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 WAVE PARADIGMS AND NEW CHALLENGES* James R. Asay Sandia National Laboratories Albuquerque, NM87185 Abstract. Modern shock wave applications require prediction of material response under complex loading conditions and are often solved with numerical simulations. Even though many of the fundamental problems have been addressed, there are still a large number that cannot be solved from a first-principles approach. To attain a fully predictive capability for these phenomena, it is necessary to understand the physical processes occurring during the shock process itself and to critically examine the fundamental postulates used to interpret shock wave phenomena. This paper examines the basic assumptions for describing shock processes in solids. strength of shocked materials, or the kinetics of shock-induced phase transitions. These properties must be evaluated experimentally for use in semiempirical theories. To realize truly predictive capabilities, it will be necessary to develop firstprinciples theories that have robust predictive capability. This will require critical examination of the fundamental postulates used to interpret shockcompression processes. Another question concerns the fundamental material processes causing the shock transition itself. Are properties, such as material viscosity, the controlling factors or are kinematic effects, such as mesoscopic-scale motions produced from local instabilities, responsible for the observed risetimes? These and other questions must be answered by thorough investigations. In particular, a strong need exists to develop in-situ diagnostics for probing these processes in real time. It is clearly time to consider explicitly the highly heterogeneous effects of shock compression and to reformulate the basic conservation equations in terms of the non-homogeneous and non-equilibrium effects produced in the shock process. A change in the fundamental assumptions, or a paradigm change, is necessary to change the way we think about shocked states and which will likely result in significantly better continuum theories. As one INTRODUCTION The field of shock compression science developed from the need to describe the highpressure response of materials in regimes inaccessible by other methods. These requirements led to the development of experimental and theoretical tools for probing the transitory nature of shock compression. The questions posed by early investigations have been answered to a large degree, with the result that a basic understanding now exists for the thermophysical and mechanical properties of shocked materials. This includes development of highly refined continuum models, which accurately predict most of the aggregate features of shock compression. Even after a half-century of intensive research, however, there are still a large number of unresolved shock wave problems. The resulting continuum theories have been incorporated into a variety of numerical techniques and used with computational methods to solve quite complex problems. However, there are still many outstanding problems that cannot be solved with this approach. For example, it is still not possible to predict shock transition times over a range of pressures, the dynamic compressive or tensile *Sandia is a multi-program laboratory operated by Sandia Corporation, a Lockheed Martin Company, for the United States Department of Energy under Contract DE-AC04-94AL85000. 26 example, most treatments of solid response are based on the elastic-plastic model with the simple assumption that the strain tensor can be separated into elastic and plastic components. This and associated models do not incorporate the known effects of mesoscopic motions and are known to be inadequate in explaining many features of shock compression, including loss and recovery of strength during shock-compression. A fundamental understanding of these processes can result in profound improvement of the continuum theories and produce vastly improved predictive capabilities for solving complex dynamic problems. In this paper, the basic assumptions and evolving capabilities that may help resolve these issues will be briefly discussed. These challenges provide exciting opportunities for the next-generation researcher who is willing to do difficult physics and chemistry research or to solve complex engineering problems. BASIC ASSUMPTIONS these scales in more detail, but the main message is that deformation in a shock wave is complex, involves different length scales, and therefore needs to be understood at all levels. The simplest description of a shock involves propagation in a fluid. There have been a large number of papers written on this topic2'3, so only a brief review will be given here. For planar onedimensional motion, it is assumed that the shock wave is a transition from an initial equilibrium to a final equilibrium state of pressure, specific volume, and internal energy. It is further assumed that the final fluid