Sei sulla pagina 1di 10

International Journal of Impact Engineering 37 (2010) 918–927

Contents lists available at ScienceDirect

International Journal of Impact Engineering


journal homepage: www.elsevier.com/locate/ijimpeng

The correct analysis of shocks in a cellular material


J.J. Harrigan a, *, S.R. Reid a, A. Seyed Yaghoubi b
a
School of Engineering, Fraser Noble Building, King’s College, University of Aberdeen, Aberdeen AB24 3UE, UK
b
School of Mechanical, Aerospace & Civil Engineering, Pariser Building, University of Manchester, PO Box 88, Manchester M60 1QD, UK

a r t i c l e i n f o a b s t r a c t

Article history: Cellular materials have applications for impact and blast protection. Under impact/impulsive loading the
Received 16 September 2008 response of the cellular solid can be controlled by compaction (or shock, see Tan et al. (2005) [3,4]) waves.
Received in revised form Different analytical and computational solutions have been produced to model this behaviour but these
19 March 2009
solutions provide conflicting predictions for the response of the material in certain loading scenarios. The
Accepted 23 March 2009
Available online 3 May 2009
different analytical approaches are discussed using two simple examples for clarity. The differences
between apparently similar ‘‘models’’ are clarified. In particular, it is argued that mass-spring models are
not capable of modelling the discontinuities that exist in a compaction wave in a cellular material.
Keywords:
Cellular material Ó 2009 Elsevier Ltd. All rights reserved.
Compaction wave
Shock wave

1. Introduction models have been applied to certain loading scenarios but with
different formulations [12–14] of the ‘‘shock theory’’. The different
Cellular materials (e.g. foams, honeycombs, etc.) are employed analytical approaches have led to some debate, e.g. Refs. [15–17].
in many applications due to their multi-functional properties, The aim of this article is not to present a new method. Rather it is
including insulation and low weight, especially for packaging and in to provide clarity in a very confused area by comparing three
light-weight sandwich panels. One of their applications is as an existing methods for the analysis of dynamic compaction of cellular
energy absorbing layer for impact and blast protection. The cellular materials. These methods are: (i) ‘‘shock theory’’; (ii) ‘‘energy
materials considered here are those that are characterised by conservation’’; and (iii) mass-spring models. The comparison is
a relatively flat plateau stress until densification. As such, they are carried out by considering the solutions to two idealised loading
often considered to be ideal energy absorbers because of their low scenarios (Examples 1 and 2 defined later). The different assump-
weight and their ability to deform over a long stroke at an almost tions inherent to each method are shown to lead to large differences
constant load (see Fig. 1). During quasi-static compression the between the predictions.
energy absorbed per unit volume is equal to the area under the Reid and Peng [1] provided the first ‘‘shock-wave’’ predictions
nominal stress–strain curve, i.e. the cellular material is considered for cellular solids to explain certain experimental results, focusing
as a continuum, not a material comprising the edges and faces of the on the enhancement of the crushing strength of wood specimens.
actual cell structure. However, under intense dynamic loads the For simplicity, the cellular solid was idealised as a rigid, perfectly-
cellular structure plays a key role and compaction waves propa- plastic, locking (R-P-P-L) material. Developments of the same
gating through the cellular array play an important role in the simple shock theory have been applied with more accurate mate-
response of cellular solids. The compression is localized at discon- rial models in order to predict better the stresses during direct
tinuity (shock) fronts and deformations propagate through the cells impact testing of cellular solids. For example, an elastic, plastic,
in a progressive manner. Various methods have been proposed to hardening material model was used in Refs. [5,6]. Lopatnikov et al.
model ‘‘shock-wave’’ (more precisely ‘‘compaction front’’) propa- [8,9] used an ‘‘elastic-perfectly-plastic-rigid’’ model and shock-
gation in cellular solids, e.g. one-dimensional shock-wave models wave analysis to predict the deformation of aluminium foam in
(e.g. see Refs. [1–9]), spring-mass models [10,11] and FE modelling both foam projectile [8] and target [9] tests. For the simple ‘‘shock’’
(e.g. see Refs. [7–9]). In addition one-dimensional ‘shock-wave’ wave analysis used in Refs. [1–4] the strain is equal to one of two
possible values, viz. zero or a value associated with the fully locked
material. In Refs. [8,9] the effect of elasticity is incorporated but the
* Corresponding author. Tel.: þ44 (0)1224 272512; fax: þ44 (0)1224 272497. strain within the compacted material is again given a value asso-
E-mail address: j.harrigan@abdn.ac.uk (J.J. Harrigan). ciated with the fully locked material, although the locking strain is

0734-743X/$ – see front matter Ó 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijimpeng.2009.03.011
J.J. Harrigan et al. / International Journal of Impact Engineering 37 (2010) 918–927 919

4.E+06 l0

E-P-P-H
R-P-P-L , A0
3.E+06 G 0
[10] a) t < 0
v0

2.E+06 u x

1.E+06 G
P b) t D , D P , 0
v

0.E+00 u+ u x+ x
0 0.2 0.4 0.6 0.8 1
D
Y

Fig. 1. Stress–nominal strain from Ref. [10] together with R-P-P-L and E-P-P-H material G
models. c) t + t

