You are on page 1of 12

Mat. Res. Soc. Symp. Proc. Vol.

800 2004 Materials Research Society

AA5.4.1

The Shock Initiation and High Strain Rate Mechanical Characterization of Ultrafine Energetic Powders and Compositions J.E. Field, S.M. Walley*, W.G. Proud, J.E. Balzer, M.J. Gifford, S.G. Grantham, M.W. Greenaway, C.R. Siviour Physics and Chemistry of Solids Group (PCS), Cavendish Laboratory, Madingley Road, Cambridge CB3 0HE, United Kingdom * Corresponding author. email: smw14@phy.cam.ac.uk ABSTRACT This paper reviews the techniques that have been developed at the Cavendish Laboratory for the study of the mechanical and ignition properties of energetic materials. HIGH-SPEED PHOTOGRAPHY A number of techniques have been developed and applied in our laboratory for the investigation of the properties of energetic materials. One method we have used in a wide range of such studies has been high-speed photography e.g. refs [1-11]. The advantage is that it is possible to see directly what is going on in, for example, hot spot initiation of energetic materials. In recent years, it has increasingly been desired to model the impact response of structures containing energetic materials using numerical methods. If meaningful numerical results are going to be obtained for, say, munitions or rockets, it is of vital importance that constitutive relations be constructed which describe the mechanical response of unreacted energetic materials over the temperature and strain rate ranges of interest. With this in mind, we have developed a range of techniques for obtaining the mechanical properties of energetic materials over a wide range of strain rates and temperatures. Examples of publications where such data have been published include refs [4, 7, 12-20]. One problem with conventional mechanical testing methods is that they only allow the measurement of the global response of a specimen. Energetic materials usually consist of a viscoelastic binder heavily loaded with explosive crystals. In order to develop realistic and physically-based constitutive models for such unusual composites, it is vital to determine how these materials deform on the mesoscale. To this end, we have developed a range of optical and microscopy techniques. These have been used for both quasistatic [14, 15, 18, 21-23] and dynamic studies [19, 20, 24-26]. HOPKINSON BARS The split Hopkinson pressure bar (SHPB) is one of the most widely used machines for obtaining the stress-strain properties of materials at high rates of deformation (103 - 104 s-1) [27, 28]. Our initial work in this area was to develop a miniaturised direct impact system (DIHB) allowing strain rates between 104 and 105 s-1 to be accessed [29, 30]. A larger DIHB was later developed to obtain the high rate mechanical properties of polymer-bonded explosives (PBXs) and polymers [7, 31]. In recent years, we have constructed a suite of conventional SHPBs of various mechanical impedances so as to be able to measure the high rate properties of materials

AA5.4.2

ranging in strength from a few MPa to a few GPa. Inconel is used as a bar material in experiments where the specimen is held at more than 300K above room temperature (the impedance of Inconel is only a weak function of temperature [32]). The effect of grain size on the high rate of strain mechanical properties of PBXs was first noticed by Field and coworkers [13]. Since that data was published, it has been observed that the effect is made more prominent by lowering the temperature at which the experiments are carried out [33] (see also figure 1). The ammonium perchlorate (AP)/hydroxyl-terminated polybutadiene (HTPB) PBX used in the work reported here consisted of 66% AP and 33% HTPB by mass. The AP was available in four different crystal sizes: 3m, 8m, 30m and 200-300m. The low temperature study was carried out using our Inconel bar system. Cooling was performed by surrounding the ends of the bars with a chamber into which helium gas was passed that had been cooled using liquid nitrogen. The temperature was monitored using chromel-alumel thermocouples. Figure 2 shows that the effect of particle size on the flow stress of the material is linear in 1/ d where d is the particle size.

