Sei sulla pagina 1di 18

Int. J. Impact Engn9 Vol 12, No, 3. pp 427-444, 1992 0734-743X/92 $5.00 + 0.

00
Printed m Great Britain Pergamon Press Ltd

PENETRATION INTO SOIL TARGETS

M. J. FORRESTALand V. K. LuK
Sandia National Laboratories, Albuquerque, NM 87185, U.S.A.

(Received 20 September 1991; m revised form 15 Aprd 1992)

S u m m a r y - - W e developed penetration equations for ogwal-nose projectiles that penetrated soil


targets after normal impact. The spherical cavity-expansion approximation simplified the target
analysis, so we obtained closed-form penetration equations. To verify our model, we conducted
field tests into soil targets with 23.1 kg, 95.25 mm diameter projectiles and obtained rigid-body
decelerations and final penetration depths. Model predictions show good agreement with measurements
for five tests conducted at an impact velocity of 280 m s - 1.

INTRODUCTION

Backman and Goldsmith [ 1] discuss many analytical and experimental methods used in
the broad field of penetration mechanics. For metal penetration, laboratory-scale experiments
play a dominant role in the research required to understand penetration phenomena.
However, experimental work on penetration into undisturbed soil targets requires field
tests. In addition, analytical models require material tests of samples cored from the
geological targets for a reasonable constitutive description. Without a companion experimental
program of material tests, penetration studies into soil targets are limited to empirical
models [2,3].
In this study, we developed penetration equations for ogival-nose projectiles that
penetrated soil targets after normal impact. Target constitutive models idealize pressure-
volumetric strain as a locking solid, whereas the Mohr-Coulomb, Tresca and
Mohr-Coulomb with a Tresca-limit yield criteria idealize the shear-strength-pressure
behavior. We derived equations for the spherically symmetric, cavity-expansion problems
for the three shear-strength models and used these results to develop our penetration
equations. To verify our penetration equations, we conducted triaxial material tests and
penetration tests into soil targets at the Sandia Tonopah Test Range, Nevada. Predictions
showed good agreement with field-test data.

FORMULATION OF THE PENETRATION MODEL

A projectile with an ogival nose impacts a uniform target at normal incidence with
velocity Vo and proceeds to penetrate at rigid-body velocity Vz. As Fig. 1 shows, an ogive
is the arc of a circle and is tangent to the penetrator shank. The ogive is commonly defined
in terms of caliber-radius-head
CRH = s/Za = ~O, (1)
where s and a are as defined in Fig. 1. For a rigid projectile, motion and final depth can
be calculated when forces on the projectile are known. Thus, we first model target resistance
and then calculate deceleration, velocity and depth.
To obtain closed-form penetration equations, we use the spherical, cavity-expansion
approximation [4]. Normal stress on the projectile nose is approximated by results from
spherically symmetric, cavity-expansion analyses that relate cavity-expansion velocity to
radial stress on the cavity surface. In particular, normal stress on the projectile nose is
taken as the stress at the cavity surface associated with the distributed expansion velocity
at the nose-target interface caused by projectile penetration. In addition, we assume a
tangential stress on the nose from interface frictional resistance. We take the simple form
427
428 M. J, FORRESTALand V. K. LUK

~S f

I
FIG. I. Ogival nose geometry.

a, =/~a.. where a,, a. are the tangential and normal stress components and/~ is the sliding
friction coefficient.
The resulting axial force on the projectile nose [5,6,7] is given by

0o s,.

The normal stress a.( V=, 40 will be approximated by results from spherically symmetric,
cavity-expansion analyses, and other terms are defined in Fig. 1.

SPHERICAL CAVITY-EXPANSION FOR A MOHR-COULOMB MATERIAL


A spherically symmetric cavity is expanded from zero initial radius at constant velocity
V. As Fig. 2 shows, this expansion produces plastic and elastic response regions. The plastic
region is bounded by the radii Vt and ct, where t is time and c is the elastic-plastic interface
velocity. As Figs 3a and c illustrate, material description for the plastic region is represented
by a locked hydrostat (pressure-volumetric strain) and a M o h r - C o u l o m b yield criterion.
Our previous model [5] used the Tresca yield criterion (Fig. 3b). The elastic region is
taken as an incompressible elastic material.

Plastic response
For a locked hydrostat, the M o h r - C o u l o m b yield criterion (Figs 3a and c), and spherical
symmetry

rl* = 1 - Po/P* (3a)


o r - a 0 = 3o + 2p (3b)
iv = (at + 2a0)/3, (3c)
where Po, P* are the initial and locked densities, q* is the locked volumetric strain, a,, a o
are the radial and tangential components of Cauchy stress (positive in compression), ~o,
2 define the yield condition, and iv is pressure. Equations (3b) and (3c) can be combined
P e n e t r a t i o n into soil targets 429

Vt ct

FIG. 2. Responseregions.