motion is one-dimensional. In this approximation, sometimes referred to as the "benign shock"4, shock compression is assumed to be equivalent to quasi-static compression. Indeed, shock compression data are often used to determine the high-pressure EOS2. The relatively simple ideas of shock propagation in fluids become much more complex in solids, as illustrated in Figure 1. In this case, it is possible to have folly elastic compression at low stresses where the material does not yield. At higher stresses, the deformation is inelastic and exhibits a rich variety of shock features that are dominated in general by rate dependent yielding. At low stresses a two-wave structure is observed that is not steady in time. The first shock is elastic and referred to as the elastic precursor. Its amplitude, referred to as the Hugoniot Elastic Limit (HEL), is generally not constant with time. The second wave representing inelastic response is usually referred to as the plastic wave, although several investigators have shown that the notion of a plastic "wave" is incorrect5'6'7 and that the process is more representative of a diffusion process. Nevertheless, in the moving coordinate system of the first wave, the second wave can assume steady propagation features. For supposedly planar shock compression, the underlying question concerns the dynamic processes that actually occur in the shock transition and the basic assumptions used to analyze the shock. In this regard, I found it useful to take a specific material and examine the predictions of specific models. This provides perspective and insight into the mechanisms that may be responsible for the observations and thus on the shock process itself I use the term "may", because without specific measurements, we can only infer what is happening in the shock transition. The results of this query raise basic questions about the assumptions used to model response at the continuum level. They also help to identify potential diagnostics that may further elucidate deformation mechanisms. Material responses such as dynamic yielding or shock-induced phase changes nucleate at microscopic scales but the effects of initial defect generation can quickly grow to dimensions of a fraction of a micrometer, normally referred to as the mesoscopic scale. There is increasing evidence that this scale has a dominating influence on shock response. For this reason, it is necessary to understand material response over the full range from the finest level, or atomic, to the microscopic, or dislocation, to the sub-granular, or mesoscopic, and finally to the continuum scale. Gupta1 discusses Figure 1. Elastic-plastic wave propagation. 27 There are several basic assumptions usually made in describing elastic-plastic response at the continuum. These can be summarized as (1) the deformation is homogeneous at a level large corresponding to the source of deformation, (2) the total strain can be separation into elastic and plastic components with no plastic dilatation, and (3) the mean longitudinal and lateral stresses are equivalent to hydrostatic pressure. Using these assumptions, the longitudinal stress is equal to the pressure plus four thirds of the resolved shear stress. Oilman6 and others have questioned these assumptions and in particular have shown that the decomposition into elastic and plastic strain is incorrect in that the separation should involve deformation gradients. Graham8 has also questioned the validity of describing the anisotropic stress state as equivalently a perturbation to the pressure state. It is important to keep in mind that even though this construction often represents the overall features of the shock wave propagation, it is misleading and diverts our thinking away from the real deformation processes that occur. There have been many studies, which show that these assumptions give approximately the right continuum response9. However, there are many counter examples to show they are incorrect and can give misleading conclusions about dynamic material response. The paper deals specifically with the following questions: (1) is the assumption of planar motion an adequate description of 1-D shocks?, (2) are homogeneous, equilibrium states produced by steady shocks?, (3) is the traditional emphasis on mechanical deformation states sufficient to describe real shock processes?, (4) is the risetime of a "plastic" wave representative of deformation mechanisms, such as viscoplastic mechanisms, and (5) are existing solid models, e.g., the elastic-plastic model, adequate descriptions of real response? We will show that the state immediately behind the shock wave is not generally homogeneous and uniform; that the variation in material properties at different locations throughout a shocked material has a strong influence on material properties; that material properties such as compressive yield behavior can be time dependent in the shocked state, a direct failure of the elastic-plastic model; and that viscoplastic or