You are on page 1of 39

 

 
Investigation of strain-rate effect on the compressive behaviour of closed-cell
aluminium foam by 3D image-based modelling

Yongle Sun, Q.M. Li, T. Lowe, S.A. McDonald, P.J. Withers

PII: S0264-1275(15)30519-0
DOI: doi: 10.1016/j.matdes.2015.09.109
Reference: JMADE 679

To appear in:

Received date: 17 May 2015


Revised date: 17 September 2015
Accepted date: 18 September 2015

Please cite this article as: Yongle Sun, Q.M. Li, T. Lowe, S.A. McDonald, P.J. Withers,
Investigation of strain-rate effect on the compressive behaviour of closed-cell aluminium
foam by 3D image-based modelling, (2015), doi: 10.1016/j.matdes.2015.09.109

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
ACCEPTED MANUSCRIPT

Investigation of strain-rate effect on the compressive


behaviour of closed-cell aluminium foam by 3D

T
image-based modelling

IP
Yongle Sun1, Q.M. Li1,3, T. Lowe2, S.A. McDonald2, P.J. Withers2

R
1
School of Mechanical, Aerospace and Civil Engineering, The University of Manchester,

SC
Sackville Street, Manchester M13 9PL, UK
2
Henry Moseley X-ray Imaging Facility, School of Materials, The University of Manchester,
Upper Brook Street, Manchester M13 9PY, UK

NU
3
State Key Laboratory of Explosion Science and Technology, Beijing Institute of Technology,
Beijing 100081, China
MA
D
P TE
CE
AC


Corresponding author, E-mail address: qingming.li@manchester.ac.uk
1
ACCEPTED MANUSCRIPT

Abstract: 3D finite element models based on computed tomography (CT) images are
developed to investigate the strain-rate effect on the compressive behaviour of closed-cell
aluminium foam (Alporas). It is found that at strain-rates below shock conditions the rate

T
dependence of the cell-wall material is the main cause of the strain-rate hardening of the

IP
compressive strength of Alporas foam. The foam exhibits slightly higher strain-rate sensitivity
than that of the cell-wall material due to localised strain-rate amplification in some critical

R
load-bearing elements. By contrast, the micro inertia of individual cell walls associated with

SC
the nonuniform deformation of Alporas foam has a negligible contribution. Under shock
conditions the stress measured at the loading end is always enhanced, but the stress measured

NU
at the supporting end is complicated, depending on the characteristics of the contiguous cells.
In general, shock leads to a concentrated large deformation in the cells at the loading end, but a
MA
restrained small deformation in the cells at the supporting end. Consequently, the hardening
effect of the rate dependence of cell-wall material on the supporting stress becomes constrained
with further increasing strain-rate, and the supporting compressive strength is limited to the
D

quasi-static level when the cell-wall material is rate-independent.


TE

Keywords: Cellular materials; Compressive properties; Strain-rate sensitivity; X-ray


tomography; Finite element modelling
P
CE
AC

2
ACCEPTED MANUSCRIPT

1. Introduction

Closed-cell aluminium foams are lightweight and multifunctional materials possessing

T
unique properties with respect to energy absorption, heat insulation, acoustic damping, fire

IP
resistance and impermeability to fluids. They are attractive for applications where there is a
need to absorb energy or to attenuate impact or blast loading because their stress is almost

R
constant over a large range of compressive deformation. Since under dynamic compression

SC
they may undergo deformation processes that differ from the quasi-static ones, the effect of
strain-rate on their compressive properties has been studied intensively in recent years.

NU
However, the actual cause and extent of the strain-rate effect for aluminium foams are still
under debate, as shown in [1-3]. Nevertheless, the strain-rate hardening of closed-cell
aluminium Alporas foam has been widely recognised [4-11]. The rate dependence of Alporas
MA
foam is evident even at low strain-rates (<0.16 s-1) [5]. A number of explanations have been
proposed. For instance, Paul and Ramamurty [5] suggested that such strain-rate hardening is
D

related to both the strain-rate sensitivity of the aluminium matrix and the micro inertia of the
TE

individual cell walls. Cady et al. [6] argued that many factors, e.g. the cell-wall interaction,
pore architecture and the rate dependence of the cell-wall material, are involved. Dannemann
P

and Lankford [7] attributed the strain-rate effect to the flow of gas through ruptured cell walls.
Mukai et al. [10] also considered the pressurisation of trapped gas in the closed cells as the
CE

possible cause. On the other hand, Elnasri et al. [12] and Merrett et al. [13] examined the
structural response during high speed impact and proposed that the strong strain-rate hardening
AC

is mainly a shock wave related phenomenon.


These and other research studies suggest that the rate dependence of the cell-wall material,
the micro inertia, the shock wave, and the trapped gas are potentially the predominant factors
responsible for the strain-rate effect. However, their individual contributions cannot be easily
distinguished in experiments. This has motivated theoretical and numerical analyses using
simplified cell geometries, e.g. cubic unit cell [1, 14] and Voronoi structures [2, 3, 15, 16].
These studies have provided general insights into the dynamic compressive behaviour of metal
foams, but they have substantial drawbacks when applied to specific closed-cell aluminium
foams since they may not be able to sufficiently capture the complexity of real cell structures.
The geometrical simplification or idealisation may introduce significant errors since some
factors such as the effects of the rate dependence of the cell-wall material and micro inertia
depend on the structural characteristics [17]. Therefore simulations considering more realistic
3
ACCEPTED MANUSCRIPT

cell structures are needed to increase the reliability of the modelling results. Recently,
numerical models based on cell structures obtained from computed tomography (CT) images
(so-called image-based modelling [18]) have been used to capture the actual meso-scale
geometry in order to investigate the strain-rate sensitivity of open-cell foams [19, 20]. Similar

T
modelling techniques have also been employed to investigate the quasi-static compressive

IP
properties of closed-cell foams [21-23]. However, the dynamic compressive behaviour of
closed-cell aluminium foams, such as Alporas foam, which are widely used in research and

R
engineering applications, have yet to be examined in this way.

SC
This study is aimed to investigate the strain-rate effect on the compressive behaviour of
closed-cell aluminium foam (Alporas), with focus on meso-scale geometrically realistic

NU
modelling to distinguish the contributions of potential physical factors. Three-dimensional (3D)
finite element models (FEM) are developed using the cell structures obtained from CT images
MA
in order to capture the actual meso-scale geometry. The uniaxial compression of the Alporas
foam is simulated over a wide range of applied nominal strain-rates between 1×10-3 s-1 and
3×103 s-1. The numerical predictions are also compared with the compression test data reported
D

in the literature.
TE

2. 3D image-based finite element models


P

2.1. Geometrical characterisation and meshing


CE

To obtain the actual meso-scale 3D geometry, a cylindrical Alporas foam sample with
dimensions of Ø15.0×14.4 mm was scanned using a Nikon Metris CT system housed in a
customised bay at the Henry Moseley X-ray Imaging Facility (HMXIF, Manchester, UK). The
AC

accelerating voltage and current of the X-rays were set as 75 kV and 125 μA, respectively.
The effective voxel size was 10.9 µm. The acquired X-ray radiographs (2000 projections with
an exposure time of 1 s for each) of the foam sample were then reconstructed into a 3D volume
using Nikon Metris CT-Pro reconstruction software. To eliminate the damaged surfaces
incurred by machining during sample preparation, the central portion of the foam sample with
dimensions of Ø11.0×13.5 mm was selected as the region of interest (ROI). The CT image of
the ROI was segmented using greyscale-based thresholding to extract the solid phase. The grey
level threshold was adjusted to select the cell walls which have higher grey values than the
background. The lower bound of the grey values was set as low as possible, which is necessary
to avoid missing the thin walls having lower contrast to the background, although this tends to
thicken the walls slightly. The ―closing‖ filter was then applied to the segmented cell walls to
eliminate the tiny pores having diameters smaller than 400 µm. The neglect of these tiny pores
4
ACCEPTED MANUSCRIPT

greatly simplified the meshing and reduced element number. Fortunately, these tiny pores
hardly affect the compressive strength of the foam since the yielding and collapse of the cell
walls occur much more easily within or near to relatively large cells [24].
In addition, the cell walls were slightly dilated (thickened by 44 µm in this case) to

T
facilitate image down-sampling which is necessary to reduce element number of FE meshing.