G r- cro

"c0

rl P

(a) (b)

T.m
G r - Go Gr - G o

~0 "co

p p

(c) (d)

FIG. 3. Soil m a t e r i a l descriptions.

to give
a, - ao = aro/2 + ear (4a)
c~ = 32/(3 + 2).). (4b)
The equations of momentum and mass conservation in Lagrangian coordinates are

(r+u) 2 +2 1 + Or/ +u)(a~-ao)+po r Ot---~

( r + u ) 3 = P ° r 2, (Sb)

where r is the Lagrangian coordinate, u is radial displacement (positive outward) and p


is current density. For this application, p = p* and ( a r - an) is given by (4a) and (4b).
The boundary condition at the cavity surface is
u ( r = O, t) = Vt. (6)
For a uniform cavity-expansion (6), we can reduce (5a,b) to ordinary differential equations
with the similarity transformation
= r/ct (7a)
and the dimensionless variables
u(r, t) -- cta(~), S = a,/~o. (7b)
430 M.J. FORRESTALand V. K. LUK

With (7a), (7b) and (4a), the conservation equations (5a) and (5b) become
dS d -2ct d po c2 44 d2t]
(¢ + f i ) ~ + 2 c t ~ ( 4 + fi)S - d - ( ¢ + fi) . . . . (8a)
,~. l"0 (¢ 4- t / ) d 4 2

ld
---[(4 + fi)3] = (1 -- q*)42, (8b)
3 de
and the boundary condition (6) transforms to
a(o) = V/c. (8c)
Equation (8b) integrates to
(3 + a) 3 = (1 - r/*)~ 3 + (V/c) 3. (9)

An exact differential for the left side of (8a) is found by multiplying both sides of (8a) by
(3 + fi)2,-~ and combining terms. Thus,
d~[(4 + a)2"S] = --~--
- 2 a (4 + /~)2a- 1 ~d( ~ + fi) _ po c2 44(¢ + f i ) 2 ~ , - ,d2O
, (lOa)
Z0 de 2

and an integration gives

2 f
(~ + a)2~S= -2----~ (3+ a)2=-,d ( ~ + f i ) - p°c2 f 4 4 ( 3 + f i ) 2 t ' - t ~-d-2 f i d ¢ + C 1, (10b)
l"0 d4 2

where C~ is an integration constant. Functions of (¢ + ~) are obtained from (9) and


substituted into (10b). Evaluation of the integrals gives
Ct 1 poc2(V/c)3[2(1 -- q*)(2 - ~ ) ~ 3 4- 3(V/c)3]
s(4) = - - +
[(1 - q,)~3 + (V/c)312~/3 2 to(1 - r/*)(1 - 2a)(2 - a)[(1 - r/*)~ 3 + (V/c)3] 4/3'
(11)
where the elastic-plastic interface velocity c and the constant C~ will be determined later.

Elastic-plastic interface conditions


The elastic and plastic regions are connected through the Hugoniot jump conditions
that conserve mass and momentum across the elastic-plastic interface. From [5]
U 2 = r/* + (1 - r/*)U1 (12a)

S 2 -- S 1 4- p°c2r/* (1 - U1) 2 (12b)


1-o

U = v/c, (12c)
where v is the radial particle velocity (positive outward). The subscripts l and 2 refer to
quantities at the interface (4 = 1) in the elastic and plastic regions, respectively.

Elastic response
The elastic response for an incompressible material is presented in [5]. Displacement,
particle velocity and radial stress are
fi- r° (13a)
2E42

3Zo
U - (13b)
2E42

2 3po c2 (13c)
s--T+ E--Z-'
where E is Young's modulus.
Penetration into soil targets 431

Response equations
Displacement is continuous at the interface (4 = 1), so the interface velocity c can be
determined from (9) and (13a). Thus,

-=?= 1+ -(1-tl*)| . (14a)


c 2E/
Radial stress in the plastic region at ~ = 1 is obtained by substituting (13b) and (13c)
into (12b) and using (14a), so

$2=~+ - +~/* 1- (14b)


72% 2E J J
The constant C~ in (1 l) can be evaluated with (14b). Thus,
(1 + ro/2E) 2~ 1 poV2 S 7312(1 - q*)(2 - ct)~ 3 + 373 ]
S(~) = ~[(1 -- F/,)~3 -1"-7312a/3 -- ~ "b - 72"f0
- ) (1 - r/*)(l - 2~)(2 - ~)[(1 - r/*)¢ 3 + .~314/3

(1 + Zo/2E)2~'[3zo/E + q*(1 - 3"Co/2E)2]


+
[(1 - q,)~3 + 7312a/3

73(1 + Zo/2E)Z~-2112(1 -- ~*)(2 - c 0 + 373]


- (5 - 2-gi - - 7'] (15)

Examination of(15) shows that the terms (2 - ~) and ( 1 - 2~) appear in the denominator.
From (4b), ( 2 - ~ ) = (6 + 2)/(3 + 22), which does not cause any problem. However,
(1 - 2 a ) = ( 3 - 42)/(3 + 22), and (1 - 2 ~ ) approaches zero as ). approaches 3/4. The
special case for 2 = 3/4 was therefore analyzed separately. Using the same techniques,

2(1 + Zo/2E ) 4 Po V2 ( 273 ~3


S(~,2 = 3/4)=
[(1 -- q*)~3 ..1_ 7311/3 3 + Zo72[( 1 -- r/*)~ 3 + 73] 1/3 13[(1 - r/*)~ 3 + 73]
2), 3 In [( 1 - r/*)~ 3 + 7 ~]
+ (1 + ro/2E)[3ro/E + q*(1 - 3%/2E) z]
3(1 - r/*)
273 273 In [(1 + "Co/2E)3]'~
(16)
3(1 + Zo/2E) 3 + 3(1 r/*) J'

The special case for 2 = 0 (Tresca criterion shown in Fig. 3b) was given in [5]. However,
we correct an error in this paper (fortunately, the error did not affect the numerical results)
and present the equation for 2 = 0. Thus,
S(~,~. = 0) = 2 2 lnl( ! _-__q*)~3 +Y3 ] poV2[
-3 k (1 + Z o / 2 E ) 3 +--72Zo 3r°/E + q * ( 1 - 3 z ° / 2 E ) 2]

