Sei sulla pagina 1di 12

Int. J. Impact En~tn#Vol. 16, No. 5/6, pp.

699 710, 1995


Copyright ~) 1995 Elsevier Science Ltd
Pergamon Printed in Great Britain. All rights reserved
0734-743X(95)00005-4 0734 743x/95 $9.50+ 0.00

PENETRATION INTO DUCTILE METAL TARGETS


WITH RIGID SPHERICAL-NOSE RODS

M. J. FORRESTALt, D. Y. TZOU~, E. ASKARIt and D. B. LONGCOPE+


~'Sandia National Laboratories, Albuquerque, NM 87185-0312, U.S.A. and :~Department of
Mechanical Engineering, University of New Mexico, Albuquerque, NM 87131, U.S.A.

(Received 16 August 1994; in revised form 5 December 1994)

Summary We developed penetration equations for rigid spherical-nose rods that penetrate ductile
metal targets. The spherical cavity-expansion approximation and incompressible and compressible
elastic-perfectly plastic constitutive idealizations simplified the target analyses, so we obtained
closed-form penetration equations. We compared predictions from our models with previously
published penetration data and results from Lagrangian and Eulerian wavecodes.

INTRODUCTION
Analytical methods for penetration mechanics began with the work of Bishop et al. El]. They
developed equations for the quasi-static expansions of cylindrical and spherical cavities and
used these equations to estimate forces on conical nose punches pushed slowly into metal
targets. Later, Goodier [2] developed a model to predict the penetration depth of rigid
spheres launched into metal targets. That penetration model [2] included target inertial
effects, so Goodier approximated the target response by results from the dynamic, spherically
symmetric, cavity-expansion equations for an incompressible target material derived by Hill
[3, 4] and Hopkins [5].
In this study, we developed closed-form penetration equations for rigid, spherical-nose
rods that impact ductile metal targets. To obtain closed-form solutions, we idealized the
constitutive target description as incompressible or compressible, elastic-perfectly plastic. In
addition, we used the spherical cavity-expansion approximation I-1,2] that approximates the
two-dimensional target response with equations derived from spherically symmetric, cavity-
expansion analyses. Thus, the spherically symmetric, cavity-expansion equations are used as
an input to the penetration equations.
First, we modified the Goodier penetration model [2] to conform with recent experimental
observations [6, 7] and obtained penetration equations for an incompressible target ma-
terial. Next, we developed closed-form, spherically symmetric, cavity-expansion equations
for a compressible material and obtained penetration equations for a compressible target
material. Finally we compared predictions from our models with previously published
penetration data and results from Lagrangian and Eulerian wavecodes.

PENETRATION MODELS FOR AN INCOMPRESSIBLE TARGET


Figure 1 shows a post-test, X-ray photograph of a maraging steel, spherical-nose rod in
a 6061-T6 aluminum target [-6]. We observed from Fig. 1 that the projectile nose remained
visibly underformed and that the projectile produced a tunnel in the target about the size of
the shank diameter. We use these observations to formulate our penetration models.
As previously mentioned, we use the spherical cavity-expansion approximation to obtain
closed-form penetration equations. Thus, our procedure begins with spherically symmetric,
cavity-expansion analyses. These cavity-expansion equations are then used as input to our
penetration models.

Cavity expansion
A spherically symmetric cavity is expanded from zero initial radius to radius a in an
incompressible elastic-perfectly plastic material. As shown in Fig. 2, this expansion produces
699
700 M.J. Forrestal et al.

Fig. 1. Post-test, X-ray photograph for a spherical-nose rod with striking velocity 1120 m/s.
Penetration into ductile metal targets 701

~.,~lTY
/ PLASTIC
a b

Fig. 2. Response regions for an incompressible target material.

L "

Fig. 3. Projectile geometry.

plastic and elastic response regions. The plastic region is bounded by the radii r = a and r = b,
where r is the radial Eulerian coordinate and b is the interface position between the plastic
and elastic response regions. The radial stress at the cavity surface [2, 3, 5] is given by
[ (2E)] [ - d Z a 3(da~2~
O'r(a)= ~ l+ln -~ +ptLadt-/+ 2\at/ ] (1)

where t is time, E is Young's modulus, Y is the yield stress, and tot is density. The first term in
(1) is called the quasi-static part and the second term is called the dynamic part [2].

