You are on page 1of 11

Journal of Applied Mechanics and Technical Physics, Vol. 54, No. 1, pp. 10–20, 2013.

Original Russian Text 


c G. Hussain, A. Hameed, J.G. Hetherington, A.Q. Malik, K. Sanaullah.

ANALYTICAL PERFORMANCE STUDY OF EXPLOSIVELY FORMED PROJECTILES

G. Hussaina , A. Hameedb , J. G. Hetheringtonb , UDC 623.4.082.6


a a
A. Q. Malik , and K. Sanaullah

Abstract: Hydrocode simulations are carried out using Ansys Autodyn (version 11.0) to study the
effects of the liner material (mild steel, copper, armco iron, tantalum, and aluminum) on the shape,
velocity, traveled distance, pressure, internal energy, temperature, divergence or stability, density,
compression, and length-to-diameter ratio of explosively formed projectiles. These parameters are
determined at the instants of the maximum as well as stable velocity during the flight towards the
target. The results of these parameters present the potential capability of each liner material used to
fabricate explosively formed projectiles. An experimental analysis is performed to study the velocity
status and the length-to-diameter ratio of explosively formed projectiles.
Keywords: velocity, yield stress, pressure, compression, divergence, stability.
DOI: 10.1134/S0021894413010021

INTRODUCTION

A shaped charge comprises a conical metallic liner, which, when deformed to form a jet, has large velocity
gradients. As a result of these velocity gradients, the jet disintegrates into a number of fragments, which restricts
the shaped charge range. In order to attack at greater standoff distances, we have to modify the design configuration
of the shaped charge, which leads to an explosively formed projectile (EFP) [1]. An EFP has a liner with a greater
apex angle. The blast force moulds the liner into a configuration determined by the geometry, casing, and initiation
system. The EFP shape, velocity, and performance depend on both geometric and material factors [2].
Liner contours, physical dimensions of the explosive charge (charge length and length-to-diameter (L/D) ratio
of the explosive charge), confinement configuration, and explosive initiation technique are some of the geometrical
factors of interest. The material factors depend on the structure, properties, and processing conditions during the
fabrication of the liner, casing, and explosive. The desirable properties of the liner in the context of the dynamic
EFP formation process and its eventual effectiveness as a penetrator are high density, high ductility, high strength,
and high enough melting temperature to prevent melting in the liner due to adiabatic heating [2]. The casing is
typically made of steel because of its relatively low cost, high strength, and high density. However, other materials
can also be used, as long as the mass is sufficient to provide the necessary confinement. The explosive properties of
importance are the explosive density, detonation velocity, and explosive energy [3]. The liner material that has a
higher velocity gives a higher value of the impact energy at the target, leading to higher penetration [4]. The L/D
ratio of the projectile is very a important parameter regarding the penetration capability. The higher the L/D
ratio, the deeper the penetration into the target [5, 6]. The solid EFP has a 10% higher penetration capability than

a
National University of Sciences and Technology, Islamabad, Pakistan. b Cranfield University at the Defence
Academy of the United Kingdom, Shrivenham, SN6 8LA UK; a.hameed@cranfield.ac.uk. Translated from Prik-
ladnaya Mekhanika i Tekhnicheskaya Fizika, Vol. 54, No. 1, pp. 13–24, January–February, 2013. Original article
submitted March 28, 2011; revision submitted August 9, 2011.
10 0021-8944/13/5401-0010 
c 2013 by Pleiades Publishing, Ltd.
the hollow EFP [7, 8]. The L/D ratio, degree of solidity, and impact velocity are important parameters of an EFP,
which play the major role in penetrating the target [9].
The shape and velocity gradients are important for performance of the projectile, which was demonstrated
in [10]. The metallurgical parameters of the liner material affect its capability for absorbing the velocity gradients by
manifesting it in the plastic work done. The velocity gradients, if not fully absorbed, would produce a continuously
stretching penetrator, which would ultimately disintegrate. The plastic work done controls the increased length of
the penetrator, stabilizing the velocity and shape of the latter.
Sharma et al. [4] expressed the energy absorbed per unit volume of the material E/V in plastic deformation
under quasi-static condition as

E K ∗ εn+1
= σ dε = ,
V n+1
where σ, ε, and n are the flow stress, strain, and strain hardening exponent, respectively; K ∗ is the strength
coefficient of the material. The properties of the material that influence E/V are the dynamic strength and ductility
of the material. This ratio is not very responsive to the strain hardening exponent. The larger the E/V ratio for
the material, the higher its capacity to absorb gradients without failure, which makes it capable of withstanding
higher explosive energy input [4]. High energy explosives, i.e., HMX, are used to produce high L/D ratios for this
particular design.
This study presents comparative parametric numerical simulations and experimental analysis of materials
used as liners in EFPs.