120

True Stress/MPa

100 80 60 40 20 0 0 0.05 0.1


3m, 4400300 s -1 8m, 3700200 s -1 30m, 3700250 s
-1 -1

110

Flow stress / MPa

100 90 80 70 60 50 0

200-300m, 3100350 s

True Strain

0.15

0.2

0.25

(Particle size) -0.5 / m-0.5

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Figure 1. Plot of the stress-strain responses of Figure 2. Plot of the flow stress data of figure 1 versus the reciprocal of the square root of an AP/HTPB with four different AP crystal the AP crystal size. sizes at -60 C obtained using an SHPB. OPTICAL TECHNIQUES When the standard SHPB equations are used to calculate sample stress and strain from the output of gauges on the bars, the values calculated have the feature that the strain is calculated using the calculated displacement of the sample/input bar interface. This means that the strain is effectively an average through the specimen thickness. There is therefore no way of knowing whether there are any inhomogeneities in the specimen deformation. Other assumptions that are made in deriving the SHPB equations are that the specimen is in force equilibrium and that the lubrication is perfect so that there is no barrelling [27]. Equilibrium can be checked by comparing the force-time traces on both sides of the specimen. Also lubrication is available that reduces friction of polymer-based materials to zero [34-36]. However, it would be advantageous to be able to obtain data from inhomogeneous, or brittle or foam specimens. Such materials do

AA5.4.3

not necessarily reach a constant strain state during an experiment. It would also be useful to be able to measure strain fields at low levels of deformation so that elastic properties might possibly be determined. For these reasons, we are in the process of developing and applying various dynamic optical techniques [20] to the measurement of specimen deformation in our compression split Hopkinson bar [19]. Digital Speckle Metrology (DSM) combined with high-speed photography has been used to follow the through thickness deformation of a conventional cylindrical compression specimen made from a sugar simulant of a PBX (PBS9501) (figure 3). Good agreement was found with the overall strain computed using the standard Hopkinson bar equations. DSM has also been used to follow the dynamic loading of PBS9501 in the Brazilian configuration (diametral loading of a disc) (figure 4). Such techniques are particularly valuable for materials where it is suspected that a significant volume change may take place during deformation (the standard Hopkinson bar equations assume volume conservation). It should be remembered that the SHPB is an excellent, and versatile, piece of apparatus. The quality of results produced by the system is high, and if the limitations are known, then very useful data can be obtained.

Figure 3. Displacement quiver plots for an SHPB compression experiment on PBS9501. The length of the arrows is proportional to the displacement at their bases. Note that there are arrows on both the input bar and the specimen.

Figure 4. Displacement quiver plots for a Brazilian experiment on PBS9501 in our SHPB

The arrows in figures 3 and 4 are examples of quiver plots. These have been presented in the zero momentum frame, i.e. the specimens frame of reference. The results are very encouraging: there is no strain concentration in the specimens, and the deformation in the Brazilian test is exactly as seen in quasi-static experiments. In particular, if the values of strain are taken across the centre of the specimen perpendicular to the loading direction, then the strain across the centre of the sample should be twice the overall strain according to theory. This was found indeed to be the case.

AA5.4.4

High-speed photography has often been applied to Hopkinson bar testing, for example to examine specimen deformation [29, 37, 38]. The use of an optical wedge can improve sensitivity for detecting barrelling and plastic wave activity [29]. This technique can give, for example, useful confirmation of strains calculated using the Hopkinson bar equations. It can also give indications of failure mechanisms and sample diameter. However, photography can only be used to measure overall deformations. DSM by contrast is a technique that allows surface deformation fields to be measured. Speckle techniques require a random pattern to be observed on a surface. The random pattern can then be tracked with time using a correlation algorithm providing a measure of both components of displacement between a reference and a deformed image. The resolution of the measurements depends on the size of the speckle pattern, and the resolution of the camera. The correlation algorithm used in these experiments was originally developed by Sjdahl and co-workers [39-41]. In the experiments reported above, the random speckle pattern was applied to the sample surface using an airbrush [17]. The Brazilian test produces tensile deformation in a compression apparatus [42-44]. It has the advantage of requiring much smaller samples than a traditional tensile test, which is important when working with explosives. The disk shaped specimen is held between two curved anvils. The curved nature of the anvils prevents localised strains at the points of contact, and instead ensures that the sample undergoes tensile strain along a line running down its centre. A Hadland Ultra-8 digital high-speed camera [45] was used for the experiments, with a Pallite VII radial light providing illumination. Use of a digital camera removes the need for fiducial markers to align the different photographs for comparison. The Ultra-8 duplicates the image onto 8 areas of a CCD array, each of which is sampled independently so that each of the 8 images can have completely different delays and exposure times. Before each experiment a set of reference images was taken. Each deformed image was compared to the reference image from the same part of the CCD array. This procedure removes the effect of any possible distortion in the images due to the internal optics of the camera. The exposure time and interframe time between the dynamic images were both set to 20s. DROP-WEIGHT The dropweight is a widely used machine to determine the impact sensitivity of energetic materials e.g. ref. [46]. A modification of this apparatus made in our laboratory many years ago was to machine a lightpath through the falling weight so that the deformation of material impacted between glass anvils can be observed using high-speed photography [2]. The confinement conditions are different to those in say a Rotter apparatus [13, 47], but it has the distinct advantage that it is possible to see what is going on during the formation of hot spots and subsequent deflagration. A number of classic papers based on work performed on energetic materials in this apparatus have been published; see, for example, refs [2-4, 6, 7, 11]. Figure 5 presents selected frames from a high-speed photographic sequence taken using an AWRE C4 rotating mirror framing camera [48] with an interframe time of 7s of the deflagration caused by the impact of our dropweight apparatus on three energetic powders. Burning can be seen to start at the interface between the AP and PETN powders. The flame generated initiated the burning of the HMX. The last to deflagrate is the AP. Note especially the prominent fracture patterns in the sintered disc of AP. The largest of these provides an easy path for the escape of gaseous products.

