You are on page 1of 9

International Journal of Solids and Structures 49 (2012) 2754–2762

Contents lists available at SciVerse ScienceDirect

International Journal of Solids and Structures


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

Inertia effects on the progressive crushing of aluminium honeycombs


under impact loading
B. Hou a, H. Zhao b,⇑, S. Pattofatto b, J.G. Liu b, Y.L. Li a
a
School of Aeronautics, Northwestern Polytechnical University, 710072 Xi’an, China
b
Laboratoire de Mécanique et Technologie, ENS-Cachan/CNRS-UMR8535/Université Paris 6, 61 avenue du président Wilson, 94235 Cachan cedex, France

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

Article history: This paper presents the test results under quasi-static and impact loadings for a series of aluminum hon-
Available online 17 May 2012 eycombs (3003 and 5052 alloys) of different cell sizes, showing significantly different enhancements of
the crushing pressure between 3003 honeycombs and the 5052 ones. A comprehensive numerical inves-
Keywords: tigation with rate insensitive constitutive laws is also performed to model the experimental results for
Honeycombs different cell size/wall thickness/base material, which suggests that honeycomb crushing pressure
Impact loading enhancement under impact loading is mostly due to a structural effect.
Lateral inertia
Such simulated tests provide detailed local information such as stress and strain fields (in the cell wall)
Kolsky’s bar
Numerical simulation
during the whole crushing process of honeycombs. A larger strain (in the cell wall) under impact loading
than for the quasi-static case before each successive folding of honeycombs is observed, because of the
lateral inertia effect. Thus, differences of the ratios of the stress increase due to strain hardening over
the yield stress between 3003 and 5052 alloys lead to the different enhancements of crushing pressure.
This result illustrates that the lateral inertia effect in the successive folding of honeycombs is the main
factor responsible for the enhancement of the crushing pressure under impact loading.
Ó 2012 Elsevier Ltd. All rights reserved.

1. Introduction For the energy absorption applications such as protective design


for accidental collisions of high speed vehicles or the bird strike of
Aluminium honeycombs are widely used in railway, automotive aircrafts, the out-of-plane behavior for large strains (up to 80%) un-
and aircraft industries because of their excellent physical and der impact loading is desired. Experimental results show that the
mechanical properties such as an interesting strength/weight ratio crushing pressure of honeycombs under impact loading is higher
and an outstanding capability in absorbing energy. Mechanical than that under quasi-static loading. For example, (Goldsmith
behavior for small strain under quasi-static loadings such as the and Sackman, 1992; Goldsmith and Louie, 1995) have reported
elastic behavior and failure strength are well investigated for struc- some experimental works on out-of-plane crushing and on the bal-
tural applications (Gibson and Ashby, 1988). Elastic and fracture listic perforation of honeycombs. They have fired a rigid projectile
models for out-of-plane crushing (Zhang and Ashby, 1992a) and to a target made of honeycomb and have shown that the mean
in-plane crushing (Zhang and Ashby, 1992b), as well as for trans- crushing pressures sometimes increase up to 50% compared to
verse shearing (Shi and Tong, 1995), have been developed. For the static results. Wu and Jiang (1997), Baker et al. (1998), Harrigan
the case of larger strain, theoretical, experimental and numerical and Reid (1999), Zhao and Gary (1998), and Zhao et al. (2005) have
studies have also been reported. Theoretical models can predict also found the similar phenomenon for metallic honeycombs with
the crushing pressure of honeycombs from its geometrical param- an enhancement ranging from 10% to 50%.
eters and wall material behavior such as the out-of-plane crushing As the aluminium alloy is hardly rate sensitive in the range of
pressure (Wierzbicki, 1983), the in-plane crushing pressure moderate impact speed, there exists no plausible explanation of
(Klintworth and Stronge, 1988), and multi-axial collapse envelope this enhancement. Indeed, the possible effect due to the eventual
(Mohr and Doyoyo, 2004a). Other related topics such as fracture shock front (Reid and Peng, 1997; Pattofatto et al., 2007), air
detection using elastic waves (Thwaites and Clark, 1995), negative trapped in the cell (Gibson and Ashby, 1988) cannot be applied
Poisson’s ratio honeycombs (Prall and Lakes, 1997), and foam-filled here. Besides, since the testing methods used in previous works
honeycombs (Wu et al., 1995), have also been reported in the open are rather different from each other, reasonable doubt also exists
literatures. on the validity of those experimental results. Enhancement or
not, how much, why, are still the open questions.
⇑ Corresponding author. Tel.: +33 1 47 40 20 39; fax: +33 1 47 40 22 40. This paper describes an experimental and numerical study of
E-mail address: zhao@lmt.ens-cachan.fr (H. Zhao). the out-of-plane compressive behavior for a series of aluminium

0020-7683/$ - see front matter Ó 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ijsolstr.2012.05.005
B. Hou et al. / International Journal of Solids and Structures 49 (2012) 2754–2762 2755