Penetration models
Goodier [2] obtained penetration equations for a rigid, spherical projectile. For a rigid
spherical-nose rod, we replace the mass of a sphere with the mass of a spherical-nose rod
given by
m = rra2pp(L + 2a/3) (2)
where pp is the projectile density and L and a are defined in Fig. 3. Goodier [2] used the
spherically symmetric, cavity-expansion equation (1) to approximate axial force on the
projectile nose. For the quasi-static part in (1), he took the radial cavity-expansion stress as
the normal stress on the projectile nose. For the dynamic part, he identified the radius a in (1)
with the projectile radius. In addition, at the tip of the projectile (0 = 0 in Fig. 3), he identified
the velocity da/dt and the acceleration d2a/dt 2 with Vz and dVz/dt, which are the rigid-body
projectile velocity and acceleration, respectively. Because the radial velocity and acceleration
are zero at 0 = ~/2 in Fig. 3, he multiplied the dynamic term in (1) by cos0 as a simple
representation of the expected variation at the projectile-target interface. With the above
assumptions, axial force on the nose of a rigid, spherical projectile [2] is given by

{ [ d'v lt
F z=rta 2 A + p t ~ a ~ + (3a)

A=~- l+ln ~ . (3b)


702 M.J. Forrestalet al.

For a rigid spherical-nose rod with mass given by (2) and Newton's second law,

na2pp( L 2a'~dV. 2[ A 2 dV.+ptVZ]"


+7) dt L + P'axe
(4a)
Collecting terms, we obtain

+ 3L/Za)J dt - [A + ptVZ]. (4b)

The second term in the bracket on the left of(4b) corresponds to the second term in the right
side of (4a). For steel projectiles and aluminum targets with L/2a = 5 and 10, the second term
in the bracket on the left side of (4a) gives 0.021 and 0.011, respectively. Thus, we neglect the
second term on the right side of (4a) and obtain

pp(L + 2a~ dVz


T/I--d-i- = - [A + ptV2]. (5)

Equation (5) is integrated to find the final penetration depth P for a striking velocity Vs. We
obtain
P
-I(pP'~ln 1+ . (6)
(L+2a/3) 2\ptJ
Equation (6) used the cavity-expan6ion approximation proposed by Goodier [2]. How-
ever, we propose a different method for the cavity-expansion approximation. Figure 1 shows
the target in contact with the spherical nose, so we equate the particle velocity in target at the
nose-target interface to the radial velocity produced by the rigid spherical nose. Thus, we
take
da
d--~= vz cos 0. (7)

Equation (7) is the same cavity-expansion approximation used in [6] where details are given
for the determination of force on the projectile nose. The equation for projectile motion is

(8)

Penetration depth is given by

P _ 2 (~tt) In ( 1 + 3ptV2~ (9)


(L+2a/3) 3 4A )
where A is given by (3b).

PENETRATION M O D E L S FOR COMPRESSIBLE TARGETS


In this section, we obtain spherically symmetric cavity-expansion results for a compress-
ible elastic-perfectly plastic material. We obtain numerical results from the full nonlinear
equations and then develop accurate, closed-form approximations. The closed-form com-
pressible cavity-expansion results are then used as input to our penetration models.

Cavity-expansion
A spherically symmetric cavity is expanded from zero initial radius at constant velocity 1/.
As shown in Fig. 4, this expansion produces plastic and elastic response regions. The plastic
region is bounded by the radii r = Vt and r = ct, where r is the radial Eulerian coordinate, t is
time, and c is the elastic-plastic interface velocity. The radii r = et and r = cdt bound the
elastic region, where cd isJ the elastic dilational velocity.
Penetration into ductile metal targets 703

t
Vt et Cd t

Fig. 4. Response regions for a compressible target material.

Material in the plastic region is described with a linear, pressure-volumetric strain relation
and the Mises yield criterion. Thus,
p = Kr/= K(1 - Po/P), (10a)
p = 1/3(a r + 2%); a o = cr~ (10b)
O"r -- O"0 = Y (10c)
where p is hydrostatic pressure; K is bulk modulus; ~/is volumetric strain; Po and p are
densities of the undeformed and deformed material; a r, ao, and a4 are the radial, hoop, and
meridian components of Cauchy stress taken positive in compression; and Y is the yield
stress. The elastic region has material properties given by Young's modulus E and Poisson's
ratio v, where E is related to K by E = 3K(1 - 2v).
Equations of momentum and mass conservation in Eulerian coordinates are
Oar 2(at - a0) (0v 0v'],
Or -t - p ~+v (lla)
r Or]

