Sie sind auf Seite 1von 32

ELSEVIER

Novel Deformation Processes and Microstructures Involving Ballistic Penetrator Formation and Hypervelocity Impact and Penetration Phenomena
L. E. Murr, E. Ferreyra T., S. Pap u, E. I?. Garcia, J. C. Sanchez, W. Huang, J. M. Rivas, C. Kennedy, A. Aya P and C.-S. Niou a,
Department of Metallurgical and Materials Engineering The University of Texas at El Paso, El Paso, Texas and Materials Research Institute,

Light metallography and transmission electron microscopy techniques affording unique observations of microstructural issues in connection with a related set of novel, high-strainrate deformation processes provide some fundamental insight into the following areas: shock-wave-induced twinning, explosive welding, shaped charge development, explosively-formed penetrator phenomena, hypervelocity impact cratering in metal targets, and long, dense rod penetration/perforation of thick metal targets. Although shock wave phenomena are precursors in all these processes, deformation twins are rarely observed in the residual, process microstructures. In the case of hypervelocity impact craters, no deformation twins are observed in the crater-related target microstructures. Microbands that appear to be related to twins are observed. Melt-related phenomena are observed only in the explosive weld-wave interfaces. Jetting phenomena related to shaped charges and crater rim formation are dominated by dynamic recrystallization, which provides a mechanism for extreme plastic flow in the solid state. Differences observed between rod penetration of rolled homogeneous armor and Ti-alloy thick targets manifest themselves in distinct microstructural differences that also do not include melt phenomena. 0 Elsevier Science Inc., 1996

INTRODUCTION There is a wide range of novel deformation processes, especially involving high-energyrate processes at high strains (true strains from 1 to 6) and high-strain rates (102-lo7 ss*), that are often both practical and unique in the context of more conventional processes, particularly those of metal working, forming, and performance modification. Many processes are driven by high explosives or other, related techniques for creating intense shock waves. Explosive fabrication processes include so-called standoff, contact, and impact operations that affect forming, drawing, hardening, compaction, welding, and cladding of metals [l-5]. In related processes, penetrating devices are
245
MATERIALS CHARACTERIZATION 37:245-276 (1996) 0 Elsevier Science Inc., 1996 655 Avenue of the Americas, New York, NY 10010

shaped, self-forged, or simply accelerated to breach a metal or material target whose deformation is intrinsically coupled to the behavior of the penetrator or to its own characteristic microstructure [l, 3, 5-71. In the past decade or so, the role of microstructure in process efficiency or in understanding process mechanisms or fundamentals began to influence design strategies and applications of such high-rate phenomena. These applications include industrial, military, and space-related structures and systems. In many high-rate deformation processes, shock heating at very high pressures, or the combined high strains and high strain rates, produces localized, adiabatic heating that can promote alternative mechanisms for deformation behavior as well as micro1044-5803/96/$15.00 PI1 SlO445803(96)00173-8

246

L. E.

Murret al.

structural modification. Shock-wave effects are precursors to most impact-related or explosive contact phenomena and can create microstructural modifications that are important in forming, working, welding, and penetration. And, though there have been decades of studies of plane shock wave modification of metals and other materials systems [l-6, 8-121, very little is known about oblique or spherical shock wave effects that characterize more practical deformation processes [3,13]. In this study, we present an overview of several related or interrelated high-strainrate and often extremely high strain deformation processes, all involving shock wave precursors as a consequence either of high explosive drive systems or of high velocity or hypervelocity impact. Of particular interest in the development of this overview is the effect of initial system microstructures on process development, microstructural alterations as a consequence of the process itself, and microstructure evolution or its unique features that characterize the final process or its products. Microstructural issues include plane-wave and oblique shock loading of both face-centered cubic (fee) and body-centered cubic (bee) metals and alloys, explosive welding, shaped charge and explosively formed penetrator development, high velocity and hypervelocity impact and impact crater formation, and thick plate penetration and perforation by explosively shaped devices and long, dense rods. Light and transmission electron microscopy techniques constitute the principal experimental approaches for illustrating and elucidating the associated microstructural features and their development, which are unique to the deformation processes or modes to be described.

SHOCK-WAVE

PHENOMENA

Essentially all explosively-driven processes and impact phenomena involve the creation of shock waves that travel as a geometrical demarcation (or shock front) in the associated materials or materials system. In

crystalline or polycrystalline materials, the advancing shock front creates lattice defects that alter the microstructure. In the simple, plane-wave shock regime illustrated in Fig. l(a), a plane flyer plate impacting a plane (parallel) crystal surface creates a shock wave having a peak shock pressure, Ps = pt (Ct + St UP,)Upm, where pt is the target density [the crystal lattice shown in Fig. l(a) 1, Ct is the bulk sound velocity in the target, St is a material constant, and UP is the modified projectile (or flyer plate) velocity in the compressed region after impact [l, 51. This shock wave actually propagates the same in both the target and the flyer plates. Consequently, the peak pressure is the same in both. As shown in Fig. l(a), the advancing shock front leaves linear dislocation arrays (as well as other defects). For many polycrystalline metals and alloys, the dislocation density increases with increasing peak pressure [1,3]. Consequently, hardness and yield strength are usually increased when a shock wave passes through a crystalline or polycrystalline material. Because a shock wave will move at bulk sound velocity, it is a precursor to plastic deformation. Consequently, the associated plastic deformation will involve the shockwave-altered microstructure, which may or may not be influenced in a significant way by the original or starting microstructure (including, of course, the grain size and dislocation density). In the arrangement shown in Fig. l(a), detonation occurs at the apex of a linewave generator that, because of its angle relative to the main explosive charge, will produce a plane front of detonation products. These in turn will detonate the main charge simultaneously over its surface, accelerating the flyer plate parallel to the plane target. If, as illustrated in Fig. l(b), the detonation wave (VD) travels parallel to the target plate surface, the collision is oblique, and the flyer and target will weld together. At this welding interface, a jet occurs that is often very small, but the associated plastic deformation at the contacting surfaces is often so severe that the flyer plate