other dissipative mechanisms may not be the controlling factor for determining shock risetime. Even the concept of a steady plastic wave in real solids can lead to erroneous interpretations and it may be more appropriate to think of the process as an ensemble of stationary waves. Finally, results will be shown to illustrate that shock-induced local motions are highly non-planar in onedimensional experiments usually assumed to produce pure translational motion. SPECIFIC EXAMPLES The advent of time-resolved measurements of shock wave profiles opened a new regime for probing shock deformations10 at a continuum level. Barker and Hollenbach11 performed a classic set of experiments on aluminum, which showed that the HEL was nearly constant and that the plastic wave was steady to within experimental resolution over the stress range of 9-90 kbar. These data were among the first to test many of the concepts of ratedependent plasticity at the continuum level. A notable observation from these experiments was that the risetime of the plastic wave varied roughly as the fourth power of the peak stress. Grady12 has further shown that this apparent fourth power dependence is more general, applying to a wide variety of metals and non-metals. A variety of rate-dependent material models were developed to describe these and similar experiments. Generally, these models are based on Maxwell construction in which the dissipative term is described by the stress difference from equilibrium9. These continuum models result in good agreement with experimental data. Grady12 has extended this concept to show that different classes of continuum models can be described by a generalized viscoplastic function. The main shortfall is that these models are not predictive and generally do not provide a complete description of loading responses, such as unloading, multiple shock loading, and multi-axial loading. Various dislocation models have also been developed to describe these data, including Johnson and Barker13, who employed a specific dislocation function based on the Orowan equation first used by Taylor14 to describe elastic precursor decay. In fitting the experimental profiles, reasonable values of dislocation parameters were obtained, which suggested that a microscopic description was sufficient to describe the shock process. However, 28 this apparent agreement points out a persistent problem that concerns the uniqueness of continuum wave analyses. Although time-resolved shock wave profiles provide useful information about the macroscopic time scale of deformation, they are not sufficient to uniquely identify actual physical mechanisms. It is also important to note that both continuum and dislocation models of aluminum shock wave profiles used in all previous analyses were based on the assumption of homogeneous response of the shocked material on a scale large compared to the deformation feature. The analyses also implicitly assumed that the deformation was one-dimensional motion on this scale. It is clear that additional information, such as real-time in-situ measurements of the deformation, is needed to discriminate between the different models. Barring this information, another approach is to use additional waves, such as unloading or reloading waves from the shocked state, to probe the properties in the shocked state. Although not a direct probe of the shock transition, this approach tests the assumptions used to describe the shock process. Comparison of these wave profiles with predictions of either the elastic-plastic (E-P) or more sophisticated models of inelastic deformation consequently provides an independent test of the models. Specifically, the simple elastic-plastic model predicts that the initial unloading is elastic, followed by plastic deformation. For reloading from the shocked state, the expected response should be completely plastic since the initial shock compression produces a state on the yield surface. These tests therefore make it possible to infer processes occurring in and behind the "plastic" shock. One set of measured unloading and reloading wave profiles15' l6 are shown in Figure 2. The unloading wave profile shows that the release from the shocked state deviates from the simple elasticplastic model, resulting through quasi elastic-plastic behavior referred to as the "Bauchinger effect". Recompression from the shocked state shows even more deviation, since an elastic precursor to the main second shock is observed, contrary to elasticplastic theory, which implies that further compression should be completely plastic17. These results can be used to determine the shear stress, TO in the shocked state and the maximum shear strength, tc, the material can support. The E-P model predicts these should be equal after initial shock compression. At low stresses (weak shocks)5, it was found that this assumption is nearly valid. For a single plastic shock wave, it was found that the initial shear stress is much less than the maximum shear strength. Detailed