IP
The dilation avoided the missing of the thinnest wall portions, which were represented by
several voxels (<50 µm) in the original segmented CT image, during the subsequent

R
down-sampling which increased the voxel size from original value of 10.9 µm to 50.0 µm.

SC
Fig. 1 shows the transverse CT slices at the mid-height of the virtual foam sample after the
reconstruction and image processing. It is evident that the cell morphology and topology are

NU
extremely complex and the cell-wall thickness is nonuniform, showing substantial differences
between the cell structures of real foams and those of the idealised foams used in previous
MA
models [1-3, 14, 15]. In Fig. 1b the colours indicate the cell-wall thickness which is measured
in 3D using a method [25] that defines the local thickness at a material point as the diameter of
the largest sphere which contains the point and is completely inside the cell wall. Due to the
D

resolution limit and the processing of the CT image, the cell structure constructed here cannot
TE

be exactly the same as that of the real foam which has many fine features (e.g. extremely tiny
pores and thin walls). Nevertheless, it maximally preserves the actual cell morphology and
P

topology which are key geometrical factors related to foam strength. Such realistic meso-scale
CE

geometry provides the basis for the analysis of the effects of physical factors (e.g. cell-wall
material properties and strain-rate) otherwise the prediction could deviate from reality solely
because of geometrical simplification.
AC

The histograms of the cell size and local cell-wall thickness of the virtual foam sample are
shown in Fig. 2. The size of a cell is obtained as the diameter of an equivalent sphere having
the same volume as the cell. The individual cells are identified using a watershed-based pore
separation algorithm (Avizo Fire, VSG, France). It is seen that the peak frequency corresponds
to a cell size in the range of 0.6-0.9 mm while 74% of cells are larger than 1.0 mm.
The statistical data for the cell size and cell-wall thickness are listed in Table 1. The
average cell size was calculated by
N
1
d
N
d
i 1
i (1)

where d i is the size of the cell i and N is the total number of cells. The volume-weighted
average of local cell-wall thickness [25] was obtained as

5
ACCEPTED MANUSCRIPT

M
1
h
M
h
j 1
j (2)

where h j is the local cell-wall thickness at the voxel j (material point in continuum) and M is

the total number of voxels.

T
Both the cell size and local cell-wall thickness exhibit high scatter (see Table 1). The

IP
maximum size is about one order larger than the minimum. The scatters in cell size and local

R
cell-wall thickness are also clearly seen in Fig. 1. The thickness of the walls at the junctions is
much larger than that at other locations, and many small cells are located in the thick walls.

SC
Such structural characteristics arise from the foaming process during which cells are formed in
the aluminium melt [26].

NU
It is challenging to generate a sufficiently fine mesh for the closed-cell foam studied here
which has extremely complex cell structure. To accomplish the FE discretisation we employed
MA
ScanIP (Simpleware Ltd, UK), which is designed for directly converting CT images into high
quality 3D meshes [27]. The profile of the foam sample after meshing is shown in Fig. 3.
Quadratic elements were used for meshing since such elements have high numerical precision
D

and are necessary to capture the bending and buckling of thin walls which are unlikely to be
TE

very finely meshed due to the geometrical complexity and the limited mesh density that can be
adopted. During the meshing, an additional surface smoothing algorithm was applied, which
P

led to a smoother surface and somewhat thinner cell walls represented by the FE mesh than that
CE

represented by the down-sampled CT image, see Fig. 3b. The relative density (i.e. the ratio of
the cell wall volume to the foam volume) of the meshed sample is 0.109, around 11% larger
AC

than that of the scanned sample. This mainly results from errors introduced by the image
processing and meshing steps. A similar volume increase of closed-cell foam model also
occurred when other image processing and meshing techniques were used (e.g. around 21%
volume increase occurred in the model by Jeon et al. [21]). A mesh sensitivity study shows that
the whole cell wall volume can be reduced by 1.1% when the element number further increases
by 74.1%, leading to around 2.5% decrease of the peak stress. However, the ratio of the
dynamic peak stress to the quasi-static one is hardly changed by the mesh refinement, which
indicates that the mesh density chosen here (see Fig. 3) is adequate for the analysis of the
strain-rate effect on the compressive strength.
Most of the simulations in this study were conducted for the relatively small sample
containing 115 cells, as shown in Fig. 3. To verify the numerical results, a larger foam sample
(Ø20.0×20.0 mm, as shown in Fig. 4), which has 390 cells and is much more computationally

6
ACCEPTED MANUSCRIPT

expensive, was also considered for several loading cases. Similar image processing and
meshing procedures were conducted to create the FE model for this large sample having a
mean cell size of 2.20 mm, a mean cell-wall thickness of 0.30 mm and a relative density of
0.166. Let l be the smallest side dimension of the foam sample, the l / d is 5 and 9 for the

T
small sample and large sample, respectively, which is acceptable to evaluate representative

IP
bulk compressive strength of Alporas foam [28]. The results presented in following sections are
obtained from the small sample, unless otherwise stated.

R
SC
2.2. Cell-wall material model
The material properties assigned to the cell walls were assumed to be isotropic and

NU
homogeneous. Hooke’s law and von Mises plasticity were adopted for the material model,
while mechanisms of material damage were not considered. A power law was used to
MA
determine the strain hardening of the yield stress. The elastic and plastic behaviour can be
described by the following stress-strain relationship under uniaxial loading

    Y0
D

(3)
E
TE

n
   
  Y0      Y0 (4)
E   Y0 
P

where  and  are the uniaxial stress and strain, respectively, E is the Young’s modulus,
CE

 Y0 is the yield strength and n is the hardening exponent. The material parameters were taken
to be E =68 GPa,  Y0 =35.5 MPa and n=8.5 [29]. The Poisson’s ratio of 0.33 was adopted.
AC

The rate dependence of the yield stress was determined by the Cowper-Symonds equation, viz.
1
 Yd   m
 1   * (5)
 Ys  

where  Yd is the dynamic yield stress,  Ys is the corresponding static yield stress, m is a

material constant,  is the strain-rate and  * is a reference strain-rate. Parameters typical of


aluminium, i.e.  * =6500 s-1 and m=4 [20], were adopted to represent the rate dependence of
the cell-wall material. These material parameters have also been used in previous simulations
based on 2D [3] and 3D [15] Voronoi tessellations. The case where the cell-wall material is
rate-independent was also considered for comparison. The loading processes were treated as
dynamic in order to incorporate inertia effects at various strain-rates, and the cell-wall density
was taken to be 2710 kg/m3.
7
ACCEPTED MANUSCRIPT