honeycombs of different size (relative densities varying from 1.78% specimen. Instead of average stress and strain in a common SHPB
to 4.72%) and different base material (3003 and 5052 alloys) under test, we use only the pressure p(t) as a function of the crush D(t)
moderate impact velocities. Tests in a range of impact speed from to give an overall idea of the behavior of the honeycombs. They
10 m/s to 28 m/s were performed using large diameter polymer are defined as follows (Eq. (2)):
Split Hopkinson bars. The accuracy of polymeric SHPB systems is
known to be adequate to obtain reliable testing results on such soft pðtÞ ¼ ðF input ðtÞ þ F output ðtÞÞ=2Ss
Z t ð2Þ
materials. The testing results confirmed that there does exist an
DðtÞ ¼ ðV output ðsÞ  V input ðsÞÞds
enhancement from 10% to 60% for aluminum 5052 and 3003 0
honeycombs.
A numerical model of the studied honeycombs (3003 and 5052 where Ss is the apparent area of the specimen face.
alloys) was performed afterward using ABAQUS commercial codes. It is worthwhile to notice that impact tests on such soft cellular
Similar enhancements were found with numerical results (>40% materials using a SHPB have two major difficulties. One is the large
for 3003 honeycombs and <20% for 5052 ones). Finally, on the basis scatter due to the small cell/sample ratio. To overcome this diffi-
of numerical models, a comprehensive numerical study of succes- culty, a large diameter pressure bar is necessary to host a larger
sive crushing process was performed in order to understand the specimen. Another is the weak signal due to the weak strength of
reason of this enhancement as well as its dependence on the base honeycombs, which leads to a low signal/noise ratio. Large diame-
materials. It shows that the lateral inertia in the successive folding ter, soft, but polymeric pressure bars are used to overcome these
of thin-wall tube structures can explain such observed enhance- difficulties. In practice, two Hopkinson bar systems were used: a
ment of out-of-plane crushing pressure. 60 mm PA6.6 Hopkinson bar system with input and output bars
of 3 m (in LMT-Cachan), and a 30 mm diameter PMMA bar system
with 2 m input and output bars (in the Laboratory of Dynamics and
2. Experimental impact rate sensitivity of 5052 and 3003 Strength, NWPU). The projectiles were made of same materials as
aluminium honeycombs pressure bars and their lengths are respectively 1.2 m and 0.5 m
(less than half input bar length to avoid superimposition of the tail
2.1. Experimental methods and procedures of incident impulse in viscoelastic bars, see Zhao et al., 1997). Thus,
the impulse is not long enough to reach the densification point in
Honeycombs of various wall thickness and cell size made of dynamic test.
5052 or 3003 alloys were tested under axial compression in the Moreover, soft polymeric bars are viscoelastic materials, and
out-of-plane direction. The quasi-static experiments were per- the wave dispersion effect increases greatly with the diameter of
formed using a universal tension/compression testing machine. the bars. Consequently, as the three waves in Eq. (1) are not mea-
The dynamic experiments were conducted on a SHPB (Split Hop- sured at bar-specimen interfaces to avoid their superimposition,
kinson Pressure Bar) apparatus (Hopkinson, 1914; Kolsky, 1949), they have to be shifted from the position of the strain gauges to
commonly used as an experimental technique to study constitutive the specimen faces. Kolsky’s original SHPB analysis on the basis
laws of materials at high strain rates. of a one-dimensional wave propagation theory is no longer valid
A typical SHPB set-up is composed of long input and output bars here. The shifting along pressure bars is performed in our experi-
with a short specimen placed between them. A projectile launched ments with Pochhammer and Chree’s harmonic wave propagation
by a gas gun strikes the free end of the input bar and develops a theory in an infinite cylindrical bar (Davies, 1948; Follansbee and
compressive longitudinal incident wave ei(t). Once this wave Franz, 1983), extended to viscoelastic bars (Zhao and Gary,
reaches the bar-specimen interface, part of it er(t) is reflected, 1995). Detailed data processing procedure can be found in Zhao
whereas the other part goes through the specimen and develops et al. (1997).
the transmitted wave et(t) in the output bar. As the stress and par-
ticle velocity of a longitudinal stress wave can be calculated from
2.2. Honeycomb specimens and testing results
the strain measured by gauges, and shifted at any other place,
the transmitted wave can be shifted to the output bar-specimen
Firstly, four 3003 alloy honeycombs referenced by the single
interface to obtain the output force and velocity, whereas the input
wall thickness h, and the minimum cell diameter S (Fig. 1) were
force and velocity can be determined via incident and reflected
studied. Samples were designed as columns of 30 mm high, with
waves shifted to the input bar-specimen interface. The forces and
hexagonal cross section containing a maximum number of cells
particle velocities can be then calculated as follows (Eq. (1)):
in a circle of 30 mm diameter (Fig. 2).
F input ðtÞ ¼ SB E ðei ðtÞ þ er ðtÞÞ V input ðtÞ ¼ C 0 ðei ðtÞ  er ðtÞÞ Table 1 provides a summary of honeycomb geometry parame-
ð1Þ ters and the relative densities (see Gibson and Ashby, 1988) of four
V output ðtÞ ¼ C 0 et ðtÞ F output ðtÞ ¼ SB E et ðtÞ
types of Al3003 honeycombs.
where F input ; F output ; V input ; V output are forces and particle velocities at Quasi-static tests with loading speed of 0.03 mm/s and SHPB
the interfaces, SB, E and C0 are respectively the cross sectional area, test between 26 and 28 m/s were performed in the out-of-plane
Young’s modulus and the longitudinal wave speed in the pressure direction (T direction in Fig. 2). For the quasi-static tests, the pres-
bars. ei ðtÞ; er ðtÞ; et ðtÞ are the strain signals at the bar-specimen sure (measured force divided by the nominal cross-sectional area)
interface.
The standard SHPB analysis (Hopkinson, 1914; Kolsky, 1949)
provides an average nominal stress–strain curve, dividing the dis-
placement and force respectively by the initial length and the
cross-sectional area of the specimen (see Zhao and Gary, 1996).
For the out-of-plane crushing tests of honeycombs, mechanical
fields are not uniform (even under quasi-static loading). Therefore,
Hopkinson bars here are considered as only a loading and measur-
ing system which can give accurately the force and displacement
time histories on the specimen faces without considering the
deforming characteristics (uniform or not) of the sandwiched Fig. 1. Geometry of the unit cell.
2756 B. Hou et al. / International Journal of Solids and Structures 49 (2012) 2754–2762

4.0

3.5 Dynamic
Quasi-static
3.0
o
No 4 (Al3003)
2.5

Pressure (MPa)
2.0

1.5

Fig. 2. Hexagonal honeycomb specimens.


