You are on page 1of 31

Pergamon

Int. J. Impact Enyno Vol. 16, No. 5/6, pp. 801-831, 1995 Copyright ~' 1995 Elsevier Science Ltd Printed in Great Britain. All rights reserved 0734-743X(95)00019-4 0734 743X/95 $9.50+ 0.00

PENETRATION OF A RIGID PROJECTILE INTO AN ELASTIC-PLASTIC TARGET OF FINITE THICKNESS


A. L. YARIN, M. B. RUBIN and I. V. ROISMAN
Faculty of Mechanical Engineering, Technion, Israel Institute of Technology, 32000 Haifa, Israel (Received 29 September 1994; in revised form 2 January 1995) Summary--This paper considers the problem of non-steady penetration of a rigid projectile into an elastic-plastic target of finite thickness. A specific blunt projectile shape in the form of an ovoid of Rankine is used because it corresponds to a reasonably simple velocity field which exactly satisfies the continuity equation and the condition of impenetrability of the projectile. The target region is subdivided into an elastic region ahead of the projectile where the strains are assumed to be small, and a rigid-plastic region near the projectile where the strains can be arbitrarily large. Using the above mentioned velocity field, the momentum equation is solved exactly in both the elastic and the rigid-plastic regions to find expressions for the pressure and stress fields. The effects of the free front and rear surfaces of the target (which is presumed not to be too thin) and the separation of the target material from the projectile are modeled approximately, and the force applied to the projectile is calculated analytically. An equation for projectile motion is obtained which is solved numerically. Also, a useful simple analytical solution for the depth of penetration or the residual velocity is developed by making additional engineering approximations. Moreover, the solution procedure presented in this paper permits a straight forward approximate generalization to accommodate a projectile with arbitrary shaped tip. Theoretical predictions are compared with numerous experimental data on normal penetration in metal targets, and the agreement of the theory with experiments is good even though no empirical parameters are used. Also, simulations for conical and hemispherical tip shapes indicate that the exact shape of the projectile tip does not significantly influence the prediction of integral quantities like penetration depth and residual velocity.

1. INTRODUCTION A comprehensive review [ 1] of the state-of-the-art of theory and experiments on penetration problems documents scientific interest from the beginning of the 19th century up until 1978. More recently, a large number of the experimental data was collected in [2]. Several analytical approaches to the penetration problem have been proposed and will be discussed presently. The hydrodynamic approach neglects the yield strength of both the target and the projectile materials (or only the target material with the projectile being rigid) and is valid for relatively high impact velocities. Under these conditions both the target and projectile can be modeled as inviscid fluids with inertial resistance to deformation. This allows one to develop a hydrodynamic theory of penetration of the type considered in [3-5]. The modified hydrodynamic approach considers a rigid (or rigid-plastic) projectile penetrating a rigid-plastic target which is modeled in a very simple one-dimensional manner by adding plastic terms in the Bernoulli equation [6-8] proportional to the yield strengths of the target and projectile. Another approach used to solve the problem of penetration of a rigid projectile into a rigid-plastic target is based on postulating a general form of the force applied to the projectile with empirical constants [1, 9]. For example, this force is assumed to consist of a Bernoulli-like term which models inertial effects and is quadratic in velocity, and a term which is independent of velocity and models the effect of perfect-plasticity yielding the Poncelet equation. The solution of the equation of motion of the projectile with the force given by the Poncelet expression yields the familiar formula for the penetration depth P [1] P = c 1In(1 +

c2U~),

(1)

where U o is the impact velocity and cl, c 2 are two constants. This approach allows one to consider relatively complicated phenomena such as oblique penetration. However, it involves two empirical constants which need to be measured for each combination of target and projectile.
801

802

A.L. Yarin et al.

Another approach also considers a rigid projectile penetrating a rigid-plastic target but it employs a one-dimensional calculation of forces applied to the projectile by a work/energy balance. Some examples of this one-dimensional force calculation are given in [10] where a combined momentum and energy balance is employed for a plugging mode of perforation, and in [11-13] where a generalized equation for the force is proposed. Another approach presumes that the normal stresses applied to the projectile surface may be approximated by the solutions of the one-dimensional problems of expansion of spherical or cylindrical cavities [ 14, 15]. Then by integrating the surface traction one can determine the resultant force applied to the projectile, which is used in solving the equation of motion [ 15]. Here it should be emphasized that this approach is based on the recognition that the shear stresses applied to the projectile surface are negligible relative to the normal stresses. The next group of analytical models employs a two-dimensional calculation of the forces imposed on a rigid (or deformable, eroding) projectile by a rigid-plastic target. In this approach hydrodynamics of an inviscid fluid is used to propose an approximate velocity field in the target. The velocity field is constructed separately in several target domains, with each of them being essentially one-dimensional. Moreover, in each domain the velocity field is presumed to be irrotational and hence derivable from a velocity potential. The composite two-dimensional velocity field is then obtained by matching the one-dimensional fields using certain geometrical constraints. The origins of this latter approach date back to [16-18]. This method was successfully exploited using an integral work-rate balance to predict penetration: of a rigid blunt cylindrical projectile in [18]; of a deformable compressible projectile in [19]; and of a deformable eroding projectile in [20]. Furthermore, this approach allows one to drastically improve the characterization of the velocity field in the plastic target [21-23]. Similar ideas were also proposed in [24] to calculate the velocity and corresponding stress fields in a rigid-plastic material for the cases of steady penetration of rigid tools with circular or spherical tip shapes. Penetration of a rigid projectile into an elastic-plastic target has also been considered using numerical methods to solve the partial differential equations of momentum in the target material. For example the work in [25, 26] uses the finite element method to solve for steady-state penetration. In principle, using numerical wave codes it is possible to consider non-steady-state penetration of a deformable projectile into a deformable target. However, the three-dimensional nature of oblique penetration still poses an extremely difficult numerical problem. The idea that under certain conditions the flow of an elastic-viscoplastic metal can be approximated by the flow of a viscoelastic liquid has been reexamined recently in [27]. In particular, it has been shown that the tensorial structure of constitutive equations for a class ofelastic-viscoplastic metals is identical to those for a class of viscoelastic liquids. Also, it has been shown [27] that for a small value of the Deborah number (which is characteristic of penetration), the constitutive equation for an elastic-viscoplastic metal may be reduced to that characterizing a rigid-viscoplastic or rigid-plastic material. Consequently, it is not surprising that hydrodynamic-type approximations of the velocity field in the target are successful in the analysis of penetration problems. The main objective of the present paper is to develop an approximate solution of the problem of non-steady normal penetration of a rigid projectile into an elastic-plastic target of finite thickness. The calculation is based on a fully two-dimensional velocity field for the whole target domain which represents a natural extension of the approach proposed in [18-22]. The target material is presumed to be thick enough that the effect of deformation of the free surface can be neglected in the prediction of integral quantities like penetration depth and residual velocity. Specifically, we consider a cylindrical polar coordinate system with coordinates {r, 0, z}, base vectors {er, e0, ez}, and a fixed origin at the initial position of the front surface of a target with thickness H (see Fig. 1). The tip of the projectile is located by z = x(t). In the following analysis we derive an equation of motion of the form [M + A(x, ~'~)].~= B ( x , k ) Y c 2 + C ( x , k ) , (2)

Penetration of a rigid projectile into an elastic-plastic target of finite thickness

803

iI
::::::::::::::::::::::::::::::

er

U0 e~

< r 2R

R/2

Fig. 1. Definition of the coordinate system.

where M is the constant mass of the projectile and the functions A(x, ~), B(x, ~), and C(x, Y~) are determined analytically. Here, a superposed dot denotes ordinary time differentiation, whereas later it will be used to denote material time differentiation. The functions A, B, and C correspond to what is commonly called the virtual mass of the target, inertial effects (quadratic in velocity), and plastic drag, respectively, and the analytical expressions for these functions are given by (53a-c). The Eqn (2) is solved subject to the initial conditions x(0) = 0, ~(0) = - Uo, (3a,b)

where Uo ( > 0) is the initial velocity of the projectile. In this solution x(t) is a monotonically decreasing function and the Eqn (2) ceases to be valid when ~ vanishes. As in [18-22], the velocity field in the target is taken to be irrotational and hence is derivable from a velocity potential. This allows us to consider a specific blunt projectile shape which corresponds to a reasonably simple velocity field that exactly satisfies the equation of continuity and the condition of impenetrability of the projectile. The target material is assumed to be incompressible and the target region is divided into an elastic region ahead of the projectile where the strains remain infinitesimal, and a rigid-plastic region near the projectile where the strains can be arbitrarily large. Furthermore, the rigid-plastic response is taken to be rate-independent. Using this velocity field, we include intertial effects in Section 2 and solve the momentum equation exactly in both the elastic and rigid-plastic regions to determine an expression for the pressure field. Then, the effects of the free front and rear surfaces of the target are modeled in an approximate manner, and the decelerating force applied to the projectile by the target is calculated in Section 3 by analytically integrating over the relevant portion of the projectile surface. The resulting equation of motion of the projectile takes the form of (2). This equation was solved numerically in Section 5 and the theoretical predictions were compared with numerous experiments. We emphasize that the agreement is good even though the formulation has no empirical parameters. A simplified approximate solution of (2) given in Section 4 yields the following analytical expression for the penetration depth P

P,~

67rYR~ J [

where p, Y and p are the density, the yield strength in uniaxial tension and the shear modulus of the target; R~ is the radius of the projectile, and U o is its impact velocity. For the case of target perforation the residual velocity of the projectile U r is also calculated analytically within the framework of the simplified approximate solution of Section 4. Moreover, in the region of validity of this approximate solution, P as given by (4) and U, are in good agreement with both the results of the numerical solution and with numerous experimental data.