analyses implied that this was not simply a rate effect, but that the yield strength in the shocked state was time dependent or that other processes were occurring. "Elastic precursor" Plastic Elastic - plastic model 0.0 1.5 16 17 IM 19 2,0 21 22 23 2.4 25 2.6 Tlme(psec) Figure 2. Unloading and reloading waves in 6061-T6 aluminum. At that time, a plausible explanation for the apparently anomalous behavior of the shear stress was made that multiple reverberations of the shock at internal grain boundaries was occurring, with the result that the shear states were not uniform throughout the shocked state at a sub-grain level. To quantitatively describe this, it was assumed that the shear stress induced behind the shock wave could be described by a Probability Distribution Function, PDF. Assuming a Gaussian PDF, the distribution was adjusted until reasonable agreement could be obtained with both the unloading and the reloading profiles15. Although this approach produced semiquantitative agreement, we thought it to be ad hoc and did not use it for further applications. However, in more recent 2-D and 3-D simulations of wave propagation in granular and polycrystalline materials18'19, this effect is observed in numerical simulations of shock propagation in polycrystalline materials. The earlier results suggested that a larger scale feature, i.e., the grain or sub-grain scale, might 29 be more important than the microscopic scale in modeling these effects. Subsequent to this investigation, Swegle and Grady employed a mixture model and included explicit temperature effects behind the shock to model the experimental results extremely well using another approach.20 Specifically, they assumed that the risetime of steady waves in aluminum are governed by the fourth power law discussed earlier and that the dissipative energy in the shock was deposited locally in small regions within grains. This effect caused a local rise in temperature due to thermal trapping21 in these regions. They assumed at the time that the local hot spots were produced by shear dissipation in micro-shear bands, but a variety of localization effects could cause such effects. These transient hot regions can have a profound effect on macroscopic properties such as yield strength, which can be temporarily decreased, as illustrated in Figure 3. Depending on the actual scale size, which is not known a priori, they found that the local temperature in these hot spots approaches melting, followed by relaxation to the bulk temperature as these regions are rapidly quenched. In correspondence, the shear strength decreases to near zero in the shock wave itself, followed by recovery at times sufficiently long for thermal equilibration. The variability shown by the bands in the figure is due to assumed variations in shear band thickness and separation20. Although the atomic and microscales are undoubtedly important contributing factors to the shock properties, the larger scale effects appear to dominate the macroscale response in aluminum, as shown by the good agreement with the continuum unloading and reloading measurements and the good agreement with the experimental shear stress states. Although the exact details of the mesoscopic scale response are not known, the ability to capture the major results of unloading and reloading from several experiments over an extended pressure range, without appealing to viscoplastic or microscopic mechanisms, is attractive. These results suggest the important point that a scale feature is established at a mesoscopic scale by the shock itself. These results also emphasize the need to identify the actual mechanical and thermal effects induced in real time during shock compression. Variable band -spacing & thickness 5.0 10.0 15.0 20.0 25.0 Shock Stress, GPa Figure 3. Shear stress and maximum shear strength versus shock pressure in aluminum. Data by AFtshuler22 are also shown. This example further emphasizes the point that existing models do not describe the response of aluminum subjected to more complex loading histories (but still simple!). In particular, the shear strength appears to be initially low during shock loading, followed by recovery later in time, which implies an explicit time dependence to the yield stress. The two interpretations discussed; i.e., softening due to local thermal heterogeneities or a distribution of shear stress states due to grain distributions, appear to explain the main features. A common thread is that good agreement is obtained with the data by using material descriptions that incorporate a scale size. It is also clear from these examples that real-time measurements of shock induced features are needed to discriminate between the different approaches. Asay and Barker23 developed an interferometer technique that allows real-time measurements of distributions (dispersion) in particle velocity during shock deformation. This technique provides new information about the nature of heterogeneous motions at a mesoscopic scale of