1.0

10 0.5

9 0.0
-3 0 3 6 9 12 15 18 21 24 27 30
8
Crush (mm)
7
Pressure (MPa)a

o
No 3 (Al3003) Fig. 4. Dynamic enhancement of honeycomb No. 4 pressure/crush curve.
6

5
Table 1a
4
Summary of 3003 honeycomb parameters.
3 Honeycomb number Material h/S (mm/mm) Relative density (%)

2 1 Al3003 0.05/5.2 2.57


2 Al3003 0.06/4.33 3.70
1 3 Al3003 0.06/3.46 4.61
4 Al3003 0.04/3.46 3.08
0
-2 0 2 4 6 8 10 12 14 16 18 20 22 24 26 28
Crush (mm)
Table 1b
Fig. 3a. Reproducibility of quasi-static experiments on honeycomb No. 3. Summary of the parameters and experimental results of 3003 honeycombs.

Honeycomb Material h/S Relative pquasi-static pimpact c


number (mm) density (%) (MPa) (MPa) (%)
11
1 Al3003 0.05/ 2.57 1.24 1.96 58.1
10 5.2
2 Al3003 0.06/ 3.70 2.79 4.06 45.5
9 4.33
3 Al3003 0.06/ 4.61 4.00 5.51 37.8
8 3.46
o
No 3 (Al3003)
4 Al3003 0.04/ 3.08 1.20 1.75 45.8
7
Pressure (MPa)

3.46
6

5 impact loading for the same kind of specimens No. 3 (Table 1).
There exist larger oscillations probably due to impact testing
4
imperfection. However, the scatter on the average plateau level is
3 small.
2 It is noticed that the crush reached in impact test is quite lim-
ited (around 6 mm) because of limited length of impulse in the
1
Hopkinson bars system, which gives also an enlarged impression
0 of this oscillations. A direct comparison between dynamic and qua-
-1 0 1 2 3 4 5 6 7 si-static pressure/crush curves is shown in Fig. 4 for the case of
Crush (mm) specimen No. 4 (Table 1a). Quasi-static specimens undergo much
larger crush compared to the dynamic ones. Oscillations under im-
Fig. 3b. Reproducibility of impact experiments on honeycomb No. 3.
pact loading in the plateau stage in such figure are not significantly
can be depicted with respect to the measured crush displacement. higher than the quasi-static case.
For example, Fig. 3a shows the pressure/crushing displacement As in the case of quasi-static loading, an average plateau value
curves for the three independent quasi-static tests on the specimen of the pressure pdynamic can be derived. The impact enhancement
No. 3 (h = 0.06 mm/S = 3.46 mm). There is a scatter on the initial ratio c is defined as follows (Eq. (3)):
buckling peak force and the locking strain, but the average plateau c ¼ ðpdynamic  pquasistatic Þ=pquasistatic ð3Þ
level is very stable. The average value of the pressure between the
initial peak and locking strain is used as the quasi-static pressure Table 1b gives the summary of testing results of all the four
pquasi-static. 3003 honeycombs. From those testing results, it is clear that the
For the SHPB tests, the pressure and the crush are calculated strength of 3003 alloy honeycomb under out-of-plane compression
using Eq. (2). Fig. 3b illustrates the three repeated tests under exhibits an important impact enhancement around 40%.
B. Hou et al. / International Journal of Solids and Structures 49 (2012) 2754–2762 2757

Table 2 aluminium sheet metals (6065 T5 in compression (Zhao, 1997),


Summary of the parameters and experimental results of 5052 honeycombs. 2024 T3 under shear loadings (Zhao et al., 2007)) reveal also a lim-
Honeycomb Material h/S Relative pquasi- pimpact c ited rate sensitivity (<10%).
number (mm/ density (%) static (MPa) (%) Therefore, the significant impact enhancement (>40%) observed
mm) (MPa) in 3003 honeycombs cannot be directly derived from the rate sen-
5 Al5052 0.076/ 1.78 1.7 2 17.6 sitivity of the cell wall base materials. There must be a structural
9.52 reason responsible for this enhancement. For this purpose, numer-
6 Al5052 0.076/ 2.76 3 3.5 16.7
ical simulation of the above honeycomb testing results is devel-
6.35
7 Al5052 0.076/ 4.72 5.1 5.7 11.8 oped. Since our study focuses on the behavior of honeycombs,
4.76 the modelling of the whole testing environment is not necessary.
Thus, only honeycomb structures were modeled here and the load-
ing environment was modeled by two rigid planes moving at the
prescribed velocities corresponding to the average value of those
measured in the experiments. Commercial FEM code of ABAQUS/
Explicit was employed for this simulated work.
In order to reduce the calculation cost with a complete honey-
comb model with the same geometry as the hexagonal honeycomb
specimen (Fig. 2) honeycomb specimen was simplified into a unit
cell consisting of three conjoint half walls in Y-shape (Fig. 5) be-
cause of its periodicity (Mohr and Doyoyo, 2004b; Hou et al.,
2011; Wilbert et al., 2011). The simplified models work with sym-
metric boundary conditions applied on the three non-intersecting
edges of each cell wall.
It is noticed that the leg of this Y-shape cell-model is a thick
wall in a real honeycomb typically made of two single-thickness
thin walls which are bonded together. In this model, we ignore
the rare delamination of the bonded interfaces and consider the
strength of the adhesive bond as infinite. Thus, the simulations
Fig. 5. Scheme of honeycomb cross section and the unit cell-model.
are carried out for a monolithic structure, where the thick walls
are represented by the same shell elements but with a double
Secondly, three Alcore 5052 honeycombs (Duracore) were also thickness value.
studied. Samples were designed as columns of 50 mm high with The model is meshed with 4-node doubly curved thick shell ele-
hexagonal cross section containing a maximum number of cells ments with a reduced integration, active stiffness hourglass control
in a circle of 60 mm diameter. For the honeycomb of largest cells, (S4R) and 5 integration points through the cell wall thickness. In
there are still 5 cells at least in one edge of the hexagon. Quasi-sta- order to determine the appropriate element size, a convergence
tic and SHPB tests (1014 m/s) were performed in the out-of-plane study was performed among different element sizes. The element
direction. Table 2 gives the characteristics of different tested hon- size is finally chosen to be 0.1 mm.
eycombs (density, cell size and wall thickness) as well as the quasi- The numerical specimen is placed between two rigid planes
static and dynamic mean crushing pressure results (Zhao et al., moving at constant velocities, which take the mean value of real
2005). tests (i.e. 27 m/s for 3003 alloy honeycomb and 15 m/s for 5052 al-
It is found that the enhancement for 5052 alloy honeycombs is loy honeycomb). In this model, general contact with frictionless
rather small (<20%), compared to that of 3003 alloy honeycombs tangential behavior is defined for the whole model excluding the
(>40%), noting that this is a general trend obtained from different contact pairs of rigid planes and tested honeycomb specimen,
wall thickness/cell size ratios in the same range of relative density which are redefined by surface-to-surface rough contact to make
(3 for 5052 and 4 for 3003). sure that no slippage occurs.
As the real honeycomb is always far from perfect, it includes all
2.3. Numerical simulation of honeycombs kinds of imperfections which affect the initial peak value, but have
little influence on the crush behavior at a large strain. These imper-
The rate sensitivity of bulk aluminium alloy is known to be fections are due to various reasons, like irregular cell geometry, un-
small (<10%, see Duffy et al., 1971). Recent testing results on thin even or pre-buckled cell walls, wall thickness variation etc. Here in