+ o°VZY-, { [373 + 4(1 - r/*)¢3] [373 + 4(1 - q---*)]'l. (17)


2%(1 - , ) [(----iZ ~/*5~g +-?3-T~ - (1 + zo/ZE) 4 J

Equations (15), (16) and (17) with (14a) give the radial stress in the plastic region
(0 < ~ < 1). Equation (13c) with (14a) gives the radial stress in the elastic region (~ > 1).
For penetration resistance, we will later use the relation between radial stress at the
cavity surface (¢ = 0) and the cavity-expansion velocity V. We write
S(~ = O) = A + BpoV2/zo . (18)
From (15) with ~ = 0,
A=!(}+Zo/2E)2~' 1 (19a)
432 M.J. FORRESTALand V. K. LUK

3
B= + 3zo/E + r/*(1 - 3Zo/2E) 2
(1 -- q*)(1 -- 2ct)(2 -- ct) )' / t.
7'3[-2(1 -- r/*)(2 -- o0 + 3y 3] "~
- (1 - ~ ] --- 2~)~--- ~ i + z~-2E) 4 J" (19b)

From (16), which is the special case for 2 = 3/4,


A = 2(1 + ro/2E)/~ -- 4/3 (20a)
-21ny (1 + ro/2E)[3ro/E + q*(1 -- 3ro/2E) 2]
B - - - +
(1 -- li*) }~3

2[ 1 3In(1 + Zo/2E)]
(1 + Zo/2E) 3 (1 -- ~7~ ]" (20b)

From (17), which is the special case for 2 = 0,

A=~{1-1n[ (1 + z ° / 2 E ) 3 - ( 1 - q * ) ]
~ + ~o/~-E~3 } (21a)

3 [(3to/E) + q*(l - 3Zo/2E) 2]


B- +
2(1 - r / * ) /-(1 + Z o / 2 E ) 3 - ( 1 - q,)]a/a

_ [(1 +ro/2E)3--(12({_+ z--~/2E~- q,)]l/3 [ 1 + 3(1-+ z-°2/E)3q*)


](1- _J' (21b)

where ct and ? are defined by (4b) and (14a). Equations for A and B involve only material
constants.
In some applications, zo/E is small compared to unity. For E --* m, there is no particle
motion ahead of the plastic region, and the response equations simplify. For E ---, m in (14a)
7 = V/c = (r/*) 1/3. (22)
For E ~ ov in (19a) and (19b)
1 1
A- - - (23a)
~(r/*) 2"/3 2

3 -- (r/*)(1-2~)/3 [(2 -- ~)(1 + 2ct) + q*(1 - 2~)(1 - ~z)]


B= (23b)
(1 - q * ) ( 1 - 2c¢)(2 - c¢)

For the special case for 2 = 3/4, and E--+ oo, Eqns (20a) and (20b) reduce to
A = 2/(q*) U3 - 4/3 (24a)
1 2 In r/*
B- (24b)
3 3(1 - q*)"
For the special case for 2 = 0 and E --* oo, Eqns (21a) and (21b) reduce to
A = (2/3)(1 - In q*) (25a)
[ 3 -- (r/*)1/3(4 -- q*)]
B = (q.)1/3 + ~ ~ . (25b)

SPHERICAL CAVITY-EXPANSION FOR A MOHR-COULOMB TRESCA-LIMIT


MATERIAL
As discussed in [8], shear-strength-pressure for many geological materials can be
approximated as shown in Fig. 3d. For this constitutive model, the shear strength z becomes
a constant value z,. for sufficiently high-pressure p. The material model shown in Fig. 3d
is called the M o h r - C o u l o m b Tresca-limit yield model.
Penetration into soil targets 433

For spherically symmetric, cavity-expansion analyses, three solutions are required. First,
for small enough cavity-expansion velocity V and large enough values of the ratio r , , / z o,
the stress state in the plastic region lies on the M o h r - C o u l o m b yield line. This solution
is given in the previous section. Second, for an intermediate value of V, the plastic region
has two zones separated at ~ = ~,.. One zone (0 < ~ < ¢,,) is bounded by the cavity surface
and has a stress state on the Tresca-limit yield line, while the other zone (¢,, < ¢ < 1) is
bounded by the elastic-plastic interface and has a stress state on the M o h r - C o u l o m b
yield line. Third, for large enough cavity-expansion velocity V, the stress state for the entire
plastic region lies on the Tresca-limit line.
If the stress state in the plastic region lies on the M o h r - C o u l o m b yield line, Eqn (3b)
applies. However, if the stress state lies on the Tresca-limit yield line
a r - a o = r,,. (26)
The magnitude of radial stress required to reach the Tresca-limit line is defined as S,. and
is obtained by equating (4a) and (26)• Thus,

S,. = _1( r m ) 1. (27)


~\r0/ 2
If S ~< S,,, the stress state lies on the M o h r - C o u l o m b line; if S >/S,,, the stress state lies
on the Tresca-limit line.
Radial stress in the plastic region monotonically decreases from the cavity surface to
the elastic-plastic interface. Since maximum radial stress is at the cavity surface ~ = 0, the
minimum value of cavity-expansion velocity Vm.o required to reach the Tresca-limit yield
line is obtained by equating S(~ = 0) and S,.. From (18), (19) and (27),

ro -c~B ~ o - 7 / d" (28)