P ~+7 =- -~+~ (llb)

where v is radial particle velocity measured positive outward. For the plastic response region,
Eqns (10a)-(1 lb) are combined to eliminate ao, p resulting in two equations in % v
2Y :po (0v 0q,
0r q r - ( 1 - - t l ) \ 0 t + v & } (12a)
Ov2v --I //O0"r _~f)
Or -t r - K ( 1 - - r / ) ~ - ~ +v ' (12b)

(7r 2Y
r/=K 3K (12c)

We define the dimensionless variables


S = (Tr/Y, U : u/c, 8 ~- V / c , T = Y/K, J=PoV2/y (13a)
and introduce the similarity transformation
= r/ct. (13b)
With (13a, b), (12a,b) transform to
dU 2U T dS
+ (14a)
d¢ (1 + 2 T / 3 - TS) (~ - U)~-~,
dS 2 fie dS
(14b)
d~ - T(1 + 2 T / 3 - TS) (~ - U ) - ~ ,
fl=c/cp, cp2 = K / p o. (14c)
704 M.J. Forrestal et al.

The boundary condition at the cavity surface is


u ( ~ = e) = e. (15)

The radial stress and particle velocity are continuous at the elastic-plastic interface at ~ = 1
[8], so
T(1 + v)
U(~=I)=U z 3(1-2v)' (16a)

2 2(1 + v) eJ
(16b)
S(¢=1)=52=~ ~ - ~ ) 1.LTr:7_ ) ej, ]J1 / 2
• F(1 - 2 v )

where Ug and S 2 are derived from the equations that govern the elastic response region [9].
First, we solve the full nonlinear equations (14a, b) numerically. For numerical evaluation,
we put the coupled equations (14a, b) in standard forms suitable for the Runge-Kutta
method [8].
2U+2T({-U'~
dU ¢ ~ \l--qJ
d--~: fl2(~-U~ 2-1 ' (17a)
\1-~;

dS j
de fl2(¢--U~ 2 --1 ' (17b)
\l-nj

We select a value offl = c/% and solve for dU/d~, dS/d~ subject to the boundary conditions at
the elastic-plastic boundary (~ = 1) given by (16a, b). The calculations proceed from ~ = 1 to
= e. When the boundary condition at the cavity surface (~ = e) given by (15) is satisfied, we
obtain the value of e = V/c corresponding to the chosen value of fl = c/%.
Next, we obtain accurate, closed-form approximate solutions for the nonlinear equations
(14a, b) with an iterative procedure. Post-analysis calculations showed that r/ is small
compared to unity, so we take (1 - q ) = (1 +2T/3 - TS) as unity. First, set the right side of
(14a) to zero, solve for U, and evaluate the integration constant from (15). The first
approximation for particle velocity is
gO(E) : g3/~ 2. (18a)

Next, substitute (18a) into the right side of (14b), integrate, and evaluate the integration
constant with (16b). The first order approximation for radial stress is

S°({) = - 2In { + J [-2e


L-~ - 41 ( 2~ ) - 2e + ~ 1 + S 2 (1 8b)

where S 2 is given by (16b) and e is found from the boundary condition at { = 1. From (16a)
and (18a),
V ~(1 + v ) T ~ 1/3
-- c = L~I--~A " (18c)

Our penetration models use radial stress at the cavity surface ~ = e, s o w e write
2 3_~{ 4e e4 4(l+v) e } (19)
S ° ( ~ = e ) = 5 [ 1 - l n e a] + 1 -~-+~-+ 9(1-v~ 1 + [(1-2v)eJ/(1 -v)] 1/z "
Penetration into ductile metal targets 705

We point out that the solutions for U° and So are for an incompressible material in the plastic
response region and a compressible material in the elastic response region.
To include the effects of compressibility in the plastic response region, we proceed by
substituting (18b) into the right side of(14a). The second approximation to particle velocity is
(1-v)T 2T~3 2TeSln¢ t_[ (6 + 464 - 464 67]
UI(~) (1_2v)~2 } ~T-'~(fle) 2 --6-~- ~3-1-~'~ . (20al
For a compressible material in the plastic response region, the interface velocity c(fl = C/Cp)
depends on the cavity-expansion velocity V(fl6 = V/cp). We obtain the equation that relates
c and V from (20a) and the boundary condition (15). Thus,
(1 - v)T + ~ _
1 (~_-2v~ E1 -31n6]
flz _
962 66 (20b)
1- - - + 463 ----
2 2
The second approximation for radial stress is obtained by substituting (20a) into the right
side of (14b) and integrating. Thus
SX(0 = St(0 + SC(~) + B, (21a)
6 -- V)
sr(~): -21n~+JT[(~)(41n~+2+ ({~2v~3)-3\eJ
1(~-'~21
+jZT[(~)(_~+8a_64)_6 f6"~z 3 6 3