about 100-jiim in diameter. If the particle velocity is not uniform over the laser-illuminated spot size, the reflected Doppler light will contain a distribution of frequencies, causing a loss of contrast, which can be analyzed to determine the average particle velocity and its variance23. Mescheryakov and Dikov24 used this effect to correlate material properties such as compressive, tensile strength, and strain-rate dependence with the degree of particle velocity dispersion in various 30 in the shocked state it is actually fluctuating from point to point on the sample surface. Another important observation is shown in Figure 4c, which illustrates that the spall strength in steel can be directly related to particle velocity dispersion. This example illustrates the importance of grain and sub-grain scale effects in the propagation of shock waves in polycrystalline materials. There have been several studies to evaluate these effects using a numerical approach l9'25. In addition, significant work is in progress at Los Alamos National Laboratory using a discrete particle approach (DM2 code) to model individual grains in 2-D motions19, as shown in Figure 5a. Impacting 0.16 0.12 (a) (A i 0.08 J 0.04 0 0.8 1.6 Time, (b) 0.04 .0.02 rf 0.8 1.6 Time, \ —3E——— Au (a) Wtoeltr, (b) Figure 5. Modeling of wave propagation in copper, (a) experimental configuration, (b) particle velocity distributions at different times behind the shock. (c) 3 I the polycrystalline sample against a rigid wall forms a reverse shock wave, as illustrated. Several notable features are apparent in this simulation. First, a microstructure corresponding to the grain size is observed. Associated with this is a non-planar shock front and highly rotational flows behind the shock front. The simulation also illustrates that the longitudinal particle velocity is not uniquely defined, but has a discrete variation near the shock front at the grain identified as "A" that persists for states well behind the shock front (grains "B" and "C"). Note that the dispersion in velocity at C is nearly of Gaussian form. The main results from these calculations are (1) that shocks in polycrystalline materials are not planar, and (2) that there is a scale effect to the deformation which is related to the grain size. The scale effect is observed to be associated with a rotational motion of mass points within grains that has some features of mesoscale eddies or the beginnings of turbulent motions. These results are in qualitative agreement with experimental results on aluminum by Lipkin Free surface velocity, m/s Figure 4. Particle velocity dispersion in metals, (a) wave profile in al. (b) dispersion velocity in Al. (c) spall strength versus velocity dispersion in Fe. Top curve is pullback velocity; the bottom is dispersion. polycrystalline metals. In their experiments, a laser beam was focused on a polycrystalline sample to determine the average free surface velocity and its variation. The resulting average particle velocity and the dispersion of particle velocity for the elastic, plastic and pullback signals observed in aluminum are shown in Figures 4a and 4b. In particular, a large dispersion of particle velocity is obtained at the onset of yielding and near the top of the plastic wave. The magnitude of dispersion decays during unloading and becomes large again during the spalling process. The implication is that although the average particle velocity appears to be constant 31 and Asay15 and in semi-quantitative agreement with the experimental results of Mescheryakov et al.24. Recent additional calculations by Yano and Hone 26 illustrate the effects of shock pressure on the local gradients near the shock front25'26. Specifically, the degree of "shock front roughness", i.e., the local velocity variations and the rotational motions, appears to decrease as the stress increases. An important implication is that measurement of shock risetime with a continuum gauge can give varying results, depending on gauge dimensions. Furthermore, the results shown in Figure 5 emphasize a critical need to understand the details of shock deformation at the mesoscale in order to develop better continuum models. Undoubtedly, the microscopic features of plastic deformation also contribute to the mesoscopic response, but local rotational motions and velocity distributions must be accounted for in realistic continuum models27' 28. This conclusion agrees with the ad hoc assumptions used by Lipkin and Asay15 to model shear states in shocked aluminum, but within a solid mechanics framework. Refinement of the approach holds considerable promise for changing the fundamental paradigms used in modeling approaches. To understand these processes in more detail, Baer 9 is studying shock propagation in granular materials, such as sugar, that are elucidating features of the deformation not heretofore considered. In particular, he is attempting to identify scalar invariants of the shock-induced flow that can be translated into continuum