(a) (b)
Fig. 6. Two honeycomb cell-model with different cell-size (with initial imperfection).
2758 B. Hou et al. / International Journal of Solids and Structures 49 (2012) 2754–2762

this work, we generated the imperfections by preloading the per- Table 3


fect specimen uniaxially by 0.1 mm. Then, the obtained displace- Bilinear material parameters.

ment of nodes are introduced geometrically into the actual Material Density Young’s Poisson’s Yield Hardening
model used for further calculation. The value of 0.1 mm is chosen q (kg/ modulus E ratio m stress rs modulus Et
to make sure that the simulated initial peak force is the same as m3) (GPa) (MPa) (MPa)

the one from experimental curve at uniaxial compression. Fig. 6 Al3003 2700 70 0.35 70 1150
illustrates two numerical models of different cell sizes as well as Al5052 2700 70 0.35 290 500

details of introduced imperfection. It shows that the imperfections


introduced by preloading is very small. hardening modulus (Table 3) were finally determined by fitting
Quasi-static simulations are almost impossible to achieve with one quasi-static calculation result (i.e. No. 1 (Al3003) and No. 6
ABAQUS/Standard which uses Newton’s method (or quasi New- (Al5052)) to the experimental one for two base materials respec-
ton’s method) as a numerical technique due to the complex nonlin- tively. These bilinear curves are also illustrated in Fig. 7.
ear effects, e.g. the geometrical and material nonlinearity, the Fig. 8 shows a comparison between experimental and simulated
complex contact conditions as well as the local instability during pressure/crush curves for honeycomb No. 1 (Al3003) under quasi-
crush. An alternative is to use also ABAQUS/Explicit for quasi-static static loading. It shows that the cell-model exhibits significant fluc-
problems. However, the explicit integration scheme of dynamic tuations at the plateau stage, which is probably due to the applica-
simulation codes usually leads to very small time step which in tion of excessive symmetric boundary constraints. Actually, it is
our simulation is around ten nanoseconds for the chosen element well known that the crushing behavior of honeycombs under
size. Thus, with the loading velocity of 0.1 mm/s, the computa- out-of-plane compression is regulated by the successive folding
tional duration for the quasi-static simulation (e.g. 130 s) will be procedure of honeycomb cell walls. With the symmetric boundary
too large. To overcome this difficulty, automatic mass scaling tech- condition on three non-intersecting edges, the cell-model is actu-
nique was employed to increase the time increment to 10 ls. The ally equivalent to a honeycomb specimen consisting of repeated
quasi-static loading conditions are guaranteed by ensuring the ra- cells with identical deforming procedure, which results in a strictly
tio of the kinetic energy to the strain energy as a small value (of the simultaneous collapse of all the honeycomb cells. Thus, in the pres-
order of 104) with the chosen time increment. It is known that sure/crush curve, each fluctuation represents one fold formation of
such mass scaling might introduce an artificial strength enhance- the cell wall in honeycomb structure. For the large size model, the
ment when the introduced imperfection is small (Langseth et al., neighboring cells interact with each other while forming the folds
1999). In our simulation, the imperfection obtained by 0.1 mm ax- and reach their local peak value at different instants, which makes
ial crushing seems to be well-adapted because the different pre- the macroscopic resulting curves smoother (Hou et al., 2011;
scribed time-step do not generate significant scatter (Hou, 2011). Wilbert et al., 2011). However, the mean crushing pressure is
A bilinear elastic-plastic material model until 20% strain and hardly affected. We finally use this cell-model for the subsequent
perfect plastic afterwards was employed to describe the cell wall calculations.
material of the aluminum honeycombs. For 5052 H38 honeycomb, Fig. 9 shows the comparison of experimental and numerical
a yield stress of 290 MPa and a small hardening was usually admit- pressure/crush curves for the honeycomb No. 1 (Al3003) under
ted in many previous works (Papka and Kyriakides, 1994). For 3003 quasi-static and impact loadings. Even though the numerical pres-
honeycombs, parameters were identified initially with the tensile sure enhancement is smaller than the experimental one, which
testing results of 1 mm thick dog bone type 3003-O sheet speci- could be due to base material rate sensitivities or other effects
men using universal testing machine and Split Hopkinson tensile not taken into account in this numerical model, the basic trend is
bar test (Fig. 7). One can see that its small rate sensitivity is preserved.
confirmed. Table 4 gives the simulated average plateau values of crushing
However, the hardening or thermal treatment of our 3003 hon- pressure under quasi-static and impact loading for all the tested
eycomb is unknown. The testing results of 3003-O give only a ref- 3003 and 5052 honeycombs. These numerical results suggest also
erence of the real base material behavior of 3003 honeycomb. The a significant enhancement under impact for 3003 honeycombs and
model parameters of the base material such as yield stress and a small enhancement for 5052 honeycombs.