Since the minimum radial stress in the plastic region is at the elastic-plastic interface,
there is a large enough value of cavity-expansion velocity Vmax such that the entire stress
state lies on the Tresca-limit yield line. Vm,x is obtained by equating S: and S,,. From (27)
and (14b),
po(Vmax)2 __ ),2 rL(l.m',~ 1 2]
(29)
ro [ 3 z o / E + r/*(1 - 3Zo/2E) z] / ~ k/ r o //
L 2 3J
In summary, for V< Vm~,the stress state in the plastic region lies on the M o h r - C o u l o m b
line, and solutions are given in the previous section. For Vm,. < V< Vmax, there are stress
states in the plastic region that lie on both the M o h r - C o u l o m b and Tresca-limit yield
lines. If V> Vmax, the entire stress state in the plastic region lies on the Tresca-limit line•

Stress state has zones on both the M o h r - C o u l o m b and Tresca-limit yield lines
The mass conservation Eqns (8b) and (9) and the radial stress at the elastic-plastic
interface in the plastic region $2 given by (14b) are independent of the yield criteria, so
these equations apply for this analysis. By contrast, the momentum Eqn (8a) was derived
by using the M o h r - C o u l o m b yield criterion (4a), so (8a) requires that the stress state lies
on the M o h r - C o u l o m b line. For S(~) < S,. and V < Vm~x, S(~) can be calculated from
(15) or (16) for ¢,. < ~ < 1, where S(~,,) = S,.. The value of ¢,, is obtained by equating
Eqns (15) and (27) and setting ~ = ~,,. Thus,
(1 + Zo/2E) 2~ OcpoV2( )'312(1 - q*)(2 -- ~)~3 + 3),3] ]
[(1 - q*)#3m + ),312,/3 + ~ - ' ~ [ ( 1< - q*)(1 -- 2~)(2 - ~)[(1 - q*)~3 + ),3]4/3

+ Zo),Z[(1 _ r/.)~3 + y3]z,/3 + q* 1 - 2 E , I

_ y312(1 - q * ) ( 2 - g ) + 3y3] ~z_~_~= 0,


(30)
(1 - q*)(1 - 2~)(2 - ~)(1 + Z o / 2 E ) 4 J zo

where ~,, is found numerically.


434 M.J. FORRESTALand V. K. LUK

For small enough values of z,,/T o, there can be stress states on both the M o h r - C o u l o m b
and Tresca-limit yield lines for the quasi-static solution ( V ~ 0). For this special case, ~,,
can be found from (30) with V= 0.
For 0 < ~ < ¢,, the stress state lies on the Tresca-limit yield line and the yield criterion
(26) governs. Using the methods outlined in the previous section, radial stress for 0 < ¢ < ~.
is given by
S(~)=l('m~12(Tm'lln[(1-rI*)~+73]
~\Zo/-2 + 3\Zo/ +7
poV27 I [ 3~_3_+_ 4( 1__- _r/*)¢3 ] [373 + 4(1 - r/*)~3]}
+ 2 , o - ~ - - ~ / , ) [ [ ( i _ rt,)¢ s + 73].)/3 - [(--~ 25~/,)---~3.~+-3-5~ 3 . (31)

Radial stress at the cavity surface ¢ = 0 is given by

S(O) = A + BpoV2/zo (32a)

A=l(z_y_~ 1 2(z,.~ in [_1 + ( l _ q,)(~./7)3 ] (32b)


~x",'ro/- ~ + 3 \ t o /
l {3 [3 + 4(1 - r/*)(~,,,/7) 3] "~
B-2(1 - q* ) ~ + ~ ~. ~ - * ~ J " (32c)

In (32) A and B depend on material parameters and cavity-expansion velocity because ~.


depends on V.

Stress state lies entirely on the Tresca-limit yield line


If S 2 given by (14b) is larger than S m given by (27), the stress state in the plastic region
lies entirely on the Tresca-limit yield line. Radial stress in the plastic region is obtained
by the methods outlined in the previous section and given by

2 2(zm~ln[(l-r/*)~3+~'l
s(o 3\ o/ j

+ P°V---~z[3zo/E + r/*(l -- 3Zo/2E) 2]


723o
+ poV2~ ~[373+4(1--_r/*)~ 3] [373 + 4(1 - r/*)]'~
(33)
2To(1 - r/*)/.1(1 - r/*)~ 3 + 73] 4/3 ;
where Eqn (33) is valid for V> Vm.,. For this case, the material in the elastic region at
= 1 satisfies the yield criterion a, - go = to. However, for V> Vm., the stress state in
the plastic region lies on the Tresca-limit line and is governed by a, - a o = Zm, which is
(26). Radial stress at the cavity surface ~ = 0 is given by

S(O) = A + BpoV2/zo (34a)

(34b)
A = 1 + 7o/2E

3 [3zo/E + r/*(1 - 3To/2E) 2]


B- +
2(1 - r/*) 3,2

_
2(1 + L7 / 2 E ) 4 [ 1+
3(1--+-z°/2E)31"(1
- r/*) _] (34c)
Penetration into soil targets 435

NUMERICAL RESULTS FROM THE CAVITY-EXPANSION EQUATIONS

In this section, we present some numerical results from the cavity-expansion equations.
As previously discussed, our penetration equations use the relationship between radial
stress at the cavity surface (¢ = 0) and cavity-expansion velocity V, so Figs 4 and 5 present
crr(~ = 0) versus V. The calculations in Figs 4 and 5 take Po = 1.86 x 103 kg m -3, r/* = 0.13,
E = 160.0 M P a and TO = 10.0 MPa.
Figure 4 presents results for the Tresca (Fig. 3b) and M o h r - C o u l o m b (Fig. 3c) yield
criteria. To check our results, we calculated at( ~ = 0) from the equations for the special

i I I

30 =0 . 7 ~

20

or/ x0

10

I I I
1 2 3 4
(PO / IO) I12 V

FIG. 4. Radial stress at the cavity surface versus cavity-expansion velocity for the Tresca and
M o h r - C o u l o m b criteria.