S°(0- 2- ~ 3\6J
21n{+(1-Zv)e3J

+J[-I +(6_'~2(L+46_5)_41"6"~31
,~,] ~,62
64
~ ) +.~(~)6]}2, (21C)

[ ~(1--V, 1 ] (! ~)
B, = S 2 - J r ( 2 3~z t-26 - j 2 T +262-

JTZ ( l +v ~ 2
+ ~ \l--(-L-~vvJ (21d)
where $2 is given by (16b).
Our penetration models require radial stress at the cavity surface ¢ = e, so we write
2(1 + v) 6J
Sl(e)-- [1-1n63]-t 3(1-v) l+[(1-2v)eJ/(1-v)] 1/2
- v) 5-4v'] 1 5
+ J T ( _ ~v~3 ~/~e 2 +41ne + ~ - 26

2 2 27 + 86-262-64 + ~ ] }
+J e2 e 5
JT2 ~ (l--v)
t- 21n6-~+ J ~-~+46-
2 [(1-2v)~ 3

+ \l (22a)

For V approaching zero, we obtain the quasi-static solution of Hill [4].

As=~{1 + l•n L[-(1-2v)K-]]


~ - - v ~ J ; = ~2 { 1 + 1 n [ 3 ( 1 _ v,d} (22b)
706 M.J. Forrestalet al.

Penetration models
To obtain closed-form penetration equations for the compressible models, we curve-fit
results from the compressible cavity-expansion models with
O'r(e) BsPt V2 (23)
y -As-~ y

where As and B s depend only on the target material properties. As is given by (22b), so only B s
is adjusted to fit the cavity-expansion results. We give values of B s for the cavity-expansion
models in the next section. From [7], penetration depth is given by

P (Pp/Pt) In 1 + (24)
( L + 2 a / 3 ) = Bs 2\As]\ Y /J"

CAVITY-EXPANSION N U M E R I C A L RESULTS
We present numerical results for 6061-T651 aluminum targets with v = l / 3 ,
E = K = 69.0 GPa, Y = 340 MPa, and Pt = 2710kg/m3. The yield stress Y was taken as an
approximation to the stress-strain data [7] for true strains to 0.8. Figures 5 and 6 present
results from cavity-expansion models. Figure 5 gives the elastic-plastic interface velocity
c versus the cavity-expansion velocity V. The zeroth order and first order solutions were
calculated from (18c) and (20b), respectively. For large values of V in the full nonlinear
solution, the interface velocity c approaches (K/po) ~/2, which is the interface velocity for the

16.0 .... , ......... : . . . . . . . .


I
iii ... - -
l 2.0 // .-""
0 / ""
II .."
8.0

//¢/ /

4.0 /if" -- zeroth order


~/ --- f i r s t order
¥ -- f u l l nonlinear

0 " 0 . 0 . . . . . . . I .0 .i0" 3 .0
. . . . . . . . . 4.0 5.0

(p0/Y) 1/2 V

Fig. 5. Elastic-plasticinterfacevelocityversus cavity-expansionvelocity.

40.0 . . . . . . . . . . . . . . . . •
/

- - incompressible ,' //
---- S0(~°), z e r o t h order , / ,
30.0 - -- St(~l), first order , / //,'~
• / • /
- - full nonhnear , //,~//
so(~) , /;~.(/"

20.0 " /-/'//""


°

I 0.0 ~.~

°'°'.0 i.o "2.0 3.0 4.0 5.0


(po/Y) 1/2 V

Fig. 6. Radial stress on the cavitysurfaceversus cavity-expansionvelocity.


Penetration into ductile metal targets 707

elastic-plastic, one-dimensional strain problem [10]. Figure 6 shows radial stress at the
cavity surface versus cavity-expansion velocity. The incompressible results were calculated
from (1) with a constant cavity-expansion velocity, so da/dt = Y and d2a/dt 2 = 0. The zeroth
order solution given by (18c) and (19) takes the elastic region as compressible and the plastic
region as incompressible. The first order solution given by (20b) and (22a) is an exact,
compressible solution for the elastic region and an approximate, compressible solution for
the plastic region. The S°(e ~) solution in Fig. 6 is given by (19) and (20b); that is, we use the
zeroth order equation for radial stress and the first order solution for the elastic-plastic
interface velocity. Figure 6 shows that the full nonlinear solution is bracketed closely by
S°(e ~) and S~(s~), which indicates that a reasonably accurate approximation for the interface
velocity c is required as the cavity-expansion velocity V increases.