Novel Deformation Processes and Microstructures

247

and target flow together to create a bond or weld zone. This zone is often characterized by waves whose period or wave spacing, X, and height are related to the detonation velocity, VD, or the associated pressure. Melting often occurs at the weld interface and within the weld-wave structure, and these features are illustrated in Fig. 2 for the explosive welding of copper. It is possible to virtually eliminate the wavy structures in explosive welding through variations of Vo and l3 (Fig. 2). Correspondingly, the melt fraction is changed. Explosive welding is unique because virtually any metal, as well as other materials, can be joined over huge surface areas, and multiple materials, composites, or laminates can be created [3,4]. The simple schematics represented in Fig. 1 illustrate some of the important features that can occur when a geometrically plane material impacts a plane target at different angles and velocities (or pressures). In Fig. l(a), a plane shock wave propagates into the target and there is no joining or welding of the two. In Fig. l(b), the plates are joined or welded, differently for different impact conditions noted earlier. It should be apparent, then, that these phenomena occur in any impact at high velocity. Moreover, if the impacting plate geometry changes, the process also changes. Finally, a thermal wave trails the shock front, whose magnitude (or dynamic temperature rise) increases with increasing pressure. The principal effect of this phenomenon is that, for very high pressures, the microstructures created in the shock front can be recovered or annihilated by the trailing thermal pulse. These thermal phenomena are further complicated by actual deformation (or residual straining) of the target (or the flyer), which produces adiabatic heating: ATx(e)(E), where AT is the temperature rise and E and i are the corresponding strain and strain-rate, respectively. Of course, for shockloading-related phenomena, the corresponding strain rates are very large (106-10s s-l). These features-high strain, high-strain rate, and associated high temperaturesare, of course, conducive to dynamic recovery and recrystallization of microstructures

created in the shock front. This occurs during plastic deformation and in static recovery and recrystallization from residual heat, either shock heating or adiabaticallydeveloped heating. DEFORMATION TWINNING

Although, as illustrated schematically in Fig. l(a), shock wave propagation in crystalline materials may create dislocations to accommodate the plastic distortion in the shock front, this simple glide phenomenon may not be able to fully accommodate large lattice shear, and localized, massive shear may occur, which is crystallographically

FIG. 1. (a) Plane shock wave generation and propagation in a crystalline solid. (b) Explosive welding of an explosively-driven flyer plate impacting a target or base plate. Explosive welding parameters and parametric relations are shown.

L. E. Murr et al.

FIG. 2. Light microscope views of explosive weld waves in a copper/copper weld zone [Fig. l(b)]. (a) Solidification microstructures in weld-wave vortex and behind the wave crest (arrows) indicative of melt zones. (b) Lower magnification view of weld waves in a copper/copper explosive weld, showing flyer plate and base plate grain structure and extreme plastic flow and grain distortion along the weld zone (interface).

Novel Deformation

Processes and Microstructures

249

manifested in deformation twin formation. In low-stacking-fault free-energy fee metals and alloys, the critical twinning stress is often very small (on the order of lOGPa peak

pressure); whereas, for metals such as copper or nickel, it exceeds 20GPa [3]. At lower peak shock pressures and low-stackingfault free energies, stacking faults and over-

FIG. 3. Plane-shock-wave-induced deformation twins/twin faults in nickel corresponding to a peak pressure of 45GPa and a 2~s pulse duration. The arrows indicate the (111) trace directions along the <022> directions. Note intersection contrast (arrow head lower left), which defines average twin width.

250

L. E. Mwr

etal.

lapping fault bundles forming twin faults are formed, and their density and spacing change with peak pressure [1, 31. In highstacking-fault free-energy metals such as copper and nickel, dislocations form cell structures whose walls densify with pressure as the cell size decreases [ 1,3]. Deformation twins are often unique for shock loading of metals and other materials because they do not form in most deformation processes, and, if they do form, their propensity is not as great as in shock loading. In contrast with fee, where deformation twins or twin faults form coincident with the (1111 slip planes, twins in bee form on { 112) planes or plane segments. And, because of the multiplicity difference between (112) and (111) planes and the glide-related twinning mechanisms, the appearance of twins is different. These features are apparent on comparing plane shock-induced deformation twins in fee nickel shown in Fig. 3 with plane shock-induced deformation twins in bee molybdenum shown in Fig. 4, for the same (100) grain surface orientation. In Fig. 3, the twin faults are narrow volumes that intersect at 90. The twin widths are relatively uniform and are apparent at the intersection that produces a recognizable contrast (arrow). The [ill} twin planes are inclined -55 with the (100) grain surface and produce a rather regular refinement of the nickel grain structure. This produces a residual hardening (AH) and increase in yield strength (a) because

and u0 refers to the unrefined structure (a single crystal), D is the grain size, A is the partitioning or grain refinement parameter noted in Fig. 3, and I< and K are material constants. These same features generally apply in Fig. 4, but the contribution of twinning to grain partitioning is different because the twin density is smaller and crystallographitally (and geometrically) different from that shown in Fig. 3. In Fig. 4, the twin boundary is generally uniform but not on the same (coherent) or parallel (112} plane. The twin boundary projection is, nonetheless, reasonably uniform and corresponds

to an angle of inclination of the (112) plane of ~66 with the (100) grain surface. Figures 5 and 6 show, for comparison, deformation twins in bee tantalum and a tantalum -2.5 wt.% tungsten alloy. In this particular case, the Ta-2.5%W alloy was the flyer plate in a plane-wave arrangement similar to that of Fig. l(a), whereas the Ta plate was the target. The peak shock pressure was -45 GPa, and the pulse duration was essentially the same as for Figs. 3 and 4 (~2 ks). Figure 5(a) shows a light (optical) metallographic view of the equiaxed Ta target after shock loading. The arrow in Fig. 5(a) indicates some nonconcurrent, long deformation twins. A crystallographically similar situation is shown in Fig. 5(b) in the transmission electron microscope (TEM). Figure 6 shows similar views for deformation twins in the Ta-2.5%W alloy. Figure 6(c, d) also shows more specific twin features, which include extra twin reflections [arrow in Fig. 6(c)] and a corresponding darkfield image utilizing these reflections and other twin reflections that coincide with the matrix reflections for the (100) surface orientation. The details of twin spots in diffraction patterns have been described by Bullough and Wayman [14,15] and Murr [16]. It might be of interest to note that the Vickers microhardness for the unshocked Ta target material corresponding to Fig. 5 averaged 109, whereas the shocked-residual hardness was 176, or a AH of 62%. It might also be important to note that shockinduced deformation twinning is grain size dependent; the propensity for twinning increases with increasing grain size (and pressure). Simultaneously imposed strain also can complicate this phenomenon. Furthermore, below some critical grain size, there is often little or no twinning [5, 17, 181. Consequently, the microstructure (grain size) can often be manipulated to avoid twinning, if desired, in certain circumstances. Other refinements to the grain size can often affect a size reduction below the critical grain size for twinning. They include dense precipitates and dispersoids as well as heavy dislocation arrangements associated with these microstructural features [19].

Novel DeformationProcessesand Microstructures Metals such as tantalum and its alloys have been of interest in recent years because their high-strain-rate formability and high density (16.7g/cm3 for Ta) are attractive in military penetration devices such as

251 shaped charges and explosively-formed penetrators (EFPs). These general concepts as well as simple impact cratering and dense rod perforation of a material target are illustrated schematically in Fig. 7. In all

FIG. 4. Plane-shock-wave-induced deformation twins in molybdenum corresponding to a peak pressure of 27GPa and a 2~s pulse duration. The arrow illustrates a principal (112) trace direction along [042].

252

L. E. Murr et al.

FIG. 5. Deformation twins in plane-wave shock-loaded tantalum at 45GPa, 2t.1~pulse duration. (a) Light microI Transim ah of polished and etched shocked sample showing equiaxed grains with crystallographic markings. (b) miS! sion electron micrographic bright-field image showing twins on two different 1112) planes in <042> dir-ections (ma rked A and B) for a (100) grain surface orientation. Two other possible <042> directions are shown by the arrows.