AA5.4.5

Figure 5. Selected frames from a high-speed photographic sequence of dropweight impact of powders of: AP (left), PETN (right), HMX (bottom). Time from moment of impact: 0, 406, 420, 455 s. From ref. [49]. PLATE IMPACT AND SMALL GAS GUNS Laboratory gas guns are widely used to produce well-characterised 1D shock waves in materials. Our laboratory has designed and built a single-stage 50mm bore gas-gun system capable of performing impacts at up to 1.2 km s-1 [50]. Recently a smaller 19mm bore facility has been constructed for studying the dynamic properties of energetic materials [51, 52]. This has been used to study hot spot formation and quenching in thin beds of ammonium nitrate (AN). The impact cell has a 2mm thick copper plate on the impact side. Behind this plate is the bed of material of interest. High-speed photography is performed through the rear window which consists of a 25mm thick glass window. An optic fibre is used to take any light emitted to a photodiode (time resolution 1 ns) and to a UV/visible spectrometer (capture time 1 ns). The cell

AA5.4.6

was shocked by a 2mm thick copper flyer mounted on the front of a nylon sabot. This produces a shock of 600 ns duration. Figure 6 presents two frames from a sequence taken using an Ultra 8 camera [45] of an impact performed on a 0.4mm thick layer of AN, grain size 150-210 m. The exposure time was 30 ns and the interframe time was zero i.e. the second frame began to be taken immediately the first frame was captured. Two regions can be identified and these are ringed in figure 6. In one region the glowing hot spots are close together and glow more intensely in frame 2. In the other region the hot spots are further apart and glow less intensely in frame 2: they are in the process of being quenched. More details of this experiment, including spectroscopic results, may be found in refs. [51, 52].

Figure 6. Selected frames from a high-speed photographic sequence of the shockloading of a 0.4mm thick layer of AN. Interframe time 30ns. From ref. [52]. LASER-FLYER A system has been developed in our laboratory for the launch of miniature flyer plates at velocities approaching 10 km/s. Laser-induced plasma is used to drive these flyers which are typically less than 1 mm in diameter and a few microns thick. High-speed streak photography has been used to measure the key parameters of the flyers as a function of launch energy, mass and distance travelled. A fibre-optic beam shaping system has proved highly effective at manipulating the shape of the flyers and excellent reproducibility is found. Using a laser as the driver to launch flyers has numerous advantages in terms of repetition, reproducibility and control. The dimensions and performance of the flyer are controlled by the optical and physical properties of the laser pulse and launch target. Precise measurement of flyer performance is achieved using high-speed streak photography. Characterisation experiments have focused on all aspects of the flyer behaviour including diameter, velocity, planarity and integrity. Beam shaping, using a fibre-optic system, enables the shape and profile of the flyer to be altered [53]. In most applications, flat planar flyers are desirable and to achieve this, a tophat light profile is necessary.