ANALYTICAL M O D E L C O M P A R I S O N S W I T H P E N E T R A T I O N DATA
AND WAVECODE SIMULATIONS
In this section, we compare predictions from our analytical model with previously
published penetration data [6, 7] and predictions from Lagrangian [ 11, 12] and Eulerian
[13, 14] wavecodes.
Spherical-nose, maraging steel rods were launched to striking velocities Vs between
350 and 1200m/s into 6061-T651 aluminum targets [6,7]. These rods had density
pp = 8000kg/m 3, shank length L = 71.12 mm, nose radius a = 3.55 mm, and nominal mass
0.0235 kg. Figure 7 shows penetration data and predictions from our analytical models that
take the target as an incompressible and compressible material. We calculated the incom-
pressible target predictions from (3b) and (9). For the compressible target predictions, we
used the cavity-expansion results that corresponded to S°(el), s l ( e l ) , and the full nonlinear
solution. Results from these three calculations showed that penetration depth predictions
using S°(e 1) and Sl(e 1) bracketed closely the penetration depth predictions using the full
nonlinear solution. The results from S°(e ~) and S~(e ~) were within one percent of the
penetration depth predicted by using the full nonlinear, cavity-expansion results for
Vs = 1200 m/s. Thus, for our compressible-target, penetration-depth predictions, we used the
cavity-expansion predictions from S°(e ~) ((19) and (20b)) as input to (24). As previously
mentioned, we curve-fit results from the compressible cavity-expansion models with (23). In
(23), A s is given by (22b), so only B s is adjusted to fit the cavity-expansion results in Fig. 6. For
S°(e 1) in Fig. 6 we take B s = 1.04.

/
=0.0 /
CompressibleModel //J
IncompressibleModel / / /
• Data, [6] /~
/
2
• Data, [7] / J /

/ •
P
(L+2aJ3)

1
/:
0 ' I ' ' I ' [ I ' I ' I

0 200 400 600 800 1000 1200 1400

V s (m/s)

Fig. 7. Analytical model predictions neglecting sliding frictional resistance and penetration data.
708 M.J. Forrestal et al.

Figure 7 shows that predictions from the incompressible target model are in close
agreement with the penetration data. By contrast, the more detailed, compressible target
model increasingly overpredicts penetration depth as striking velocity increases. We specu-
late from previous work [6] that the results of Fig. 7 neglect a tangential stress on the
projectile nose resulting from sliding frictional resistance. In [6], the authors took photomic-
rographs of the material adjacent to the tunnel (see Fig. 1) of some targets. These photomic-
rographs showed that a 5-15#m layer normal to the tunnel surface had undergone
microstructural changes that strongly suggested a thin melted region of the aluminum target.
With this limited information, the authors [6] assumed a tangential stress on the nose o- t
proportional to the normal stress a r and took

a t = #a, (25)
where # is the sliding friction coefficient. With both normal and tangential stress components
on the projectile nose, penetration resistance increased [6, 7]. To include the effects of sliding
frictional resistance [6, 7], replace As and Bs in (24) with -4s and/3s given by

As = (1 + rc#/2)As (26a)

/~s = (1 + r~#/4)Bs. (26b)

In Fig. 8, we compare predictions from our compressible, analytical penetration model


with # = 0 and 0.10 and penetration data. With # = 0.10, model predictions agree well with
penetration data. Since we only speculate on the effect of sliding friction and have no
laboratory data for # with large sliding velocities and large normal stresses, we also compare
results of our analytical model with wavecode predictions. Figure 9 shows predictions from
our compressible, analytical model and finite-element simulations calculated by Hallquist
[11] and Chen [12]. Similarly, Fig. 10 shows predictions from our compressible, analytical
model and finite-difference simulations calculated by McGlaun et al. [13] and Silling [14].
Figures 9 and 10 show good agreement between the three prediction methods, and Fig. 8
shows good agreement with penetration data.