Novel Deformation Processes and Microstructures

253

FIG. 6. Deformation twins in plane-wave shock-loaded Ta-2.5%W alloy at 45GPa, 2~s pulse duration, as in Fig. 5. (a) Light micrograph of polished and etched shocked sample showing crystallographic markings. (b) Transmission electron micrographic bright-field image showing deformation twins in a (100) grain surface orientation similar to that of Fig. 5(b). (c) Selected-area electron diffraction (SAD) pattern for (a), showing [loo] zone. (d) Transmission electron micrographic dark-field image of (a), using twin reflections noted in (c) by arrows. The small spot is an extra twin reflection, whereas the primary twin reflection is coincident with a [loo] zone matrix reflection.

254

L. E. Murr et al.

of the deformation processes illustrated in Fig. 7(a-e), a shock wave is a precursor. In Fig. 7(b-c), the arrows denoting the detonation (or shock) front are variously oblique to the liner material, with the EFP in Fig. 7(c) being closest to a plane shock wave. The impact conditions in Fig. 7(d, e) produce an even more complex, spherical shock wave, which forms the penetrating crater whose depth is related generally to 2, mt, where pP and pt are the penetrator and target densities, respectively, and I, is the penetrator dimension (length or diameter). It is thus apparent why heavy penetrator materials are of interest. Figure 7(a) shows a special shock loading arrangement for creating an oblique shock wave (-55) relative to the axis of metal rods, which can be simultaneously shock loaded and uniaxially strained. This arrangement can provide some indications of microstructures created by shock precursor geometries more realistically related to the deformation processes illustrated in Fig. 7(b-e), and recent results for copper rods having different grain sizes have been reported by Sanchez et al. [20]. Figure 8 shows some examples of deformation twins in copper rods shock loaded in the oblique arrangement shown schematically in Fig. 7(a). Figure B(a, b) attests to the increasing deformation twin density with increasing grain size noted previously, whereas Fig. 8(c) and its selected-area diffraction (SAD) pattern inset illustrate the twin nature in the TEM. The twins in Fig. B(c) are actually tilted between 5 and 10 [the (110) surface is tilted by this amount]. They would normally appear as thin strips because the { 111) twin planes are ideally perpendicular to the (110) surface orientation. These crystallographic features are demonstrated in a more exact (110) grain surface orientation with deformation twins coincident with two nonconcurrent { 111) slip systems along <112> directions (Fig. 9). The SAD pattern inset in Fig. 9(a) and its simulated pattern inset in Fig. 9(b) show prominent twin reflections at <111>/3 positions. The corresponding dark-field image in Fig. 9(b) originates for two unique

<111>/3 twin reflections near the (113) matrix reflection in the simulated diffraction pattern inset. One of the interesting features of Figs. 8 and 9 is that the deformation twinning is very similar to that observed for planewave shock loading, although there are exceptions: the peak pressure is lower than the critical twinning pressure in planewave shock loading (N12GPa in contrast with N20GPa). Sanchez et al. [20] have also noted that there is some intermixing of microbands [21, 221 with the deformation twins, which also seem to increase with increasing grain size and residual strain (at constant grain size). Nonetheless, planewave shock loading appears to provide at least phenomenological details of deformation twinning in a variety of processes involving high-pressure shock waves. There are, however, some intriguing exceptions, which we will describe later.

SHAPED CHARGE DEFORMATION PROCESSES As shown schematically in Fig. 7(b), a shaped charge involves a conical liner with apex angles ranging from 42 to 68. The shock or detonation front (with pressures of 2O-80GPa) initially obliquely shock loads the liner prior to its collapse. The outer 80% of the collapsed liner flows to form the slug while the inner 20% of the liner flows to form an elongating, high-speed jet that elongates under the velocity gradient and necks into periodically particulated fragments. In the jet elongation, the strains can exceed lOOO%, and the associated strain rates are usually in the range lo4 to lo5 s-i. These extreme deformation features contribute to adiabatic heating, with temperatures in the deforming and elongating jet exceeding 0.7T~ in metals such as copper. Under these processing conditions, the deforming jet can extend by dynamic recrystallization that, as Chokshi and Meyers [23] originally proposed, could be phenomenologically described as superplastic stretching as the recrystallized grains are refined to sizes

Novel Deformation Processes and Microstructures

255

~lprn. In reality, this process could be considered a continuous dynamic recrystallization process with recovery and recrystallization actually providing a flow mechanism. Figure 10 shows, for comparison, the grain structure for a starting, undeformed, equiaxed tantalum shaped charge liner

cone [Fig. 10(a)], along with a typical residual microstructure of a soft-recovered jet fragment [Fig. 10(d)] [24,25]. It is apparent on comparing Fig. 10(b) with Fig. 10(f) that there is a grain size reduction from roughly 40Frn to 1u.m in the jet center. However, the jet cross section actually exhibits con-

DETONATOR

FIG. 7. Deformation process schematics. (a) Cylindrical/oblique shock loading of copper rods placed in a cylindrical metal holder. The explosive is C-4. (b) Detonating shaped charge. Dashed lines depict the metal liner cone. Arrows denote the shock front or detonation propagation. Shading represents the high explosive (H.E.). (c) Simple backward folding EFP. Arrows denote detonation (shock) front propagation. The liner (black) is a circular saucershaped plate. (d) High velocity/hypervelocity impact and crater formation; 1.4~ the impact velocity. (e) Sequenis tial views of long rod (length/diameter = lo), high-velocity penetration and perforation of thick (T) plate.

256

L. E. Murr et al.

centric zones of very small grains, with average grain sizes less than 0.4pm (a reduction from the original cone of lo*). This feature is also observed to some extent in the recovered slug, but the slug also exhibits an evolutionary microstructure across its cross section that does not experience repeated deformation in the same context as the elongating jet. Figure 11 shows these microstructural features where only at the

slug center is the microstructure characteristic of that in the jet cross section and dynamically recrystallized [compare Figs. 10(f) and 11(d)]. However, the slug represents an evolutionary and connected microstructure as a consequence of temperature and strain variations from the slug center to its perimeter. At the slug perimeter [Fig. 11(b)], the microstructure is characterized by large dislocation cells with misorientations Cl.

copper FIG. 8. Deformation twins in oblique-shock-loaded graph; (b) large grain size (D = 375km) light micrograph; (c) age with superimposed SAD pattern showing <111>/3 twin rection. The (110) grain surface is tilted to provide a projection

rods: (a) small grain size (D = 29km) light microtransmission electron micrographic bright-field imreflections perpendicular to the [li2] twin trace diof the twins.

Novel Deformation Processes and Microstructures

257

FIG. 9. Deformation twins in oblique-shock-loaded copper rod. (a) Transmission electron micrographic brightfield image showing two sets of twins along the [i12] and [li2] directions indicated by dashed lines in the diffraction pattern simulation inset in (b). The SAD pattern inset in (a) shows <111>/3 twin reflections, which are also noted in the pattern simulation. (b) Dark-field transmission electron micrographic image obtained using the (173) reflection and its associated <111>/3 twin reflections.