AA5.4.7

Low light fluences are required for flyer launch, typically less than 20 J/cm2 and highspeed diagnostics ensure good characterisation. The system is suitable for producing single or arrays of flyers of well-defined size, shape and velocity [54, 55]. We have used this system to study the detonics of explosives [56-58]. A Nd:YAG laser is used, producing 9 ns (full-wave half maximum) duration pulses at its fundamental wavelength, 1064 nm. The laser energy is concentrated onto a substrate-backed metal film. The optical radiation is readily absorbed at the substrate-film interface to produce hot plasma (~ 8000 K). The rapid expansion of this plasma is sufficient to launch the remaining depth of film as a flyer plate. The flyers reach their maximum velocity within 100 ns and can traverse a few hundred m in air. The effect of friction heating, velocity gradients within the flyer and heating from the plasma cause the flyer to effectively burn up. The first of these is believed to be the dominant mechanism so that in a vacuum, the flyers would be expected to be able to travel much further. Flyer characterization has been performed using high-speed streak photography [55, 59, 60]. The flyer was allowed to impact a transparent window held on the optical axis and parallel to the launch substrate. The gas trapped between the flyer and window is compressed adiabatically. This process liberates light which is photographed by a high-speed camera looking back down the optical path. Streak photography was used to obtain a record of the flyer impact and a timeof-flight measurement. In this technique, the impact event was viewed through a 100 micron slit, which was moved at 10 ns/mm across a photographic film to produce the streak records. Example flyers produced with this system are shown in figure 7. This laser-driven flyer plate system was found to be capable of initiating fine grain HNS and PETN. HNS was found to be more insensitive to this type of initiator. In both cases, the sensitivity of the material was heavily dependent on particle size. Neither the coarse grained HNS nor the PETN equivalent could be detonated [58].

Figure 7. Examples of two laser-drive flyer plates moving at ca. 2 km/s. The LH one was produced using beam shaping. The RH one was Figure 8. Schematic comparison of the width produced without beam shaping. of the shock produced by a laser flyer with conventional (LH) and ultrafine (RH) energetic crystals. From ref. [58]. The large dependence on particle size is explained by the size of the critical hot spots. The shock width is of the order of a few microns, which is comparable to the grain size (and hence gas spaces) of the fine materials but is much smaller than the grain size of the coarse grain

AA5.4.8

materials. Thus, in these materials, the gas spaces cannot be entirely collapsed during the short duration of the shock (see figure 8). In addition, there is better mixing of the hot spots and explosive particles in a fine grain material. These differences offer an explanation as to why the fine grain materials are more sensitive to these very short shocks. A number of possibilities for increasing the mode generation within the fiber were investigated: bending and looping the fiber, improving the injection optics, incorporating a mode scrambler, and increasing the fibre length [53]. Bending and looping alters the angle of the internal reflections within the fiber and thus can generate more modes if a multimode laser is used. However, as we are using a single-mode laser, this method was found not to work. Some examples of the emerging spot of light are given in figure 9 (increasing top hat factor from the left- to the righthand side). The profile with the lowest top hat factor (lefthand side) shows the largest fluctuations and profile with the highest top hat factor (right-hand side) shows a much more even distribution of energy.

1m straight 10m coiled 20m coiled Figure 9. Effect of pulse shaping on the intensity distribution of light across a laser pulse. From ref. [53]. DEFLAGRATION-TO-DETONATION TRANSITION (DDT) Two types of deflagration-to-detonation transition (DDT) have been identified in columns of explosive materials composed of small crystals [61-63]. Which one occurs has been found to depend upon the density of packing (porosity). We have found that classic type I DDT occurs in columns of PETN with a density greater than 50% TMD (see figure 10). What happens is that convective burning from one end compresses the material ahead of it into an impermeable plug which at some critical pressure and velocity shocks the material the other side into detonating. Figure 11 shows a streak record of type II DDT process occurring in a charge of ultrafine PETN pressed to a density of 29 1% TMD. A schematic of the confinement is shown next to the streak record at the same scale as the photograph. The light output detected using optical fibres and photodiodes showed that the initial convective burning stage took some 600 s to propagate along the length of the column. The column then continued to burn for a period of around 70 s before the detonation broke out. The streak record gives detail of the events in the final 40 s prior to the detonation and the following 30 s.