3.0
600

Al3003 Static Exp. 2.5


500
Al3003 Dyn. Exp. Cell-model
Al3003 Moel Experiment
2.0
Pressure (MPa)

Al5052 Model
400
Stress (MPa)

500MPa
1.5
300
290MPa

1150MPa 1.0
200
Quasi-static
0.5
100 o
No 1 (Al3003)
70MPa
0.0
-1 0 1 2 3 4 5 6 7 8 9
0
−0.05 0 0.05 0.1 0.15 0.2 0.25 0.3
Crush (mm)
Strain
Fig. 8. Comparison between numerical and experimental results for Al3003 cell-
Fig. 7. Prescribed constitutive relation of 5052 and 3003 alloys. model under quasi-static loading.
B. Hou et al. / International Journal of Solids and Structures 49 (2012) 2754–2762 2759

b=1.833mm
3.5
h=152 m
Static Exp.
3.0
Static Cal. L=0.5mm
Dynamic Exp. v
2.5
Dynamic Cal.
1
Pressure (MPa)

2.0

1.5 3
2
1.0

0.5

0.0 o
No 1 (Al3003)
Fig. 11. Scheme of double-plate model.
-0.5
-1 0 1 2 3 4 5 6 7 8 9
2.4. Lateral inertia effect on the successive folding mechanism of
Crush (mm)
honeycombs
Fig. 9. Comparison between the calculating and experimental results.
2.4.1. Lateral inertia effect in a simple double-plate model
The early theoretical work on the lateral inertia effect was re-
ported by Budiansky and Hutchinson (1964). Gary (1983) showed
Table 4
Simulated pressure enhancements of 3003 and 5052 honeycombs. experimentally that the buckling of a column under compressive
impact occurs at a larger plastic strain. Calladine and English
Honeycomb Material h/S (mm/ pquasi-static pimpact c (%)
(1984) identified velocity-sensitive type II structure and Tam and
number mm) (MPa) (MPa)
Calladine (1991) revealed that there exists, under dynamic loading,
1 Al3003 0.05/5.2 1.24 1.83 48
an initial phase where the compression is dominant before a sec-
2 Al3003 0.06/4.33 2.27 3.57 57.1
3 Al3003 0.06/3.46 3.61 4.92 36.3 ond phase of bending. More sophisticated models were also re-
4 Al3003 0.04/3.46 1.80 2.49 38.6 ported (Karagiozova and Jones, 1995; Su et al., 1995). It leads to
5 Al5052 0.076/9.52 1.55 1.66 7.38 the fact that the buckling of an elastic-plastic column under com-
6 Al5052 0.076/6.35 2.99 3.58 19.8 pressive impact occurs at a larger strain (than under quasi-static
7 Al5052 0.076/4.76 4.62 5.39 16.7
loading) because of necessary transverse acceleration. If the elas-
tic-plastic column has a strain-hardening behavior, the buckling
peak force will also increase as shown in the numerical work of
Webb et al. (2001).
70 In order to illustrate this lateral inertia effect without long ana-
Al3003 lytical formulas and to quantitatively evaluate the magnitude of
h/S=0.06/4.33
60 Experiment potential augmentation of dynamic buckling force due to inertia
Calculation effect in the honeycombs, a double-plate numerical model with
Dynamic enhancement (%)

Al3003
50 h/S=0.04/3.46 dimensions comparable to honeycomb is built using ABAQUS.
The scheme of this model is shown in Fig. 11 (like most previous
Al3003
40 h/S=0.05/5.2 works cited above), which is composed of two plates connected
with an angle (for the initial imperfection). The size of the model
Al3003 is in the same order with honeycomb cell walls with the plate
30
h/S=0.06/3.46 Al5052 thickness h = 152 lm, plate width b = 1.833 mm and height of
h/S=0.076/6.35 one plate L = 0.5 mm, d is the maximum deviation of plates from
20
the vertical line, which represents the magnitude of initial imper-
Al5052 fection of this model.
10 h/S=0.076/9.52
Al5052
In this study, the initial imperfection employed is fixed to
h/S=0.076/4.72 3.2 lm (much more exaggerated in Fig. 11) in order to avoid the
0 undesirable elastic buckle before the plastic collapse. The double-
0 1 2 3 4 5 6 7 8
plate model is sandwiched between two parallel rigid walls (one
Specimen Number fixed and another moving at prescribed velocity). The loading
Fig. 10. Summary of the calculated and experimental dynamic enhancement for
velocity is 0.1 mm/s for quasi-static case and 15 m/s for dynamic
both 3003 and 5052 alloys. loading. A surface-to-surface rough contact is defined at the inter-
faces of double-plate model and rigid walls to make sure that no
slippage occurs. Such simulation permits to obtain the force and
Fig. 10 depicts comparison of the simulated and experimental displacement time histories, which can be converted to nominal
enhancement ratios for all the 5052 and 3003 honeycombs. stress and strain by being divided by the plate cross sectional area
The numerical models follow well the experiments. It shows and the initial distance between rigid walls.
that the numerical enhancement is around 40% for Al3003 honey- Fig. 12 shows the calculated quasi-static and impact nominal
combs and less than 20% for 5052 ones. Thus, this enhancement stress-strain curves for both 3003 and 5052 alloys, compared with
has its origin in the structural response and this structural re- the prescribed constitutive relations. It can be found that the col-
sponse should be related to the constitutive relationship of two lapse point (at which the curve begin to decrease rapidly) of the
materials because the geometry and loading conditions are similar quasi-static curve coincides with the yield point. However, under
in numerical models for 3003 and 5052 honeycombs. impact loading, the collapse cannot take place at yield point and
2760 B. Hou et al. / International Journal of Solids and Structures 49 (2012) 2754–2762