258 These cells become elongated and the misorientations increase to -2 at e in Fig. 11(b). From this zone [e in Fig. 11(b)], the dislocation cells begin to recover and recrystallize, and the misorientations average from 9 to 12 at the slug center [Fig. 11(d)]. Dynamic recrystallization therefore plays a minor role in the slug development, whereas it is the dominant feature of jet plasticity. It should also be noted that, unlike his-

L. E. Murr et al.

torical models for shaped charge jetting, there are no melt- or solidification-related microstructures; refer to Fig. 2 for example. Consequently, shaped charge jetting is characterized by extreme solid-state plastic flow, with melt fragments occurring (intermittently) for certain process circumstances along the jet axis [26]. It is also notable that no deformation twins are observed in either the Ta jet fragments or slug (Figs. 10

FIG 10. Shaped charge components and their microstructures: (a) liner cone section view showing isolation of specimen from inner cone wall for light microscope and transmission electron micrographic imaging: (b) light micrograph of tantalum starting liner cone; (c) transmission electron micrographic image of liner cone in (b); (d) softrecovered Ta jet fragment; (e) light micrograph of jet fragment center; (f) transmission electron micrographic brightfield image of jet fragment center in (e). Compare (b) and (c) and (e) and (f). Magnifications are corresponding.

Novel Deformation Processes and Microstructures

FIG. 11. Shaped charge components and their microstructures (continued): (a) tantalum slug (scanning electron micrographic composite view); (b) light micrographic view for strip through slug cross section (arrow) [lowercase c and e show positions of microstructures shown correspondingly in (c) and (e).]; (c) light micrographic view of slug center zone; (d) transmission electron micrographic image of (c); (e) light micrograph of slug zone marked e in (b); (f) transmission electron micrographic image of (e).

260

L. E. Murr et al.

and 11). They are annihilated in the dynamic recovery and recrystallization processes. However, Meyers et al. [17] have argued recently that, because of the grain size effect on twinning, this phenomenon contributes to the improved jet stability as the grain size is reduced. Alternatively, as suggested by Murr et al. [25, 271, the use of very small grain size liner cones or highly deformed starting liner cones also may accelerate the dynamic recrystallization process, thereby promoting enhanced jet stability and longer breakup (particulation) behavior by providing significant stored energy to the plastic deformation of the collapsed liner. Regardless of the specificity of the arguments for shaped charge jet development, the experimental realities can only provide us with a view of the starting liner microstructure and the residual, recovered jet fragment microstructure, between which the dynamic event (or events) must be reconstructed. Figure 12 shows a schematic cartoon of how this process might develop in the shaped charge. It is worth reiterating that the remarkable feature of shaped charge jet development is that the enormous deformation associated with the elongating jet is accomplished primarily by solid-state plastic flow. This is also true of extremely fine wire drawing of, say, lcm diameter copper rod to 30p.m wire with only infrequent annealing, which is a lower strainrate analogue.

EXPLOSIVELY FORMED PENETRATOR PROCESSING

Explosively-formed penetrators are one of numerous self-forged projectiles that differ from shaped charges primarily in the very large angle of the liner cone (>120). The explosively formed penetrator can be forward or backward folding, and there are special cylindrical devices for assuring symmetrical folds, particularly in the tail of the EFP. Figure 7(c) shows a simple modification of the shaped charge detonation regime in Fig. 7(b). However, in contrast with the shaped charge, the EFP is dominated by

what is equivalently a self-formed slug. This slug does not form by the uniform, extreme plastic strain characteristic of the shaped charge, and the strains within the EFP usually do not exceed a true strain of 4. The strain rate is somewhat lower than the shaped charge jet (~10~ s-l). There is also a velocity gradient along the forming EFP, which produces tensile stresses along its axis in flight. Although it might be expected that there may be some microstructural similarities to the shaped charge, there is only a meager volume fraction of dynamic recrystallization. Dynamic recovery and evolutionary dislocation cell structures dominate in tantalum EFPs, for example. There are isolated examples of recrystallized grains intermixed with recovery microstructures. The extremes in microstructure, illustrated in the light metallographic views in Fig. 13, are not observed in the shaped charge, although there are some similarities between the Ta shaped charge slug microstructures-Fig. 11(e) for example-in contrast with Fig. 13(b) [28-321. Figure 13(d) illustrates a Ta-2.5%W EFP that revealed a few isolated examples of residual deformation twins in the tail section [to the left in Fig. 13(d)]. Figure 14 reproduces a sequence of transmission electron micrographic bright- and dark-field images illustrating these features. These twins should be compared with those formed by planewave shock loading illustrated in Fig. 6. The general absence of deformation twins in tantalum and their apparent alteration in Ta-2.5%W EFPs (Fig. 14) may be an indication of their thermal instability, because the EFP temperature overall does not exceed 0.4TM (where TM, the melting point, is -3000C for Ta) and could reach this temperature only in isolated, high-strain regions.

HIGH-VELOCITY AND HYPERVELOCITY IMPACT CRATER PHENOMENA

Just as the shock wave collapses the liner cone in the shaped charge or drives the self-forging EFP, a spherical shock wave as-

Novel Deformation Processes and Microstructures

261

6 1

FIG.12. Cartoon and associated transmission electron micrographic bright-field images depicting the dynamic recrystallization features implicit for the shaped charge jetting process. The schematic part is from Ref. [25]. The cartoon shows continual (repeated) deformation (shading) and recrystallization with the straining jet. At 2, the original cone is shocked. At 3, the jet begins to recrystallize, At 7, the jet necks and, at 8, it particulates. The transmission electron micrographic images are of tantalum.

262

L. E. Murr et al.

EFP components and microstructures. (a) Soft-recovered tantalum EFP cut in half and polished to reveal FIG.,13. mic restructures observed in the light microscope in (b) and (c). (d) Soft-recovered Ta-2.5%W symmetrical (threefold ) EFP.

Novel Deformation Processes and Microstructures

263

FIG. 14. T ransmission electron micrographic bright-field (a) and dark-field (b) images showing defo rmation twir IS in the3tail section (arrow) in Fig. 13(d). The dark-field image was made by using apertured reflec :tion (arrow) in tl-re SAD pattern inset; a and b show (ilO) and (002) reflections, respectively, for [IlO] zone. 7 win , double diffr action, and extra (kinematical) twin reflections occur between the two matrix reflections.