804

A. L. Yarin et al. Table 1. A brief summary of various parameters and the equations where they are defined Parameter fl f2 f3 f4 /4 r* 3 ~* 3 /~3 /~4 /~5 /~6 /~L 2 Equation (39b) (44a) (39c) Parameter x* ,~* x** :~** x*** f*** 0b ~ Equation (31b) (31 a) (34) (60b) (57) (57) (68) ( 11b) (27b) (29b) (40) (30b) (36) (24f) (27c) (29c)

(44b)
(39d) (44c) (39e) (44d) (64c) (42a) (42b) (41 a,b) (48) (45a) (47) (46a) (30d) (30c)

$~
~-2 ~2 ~ ~-1 ~2

Furthermore, the present approach is generalized in Section 6 for analyzing projectiles with general tip shapes. This is accomplished by using the method of singularities to calculate the velocity field. Simulations for conical and hemispherical tip shapes indicate that the exact shape of the projectile tip does not significantly influence the prediction of integral quantities like penetration depth and residual velocity. Section 2 describes the analysis of the target. In order to aid the reader, Table 1 provides a brief summary of various parameters and the equations where they are defined. Section 3 develops the equation of motion for the projectile, and Section 4 describes an engineering approximation which simplifies this equation and yields an analytical expression for the penetration depth or the residual velocity. Section 5 presents a number of examples which show good comparison of the theoretical predictions with experimental data. Section 6 briefly describes a generalization for the case of a projectile with general tip shape and discusses simulations of projectiles with conical and hemispherical shaped tips. Finally, Section 7 discusses the main conclusions of this work. The most important results obtained in the present work are given by equations (49) supplemented by (52) and (53) as well as by (27), (30), (31), (36), (39), (40), (42), (44) and (48) for numerical calculation of the projectile motion. Also, an approximate analysis yields the simplified analytical expressions (4), (62)-(66) and (68). 2. ANALYSIS OF THE TARGET In the present section a blunt projectile in the form of an ovoid of Rankine is considered and the associated velocity field is used to calculate the stress field in the target. Specifically, the target material is assumed to have a constant density p and to remain incompressible. This means that the conservation of mass and the balance of momentum equations may be written in the forms divv = 0, p~, = - Vp + dive', (5a, b)

where the gradient operator V and the divergence operator div are defined with respect to the present position of a material point, v is the absolute velocity of the material point in the target, and the Cauchy stress a has been separated into a pressure p and its deviatoric part ty',

Penetration of a rigid projectile into an elastic-plastic target of finite thickness

805

such that ~ = - p l +o', a'.l =0. (6a, b) Also, I denotes the unit tensor, the notation a. b denotes the usual scalar product between two vectors a, b, and the notation A" B = tr(AB T) denotes the inner product between two second order tensors A, B. In the following solution we determine a velocity field that exactly satisfies the continuity Eqn (5a) and the condition of impenetrability at the surface of the projectile. At the instant that the projectile touches the target's front surface the target material begins to deform elastically. However, at some point during the penetration process the target material begins to deform plastically and an elastic-plastic boundary propagates away from the tip of the projectile. In the elastic region the strains remain small whereas in the plastic region the strains can be very large. Furthermore, for simplicity we assume that the material response in the plastic region is rate-insensitive and rigid-plastic so that the deviatoric stress may be approximated by a' =
( D " D ) 1/2

D,

(7)

where Y is the constant yield stress in uniaxial tension and D is the symmetric part of the velocity gradient. Also, in the elastic region we assume the material response is linear elastic and isotropic so that for isochoric motion the deviatoric stress is related to the linear strains by the expression o ' = 2/~, (8)

where/~ is the constant shear modulus. Here, we will develop expressions for the pressure in both the plastic and elastic regions that exactly satisfy the equation of motion (5b). Although the equations of motion are satisfied exactly, the boundary conditions associated with the free front and rear surfaces of the target and the free surface that develops near the projectile will be satisfied approximately. In order to impose these boundary conditions and to determine the location of the elastic-plastic boundary it is convenient to divide the region into four parts separated by the boundaries z = :z 1 = - H associated with the target's rear surface; z 2 associated with the elastic-plastic boundary; z 3 associated with the beginning of the region of contact between the target and the projectile; and z4 associated with the separation line of the target material from the projectile surface (see Fig. 2).
10.0

Elastic-Plastic Boundary
5.0-

Front Surface Rear __~ Surface

- " ~ ZI
., Pr I

~0.0-

. _ . ~ ~...

-5.0.
Z -'--H
1

-10.0 -15.0

z4

-10.0

-5.0

0.0

5.0

Fig. 2. Definition of the axial location of: the target's rear surface (z 0 and front surface (z = 0); the elastic-plastic boundary (z2); the projectile tip (z3); and the separation line of the target material from the projectile surface (z4).

806
Velocity field and projectile shape

A.L. Yarinet al.

It is well known that for axisymmetric irrotational motion, the velocity field is determined by a potential function ~b(r,z, t) such that
v = V~b. (9)

Then, the continuity Eqn (5a) and the balance of momentum (5b) reduce to
V 2 ( ~ = O,

V [ p {~-~t+ 1V~b.V~b} + p ] = div 6'.

(10a, b)

Here, we take the projectile to be an ovoid of Rankine because the associated velocity field which satisfies the condition of impenetrability is known to be characterized by a combination of a single source and uniform flow. Specifically, the ovoid is a body of revolution of length L whose lateral surface is defined by r = R [28-30] where -R2+~ -2)1/2+1
~=z-x-

=0,

for2

<~<-

+L,

(|la) (llb)

In (11) Ro~ is a constant controlling the radius of the projectile and ~ is measured relative to a material point in the projectile (see Figs 1 and 3). It can be shown that the surface defined by (1 la) can be expressed in the equivalent parametric forms
2R 2 _ R 2

= E(R) = 2(R2 _ RE)l/2 , ~ r=R(0= ~+ (42+2R2) u2

(12a)

Now, the velocity potential ~b,the radial component vr and axial component v=of the velocity field, which satisfy the continuity Eqn (10a), and the condition that (lla) [or 02)] is a material surface (i.e. the projectile is impenetrable) are given by = +x (~2 + r2)1/2 , (13a)

.F

(13b,c)

1.5

, , , I , , , , I , , , ,

1.00.50.0-0.5-

-1

'ii1
-1.0

i,i

i i i i 1 , , ~ 1

0.0 ~/R

1.0

2.0

Fig. 3. Actualscaledshapeof the projectiletip.

Penetrationof a rigidprojectileinto an elastic-plastictarget of finitethickness

807

Before proceeding further it is worth discussing the motivation for using the above velocity field. In this regard, it is noted that the perforation process of a plate is divided into several stages including: dynamic penetration, bulge formation, bulge advancement, plug formation, and exit [18]. For relatively thick ductile metal plates, which are of interest in the present study, the dominating stage of longest duration is the dynamic penetration stage. Consequently, it is unlikely that the other terminal stages (with a cumulative duration shorter than that of the penetration stage) can significantly effect the prediction of integral quantities like penetration depth and residual velocity. Moreover, the general pattern of the velocity field during the stages of bulge formation and advancement does not differ significantly from the velocity field during dynamic penetration even though different formulae are used for their descriptions [18]. Also, note that for sufficiently thick targets the dynamic penetration stage which is modeled here is the only stage of the penetration process. Simple geometrical reasoning demands that the velocity field should satisfy the condition of impenetrability at the projectile surface and degenerate towards a uniform flow far from it. Consequently, the fully two-dimensional velocity field (13b,c), which satisfies these two conditions, is expected to adequately model the kinematics of the target material during the dynamic penetration stage of a relatively thick target. The potential (irrotational) character of the velocity field (13b,c), like the potential character of all other velocity fields used in the literature [15, 18-23], does not predict the detailed effects of shearing near the projectile surface. Nevertheless, predictions of integral quantities like penetration depth and residual velocity using this velocity field are expected to be good because the decelerating force of the target is dominated by the normal stresses applied to the projectile surface and not by the shear stresses. This fact was clearly recognized in the classical works [14, 16, 31] and is continually exploited to construct potential velocity fields appropriate for analytical models of the penetration process [15-20, 32, 33]. All of these cited velocity fields are based on the models of expansion of cylindrical or spherical cavities, which model the main kinematics responsible for the development of normal stresses. Moreover, these cited velocity fields do not model details of the slip process near the surface of the projectile which is responsible for the development of shear stress there.

The rigid-plastic region of the target Using the velocity field (13b,c) it can then be shown that the rate of deformation tensor D becomes D=
-

--

4(~2 -{-

r2)S/2 [(~2 _ 2r2)e, e, + (~2 + r2)e0 eo + (r 2 _ 2~2)ez@ ez


(14)

3r~(er e z + ez er)],

where the symbol denotes the usual tensor product. Thus, recalling that is non-positive it can also be shown that whenever the projectile is moving (~ < 0), the deviatoric stress (7) corresponding to the velocity field (13b,c) may be written in the form O'
-

3(~2 ~_r2 ) [(~2--2r2)e,e,+(~Z+rZ)eoeo+(r2--2~2)ezez 3r~(e, ez + e, @ e,)]. (15)

Furthermore, it can be shown using (15) that diva'=V-Ylnk R2 ]j. (16)

This simple result indicates that the balance of momentum (10b) can be solved exactly in the

808

A.L. Yarin et

al.

rigid-plastic region to obtain an expression for the pressure of the form

p = f(t) + p22 [

R2~ 4( 2 + r2)3/2 32(~ 2 + rZ)2J


R~

- p:~R~ [4(~2 ~r2),/2]

Yln{~2+r2)
\ R~

/'

(17)

where f(t) is a function of time only which must be determined by a boundary condition. In order to determine the stress acting on the surface of the projectile, we note that the unit normal n (outward normal from the projectile), the unit tangent vector r, and the element of area da may be expressed in the forms

n=

/~(3 - 2/~2)e, - 2(1

R2)3/2e

(;4 -- 3/~2) 1/2


2(1 -/~2)3/2e, +/~(3 - 2/~2)ez (4 - 3/~2) 1/z

(18a)

(18b) (18c)

Ly6--

j R dRdO,

where/~ is the normalized value of the radius (12b) /~ = - - . R~


R

(19)

Now, using the expressions (6a), (11), (17), (18) and (19) we may evaluate (15) on the surface of the projectile and express the resulting stress in terms of the vectors {n, e0, z} to deduce that
= o.,n n +

aooeoeo+

a~

r + o.,(n + ~n),

(20a) (20b,c) (20d,e)

y [ - ( - 8 + 9/~2)]

Y
%0 = - P + ~-,

-p+

L r[

j,

[- -- 211~(!"~ J~2)1/2 ]
cr = Y (4 - 3/~ z) J'

p= f(t)+p22I(1- 3R2)(1 - R2)] [(1 -/~2)'/2-[ 2 - pY~R~ ~J + rln[4(1 -/~2)3.