2.3. Numerical model setup


The foam sample was loaded on the top surface by applying a constant downward velocity
to the associated nodes, while for the bottom surface the displacement was fixed in the vertical
direction. In-plane motion was allowed on the end surfaces and no constraints were exerted on

T
the lateral sides of the sample. For comparison purpose, the upward loading was also

IP
considered in some cases and the compressive load was applied by exerting a constant upward
velocity on bottom nodes. The nominal strain  nom and stress  nom were defined as the

R
SC
displacement divided by the original sample height and the end-surface reaction divided by the
original sample cross-sectional area, respectively. The nominal strain-rate  nom was defined
as the ratio of the loading speed to the original sample height and varied from 1×10-3 s-1 to

NU
3×103 s-1 in the simulations. Following the method used in Refs. [4, 5, 7, 8], the compressive
strength was defined as the first peak stress or collapse stress measured at the supporting end.
MA
The general purpose FE code, Abaqus, was employed to perform the simulations
considering large deformation effect. The implicit solver of Abaqus/Standard for dynamic
problems was adopted. According to Abaqus documentation [30], both Abaqus/Standard and
D

Abaqus/Explicit are capable of accurate dynamic modelling, but the size of the time increment
TE

in an explicit dynamic analysis is limited to a very small value determined by the smallest
element in the mesh, and consequently Abaqus/Explicit is usually more suitable for the
P

simulations when loading speed is high (or short period of loading). As in this study the
CE

applied nominal strain-rate is in a wide range of 1×10-3-3×103 s-1, the implicit solver, which is
unconditionally stable and does not have such size limit of time increment, was chosen in order
AC

to efficiently simulate the compression at low and intermediate strain-rates. Moreover,


Abaqus/Standard allows the use of standard quadratic elements which have high numerical
precision and are necessary to capture the deformation of the thin walls in the present FE
model. Although Abaqus/Standard has these advantages, it may not be able to converge if
extreme impact loading and complex contact interaction are encountered. Therefore, the
loading speed in the present FE model is restricted to relatively low level (the highest loading
speed is 40.5 m/s and 60.0 m/s for the small sample and large sample, respectively) and the
contact between cell walls is not considered. By checking the foam deformation, it is found
that the cell walls do not come into contact before a nominal strain of 7.5% which is sufficient
to obtain the collapse stress of primary interest in this study.

3. Results

8
ACCEPTED MANUSCRIPT

3.1. Stress-strain relationship and deformation mode


For dynamic compression, it is necessary to distinguish between the stress measured at the
loading end and that at the supporting end, since macro inertia effects may cause stress
enhancement at the loading end. For instance, at  nom  3 103 s1 , as shown in Fig. 5, the stress

T
at the loading end is significantly larger than that at the supporting end, whether the cell-wall

IP
material is rate-dependent or not. This phenomenon is referred to as shock enhancement for

R
cellular solids and has been widely observed in experiments [e.g. 12, 13, 31] and simulations

SC
[e.g. 2, 3, 32]. It should be noted that the critical impact speed to initiate shock in Alporas foam
is reported to be 40 m/s [13], which corresponds to a nominal strain-rate of 2963 s-1 and 2000
s-1 for the small sample and large sample, respectively. In addition, as Tan et al. [31] pointed

NU
out, the loading stress under shock is essentially correlated with impact speed rather than
strain-rate since it arises from macro inertia effects. Therefore, the shock enhancement should
MA
be evaluated separately from the strain-rate effect on the compressive strength without
considering contribution from macro inertial force. This has been thoroughly discussed in
previous studies based on experiments [31, 33] and simulations [2, 3]. Another observation
D

from Fig. 5 is that both the loading and supporting stresses for the foam with rate-dependent
TE

cell walls are larger than those for the foam with rate-independent cell walls.
To exclude the shock enhancement occurring at the loading end, the stress measured at the
P

supporting end is used to characterise the compressive strength at the high strain-rates. It
CE

should be noted that at low and intermediate strain-rates (1×10-3-1×102 s-1), the loading and
supporting forces (and stresses) are equal. This balancing of forces at the macro-scale, however,
AC

can co-exist with the micro inertia of individual cell walls arising from the nonuniform
deformation of the heterogeneous cell structure. Fig. 6 shows the compressive stress-strain
curves at different strain-rates. These curves are typical of foams under uniaxial compression
[5, 7, 9, 28], i.e. an essentially linear relationship at low stress before attaining a peak stress
after which the compressive stress drops to a plateau due to cell crushing. The quasi-static (
 nom  1103 s1 ) compressive strengths obtained for foams with cell-wall materials being
rate-dependent and rate-independent are 1.61 MPa and 1.57 MPa, respectively, which lie
within the scatter band of the compressive strength (1.3-2.3 MPa) reported for Alporas foam of
similar density [34, 35]. For the foam with rate-independent cell walls the compressive
stress-strain curves are identical at the strain-rates between 1×10-3 s-1 and 1×102 s-1 (see Fig.
6a), indicating the negligible influence of micro inertia for Alporas foam in this strain-rate
range. By contrast, differences arise at strain-rates above 1×103 s-1 due to the shock effect. For
9
ACCEPTED MANUSCRIPT

instance at  nom  3 103 s1 , the elastic precursor wave does not arrive at the supporting end
until the nominal strain reaches 0.007, after which the stress becomes non-zero and increases
rapidly as a result of the transmission of load through the deforming cell walls. The result is
different if the cell-wall material is taken to be rate-dependent. In this case, the compressive

T
strength generally increases with strain-rate (Fig. 6b), but the increase becomes subtle from

IP
1×103 s-1 to 3×103 s-1. This is because when shock initiates at high strain-rates, the compressive
strength measured at the supporting end is not affected solely by the rate dependence of the

R
cell-wall material, which will be further discussed later.

SC
The cell deformations at different strain-rates are compared in Figs. 7 and 8. Fig. 7 shows
the plastic strain distribution in the internal cell walls at  nom  0.04 when the cell-wall

NU
material is taken to be rate-independent. It is seen that plastic bending and buckling are
dominant deformation modes at all the considered strain-rates. However, at the low and
MA
intermediate strain-rates the significant plastic strain (>2%) is mainly in ―weak‖ cells close to
the supporting end, whereas at the high strain-rate the plastic deformation is concentrated close
to the loading end due to the shock effect. Such deformation feature is consistent with the
D

experimental observations on the quasi-static and dynamic compressions of Alporas foam [6, 9,
TE

12, 13, 36, 37]. The localisation of cell deformation at the loading end at the high strain-rate in
Fig. 7 leads to a stress-strain curve distinct from those at the low and intermediate strain-rates,
P

as shown in Fig. 6a.


