bar of the final shocked internal energy.
Because Eq. (4) relates pressure on the Hugoniot to pressure on the reshock curve at the same volume, the model fails beyond maximum compression of the principle Hugoniot. For this reason, Eq. (4) cannot be used to determine γ from the reshock data in Ref. 13. There are reshock data reported on CO2 in Ref. 10 from aluminum and stainless-steel anvils, but the data have significant scatter and no reported uncertainty, so were not included here.
VI. CONCLUSIONS
To summarize, this work provides additional details on recently published14 equation-of-state measurements of shock compressed CO2 to 1 TPa and 93 000 K from varying initial densities, and presents new information on the Grüneisen parameter, sound speed, and reshock behavior of high-pressure shocked CO2. We find that the compressibility, Grüneisen parameter, and sound speed of shocked CO2 are well represented by LEOS;33 this work does not support the extreme curvature in compressibility modeled by DFT.32 Notably, lower-pressure gas-gun data support DFT over LEOS. We discuss an anomaly in the lower-pressure CO2 data, which has led to disagreement among models. This complexity in the compressibility behavior of shocked CO2 warrants further study, since there is currently a gap between 71 and 189 GPa where no data exist to constrain theory. We report four reshock states of CO2, and discuss the effect of the Grüneisen parameter on the reshock curve. This work provides significant new benchmarks for theoretical calculations of fluids in the warm-dense-mater regime.
ACKNOWLEDGMENTS
This material is based upon the work supported by the Department of Energy National Nuclear Security Administration under Award No. DE-NA0003856, the University of Rochester, the New York State Energy Research and Development Authority, NNSA support to the University of California, Berkeley, and NSF 19–528 Physics Frontier Centers, Award No. 2020249, Center for Matter at Atomic Pressures. This work was performed under the auspices of the U.S. Department of Energy by Lawrence Livermore National Laboratory under Contract No. DE-AC52–07NA27344 and was supported by the LLNL-LDRD Program under Project No. 12-SI-007.
This report was prepared as an account of work sponsored by an agency of the U.S. Government. Neither the U.S. Government nor any agency thereof, nor any of their employees, makes any warranty, express or implied, or assumes any legal liability or responsibility for the accuracy, completeness, or usefulness of any information, apparatus, product, or process disclosed, or represents that its use would not infringe privately owned rights. Reference herein to any specific commercial product, process, or service by trade name, trademark, manufacturer, or otherwise does not necessarily constitute or imply its endorsement, recommendation, or favoring by the U.S. Government or any agency thereof. The views and opinions of authors expressed herein do not necessarily state or reflect those of the U.S. Government or any agency thereof.
DATA AVAILABILITY
The data that support the findings of this study are available from the corresponding author upon reasonable request.