v+ v
defined in a different way. However the basic assumptions in the
analysis remain and provide good agreement with FE predictions
[8,9]. The predictions in Refs. [1–9] differ only in detail (see
Ref. [17]) and will be grouped together as ‘‘shock-wave’’ models. Fig. 2. Example 1: a mass G impacting a foam column.
FE modelling of the dynamic compaction of cellular solids has
been carried out in two ways. In certain investigations the material
has been treated as a continuum [7–9]. In other simulations which will be the subject of a future paper. In the R-P-P-L model
approximations to the actual structure of either two-dimensional quasi-static material behaviour is characterised by two parameters,
[4,20,21] or three-dimensional [22] cellular solids are modelled. As i.e. the plateau stress sP and the locking (densification) strain 3D.
continuum-based FE predictions have been shown to agree with Example 2 builds on the theory developed in Example 1 to
‘‘shock-wave’’ theory, they are discussed only briefly here. explain the large differences between the predictions of mass-spring
An alternative solution to dynamic loading of foams wherein the models described in Refs. [10,11] and ‘‘shock theory’’. In Example 2,
deformation is governed by a compaction wave was reported in the boundary condition corresponding to the impact of a rigid mass
Ref. [12] giving the equation of motion and energy absorption for on the foam column (Example 1) is replaced by a constant pressure
the case of a rigid mass striking a foam column. This formulation applied over a short duration. The intensive pressure pulse load from
has since been repeated or used in Refs. [13,14,18], and discussed in a blast is another possible source for shock propagation in a cellular
Refs. [15,16]. In Refs. [12,13] the case of a rigid mass striking a foam material. Li and Meng [10] studied the ‘‘compressive shock-wave
layer with the distal face fixed rigidly was considered. In Ref. [14] propagation in the solid phase of a cellular material’’ using a one-
the analysis was extended to that of a sandwich panel by consid- dimensional mass-spring model. A foam column was subjected to
ering the overall motion of both facesheets and the compression of a rectangular pressure pulse at one end (simplified ‘blast’ loading)
the internal foam layer when the panel is subjected to blast. The while the other end was fixed to a rigid support. Stress enhancement
procedure used in Ref. [14] was then applied to a clamped, circular, (i.e. stresses greater than the applied pressure [10]) within the foam
sandwich plate in Ref. [18] to predict the energy absorbed and was predicted during the propagation of the compaction wave
timescales associated with the core compression stage. through the cellular material. Further, a series of pressure pulses
Tan et al. [4] argued that the equations of motion derived in were considered in order to predict the critical pulses delivered to
Ref. [12] are incorrect. The ‘‘steady-state shock’’ equations in Ref. [4] the foam core, beyond which the maximum pressure pulse applied
were derived from a thermo-mechanical approach. Whilst providing to the rigid support was greater than the original, applied pressure.
useful insights, this approach clouds the simplicity of a derivation An R-P-P-L idealisation is again used to obtain closed-form
based on a purely mechanical approach using conservation of mass solutions. A more realistic elastic, perfectly-plastic, hardening (E-P-
and momentum only. In this paper, this simpler approach is used in P-H) material description is used later for one loading scenario for
order to illustrate the essence of the ‘shock’ model for impact/blast comparison with the R-P-P-L model and the results in Ref. [10].
loading and to clarify certain features of the solution of such problems Gao and Yu [11] used a one-dimensional mass-spring system to
that have appeared in the recent literature, two examples are pre- model the response of a cellular material to a short duration pressure
sented. In Example 1, conservation of momentum is applied to predict pulse in order to investigate the effect of material properties on the
the compaction of an energy absorbing foam column which is struck response. The speed of the compaction wave predicted by the mass-
by a rigid mass as shown in Fig. 2, i.e. the same example as in Ref. [12]. spring model is shown to agree well with that predicted by ‘‘shock
In doing so, it will be made clear that the solutions based on ‘‘shock- theory’’ [11]. However, typical pressures considered in Ref. [10] were
wave’’ models [1–9] are fundamentally different to those based on the 2–5 times the plateau stress in the material whereas in Ref. [11] the
formulations in Refs. [12–14,18]. The R-P-P-L idealisation is used for maximum pressure considered is only 2.5 times the ‘‘collapse stress’’
Example 1. The aim is to derive the governing equations in a simpler of the material. Additionally, Li and Meng provide an attenuation/
way than in Ref. [4] as well as to illustrate the differences between the enhancement boundary in Ref. [10] (discussed in Example 2) that is
‘‘shock-wave’’ theory and the predictions in either Refs. [10,11] or not provided in Ref. [11]. For these reasons, the predictions in
Ref. [12], rather than to predict more accurately experimental data, Ref. [10] are used for comparison in Example 2.
920 J.J. Harrigan et al. / International Journal of Impact Engineering 37 (2010) 918–927

2. Example 1: compaction wave in cellular solid subjected by applying the classical Rankine–Hugoniot relationships (e.g. see
to impact by a rigid mass Ref. [4]). All the referenced works use this equation in one form or
another. However, after this point different models diverge. In Section
Under a high speed impact or high magnitude impulsive load, 3.1, conservation of momentum is applied to the whole system to
a cellular material deforms in a progressive manner with the cells derive an equation of motion. In Section 3.2, the alternative ‘‘energy
adjacent to the loaded surface compacting first and the deformation balance’’ approach is employed to derive a different equation of motion.
then passing through the material in a wave-like manner. Whilst this
problem has been treated previously (e.g. Refs. [6,7]), for clarity, the
steps involved in developing the basic theory are repeated below. 3. Two different solution schemes in the literature
The dynamic crushing model for a rigid projectile of mass G
striking an R-P-P-L target of mass Mcore, initial length l0, cross- 3.1. Scheme 1: ‘‘Conservation of momentum’’ approach
sectional area A0 and density r0 is illustrated in Fig. 2. The mass
G has an initial velocity v0 and strikes the cellular column at time Conservation of momentum for the whole system shown in Fig. 2
t ¼ 0. Fig. 2(a) shows the situation just before impact while Fig. 2(b) over the period t to t þ dt gives
and (c) illustrates the compression at times t and t þ dt. As cellular
solids can undergo large plastic compression in one direction ½G þ rD ðx þ dxÞA0 ½v þ dv  ½G þ rD xA0 v ¼ sP A0 dt: (6)
without lateral expansion, the target is assumed to deform under Thus, substituting Eqs. (1) and (2) into Eq. (5), the governing
uniaxial compressive strain conditions. As the compaction wave equation of motion is
travels through the material a strong discontinuity is assumed to
  
exist between the deformed and undeformed regions. In the dv d2 u dv r0 2 G r0
undeformed region ahead of the wave, the stress is equal to the ¼ ¼ v ¼  sP þ v þ u : (7)
dt dt 2 du 3D A0 3D
plateau stress sP (essentially assuming an infinite elastic wave
speed) while the strain and particle velocity are zero. After Integrating Eq. (7) with the initial condition v ¼ v0 at u ¼ 0 gives
compression the stress can take any value. The compacted region vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u "   2  2 ! #ffi
travels with the same (reducing) velocity as the mass and has been u3D r G G r
compressed to the densification strain, 3D. The stress in this region v ¼ t 0
v2 þ sP þ u 0
 sP : (8)
r0 3D 0 A0 A 0 3D
is now considered.
From kinematic considerations it can be shown in Ref. [1] Thus for sD > sP, from Eqs. (5) and (8) the stress just behind the
that the speed of the compaction front relative to the projectile compaction front can be found as a function of the displacement of
(see Fig. 2) is the rigid mass, i.e.
dx 1  3D   2  2
¼ v; (1) r0 2 G G r0
dt 3D sD ¼ v þ sP þ u : (9)
3D 0 A0 A0 3D
where v is the instantaneous velocity of the rigid mass with respect
The compaction front will continue to propagate through the
to a stationary frame. Conservation of mass [1] gives
foam until either the mass is brought to rest or the compaction
r0 x front reaches the distal end, which results in rigid impact with the
¼ 1  3D ¼ ; (2) support for the R-P-P-L shock model. The impact velocity, vD, for
rD xþu
which the cellular specimen is just fully crushed before the end
where r0 is the initial density, rD is the density when fully crushed, mass is brought to rest, can be found by setting the initial velocity
u is the displacement of the rigid mass at time t, and x is the as v0 ¼ vD and the final velocity as v ¼ 0 when u ¼ umax in Eq. (8), i.e.
deformed length of the crushed cellular material. Over a small time
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
" ffi
interval dt (see Fig. 2) the compaction wave travels along the u
usP 3D  2 #
A r u
¼ t
max
column so that a small element of material of mass dm is added to vD 1þ 0 0 1 : (10)
the front of the compacted portion which increases in length by r0 G3D
a small amount dx. The mass of this element is
The maximum displacement of the proximal end is umax ¼ l0 3D ,
where l0 is the original length of the column. Therefore, Eq. (10) can
r0 A0 dx
dm ¼ : (3) be written in terms of the mass ratio for full specimen locking as
1  3D
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
" #
u
usP 3D  Mcore 2