30 I I I

Crr/ TO20 /-

, MOHR-COULOMB o
TRESCA

0 I I I
0 1 2 3 4

(p0/ t0)1/2V
FIG. 5. Radial stress at the cavity surface versus cavity-expansion velocity for the Tresca and
M o h r - C o u l o m b , Tresca-limit criteria.
436 M.J. FORRESTALand V. K. LUK

cases when 2 = 0 and 0.75 and then used the more general M o h r - C o u l o m b equations and
calculated a,(~ = 0) for 2 = 0.0001 and 2 = 0.74, 0.76. No differences in the plotted results
could be detected between the 2 = 0 and 2=0.0001 results, and the 2 =0.74,
0.76 calculations closely bracketed the 2 = 0.75 calculations.
Figure 5 compares results from the equations for the Tresca yield criterion (Fig. 3b,
Zo = 10 MPa) and the M o h r - C o u l o m b with a Tresca-limit criterion (Fig. 3d, z o = 10 MPa,
r,, = 20 MPa, 2 = 0.375). As previously discussed, the latter model has several response
regions that depend on the cavity-expansion velocity. For V/(zo/Po) 1/2 < 1.04, the stress
state in the plastic region lies on the M o h r - C o u l o m b yield line and is the same as that
presented in Fig. 4 for 2 = 0.375; for 1.04 < V/(zo/Po) 1/2 < 2.05, the stress state in the
plastic region lies on both the M o h r - C o u l o m b and Tresca-limit lines; and for
V/(ro/Po) 1/2 > 2.05, the stress state in the plastic region lies on the Tresca-limit line.

PENETRATION EQUATIONS
Axial force on the projectile nose is given by the integral Eqn (2a). This integral contains
tr,( V..,4~), which is normal stress on the ogival nose from target resistance. We approximate
these stresses by using results from the spherically symmetric, cavity-expansion analyses
that relate radial stress at the cavity surface to cavity-expansion velocity. From the nose
geometry shown in Fig. 1, the particle velocity at the ogival nose-target interface caused
by projectile, rigid-body velocity V: is
v(V_., ~b)= Vzcos ~b. (35)
We now approximate normal stress distribution on the ogival nose by replacing the
spherically symmetric, cavity-expansion velocity V with the particle velocity v given by
(35). From (18) and (35), normal stress distribution around the ogival nose is taken as
trn( Vz, C~) = zoA + p o B ( ~ c o s q~)2. (36)
Substituting (36) into (2a) and integrating gives
(37a)
cq = rca2zoA[1 + 4#¢2(rc/2 - 0o) - #(2¢ - 1)(4¢ - 1) l/z] (37b)
2 #(2¢- 1)(6¢ 2 + 4¢ - 1)(4¢ - 1) ~/2]
#, = po,.[-(8¢
l_ ~-4~---2 1) + .02( 12 - 0o)- 24¢ 2 J
(37c)

0o sin (37d)

where ¢ is the caliber-radius-head for the ogive given by (1). Equations (36) and (37) were
derived in [7].
In the field tests, we measure rigid-body deceleration with an onboard, digital-recording
system. In addition, velocity-time and displacement-time are obtained by integrating the
acceleration-time data. We derived equations for force on the ogival nose, so we now
present equations for projectile motion. From Newton's law and (37a)

m-- = + #sVz ), (38)


dt
where m is the projectile mass. Integration of (38) with the initial condition V~(t = O) = Vo
gives

= 121 tan tan-~ . (39)


Penetration into soil targets 437

Acceleration-time is
dV~ (~,/m)
c°s {ta (40)
a -- - - --

dt

and displacement-time is

z= rnlnF c°s { tan--~-[(fl'/~')'/2V°] --(~'fl')uZt/m} l


(41)
~ [_ cos{tan-'[(fl,/~,)'/2Vo]} 3"
Final depth of penetration P is given by

P = m In(1 + /~v g) . (42)


2fl~ \ ~

SOIL MATERIAL AND PENETRATION TESTS


We conducted penetration tests at the Sandia T o n o p a h Test Range, Nevada, into a
soil target at a site called Antelope Lake. Figure 6 is a m a p that shows the locations of
the six penetration tests and the three cores taken for soil material properties tests.

Soil tests
Kraatz [ 9 ] obtained three soil cores to a depth of 6.1 m and performed standard soil
analyses from the core labeled 5 in Fig. 6. The cores labled 3 and 4 were sent to the
Geomechanics L a b o r a t o r y of Waterways Experiment Station, Vicksburg, MS, where
Ehrgott and Kinnebrew [ 10] conducted standard soil analyses and triaxial material tests.

20m 10m 10m 20m


I I I I

20m
50

3• 5_1 40m 4 [] 30

60m

40
-- 80m

O Penetration Tests
[] Soil Cores

100m

FIG. 6. Map showing the locations of penetration tests and soil cores.
438 M.J. FORRESTAL and V. K. LUK

30

25

Q.
a..
¢/)
co
LM
rr
O.
20

0
0.0 0.5 0.10 0.15 0 20

VOLUMETRIC STRAIN, q

FIG. 7. Pressure versus volumetric strain for the soil target.