CE

Fig. 8 shows the cell deformation of the foam with rate-dependent cell walls. Here the
cell-wall material is hardened by not only plastic strain but also strain-rate. The pattern of
AC

plastic deformation is very similar to that shown in Fig. 7. However, the increase in yield stress
with strain-rate implies that the magnitude of the plastic strain in the cell walls is reduced at
intermediate and high strain-rates. This leads to the strain-rate hardening of the compressive
stress of the foam (see Fig. 6b) and more distributed plastic deformation in cell walls (see
comparison between Figs. 7 and 8 for the strain-rates of 1×102 s-1 and 3×103 s-1). Recalling the
stress-strain curves at the strain-rates of 1×103 s-1 and 3×103 s-1 shown in Fig. 6b, the subtle
rises in the peak stress at the supporting end at high strain-rates can be attributed to the
concentrated large cell deformation at the loading end and constrained small cell deformation
at the supporting end under shock compression.
As the cell deformation depends on loading direction when shock occurs, it may be
necessary to also examine the compressive strength measured at the supporting end when the
compressive load is applied upward. Fig. 9 compares the stress-strain curves when loading

10
ACCEPTED MANUSCRIPT

direction changes. It is evident that the compressive stresses are different between downward
compression and upward compression at the strain-rates of 1×103 s-1 and 3×103 s-1. For the
downward compression, the cells adjacent to the supporting end are large and unstable (see
Figs. 7 and 8), in contrary to the case for upward compression, and consequently the peak

T
stress is smaller for the former. This implies that the measured compressive strength of the

IP
foam sample under shock is dependent on the characteristics and arrangement of cells near to
the sample end, which may be different from sample to sample. However, it has been verified

R
that compressive strength is independent of loading direction at low and intermediate

SC
strain-rates.

3.2. Strain-rate sensitivity of the compressive strength

NU
To identify the strain-rate effect, the compressive strength (at the supporting end) is
normalised by that at the strain-rate of 1×10-3 s-1 (regarded as quasi-static). The normalised
MA
strength is plotted against strain-rate in Fig. 10 alongside compression test data from the
literature. Reasonable agreement is achieved between the simulations (for rate-dependent cell
walls) and experimental tests although the experimental strain-rate sensitivity appears higher.
D

Furthermore, the foam has a slightly higher strain-rate sensitivity than the cell-wall material
TE

itself (Cowper-Symonds curve in Fig. 10) until the shock regime wherein the measured
compressive strength depends on loading directions. When the cell-wall material is
P

rate-independent, the compressive strength is insensitive to strain-rate although variation


CE

occurs after  nom  1103 s1 . In general, the strain-rate hardening is somewhat constrained by
the reduced cell deformation near to the supporting end with further increasing strain-rates
AC

under shock compression (e.g. from 1000 s-1 to 3000 s-1), which is consistent with the test data
trend shown in Fig. 10.
In order to evaluate the extent to which the results on the small foam sample could be
generalised more widely, the strain-rate effect was also examined for the larger foam sample.
The normalised strain-rate effects on the compressive strength at the supporting end are
summarised in Table 2. Basically the results show that the compressive strength is independent
of strain-rate, unless the rate dependence of the cell-wall material is introduced in which case
the foam shows a slightly stronger strain-rate effect than the cell-wall material until the shock
regime at  nom  3 103 s1 . This confirms the observation from the small sample.

4. Discussion

From our simulations based on the actual cell structure of Alporas foam we can clarify the
11
ACCEPTED MANUSCRIPT

contributions of three factors to the strain-rate effect on the compressive behaviour of the foam,
namely, the micro inertia, the rate dependence of cell-wall material and the shock wave.
Firstly, we analysed the case where the cell-wall material has no rate dependence and no
shock response (strain-rates between 10-3-102 s-1). It is evident that the micro inertia, which

T
depends on the local acceleration of individual cell walls, is not the cause of the strain-rate

IP
hardening of Alporas foam, since if it had any effect, the foam strength should depend on the
strain-rate even when the rate dependence of the cell-wall material is not included, which

R
evidently is not the case in the simulations (see Figs. 6a and 10 and Table 2). This agrees with

SC
previous work using Voronoi structures where micro inertia was observed to have a negligible
effect [2]. This confirmation from the image-based modelling is important as there is a

NU
question as to whether observations from idealised cell structures are applicable to real
closed-cell foams, since the effect of micro inertia is dependent on structural characteristics
MA
[17]. To quantitatively evaluate the extent of micro inertial effect, we adopt the
transverse-to-longitudinal acceleration ratio defined in Ref. [2], viz.
 X  AX AZ (6)
D

Y  AY AZ (7)
TE

1/2
 n 
where A    ai2 / n  is the mean square root of the acceleration with ai being the
 i 1 
P

acceleration of the i-th node in Z direction (longitudinal loading direction) or X (or Y)


CE

direction (transverse direction) and n being the total number of nodes of the foam model. The
typical values of the acceleration ratios at different strain-rates are listed in Table 3 (the results
AC

at high strain-rates are also given), which indicates that the transverse acceleration (i.e. micro
inertia) is close to the longitudinal acceleration for the simulated foam samples and thus the
Alporas foam cannot exhibit Type-II structure behaviour which is characterised by much
higher transverse acceleration [2].
Secondly, we analysed the case that the cell-wall material is rate-dependent but the shock is
absent (strain-rates between 10-3-102 s-1). The corresponding FE results are presented in Figs.
6b, 8 and 10 and Table 2. This shows a considerable strain-rate hardening of the compressive
strength of the foam, but reduced plastic deformation in cell walls due to the strain-rate
hardening of the cell-wall material itself. Furthermore, the comparison between the rate
dependence of foam and that of cell-wall material shows that the foam exhibits slightly higher
strain-rate sensitivity than that of the cell-wall material. This is attributed to the localised
strain-rate amplification in some critical load-bearing elements wherein cell-wall deformation
12
ACCEPTED MANUSCRIPT

is concentrated, as shown in Fig. 11. However, it should be noted that the local strain-rate
depends on the structural characteristics of closed-cell foams. For instance, for 2D and 3D
Voronoi closed-cell foams [3, 15] the rate dependence of aluminium cell walls with same
Cowper-Symonds parameters as those used here has less effect on the compressive stress than

T
that observed here, implying that the local strain-rate amplification in critical load-bearing

IP
elements may be absent or weak for these kinds of idealised foam structures. For cubic cell,
Deshpande and Fleck [14] estimated that the local strain-rate of cell walls is about one order of

R
magnitude lower than the nominal strain-rate. These differences thus demonstrate the

SC
importance of using actual cell structure in modelling to investigate the material properties of
real foams. Another question arises when we note that the numerical prediction underestimates

NU
the strain-rate sensitivity of the compressive strength even when the rate dependence of the
cell-wall material is incorporated in modelling. This underestimation is mainly due to the
MA
assumption of the rate dependence of the cell-wall material described by the Cowper-Symonds
equation with typical parameters for aluminium. The cell-wall material may exhibit a higher
rate dependence than that assumed in the simulations, since the components and microstructure
D

of the cell-wall material (with aluminium matrix) of Alporas foam is complicated due to the
TE

foaming process during which additional material elements are introduced and special
processing is applied [26]. Unfortunately, it is still challenging to conduct dynamic tests
P

directly on the cell-wall material and thus the accurate determination of the rate dependence of
CE

the cell-wall material has not been reported in the literature.