1. SIMULATION SCHEME

The formation, elongation, and performance of EFPs made of different materials are modeled by using the
Autodyn-2D hydrocode. The reason for selecting the Autodyn-2D finite-difference code is its versatility with respect
to its solving processors. These processors are well capable of tackling different kind of problems and also improving
the formation and performance of the self-forging devices [11]. The Euler hydrocode scheme is adopted to avoid
the problem of eroding cells [12]. The significance of the use of the Autodyn-2D hydrocode in the field of EFPs was
demonstrated in the literature [13–16].
The Johnson–Cook model is used as a strength model in the simulation. It expresses the flow stresses in
terms of the equivalent plastic strain, plastic strain rate, and homologous temperature [17]:
σy = (A + Bεnp )(1 + C ln ε∗ )(1 − (T ∗ )m ).
The expression in the first brackets gives the stress as a function of strain; the expressions in the second and third
brackets represent the effect of the strain rate and temperature, respectively, εp is the effective plastic strain, ε∗ is
the strain rate, A, B, n, C, and m are the material constants, and T ∗ = (T − Tref )/(Tmelt − Tref ) is the homologous
temperature (Tref and Tmelt are the reference and melting temperatures of the material).
Heat is generated in an element by plastic work, and the resulting rise in temperature is computed by using
the specific heat for the material c. We applied the Jones–Wilkins–Lee (JWL) equation of state to HMX, which is
given below:
 ωη   R   ωη   R 
1 2
P = AJWL 1 − exp − + BJWL 1 − exp − + ωηρref e.
R1 η R2 η
Here P is the pressure, η = ρ/ρref , ρ is the current density, ρref is the reference density, e is the specific internal
energy, AJWL , BJWL , R1 , R2 , and ω are the material properties of the chemical high explosive [12, 18]. The
initial values of these constants used in the simulation are ρref = 1.89 g/cm3 , AJWL = 9.4334 · 10−1 TPa, BJWL =
8.8053 · 10−3 TPa, R1 = 4.7, R2 = 0.9, ω = 0.35, and e = 1.02 · 10−2 kJ/mm3 [19–21].
The input parameters of materials are listed in Table 1 (ρref is the reference density, K is the bulk modulus,
c is the specific heat, G is the shear modulus, A is the yield stress, B is the hardening constant, n is the hardening
exponent, C is the strain rate constant, m is the thermal softening exponent, and Tmelt is the melting temperature).
It should be noted that all liner materials are described by the Johnson–Cook strength model and the linear equation
of state. The instantaneous geometric strain due to erosion is selected to be 2.5. The reference temperature for all
liner materials is selected to be Tref = 300 K.
11
43.5
39

40.5

30
36
3
33.5 2

Fig. 1. Geometric shape of the EFP warhead.

2. DESIGN CONFIGURATION

Numerical simulation is carried out using the Ansys Autodyn (version 11.0) hydrocode to study different
material parameters used as liners in the configuration of the tapered HMX explosive (density 1.89 g/cm3 ) with
4340 steel as a casing material. Geometrical trough-shaped liners and explosive configurations with the length-
to-diameter ratio of 1.3 are used with a constant liner wall thickness of 2.0 mm. The diameter of the modeling
configuration is 36.0 mm with confinement of mild steel included, as is shown in Fig. 1. A symmetric configuration
of the design is selected, since it supports axial plastic deformation in the EFP [22]. The point initiation method is
used in the simulation.

3. EXPERIMENTAL SETUP

The EFP assembly used in the experiments is shown in Fig. 2. The EFP warhead and sand bags, for
soft recovery of the warhead, are located 5 m apart. The flight trajectory is captured by a high-speed camera to
determine the EFP velocity, shape, and L/D ratio.