AA5.4.9

Figure 10. A type I DDT event in ultrafine PETN. From ref. [64]. At the time indicated by the dashed line A, it can clearly be seen that reaction is taking place at a number of points which appear stationary on the time scale of the streak record. As the camera can only image the surface of the column it can be surmised that these reaction sites are in the part of the charge in contact with the window. The luminosity of the reaction at these sites continues to increase until at site B the conditions reach the critical parameters for initiation. There appears to be no shock wave associated with the build up to the initiation event. A detonation wave (C) then propagates downstream (relative to the ignition site) at an average velocity of 5.4 0.2 mm s-1. There is also an upstream retonation wave (D), that propagates from the point B, but the proximity of the initiation to the upstream end of the column makes any calculation of the velocity of this wave difficult. A reflected shock (E) travelling at 2.23 0.09 mm s-1 can also be seen propagating upstream from the point at which the detonation (C) reached the witness plate at the downstream end of the column.
O C E 20mm S 10s A B D I O P

Figure 11. Streak photograph of a type II DDT event in ultrafine PETN. A schematic of the column configuration is shown to scale to the right of the photograph. P - PETN column; O optical fibres; F - Fuse. From ref. [65]. Other experiments in which type II DDT occur are all similar in the way in which the reaction proceeds, but the point at which the detonation breaks out can be anywhere along the length of the column. The first step appears to be the formation of a channel in the low density material during the convective burn stage. There appears to be no correlation between either column length or density and the position of the detonation break-out.

AA5.4.10

ACKNOWLEDGEMENTS We would like to thank the following for financial support of this work: QinetiQ, [dstl], AWE, EPSRC, European Office of the US Army, Airforce, and Navy. REFERENCES 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. M. M. Chaudhri and J. E. Field, Proc. R. Soc. Lond. A 340 113 (1974) S. N. Heavens and J. E. Field, Proc. R. Soc. Lond. A 338 77 (1974) J. E. Field, G. M. Swallowe, and S. N. Heavens, Proc. R. Soc. Lond. A 382 231 (1982) G. M. Swallowe and J. E. Field, Proc. R. Soc. Lond. A 379 389 (1982) N. K. Bourne and J. E. Field, Proc. R. Soc. Lond. A 435 423 (1991) J. E. Field, N. K. Bourne, S. J. P. Palmer, and S. M. Walley, Phil. Trans. R. Soc. Lond. A 339 269 (1992) S. M. Walley, J. E. Field, and S. J. P. Palmer, Proc. R. Soc. Lond. A 438 571 (1992) P. M. Dickson and J. E. Field, Proc. R. Soc. Lond. A 441 359 (1993) P. E. Luebcke, P. M. Dickson, and J. E. Field, Proc. R. Soc. Lond. A 448 439 (1995) N. K. Bourne and J. E. Field, Proc. R. Soc. Lond. A 455 2411 (1999) S. M. Walley, J. E. Balzer, W. G. Proud, and J. E. Field, Proc. R. Soc. Lond. A 456 1483 (2000) J. E. Field, G. M. Swallowe, P. H. Pope, and S. J. P. Palmer, Inst. Phys. Conf. Ser. 70 381 (1984) J. E. Field, S. J. P. Palmer, P. H. Pope, R. Sundararajan, and G. M. Swallowe, "Mechanical properties of PBXs and their behaviour during drop-weight impact", Proc. Eighth Symposium (Int.) on Detonation, ed. J. M. Short (Naval Surface Weapons Center, White Oak, Maryland, USA, 1985), p. 635 S. J. P. Palmer, J. E. Field, and J. M. Huntley, Proc. R. Soc. Lond. A 440 399 (1993) H. T. Goldrein, J. M. Huntley, S. J. P. Palmer, M. B. Whitworth, and J. E. Field, "Optical techniques for strength studies of polymer bonded explosives", Proc. 10th Int. Detonation Symposium, ed. J. M. Short and D. G. Tasker (Office of Naval Research, Arlington, Virginia, 1995), p. 525 P. J. Rae, H. T. Goldrein, S. J. P. Palmer, J. E. Field, and A. L. Lewis, Proc. R. Soc. Lond. A 458 743 (2002) P. J. Rae, S. J. P. Palmer, H. T. Goldrein, J. E. Field, and A. L. Lewis, Proc. R. Soc. Lond. A 458 2227 (2002) H. T. Goldrein, P. J. Rae, S. J. P. Palmer, and J. E. Field, Phil. Trans. R. Soc. Lond. A 360 939 (2002) C. R. Siviour, S. M. Walley, W. G. Proud, and J. E. Field, "Hopkinson bar studies on polymer bonded explosives", Proc. 6th Seminar on New Trends in Research of Energetic Materials, ed. J. Vgenknecht (University of Pardubice, Pardubice, Czech Republic, 2003), p. 338 S. G. Grantham and J. E. Field, Proc. SPIE 4933 27 (2003) H. T. Goldrein, S. J. P. Palmer, and J. M. Huntley, Optics Lasers Engng 23 305 (1995) W. G. Proud, S. J. P. Palmer, J. E. Field, G. Kennedy, and A. Lewis, Thermochim. Acta 384 245 (2002)