450

400 5052 material


5052 dynamic
350 5052 static
3003 material
300 3003 dynamic
Stress (MPa)

3003 static
250

200

150

100

50

0
-0.025 0.000 0.025 0.050 0.075 0.100 0.125 0.150

Strain

Fig. 12. Force divided by cross section area vs. displacement divided by length.

the double-plate model is further compressed. Therefore, the mod-


el undergoes larger plastic strains in axial direction under impact
loading before the collapse occurs. As the strain hardening curve
is different, the stress increase due to this larger strain is different
and especially the ratio of this stress increase over the yield stress
is different.
From constitutive relation shown in Fig. 7, for the 3003 alloy,
with a yield stress of 70 MPa and a hardening modulus of
1150 MPa, 5% of strain enhancement induces more than 80% stress
increase. For the 5052 alloy, with a yield stress of 290 MPa and a
hardening modulus of 500 MPa, 5% of strain enhancement induces
only less than 9% of rise in stress. This is considered to be the rea-
Fig. 14. Formation of the second fold of honeycomb cell-model.
son for the pressure enhancement difference between 5052 and
3003 honeycombs.
of this corner part determines the successive peak load. For the
2.5. Lateral inertia effect in the honeycomb cell-models honeycomb structure, this corner part corresponds to the region
near the intersectional line. In our Y-shape cell-model (as shown
Such a lateral inertia effect under impact loading exists also in in Fig. 13(b) the deformed profile), the analysis should be then
the successive folding of tube-like hollow structures (Langseth focused on the central intersectional line of three cell walls.
and Hopperstad, 1996; Zhao and Abdennadher, 2004; Karagiozova Fig. 13(a) shows the pressure/crush curve of a cell-model for
and Alves, 2008). It has been found that in the successive crushing honeycomb No. 1 (Al3003) taken as an example. Large fluctuation
process of square tube, the corner region (intersection of two flat with each wave representing one fold formation is observed.
plates) supports most of the external loadings and the buckling We consider the formation of the second fold to illustrate the

4.0
o
3.5 No 1 (Al3003) Thick wall
Dynamic Vimpact=27m/s
3.0

2.5
B
Pressure (MPa)

D
2.0 d
b c
1.5
a
C Thin wall
1.0 A

0.5 first fold second fold third fold fourth fold

0.0

-0.5
-1 0 1 2 3 4 5 6 7 8 9
Crush (mm)
(a) (b)
Fig. 13. Deformation profile (a) and pressure/crush curve (b) of honeycomb cell-model.
B. Hou et al. / International Journal of Solids and Structures 49 (2012) 2754–2762 2761

successive folding system of honeycomb under out-of-plane com- and will repeat the above-mentioned process, i.e. the C-c-D-d pro-
pression. The deformed profile of cell-model at point B in Fig. 13(a) cess in Fig. 13(a).
is shown in Fig. 13(b). The deformation sequence of the basic cell- The successive folding is controlled by the successive buckling
model is shown in Fig. 14 (only the thick wall is displayed for sake of region near the central intersectional line. It is important to no-
of illustration’s clarity). tice that the intersectional line remain rather straight (Fig. 14(c))
The formation of the second fold begins from point A in before its buckling. Thus, the initial imperfection is small enough
Fig. 13(a). At this moment, the first fold has completely collapsed for each single successive fold so that the lateral inertia effect like
and the material of the second fold begins to support loading in the double-plate model will apply.
(Fig. 14(a)). The continuous axial deformation of the second fold Figs. 15a and 15b depicts the equivalent strain profiles (still
enables the carrying capacity of the cell-model to increase gradu- considering that the strain here is the calculated wall material
ally (Segment a in Fig. 13(a) and the deformation image in equivalent strain) along the central line of one fold (see
Fig. 14(b)). During this process, the intersectional line (as shown Fig. 14(d)) in both the double thickness wall and the single thick-
in Fig. 14(a)) and its adjacent region remains straight, while the ness wall just before the buckling (at the successive force peaks,
plate region has been buckled. The peak load of the second fold points B, D in Fig. 13(a)). For honeycomb No. 1 (Al3003) model
is reached (Point B in Fig. 13(a)) when the intersectional line and (Fig. 15a) as well as No. 6 (Al5052) model (Fig. 15b), one can see
its adjacent region begin to buckle (as show in Fig. 14(c)). After this that the strain reached before buckling is higher under impact
peak point, the overall carrying capacity decreases dramatically loading (27 m/s for Al3003 and 15 m/s for Al5052) than the case
(segment b in Fig. 13(a)) and the corresponding deformation of under quasi-static loading. The closer is the position to the inter-
the cell-model is characterized with the bending of intersection re- sectional line; the larger is the increase in strain. The lateral inertia
gion (as shown in Fig. 14(d)). When the carrying capacity of the effect is then clearly seen in the successive buckling of the central
cell-model reaches the trough C in Fig. 13(a), the third fold initiates intersectional line of the Y-shape cell-model.