Based on density and water content, Ehrgott and Kinnebrew [ 10-1 qualitatively divided
the soil depth of 6.1 m into three layers: layer 1 (0-1.8 m), layer 2 (1.8-3.0 m), and layer
3 (3.0-6.1 m).
As previously mentioned, analytical models require triaxial material test data of samples
cored from the target area. These tests are conducted in two phases: isotropic compression
followed by triaxial shear. Soil cylinders are loaded with axial stress tr, and radial stress
tr,. For isotropic compression tr~--¢r,, so pressure is given by p = 1/3 (a, + 2a~). For
isotropic compression or uniform pressure, deformation measurements of the length and
diameter changes during loading were performed and volumetric strains were calculated.
Figure 7 shows pressure versus volumetric strain data for samples taken from the three
layers. For the triaxial shear tests, the soil cylinders are first loaded with isotropic
compression, and then the axial stress fro is increased while the radial stress ~L is held
constant. Values of the maximum stress difference z = (tr= - ar) or shear strength versus
pressure p = 1/3 (or, + 2try) are given in Fig. 8. Data in Fig. 8 show that shear strength is
nearly constant, so we later use the Tresca yield criterion (Fig. 3b) for penetration modeling.
In addition, one-dimensional strain tests were conducted [ 10]. For these tests e, = 0,
and fro, ¢r,, and e° were measured. We estimated Young's modulus E from the small strain
data from these one-dimensional strain tests. For an incompressible elastic material,
(tr~ - try) = ( 2 / 3 ) E e = . For layer 1, E = 120 MPa; for layer 2, E = 150 MPa; and for layer
3, E = 210 MPa.

Penetration tests
We conducted six penetration tests with a mobile, 152 mm diameter, smooth-bore gas
gun [ 11 ]. Figure 9 shows the projectile geometry with a ~b = 3 (caliber-radius-head) ogival
nose and mass m = 23.1 kg. We designed the projectile with an internal supporting structure
that attached an onboard recording package to the walls of the projectile. Location of the
a c c e l e r o m e t e r s is shown in Fig. 9. Some of the tests results are summarized in Table 1,
where Vo is the impact velocity, ~b1 is the angle of obliquity, ~b2 is the yaw angle and P is
the path length of the projectile. Definitions of ~b1 and ~b2 are given in [12].
Penetration into soil targets 439

25
Z~ * LAYER 1

FI - LAYER 2

O - LAYER 3
20 - - TRESCA IDEALIZATION

"I-
A
I-- O
L~ 0
z 10 H o V1
uJ
n,,, o 0
t-
X-
w 5

ffl

0 I I I I I
0 10 20 30 40 50 60

PRESSURE, p (MPa)

FIG. 8. Shear-strength versus pressure for the soil target.

--~- ACCELEROMETER LOCATIONS

- - 158 mm

95.2 m r n ~

525 mm

FIG. 9. Projectile geometry and accelerometer locations.

TABLE 1. TEST RESULTS

Test Vo ~bl, q~2 P


number (m s - l) (degrees) (m)

1 280 0,0 6.48


2 280 0,0 4.98
3 278 30,0 5.18
4 280 30,0 5.18
5 280 0,3.5 5.02
6 280 0,4.0 4.82

Modeling for this study is limited to penetration at normal incidence, so angles of


obliquity and yaw were neglected when we later compared data and predictions. For Tests
3 and 4 with ~b1 = 30 °, we measure and calculate path length. As shown by the data in
Table 1, the impact velocity for all tests can be taken as Vo = 280 m s- 1. The penetration
depths from Tests 2 - 6 are in close agreement with an average depth or path length of
5.04 m. Thus, the penetration depth from Test 1 is about 30% greater than for the other
five tests. The map in Fig. 6 shows that Test 1 was conducted at the north end of the site.
With the penetration depths in close agreement away from the location of Test 1, we
assumed that the soil data taken from the cores were representative for Tests 2 - 6 ( see Fig. 6).
440 M.J. FORRESTALand V. K. LUK

The 95.2 mm diameter projectile shown in Fig. 9 was accelerated to an impact velocity
of 280 m s-1 with a 152 mm diameter gas gun [111. Since the projectile had a smaller
diameter than the gun bore, the shank of the projectile was fitted with two 0.35 kg foam
bore-riders to keep the projectile in proper alignment during launch. In addition, a
pusher-plate was attached to the base of the projectile. For Tests 1-4, the pusher plate
was a 0.73 kg aluminum disk and the centerline of the projectile was located along the
centerline of the gun bore. For Tests 5 and 6, the pusher-plate was a 1.0 kg phenolic disk,
and the centerline of the projectile had a 4 ° angle with the bore centerline. As indicated
in Table 1, two tests were conducted with q~l = q~2 = 0, two tests were conducted with
q~l = 30 ° and q~2 = 0, and two tests were conducted with q~x = 0, and ~b2 = 3.5 and 4.0 °.
Measured values of Vo and q~ and ~b2 in Table 1 were obtained from streak and framing
camera data.
In addition to the data in Table 1, we measured axial deceleration with an onboard
recorder and two accelerometers (see Fig. 9). The data recording system had a 3.5 kHz
resolution [131; but this study is concerned with rigid-body deceleration, so data were
filtered to 1.0 kHz. The 1.0 kHz filter removed much of the higher frequency responses
associated with the vibration of the projectile case and internal supporting structure that
supported the instrumentation package. When filtered to 1.0 kHz, no difference could be
detected in the plotted results for the forward and aft deceleration records. Deceleration-time
data are presented in the next section and compared with model predictions. Reports by
Wood [ 131 present more detail on the instrumentation system and deceleration data.