4. RESULTS AND DISCUSSION

The velocity gradient profiles and decay rates of the liner materials are analyzed by comparing the results of
simulation and experiments. Also, the pressure, yield stress, internal energy, density, temperature, and L/D ratio
of EFPs are investigated.

4.1. Output Parameters of the EFP at the Maximum Velocity

The input parameters of the materials are summarized in Table 1. The output parameters of the liner
materials, such as mild steel, copper (Cu), armco iron (Fe), tantalum (Ta), and aluminum (Al) at the instant of
the maximum velocity have never been reported in the literature. The values of these parameters for different
materials are listed in Table 2 (v is the velocity, t is the time, S is the distance traveled, ρ is the density, T is
the temperature, ΔS is the divergence, σt max is the maximum yield stress, ΔV is the maximum compression, e is
the internal energy, and P is the pressure). Aluminum is the liner material that provides the highest values of the
distance traveled, compression, internal energy, and L/D ratio of the EFP in the least time. The outcomes are the
result of Al material characteristics, i.e., softness, low density, low melting point, and inter atomic structure (atoms
are not densely packed like in other liner materials). The initial compression of Al is mostly affected by sudden
shock waves of the explosive due to its low density and comparatively weak interatomic structure, resulting in more
axial stretching leading to a longer projectile with a higher L/D ratio.
12
13
Fig. 2. Schematic diagram of the experimental setup.

(a) (b) (c) (d) (e)

Fig. 3. EFP shapes at the maximum velocity: (a) mild steel; (b) copper; (c) tantalum;
(d) aluminum; (e) armco iron.

The velocity and traveled distance of Ta are the lowest with the highest time consumption and the greatest
change in density, which results in the highest temperature rise due to the low thermal softening exponent, high
density, and large hardening exponent.
The divergence amplitude is the lowest for mild steel and the highest for Cu due to their respective charac-
teristics. The highest yield stress and pressure are observed for the EFP made of armco iron.
The EFP shapes at the maximum velocities are given in Fig. 3. We observed curved profiles for projectiles
made of low-density materials and straighter ones for high-density materials. The pressure intensity variation in
the central region for mild steel and aluminum was reported in [23]. This study causes a broad spectrum of velocity
variation for all liner materials, which show a high velocity in the central region, decreasing toward the edges.

4.2. Output Parameters of the EFP at the Stable Velocity

The output parameters of mild steel, Cu, Ta, Al, and Fe at the stable velocities are given in Table 3. At this
instant, the stable velocity of the EFP made of aluminum is the highest. This is one of the reasons why the effective
standoff distance is greater for Al than for more dense metals. Also, the divergence amplitude or instability, internal
energy, and density change are the highest in Al due to its softness, low density, low melting point, and interatomic
structure. The time to obtain the stable velocity and the distance traveled are the lowest for Al as well.
The Ta EFP has a larger diameter, which would produce a wider crater in the target. The Fe EFP has a
diameter very close to the Ta EFP, which makes it attractive for producing a wide crater due to its economical
market value and easy machining capability. The temperature generated within the Ta EFP is the highest, which
indicates the potential to produce hot fragments.
The L/D ratio for the Cu EFP at the stable velocity is the highest; hence, the reason to prefer Cu for
achieving the maximum possible penetration. The EFP made of mild steel is the most reliable or stable liner
material due to the minimum divergence. So this material is a very effective replacement for Fe in the multiliner
EFPs with Ta as a penetrator and Fe as a stabilization base [2]. The yield stress developed in the Fe EFP is the
highest among the liner materials.
14
15
(a) (b) (c)

(d) (e)

Fig. 4. Experimental EFP shape (left) and simulated EFP shape (right) at stable velocities:
(a) mild steel; (b) copper; (c) tantalum; (d) aluminum; (e) armco iron.

These parameters are settled at the stable velocity and have infinitesimal effects until the impact with the
target. The EFP shapes of the simulation and the experiments are shown in Fig. 4. Good consistency is found
between the shapes of the EFPs for the simulation and the experiments. Mild steel and copper are preferable
materials for solid EFPs, while tantalum, aluminum, and armco iron are preferable materials for hollow EFPs. It
should be noted that solid EFPs made of mild steel and copper offer a 10% improvement in penetration efficiency.