Thirdly, we analysed the case that the cell-wall material is rate-independent but shock
occurs (e.g.  nom  3 103 s1 ). Unsurprisingly the shock increases the stress measured at the
AC

loading end (see Fig. 5). This results from the macro inertial force which enhances the loading
stresses even if the foam material is inherently insensitive to strain-rate, as discussed in the
literature [2, 3, 12, 13, 32, 33]. The shock effect on the compressive strength measured at the
supporting end is complicated. For downward compression, slight strength reduction occurs for
the small sample (see the result for  nom  3 103 s1 in Fig. 10) but the strength does not

change for the large sample (see the data for  nom  3 103 s1 in Table 2). Similar observations
were also reported for other types of cellular solids. For instance, Zou et al. [38] reported that
the dynamic supporting stress of honeycombs remains more or less unchanged. Hu and Yu [39]
observed that the average supporting stress of honeycombs decreases with impact speed. Liu et
al. [2] and Liao et al. [40] also found the decreasing trend of supporting stress for 2D Voronoi
foams with increasing impact speed. For open-cell aluminium Duocel foam, Barnes et al. [41]
13
ACCEPTED MANUSCRIPT

concluded that the dynamic supporting stress is at relatively low level and bounded by the
quasi-static one. To explain this phenomenon, the distinct cell deformation under shock must
be invoked. As shown in Fig. 7, the shock leads to a concentrated large cell deformation at the
loading end and leaves cells elsewhere much less deformed, in contrast to the localised

T
deformation occurring in ―weak‖ cells under low and intermediate strain-rate compressions.

IP
We noticed that when compressive load is applied downward the ―weakest‖ cells controlling
the collapse stress are located close to the supporting end for the small sample but around the

R
middle part for the large sample. This structural difference may be the cause of the different

SC
strain-rate sensitivities of the two samples at  nom  3 103 s1 . However, the measured
compressive strength can be sensitive to the characteristics and arrangement of the cells

NU
adjacent to the sample end due to the localised cell deformation under shock, which is
supported by the results shown in Figs. 7, 9 and 10.
MA
Fourthly, we analysed the case where the cell-wall material is rate-dependent and shock

occurs (e.g.  nom  3 103 s1 ). The corresponding FE results are presented in Figs. 5, 6b, 8-10
D

and Table 2. This is the most complicated case in our study, since the shock effect can interact
TE

with the effect of cell-wall rate dependence. We found that since shock restrains the cell
deformation (thus strain and strain-rate) close to the supporting end (see Fig. 8), the rate
P

dependence of the compressive strength at the supporting end is constrained with further
CE

increasing strain-rate (see Figs. 6b and 10 and Table 2). However, the stress at the loading end
is enhanced by both the shock effect (e.g. large inertial force) and the cell-wall rate dependence
AC

(see Fig. 5) as cell deformation is concentrated at the loading end (see Fig. 8).
Although the effect of internal gas is neglected in this study, it has been confirmed that the
trapped gas hardly affects the compressive strength (i.e. collapse stress) of aluminium foams
according to the previous analytical estimate [2, 14] and numerical simulation [42, 43]. The gas
effect could be significant at the densification stage [42, 43] or when the initial gas pressure is
much higher than atmosphere pressure [44, 45].
Some other physical factors, such as cell membrane damage mechanisms and
temperature-dependent material properties, may also affect the strain-rate sensitivity of the
compressive strength of foams. However, previous experimental observations [36, 46] show
that the plastic buckling of cell walls controls the cell collapse of Alporas foam and other
ductile foams. Consequently, the plasticity of cell walls is the dominant factor determining the
compressive strength of the Alporas foam studied here. Membrane damage mechanisms (e.g.
14
ACCEPTED MANUSCRIPT

cracking) and other irreversible processes (e.g. friction between contacting cell walls) are likely
to affect the compressive response at much larger strains after cell collapse for ductile foams.
The temperature effect can be important when foams are used in extreme temperature
environments or subjected to high speed impact. The foam strength will approach zero when

T
the environmental temperature approaches the melting point of the cell-wall material, whether

IP
the loading speed is low or high. On the other hand, the cell-wall temperature will rise due to
the conversion of plastic work into heat. As the heat transfer will be constrained at high

R
strain-rates, adiabatic condition will be applicable to the cell-wall material subjected to high

SC
speed impact [47]. By neglecting the elastic contribution and assuming perfectly plastic
deformation, the temperature change rate in cell walls can be obtained as    (  c)

NU
where  ,  ,  and c are cell-wall stress, strain-rate, mass density and specific heat
capacity, respectively, and  is the Taylor-Quinney coefficient representing the fraction of
MA
plastic work converted into thermoplastic heating. For the aluminium-based cell walls, the
typical values of the parameters are   0.9 ,   35.5 172 MPa ,   2.6  2.9 Mg  m3 and

c  920  1080 J  kg-1  K -1 [1, 29, 37, 47]. Then after 200 µs (i.e.   0.2 at   1103 s1 ) the
D
TE

temperature rise will be 2.0-12.9 K, which will cause negligible softening in comparison with

the strong strain-rate hardening (i.e.  Yd  Ys  163% at   110 s ).


3 1
P

5. Conclusions
CE

Three-dimensional FE models with high fidelity of meso-scale cell structures based on CT


AC

images have been developed to investigate the strain-rate effect on the compressive behaviour
of closed-cell aluminium foam (Alporas). Uniaxial compression has been simulated at different
nominal strain-rates (1×10-3-3×103 s-1). The numerical prediction of the rate dependence of the
compressive strength (i.e. first peak stress) is in reasonable agreement with the reported test
data. The simulations show that the rate dependence of the cell-wall material dominates the
macro strain-rate effect on the compressive strength measured at the supporting end, whereas
micro inertia has negligible contribution for Alporas foam. At sufficiently high strain-rates,
shock is initiated, which leads to higher stresses measured at the loading end, but it has a
complex effect on the stress at the supporting end, depending on the strain-rate sensitivity of
the cell-wall material and the characteristics of the cells contiguous to the sample end. In
general, shock restrains the development of cell deformation close to the supporting end when
foam collapse occurs, thereby limiting the stress level and the strain-rate sensitivity there. The

15
ACCEPTED MANUSCRIPT

numerical results also demonstrate that foams can be more strain-rate sensitive than their
cell-wall materials. This macroscopically observed higher strain-rate sensitivity is attributed to
the enhanced magnitude of local strain-rate in critical load-bearing elements in the cell walls.

Acknowledgements: The authors would like to acknowledge the assistance given by the IT Services

T
and the use of the Computational Shared Facility at The University of Manchester (UoM). The first

IP
author is grateful for the research scholarship offered by the School of Mechanical, Aerospace and Civil

R
Engineering at UoM and the Henry Lester Scholarship. Dr Rob Bradley’s help of CT image analysis is
also appreciated. The authors are grateful to the Engineering and Physical Science Research Council

SC
(EPSRC) for funding the Henry Moseley X-ray imaging facility through grants EP/F007906/1 and
EP/F028431/1. The second author, as an adjunct professor at Beijing Institute of Technology,

NU
acknowledges the support from State Key Laboratory of Explosion Science and Technology
(ZDKT11-03).
MA
D
P TE
CE
AC

16
ACCEPTED MANUSCRIPT

T
IP
R
SC
NU
MA
D
PTE
CE
AC

17
ACCEPTED MANUSCRIPT

References

[1] Gibson LJ, Ashby MF. Cellular solids: structure and properties. 2 ed: Cambridge University Press; 1997.