14. 15.

16. 17. 18. 19.

20. 21. 22.

AA5.4.11

23.

24. 25. 26. 27.

28.

29. 30. 31. 32. 33. 34. 35. 36. 37.

38.

39. 40. 41.

42. 43. 44. 45.

H. T. Goldrein, P. J. Rae, S. J. P. Palmer, A. L. Lewis, G. Miles, and N. Zahlan, "Wholefield displacement measurement with moir interferometry as a basis for microstructural FE model of a polymer bonded explosive", Field Analyses for Determination of Material Parameters: Experimental and Numerical Aspects, ed. P. Sthle and K. G. Sundin (Kluwer, Dordrecht, The Netherlands, 2003), p. 1 J. M. Huntley and J. E. Field, Engng Fract. Mech. 30 779 (1988) M. B. Whitworth, J. M. Huntley, and J. E. Field, Proc. SPIE 1358 677 (1990) J. M. Huntley and J. E. Field, Opt. Engng 33 1700 (1994) G. T. Gray III, "Classic split-Hopkinson pressure bar testing", ASM Handbook. Vol. 8: Mechanical Testing and Evaluation, ed. H. Kuhn and D. Medlin (ASM International, Materials Park, Ohio, 2000), p. 462 J. E. Field, W. G. Proud, S. M. Walley, and H. T. Goldrein, "Review of experimental techniques for high rate deformation and shock studies", New Experimental Methods in Material Dynamics and Impact, ed. W. K. Nowacki and J. R. Klepaczko (Institute of Fundamental Technological Research, Warsaw, Poland, 2001), p. 109 D. A. Gorham, Inst. Phys. Conf. Ser. 47 16 (1980) D. A. Gorham, P. H. Pope, and J. E. Field, Proc. R. Soc. Lond. A 438 153 (1992) S. M. Walley and J. E. Field, DYMAT Journal 1 211 (1994) R. Kandasamy and N. S. Brar, Exper. Mech. 35 119 (1995) J. E. Balzer, C. R. Siviour, S. M. Walley, w. G. Proud, and J. E. Field, Proc. R. Soc. Lond. A (accepted for publication) (2003) B. J. Briscoe and R. W. Nosker, Wear 95 241 (1984) S. M. Walley, J. E. Field, P. H. Pope, and N. A. Safford, Phil. Trans. R. Soc. Lond. A 328 1 (1989) S. M. Walley, J. E. Field, P. H. Pope, and N. A. Safford, J. Phys. III France 1 1889 (1991) G. Gary, L. Rota, and H. Zhao, "Testing viscous soft materials at medium and high strain rates", Constitutive Relation in High/Very High Strain Rates, ed. K. Kawata and J. Shioiri (Springer-Verlag, Tokyo, 1996), p. 25 C. R. Siviour, S. M. Walley, W. G. Proud, and J. E. Field, "Are low impedance Hopkinson bars necessary for stress equilibrium in soft materials?" New Experimental Methods in Material Dynamics and Impact, ed. W. K. Nowacki and J. R. Klepaczko (Institute of Fundamental Technological Research, Warsaw, Poland, 2001), p. 421 M. Sjdahl and L. R. Benckert, Appl. Opt. 32 2278 (1993) M. Sjdahl, Appl. Opt. 33 6667 (1994) M. Sjdahl, "Strain field measurements using electronic speckle photography: A comparison", Recent Advances in Experimental Mechanics, ed. S. Gomes (Balkema, Rotterdam, 1994), p. 325 F. L. Carneiro and A. Barcellos, RILEM Bull. 13 98 (1949) G. Hondros, Austral. J. Appl. Sci. 10 243 (1959) A. B. Fadeev, Sov. Mining Sci. 10 378 (1974) G. W. Smith, M. J. Riches, R. B. Huxford, P. A. Smith, C. L. G. Seymour, and J. D. Bell, Proc. SPIE 4183 105 (2001)

AA5.4.12

46.