4.3. Velocity Profiles

The variation in the velocities of the liner materials shows how the velocity, for a particular liner material,
rises to a maximum value and then falls to a stable value (Fig. 5). The maximum and stable velocity of each liner
depends on material properties.
The Ta EFP provides the lowest maximum and stable velocities in the longest time, whereas the Al EFP
ensures the highest maximum and stable velocities in the shortest time. So the liner density is proportional to the
EFP velocity.
The time tst necessary to stabilize the EFP from its maximum velocity, the maximum velocity vmax and
stable velocity vst , the deceleration or velocity decay rate a, and the L/D ratio of EFPs for each liner material
(δ is the error) are tabulated in Table 4. The time necessary to reach the stable velocity is directly proportional
to the liner material density. The highest percentage difference for the maximum velocity is 3.5% and the highest
percentage difference for the stable velocity is 1.5% for the Ta EFP.
The highest percentage difference for deceleration or the velocity decay rate for the liner material is l3.0%
for Ta. The comparison clearly shows that Cu is the liner material which has the largest L/D ratio, whereas Ta
has the smallest L/D ratio. The maximum difference between the numerical simulation and experiment is 9.7% for
the Cu EFP.
The velocity decay rate for the liner material is inversely proportional to its density. Other material param-
eters also affect the velocity decay rate (Fig. 6). Although these materials have approximately similar densities,
their velocity decay rates are different. The Fe EFP has a higher velocity decay rate than the EFP made of mild
steel because of its higher yield stress, hardening constant, strain rate constant, and melting temperature.

4.4. Liner Materials Output Characteristics

The EFP velocity varies along its trajectory to achieve its stable value. The other parameters, such is the
pressure, density, temperature, and internal energy show a similar behavior. Figure 7 shows the variation in the
pressures of the liner materials, displaying a pressure rise to a maximum value and then attenuation to a stable
value. The stabilized pressures in the liner materials are proportional to their densities.
16
v, m/s
3500
1
3000
2500
2000 2
3
1500 4
5
1000
500
0 20 40 60 80 100 120 t, ms

Fig. 5. EFP velocity profiles for aluminum (1), mild steel (2), armco iron (3), copper (4), and tantalum (5).

a, m/s2 P .104, TPa


21 1 16
1
2 14 2
18 3
12 3
15 4 4
5 10 5
12 6 8
9 6
4
6
2
3 0
_2
0 3 6 9 12 15 r, g/cm3 0 20 40 60 80 100 120 t, ms

Fig. 6. Fig. 7.

Fig. 6. Deceleration of the liner materials against their densities: points 1–5 show the calculated results for alu-
minum (1), mild steel (2), armco iron (3), copper (4), and tantalum (5); points 6 show the experimental data for
the corresponding materials.
Fig. 7. Pressure variations in the liner materials for aluminum (1), mild steel (2), armco iron (3), copper (4), and
tantalum (5).

The maximum generated shock pressure, internal energy change, density change, and temperature change of
the liner materials are tabulated in Table 5. Aluminum is softer, less dense, and less densely packed, so its internal
energy is the highest upon experiencing the action of a shock wave of the same strength. The Ta EFP has the
lowest internal energy change due to its higher density and closely-packed atomic structure. The internal energy of
the EFPs is inversely proportional to the density of the liner material. The density change of the liner materials
depends on the properties of the materials including their reference density and hardness (interatomic structure).
It is suggested that, as the density of the material rises, atoms become more closely packed and each atom
influences its surrounding atoms more actively, resulting in immediate transfer of its energy to others, increasing
the temperature of the material. This phenomenon proved true in the EFPs, resulting in the highest temperature
change in the Ta EFP, which had the maximum density change.
The pressure, internal energy, and temperature variations within the EFPs and their L/D ratio variations
at the stable velocity are plotted in Figs. 8–11. The maximum fluctuations in the value of pressure are within the
EFP made of mild steel, whereas Cu shows the least variation. A significant variation in internal energy is found
in the low-density materials (aluminum, mild steel, armco iron). For all materials, the L/D ratio rises initially and
then becomes constant as the EFPs reach their stable velocities.
17
Table 5. Maximum changes of the liner material parameters
Material Pmax · 104 , Pa emax , kJ/g ρmax , g/cm3 Tmax , K
Aluminum 8.206 0.8970 0.040 177
Mild steel 8.910 0.2190 0.036 148
Armco iron 14.600 0.2270 0.071 203
Copper 12.190 0.1720 0.088 151
Tantalum 12.300 0.0946 0.141 319

e, kJ/g
P .104, TPa 1.0 1 2 3 4 5
0.6 1 2 3 4 5
0.8
0.4
0.6
0.2

0 0.4

_0.2 0.2

_0.4
0 1 2 3 x, mm 0 20 40 60 80 100 120 t, ms

Fig. 8. Fig. 9.