(20f)

Since we have not imposed any boundary condition on the slip velocity or the shear stress at the projectile surface we do not expect that the value of the shear stress a,~ predicted by (20) will be very accurate. In fact, it can be seen from (20) that since a.~ is non-positive, the shear stress applied to the projectile surface acts in the negative z direction, which seems physically incorrect because it tends to accelerate the projectile instead of decelerate it. On the other hand, the diagonal components a.,, %0, o~ of stress are predominantly determined by the i~inematics of the flow past the projectile. Consequently, since we have satisfied the condition of impenetrability of the projectile and the balance laws (10) exactly, we expect these diagonal components of stress to be reasonably accurate (in conformity with the ideas of [14, 16, 31]). For these reasons, we will ignore the shear stress component a,~ in the subsequent analysis. Once an appropriate boundary condition is imposed to determine the function f(t) in the expression for the pressure (17) and (20f), the value of the stress on the boundary of the projectile (20) will be known. This means that the axial force F which decelerates the projectile can be determined by integrating the appropriate expression for the traction vector applied to the projectile surface.

Penetration of a rigid projectile into an elastic-plastic target of finite thickness

809

Elastic region of the target


Next, we determine a solution in the elastic region of the target where the strains remain small. The velocity field (13) is also assumed to approximate the motion in the elastic region because it has the main physical features that the velocity vanishes as r >>R~ or z -~ - ~ . Also, it will be shown that the results based on this assumption are consistent with results of the punch problem [31]. Using this assumption the infinitesimal strain e can be determined by integrating the equations
-

Og(r, z, t) -D. gt

(21)

where the strain e admits the representations

e = g(r, z, t) = ~.(r, ((z, x(t)) = Cz(r, z - x(t) - R~),


2

(22)

and the partial derivative with respect to time is used instead of the material derivative because the deformations are assumed to be infinitesimal in the elastic region. Noting that
Og(r, z, t)
-

. O~(r, ~)
= x

Ot

g~

'

(23)

we may use (14), (21) and integrate (23) subject to the initial condition that ~ vanishes at the initial time of impact when x(0) vanishes. This procedure yields the expressions

= er~e,e, + eooeoeo + e=e~ e~ + e~(e, ez + e~ e,), R2[ ~(~2+2r 2) ( ( ( 2 + 2 r 2 ) ] e,, = ~ (~2 + r2)3/2 -} ((2 + rZ)a/2j,
CO0= ~ r 2 (~2 +r2)1/2 ((2

(24a) (24b)

+r2)1/2

(24c)

~ =-~- (~2 + r2)3/2


=R~r~
er~
4

((2 ~_(-r2)3/2 ],
1 ]
((2 q_-r2)3/2 ,

(24d)

1
L(~ 2 q.- r2) 3/2

(24e) (24f)

( = z - -2'

RoC

where ~ is given (1 lb). Now, using the expression (8) for the stress and noting that divD vanishes, it follows that in the elastic region div ~' = 2p dive = 0, so that the equation of motion (10b) yields an expression for the pressure of the form (25)

P=g(t)+P2214(~.2+r2)3/2

32(~2+r2) 2

4(~2~r2) 1/2- '

(26)

where 9(t) is another function of time to be determined in the elastic region.

Boundary and matching conditions


Although we cannot satisfy the boundary condition that the rear surface of the target is stress free pointwise using the assumed velocity field (13b,c), we can approximate this condition by requiring the stress t~= to vanish at r -- 0 and z = z 1 = - H. Thus, it follows from

810

A . L . Y a r i n et al.

(8), (24), and (26) that for the elastic solution 1


9(t) = - p x 2
2-

4~ 1

32(~
x R~,

- p~R~o

H 1 2

(27a)

H R oo

1 2' ~-

~-

R~

(27b,c)

Given this value of 0(t ), the stress in the elastic region is determined by Eqns (6.), (8), (24), and (26). This approximation of the boundary condition at the rear target surface is not expected to be too restrictive because the stress oz: at the rear target surface remains relatively small for r > 0 until the projectile comes close to the rear surface. Next, we determine the function f ( t ) in (17) by a matching condition at the elastic- plastic boundary. To this end, we note that the elastic-plastic boundary is determined by the values of r and z which cause the elastic solution to satisfy the yield condition that
G(r,z,x) = [~,~'~] 3 ' '
1/2

- Y=#[6~'~.] ti2- Y=0.

(28)

For an exact solution it is necessary to require continuity of surface tractions pointwise on t h e elastic-plastic boundary. However, for the approximate velocity field we only require continuity of the axial stess G= at the intersection z = z 2 of the elastic-plastic boundary with the r = 0 axis. Thus, for a given value of x(t) we can evaluate (28) at r = 0 and z = z 2 to obtain an equation of the form [~z2 - 1
4r -~,
_ z2 1 Y x

(29a)

z2

~-2

2'

(29b,c) (29d)

for - H ~<z2 ~<x ~<0.

Alternatively, we can substitute (29c) into (29a) and solve the result for x to deduce that
/ ~2 \1/2

/ 4 Y ' ~ '/2 ~-

f 4 Y " l '/2

where the normalized variables (30b,c,d) have been introduced to simplify Eqn (30a). This approximation of the boundary condition at the elastic-plastic boundary is not expected to be too severe because the shape of the elastic-plastic boundary ahead of the projectile is relatively flat in the range r ~< R~o [(see Fig. 13(c)]. Moreover, these types of approximations are common in the solution of plasticity problems using approximate velocity fields (e.g. see the penetration problem in [34]). Yielding first occurs at the tip of the projectile (z = x, ~-2 = - 1/2) when ~ takes the value Y* (and x takes the value x*) given by

2\3,)

1- 3 . ) \4Y)

Z~I/2

~--~\-~)

(31a)

Ro~

16\3/i)'

where we have used the fact that for most metals Y / # is less than about 0.01. Thus, for .,7 ~<.f* the elastic-plastic boundary occurs ahead of the projectile. Alternatively, we may use (29c)

Penetration of a rigid projectile into an elastic-plastic target of finite thickness

811

and (30) to deduce that

2+x= .

.v .

\3.)

~ ,

,)

(32)

so that x can be written as a function of z z of the form

--R-~

1-

(4y,](l_z2,]2

112

(33)

R~

1 + \3/x)\2

RooJ

It follows that the elastic-plastic boundary reaches the rear target surface when z 2 = - H and x = x**, where

j> J
Notice also from (30a), that ~ approaches negative infinity when ~2 approaches negative unity. Thus, for large values of [xl the solution (30a) yields

~z ~ - 1 ,

~2=-\~j

(35a,b)

For the general case, the value of ~z associated with the elastic-plastic boundary can be determined analytically as a function of~ by rewriting (30a) as a quartic equation of the form
~2 + 2~23 _~_3~2~'2~2 2X~2 - ~2 = 0. 4 -

(36)

Using Descarte's rule of signs r35] and the fact that ~ is negative it can be shown that (36) has only one real negative root, and that this root corresponds to the elastic-plastic boundary. Furthermore, using the expressions (8), (24), (26), (27), and (29), the axial stress in the elastic region at the elastic-plastic boundary may be expressed in the form

az=(O'z2't)=--P2

4~

4~

32~-24+
9

-pS~R~

4? 2

4~ 1 (37)

-?

On the other hand, we can use (6), (15), and (17) to evaluate the axial stress in the plastic region at the elastic-plastic boundary to obtain

a==(O,z2, t ) = - f ( t ) - p 2

1 2 2 [4~ 3~'~-24}-P'fR~[~z}+ Y[ln~2-3]"

(38,

Since our matching condition requires the axial stress to be continuous at the elastic plastic boundary we may equate (37) and (38) to obtain f(t) in the form

f(t) = pYR~ f l(t) - p22 f z(t) + Y f 3(t) - p f 4(t),


f,(t)= 1 4~ 1' f2(t)= [ 1 -2 4~ 3 2_1_141 '

(39a) (39b,c) (39d,e)

f3(t) = In ~~,

/4(t) = ~ [-~ - ~1.