[2] Liu YD, Yu JL, Zheng ZJ, Li JR. A numerical study on the rate sensitivity of cellular metals. International
Journal of Solids and Structures. 2009;46:3988-98.

T
IP
[3] Ma GW, Ye ZQ, Shao ZS. Modeling loading rate effect on crushing stress of metallic cellular materials.
International Journal of Impact Engineering. 2009;36:775-82.

R
[4] Mukai T, Kanahashi H, Miyoshi T, Mabuchi M, Nieh T, Higashi K. Experimental study of energy absorption

SC
in a close-celled aluminum foam under dynamic loading. Scripta Materialia. 1999;40:921-8.

[5] Paul A, Ramamurty U. Strain rate sensitivity of a closed-cell aluminum foam. Materials Science and

NU
Engineering: A. 2000;281:1-7.

[6] Cady CM, Gray Iii GT, Liu C, Lovato ML, Mukai T. Compressive properties of a closed-cell aluminum foam
MA
as a function of strain rate and temperature. Materials Science and Engineering: A. 2009;525:1-6.

[7] Dannemann KA, Lankford Jr J. High strain rate compression of closed-cell aluminium foams. Materials
Science and Engineering: A. 2000;293:157-64.
D

[8] Ramachandra S, Sudheer Kumar P, Ramamurty U. Impact energy absorption in an Al foam at low velocities.
TE

Scripta Materialia. 2003;49:741-5.

[9] Shen J, Lu G, Ruan D. Compressive behaviour of closed-cell aluminium foams at high strain rates.
P

Composites Part B: Engineering. 2010;41:678-85.


CE

[10] Mukai T, Miyoshi T, Nakano S, Somekawa H, Higashi K. Compressive response of a closed-cell aluminum
foam at high strain rate. Scripta Materialia. 2006;54:533-7.
AC

[11] Irausquín I, Pérez-Castellanos JL, Miranda V, Teixeira-Dias F. Evaluation of the effect of the strain rate on
the compressive response of a closed-cell aluminium foam using the split Hopkinson pressure bar test. Materials
& Design. 2013;47:698-705.

[12] Elnasri I, Pattofatto S, Zhao H, Tsitsiris H, Hild F, Girard Y. Shock enhancement of cellular structures under
impact loading: Part I Experiments. Journal of the Mechanics and Physics of Solids. 2007;55:2652-71.

[13] Merrett RP, Langdon GS, Theobald MD. The blast and impact loading of aluminium foam. Materials &
Design. 2013;44:311-9.

[14] Deshpande VS, Fleck NA. High strain rate compressive behaviour of aluminium alloy foams. International
Journal of Impact Engineering. 2000;24:277-98.

[15] Li Z, Zhang J, Fan J, Wang Z, Zhao L. On crushing response of the three-dimensional closed-cell foam based
on Voronoi model. Mechanics of Materials. 2014;68:85-94.

[16] Song Y, Wang Z, Zhao L, Luo J. Dynamic crushing behavior of 3D closed-cell foams based on Voronoi
18
ACCEPTED MANUSCRIPT

random model. Materials & Design. 2010;31:4281-9.

[17] Calladine CR, English RW. Strain-rate and inertia effects in the collapse of two types of energy-absorbing
structure. International Journal of Mechanical Sciences. 1984;26:689-701.

[18] Maire E, Withers PJ. Quantitative X-ray tomography. International Materials Reviews. 2014;59:1-43.

T
IP
[19] Brydon AD, Bardenhagen SG, Miller EA, Seidler GT. Simulation of the densification of real open-celled
foam microstructures. Journal of the Mechanics and Physics of Solids. 2005;53:2638-60.

R
[20] Vesenjak M, Veyhl C, Fiedler T. Analysis of anisotropy and strain rate sensitivity of open-cell metal foam.

SC
Materials Science and Engineering: A. 2012;541:105-9.

[21] Jeon I, Asahina T, Kang K-J, Im S, Lu TJ. Finite element simulation of the plastic collapse of closed-cell

NU
aluminum foams with X-ray computed tomography. Mechanics of Materials. 2010;42:227-36.

[22] Veyhl C, Belova IV, Murch GE, Fiedler T. Finite element analysis of the mechanical properties of cellular
aluminium based on micro-computed tomography. Materials Science and Engineering: A. 2011;528:4550-5.
MA
[23] Sun YL, Lowe T, McDonald SA, Li QM, Withers PJ. In Situ Investigation and Image-Based Modelling of
Aluminium Foam Compression Using Micro X-Ray Computed Tomography. Visual Computing: Scientific
Visualization and Imaging Systems: Springer; 2014. p. 189-97.
D
TE

[24] McDonald SA, Mummery PM, Johnson G, Withers PJ. Characterization of the three-dimensional structure of
a metallic foam during compressive deformation. Journal of Microscopy. 2006;223:150-8.
P

[25] Hildebrand T, Rüegsegger P. A new method for the model-independent assessment of thickness in
three-dimensional images. Journal of Microscopy. 1997;185:67-75.
CE

[26] Simone AE, Gibson LJ. Aluminum foams produced by liquid-state processes. Acta Materialia.
1998;46:3109-23.
AC

[27] Young PG, Beresford-West TBH, Coward SRL, Notarberardino B, Walker B, Abdul-Aziz A. An efficient
approach to converting three-dimensional image data into highly accurate computational models. Philosophical
Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences. 2008;366:3155-73.

[28] Jeon I, Asahina T. The effect of structural defects on the compressive behavior of closed-cell Al foam. Acta
Materialia. 2005;53:3415-23.

[29] Jeon I, Katou K, Sonoda T, Asahina T, Kang K-J. Cell wall mechanical properties of closed-cell Al foam.
Mechanics of Materials. 2009;41:60-73.

[30] Simulia, ABAQUS Analysis User's Manual. ABAQUS v612 Documentation. Dassault Systémes, V6.12.

[31] Tan PJ, Reid SR, Harrigan JJ, Zou Z, Li S. Dynamic compressive strength properties of aluminium foams.
Part I—experimental data and observations. Journal of the Mechanics and Physics of Solids. 2005;53:2174-205.

[32] Zheng Z, Wang C, Yu J, Reid SR, Harrigan JJ. Dynamic stress-strain states for metal foams using a 3D

19
ACCEPTED MANUSCRIPT

cellular model. Journal of the Mechanics and Physics of Solids. 2014;72:93-114.

[33] Pattofatto S, Elnasri I, Zhao H, Tsitsiris H, Hild F, Girard Y. Shock enhancement of cellular structures under
impact loading: Part II analysis. Journal of the Mechanics and Physics of Solids. 2007;55:2672-86.

[34] Ramamurty U, Paul A. Variability in mechanical properties of a metal foam. Acta Materialia.

T
2004;52:869-76.

IP
[35] Ashby MF, Evans AG, Fleck NA, Gibson LJ, Hutchinson JW, Wadley HNG. Metal foams: a design guide:

R
Elsevier; 2000.

SC
[36] Markaki AE, Clyne TW. The effect of cell wall microstructure on the deformation and fracture of
aluminium-based foams. Acta Materialia. 2001;49:1677-86.