Fig. 8. Pressure fluctuations within the EFP for mild steel (1), copper (2), armco iron (3), alu-
minum (4), and tantalum (5).
Fig. 9. Internal energy profiles of the liner material for aluminum (1), mild steel (2), armco iron
(3), copper (4), and tantalum (5).

T, Ê L/D
650 1 2 3 4 5 2.0 1 2 3 4 5
600
550 1.5
500
1.0
450
400 0.5
350
300
0 20 40 60 80 100 120 t, ms 0 20 40 60 80 100 120 t, ms

Fig. 10. Fig. 11.

Fig. 10. Temperature profiles of the liner materials for aluminum (1), mild steel (2), armco iron
(3), copper (4), and tantalum (5).
Fig. 11. L/D ratio profiles of the liner materials for aluminum (1), mild steel (2), armco iron (3),
copper (4), and tantalum (5).

CONCLUSIONS

The potential capability of a range of liner materials (mild steel, copper, tantalum, aluminum, and armco
iron) for EFPs is investigated by comparing their output parameters: shape, velocity, pressure, density, internal
energy, temperature, and L/D ratio. These parameters are measured at the instants of their maximum and stable
velocities in flight.
18
At the instant of the maximum velocity, Al is the liner material that shows an immediate response to the
impacted shock wave, resulting in the highest velocity, traveled distance, compression, density change, internal
energy change, and L/D ratio due to its softness, low initial density, low melting temperature, and interatomic
structure. The greatest velocity gradient is in the central region of the EFP and decreases toward its edges. In the
stable regime, Al has the highest velocity, divergence amplitude or instability, internal energy, and density change.
Al is the liner material that is most unstable during the flight trajectory. The Ta EFP has a large diameter, which
enables it to produce a wider crater in the target. The Fe EFP has a diameter very close to that of the Ta EFP,
making it attractive due to cost and easy machining capability. The temperature generated within the Ta EFP
is the highest, which can produce hot fragments. The L/D ratio for the Cu EFP at the stable velocity is the
highest, which ensures the maximum penetration. Mild steel is the most stable liner material due to the minimum
divergence. This material may be an effective replacement of Fe in multiliner EFPs with Ta as a penetrator and Fe
as a stabilization base.
Mild steel and copper are preferable liner materials for solid EFPs, whereas tantalum, aluminum, and armco
iron are preferable materials for hollow EFPs. Keeping the other parameters in mind for this particular design,
EFPs made of mild steel and copper are preferable materials for deeper penetration.
The time necessary to reach the stable velocity is directly proportional to the liner material density. The
velocity decay rate of the liner material is inversely proportional to its density.
The authors should like to thank Mr. Aslam Hayat, Dr. Ghani Akram, and Dr. Shakeel Abbas Raufi (Al-
Technique Corporation of Pakistan) and Dr. Mike Gibson (Cranfield University) for their help in the Autodyn
simulation and experimentation.