In view of the result (31) it may be seen that for .-7" < ,-7~ 0 the target material on the r = 0 axis remains elastic. However, since for most metals Y/'p is less than about 0.01 it follows from (30) and (3l) that Ix*/R=l is less than about 0.001. For this reason we assume that the elastic-plastic boundary is located at the tip of the projectile until .-7is less than .-7* so that f(t)

812 is determined by (39) with

A. L, Yarin et al.

for

<

(40)

This causes a negligible error for very small values of Ix]. Now, for ~ less than ~* and greater than some critical value, the quantity ~'2 is determined by the solution of(36). This solution is valid until ~-2 equals the value ~-1 and the elastic-plastic boundary reaches the rear surface of the target. Once this occurs we can determine f ( t ) directly by requiring the axial stress azz in (38) at r = 0 to vanish when ~-2 is replaced by ~-1. However, the results of simulations suggest that it is a better approximation to require the average axial stress on the rear surface in the plastic region to vanish. Depending on the thickness of the target and the value of the impact velocity U o, the tip (z = x) of the projectile may reach the rear surface of the target. If we neglect the motion (bulging) of the target's rear surface (which is practically insignificant for relatively thick targets) then this occurs when x = - H. Furthermore, if the penetration process continues then part of the projectile tip protrudes from the target. To model this process it is convenient to define the radius R 3 and the normalized radius/~a of the projectile at the first point of contact of the projectile with the target material by the formulae R3 I1~3~---R-~ = 0 R3 /~3 ----~-~ ----I ~ for -- H ~<x, forx< -H, (41a) (41b)

~ 11/2 + - (~-2 + 2 ) 1 / 2

where use has been made of the expression (12b) in obtaining (41b). Also, we denote the location of the intersection of the elastic-plastic boundary with the rear surface (z = - H) by the radius r* and the normalized radius ~ , where r* is determined by using (28) and solving
G(r~, - H, x) = O, ~ = r~/R~.

(42a,b)

Then the condition of vanishing average axial stress on the rear surface in the plastic region requires
r3 - R 3

,2

a=(r, - H, t)r dr = 0.

(43)

Using Eqns (6), (15), and (17) for the stress in the plastic region, the condition (43) yields an expression for f ( t ) in the form (39a) with
fa(t) - 2(~32 _ 1~2) [(~2 + .q2)1/2 _ (~1 ~- ~2)1/2],

(44a) 1 ] 16(~ q-/~2) ,

1
f2(t) -- 2 (.--,2- 62, "3 r'3)

~-1 ( ~ + r--,2/,/2 "3 ,

~-1 1 ( ~ +/~2)1/2 I- 16(~ + r~ 2)

(44b) (44c) (44d)

2 1 -*2 ~-2 .--.2 -f3(t) = - ~ + (~2 -- R3) Era ln(~a + r 3 ) - R2 In (~2 +/~2)], -2 3
f4(t) = 0.

Furthermore, it can be shown using the expressions (39) and (43) that fx, f2, and ( Yf3 -/~f4) are continuous functions as the elastic-plastic boundary approaches the rear surface.
Separation line

In order to determine the position of the line where the target material separates from the projectile surface we first determine the normalized radius/~5 of the projectile at the target's front surface 11~5= L - - ~~5 + ~-tt ~5 + 2) 1/2 ],/2, [-1 ' ~-5 ~-5 = R--~ X 7.' 1 (45a,b)

Penetration of a rigid projectile into an elastic-plastic target of finite thickness

813

the normalized radius/~L at the end of the projectile

J~L =

[ --~

~- (~'~L+

2) '/2

11/2'~-L-- R~oc 2'


L 1
-

(46a,b)

and the normalized radius/~6 of the first point (smallest value of/~6) at which the normal stress a., in (20b) vanishes ""(/~6) = - f ( t ) - p x 2 [
(1 - 3/~2)(1 ) ] 2 -/~
+

+L +

I pR~(B~Z

K2)l/2q
~- 1 - Yln [4(1 -/~2)] (47)

(-M+/I~ C)] [ (1

[_ 3 ~ - ~ J = O '

where use has been made of the equation of motion (2) to eliminate the term ~ in terms of 2 and the functions A, B, C which will be determined later [see (53)]. Then the separation line of the target material from the projectile surface is determined by the normalized radius/~4 defined by R4 = Min [R 5, R L, R6]. (48) 3. EQUATION OF MOTION OF THE PROJECTILE The equation of motion of the projectile in the axial direction is obtained from rigid body dynamics and may be written in the form MY = F, M= 2 (1 -- ~2~3/2_(1 -- "'L, + (1 ~z~1/2 'L,
-

(4%) -2 RL)1/2
'

(49b)

where M is the mass of the projectile, pp is its density, and F is the resultant axial force acting in the positive ez direction. This force can be calculated by integrating the traction vector t applied by the target to the projectile, which is given by
t = ~ n = a . . ( / ~ ) n + a.~(/~R. (50)

In view of the discussion in the last section, we neglect the effect ofa,, on the value ofF. Also, we recall that the normal stress a.. is compressive over the region/~3 ~</~ ~</~4 where/~3 and /~4 are given by (41) and (48), respectively. Thus, F= t'%da~ -R 2
~,.(R)RdRdO, do JR3

(51a) (51b)

F = - 2~R 2

f?3

a.,(R)RdR.

Substitution of the expression (19), (20) and (39a) into (51b) yields the result that
F = - A(x, 2)2 + B(x, 2)22 + C(x, 2),

(52a)

A(x,2)=2gpR~Jg~ C (1- ~2)1/2 fR'[ 2


B(x, 2) = 2~pR 2

fl(t)

] RdR,
fz(t) R dR,

(52b)

;?[

(1 - 3/?,2)(1 -/~2) 2

(52c)

C ( x , 2 ) = 2 r c Y R 2 f k " I ln{4(1 -/~2)} q (8 -- 9/~2) ~-f3(t) 1 R d R


J R j L_

3(4

3/~ 2)

-- 2~l~R 2

f?3

f 4(t)R dR.

(52d)

814

A.L. Yarin et

al.

In the above we have assumed that limits of integration/~3 and/~4 may depend on x and ~ so the functions A, B, C also depend on x and ~. Furthermore, the integrals in (52) may be evaluated analytically to obtain
h = r~pR 3 []{(11 __ /~2)3/23 . (1 . 1~2).3/2} . f a ( t ) ( R ]

/~2)],

(53a) (53b)

B :- ~ p R ~ [/~.(1 --/~2)2 _ /~(1 _ R3 ) _ f2(t)(/~24 _/~2)], --22

C = rt y R 2 [(/~2 -/~2) In ( 4 ) - (1 _/~2) in ( 1 _/~42) + (1 _ R3)ln ( _ R3 1-2 )-2 [ ~ J + f3(t)(/~2 _ R3 ) _ 7 z # R 2 f a ( t ) ( ~ 2 _ ~2). (53c)

Note that A is often referred to as the virtual mass of the target material associated with the projectile. Given values for # and Y, the critical value 2" is determined by (31). Then, when Y* < ff ~<0, the functions f l , f2, f3, and f4 are given by (39b-e) with ~2 specified by (40); whereas when ff ~<if*, the functions fa, f2, f3, and f4 are given by (39b-e) with ~-2 specified by the solution of (30b) and (36) until the elastic-plastic boundary reaches the rear surface with ~-2 = ~-1 [with ~-1 specified by (27b)]. When :~ ~<2" and the elastic-plastic boundary has reached the rear surface so that r~ determined by (42) is greater than zero, the functions f l , f2,f3, and f4 are given by (44a-d) with/~3 specified by (41) and/~4 specified by (48). 4. AN E N G I N E E R I N G A P P R O X I M A T I O N The equation of motion of the projectile (49), supplemented by the expressions for the decelerating force (52a) and (53), may be integrated for the general case only numerically. However, for some important particular cases, reasonable approximations can be introduced which simplify the equation of motion to a form that can be integrated analytically. This approximate analysis developed below yields analytical expressions for the penetration depth when the target is not perforated and the residual velocity when the target is perforated. The analysis of this paper assumes that the projectile remains rigid. Consequently, for a metal projectile penetrating a metal target this assumption limits the impact velocity to be less than about 1.5 km/s. The results of preliminary simulations using impact velocities in this range indicated that the resisting force F remains reasonably constant when a projectile penetrates a semi-infinite target. For this situation it is a good approximation to set/~3 equal to zero and/~4 equal to unity in (53). Furthermore, for a semi-infinite target, ~-1 and (-1 in (39) approach negative infinity, so that the expressions (53) reduce to

1 3 A = A~ = ]ltpR~,

B = 0,

(54a,b) (54c)

C = rcYRo~[~ ln(4) + in ~-2]. 2 5

In general, the value of ~-2as a function of x can be determined by using (30b) and solving (36). Alternatively, we can approximate the solution of(30a) by noting that for small values ofl ~2[ (30a) yields 2 z. t55)

Furthermore, since the exact solution of (30a) requires ~2 to approach negative unity when approaches negative infinity, this suggests the approximation 2:~ ~2 ~

23~ ~2/3

for :~ ~<~* < O,

(56a,b)

which has the correct asymptotic form for small values of lxl and approaches the correct limit for large values of Ix I. More specifically, it will be shown later in Fig. 9 that the approximate solution (56a) is quite close to the exact expression (30a).

Penetration of a rigid projectileinto an elastic plastictarget of finitethickness

815

Using (54c) and the approximate solution (56a) it follows that the function C passes through zero and becomes positive when x = x*** with

,
R~ ~ , (,4r,]~/~/. 1 -~t-~,,/
kVr-i

l
is7)
J

With the help of(31b) and the fact that Y/# << 1, it can be shown that x*** is about 2.5 times the value of x* (associated with the onset of yielding at the projectile's tip) since x*** R~ -1 /'4Y'~ 8 x* x* (58)

2x -45/-----~ t - ~ ) ~ 4-gTg~ ,~ 2 . 5R~.

It should be noted that the elastic-plastic boundary reaches the rear target surface when x = x** which is defined by (34). Simulations indicate that the value of C tends not to increase once the elastic-plastic boundary reaches the rear target surface. For this reason and the fact that the functional form for C must be non-negative for the target to cause deceleration of the projectile, we approximate C(x) by

C(x) = 0 C(x) = nYR2C(yc) C(x) = nYR2C(ff **) C(x) = 0

for for for for

x*** < x ~<0, x** ~<x ~<x***, - H ~<x < x**, x < - H,

(59a) (59b) (59c) (59d)

where the function (~(#) and the value ~** associated with (34) are defined by C(#) = In [45/9 ( 4 3 ~ ) ( - 2 ~

kl --~37) j

~2/31,

(60a)

~**--

\-~#,/ -~-~.