264

L. E. Murr et al.

sociated with a high-velocity or hypervelocity (impact velocity, u0 > 5km/s) particle impacting a target drives the crater that forms. Because, as noted earlier, the shock pressure is the same in both the impacting particle and the target at the instant of contact, the reflection of shock waves in a finite (and ideally spherical) impactor can create extremely high temperatures as well as complete spallation. Melting, fragmentation, and even vaporization can be competing mechanisms affecting the impacting particle. A semi-infinite target, on the other hand, will crater with the symmetry of the impact-generated shock wave, which is ideally hemispherical in the target, with the shock wave attenuation contributing to the actual crater size. As the actual crater forms, the target material flows plastically up along the wall to form a jet, which creates a rim above the target surface, a process that is phenomenologically similar to explosive weld jetting (Fig. 2) and shaped charge jetting [Fig. 7(b)]. These features are represented very generally in Fig. 15. Figure 15(b) shows a computer simulation for a steel impactor striking an aluminum target at 25km/s. The impactor is vaporized in this model, and the rim jetting is apparent. Figure 15(c, d) shows a normal view into a crater in copper and a corresponding sectional view through the same crater, respectively, made by a soda-lime glass sphere (3.2mm diameter) impacting at 5.8km/s. Figure 15(d) shows the impact axis through the crater half-section, with sequential notations a, b, c, and d below the crater wall. Figure 16 shows a corresponding sequence of transmission electron micrographic brightfield images taken along the impact axis, as noted by extracting and polishing 3mm discs sequentially from the crater wall bottom. Figure 16(e) corresponds to a zone roughly 14mm below the crater bottom in Fig. 15(d), whereas Fig. 16(f) is typical of the unimpacted copper plate, which had a large grain size of 763pm. Unlike the computer-simulated crater in Fig. 15(b), the transmission electron micrographic sequence in Fig. 16 illustrates that a residual microstructure below a hypervelocity cra-

ter in a copper target extends far beyond the crater wall region and is a consequence of the spherical shock wave. Near the crater wall (0.2-0.4mm), as shown in Fig. 16(a), there is a zone of dynamically recrystallized and considerably refined grains (0.7pm in contrast with ~700~m for the original target plate). This is an even larger grain refinement than that observed in the tantalum shaped charge-compare Fig. 10 for example. Beyond the recrystallized zone, there is an extensive zone of severe plastic deformation that contains very fine dislocation cells and a high dislocation density, as shown typically in Fig. 16(b). This zone becomes intermixed with microbands and a wide region where profuse microbanding is observed. Beyond this microband zone [from about 3 or 4mm in Fig. 15(d)], the microstructure is characterized by evolutionary dislocation cell structures that increase in cell size and decrease in dislocation density. There is a corresponding variation in hardness (Vickers) with considerable softening in the dynamically recrystallized zone near the crater wall, rising to a peak in the microband zone and then decreasing somewhat regularly as the dislocation cell size increases [Fig. 16(d, e)]. These features were originally observed by Quinones et al. [33] and Rivas et al. [34,35] for copper and aluminum craters. Interestingly enough, similar phenomena were observed even earlier in the examination of laser-induced cratering in single-crystal and polycrystal iron targets [36]. There are several interesting features included in the microstructural evolution below the copper crater in Fig. 16. The first is the zone of dynamically recrystallized and ultrarefined grains at the crater wall. The second is the observation of a zone of microbands somewhat removed from the crater wall, and the absence of deformation twins. The third is the absence of any meltrelated phenomena at or near the crater wall. The observation of a zone of microbands [Fig. 16(c)] rather than shock-induced deformation twins or twin faults associated with hypervelocity impact craters in copper is especially intriguing because it has

Novel Deformation Processes and Microstructures

265

also been demonstrated that this zone is velocity (or pressure) dependent and increases with impact velocity, u,,. Furthermore, as shown in Fig. 17 and 18, crater-related microbands tend to increase in density with increasing grain size, are oriented coincident with a primary slip system ((ill}), and have misorientations of -2-3 and widths ranging from about 0.2 to 0.4Fm. Huang and Gray [22] believe that the microband boundary is a dislocation double wall. It is apparent from Fig. 18 that microbands are somehow related to deformation twins in terms of primary (111) slip, but their mechanism of formation is obviously different. It is interesting to note that the dislocation

cells in Fig. 11(f) look very similar to the microbands in Figs 16(c) and 17(f). Even the width and misorientation of the cells are the same. However, the cells are not coincident with a slip system (which, of course, would be {llO) for the Ta in contrast with (111) for copper). Microbands have also been identified in other cratered fee targets which, as a consequence of their propensity to twin, might normally be expected to contain deformation twins, as in the case of copper. Figure 19 shows a hypervelocity impact crater in a 6061-T6 aluminum target. Note the absence of any significant rim (owing to tensile fracture in the rim jet) and the crater surface

FIG. 15. Hypervelocity impact crater phenomena. (a) Hypervelocity impact crater schematic [Fig. 7(d)]. (b) Computer-simulated impact crater half-section for a steel projectile impacting an aluminum target at 25km/s. The projectile is assumed to be vaporized. (c) Normal view of impact crater in a large-grain (735pm) copper target for 3.2mm diameter soda-lime glass sphere (u, = 5.8km/s). (d) Sectioned crater in (c). The notations along the impact axis [dashed line in (d)] correspond to the specific zone of microstructure illustrated in Fig. 16.

266

L. E. Murr et al.

FIG. 16. Sequential and evolutionary residual transmission electron micrographic images of microstructures along the impact axis [in Fig. 15(d)] below the crater wall bottom in copper. (a) Dynamic recrystallization zone near the crater wall. (b) Heavy dislocation zone beyond (a). (c) Microband zone beyond heavy grain distortion. Microbands are oriented along the trace of (111) for a (110) oriented grain. (d) Zone of dislocation cells beyond the microband zone, which continues beyond about 12mm from the crater bottom in Fig. 15(d). Note the dislocation cell changes from (d) to (e). (f) Undeformed copper starting-target microstructure for reference. The centerline is a coherent annealing twin boundary with dislocations at the boundary plane.

Novel Deformation

Processes and Microstructures

267

FIG. 17. Comparison of microbanding associated with hypervelocity impact cratering in copper targets of different grain size: (a) original target microstructure, D = 38um; (b) microband zone about 2mm from the crater bottom formed by a 3.2mm aluminum sphere impacting the plate in (a) at 6.lkm/s; (c) transmission electron micrographic view of microbands in (b); (d) original target microstructure, D = 763um; (e) microband zone about 3mm from the crater bottom in Fig. 15(d); (f) transmission electron micrographic image of microbands in (e). Microbands in (c) and (f) are along traces of 1111). Magnification of(f) is the same as that of (c).

268

L. E. Murr et al.

differences in Fig. 19(a, b) in contrast with Fig. 15(c, d). Note also the microband similarities on comparing Fig. 19(c, d) with Figs. 16(c) and 17(c, f), as well as the general ap-

pearance of the distortions of grain flow lines below the sectional view in Fig. 19(b) in contrast with the computer simulation in, Fig. 15(b).