REFERENCES

1. M. Arshad, “Design and Simulate Explosively Formed Projectiles (EFPs),” MS Thesis (Pakistan Inst. Eng.
Appl. Sci., Islamabad, Pakistan, 2004).
2. K. Weimann and A. Blache, “Explosively Formed Projectile with Tantalum Penetrator and Armco Iron Stabi-
lization Base,” in Proc. 18th Int. Symp. on Ballistics, San Antonio, November 15–19, 1999, pp. 603–608.
3. I. G. Cullis, “Experiments and Modeling in Dynamic Materials Properties: Explosively Formed Projectile
Research in a European Collaborative Forum,” in Proc. 13th Int. Symp. on Ballistics, Stockholm, June 1–3,
1992, pp. 457–464.
4. V. Sharma, P. Kishore, and S. Singh, “An Analytical Approach for Modeling EFP Formation and Estimation
of Confident Effect on Velocity,” in Proc. 16th Int. Symp. on Ballistics, San Francisco, September 23–28, 1996,
pp. 585–595.
5. W. Lanz and W. Odermatt, “Penetration Limits of Conventional Large Caliber Anti Tank Guns/Kinetic Energy
Projectiles,” in Proc. Proc. 13th Int. Symp. on Ballistics, Stockholm, June 1–3, 1992, pp. 225–233.
6. T. W. Bjerke, G. F. Silsly, and D. R. Scheffler, “Yawed Long Rod Armour Penetration,” Int. J. Impact Eng.
12, 281–292 (1992).
7. K. Weimann, “Performance of Ta, Cu and Fe EFPs Against Steel Targets,” in Proc. 15th Int. Symp. on
Ballistics, Jerusalem, May, 1995, Vol. 2, pp. 399–404.
8. F. Rondot, “Performance of Ta EFP Simulants,” in Proc. 17th Int. Symp. on Ballistics, 1998, Vol. 3, pp. 81–88.
9. F. Rondot, “Terminal Ballistics of the EFPs-A Numerical Comparative Study between Hollow and Solid Sim-
ulants,” in Proc. 19th Int. Symp. on Ballistics, Interlaken, May 7–11, 2001, pp. 455–461.
10. G. Gazeaud, “Explosively Formed Projectile: Optimization,” in Proc. 13th Int. Symp. on Ballistics, Stockholm,
June 1–3, 1992, pp. 473–479.
11. B. Janzon, N. Burman, J. Forss, et al., “EFP Modeling by Numerical Continuum Dymanics on Personal
Computers—A Comparison between PC-Dyna2D, ZEUS and Autodyn-2D,” in Proc. 13th Int. Symp. on Bal-
listics, Stockholm, June 1–3, 1992, pp. 457–464.
12. J. W. Hermann, Ammunition Development and Engineering (Directorate US Army, S. a. 07801).
13. G. Hussain and K. Sanaullah, “Simulation Studies of Explosive Formed Projectiles (EFPs),” J. Eng. Appl. Sci.
28 (2), 11–22 (2009).
14. D. Davison and A. Nordell, “Hydrocode Analysis of Acceptable Limits on Fabrication of EFP Liners,” in Proc.
15th Int. Symp. on Ballistics, Jerusalem, May 1995, pp. 235–242.
19
15. G. Wijk and R. Amiree, “Simulation of High Velocity Impact,” in Proc. 21st Int. Symp. on Ballistics, April
19–23, 2004, pp. 838–844.
16. H. E. V. Karlsson, “Computer Simulations of Shaped Charge Jet Fragmentation,” in Proc. 20th Int. Symp. on
Ballistics, September 23–27, 2002, pp. 557–564.
17. J. Bucharr, S. Rolc, and J. Pechacek, “Numerical Simulation of the Long Rod Interaction with Flying Plate,”
in Proc. 21st Int. Symp. on Ballistics, Adelaide, Australia, April 19–23, 2004, pp. 196–199.
18. Rondot F., Berner C. “Performance of Aerodynamically Optimized EFP Stimulants,” in Proc. 17th Int. Symp.
on Ballistics, Midrand, South Africa, 1998, Vol. 3, pp. 225–232.
19. S. Pappu and L. E. Murr, “Hydrocode and Micro-Structural Analysis of Explosively Formed Penetrators,”
J. Mater. Sci. 37 (2), 233–248 (2002).
20. G. Hussain and K. Sanaullah, “Gradient Valued Profiles of EFP’s Liner Materials with Modified Johnson Cook
Model,” NUST J. Eng. Appl. Sci. 2, 78–87 (2009).
21. G. Hussain and K. Sanaullah, “Computer Simulation Profiles and Velocity Decaying Rates of Explosively
Formed Projectiles,” in Proc. 25th Int. Symp. on Ballistics, Beijing, China, 2010, Vol. 1, pp. 714–727.
22. J. Macmahon, P. Church, M. Filer, et al., “Use of Hydrocode Simulations in the Design of Tantalum EFPs,”
in Proc. 18th Int. Symp. on Ballistics, San Antonio, November 15–19, 1999, pp. 520–527.
23. R. J. Almond and S. G. Murray, “Projectile Attack of Surface Scattered Munitions: Prompt Shock Finite
Element Models and Live Trials,” Propellants, Explosives, Pyrotechnics 31 (2), 83–88 (2006).

20

You might also like