(60b)

It is of interest to note that for the limiting case when the target is semi-infinite and the penetration is deep ( - # >> 1), C is given by the constant

This form is seen to have the same functional dependence on the material constant #/Y as the solution of the punch problem [31 ], which supports the validity of the approximate velocity field and of the approximate boundary condition on the elastic-plastic boundary used above. Now, with the help of (52a), (54a,b) and (57), the equation of motion of the projectile (49a) becomes

(M + A oo):f = C(x),
which can be integrated analytically to deduce that
~ = - t?o .~ = - [ 0 ~ - $ ( ~ ) ] 1 / 2 = - [0~ x = - ~,(~**) - (~** - ~)~(~**)],/2 - [t?g -~**)0z** +/~)~0z**)] '/2

(62)

for x*** ~<x ~<0, for x** ~<x ~<x***, for - H ~<x ~<x**, for x ~< - H,

(63a) (63b) (63c) (63d)

O, =

where Uo, x,/~, t],, and the function $(~) are defined by (64a)

816 x=L\~-~-~,

A.L. Yarinet al:

r(M+A~)f4Y'~I/2]
~-~}

u2

~,

(64b) (64c)

( 4 Y ~ u2 H

O, = L\

[{M+A.)

d v,,

(64d)

~(#} = (#*** -- #) In {4519(43---~) } +~ . + 2x*** In L.I _~) j. (64e)

Equation (63) is used to determine either the depth of penetration or the residual velocity. When the target is thick enough, the depth of penetration P = - x is determined by solving (63) with ~ = 0. Alternatively, when the target is perforated, the residual velocity U, is determined by (63d) and (64d). Notice that for deep penetration [t~l >> 1] into a semi-infinite target the function t~(~7)may be approximated by 3# so that the depth of penetration determined by (63b) is given by

~=(4Y~"2 P =
\3/~] R~

02
lni4S/9{3#'~' (66)

}.''\4Y/IJ
or by (4) in dimensional form. The analytical expression (4) can be compared with the solution of the Poncelet Eqn (1) which has two empirical constants c~ and c2. Now it is important to emphasize that our assumption that the projectile remains rigid is only valid when the impact velocity U o is below a critical value. For many practical situations when a metal projectile penetrates a metal target this causes the value of c 2U 2 to be much smaller than unity so that (1) may be approximated by

P ~ clc2U 2.

(67)

Although (4) and (67) both predict that the penetration depth is a quadratic function of impact velocity, it is important to emphasize that the analytical formula (4) expresses the penetration depth in terms of measurable physical properties of the projectile and the target without the use of any empirical constants. The ballistic limit is defined as the minimum value U b of the impact velocity U o which causes the projectile to perforate the target. Using the analytical solution (63d) we may determine the ballistic limit by setting U o -- U b and U, = 0 to obtain the expression /~2 = ~(~**) + (~** +/4)t~(~**). (68) Later it will be shown in Fig. 14 that the predictions of the ballistic limit determined by the approximate analytical solution (68) compare well with calculations of the more general equations in Section 3. 5. COMPARISON WITH EXPERIMENTS, AND DISCUSSION Using the formulation described at the end of Section 3 we developed a computer program to numerically integrate Eqn (2) or (49a), subject to the initial conditions (3) with U o being the impact velocity. When the impact velocity U o is small enough or the target thickness H is large enough the projectile attains a maximum depth of penetration P which is given by the

Penetration of a rigid projectile into an elastic-plastic target of finite thickness

817

Table 2. A list of experiments, references, and relevant material and geometrical properties, t T h e material was the same as that used in Exp 1, but the yield stress was not given. SThe masses of the projectiles ranged between 24 and 25 g. T h e material parameters correspond to the middle soil layer Projectile M (g) 24.8 23.3 23.8 11.8 23.4 12.1 110 110 110 24.5:~ 78.0 23,100 L (ram) 82.9 74.7 81.8 73.7 74.7 39.1 ---81.8 146.0 674 R~ (ram) 3.55 3.55 3.55 2.54 3.55 3.55 10.0 10.0 10.0 3.55 4.765 47.6 p (kg/m 3) 2710 2710 2710 2710 2710 2710 7800 7800 7800 2710 2710 1860 Y (MPa) 396 396 396t 396t 396t 396t 860 920 1080 305 305 10 Target /~ (GPa) 28 28 28 28 28 28 83 83 83 28 28 0.05 H (mm) oo oc oc oc oc 80 40 20 25.4 25.4 oo

EXP 1 2 3 4 5 6 7 8 9 10 11 12

Ref. 11 11 11 36 36 36 37 37 37 38 39 40

Shape Ogive Sphere Cone Sphere Sphere Sphere Ogive Ogive Ogive Cone Cone Ogive

Table 3. A list of computations correlated with the experiments described in Table 2 ANAL COMP
1 2

EXP
1 --

Changes for the calculations, with the other parameters of the corresponding experiment being

fixed

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24

4 5 6 7 7 8 8 9 9 10 11 1 1 1 1 1 10 10 10 10 12 2 3

--H=oo -H=~ -H=oo


--

-H =0.06m H = 0.04 m U o = 0.4 km/s; H = 0.02-0.1 U o = 0.6 km/s; H = 0.02-0.1 U o = 0.7 km/s; H = 0.02-0.1 -R~ = 7 . 0 m m Y = 500 M P a Y = 1000 M P a ; / t = 83 GPa; 'Uo = 0.28, 0.50, 1.0 km/s H=oo H = oo

m m m

p = 7800 kg/m 3 (Steel target}

maximum value of Ixl e = Ixlmax(69) On the other hand, if the impact velocity is large enough then the projectile will exit the target with a residual velocity U r. In order to examine the accuracy of the numerical solution of Section 3 and the approximate solution developed in Section 4 we have compared computations with a number of experiments with projectiles of different shapes, lengths, radii, masses and targets with different material properties and thicknesses. A list of these experiments, together with the associated references and relevant material and geometric properties is recorded in Table 2. Also, Table 3 provides a list of computations which is correlated with the experiments

818
0.50 0.40' ~ 0.30. 0.20 0.10 0.00 0.2 0.4

A.L. Yarin et
. . . j . * . l . . . I . . . I , ,

al.
.I...1,. I1...

o + n

COMP1 o~++ ~ o

EXP 1 EXP 2 EXP 3

o/~/// /

0.6

0.8

1.0

1.2

1.4

1.6

1.8

~knvs)
Fig. 4. Semi-infinite target. Comparison of the computation with experimental results for three different projectile tips; ogive (EXP 1); sphere (EXP 2); and cone (EXP 3).

described in Table 2. For convenience in presenting this data in the following figures we use a numbering system which indicates that for example EXP 2 corresponds to experiment number 2, C O M P 1 corresponds to computation number 1, and ANAL 14 corresponds to the approximate analysis of Section 4 with the values associated with C O M P 14. We emphasize that in performing the calculations we have merely used the published values of the material properties and have made no attempt to adjust material parameters to match the experimental data. Figures 4 and 5 compare the computational results with experiments for a semi-infinite target. Figure 4 shows that the effect of the shape of the projectile tip is relatively small. However, the computation seems to compare better with the ogive projectile of EXP 1 because the ogive is a closer approximation to the ovoid of Rankine than the sphere of EXP 2 or the cone of EXP 3. Figure 5 shows the effect of changing the mass and dimensions of the projectile while keeping the tip shape constant. Figure 6 shows the effect on the penetration depth of targets of different finite thicknesses and Fig. 7 shows the effect of changing the mass and dimensions of the projectile for the same target thickness. Figure 6 also shows that the effect of the free rear target surface is only felt when the projectile tip comes close to the rear surface because the predictions of the main penetration stage of the projectile are nearly the same for finite and infinite thickness targets. This supports the validity of the approximate boundary condition used above at the rear target surface. Figure 8(a) shows the propagation of the projectile tip and the elastic-plastic boundary as functions of time, and Fig. 8(b) shows the drag force for finite and semi-infinite targets. Notice from Fig. 8(b) that the drag force remains reasonably constant until the projectile tip leaves the target and the surface area of the projectile in contact with the target begins to shrink rapidly. Also, notice from Fig. 8(b) that the drag force exhibits a kink when the elastic-plastic boundary reaches the target's rear surface and the form of the stress-free boundary condition is changed to that described by (43). Figure 9 shows that the approximate solution (56) of the location of the elastic-plastic boundary compares well with the exact solution obtained by solving (36). This suggests that the approximate analytical formulas developed in Section 4 may produce reasonably accurate predictions of both penetration depth and residual velocity. Figure 10 supports this conclusion because it shows that the penetration depth predicted by (66) compares well with a number of experiments. Additional figures examine various predictions of the numerical computations of Section 3 and the analytical solution of Section 4. For example, Fig. 11 shows that the predictions of the computations and the analytical solution are very close and that the effect of the finite target thickness is not significant for this case. The effect of the finite target thickness is explored more carefully in Fig. 12 and is again observed to be not too significant except near the ballistic limit.

Penetration of a rigid projectileinto an elastic-plastictarget of finite thickness


0.20
, , , I , , I , , , I , , , I , . ,

819

(a)
0.15
o

EXP 4

COMP 2
~" 0.10

0.05

0.00 0.2

'

'

'

'

'

'

'

'

'

'

'

'

0.4

0.6

0.8

1.0

1,2

u o (kaVs)

0.20

(b)
0.15

_.