models. In his recent simulations29, a shock wave was generated in a three-dimensional sample of 65% dense sugar sample containing representative grain distributions. A hydrodynamics code was used to simulate the propagation of the shock motion created by impacting this sample into a rigid plate, as shown in Figure 629. Figure 6 shows the resulting pressure and temperature fields induced by the shock. As observed, there is considerable variation, which must be quantified in order to link mesoscopic scale response to continuum models. Baer has made significant progress in this endeavor by synthesizing large sets of numerical output, typically fewgigabyte files, to determine the spatial and temporal distributions in shock and thermodynamic variables. PfWMA .Pressure Kapton sugar Temperature (a) (b) Figure 6. (a) Numerical configuration used for 3-D modeling a granulated sugar mixture, (b) pressure and temperature variations. The results for particle velocity distributions, or probability distribution functions (PDFs), are illustrated in Figure 7, which shows the distribution in longitudinal, vz, and lateral , v xy , particle velocities for a micron-thick layer at the mid-plane of the sugar sample for times of 500 ns, 800 ns, and 1000 ns after impact. In the first frame, the wave has not yet reached the mid plane. At 800 ns, the wave has passed the mid-plane, inducing a distribution in the longitudinal velocity, but the average velocity has still not achieved the boundary condition value (zero velocity). Even at 1000 ns, the longitudinal velocity still has a considerable spread, which is nearly Gaussian. Similarly, the lateral velocity develops a distribution about zero, which is not predicted for strictly one-dimensional motion of the shock. Presumably, the lateral velocity averaged over the complete volume of the sample for all times would be exactly zero, since the initial loading is 1D. Note also that even late in time, lateral motions on the mid-plane are appreciable. More details of this calculation can be found in Baer's paper29, but it is instructive to comment briefly on the temperature distributions since they are important in thermodynamic calculations. Baer shows that the temperature PDF for this problem has very interesting features, which includes four distinct regions: (1) a precursor region, (2) a bulk region, which represents bulk response, (3) a hot spot region due to local jetting, and (4) a gradient region, which will be important for regions of 32 A typical result30 obtained on sugar is shown in Figure 8, which indicates that spatial irregularities in particle velocity persist for long times behind shock, in agreement with the numerical simulations by Baer. reactivity growth in energetic materials. Although an exact correspondence cannot be made, it is interesting to note that a small fraction of material in his simulations is at high temperature, which is in qualitative agreement with the thermal model of Swegle and Grady20. It is also apparent that these temperature variations will significantly influence most thermodynamic or phase transition properties in non-reactive materials. (a) 800ns 1000ns (b) Figure 8. Experimental technique for Line VISAR measurements, (a) experimental setup, (b) typical velocity result obtained on sugar. The particle velocity records from a Line VISAR can be integrated to provide spatial displacements versus time along a projected line on the target, which for sugar shows a distinct periodicity to the displacement that is finer than the grain scale. Furthermore, these features appear to be nearly independent of time, in agreement with the concept of a scalar invariant of the motion proposed by Baer. In addition, the experimental results are in semiquantitative agreement with predictions of the 3-D granular calculations by Baer29. Baer is continuing these studies in order to investigate the possibility of scalar invariants that can be used in developing generalized continuum models. This approach is similar that used to predict distributed shear stresses in aluminum15. The observations from numerical and experimental measurements on granular materials can be summarized as follows. The common feature for "planar" shock propagation in either polycrystalline metals or granular sugar is that shock wave variables, such as pressure, particle velocity and density, are not distinct values behind the shock, but are distributed in 3-D space and time. Initial studies indicate that these distributions are nearly invariant in time but further work is needed to clarify this. For planar loading of a granular or polycrystalline material, numerical studies to date indicate that large transverse components of velocity are produced, giving rise to localized rotations that V L _ n A ——23*——- 0.0 0.4 -0.2 0.0 0.2 Vz (km/s) VX,Y (km/s) Figure 7. Probability distribution functions of particle velocity for shock propagation in sugar. As mentioned, it is necessary to develop better diagnostics for in-situ. real-time