FIG. 18. Comparison of (a) microband features with (b) shock deformation-induced microtwin features. The microbands in (a) are coincident with the [1?2] trace for (ill) planes (arrow), which are perpendicular to the [ill] direction shown within the SAD pattern inset. The grain surface orientation is (110). The microbands exhibit a 2 overall misorientation shown by spot splitting (arrow) in the SAD pattern inset in (a). (b) T win-faults coincident with (ill) planes along [li2] (arrow) are exactly coincident with microbands in (a). The SAD pattern inset shows prominent <111>/3 twin reflections (arrow). The [ill] direction is perpendicular to the [1?2] arrow. The grain surface is (110).

Novel DeformationProcessesand Microstructures Finally, Fig. 20 summarizes hypervelocity impact cratering in metal targets, especially in the context of the important microstructural issues associated with the cratering process. The light micrograph in Fig. 20 shows a cross section of a rim area in the crater of Fig. 15(c, d). The upper-rim surface forms by a kind of shear band flow, and these overlapping flow zones consist of the dynamically recrystallized (and refined) grain structure. It appears that, just like shaped charge jetting, the rim jetting represented in Fig. 20 occurs by a velocity gradient, which imposes a tensile stress in the flow direction. The rim edges can therefore neck down and eject or particulate a circumferential rim zone. There is no evidence for melting, and crater flow (or, more specifically, target material flow) into the jetting rim is a solid-state plastic flow phenomenon identical with shaped charge jet flow and elongation. Or, in the case of cratering in 6061-T6 aluminum shown in Fig. 19(a, b), the rim can fracture extensively under this resolved tensile stress and eject the bulk of the rim material. This process of rim fracture accounts for essentially all of the crater-related (target) mass loss. Note the microbands and original grain structure in the underside of the rim section shown in Fig. 20. This relatively undeformed and lifted target surface area is also illustrated in the computer-simulated crater in Fig. 15(b) and accounts for the balance of displaced target mass in forming the crater. The schematic representation shown in Fig. 20 is an effort to depict very generally the hypervelocity cratering process in thick metal targets (where there is no spa11 or perforation). The microstructural zones depict those prominent features shown in Fig. 16 and the rim section view shown in Fig. 20. An attempt is made to depict the fate of the impacting projectile, which, in the case of aluminum impacting a copper target, produces a thin, melt/solidification film on the crater wall. However, there is no remnant glass for a soda-lime glass projectile. There have been no examples of projectile material interacting with (alloying with) the target. There are examples of the projectile spalling into many fragments, some of which are embedded in the crater wall. We have observed this phenomenon for 3.2mm diameter stainless steel projectiles impacting the copper targets shown in Fig. 17(a, d) for example (u, > 5km/s).

LONG ROD PENETRATION OF THICK TARGETS The novel features of long rod penetration, as illustrated in Fig. 7(e), in contrast with simple cratering shown in Fig. 7(d), include the extension of the penetrator and the complete perforation of a finite-thickness target. Grace [37] has recently examined the interaction of a dense, penetrating rod with the target in the context of penetrator erosion and has applied this approach in explaining differences in penetration/perforation between conventional rolled, homogeneous armor (RHA) plate and a Ti6AI-4V target plate [38]. In addition, Huang et al. [39] have shown that there is a significant microstructural difference associated with the residual penetration channel region. These differences have also been augmented by TEM observations of channel-related microstructures, and Figs. 21 and 22 illustrate some of these features. In Figs. 21(a) and 22(a) are shown comparative cross sections resulting for a W5%Ni, 2%Fe alloy rod (7.8mm in diameter and 78mm in length; with a density of 17.6g/cm3), impacting at 1.5km/s. These two targets, RHA [Fig. 21(a)] and Ti-alloy [Fig. 22(a)], were w45mm and 70mm in thickness, respectively. There is a readily apparent difference in the cratering process in Fig. 21(a) in contrast with Fig. 22(a). The target surface is lifted three times as much in forming the RHA crater and perforation in comparison with the Ti-alloy. In addition, the light micrograph of the RHA microstructure in Fig. 21(b) is considerably different and deformed in comparison with the corresponding images in Fig. 22(b, c) for the Tialloy, where there is no apparent grain or phase distortion even 5 to 10p.m from the penetration channel wall [compare Fig.

270

L. E. Murr et al.

21(b, d) and Fig. 22(b, c)]. The transmission electron micrograph in Fig. 21(c) shows the complex and heavily deformed RHA (carbon steel) starting plate microstructure; because of the heavy deformation in the channel wall region creating highly magnetic thin foils, no corresponding transmission electron micrographs have been obtained for this region. However, the transmission electron micrographs of the Ti-alloy channel wall region shown in Fig. 22(e) exhibit a significantly increased dislocation density in the Ti-alloy duplex (CI + p) microstructure, possibly shock-wave-induced [compare with Fig. 22(d, f)]. Because the microstructures associated with different penetration channels in Figs. 21 and 22 are so different, it seems likely that the actual mechanism of penetration might be different. Such differences are

somewhat apparent in hypervelocity cratering on comparing, in retrospect, Fig. 15(c, d) for copper with Fig. 19(a, b) for a 6061-T6 aluminum target.

SUMMARY AND CONCLUSIONS We began this study of novel, high-strainrate deformation processes by focusing on their common feature of shock wave or detonation propagation, either as a means to drive the process or as a consequence of the process. Because, ultimately, this shock wave phenomenon affects a metal or metal system, a brief survey of shock wave-induced deformation twinning unique to many shock processing regimes was presented, which included twinning in both fee and bee metals and which also included plane-

FIG. 19. Microbands associated with a hypervelocity impact crater in a 6061-T6 aluminum target for a 6.4mm diameter soda-lime glass sphere impacting at 5.2km/s. (a, b) An example of a crater at 4.6km/s. (c, d) Microbands below the 5.2km/s crater wall. (c) A light micrographic image. (d) A transmission electron micrographic brightfield image. The orientation in (d) is (110), and microbands are coincident with the trace of [li2].

Novel DeformationProcesses and Microstrtlctures wave and oblique shock wave arrangements. Deformation twinning increases with grain size in both fee and bee metals and can refine or partition polycrystalline metals, resulting in significant increases in both hardness and yield stress. It can significantly increase the initial stored energy and can thereby affect microstructure development, recovery, and recrystallization phenomena. In oblique shock loading of copper rods, Sanchez et al. [20] have also noticed that there is an intermixing with microbands, particularly at large grain sizes and when there is a simultaneous (and residual) strain. Gray [40] has also shown that twinning in plane-wave shock-loaded copper can change

271 predominantly to microbands when there is a large, associated strain. The explosively-welded system interface in copper, for example, was shown to consist of prominent melt/solidification zones associated with the vortex and crest of weld waves. There is also a narrow region of extreme deformation characterized by recrystallization phenomena, probably a mixing of dynamic and static recrystallization. Although shaped charges can be effectively examined only as recovered fragments (jet fragments and the slug), the comparison of starting liner cone microstructures with the residual component microstructures has suggested that dynamic recrystallization plays a prominent role in

_r_,. - ._,j

crywr &+le

metbsjectc

FIG. 20. Hypervelocity impact crater rim flow and a generalized cartoon showing the dynamically recrystallized zone that this represents, as well as other crater-related microstructures. The peak shock pressure at the point of impact is denoted P, and the shading represents its amplitude attenuation into the target. The projectile can melt, spa11 into fragments, vaporize, or undergo combinations of these phenomena. Note, in the light micrograph of the rim section, that the outer rim (surface) flow zone is characterized by extremely small grain structure (by dynamic recrystallization). This is in contrast with the large grain structure below the rim, which contains some groups or bundles of microbands.