Over the time interval dt this element has stresses sD and sP on vD ¼ t 1þ 1 ; (11)
its left and right faces respectively, where sD is the stress just inside
r0 G
the compacted region. The element has zero velocity at time t and
where Mcore is the mass of the cellular material cylinder. It follows
a velocity v at time t þ dt. Conservation of momentum for the
that for full locking of the specimen, the ratio of the specific energy
element of mass dm over the time increment dt during which it
absorbed under conditions where the response is governed by the
undergoes compaction gives
propagation of a compaction (structural shock) wave to that under
ðsD  sP ÞA0 dt ¼ vdm: (4) quasi-static compression is given by

Substituting Eqs. (1) and (2) into Eq. (4) leads to Mcore
j¼ þ 1: (12)
2G
r0 2
sD ¼ sP þ v : (5) Eq. (12) shows that for impacting masses with equal kinetic
3D
energy but different masses (and impact velocities), the ratio of the
This is the original, simple, stress enhancement formula (cf. [1]) for specific energy absorption can be defined in terms of the mass ratio
the stress just inside the discontinuity and can also be derived only. This clearly demonstrates the shock enhancement in the
J.J. Harrigan et al. / International Journal of Impact Engineering 37 (2010) 918–927 921

dynamic energy absorbing mechanism. Additionally, from Eq. (11) is incorrect. In Scheme 1 no assumptions are made regarding the
we can find the critical length of the foam column required to just energy absorption capacity of the column of cellular material whereas
bring a mass G with an impact velocity v0 to rest as this was done in Scheme 2. Energy is conserved in Scheme 1 and not
0sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1 in Scheme 2 despite the ‘‘energy balance’’ approach used, as can be
G @ r0 v20 illustrated by considering the hatched element in Fig. 2. Over the time
lCR1 ¼ 1þ  1A: (13) increment dt, the work done on this element by external forces is
r0 A0 sP 3D
WE ¼ sD A0 du; (20)
3.2. Scheme 2: ‘‘Energy balance’’ approach and the gain in kinetic energy of the element is

However a different development of the equations is found in 1 A0 r0 v2 dx


Refs. [12–14]. Assuming that the energy absorbed during the TE ¼ : (21)
2 ð1  3D Þ
dynamic compression is equal to that for quasi-static compression
as in Refs. [12–14], ‘‘conservation of energy’’ for the element of mass By conservation of energy, the energy absorbed by plastic
dm over the time increment dt during which it undergoes deformation per unit volume of the element is therefore
compaction (Fig. 2) gives
1
EV ¼ ðs þ sP Þ3D : (22)
r
sD ¼ sP þ 0 v2 : (14) 2 D
23D
Eq. (22) is consistent with the Scheme 1 solution and can be
A similar ‘‘energy conservation’’ approach for the whole system derived from the Rankine–Hugoniot jump conditions, as has been
between times t and t þ dt (Fig. 2) gives, done elsewhere [4]. The key point is that the conservation equations
that should be applied when dealing with wave motion that involves
1 discontinuities in density, particle velocity and stress, refer to initial and
½G þ rD A0 xvdv þ rD A0 v2 dx ¼ sP 3D A0 dl: (15)
2 final values across the shock and are not a function of the path between
Substituting Eqs. (1) and (2) into Eq. (15), the governing equa- the initial and final values. This creates no problem for equations
tion of motion is involving mass and momentum conservation. The assumption in
Scheme 2 that the energy absorbed per unit volume is equal to the
  
dv dv r G r0 area under the stress–strain curve (i.e. EV ¼ sP 3D ) is erroneous as
¼ v ¼  sP þ 0 v2 þ u : (16) this clearly implies that the material follows a certain path during
dt du 2 3D A0 3D
the compression. The correct ‘‘energy balance’’ approach would be to
Integrating Eq. (16) with the initial conditions v ¼ v0 at u ¼ 0 use the Rankine–Hugoniot equation that ensures energy conserva-
gives, tion across a discontinuity. Scheme 1 makes no assumptions
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi regarding path-dependency and agrees with the ‘‘shock theory’’ in
20
u
u r0 v20 1 3
Refs. [1–9]. However, the differences between solutions available in
u s P þ
u23D 6B 2 3D C 7 the literature are not so apparent as the differences between
v ¼ u 6B C  sP 7: (17)
t r 4@ r0 A0 uA 5 Schemes 1 and 2 above. That is because in Refs. [12–14,18], Eq. (5) is
0

G3D used to describe the stress jump across the shock. However, the
equation of motion used in Refs. [12–14,18] corresponds to the
From this ‘‘energy balance’’ approach the impact velocity that Scheme 2 approach, i.e. Eq. (16).
just causes full crushing of the core by the end mass can be found The differences between the two solutions are illustrated in
either by setting v0 ¼ vD and v ¼ 0 when u ¼ umax in Eq. (17), or Fig. 3, which shows the critical length of a foam column struck by
more simply by equating the initial kinetic energy of the mass with a 40-kg m2 mass over a range of velocities according to Eqs. (13)
the maximum energy absorbed by the cellular column, giving and (19). The R-P-P-L foam properties used in the analysis are given
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi in the figure title. Clearly the differences between solution Schemes
2A0 umax sP 1 and 2 become large with increasing impact velocity. Note that no
vD ¼ : (18)
G energy absorbing mechanism has been defined that is a function
It follows that the critical thickness of cellular solid required to
just bring a mass G with an impact velocity v0 to rest is
18

Gv20 16
lCR ¼ : (19)
2A0 sP 3D 14
Eqn. (19)
That is, j ¼ 1 and, as expected, there is no velocity effect on 12
energy absorption.
LCR, m