NU
[37] Idris MI, Vodenitcharova T, Hoffman M. Mechanical behaviour and energy absorption of closed-cell
aluminium foam panels in uniaxial compression. Materials Science and Engineering: A. 2009;517:37-45.

[38] Zou Z, Reid SR, Tan PJ, Li S, Harrigan JJ. Dynamic crushing of honeycombs and features of shock fronts.
MA
International Journal of Impact Engineering. 2009;36:165-76.

[39] Hu LL, Yu TX. Mechanical behavior of hexagonal honeycombs under low-velocity impact – theory and
simulations. International Journal of Solids and Structures. 2013;50:3152-65.
D
TE

[40] Liao S, Zheng Z, Yu J. Dynamic crushing of 2D cellular structures: Local strain field and shock wave
velocity. International Journal of Impact Engineering. 2013;57:7-16.
P

[41] Barnes AT, Ravi-Chandar K, Kyriakides S, Gaitanaros S. Dynamic crushing of aluminum foams: Part I –
Experiments. International Journal of Solids and Structures. 2014;51:1631-45.
CE

[42] Sun Y, Li QM. Effect of entrapped gas on the dynamic compressive behaviour of cellular solids.
International Journal of Solids and Structures. 2015;63:50-67.
AC

[43] Fang Q, Zhang J, Zhang Y, Gong Z, Wu H. A 3D mesoscopic model for the closed-cell metallic foams
subjected to static and dynamic loadings. International Journal of Impact Engineering. 2015;82:103–12.

[44] Zhang W, Xu Z, Wang TJ, Chen X. Effect of inner gas pressure on the elastoplastic behavior of porous
materials: A second-order moment micromechanics model. International Journal of Plasticity. 2009;25:1231-52.

[45] Xu ZM, Zhang WX, Wang TJ. Deformation of closed-cell foams incorporating the effect of inner gas
pressure. International Journal of Applied Mechanics. 2010;02:489-513.

[46] Bastawros AF, Bart-Smith H, Evans AG. Experimental analysis of deformation mechanisms in a closed-cell
aluminum alloy foam. Journal of the Mechanics and Physics of Solids. 2000;48:301-22.

[47] Meyers MA. Dynamic behavior of materials: John Wiley & Sons; 1994.

20
ACCEPTED MANUSCRIPT

T
IP
R
SC
NU
MA
D
PTE
CE
AC

21
ACCEPTED MANUSCRIPT

List of tables

Table 1 Geometrical data of the foam sample

Standard

T
Average Maximum Minimum
deviation

IP
Cell size (mm) 2.01 1.25 5.01 0.43

R
Cell-wall thickness (mm) 0.28 0.30 0.81 0.10

SC
NU
MA
D
P TE
CE
AC

22
ACCEPTED MANUSCRIPT

Table 2 Dynamic compressive strength normalised by the quasi-static one (  nom  1103 s1 )
for the large foam sample (Ø20.0×20.0 mm, loaded downward) and its cell-wall material

Foam with Foam with

T
Cell-wall
 nom rate-independent rate-dependent
material

IP
cell walls cell walls
1×100 s-1 1.00 1.10 1.09

R
1×102 s-1 1.00 1.36 1.33

SC
3×103 s-1 1.00 1.77 1.79

NU
MA
D
P TE
CE
AC

23
ACCEPTED MANUSCRIPT

Table 3 Ratio of transverse acceleration to longitudinal acceleration

 nom  1100 s1  nom  1102 s1  nom  3 103 s1

T
 nom  0.01  nom  0.07  nom  0.01  nom  0.07  nom  0.01  nom  0.07

IP
AX/AZ 1.17 1.23 1.23 1.62 1.02 1.00

R
AY/AZ 1.43 1.25 1.53 2.73 0.82 0.60

SC
NU
MA
D
P TE
CE
AC

24
ACCEPTED MANUSCRIPT

List of figure captions

Fig. 1. (a) Transverse slice through the 3D CT image of the scanned foam (Ø15.0×14.4 mm)
with a voxel size of 10.9 µm; (b) corresponding slice CT image of the ROI (Ø11.0×13.5

T
mm) with a voxel size of 50.0 µm after the image processing with the local cell-wall

IP
thickness (measured in 3D) represented as a colour-scale.
Fig. 2. Histograms of the cell size (a) and the local cell-wall thickness (b) of the virtual foam

R
sample (Ø11.0×13.5 mm) which has 115 cells in total.

SC
Fig. 3. Three-dimensional FE mesh (493150 quadratic tetrahedral elements and 957697 nodes)
of the foam sample created from the CT image: (a) perspective view of the 3D foam
body and enlarged view of the surface cells to show the mesh density and quality; (b)

NU
comparison of the cell structure before meshing and after meshing (the green contour
shows the boundary of the meshed surfaces and the white regions are the cell walls
segmented from CT image).
MA
Fig. 4. Large foam sample (Ø20.0×20.0 mm) used for numerical simulations: (a) transverse CT
slice for which the local cell-wall thickness is colour-coded; (b) 3D body of the foam
sample after meshing (2125544 quadratic tetrahedral elements and 4063952 nodes).
Fig. 5. Compressive stress-strain curves at a strain-rate of 3×103 s-1.
D
TE

Fig. 6. Compressive stress-strain curves measured at the supporting end at different strain-rates
with the yield stress of the cell-wall material being rate-independent (a) and
rate-dependent (b). Note that the stress-strain curves overlap at the strain-rates of
1×10-3-1×102 s-1 in (a).
P

Fig. 7. Equivalent plastic strain distribution in the internal cell walls at  nom  0.04 when the
CE

yield stress of cell-wall material is rate-independent. Half the compressed sample is


shown (top) whereas a corresponding vertical cross-section is depicted (bottom). The
AC

compressive load is applied downward from the top sample end. Also note that the
same plastic strain range 0.02-0.20 is used in the legend for the three loading cases, and
the maximum plastic strain is 1.198, 1.207 and 2.485 at the strain-rate of 1×10-3 s-1,
1×102 s-1 and 3×103 s-1, respectively.

Fig. 8. Equivalent plastic strain distribution in the internal cell walls at  nom  0.04 when the
yield stress of cell-wall material is rate-dependent. Half the compressed sample is
shown (top) whereas a corresponding vertical cross-section is depicted (bottom). The
compressive load is applied downward from the top sample end. Also note that the
same plastic strain range 0.02-0.20 is used in the legend for the three loading cases, and
the maximum plastic strain is 1.181, 0.970 and 1.468 at the strain-rate of 1×10-3 s-1,
1×102 s-1 and 3×103 s-1, respectively.
Fig. 9. Compressive stress-strain curves measured at the supporting end at high strain-rates
when compressive loads are applied downward and upward.
Fig. 10. Predicted strain-rate sensitivity of the compressive strength (collapse stress measured
at the supporting end) of closed-cell aluminium Alporas foam for rate dependent and
independent material models alongside available test data [4, 5, 8]. The dashed curve
25
ACCEPTED MANUSCRIPT

corresponds to the rate dependence of the cell-wall material. The dynamic compressive
strength is normalised by the quasi-static one (  nom  1103 s1 ).