measurements of deformation states. Although the velocity dispersion technique described earlier provides information about the magnitude of the distribution, it contains no information about the spatial nature. A recent extension of the VISAR has recently been developed that provides information about both the magnitude and the spatial distribution of velocity dispersion. The technique has been referred to as a Line Imaging Velocity interferometer and alternatively as a Line ORVIS or a Line VISAR30'31. The technique is applied by projecting a line of laser light onto the target, in contrast to a regular VISAR that uses a bulls-eye pattern of focused light typically 100 f^m in diameter. During target motion, fringes displace along the line in proportion to the local corresponding particle velocity at that point and are recorded by a streak camera30. Measurements have been made with the technique for line lengths ranging from a fraction of one millimeter to about 20 mm. Typically, for a ~200um line, the spatial resolution is about 20 microns. 33 are reminiscent of turbulent effects that may produce localized stress concentrations, including grain fracture and thermal hot spots. The temperature distributions persist for long periods of time after the material has been shock compressed. Similarly, pressure and density variations persist for long times after the shock. Given these variations, it is important to investigate the resulting effects on other property measurements, such as equations of state, that are routinely made to high precision. In addition, these heterogeneous effects are likely responsible for apparent anomalies, such as phase transitions8, optical properties4'32'33, electrical properties4, and energetic materials29. SUMMARY AND FUTURE DIRECTIONS We've discussed the effects of nonhomogeneous deformations of shock compression in polycrystalline materials. As implied by the observed yield behavior of aluminum that may be partially due to a distribution of shear states, the computationally predicted distribution of states in particle velocity, density and temperature, may likely influence all physical phenomena traditionally studied in shock physics. The effects of heterogeneous deformation on mechanical properties have been emphasized in this paper, although there are likely many other dynamic properties that are also influenced. A few of these are summarized in the following paragraphs. In phase transition studies, there has been a longstanding issue about why some transitions proceed extremely rapidly in dynamic experiments but are sluggish in static experiments8. The iron transition is a well-known example. When this was first identified in shock wave experiments at LANL in 195434, Bridgman was skeptical about these results and expressed concern whether traditional deformation mechanisms could result in the fast transition kinetics observed8. Recent MD atomic level and particle mesoscopic-level calculations at Los Alamos National Laboratory are shedding light on the role of defects and non-homogeneous motions in the transformation18'19. The principal conclusion is that the transformation appears to initiate within defect regions induced by the shock and grow from these regions. Another example includes porous materials that are often used to determine off-Hugoniot properties, 34 especially the thermal component of the EOS. However, recent 3-D computer simulations for porous granular sugar by Baer29 illustrate that the state behind the shock front is definitely not in equilibrium over relatively long times. Furthermore, the shock compression states are characterized by distributions in thermodynamic properties that persist for long periods of time. These effects raise serious concerns as to whether meaningful thermodynamic properties can be accurately determined by shock loading of porous materials. A systematic study of the actual distributed states produced by shock compaction and the states expected from thermodynamic properties should be undertaken using 3-D simulations. Numerous other examples can be cited, including anomalous emissions observed in transparent materials above their yield point, electrical conductivities, chemical reactions and so on. In addition to the usual problem of correcting for strength effects, it is useful to keep in mind whether non-homogeneous response will be a problem in obtaining accurate thermodynamic data from shock wave experiments. Continuum measurements continue to be the mainstay of shock physics work, but it is becoming increasingly clear that present models are far from a predictive capability. Combining continuum property measurements with other mesoscopic scale information, such as real-time PDF measurements of the important variables would be powerful for understanding the processes and replacing the existing continuum models. Developing the right experimental and computational tools to accomplish this goal is probably one of the most challenging problems that the shock community faces, but it is absolutely essential that this problem be addressed as soon as possible. Otherwise, we will continue to use the inadequate models developed decades ago. As modeling abilities improve, it will become necessary to include microscale and perhaps atomic scale effects into mesoscale models. Significant progress has already been made in the initial steps of this work27'28. For the most part, shock wave experiments have been limited to planar loading conditions. This type of experiment does not really give much information about the full anisotropic stress tensor, especially if the gauges themselves are essentially onedimensional. If we are to extend the present model 12. D.E. Grady, J. Mech. Solids 10, 2017-2032, 1998. 