10

8
3.3. Comparison of the solution Schemes 1 and 2
6
It should be noted that Eqs. (5) and (7) are consistent and that Eq. 4
Eqn. (13)
(7) can be derived from Eq. (5) by considering the motion of the end
2
mass and compressed material to the left of the shock front. In the
same way Eqs. (14) and (16) form a consistent pair in that Eq. (16) can 0
0 100 200 300 400
be derived from Eq. (14). However, solution Scheme 2 violates v0, m.s-1
momentum conservation for both the prediction of the stress just
inside the compaction front (cf. Eqs. (14) and (5)) and the governing Fig. 3. Comparison between critical lengths for the two solution schemes for
equation of motion (cf. Eqs. (16) and (7)); therefore Scheme 2 G ¼ 40 kg, A0 ¼ 1 m2, 3D ¼ 0.67, s0 ¼ 0.3 MPa and r0 ¼ 155 kg m3.
922 J.J. Harrigan et al. / International Journal of Impact Engineering 37 (2010) 918–927

of strain-rate or velocity, yet Scheme 1 analysis predicts that 6 Densification


the energy absorbed increases with increasing impact velocity R=1
5
(substitute for sD from Eq. (5) in Eq. (22)). II (Densification) Eq.(16)
4

p
4. Example 2: comparison between mass-spring models 3
and ‘‘shock-wave’’ predictions for a foam column
2 I (Plateau)
subjected to high pressure over a short duration
1
The use of foam has been suggested as a possible way of miti- 0.25 0.5 0.75 1 1.25 1.5 1.75
gating blast loads. This important application has been modelled τd
using mass-spring idealisation in recent work [10]. There are
several interesting features in this application of cellular materials Fig. 5. Non-dimensional enhancement/attenuation boundary according to Li and
for blast protection including a possible shock enhancement Meng [10]. Squares denote densification. Triangles denote distal stresses greater than
the applied pressure.
resulting from the use of a ‘‘protective’’ foam layer. Such an event
is problematic and counterintuitive, considering the use of such
a foam as a protective measure.
The experimental evidence for this enhancement was discussed essentially the same loading condition as considered by Tan et al. [4]
in Ref. [10], and in that paper, the characteristics of compressive to derive the ‘‘steady-state shock model’’. In this stage there is
shock-wave propagation in a cellular material were treated using a steady shock with the two states on either side of the shock
a one-dimensional mass-spring model. For illustrative purposes, remaining constant with time. The compaction wave starts from the
the foam core was considered to be subjected to a rectangular loaded end and the stress within this compacted region is raised to
pressure pulse at one end with magnitude P and duration T P. From Eq. (5), the particle velocity behind the shock during stage 1 is
(see Fig. 4), while the other end of the foam core was fixed and this rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
simplification is followed herein for comparison purposes.
3D
vI ¼ ðP  sP Þ: (25)
Typical pressures considered were 2–5 times the plateau stress in r0
the material and it was argued in Ref. [10] that, although a one- At time t ¼ T, there is a compacted portion at the left end of the
dimensional mass-spring model is more suitable for a system column with an undeformed length given by
consisting of periodic structures, such as a ring or tube system, it is
still capable of representing a macroscopically continuous cellular vI
lI ¼ T : (26)
material. The three-stage stress–strain curve used in Ref. [10] is 3D
shown in Fig. 1. The material response is first linear elastic, then
From Eqs. (25) and (26),
perfectly-plastic before hardening begins at a ‘‘lock-up’’ strain. The
findings were that stress enhancement may occur during the prop- rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
T 3D
agation of a compaction wave through a cellular material [10], where lI ¼ ðP  sP Þ: (27)
3D r0
stress enhancement was defined as stresses anywhere along the foam
column that are greater than the applied pressure. Additionally it was Stage 2. This is equivalent to Example 1, with a mass G impac-
shown that the conditions for full densification of the mass-spring ting the foam column of length lII, where G ¼ r0A0lI, and lII ¼ l  lI is
system (i.e. compression of every spring beyond the ‘‘lock-up’’ strain) the undeformed length of the foam at time t ¼ T. The mass ratio of
and conditions for blast enhancement at the rigid support (i.e. loads the foam core to the rigid mass is equal to the ratio of lII to lI. So,
greater than the applied load) are almost the same. A series of pres- considering only the critical case where the length lII is fully com-
sure pulses were considered in order to predict the pulse conditions pacted at the end of stage 2, Eq. (11) gives
for full densification and therefore stress enhancement at the rigid sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
support. This is shown in Fig. 5 in non-dimensional form where lII r0 v2I
¼ 1þ  1: (28)
lI s P 3D
P
p ¼ ; (23)
sP Since the total critical length of the column in this scenario is
sffiffiffiffiffi lCR2 ¼ lI þ lII and substituting for the velocity from Eq. (25) and for
T E lI from Eq. (27), the critical initial length of the foam column can
sd ¼ ; (24)
L r0 be defined in terms of the loading parameters and material
properties as
and E is the elastic modulus of the foam core.
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
For the same loading condition, it is straighforward to use
ðP  sP Þ
‘‘shock theory’’ with an R-P-P-L material model (see Ref. [7] for lCR2 ¼ T P : (29)
similar example). The pressure pulse is applied at t ¼ 0 and there
sP r0 3D
are two distinct stages to the solution. Because Eq. (29) is defined for an R-P-P-L material, it is not
Stage 1: t < T. The boundary condition during this stage is that possible to non-dimensionalise the solution as has been done in
there is a high pressure P at the left end of the column. This is Ref. [10]. However, the two predictions can be compared by
approximating the properties used in Ref. [10] for Rohacell-51 WF
P(t) foam reported in Ref. [19]. For this material, the properties are
P(t) summarised in Table 1, the only difference between the properties
in Table 1 and those tabulated in Ref. [19] is that, following Ref. [10],
the material is assumed to be linear elastic up to the plateau stress
T so that the yield strain 3Y is equal to 3.64%. Therefore, taking
t
a column length of 1 m for simplicity, a dimensionless loading
Fig. 4. Example 2: rectangular pressure pulse applied to foam column. period sd ¼ 1 corresponds to a loading time of 1.53 ms.
J.J. Harrigan et al. / International Journal of Impact Engineering 37 (2010) 918–927 923

Table 1
Mechanical properties of Rohacell-51 WF foam.
B cS A
r0 (kg m3) E (MPa) sP (MPa) 3Y (%) 3D (%)
51.6 22.0 0.8 3.64 68.9 B A

vB vA
For the ‘‘shock theory’’ model the material is assumed to be R-P-
P-L with a plateau stress of 0.8 MPa and a locking strain of 68.9%. Fig. 7. States immediately ahead of and behind a one-dimensional shock.
Eq. (29) can be rearranged so that for an initial length of 1 m, the
critical period of loading for a given pressure can be calculated equations for the theory given above, following Ref. [6]. Consider
from, a one-dimensional, compressive shock-wave propagating through
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi a cellular material as illustrated in Fig. 7. At the shock front there
sP r0 3D is a discontinuous change in properties. The material state ahead of
TCR ¼ : (30)
PðP  sP Þ the shock is denoted by subscript A and the compacted material
Eq. (30) has been used to plot the enhancement/attenuation behind the shock has the subscript B. Conservation of mass and
boundary for a 1-m long column of Rohacell-51 WF in Fig. 6. Both of momentum across the shock gives:
the curves in Fig. 6 separate the pressure pulses that lead to shock  
r
enhancement at the distal boundary due to full compaction of the vB  vA ¼ cS 1  A ; (31)
foam core from those pulses that are attenuated by the foam. The
rB
regions above and to the right of the lines experience enhancement.
The differences between the two predictions are extremely large. sB  sA ¼ cS rA ðvB  vA Þ; (32)
The material model used in Ref. [10] incorporates both elasticity
where v is absolute particle velocity and cS is the shock speed
and a hardening régime beyond the densification strain (defined as
relative to the material just ahead of the shock, so that the absolute
‘‘lock-up’’ strain in Ref. [10]). This more accurate material descrip-
shock speed is
tion requires an extended shock model which is dealt with in the
next sub-section. VS ¼ vA þ cS : (33)