EXP 3

/
/

~" 0.10

0.05

0.00 0.2

'

'

'

'

'

'

'

'

'

'

'

'

'

'

0.4

0.6

0.8

1.0

1.2

U (km/s)

0.20

'

' (c)

" '

' ' ' '

~ ,

0.15 ~0. 10
0.05

o
~

EXP 6
COMP 4

0.00 0.2

0.4

0.6

0.8

1.0

1.2

u o (k.m/s)

Fig. 5. Semi-infinite target. Comparisonof the computationwithexperimentalresultsfor projectiles with the same tip shape but differentmassesand dimensions.

Figure 13 shows the computational predictions of the elastic-plastic boundary for penetration into a semi-infinite target and for different values of the penetration depth. For convenience we have plotted the values {f, ~, ~ of {r,z, 4} normalized by the projectile dimension R oo. Figures 13(a,b) show the elastic-plastic boundary propagating relative to the fixed spatial coordinate 2, whereas Fig. 13(c) shows the elastic-plastic boundary relative to the coordinate ~- moving with the projectile. The vertical lines in Fig. 13(c) indicate the location of the targers front surface. Notice from Fig. 13(c) that the elastic-plastic boundary

820
0.06 0.05
0.04
, ,

A.L. Yarin et al.


I I , I ,

(a)
'o
--

EXP 7
coMPs

~ .
/f.-

. . . . . . COMP 6

.,~."

0.03 0.02 0.01 0.00 0.2

(H=

""

'

0.4
Uo

0.6

0.8

.0

(km/s)

0.06 0.05 0.04 0.03 0.02 0.01 0.00 0,2

(b) EXP 8 COMP 7 . . . . . . COMP 8


o

/ / ~. ./.-"" ""

'

'

'

0.4

0.6

0.8

1.0

Uo (km/s)

0.06 0.05 0.04 =., 0 . 0 3 O.02 0.01 0.00 0.2


I I

EXP 9 COMP 9 . . . . . . COMP i 0

I
I

(c)

(H = **)

O0 0

.-"

0.4

0.6

0.8

1.0

U (kin/s)
Fig. 6. Finite target. Comparison of the computation with experimental results for the same projectile penetrating targets of different thicknesses. The computations for the finite thickness targets are shown by the solid lines (COM P 5, 7, 9), whereas those shown by the dashed lines (COM P 6, 8, 10) correspond to a semi-infinite target with the other parameters of EXP 7, 8, 9 being fixed.

quickly propagates ahead of the projectile and attains a steady state shape relative to the projectile, with the coordinate ~-approaching the value ~-2 in (35b) at the intersection of the elastic-plastic boundary with the r = 0 axis. Figure 14 shows that both the computations and the analytical solution predict reasonably close values for the normalized ballistic limit. In this figure the values associated with the analysis were determined using Eqns (60a), (64c) and (68).

Penetration of a rigid projectile into an elastic-plastic target of finite thickness


2.00

821

(a)
1.50
o E X P 10

~'~'1,00
0.50

0 , 0 0

,,,/',l,t,,i,,,,i,r,,
0.0 0.5 1.0 1.5 2.0

u (~s)

1.00

(b)
o

= ,

J ,

0.80
"~ 0 . 6 0 ~0.40

E X P 11

0.20 0.00 0.2 . . . . . . 0.6 0.8 .0

'

'

'

0.4

u 0 (~Jnls) Fig. 7. Finite target (H =0.254m).Comparisonofthecomputationwithexperimentalresults for two


projectiles with the same tip shape but different dimensions.

With regard to the formulation of the computations of Section 3 we note that in addition to the mass of the projectile M in (49), the inertia of the target enters the analysis through the expressions for A and B in (52a), as well as in the determination of the position of the separation line from the solution of(47). However, the approximations of Section 4 indicate that for a metal projectile penetrating a metal target and for impact velociiies below about 1.5 km/s, the inertial effects of the target can be modeled simply by taking A constant and neglecting B. Although the computations of Section 3 would predict significant inertial effects for higher impact velocities, it is no longer possible to ignore the ei-osion of the projectile as it penetrates the metal target. However, by considering the penetration of a metal projectile into a soil target with a relatively low yield strength it is possible to observe significant inertial effects while still ensuring that the projectile remains rigid. For example, we used the material properties of soil given in [40] to perform the calculatioiis presented in Fig. 15 for a semi-infinite soil target. Figure 15(a) shows the deceleration of the projectile as a function of time (which agrees relatively well with the experimental data of [40]) and Fig. 15(b) shows the resulting penetration depth, each for different impact velocities. In particular, notice that the inertial effects for soil targets become signifie~tit for impact velocities as low as about 0.5 km/s. These inertial effects manifest themselves iti the deviation of:f from a constant as well as in the large differences between the computiati0fis of Section 3 and the analytical solution of Section 4 [see Fig. 15(b)].

822
0,02

A.L. Yarin et al.

(a) C O M P ]4
0.00-0.02
. . . . . . . . . . . . . . . . . . . . . . . . . . . . .

~ - - Front Surface
. . . . . . . . .

,~E - 0 . 0 4 N -0.06
-0.08

Elastic-Plastic Boundary Projectile Tip


,
1

/ !

-0.10 0.0

'

'

'

'

'

'

'

'

'

'

'

'

'

'

'

20.0

40.0

60.0 80.0 t (I.tsec)

100.0 120.0

80.0

''

' l ' ' ' l = l i l = = l i l '

' 1 ' ' '

(b)
60.0

...........

,--, 40.0
u. 20.0

O.0
-20.0 0.0' ''

- I ' '

COMP 14
(H = 0.04 m)
' I ' . ' I ' ' ' i . . . I ' ' '

20.0

40.0

60.0

80.0

100.0 120.0

t (p~seaz) Fig. 8. Finite target (H = 0.04 m). Locations of the projectile tip and the elastic-plastic boundary as well as plots of the drag force for finite thickness and semi-infinite targets. Apart from the target thickness, the other geometrical and material parameters of COMP 1 and COMP 14 are the same.
- 0 . 5 , , . I , , , i , , , i , , , i , , ,

i
I [
i

-o.6 . . . . . -O.7.

EXACT (36) - APPROXIMATE

-0.9 -1.0
- 1 . 1 , ' i ' ' ' I ' ' I ' ' I

-' 0 . 0

-8.0

-6.0

-4.0

-2.0

0.0

2.0

Fig. 9. Comparison of the exact (36) and approximate (56) expressions for the elastic-plastic boundary.

6. PENETRATION OF P R O J E C T I L E OF R E V O L U T I O N WITH ARBITRARY SHAPED TIP To generalize the above theory for the case of a projectile of revolution with arbitrary shaped tip it is necessary to construct the potential flow past such a body. One of the several methods for constructing the desired velocity field is the method of singularities [28, 29, 41]. The velocity field associated with the ovoid of Rankine is the simplest of such fields because it

Penetration of a rigid projectile into an elastic-plastic target of finite thickness


15.0
, , -

823

o
10.0. [] x

A ~
5 0 1

, , l , (66a) EXP 1 EXP 2 EXP 4 EXP 5 EXP 6 ~


~

, /

, o

~~

0.0
0.0

5.0
0

10.0

15.0

0 2 / In [45/9(31a/4Y)]

Fig. 10. Comparison of the analytical approximation (66) of the normalized depth of penetration with experimental data.

0.07
0.06. 0.05,
_ o.o4.

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

1.0
0.80.60.4-

(a) H =

0.06 m / / // i S'/ 13

,)..

(b) H = 0

"
--

COMP13
ANAL

=,-, 0 . 0 3 0.02.

0.01 0.00

2Y
. . . . . . COMP !
. . . . I . . . . I . . . . I . . . .

0.20.0 0.60

---

COMP 13

. . . .

. . . .

0.00 0.10

0.20

0.30
U o (knds)

0.40

0.50

0.40

0,60

0.80

1 .O0

1.20

U (kin/s)

0.07
0.06

. * , * I , , , , I l l , l l , l l , l l . l

llll.

1.0

(c) H = 0.04 m -COMP 14 ~ ." ~.'" ."


~' ~
0.6-

(d) H

0.05
.~ 0.04

- - - ANAL 14 . . . . . . COMP 1

0.6.

~" 0.03 0.02 0.01 0.00 ~."r" I .00 0.10


I I I I . . . .

0.4-"

-0.20.0 0.40

- COMP 14

0.20

0.30 0.40 u o (k.es)

0.50

0.60

. . . .

0.60

;Ti??,'i...
0.80 U o (kin/s) 1.00

1.20

Fig. 11. Comparison of the predictions of the computations (COMP) of the equations in Section 3 with the analytical solution (ANAL) of Section 4 for two targets of finite thickness. The dotted lines show the results for a semi-infinite target with the other parameters being fixed.

requires only a single source-type singularity. More generally, within the context of this method it is possible to satisfy the impenetrability condition at the surface of an arbitrary body of revolution by introducing additional singularities. The distributed singularities may be either sources, sinks, doubles, or vortex rings. Here, we consider the simple case of a distribution of source-sink singularities along the axis of the body.

824
0.70 0.60 0.50
~ 0.40.
. r . i . i I ~ i i I i i

A.L. Yarin et al.


0.10 . = , , , J , , , t , . , t , , ,

(a) U = 0.4 knVs


0.08

(b) U = 0.4 knv's -COMP 15

--

COMP 15

. . . . . . ANAL 15
.

g
~"

0.06

. . . . . . ANAL 15

~0

30"

0.04
. . . . . . . . . . . . . . . . . . . . . .

o.2o~ N ~ i
0.10 0.00 0.020
~ ~ .

0.02

0.00 0.02

'

'

'

'

'

'

'

'

'

'

'

0,040

0060

0.080

0.100

0.04

0.06

0.08

0.10

0.12

H (m)

H (m)

0.70 0.60. O.SO.


~ 0.40.