0.28 350

0.24
o Quasi-static 300
No 1 (Al3003) Dynamic o Quasi-static
No 1 (Al3003)
0.20 Dynamic
250
Mises stress (MPa)
Equivalent strain

0.16
200
0.12

150
0.08

100
0.04

0.00 50
Half width of thick wall Half width of thin wall
-0.04 0
-1.5 -1.0 -0.5 0.0 0.5 1.0 1.5 -1.5 -1.0 -0.5 0.0 0.5 1.0 1.5
Distance from the intersection line (mm) Distance from the intersection line (mm)

Fig. 15a. Strain distribution in the central line of a fold under quasi-static and Fig. 16a. Stress distribution in the central line of a fold under quasi-static and
dynamic loading, honeycomb No. 1 (Al3003). dynamic loading, honeycomb No. 1 (Al3003).

0.12 400
Quasi-static
o
o
350 No 6 (Al5052) Dynamic
0.10 No 6 (Al5052) Quasi-static
Dynamic
300
0.08
Mises stress (MPa)
Equivalent strain

250
0.06
200

0.04
150

0.02
100

0.00 50
Thick wall Thin wall

-0.02 0
-1.5 -1.0 -0.5 0.0 0.5 1.0 1.5 -1.5 -1.0 -0.5 0.0 0.5 1.0 1.5

Distance from the intersection line (mm) Distance from the intersection line (mm)

Fig. 15b. Strain distribution in the central line of a fold under quasi-static and Fig. 16b. Stress distribution in the central line of a fold under quasi-static and
dynamic loading, honeycomb No. 6 (Al5052). dynamic loading, honeycomb No. 6 (Al5052).
2762 B. Hou et al. / International Journal of Solids and Structures 49 (2012) 2754–2762