4.1. Shock-wave models based on R-P-P-L and E-P-P-H Using the more general form of Eq. (2) to define the density r in
material descriptions terms of strain 3,

r0
The generation of attenuation/enhancement boundaries from rð3Þ ¼ ; (34)
13
shock theory is more difficult when using an E-P-P-H material
model than is the case for the R-P-P-L idealisation. A single case is from Eq. (31) the shock speed is given by (see Refs. [6,17])
used to illustrate the effect of the inclusion of elasticity and hard-
ening beyond the onset of densification. The R-P-P-L idealisation
½v dZ
cS ¼ ð1  3A Þ ¼ ; (35)
used here employs a locking strain definition that is consistent with ½3 dt
that in Refs. [2,3]. However, this definition is unrealistic for large where [ ] denotes the change in quantity across the shock front and
velocity changes wherein the stress jump is large and strain jumps Z is the displacement of the shock front (used later) relative to the
can be significantly greater than the locking strain defined in Fig. 1. particle displacement just ahead of the shock and is defined
The locking strain used in the previous section and in Refs. [2,3] is according to Eq. (35) such that a steady shock with a speed cS would
essentially the strain at the onset of densification (see Ref. [23]) and traverse a length dZ of material ahead of the shock over a time
its use for more intense shocks will in general lead to an over- increment dt. The mass of material dm that is compacted by the
prediction of stresses due to compaction waves (see Refs. [4,24]). shock over this time increment can therefore be calculated using
In order to consider a shock-wave model based on a fuller
E-P-P-H material description, it is necessary to generalise the basic dm ½v dZ
¼ A0 rA cS ¼ A0 r0 ¼ A0 rA : (36)
dt ½3 dt
10 Combining Eqs. (31) and (32) gives

r0 ½v2
8 ½s ¼ : (37)
½3
Pressure, MPa

The E-P-P-H material model used here is based on that used in


6
Ref. [10]. The elastic and perfectly-plastic parts of the stress–strain
response are the same (see Fig. 1) as those in Ref. [10]. However, for
4 compression beyond the densification strain, a slightly different
Eqn. (30)
and more convenient stress–strain relationship than that in
Ref. [10] is used. This is because the densification behaviour defined
2 mass-spring [10] in Ref. [10] according to Eq. (38) below is difficult to incorporate in
the set of differential equations that must be solved for the
0 extended shock-wave solution:
0 1 2 3 4
     
Time, ms p  3  3D p
sð3Þ ¼ 1 þ b tan 1g þg  tan ð1  gÞ :
2 1  3D 2
Fig. 6. Enhancement/attenuation boundaries for a 1-m long Rohacell-51 WF foam
column according to ‘‘shock theory’’ and a mass-spring model [10]. (38)
924 J.J. Harrigan et al. / International Journal of Impact Engineering 37 (2010) 918–927

1.0 E-P-P-H material model. From Fig. 1 it can be seen that this E-P-P-H
model is a reasonable approximation to the material model used in
Ref. [10] and defined in Eq. (38).
0.9

4.1.1. Stage 1 for E-P-P-H shock-wave model


0.8 As with the R-P-P-L solution, the E-P-P-H analysis is carried
E-P-P-H
out in two stages. Unfortunately, it is not possible to obtain
Eqn. (38) closed-form solutions for the E-P-P-H material model. For illus-
0.7
tration, a single case is considered for comparison with the
predictions in Ref. [10]. The case considered is for an applied
0.6 pressure P of 3.6 MPa (non-dimensional pressure p ¼ 4.5). A foam
column of 1 m in length is assumed so that the non-dimensional
0.5
loading duration in Ref. [10] of s ¼ sd corresponds to a time T of
1.53 ms. All other material properties correspond to those listed
in Table 1 and plotted in Fig. 1.
0.4 Consider the situation illustrated in Fig. 4 for time t such that
0 50 100 150 200 250
0  t  T. There are two waves travelling through the foam core. The
[v], m.s-1
elastic precursor travels through the foam and loads it to the elastic
Fig. 8. Strain–particle velocity relationship corresponding to Eq. (38) and E-P-P-H limit, sP of Fig. 1. The speed of this wave is c0 and the change in
material models. particle velocity associated with this wave is VY, where
sffiffiffiffiffi
To incorporate an approximate hardening behaviour for the E-P- E
c0 ¼ z 653 m s1 ; (40)
P-H model, consider a compaction wave travelling through the r0
foam, with the material ahead of the foam already loaded to the
elastic limit by an elastic wave. Eqs. (37) and (38) have been P s
combined to plot the relationship between the particle velocity ffi z 24 m s1 :
VY ¼ pffiffiffiffiffi (41)
Er
jump across the compaction front and the strain just inside the
compaction front in Fig. 8. Note that Eq. (38) is in terms of strain Behind this elastic precursor there is a slower travelling
whereas Eq. (37) is in terms of strain jump. However, converting compaction front that has a speed relative to the elastically loaded
from absolute strain to strain jump is simple as the strain ahead of material defined from Eq. (35). The material in the compacted region
the compaction wave is 3Y. In Fig. 8, the E-P-P-H behaviour is is loaded to a stress equal to P. Eqs. (37) and (39) give the strain and
a straight line approximation of the curve derived from Eq. (38) and particle velocity behind the compaction front. The position of the
is given by compaction front at the end of stage 1 can be calculated using
 
3 ¼ 3D þ 0:00105½v: (39) ½v
X1 ¼ VS T ¼ T VY þ ð1  3Y Þ ; (42)
½3
Note that Eq. (39) relates the change in particle velocity to the
change in strain (from 3D to 3) through the gradient of the straight where VS is the absolute velocity of the shock defined in Eq. (33)
line in Fig. 8. The constant in Eq. (39) therefore has units of m1 s. and the quantity changes are those associated with the compaction
Based on the two curves plotted in Fig. 8 and Eq. (35), it is possible wave. The displacement of the proximal end at the end of phase 1 is
to plot a relationship between the shock speed and the change in
particle velocity. This is plotted for both the material defined by u1 ¼ ðVY þ ½vÞT: (43)
Eq. (38) and the E-P-P-H models in Fig. 9. Clearly there is little
The total mass of the compacted portion of the material at the
difference between the shock properties of the two material
end of stage 1 is
models. Eq. (37) was used to plot a stress–strain curve for the
½v
m1 ¼ A0 rA cS T ¼ A0 r0 T: (44)
½3
250 E-P-P-H Eqn. (38)
4.1.2. Stage 2 for E-P-P-H shock model
200
At s ¼ 1, the elastic precursor reaches the distal end of the foam
column and is reflected back as a compaction wave. The jumps in
stress, strain and particle velocity and shock-wave speed at this
cS, m.s-1