0.10

(c) U = 0.6 km/s


0.08 -

(d) Vo =0.6 knV


A

COMP 16

ANAL 16

0,06-

~., ---. . . . . . . . . . . . . . . . -COMP 16


-

E
0.04

~ 0.30" 0.20'

. . . . . . ANAL 16

0.02 0.10 0.00 O.02O 0.040 0,060 0.080 0.100 0.00 0.02
. . . ~ . . . , . . ; ~ . . . ~ .

, , 0.1

0.04

0.06

0.08

0.10

(m)

H (m)

0.10

,
M

. i

, ,

0.60- ~
0.50~ 0.40-

km/s

0.08 _

(f) oo = 0.7
v

k~

-............

COMP 17

g
"'",
COMP 17 ",N N

0.06.

. . . . . . ANAL 17
0.04-

~0.300.20-

0.02 -

ot.oo "..... : A L l 7 ,
0.020 0.040 0.060

,"i~
0.00
0.080 0.100 0.02
' ' ' ~ ' ; ~ ' ' ~ ' ' ' i ' , ,

0.04

0.06

0.08

0.10

0.12

H (m)

H (m)

Fig. 12. Comparison of the predictions of the computations (COMP} of the equations in Section 3 with the analytical soluiion (ANAL) of Section 4 for a range of target thicknesses.

Using this method the velocity potential ~b is given by ~b(r, ~, t) = ~(t) R ~ f ~ L - ~ ' / ~ F
q(fl)

q--J-~/"~

Ll-(~-n~+:]~/~J dO'

(70a) (70b)

= z - x(t) - c<,

where ~ is the Velocity of the body in the positive e z direction, R ~ is a parameter characteristic of the body's radius, L is the length of the body, the tip of the body is located at ~ = - ct (with 0 < ~ < L), arid q is the distribution of sitigiilarities. Oiven a projectile shape characterized by the radius r = R() it may be shown that the distribution q of sourCe-sink singularities which enSUres impenetrability is determined by the Fredhoim integral eqUation of the

100.0

15.0

'

'

'

825

(a)
80.0:
60.0 L 10.0-

(b) -~ 0.5

-x/R

5.0-

1,
40.00.0-

20.0-

10 fl00

-5.0-

0.0:
-20.0 -120.0 ", ,
,

-10.0-

'

, 0.0

-15.0 -30.0

'

'

'

'

. . . .

I -10.0

. . . .

90.0

-60.0

-30.0

-20.0

0.0

15.0J

'

'

'

, I a i J i

i I J , , , ,

(c)

Projectile \

5.0-1 i

0.1.,~/ 5

"

I\

ooJ

":

] .................

.... J ..J .......

-lO.O:
-15.0: -15.0
. . . . .

-5.0

5.0

15.0

Fig. 13. Semi-infinite target. Computation of the elastic-plastic boundary relative to the front of the target (a) and (b), and relative to the tip of the projectile (c), for the data associated with experiment EXP 1. The values o f ( - x / R 2) indicate the positions of the projectile tip.
5.0
5.0
,,,,I,,, ,I,,, ,I,,,, I,
, ,

(a)
4.03.0

(b)
4.0

/
~

3.0

.-,~ ..'7" j ANAL 18 ANAL 19 . . . . . . ANAL 20 ANAL 21


I . . . . I . . . . I . . . . I . . . .

2.0

1.0

- - - - COMP 18 COMP 19 . . . . . . COMP 20 COMP 21


1 . . . . I ' ' ' ' I . . . .

2.0

1.0-

0.0
0.0

0.0

. . . .

1.0

2.0

3,0

4.0

5.0

0.0

1.0

2.0

3.0

4.0

5.0

2 /In 14519(3p./4Y)]
5.0 4.0. 3.0. 20 1.0
. . . . I . . . . I . . . . I . . . . I

Ub2 /In Hslg(3p/4Y)I


. . . .

(c)

- - - -- ANAL 20
I . . . . I . . . . I . . . . I . . . . I . . . .

0.0

0.0

1.0
b

2.0

3.0

4.0

5.0

2 /In [45;9(-'~tl/4Y)] Fig. 14. The normalized ballistic limit predicted by the computations of the equations of Section 3 (COMP), and the analytical solution of Section 4 (ANAL).

826

A.L. Yarin et al.

0.0
-10.0

~
, _

-20.0

.'
,

COMP 22 --U --o

0.28km/s

-30.0

U = 0.5kmls
o

-40.0,

::'

- . . . . . U o = 1.0 km]s

-50.0

-20.0

0.0

20.0

40.0

60.0

80.0

100.0

t (msec)

100.0 80.0 ~" 60.0 40.0

a , i I , , , l i , a l , , , I , , , I , , ,

(b)
/

/ /
/

- - - ANAL 22 - COMP 22
j f

/ /

200

0.0

'

'

'

'

'

'

'

"

'

'

'

'

0.0

0.2

0.4

0.6

0.8

1.0

1.2

U (kin/s)
Fig. 15, Semi-infinite soil target. Predictions of the deceleration of the projectile and its depth of

penetration for differentinitial velocities.

first kind [28, 29, 41, 42].

A computer program was developed in [42] to solve (71) for the distribution of singularities q and the associated velocity field. This program was first tested on the ovoid of Rankine (1 la) which corresponds to the analytical solution of(71) with q being the delta function 6(f/) and ct=R~/2 [this form for q also yields (13a) from (70a)]. Figure 16(a) shows that the calculated distribution is very close to the delta function of the analytical solution and Fig. 16(b) shows the numerically generated velocity field Figures 17 and 18 show similar results for hemispherical and conical shaped projectiles, respectively. Once the velocity field has been calculated it is possible to determine the deviatoric stresses in the rigid-plastic and the elastic regions of the target using the formulae of Section 2 (for details see [42]). However, in contrast with the solution for the ovoid of Rankine, the divergence of these deviatoric stresses in general will not be the gradient of a scalar so the momentum equation cannot be satisfied exactly pointwise in the target region. Nevertheless, it is possible to determine the pressure field by integrating along the streamline which follows the r = 0 axis ahead of the projectile and follows the projectile surface behind the projectile tip. This integration also uses appropriate matching conditions at the elastic-plastic boundary and boundary conditions at the rear target surface. Then, the force applied to the projectile may be determined and an equation of motion of the projectile similar to (2) may be numerically integrated.

Penetration

of a rigid projectile into an elastic-plastic target of finite thickness 50.0 40.0 : 30.0. .... t .... t .... t .... i .... t..,,

827

(a)

"~

20.0.
10.0 0.0. -10.0 -1.0 .... I .... -0.5 , ~"--~ I .... 0.0 I .... 0.5 a .... 1.0 I .... 1.5 2.0

11

1.5

,_

0.5 0.0
-1.0 0.0 1.0 ~ 2.0 3.0 4.0

F i g . 16. S o u r c e d i s t r i b u t i o n

a n d f l o w field a b o u t

t h e o v o i d o f R a n k i n e . (a) s o u r c e d i s t r i b u t i o n ;

(b) s t r e a m l i n e s a n d t h e b o d y s h a p e . 200.0

(a)
150.0

100.0
~v

50.O

0.0

-50,0 -1.5

....

~ .... i .... w. . . . -1.0 -0.5 0.0:0.5 q

i ....

i .... 1.0

~ .... 1.5 2.0

0.0

'

'

'

'

'

'

. . . .

-2.0

-1.0

0.0

_-

1.0

2.0

3.0

Fig.~17. S o u r c e

distribution

and

f l o w field a b o u t

a hemispherical

shaped

p r o j e c t i l e . (a) s o u r c e

d i s t r i b u t i o n ; (b) s t r e a m l i n e s a n d t h e b o d y s h a p e .

Using this procedure the penetration depth was calculated for projectiles with hemispherical (COMP 23) and conical (COMP 24) tip shapes and the computational results are compared with associated experimental data in Fig. 19. Also, shown in this figure are the computational results based on the model of Sections 2 and 3 using the ovoid of Rankine.

828
10.0

A.L. Yarin et al.

5.0-

0,0

(a)

-5.0"

-10.0

.... I .... I .... I .... I .... I .... I .... -1.0 0,0 1.0 2 . 0 ~] 3.0 4.0 5.0 6.0

0.0 -1.0 0.0 1.0 ~ 2.0 3.0 4.0

Fig. 18. S o u r c e d i s t r i b u t i o n a n d f l o w field a b o u t a c o n i c a l s h a p e d p r o j e c t i l e ( w i t h s l i g h t l y r o u n d e d tip). (a) s o u r c e d i s t r i b u t i o n ; (b) s t r e a m l i n e s a n d t h e b o d y s h a p e .

These results, as well as the experimental data [15], indicate that the exact shape of the projectile tip does not significantly influence the prediction of the penetration depth. The fact that integral quantities such as the penetration depth and the residual velocity are only slightly affected by differences in the shape of the projectile tip is well known [16]. Even the extreme case of modeling a conical projectile as a flat cylinder seems to be adequate to predict certain integral quantities [18]. Note also, that a similar conclusion follows from works on drawing and extrusion of ductile metals since such integral quantities as drawing pressure and drawing tension appear to be insensitive to details of the velocity field assumed in the drawing zone [43-46]. 7. CONCLUSIONS The model developed in the present work accounts for the inertial and elastic-plastic drag imposed by a target on a penetrating projectile, as well as the finite thickness of the target. The solution is determined completely by material and geometrical parameters and, in particular, it does not involve any empirical constants or functions. It has been shown that when the inertial effects of a semi-infinite metal target are negligible (typically for impact velocities less than about 1.5 km/s) the analytical solution of Section 4 for the penetration depth (4) reduces to the same form as the result derived from solution (67) of the empirical Poncelet equation. In fact, the expression (4) allows one to determine the empirical constant c l c 2 in (67) in terms of measurable physical properties of the projectile and the target. Analytical predictions of the ballistic limit and the residual velocity of a projectile penetrating a target of finite thickness agree well with numerous experimental data. This is in spite of the fact that the calculations based on the analytical solution of Section 4 involved only published values of material and geometric properties of the projectile and target. The inertial effects of a metal target remain relatively small for impact velocities below about 1.5 km/s, which is near the limit that causes significant erosion of a metal projectile.