L. E. Murr et al.

FIG. 21. W-alloy rod penetration of an RHA target at 1.5km/s: (a) target sectional views showing penetration channel; (b) light micrographic view of channel-related microstructure in (a); (c) transmission electron micrographic bright-field view of the complex RHA microstructure containing lamellar and other distributed carbides and dislocations; (d) light micrograph corresponding to a zone in the target far removed from the penetration channel: (c) represents a region in this zone. Compare (b) and (d) where the magnification is the same as that of (b).

Novel Deformation Processes and Microstructures

273

FIG. 22. W-alloy rod penetration of a Ti-6AI-4V target at 1.5km/s: (a) target sectional view showing penetrai tion channel; (b) light micrographic image of channel-related microstructure in (a); (c) typical target microstructure I far removed from the penetration channel; (d) transmission electron micrographic bright-field image from regionn in (c); (e) transmission electron micrographic bright-field image essentially coincident with the channel wall in (b) through a tangential section; (f) transmission electron micrographic bright-field image near the bottom of the micrograph in (b), roughly lmm from the channel wall.

274

L. E. Murr et al.

jet development while the slugs are dominated by dynamic recovery and exhibit dynamic recrystallization only along the slug center (axis). There are no significant melt phenomena, and the shaped charge process occurs by extreme plastic flow in the solid state. Dynamic recrystallization can provide a unique mechanism for this plastic flow. Explosively-formed penetrators or selfforming or forging projectiles differ from shaped charge phenomena because of a wider distribution of smaller strains and slightly lower strain rates. The microstructures associated with EFPs are widely distributed and contain significantly less evidence for dynamic recrystallization than even the shaped charge slug. Residual microstructures in the EFP are dominated by dynamic recovery microstructures. These include prominently elongated dislocation cells, but there are a significant evolution of dislocation cells to recovery microstructures and some intermixing of dynamically recrystallized grains. There have been some limited observations of deformation twins in EFPs, but they appear to be thermally altered and may be unstable at relatively low processing temperatures. Hypervelocity impact craters in metal targets such as copper provide a unique opportunity to link high-strain-rate and highstrain deformation microstructures with the undeformed target. There is a narrow zone near copper crater walls that looks exactly like shaped charge jet fragments and is characterized by dynamic recrystallization. In addition, there are also no significant melt phenomena in copper crater microstructures, and the flow of material out of the target and into the rim is identical with the extreme solid-state plastic flow that characterizes shaped charge jet development. Dynamic recrystallization also provides a mechanism for this solid-state flow. Below this recrystallized zone, there is a zone of heavy deformation and grain distortion, with high dislocation densities and small dislocation cell sizes. This zone becomes an extensive zone of microbands. No deformation twins have been observed, and dislocation cells with increasing size and decreasing dis-

location density characterize the microstructure beyond the microbands. Microbands are coincident with traces of (111) planes in fee metal targets and are apparently related to deformation twins. Microbands are also apparently related to grain size and are impact-velocity (or pressure) dependent. Residual microstructural effects in hypervelocity impact craters can extend well beyond the crater wall, to distances of roughly one crater diameter below the crater bottom, for example. Finally, we observed significant microstructural differences between long-rod penetration of an RHA target and that of a Ti6AI-4V target. There was a zone of extreme plastic deformation surrounding the RHA, whereas only dense dislocation arrays very close to the penetration channel wall were observed in the Ti-alloy target. There was no evidence for melt phenomena in either target; nor was there evidence for recrystallization near the wall in the Ti-alloy target. This research was supported in part by Army Contract DAAA21-94-C-0059 through the Army Research, Development, and Engineering Center, Picatinny Arsenal, New Jersey, and NASAJohnson Space Center Grant NAG-9-481. A Patricia Roberts Harris Fellowship supported J. C. Sanchez, and other students taking part in this research have been supported by a Mr. and Mrs. Macintosh Murchison Endowed Chair. We thank Mike Hespos of ARDEC, Picatinny, for providing EFPs, Dr. Lou Zemow, for providing shaped charge components, and Dr. Fred H&z of NASAJSC, for providing archived crater samples. Dr. Fred Grace of the Army Research Laboratory, Aberdeen, Maryland, provided the thick, rodpenetrated targets examined in this research.

References
1. Shock Waves and High-Strain-Rate Phenomena in Metals. M. A. Meyers and L. E. Murr, eds., Plenum, New York (1981). 2. Metallurgical Applications of Shock-Wave and HighStrain-Rate Phenomena: L. E. Murr, K. P. Staudhammer, and M. A. Meyers, eds., Marcel Dekker, New York (1986).

Novel Deformation Processes and Microstructures


3. Shock Waves for Industrial Applications. L. E. Murr, ed., Noyes Publications, Park Ridge, NJ (1989).

275 ena, L. E. Murr, K. I. Staudhammer, and M. A. Meyers, eds., Elsevier Science, B.V., Amsterdam, p. 801-810 (1995). 21. I?. J. Jackson: The role of cross-slip in the plastic deformation of crystals. Mater. Sci. Eng. 5739-47 (1983). 22. J. C. Huang and G. T. Gray III: Microband formation in shock-loaded and quasi-statically deformed metals. Acta Metall. 37(2):3335-3347 (1989). 23. A. Chokshi and M.A. Meyers: The prospects of superplasticity at high strain rates. Ser. Metall. 24: 605-611 (1990). 24. A. C. Gurevitch, L. E. Murr, H. K. Shih, C.-S. Niou, A. H. Advani, D. Manuel, and L. Zemow: Characterization and comparison of microstructures in the shaped charge regime. Muter. Char. 30~201-216 (1993). 25. L. E. Murr, H. K. Shih: and C-S. Niou: Dynamic recrystallization in detonating tantalum shaped charges. Mater. Char. 33:65-74 (1994). 26. D. H. Lassila, W. I. Walters, D. J. Nikkel, Jr., and R. I?. Kershaw: Evidence of melt in soft recovered copper jets. In Metallurgical and Materials Applications of Shock-Wave and High-Strain-Rate Phenomena, L. E. Murr, K. I. Staudhammer, and M. A. Meyers, eds., Elsevier Science, B.V., Amsterdam, p. 503-509 (1995). 27. L. E. Murr, C.-S. Niou, J. C. Sanchez, H. K. Shih, L. DuPlessis, S. Pappu, and L. Zernow: Comparison of beginning and ending microstructures in metal shaped changes as a means to explore mechanisms for plastic deformation at high rates. I. Muter. Sci. 30:2747-2758 (1995). 28. L. E. Murr, C.-S. Niou, S. Pappu, J. M. Rivas, and S. A. Quinones: LEDS in ultra-high strain-rate deformations. Phys. Stat. Sol. (a) 149:253-274 (1995). 29. S. Pappu, C.-S. Niou, C. Kennedy, L. E. Murr, L. DuPlessis, and M. A. Meyers: High-strain-rate behavior of pure tantalum in explosively formed penetrator and shaped charge regimes. In Metallurgical and Materials Applications of Shock-Wave and High-Strain-Rate Phenomena, L. E. Murr, K. I. Staudhammer, and M. A. Meyers, eds., Elsevier Science, B.V., Amsterdam, p. 495502 (1995). 30. C.-S. Niou, L. E. Murr, C. Feng, and S. Pappu: Deformation microstructures in tantalum explosively formed penetrators. In Microstructural Science, Vol. 22, M. T. Shehata, T. R. Leduc, I. LeMay, and M. R. Louthan, Jr., eds. ASM International, Materials Park, OH, pp. 73-85 (1994). 31. L. E. Murr, C.-S. Niou, and C. Feng: Residual microstructures in explosively formed tantalum penetrators. Ser. MetaR. Mater. 31(3):297-302 (1994). 32. L. E. Murr, S. Pappu, C. Kennedy, C.-S. Niou, and M. A. Meyers: Tantalum microstructures for highstrain-rate deformation: shock loading, shaped charges, and explosively formed penetrators. In