COMPARISONS OF MODEL PREDICTIONS AND PENETRATION DATA


Post-test, we observed a crater about one nose-length long followed by a circular tunnel
nearly the diameter of the projectile shank. Because the projectiles traveled about thirty
nose-lengths, we neglected the crater phase of penetration in the model predictions. In
addition, the target stripped the bore-riders and pusher plate from the projectile. Since the
projectile had mass 23.1 kg and the total mass of the bore-riders and pusher plate was
1.4kg, we ignored the added mass of the bore-riders and pusher plate in the model
predictions.
Our penetration equations (39)-(42) depend on the projectile mass (23.1 kg), shank
diameter (95.2 ram), nose shape (~ = 3.0) and material properties of the soil target. For a
previous study on penetration into dry porous rock with water content less than 1% [ 141,
we observed that the projectile nose lost material through melting, so we included the
effect of interface friction in the modeling of those test results. For these tests, the target
had a water content of about 20% and we did not observe any nose erosion. Thus, for
modeling the test results in this study, we neglect interface friction between the projectile
nose and the soil target (/~ = 0). From (37a), (37b) and (37c) with ~ = 0, force on the
projectile nose reduces to

Fz=% + flsV~=r:a 2 zoA + ~ ) B p o V ~ . (43)

As indicated by the data in Fig. 8, the shear strength of the soil target can be closely
approximated with the Tresca yield criterion (Fig. 3b), so we use (21a) and (21b) for A
and B, which are dimensionless quantities that depend on ~/* and %/E.
As previously discussed, Ehrgott and Kinnebrew [ 101 qualitatively divided the cored
soil depth of 6.1 m into layer 1 (0-1.8m), layer 2 (1.8-3.0m) and layer 3 (3.0-6.1 m).
Figure 7 shows the pressure-volumetric strain data and the locking solid idealization
values of r/* (Fig. 3a). For layers 1, 2 and 3, we take r/*=0.17, 0.10 and 0.13,
respectively. Figure 8 shows the shear strength-pressure data, which we model with the
Tresca yield criterion (Fig. 3b) and take % = 10.0 M P a for all three layers. In addition,
values of E for layers 1, 2 and 3 are E = 120, 150 and 210 MPa, respectively. We show
later that deceleration-time predictions have a weak dependence on r/* and E, so we take
Penetration into soil targets 441

500 I I I I I I I I I I I I

l
t-
O -500

a)
¢..) -1000
u

-1500 I I I I I I I I I I t I
-5 0 5 10 15 20 25 30 35 40 45 50 55 60

Time (ms)

(a)

500 I I
I I I I I I I I l I

0 Ix__

'i -500

"~ -1000

-1500 I I I I I I I I I I I I
-5 0 5 10 15 20 25 30 35 40 45 50 55 60

Time (ms)
(b)
FIG. 10. Model predictions and penetration deceleration data for Tests 1 and 2.

average values and use Po = 1.86 x 103 kg m-3; z o = 10.0 MPa, q* = 0.13 and E = 160.0 M P a
for the predictions that we compare with measurements. Figures 10, 11 and 12 show
predictions and deceleration data filtered to 1.0kHz. Other than Fig. 10a for Test 1
(discussed in the previous section), rigid-body deceleration predictions are in good
agreement with measurements. Neglecting the low-amplitude oscillations in the data that
we associate with structural vibrations caused by target impact and the nonhomogeneous
soil target, deceleration measurements and predictions show a monotonic decay to a nearly
constant value followed by a j u m p at the end of the trajectory.
In addition to the predictions shown in Figs 10, 11 and 12, we calculated three
deceleration-time profiles with values of r/* corresponding to the three soil layers and held
the other three material parameters fixed at Po = 1.86 x 10 3 k g m -3, r o = 10.0 M P a and
E = 160.0 MPa. Table 2 compares m a x i m u m decelerations at t = 0, minimum decelerations
at the end of the event t = t I, and final penetration depths P. Thus, q* is a weak parameter,
and Table 2 justifies using a single value of r/* --- 0.13 for predictions in Figs 10, 11 and 12.
We further justify the locking solid idealization for the pressure-volumetric strain data
(Fig. 7) by calculating the minimum pressure on the projectile nose when V~ = 0. From
(18), (21), (3b) and (3c) with 2 = 0 , Z o = 10.0MPa, r / * = 0 . 1 3 and E = 160MPa,
p = 10.5 MPa, so our idealization is for p > 10.5 MPa.
We also calculated three deceleration-time profiles with the values of E corresponding to the
three soil layers and held the other three material parameters fixed at Po = 1.86 x 103 kg m-3,
z o = 10.0 M P a and q* = 0.13. Table 3 shows that E is a weak parameter and justifies using
a single value of E = 160.0 M P a for the predictions in Figs 10, 11 and 12.
For 1/* = 0.13, E = 160.0 M P a and To = 10.0 MPa, our model predicts a final penetration
depth of P = 4.98 m, which is in good agreement with the data in Table 1 for Tests 2-6.
442 M.J. FORRESTALand V. K. LUK

500 I I I I I I I I I

0 •"% ~ ~v A^

C
o -500

w -1000
o
o
,<
-1500 I I I I I I I
-5 0 5 10 15 20 25 30 35 40 45 50

Time (ms)

(a)

500 I I I I I I I I I

0 r-!