13. J.N. Johnson and L.M. Barker, J. Appl. Phys. 40, 4321-4334, 1969. 14. J.W. Taylor, J. Appl. Phys. 36, 1965. 15. J. Lipkin and J. R. Asay, J. Appl. Phys., 48, 182 1977. 16. J. R. Asay and L.C. Chhabildas, Shock Waves and High-Strain-Rate Phenomena in Metals: Concepts and Applications, Marc A. Meyers and Lawrence E. Murr, Plenum Publishing Corporation, NY, 1981. 17. G.R. Fowles, J. Appl. Phys., 32, 1475-1487, 1961. 18. B.L Holian and P.S. Lomdahl, Sci. 280, 2085, 1998. 19. K. Yano and Y. Horie, Phys. Rev. B, 59, 1999. 20. J.W. Swegle and D. E. Grady, Metallurgical applications of shock-wave and high-strain-rate phenomena, Eds. Lawrence E. Murr, Karl P. Staudhammer and Marc A. Meyers, NY, 1986. 21. D. E. Grady and J. R. Asay, J. Appl. Phys. 53, 7350, 1982. 22. L.V. APtshuler, M.N. Pavlovskii, V.V. Komissarov, and P.V. Makarov, Combustion, Explosion and Shock Waves, 55, 92-96, 1999. 23. J. R. Asay and L. M. Barker, J. Appl. Phys., 45, 2540-2546,1974. 24. Yu. I. Meshcheryakov and A. K. Dikov, Sov. Phys. Tech. Phys. 30,348-350, 1985. 25. M. A. Meyers and M. S. Carvalho, Materials Science and Engineering, 24, 131-135, 1976. 26. Horie, private communication, 2001 27. P. V. Makarov, Physical Mesomechanics /, 57, 1998. 28. V. E. Panin, Yu. V. Grinyaev, T. F. Elsukova, and A. G. Ivanehin, Izvestiya Vysshikh Uchebnykh Zavedenii, Fizika, 6, 5-27, 1982. 29. M.R. Baer, this conference. 30. W. M. Trott, J. N. Castaneda, J. J. O'Hare, M. R. Baer, L. C. Chhabildas, M. D. Knudson, J. P. Davis, and J. R. Asay, Dispersive velocity measurements in heterogeneous materials. SAND2000-3082, 2000. 31. K. Baumung et al, in Laser and Particle Beams. Cambridge University Press, 14, 181-209, 1986. 32. P.J. Brannon, C. Konrad, R.W. Morris, E.D. Jones and J. R. Asay, J. Appl. Phys. 54, 6374-6381, 1983. 33. David Hare, this conference 34. J.W. Taylor, Shock Waves in Condensed Matter 1983, ed. J. R. Asay et al., 1984. 35. Yu. Mescheryakov, S.A. Atroshenko, V. B. Vasilkov and A. I. Chernyshenko, Shock Compression of Condensed Matter, ed. S.C. Schmidt et al., 1991. for solids to a more realistic description, we must define experiments that exercise the lateral components of stress. These might include planar compression-shear or off-axis impact. I believe there is much to learn by studying combined states of loading, such as shock followed by unloading or reloading, ramp loading, ramp followed by shocking, etc. These additional complexities allow a test of the assumed state produced by the initial process and will be needed to develop fully validated material models for use in the next generation computer codes. Several investigators have recently emphasized that taking the next step in material modeling will require detailed studies at the mesoscopic level19'27'28'29'35. Personally, I believe this is the right approach, that it is the most challenging problem in shock physics today, and that it is also one of the most exciting challenges for the shock physics community. I believe the challenge of developing the appropriate tools and using them to understand the actual shock deformation mechanisms at multiple levels will draw new talents to the field. REFERENCES 1. Y.M Gupta, Shock Compression of Condensed Matter-1999, M.D. Furnish, L.C. Chhabildas, R. S. Hixson (eds), 3-10, North Holland, 2000. 2. L. Davison and R.A. Graham, Physics Reports, 55, 1979. 3. G.W. Swan, G. E. Duvall and C.K. Thornhill,, J. Mech. Phys. Solids, 21, 215-227, 1973. 4. R.A. Graham, Solids Under High-Pressure Shock Compression: Mechanics, Physics, and Chemistry, Springer-Verlag, New York, 1993. 5. D. C. Wallace, Phys. Rev. B, 22, 1495-1502, 1980. 6. J. J. Gilman, Fundamental Issues and Applications of Shock-Wave and High-Strain-Rate Phenomena, K.P. Staudhammer, L.E. Murr, and M.A. Meyers (editors), Elsevier Sciences Ltd, 2001. 7. Y. Hone, private communication, 2001. 8. R.A. Graham,," High Pressure Science and Technology, Part 1, Eds. S.C. Schmidt et al, American Institute of Physics, 1993. 9. W. Herrmann and R J. Lawrence, J. Eng. Mat. Tech., 7,84, 1978. 10. L.C. Chhabildas and G.A. Graham, Amer. Soc. of Mech. Engineers, NY, AMD 18, 1-18, 1987. 11. L.M. Barker, in Behavior of Dense Media Under High Dynamic Pressures. Gordon & Breach, N.Y, 483, 1968. 35
© Copyright 2025 Paperzz