3-
/

Model, p = 0.0 / /
Model, P=O.I / /
/ /
• Data, [6] / / /
• Data, [7] / / /
2

P
(L+2-a/3~

//

/ / /

0 i i t i , i • i
200 400 600 800 1000 1200 1400

V s (m/s)

Fig. 8. Compressible analytical model predictions that include and neglect sliding frictional resis-
tance and penetration data.
Penetration into ductile metal targets 709

p_= 0.1
_ _ AnalyticalModel /

P
(L+2a/3)

0
0 200 400 eoo 800 1000 1200 1400

Vs (m/s)

Fig. 9. Compressible analytical model predictions and Lagrangian wavecode simulations that
include sliding frictional resistance.

~t = 0 . 0 6 /
AnalyticalModel /

(L+2a/3)

0 200 400 600 800 1000 1200 1400

V s (m/s)

Fig. 10. Compressible analytical model predictions and Eulerian wavecode simulations that include
sliding frictional resistance.

DISCUSSION AND CONCLUSIONS


We developed a closed-form, experimentally verified penetration model. The model is in
good agreement with penetration data for an assumed sliding friction coefficient of # = 0.10.
Lagrangian and Eulerian wavecode solutions are also in good agreement with our compress-
ible, analytical penetration model for sliding friction coefficients of # = 0.10 and # = 0.06,
respectively.
We also point out a similarity of our work and the computational studies for eroding rods
published by Anderson et al. [15, 16]. Radial stress at the cavity surface for an incompressible
target (1) has a quasi-static term independent of cavity-expansion velocity. However, the
corresponding first term in (22a) for a compressible target depends on e, which, in turn,
depends on the cavity expansion velocity.

Acknowledgement This work was supported by the DoD/DOE Munitions Technology Program.
710 M.J. Forrestal et al.

REFERENCES

1. R.F. Bishop, R. Hill and N. F. Mott, The theory of indentation and hardness. Proc. Roy. Soc. 57, Part 3, 147 159
(1945).
2. J. N. Goodier, On the mechanics of indentation and cratering in solid targets of strain-hardening metal by
impact of hard and soft spheres. Proc. 7th Symposium on Hypervelocity Impact III, pp. 215 259. AIAA, New
York (1965).
3. R. Hill, Atheoryofearthmovementnearadeepundergroundexplosion. MemoNo.21 48, Armament Research
Establishement, Fort Halstead, Kent, U.K. (1948).
4. R. Hill, The Mathematical Theory of Plasticity. Oxford University Press, London (1950).
5. H. G. Hopkins, Dynamic expansion of spherical cavities in metals. Progress in Solid Mechanics, Vol. 1,
Chapter III (Edited by I. N. Sneddon and R. Hill). North-Holland Publ. Co., Amsterdam, New York (1960).
6. M.J. Forrestal, K. Okajima and V. K. Luk, Penetration of 6061-T651 aluminum targets with rigid long rods.
J. Appl. Mech. 55, 755 760 (1988).
7. M.J. Forrestal, N. S. Brar and V. K. Luk, Penetration of strain-hardening targets with rigid spherical-nose rods.
J. Appl. Mech. 58, 7-10 (1991).
8. W.H. Press, B. P. Flannery, S. A. Teukolsky and W. T. Vetterling, Numerical Recipes, the Art of Scientific Com-
puting. Cambridge University Press, New York (1989).
9. M. J. Forrestal and V. K. Luk, Dynamic spherical cavity-expansion in a compressible elastic plastic solid.
J. Appl. Mech. 55, 275-279 (1988).
10. P. C. Chou and A. K. Hopkins, Dynamic Response of Materials to Intense Impulse Loading, p. 68. Air Force
Material Laboratory, Wright-Patterson AFB, Ohio, U.S.A.
11. J. O. Hallquist, LS-DYNA2D, an explicit two-dimensional hydrodynamic finite element code with interactive
rezoning and graphical display. Livermore Software Technology Corporation, Livermore, CA, U.S.A. (1990).
12. E. P. Chen, Numerical simulation of penetration of aluminum targets by spherical-nose steel rods. Theoretical
and Applied Fracture Mechanics, 22, 159-164 (1995).
13. J. M. McGlaun, S. L. Thompson and M. G. Elrick, CTH: a three-dimensional shock wave physics code. In. J.
Impact Engng 10, 351-360 (1990).
14. Private communication from S. A. Silling, Sandia National Laboratories, Albuquerque, NM 87185-0820,
U.S.A.
15. C. E. Anderson and J. D. Walker, An examination of long-rod penetration. Int. J. Impact Engng 11,481 501
(1991).
16. C. E. Anderson, J. D. Walker and G. E. Hauver, Target resistance for long-rod penetration into semi-infinite
targets. Nuclear Engng Design 138, 93-104 (1992).

Potrebbero piacerti anche