47. 48.

49. 50. 51.

52. 53. 54. 55. 56. 57.

58.

59. 60. 61. 62. 63. 64.

65.

C. S. Coffey and V. F. DeVost, "Drop weight impact machines: A review of recent progress", JANNAF Propulsion Systems Hazards Subcommittee Meeting (CPIA Publ. no. 446 vol. 1), (Chemical Propulsion Information Agency, Columbia, Maryland, 1986), p. 527 H. N. Mortlock and J. Wilby, Explosivstoffe 14 49 (1966) K. R. Coleman, "The photography of high temperature events", Proc. Fourth Int. Kongress Kurzzeitphotographie, ed. H. Schardin and O. Helwich (Verlag Dr. Othmar Helwich, Darmstadt, Germany, 1959), p. 32 J. E. Balzer, Low-level impact loading of explosives, PhD, Univ. of Cambridge (2001) N. K. Bourne, Z. Rosenberg, D. J. Johnson, J. E. Field, A. E. Timbs, and R. P. Flaxman, Meas. Sci. Technol. 6 1462 (1995) W. G. Proud, "The measurement of hot-spots in granulated ammonium nitrate", Shock Compression of Condensed Matter - 2001, ed. M. D. Furnish, N. N. Thadhani, and Y. Horie (American Institute of Physics, Melville, NY, 2002), p. 1081 W. G. Proud, E. J. W. Crossland, and J. E. Field, Proc. SPIE 4948 510 (2003) M. W. Greenaway, W. G. Proud, J. E. Field, and S. G. Goveas, Rev. Sci. Instrum. 73 2185 (2002) S. Watson and J. E. Field, J. Phys. D: Appl. Phys. 33 170 (2000) S. Watson and J. E. Field, J. Appl. Phys. 88 3859 (2000) S. Watson, M. J. Gifford, and J. E. Field, J. Appl. Phys. 88 65 (2000) M. W. Greenaway, M. J. Gifford, W. G. Proud, J. E. Field, and S. Goveas, "Laser-driven flyer plate initiation of explosives", Theory and Practice of Energetic Materials, ed. L. Chen and C. Feng (China Science and Technology Press, Beijing, 2001), p. 451 M. W. Greenaway, M. J. Gifford, W. G. Proud, J. E. Field, and S. G. Goveas, "An investigation into the initiation of hexanitrostilbene by laser-driven flyer plates", Shock Compression of Condensed Matter - 2001, ed. M. D. Furnish, N. N. Thadhani, and Y. Horie (American Institute of Physics, Melville, NY, 2002), p. 1035 S. Watson, The production and study of laser-driven flyer plates, PhD thesis, Univ. of Cambridge (1998) M. W. Greenaway, The development and characterisation of a laser-driven flyer system, PhD thesis, Univ. of Cambridge (2001) B. S. Ermolaev, A. A. Sulimov, V. A. Okunev, and V. E. Khrapovskii, Combust. Explos. Shock Waves 24 59 (1988) V. E. Khrapovskii, Combust. Explos. Shock Waves 29 129 (1993) D. Price and R. R. Bernecker, Propell. Explos. 6 5 (1981) M. J. Gifford, A. Chakravarty, M. W. Greenaway, S. Watson, W. G. Proud, and J. E. Field, "Unconventional properties of ultrafine energetic materials", Energetic Materials: Ignition, Combustion and Detonation, ed. (DWS Werbeagentur und Verlag GmBH, Karlsruhe, 2001), p. paper 100 M. J. Gifford, P. E. Luebcke, and J. E. Field, J. Appl. Phys. 86 1749 (1999)

You might also like