150 distal end are simple to calculate since the elastic precursor has
increased the particle velocity to VY and the compaction wave
brings the material back to rest. Note that this solution is an ide-
100
alisation based on the E-P-P-H model and that experimentally,
distal end compaction is often not present. Note also that the stress
50 at the distal end will be greater than sP. However, the stress jump at
this end is small and corresponds to an increase in stress from
0.8 MPa to 0.84 MPa for this material.
0 At the same instant that the elastic wave is reflected back as
0 50 100 150 200 250
a compaction front at the distal end, the pressure is released at the
[v], m.s-1
proximal end. During stage 2, the proximal end can be considered as
Fig. 9. Shock speed–particle velocity relationship corresponding to Eq. (38) and E-P-P-H consisting of an unloading rigid mass of compacted foam that has
material models. sufficient velocity to cause further material to compress at this end,
J.J. Harrigan et al. / International Journal of Impact Engineering 37 (2010) 918–927 925

of the compaction front through the material. The maximum stress


in the foam is equal to the applied pressure during stage 1 and
IV reduces to the plateau stress as the velocity reduces in stage 2. For
t u VY
‘‘shock-wave’’ models the only source of stress enhancement is
III II
m1 cS1 cS2 reflection of the compaction front from the distal end. Other
m
differences between the ‘‘shock-wave’’ and mass-spring predictions
are highlighted by inspection of the wave propagation predicted by
the two types of analysis. For Example 2, an R-P-P-L shock-wave
v solution is plotted in Fig. 11(a). This solution was obtained by
solving for u in Eq. (7) and using Eq. (2) to find the position of the
compaction front. The material in region I of Fig. 11(a) is not
t+δ t u+δ u VY compressed while the material in region II is compacted to the
m1 III II IV densification strain. Full densification does not occur for this case
cS1 cS2
m+δ m since nearly 10% of the column is still in region I when the com-
pacted portion of the material is brought to rest after about 6.9 ms.
The R-P-P-L shock-wave solution therefore predicts that the pres-
v+δ v sure applied to the proximal end is attenuated by the foam column
and the stress at the distal end never rises above the foam’s plateau
Fig. 10. Example 2: stage 2 for E-P-P-H material model. stress. The solutions in Fig. 11 were obtained via a coordinate
system in which the proximal end at t ¼ 0 is the origin and the
i.e. the initial mass at the proximal end is defined in Eq. (44). The distal end of the column is 1 m from the impact. This axis has been
situation in stage 2 is illustrated in Fig. 10. The compaction front at reversed for the plots in Fig. 11 so that the results can be compared
the proximal end is travelling into material in region II that has been more easily with those in Fig. 12 [10]. In Fig. 12, the dimensionless
loaded to the yield stress and has a particle velocity defined in displacement of each mass in a mass-spring model with ten
Eq. (41). The simplest way to obtain the end displacement and wave
positions as a function of time is to consider the material inside the
dashed box in Fig. 10. This box moves to the right with a constant a 1
R-P-P-L
speed equal to VY so that no material enters or leaves the box over
the time increment dt. The compaction front travels to the right with
a speed greater than VY so that the mass of the compacted portion of 0.8
the foam increases by an amount dm over the time increment dt. The
material in region II has been raised to the yield stress so that the
Position, m

change of momentum of the material inside the box can be related to 0.6
the impulse delivered at the boundary of the box as

ðm1 þ m þ dmÞðv þ dvÞ  ðm1 þ mÞv  dmVY ¼ A0 sP dt: (45) 0.4 II u

Re-arranging Eq. (45) and combining this expression with Eqs.


(36) and (39) it is possible to define velocity changes as a function of
0.2
v and Z only, i.e. I
X
 
r0 ðv  VY Þ2
 sP þ 0
dv ð3D  3Y þ 0:00105ðv  VY ÞÞ 0 0.001 0.002 0.003 0.004 0.005 0.006 0.007
¼   : (46)
dt m1 r0 Time, s
þ Z
A 0 1  3Y
A solution can then be obtained by solving three first-order b 1
E-P-P-H
partial differential equations in terms of the proximal end velocity v
(Eq. (46)), the length Z (Eq. (36)) and the displacement of the
0.8
proximal end uðdu=dt ¼ vÞ. The initial conditions for stage 2 come
from the final conditions of stage 1. The compaction front at the
proximal end will continue to propagate until either full densifi-
Position, m

0.6
cation occurs or the particle velocity in region III has reduced to the
level in region II (see Fig. 10). As the compaction wave propagates,
its position can be calculated using
0.4

X ¼ Z þ VY ðt  TÞ þ X1 ; (47)
u
where X1 is the position at the end of stage 1. 0.2 I II III

X1
IV
4.2. Compaction waves predicted by the R-P-P-L and the X2
E-P-P-H shock waves and the mass-spring model 0
0 0.001 0.002 0.003 0.004 0.005 0.006 0.007
Time, s
As mentioned previously, there are large differences between
the ‘‘shock-wave’’ and mass-spring predictions. For ‘‘shock-wave’’ Fig. 11. Propagation of elastic and compaction waves according to ‘‘shock-wave’’
models, stress enhancement never occurs during the propagation theory for (a) R-P-P-L and (b) E-P-P-H material models.
926 J.J. Harrigan et al. / International Journal of Impact Engineering 37 (2010) 918–927

1 discretisation levels used. Large amplitude vibrations may result


9 10 from insufficient discretisation and these vibrations can be mis-
0.9
interpreted as stress enhancements.
8 Another obvious difference is that Fig. 12 shows masses ‘‘pulling
0.8
7 away’’ from the back of the proximal end. This is not seen in the
0.7 results reported in Ref. [11], nor in the ‘‘shock-wave’’ solutions, where
6 unloading behaviour is not considered and compacted portions of
0.6
material will remain at their maximum strain throughout the solu-
yi 5
0.5 tion, even when the stress reduces (rigid unloading). This omission in
L
4 the ‘‘shock-wave’’ solution may lead to errors for the pressure pulse
0.4
loading situation, however this should not be significant for other
3
0.3 loading situations where impact masses, facesheets, etc. prevent
2 separation of the material during unloading.
0.2 The inclusion of elasticity is unlikely to lead to large differences in
1 the ‘‘shock-wave’’ prediction of Fig. 6, but will lead to a decrease in the
0.1
critical time period calculated for an R-P-P-L material from Eq. (30). On
0
0 1 2 3 4 5
the other hand, the inclusion of a more accurate model beyond the
τ onset of densification will lead to increased predictions for this critical
period. Overall, the large differences between the predictions for the
Fig. 12. Propagation of elastic and compaction waves according to mass-spring model enhancement/attenuation boundary shown in Fig. 6 are not likely to be
[10]. Each line shows the dimensionless position of a point mass as a function of the result of the simplified R-P-P-L material model used in the shock
dimensionless time.
theory. Rather, the difference is due to the fundamental difference
(explained below) in the way that the compaction wave is modelled.