Penetration of a rigid projectile into an elastic-plastic target of finite thickness

829

0.40

(a)
0.30 + EXP2 C O M P23 ,/ ,~

../~
I

_,

0.20

0.10

0.00 0.3 0.5 0.7


0

0.9

1.1

1.3

1.5

U (kin/s)

0.40

(b)
0.30-

EXP3

COMV24
,--. 0.20h 0.10. . . . . . COMP 1 /

/u
~

,/kf'_"

[t

0.00 0.3
0.5

0.6

0.8

1.0

1.2

1.3

1.5

U (kin/s)
Fig. 19. Comparison of computations with experimental data [l 1] for two different projectile tips penetrating a semi-infinite target; (a) hemispherical (COMP 23 and EXP 2); (b) conical (COMP 24 and EXP 3). The computations (COMP 1) are based on the model of Sections 2 and 3 using the ovoid of Rankine as the projectile.

For these cases, the approximate analytical solution of Section 4 seems adequate. However, when the inertial effects of the target become significant, the more complete computations of Section 3 are required. This can occur for impact velocities as low as about 0.5 km/s without erosion of the projectile when a metal projectile penetrates a relatively low strength soil target. The method of singularities, briefly reviewed in Section 6, was used to determine the velocity field for hemispherical and conical projectiles. Subsequent calculations indicated that the prediction of such integral quantities as penetration depth and residual velocity are relatively insensitive to the exact shape of the projectile. Thus, an accurate description of the penetration process can be obtained using the solution developed here for a projectile in the form of an ovoid of Rankine. Finally, we note that the solution procedure outlined in Section 6 can be generalized for the case of oblique penetration and details will be presented in a future publication.
Acknowledgements--The authors would like to thank Prof. S.R. Bodner and Prof. D. Durban for helpful discussions. A. L. Yarin is a recipient of the Guastalla Fellowship established by Foundation Rashi, Planning and Grants Committee of the Council of Higher Education, the Israel Academy of Sciences and Humanities. This work was also partially supported by the Technion V.P.R. Fund.

REFERENCES
1. M. E. Backman and W. Gdsmith The mechanics f penetratin f prjecties int targets` nt. J. Enyn9 Sci. 16, 1-99 (1978). 2. C. E. Anderson Jr, B. L. Morris and D. L. Littlefield, A penetration mechanics database. Southwest Research Institute Report 3593/001 11992).

830

A.L. Yarin et al.

3. G. Birkhoff, D. F. Macdougal, E. M. Pugh and G. I. Taylor, Explosives with lined cavities. J. Appl. Phys. 19, 563-582 (1948). 4. A. Ya. Sagomonyan, Penetration. Moscow University Press, Moscow (1974) (in Russian). 5. I. Frankel and D. Weihs, Hydrodynamic theory of glancing impact. J. Fluid Mech. 216, 213-223 (1990). 6. V.P. Alekseevskii, Penetration of a rod into a target at high velocity. Combustion, Explosion and Shock Waves 2, 63-66 (1966). 7. A. Tate, A theory for the deceleration of long rods after impact. J. Mech. Phys. Solids 15, 387-399 (1967). 8. A. Tate, Further results in the theory of long rod penetration. J. Mech. Phys. Solids 17, 141-150 (1969). 9. J. Falcovitz, M. Mayseless, Z. Taubes, D. Keck, R. Kennedy, K. Oftedahl and P. Singh. A computer model for oblique impact of a rigid projectile at ductile layered targets. 1 lth Int. Symp. on Ballistics, pp. 1-11, Brussels (1989). 10. R. F. Recht and T. W. Ipson, Ballistic perforation dynamics. J. Appl. Mech. 30, 384-390 (1963). 11. J. Awerbuch and S. R. Bodner, Analysis of the mechanics of perforation of projectiles in metallic plates. Int. J. Solids Struct. 10, 671-684 (1974). 12. J. Awerbuch and S. R. Bodner, Experimental investigation of normal perforation of projectiles in metallic plates. Int. J. Solids Struct. 10, 685-699 (1974). 13. J. Liss, W. Goldsmith and J. M. Kelly, A phenomenological penetration model of plates. Int. J. Impact Engng 1, 321-341 (1983). 14. G.I. Taylor, The formation and enlargement of a circular hole in a thin plastic sheet. Quart. J. Mech. Appl. Math. 1, 103-124 (1948). 15. M. J. Forrestal, K. Okajima and V. K. Luk, Penetration of 6061-T651 aluminium targets with rigid long rods. J. Appl. Mech. 55, 755-760 (1988). 16. R. Hill, Cavitation and the influence of headshape in attack of thick targets by non-deforming projectiles. J. Mech. Phys. Solids 28, 249-263 (1980). 17. A. Tate, A comment on a paper by Awerbuch and Bodner concerning the mechanics of plate perforation by a projectile. Int. J. Engng Sci. 17, 341-344 (1979). 18. M. Ravid and S. R. Bodner, Dynamic perforation of viscoplastic plates by rigid projectiles. Int. J. Engng Sci. 21, 577-591 (1983). 19. M. Ravid S. R. Bdner and I. H crnan Analysis f very high speed impact. nt. J. Engng Sci. 25 473-482 ( 987). 20. M. Ravid, S. R. Bodner and I. Holcman, A two-dimensional analysis of penetration by an eroding projectile. Int. J. Impact Engng 15, 587-603 (1994). 21. A. Tate, A simple hydrodynamic model for the strain field produced in a target by the penetration of a high speed long rod projectile. Int. J. Engng Sci. 16, 845-858 (1978). 22. A. Tate, Long rod penetration models--Part I. A flow field model for high speed long rod penetration. Int. J. Mech. Sci. 28, 535-548 (1986). 23. A. Tate, Long rod penetration models--Part II. Extensions to the hydrodynamic theory of penetration. Int. J. Mech. Sci. 28, 599-612 (1986). 24. M. R. Sitzer and D. Durban, Steady penetration of rigid tools through plastic media. Acta Mech. 40, 209-220 (1981). 25. T. Gobinath and R. C. Batra, A steady state axisymmetric penetration problem for rigid/perfectly plastic materials. Int. J. Engng Sci. 29, 1315-1331 (1991). 26. R. C. Batra and A. Adam, Effect of viscoplastic flow rules on steady state penetration of thermoviscoplastic targets. Int. J. Engng Sci. 29, 1391-1408 (1991). 27. M. B. Rubin and A. L. Yarin, On the relationship between phenomenological models for elastic-viscoplastic metals and polymeric liquids. J. Non-Newtonian Fluid Mech. 50, 79-88 (1993). Corrigendum 57, (2/3) (1995). 28. N. E. Kochin, I. A. Kibel and N. V. Roze, Theoretical Hydrodynamics. Interscience, New York (1964). 29. G. K. Batchelor, An Introduction to Fluid Mechanics. Cambridge University Press, Cambridge (1967). 30. L. M. Milne-Thomson, Theoretical Hydrodynamics. MacMillan, New York (1968). 31. R. F. Bishop, R. Hill and N. F. Mott, The theory of indentation and hardness tests. Proc. Roy. Soc. Lond. 57, 147-159 (1945). 32. H.G. Hopkins, Dynamic expansion of spherical cavities in metals, Chapter III in Progress in Solid Mechanics, Vol. 1 (Edited by I. Sneddon and R. Hill). North-Holland, Amsterdam (1960). 33. M. J. Forrestal and V. K. Luk, Dynamic spherical cavity expansion in a compressible elastic-plastic solid. J. Appl. Mech. 55, 275-279 (1988). 34. D. Durban and N. A. Fleck, Singular plastic fields in steady penetration of a rigid cone. J. Appl. Mech. 59, 706-710 (1992). 35. James & James Mathematics Dictionary. Van Nostrand Reinhold, New York (1968). 36. 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). 37. S. N. Dikshit and G. Sundararajan, The penetration of thick steel plates by ogive shaped projectiles-experiments and analysis. Int. J. Impact Engng 12, 373-408 (1992). 38. Z. Rosenberg and M. J. Forrestal, Perforation of aluminium plates with conical-nosed rods--additional data and discussion. J. Appl. Mech. 55, 236-238 (1988). 39. M. J. Forrestal, Z. Rosenberg, V. K. Luk and S. J. Bless, Perforation of aluminium plates with conical-nosed rods. Report No. SAND 86-0292J, Sandia National Laboratory, Albuquerque, NM (1986).

Penetration of a rigid projectile into an elastic-plastic target of finite thickness

831

40. M. J. Forrestal and V. K. Luk, Penetration into soil targets. Int. J. Impact Engng 12, 427-444 (1992). 41. M. Rouse, Advanced Mechanics of Fluids. Wiley, New York (1964). 42. I.V. Roisman, Description of projectile penetration with small Deborah number, M. Sc. Thesis, Technion, Haifa (1994). 43. D. Durban and B. Budiansky, Plain-strain radial flow of plastic materials. J. Mech. Phys. Solids 26, 303-324 (1979). 44. D. Durban Drawing and extrusin f cmpsite sheets wires and tubes. nt. J. Slids Struct. 2 649-666 ( 984). 45. R.M. Govindarajan and N. Aravas, Asymptotic analysis and numerical simulation ofdeformation processing of porous metals. In Finite Inelastic Deformations-- Theory and Applications, pp. 58-66 (Edited by D. Besdo and E. Stein), Springer, Berlin (1992). 46. J. Tirsh and D. ddan The dynamics f fast meta frming prcesses. J. Mech. Phys. Slids 42 6 -628 ( 994).

You might also like