4. Explosive Welding, Forming, and Compaction. T. Z.


Blazynki, ed., Applied York (1983). Science Publishers, New

5. M. A. Meyers: Dynamic Wiley, New York (1994).

Behavior

of Materials.

6. Metallurgical and Materials Applications of ShockWave and High-Strain Rate Phenomena. L. E. Murr, K. I. Staudhammer, and M. A. Meyers, eds., Elsevier Science, B.V., Amsterdam (1995). 7. W. P. Walters and J. A. Zukas: Fundamentals Shaped Charges. Wiley, New York (1989). of

8. L. E. Murr: Metallurgical effects of shock and high-strain-rate loading. In Materials at High Strain Rates, T. Z. Blazynski, ed., Elsevier Science, New York, p. 1-46 (1987). 9. L. Zernow: High-Velocity Forming of Metals. Prentice-Hall, Englewood Cliffs, NJ (1964). 10. A. A. Ezra: Principles and Practice of Explosive Metal Working. Garden City Press, London (1973). 11. J. Rinehart and J. Pearson: Explosive Metats. MacMillan, New York (1963). Working of

12. Shock Wave and High-Strain-Rate Phenomena in Materials. M. A. Meyers, L. E. Murr, and K. I. Staudhammer, eds., Marcel Dekker, New York (1991). 13. K. I. Staudhammer and L. E. Murr: Dynamic consolidation of cylinders by oblique shock loading. I. Mater. Sci. 25:2287-2298 (1990). 14. R. Bullough and C. M. Wayman: Twinning and some associated diffraction effects in cubic and hexagonal metals: I. Trans. AIME 236:1704-1710 (1966). 15. C. M. Wayman and R. Bullough: Twinning and some associated diffraction effects in cubic and hexagonal metals: II. Trans. AZME 236:1711-1715 (1966). 16. L. E. Murr: Electron and Ion Microscopy and Microanalysis: Principles and Applications. 2nd ed., Marcel Dekker, New York (1991). 17. M. A. Meyers, U. R. Andrade, and A. H. Chokshi: The effect of grain size on the high-strain, highstrain-rate of copper. Metall. Mater. Trans. (A) 26: 2881-2893 (1995). 18. L.E. Murr, M. A. Meyers, C.-S. Niou, Y. J. Chen, S. Pappu, and C. Kennedy: Shock-induced twinning in tantalum. Acta Mater. 45(1):157-175 (1997). 19. L. E. Murr: Residual microstructure-mechanical property relationships in shock-loaded metals and alloys. In Shock Waves nnd High-Strain-Rate Phenomena in Metals. M. A. Meyers and L. E. Murr, eds., Plenum, New York, pp. 607-673 (1981). 20. J. C. Sanchez, K. I. Staudhammer, and L. E. Murr: Oblique shock loading of copper rods with different grain sizes. In Metallurgical and Materials Applications of Shock-Wave and High-Strain-Rate Phenom-

276
Tantalum, A. Crawson and E. Chen, Warrendal(l996) in press. eds. TMS,

L. E. Murr et al.
pact Loading and Dynamic Behavior of Materials, vol. 2, C. Y. Chiem, H. D. Kunze, and L. W. Meyer, eds., DGM Informationsgesellschaft GmbH, Oberusel, Germany, pp. 1051-1056 (1988). 37. F. I. Grace: Long-rod penetration nite thickness at normal impact. 16(3):419-428 (1995). into targets of fi-

33. S. A. Quinones, J. M. Rivas, E. P. Garcia, L. E. Murr, F. Horz, and R. P. Bernhard: Microstructure evolution associated with hypervelocity impact craters in OFHC copper. In Metallurgical and Materials Applications of Shock-Wave and High-StrainRate Phenomena, L. E. Murr, K. P. Staudhammer, and M. A. Meyers, eds., Elsevier Science, B.V., Amsterdam, p. 293-312 (1995). 34. J. M. Rivas, S. A. Quinones, and L. E. Murr: Hypervelocity impact cratering: microstructural characterization. Ser. Metall. Mater. 33(1):101-107 (1995). 35. J. M. Rivas, S. A. Quinones, E. I. Garcia, and L. E. Murr: Microstructural evolution associated with hypervelocity impact crater formation in metallic targets. In Metallurgical and Materials Applications of Shock-Wave and High-Strain-Rate Phenomena, L. E. Murr, K. P. Staudhammer, and M. A. Meyers, eds., Elsevier Science, B.V., Amsterdam, p. 313323 (1995). 36. M. Hallouin, F. Cottet, J. I. Romain, L. Monty, and M. Gerland: Microstructural transformations in iron induced by intense laser shock loading. In Im-

ht. 1.Impact Eng.

38. N. L. Rupert and F. I. Grace: Impact studies of rod projectiles against titanium and steel targets. In Metallurgical and Materials Applications of ShockWave and High-Strain-Rate Phenomena, L. E. Mm-r, K. P. Staudhammer, and M. A. Meyers, eds., Elsevier Science, B.V., Amsterdam, pp. 257-264 (1995). 39. W. Huang, L. E. Murr, C.-S. and F. I. Grace: Metallurgical tion and failure patterns from long-rod penetrators. Ibid, pp. Niou, N. L. Rupert, studies of deformatargets impacted by 265-272 (1995).

40. G. T. Gray, III: Shock loading of copper at different residual strains. In High Pressure Shock Compression of Solids, J. R. Asay and M. Shahinpoor, eds., Springer-Verlag, New York, pp. 409316 (1993). Received March, 2996; accepted June 1996.

Das könnte Ihnen auch gefallen