elements is plotted against dimensionless time. The mass-spring 4.3. Discussion of mass-spring and alternative models
prediction is that the compaction wave reaches the distal end and for shock propagation
the distal stress is enhanced by the foam column.
The wave locations for the E-P-P-H solution are shown in Mass-spring models have certain similarities to those FE models
Fig. 11(b). The material in region I is unstressed and has no particle that incorporate a cellular structure for the material [4,20,21]. Both
velocity. The material in region II has been stressed to the elastic types of analyses make use of lumped masses as well as elements
limit and has a particle velocity VY. The material in regions III and IV that represent the strength of the solid phase of the material.
has been compressed beyond 3D. The solution has been determined Furthermore, some FE predictions (for example those based on
until the particle velocity in region III has been reduced to that in two-dimensional honeycombs) have been shown to agree well
region II at a time of approximately 6.1 ms. At this time 8% of the with ‘‘shock-wave’’ theory [4,24].
foam column has no permanent deformation. An elastic unloading However the fundamental difference between the mass-spring
wave is now released from the location of the proximal end and FE models that incorporate a cellular structure for the material is
compaction front and although further compression will occur (see that the loading path is defined in the mass-spring models that have
Ref. [6]), equating the kinetic energy remaining in the system to the been described in the literature to date, wherein the quasi-static
‘‘quasi-static’’ energy absorption capacity remaining in the foam stress–strain curve is always assumed to be followed within each
shows that full densification will not take place. individual cell. This is similar to the Scheme 2 prediction considered
The E-P-P-H solution differs from the R-P-P-L solution in two for Example 1 and herein lies the error. As noted earlier, the speed of
obvious ways. First, reflection of the elastic precursor results in the compaction wave derived from the mass-spring model in
compression from the distal end. Second, the compressive strains in Ref. [11] is compared to the speed predicted by ‘‘shock theory’’.
region III of Fig. 10(b) are clearly much greater than those in region The argument in Ref. [11] is that the close agreement between the
II of Fig. 10(b). However, there are clear similarities in the solutions compaction wave speeds predicted by the two methods verifies the
and both produce the same conclusion for Example 2, i.e. full accuracy of the mass-spring system. However, only low pressures
densification does not occur so that the pressure pulse applied to are considered in Ref. [11] and differences in predictions will
the proximal end is attenuated by the foam column. increase with increasing ‘‘jumps’’ in particle velocity. Furthermore,
Clearly the results differ greatly with those for the mass-spring the close agreement in terms of compaction wave speeds is simply
model. The compaction wave in Fig. 12 shows very little reduction because this wave speed is highly dependent on the impact velocity
in velocity. However, the point masses within the compacted region and the stress–strain relationship within the densification region.
apparently change speed a number of times after the compaction The ‘‘extra’’ energy absorbed due to dynamic compression in Ref. [11]
wave has passed beyond the mass. This oscillation is not seen in the is simply the consequence of the cell being compressed further along
‘‘shock-wave’’ models and is the cause of the stress histories shown the quasi-static loading path. Alternatively, the ‘‘extra’’ energy
in Ref. [10] where there is more than one peak stress within the absorbed in a ‘‘shock’’ results from the material following a different
compaction wave. Early studies on mass-spring models [25] show loading path to the quasi-static response. As such, the compaction
a succession of saddles and peaks that correspond to the post-yield wave in the mass-spring models is of a different nature to that in
deformation characteristics of the cells, even for low velocity ‘‘shock-wave’’ theory. The one-dimensional mass-spring models
impacts. The number of peaks in the load-history is directly related reported in Refs. [10,11] are not capable of modelling the disconti-
to the number of cells that are crushed. However, the oscillations in nuities that would exist in a shock in a cellular material. Even if
stress and displacement reported in Ref. [10] are unusual. The viscous damping effects are included in a mass-spring model, the
strength of these mass-spring models is their ability to incorporate result is that a loading-path-dependent solution is defined. In
a wide variety of material properties and a variation in strength contrast, when FE models that incorporate a cellular structure for the
from one cell to the next. However, their ability to simulate stress material are loaded dynamically, inertial effects can lead to defor-
wavefronts (shock or non-shock) in a material depends on the mation modes that differ from those for quasi-static compression.
J.J. Harrigan et al. / International Journal of Impact Engineering 37 (2010) 918–927 927