C
o -500

"~ -1000
0
0
,<
-1500
-5 0 5 10 15 20 25 30 35 40 45 50

Time (ms)

(b)
FIG. 1 1. Model predictions and penetration deceleration data for Tests 3 and 4.

500 I I I I I I I

A
\i
i -500 -

-1000
0
U

I I I I I I I I I I
-1500
-5 0 5 10 15 20 25 30 35 40 45 50

Time (ms)
(a)

500 k 1 I I I I I I ~ 1

0
A

C
.~ -500

• -1000
0
0
,<
-1500 I I I I I I I I I I
-5 0 5 10 15 20 25 30 35 40 45 50

Time (ms)

(b)
FIG. 12. Model predictions and penetration deceleration data for Tests 5 and 6.
Penetration into sod targets 443

TABLE 2. MODEL SENSITIVITY RESULTS WITH VALUES OF r/*

Maximum Minimum t/ P
q* deceleration (g) deceleration (g) (ms) (m)

0.10 1187 569 39 4.75


0.13 1137 540 41 4.98
0.17 1083 506 43 5.27

TABLE 3. MODEL SENSITIVITY RESULTS WITH VALUES OF E

E Maximum Minimum t/ P
(MPa) deceleration {g) deceleration (g) (ms) (m)

120 1128 517 42 5.10


160 1137 540 41 4.98
210 1145 558 40 489

CONCLUSIONS

W e d e v e l o p e d p e n e t r a t i o n e q u a t i o n s for o g i v a l - n o s e projectiles that p e n e t r a t e soil targets.


T a r g e t constitutive m o d e l s idealized p r e s s u r e - v o l u m e t r i c strain as a l o c k i n g solid, whereas
the M o h r - C o u l o m b , T r e s c a a n d M o h r - C o u l o m b with a Tresca-limit yield criteria idealize
the s h e a r - s t r e n g t h - p r e s s u r e behavior. W e d e r i v e d e q u a t i o n s for the spherically symmetric,
c a v i t y - e x p a n s i o n p r o b l e m for the three s h e a r - s t r e n g t h m o d e l s a n d used these results to
d e v e l o p o u r p e n e t r a t i o n equations.
T o verify o u r p e n e t r a t i o n e q u a t i o n s , we c o n d u c t e d p e n e t r a t i o n tests into soil targets
a n d o b t a i n e d r i g i d - b o d y d e c e l e r a t i o n s a n d final p e n e t r a t i o n depths. In a d d i t i o n , we t o o k
soil cores from the test a r e a a n d o b t a i n e d the triaxial m a t e r i a l d a t a required for the
constitutive model. W e idealized the shear strength d a t a in Fig. 8 with the Tresca yield
c r i t e r i o n (Fig. 3b), so we used the m o s t simple of the d e s c r i p t i o n s for shear strengths s h o w n
in Fig. 3. O t h e r soil targets m a y require the shear strength d e s c r i p t i o n s s h o w n in Figs 3c
a n d 3d.

Acknowledgements--This work was supported by the Joint DoD/DOE Munitions Technology Development
Program. The authors thank R. G. Lundgren and W. R. Wood for conducting the field tests.

REFERENCES

1. M. E. Backman and W. Goldsmith, The mechanics of penetration of projectiles into targets. Int. J. Engng
Sci. 16, 1-99 (1978).
2. C. W. Young, Depth prediction for earth penetrating projectdes. ASCE J. Soil Mech. Found., 95 (SM3),
803-817 (1969).
3. C.W. Young and E. R. Young, Simplified Analytical Model of Penetration with Lateral Loading, SAND84-1635,
Sandia National Laboratories, Albuquerque, NM (1985).
4. R. F. Bishop, R. Hill and N. F. Mott, The theory of indentation and hardness tests. Proc. Physical Society
57 (3), 147-155 (1945).
5. V. K. Luk and M. J. Forrestal, Penetration into semi-infinite reinforced concrete targets with spherical and
ogival nose projectiles. Int. J. Impact Engng 6 (4), 291-301 (1987).
6. V. K. Luk and M. J. Forrestal, Comment on "Penetration into semi-infinite reinforced-concrete targets with
spherical and ogival nose projectiles". Int. J. Impact Engng 8 (1), 83-84 (1989).
7. 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).
8. I. S. Sandier, F. L. Di Maggio and G. Y. Baladi, Generalized cap model for geological materials. J. Geotechn.
Engng Division, ASCE, GT7, 683-699 (1976).
9. P. R. Kraatz, Subsurface Investigation Ballistic Experiments at the Tonopah Test Range, Nye County, Nevada,
SHB Job No. E90-8039, Sergent, Hauskins, and Beckwith, Reno, Nevada (1990).
444 M.J. FORRESTALand V. K. LUK

10. J. Q. Ehrgott and S. L. Kinnebrew, Mechanical Response oJ Subsurface Soils from Tonopah Test Site, Waterways
Experiment Station, Vicksburg, MS (1991).
11. R. G. Lundgren, M. M. Hightower and B. K. Christensen, Development of a mobile gas gun. Proc. 39th
Aerobalhstic Ran#e Association, Vol. 2 (1988).
12. J. A. Zukas et al., Impact Dynamics. pp. 203-206. Wiley, New York.
13. W. R. Wood, MOU Penetrator Acceleration Data, Test Numbers MOU 65-70, Sandia National Laboratories
Test Reports, July (1990).
14. M J. Forrestal, Penetration into dry porous rock. Int. J. Solids Struct. 22 (12), 1485-1500 (1986).

Potrebbero piacerti anche