The stress profiles just before the buckling are also shown in Harrigan, J.J., Reid, S.R., Peng, C., 1999. Inertia effects in impact energy absorbing
materials and structures. Int. J. Impact Engng. 22, 955–979.
Figs. 16a and 16b. On the one hand, there is a larger difference of
Hopkinson, B., 1914. A method of measuring the pressure in the deformation of high
stress levels between impact and quasi-static loadings for 3003 explosives by the impact of bullets. Phil. Trans. Roy. Soc. A213, 437–452.
model which is due to the important strain hardening behavior Hou B. 2011. Dynamic enhancement and multi-axial behavior of honeycombs under
of prescribed 3003 constitutive law, especially the ratio between combined shear-compression. Ph.D. Thesis of ENS Cachan.
Hou, B., Pattofatto, S., Li, Y.L., Zhao, H., 2011. Impact behavior of honeycombs under
the stress increase due to strain hardening and the yield stress. combined shear-compression, Part II Analysis. Int. J. Solids Struct. 48, 698–705.
On the other hand, because of prescribed flat strain hardening Karagiozova, D., Jones, N., 1995. Some observations on the dynamic elastic-plastic
behavior for 5052 model, the stress level differences for 5052 are buckling of a structural model. Int. J. Impact Engng. 16, 621–635.
Karagiozova, D., Alves, M., 2008. Dynamic elastic-plastic buckling of structural
less significant even though a noticeable strain difference due to elements: a review. Appl. Mech. Rev. 61 (4), 040803.
lateral inertia is observed. Klintworth, J.W., Stronge, W.J., 1988. Elasto-plastic yield limits and deformation
laws for transversely crushed honeycombs. Int. J. Mech. Sci. 30, 273–292.
Kolsky, H., 1949. An investigation of the mechanical properties of materials at very
3. Conclusion high rates of loading. Proc. Phys. Soc. B62, 676–700.
Langseth, M., Hopperstad, O.S., 1996. Static and dynamic axial crushing of square
thin-walled aluminium extrusions. Int. J. Impact Engng. 18, 949–968.
Quasi-static and impact tests with large diameter polymer SHPB
Langseth, M., Hopperstad, O.S., Berstad, T., 1999. Crashworthiness of aluminium
were performed on 7 kinds of aluminium (5052 or 3003) honey- extrusions: validation of numerical simulation, e!ect of mass ratio and impact
comb with a relative density varying from 1.78% to 4.72%. The re- velocity. Int. J. Impact Engng. 22, 829–854.
sults show that the enhancement under impact loading of the Mohr, D., Doyoyo, M., 2004a. Experimental investigation on the plasticity of
hexagonal aluminium honeycomb under multi-axial loading. J. Appl. Mech. 71,
crushing pressure depends on the base material. In fact, 4 different 375–385.
3003 honeycombs have an enhancement larger than 40% while the Mohr, D., Doyoyo, M., 2004b. Deformation-induced folding systems in thin-walled
3 types of 5052 honeycombs exhibit an enhancement less than monolithic hexagonal metallic honeycomb. Int. J. Solids Struct. 41, 3353–3377.
Papka, S.D., Kyriakides, S., 1994. In-plane compressive response and crushing of
20%. honeycomb. J. Mech. Phys. Solids 42, 1499–1532.
Abaqus models using a Y-shape cell-model with rate insensitive Pattofatto, S., Elnasri, I., Zhao, H., Tsisiris, H., Hild, F., Girard, Y., 2007. Shock
constitutive law were performed to simulate the studied sets of enhancement of cellular structures under impact loading: Part II Analysis. J.
Mech. Phy. Solids 55, 2672–2686.
honeycombs under quasi-static and impact loadings. This result Prall, D., Lakes, R.S., 1997. Properties of chiral honeycomb with a Poisson’s ratio of -
confirmed numerically this difference of enhancement. 1. Int. J. Mech. Sci. 39, 305–314.
Finally, strain and stress profiles in single Y-shape cell during Reid, S.R., Peng, C., 1997. Dynamic uniaxial crushing of wood. Int. J. Impact Engng.
19, 531–570.
the successive folding (especially at the force peaks) were ana-
Shi, G.Y., Tong, P., 1995. Equivalent transverse shear stiffness of honeycomb cores.
lysed. The strain reached before collapse (successive force peak) Int. J. Solids Struct. 32, 1383–1393.
under impact loading is bigger than that under quasi-static load- Su, X.Y., Yu, T.X., Reid, S.R., 1995. Inertia-sensitive impact energy-absorbing
structures. Part I: Effects of inertial and elasticity. Int. J. Impact Engng. 16 (4),
ing. This is considered to be due to the lateral inertia effect of type
651.
II structure, well studied in the past. As the 3003 alloy has a bigger Tam, L.L., Calladine, C.R., 1991. Inertia and strain rate effects in a simple plate
ratio between the stress increase due to strain hardening and the structure under impact loading. Int. J. Impact Engng. 11, 689–701.
yield stress than that of 5052 alloy, the crushing pressure enhance- Thwaites, S., Clark, N.H., 1995. Non-destructive testing of honeycomb sandwich
structures using elastic waves. J. Sound Vib. 187, 253–269.
ment under impact loading of 3003 honeycombs is therefore big- Webb, D.C., Kormi, K., Al-Hassani, S.T.S., 2001. The influence of inertia and strain-
ger than 5052 honeycombs. rate on large deformation of plate-structures under impact loading. Comput.
Struct. 79, 1781–1797.
Wierzbicki, T., 1983. Crushing analysis of metal honeycombs. Int. J. Impact Engng. 1,
Acknowledgement 157–174.
Wilbert, A., Jang, W.Y., Kyriakides, S., Floccari, J.F., 2011. Buckling and progressive
The authors would like to thank 111 project of China (Contract crushing of laterally loaded honeycomb. Int. J. Solids Struct. 48, 803–816.
Wu, C.L., Weeks, C.A., Sun, C.T., 1995. Improving honeycomb-core sandwich
No.1307050) for funding the cooperation between NPU and LMT. B. structures for impact resistance. J. Adv. Mater. 26, 41–47.
Hou and Y. L. Li would also like to thank the supports of the Na- Wu, E., Jiang, W.S., 1997. Axial crush of metallic honeycombs. Int. J. Impact Engng.
tional Science Foundation of China (Contract Nos. 10932008 and 19, 439–456.
Zhang, J., Ashby, M.F., 1992a. The out-of-plane properties of honeycombs. Int. J.
11072202).
Mech. Sci. 34, 475–489.
Zhang, J., Ashby, M.F., 1992b. Buckling of Honeycombs under in-plane biaxial
References stresses. Int. J. Mech. Sci. 34, 491–509.
Zhao, H., Gary, G., 1995. A three dimensional analytical solution of the longitudinal
Baker, W.E., Togami, T.C., Weydert, J.C., 1998. Static and dynamic properties of high- wave propagation in an infinite linear viscoelastic cylindrical bar. Application to
density metal honeycombs. Int. J. Impact Engng. 21, 149–163. experimental techniques. J. Mech. Phys. Solids 43, 1335–1348.
Budiansky, B., Hutchinson, J. W., 1964. Dynamic buckling of imperfection sensitive Zhao, H., Gary, G., 1996. On the use of SHPB techniques to determine the dynamic
structures. In: Proceedings of 11th international congress of Applied Mechanics. behavior of materials in the range of small strains. Int. J. Solids Struct. 33 (23),
Springer Verlag, Munich. 3363–3375.
Calladine, C.R., English, R.W., 1984. Strain-rate and inertia effects in the collapse of Zhao, H., Gary, G., Klepaczko, J.R., 1997. On the use of a viscoelastic split Hopkinson
two types of energy-absorbing structures. Int. J. Mech. Sci. 26 (11–12), 689–701. pressure bar. Int. J. Impact Engng. 19, 319–330.
Davies, R.M., 1948. A critical study of Hopkinson pressure bar. Phil. Trans. Roy. Soc. Zhao, H., 1997. A constitutive model for metals over a large range of strain rates.
A240, 375–457. Identification for mild-steel and aluminium sheets. Mater. Sci. Eng. A320, 95–
Duffy, J., Campbell, J.D., Hawley, R.H., 1971. On the use of a Torsional Split 99.
Hopkinson Bar the study rate effects in 11000 Aluminum. J. Appl. Mech. 38, 83– Zhao, H., Gary, G., 1998. Crushing behavior of aluminium honeycombs under impact
91. loading. Int. J. Impact Engng. 21, 827–836.
Follansbee, P.S., Franz, C., 1983. Wave propagation in the split Hopkinson pressure Zhao, H., Abdennadher, S., 2004. On the strength enhancement under impact
bar. J. Engng, Mater. Tech. 105, 61–66. loading of square tubes made from rate insensitive metals. Int. J. Solids Struct.
Gary, G., 1983. Dynamic buckling of an elastoplastic column. Int. J. Impact Engng. 2, 41, 6677–6697.
357–375. Zhao, H., Elnasri, I., Abdennadher, S., 2005. An experimental study on the behavior
Gibson, L.J., Ashby, M.F., 1988. Cellular Solids. Pergamon Press, Oxford. under impact loading of metallic cellular materials. Int. J. Mech. Sci. 47, 757–
Goldsmith, W., Sackman, J.L., 1992. An experimental study of energy absorption in 774.
impact on sandwich plates. Int. J. Impact Engng. 12, 241–262. Zhao, H., Elnasri, I., Girard, Y., 2007. Perforation of aluminium foam core sandwich
Goldsmith, W., Louie, D.L., 1995. Axial perforation of aluminium honeycombs by panels under impact loading: an experimental study. Int. J. Impact Engng. 34
projectiles. Int. J. Solids Struct. 32, 1017–1046. (7), 1246–1257.

You might also like