This is a fundamental problem and applies to all situations References


where shock behaviour is required by the physics of the problem.
The examples above have been kept simple in order to highlight [1] Reid SR, Peng C. Dynamic uniaxial crushing of wood. Int J Impact Eng 1997;
19(5–6):531–70.
the differences between the different analytical approaches. [2] Tan PJ, Harrigan JJ, Reid SR. Inertia effects in uniaxial dynamic compression of
For example, the boundary conditions affecting the interaction a closed cell aluminium alloy foam. Mater Sci Technol 2002;18:480–8.
between the foam and the agent producing the blast are unrealistic [3] Tan PJ, Reid SR, Harrigan JJ, Zou Z, Li S. Dynamic compressive strength prop-
erties of aluminium foams. Part I – Experimental data and observations.
compared to those for a real blast situation. Further details on the J Mech Phys Solids 2005;53:2174–205.
effect of fluid–structure interaction on the pressure pulses applied [4] Tan PJ, Reid SR, Harrigan JJ, Zou Z, Li S. Dynamic compressive strength prop-
to structures during blast loading can be found in Ref. [26] for blasts erties of aluminium foams. Part II – Shock theory and comparison with
experimental data. J Mech Phys Solids 2005;53:2206–30.
in water and in Ref. [27] for blasts in air. Also herein, the cellular [5] Harrigan JJ, Reid SR, Reddy TY. Inertia effects on the crushing strength of wood
materials have been treated as a continuum and deformations are loaded along the grain. In: Allison IM, editor. Experimental mechanics.
assumed to occur at wavefronts. Clearly this is not always the case Rotterdam: Balkema; 1998. p. 193–8.
[6] Harrigan JJ, Reid SR, Tan PJ, Reddy TY. High rate crushing of wood along the
(e.g. see Ref. [28]) and sufficient particle velocity must be reached
grain. Int J Mech Sci 2005;47:521–44.
before a ‘‘steady’’ compaction wave is formed [4]. A recent FE study [7] Hanssen AG, Enstock L, Langseth M. Close-range loading of aluminum foam
that employs a hexagonal cell structure [24] gives predictions that panels. Int J Impact Eng 2002;27:593–618.
[8] Lopatnikov SL, Gama BA, Haque MJ, Krauthauser C, Guden M, Hall IW, et al.
agree closely with ‘‘shock-wave’’ theory and suggests that the
Dynamics of metal foam deformation during Taylor cylinder–Hopkinson bar
compaction front has a dimension of approximately one cell size. impact experiment. Compos Struct 2003;61:61–71.
This is in agreement with the experimental evidence available [29]. [9] Lopatnikov SL, Gama BA, Haque MJ, Krauthauser C, Gillespie Jr JW.
The material properties that are needed to predict the shock High-velocity plate impact of metal foams. Int J Impact Eng 2004;30(4):
421–45.
compression of cellular materials are ill-defined. ‘‘Equations of state’’ [10] Li QM, Meng H. Attenuation or enhancement – a one-dimensional analysis on
for solid materials are normally derived from plate-impact test data. shock transmission in the solid phase of cellular material. Int J Impact Eng
Equivalent data for compaction waves in cellular materials are not 2002;27:1049–65.
[11] Gao ZY, Yu TX. One-dimensional analysis on the dynamic response of cellular
available due to the difficulties associated with measuring the ‘‘states’’ chains to pulse loading. Proc Inst Mech Eng C J Mech Eng Sci 2006;220:
on either side of a compaction wave. Currently, quasi-static stress– 679–89.
strain curves are used to predict stresses under impact conditions. [12] Ashby MF, Evans A, Fleck NA, Gibson LJ, Hutchinson JW, Wadley HNG. Metal
foams: a design guide. London: Butterworth Heinemann; 2000. p. 163–170.
These quasi-static material properties are usually measured over [13] Guoxing L, Tongxi Y. Energy absorption of structures and materials. Cam-
length scales over which several cells are present in any direction and bridge: Woodhead Publishing Ltd.; 2003. p. 289–304.
deformation modes are very different from those during dynamic [14] Fleck NA, Deshpande VS. The resistance of clamped sandwich beams to shock
loading. J Appl Mech 2004;71(3):386–401.
compaction. This issue will be the subject of future research.
[15] Tan PJ, Reid SR, Harrigan JJ. Discussion: ‘‘The resistance of clamped sandwich
beams to shock loading’’ (Fleck NA, Deshpande VS. ASME J Appl Mech
5. Conclusions 2004;71:386–401). J Appl Mech 2005;72:978–9.
[16] Fleck NA, Deshpande VS. Closure to ‘‘Discussion of ‘The resistance of clamped
sandwich beams to shock loading’’’ (ASME J Appl Mech 2005;72:978–9).
Two examples of the dynamic compression of a cellular material J Appl Mech 2005;72:980.
have been analysed. It has been shown that the ‘correct’ shock [17] Li QM, Reid SR. About one-dimensional shock propagation in a cellular
theory solution can be derived purely by applying the conservation material. Int J Impact Eng 2006;32:1898–906.
[18] Qiu X, Deshpande VS, Fleck NA. Dynamic response of a clamped circular
equations for mass and momentum and this derivation has been sandwich plates to shock loading. J Appl Mech 2004;71:637–45.
used to explain differences between ‘‘shock’’ solutions available in [19] Li QM, Mines RAW, Birch RS. The crash behaviour of Rohacell-51WF foam. Int
the literature. An extended shock theory, which accounts for the J Solids Struct 2000;37:6321–41.
[20] Ruan D, Lu G, Wang B, Yu TX. In-plane dynamic crushing of honeycombs –
non-linearity in the post-locking behaviour of the material, has finite element study. Int J Impact Eng 2003;28:161–82.
been derived and compared to the predictions of a mass-spring [21] Zheng Z, Yu J, Li J. Dynamic crushing of 2D cellular structures: a finite element
model for the case of rectangular pressure pulse loads. study. Int J Impact Eng 2005;32:650–64.
[22] Harrigan JJ, Hung YC, Tan PJ, Bourne NJ, Withers PJ, Reid SR, et al. High rate
The mass-spring model in Ref. [10] predicts that stress enhance- compaction of aluminium alloy foams. In: Furnish MD, Elert M, Russell TP,
ments can occur during the propagation of a compaction wave White CT, editors. Shock compression of condensed matter – 2005. Melville:
through a cellular material. The ‘‘shock theory’’ prediction is that the American Institute of Physics; 2006. p. 1519–22. ISBN: 0735403414.
[23] Li QM, Magkiriadis I, Harrigan JJ. Compressive strain at the onset of the
stress in the compacted region is constant at P during stage 1 and then
densification of cellular solids. J Cellular Plastics 2006;42(5):371–92.
reduces below this value during stage 2, i.e. the stress in the foam only [24] Zou Z, Reid SR, Tan PJ, Li S, Harrigan JJ. Dynamic crushing of honeycombs and
exceeds the magnitude of the applied pressure once the entire foam features of shock fronts. Int J Impact Eng 2009;36:165–76.
[25] Shim VPW, Tay BY, Stronge WJ. Dynamic crushing of strain-softening
column is fully compacted. For pressures between two and five times
cellular structures – a one dimensional analysis. J Eng Mater Technol 1990;
the plateau stress in the foam, there is a large difference in the pre- 112:398–405.
dictions of the ‘‘mass-spring’’ and ‘‘shock-wave’’ models. The predic- [26] Deshpande VS, Fleck NA. One-dimensional response of sandwich plates to
tions are expected to diverge as the loading intensity increases. underwater shock loading. J Mech Phys Solids 2005;53:2347–83.
[27] Kambouchev N, Noels L, Radovitzky R. Nonlinear compressibility effects in
‘‘Shock theory’’ has been shown to predict well experimental fluid-structure interaction and their implications on the air-blast loading of
results for a number of cellular materials (e.g. see Ref. [6]) and to be in structures. J Appl Phys 2006;100:063519.
close agreement with FE simulations (e.g. see Ref. [24]). However [28] McKown S, Shen Y, Brookes WK, Sutcliffe CJ, Cantwell WJ, Langdon GS, et al.
The quasi-static and blast loading response of lattice structures. Int J Impact
‘‘mass-spring’’ models have not been calibrated against test data for Eng 2008;35:795–810.
cases where compaction waves are likely to occur. As such, the authors [29] Radford DD, Deshpande VS, Fleck NA. The use of metal foam projectiles to
believe that ‘‘mass-spring’’ models should be treated with caution. simulate shock loading on a structure. Int J Impact Eng 2005;31:1152–71.

Potrebbero piacerti anche