Fig. 11. Local maximum principal strain-rate distribution in the internal cell walls at
 nom  0.04 and  nom  1102 s1 : (a) 3D view; (b) cross-sectional view.

T
R IP
SC
NU
MA
D
P TE
CE
AC

26
ACCEPTED MANUSCRIPT

Region of
interest (ROI)

ROI selection,
segmentation,

T
pore closing,
dilation,

IP
down-sampling

R
SC
NU
MA
Fig. 1. (a) Transverse slice through the 3D CT image of the scanned foam (Ø15.0×14.4 mm)
with a voxel size of 10.9 µm; (b) corresponding slice CT image of the ROI (Ø11.0×13.5 mm)
D

with a voxel size of 50.0 µm after the image processing with the local cell-wall thickness
TE

(measured in 3D) represented as a colour-scale.


P
CE
AC

27
ACCEPTED MANUSCRIPT

(a) 16 (b) 21

14 18

T
12
15

IP
Frequency (%)
10
Frequency (%)

12
8

R
9
6

SC
4

2 3

0 0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

NU
Cell size (mm) Local cell-wall thickness (mm)

Fig. 2. Histograms of the cell size (a) and the local cell-wall thickness (b) of the virtual foam
MA
sample (Ø11.0×13.5 mm) which has 115 cells in total.
D
P TE
CE
AC

(a) (b)

Fig. 3. Three-dimensional FE mesh (493150 quadratic tetrahedral elements and 957697 nodes) of the foam
sample created from the CT image: (a) perspective view of the 3D foam body and enlarged view of the
surface cells to show the mesh density and quality; (b) comparison of the cell structure before meshing and
after meshing (the green contour shows the boundary of the meshed surfaces and the white regions are the
cell walls segmented from CT image).

28
ACCEPTED MANUSCRIPT

T
R IP
SC
NU
MA

(a) (b)
D
TE

Fig. 4. Large foam sample (Ø20.0×20.0 mm) used for numerical simulations: (a) transverse CT
slice for which the local cell-wall thickness is colour-coded; (b) 3D body of the foam sample
P

after meshing (2125544 quadratic tetrahedral elements and 4063952 nodes).


CE
AC

29
ACCEPTED MANUSCRIPT

18

15 Supporting end, rate-dependent cell walls


Loading end, rate-dependent cell walls

Nominal stress (MPa)

T
Supporting end, rate-independent cell walls
12 Loading end, rate-independent cell walls

IP
9

R
6

SC
3

NU
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07
Nominal strain

Fig. 5. Compressive stress-strain curves at a strain-rate of 3×103 s-1.


MA
D
P TE
CE
AC

30
ACCEPTED MANUSCRIPT

T
(a) 3.0 -3 -1
(b) 3.0
Strain-rate = 1×10 s

IP
0 -1
Strain-rate = 1×10 s 3 -1
2.5 2 -1 2.5 3×10 s
Strain-rate = 1×10 s
3 -1
Nominal stress (MPa)

Nominal stress (MPa)


3 -1
Strain-rate = 1×10 s 1×10 s
3 -1

R
2.0 Strain-rate = 3×10 s 2.0
2 -1
1×10 s
0 -1
1×10 s

SC
1.5 1.5

-3 -1
1×10 s
1.0 1.0

NU
0.5 0.5

0.0 0.0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07
Nominal strain Nominal strain
MA
Fig. 6. Compressive stress-strain curves measured at the supporting end at different strain-rates
with the yield stress of the cell-wall material being rate-independent (a) and rate-dependent (b).
Note that the stress-strain curves overlap at the strain-rates of 1×10-3-1×102 s-1 in (a).
D
P TE
CE
AC

31
ACCEPTED MANUSCRIPT

T
R IP
SC
NU
MA
D

 nom  1103 s1  nom  1102 s1  nom  3 103 s1


TE

Fig. 7. Equivalent plastic strain distribution in the internal cell walls at  nom  0.04 when the
P

yield stress of cell-wall material is rate-independent. Half the compressed sample is shown
CE

(top) whereas a corresponding vertical cross-section is depicted (bottom). The compressive


load is applied downward from the top sample end. Also note that the same plastic strain range
AC

of 0.02-0.20 is used in the legend for comparison purpose in the figure, while the maximum
plastic strain is 1.198, 1.207 and 2.485 at the strain-rate of 1×10-3 s-1, 1×102 s-1 and 3×103 s-1,
respectively.

32
ACCEPTED MANUSCRIPT

T
R IP
SC
NU
MA
D
TE

 nom  1103 s1  nom  1102 s1  nom  3 103 s1


P
CE

Fig. 8. Equivalent plastic strain distribution in the internal cell walls at  nom  0.04 when the
yield stress of cell-wall material is rate-dependent. Half the compressed sample is shown (top)
AC

whereas a corresponding vertical cross-section is depicted (bottom). The compressive load is


applied downward from the top sample end. Also note that the same plastic strain range of
0.02-0.20 is used in the legend for comparison purpose in the figure, while the maximum
plastic strain is 1.181, 0.970 and 1.468 at the strain-rate of 1×10-3 s-1, 1×102 s-1 and 3×103 s-1,
respectively.

33
ACCEPTED MANUSCRIPT

(a) 3.5 (b) 3.5

3.0
Loaded downward 3.0 1×103 s-1
Loaded upward
Nominal stress (MPa)

Nominal stress (MPa)


2.5 2.5

T
3 -1
1×10 s
2.0 2.0
-3 -1 3×103 s-1
1×10 s

IP
1.5 1.5
1×10-3 s-1
1.0 1.0

R
3×103 s-1
Loaded downward
0.5 0.5

SC
Loaded upward

0.0 0.0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07
Nominal strain Nominal strain

Fig. 9. Compressive stress-strain curves measured at the supporting end at high strain-rates
when compressive loads are applied downward and upward. NU
MA
D
P TE
CE
AC

34
ACCEPTED MANUSCRIPT

T
Loaded upward
2.0
Cell-wall material

IP
(Cowper-Symonds equation)
Normalised compressive strength

1.8
FEM (rate-indepedent cell walls)
FEM (rate-dependent cell walls)

R
Mukai et al., 1999
1.6 Paul & Ramamurty, 2000 Loaded

SC
Ramachandra et al., 2003 downward

1.4

NU
1.2 Loaded upward

Loaded downward
1.0
MA
Small inertia Large inertia
0.8
-3 -2 -1 0 1 2 3
10 10 10 10 10 10 10
-1
Nominal strain-rate (s )
D

Fig. 10. Predicted strain-rate sensitivity of the compressive strength (collapse stress measured
TE

at the supporting end) of closed-cell aluminium Alporas foam for rate dependent and
independent material models alongside available test data [4, 5, 8]. The dashed curve
P

corresponds to the rate dependence of the cell-wall material. The dynamic compressive
CE

strength is normalised by the quasi-static one (  nom  1103 s1 ).


AC

35
ACCEPTED MANUSCRIPT

T
R IP
(a)

SC
NU
MA
D

(b)
TE

Fig. 11. Local maximum principal strain-rate distribution in the internal cell walls at
 nom  0.04 and  nom  1102 s1 : (a) 3D view; (b) cross-sectional view.
P
CE
AC

36
ACCEPTED MANUSCRIPT

T
IP
R
SC
NU
MA
D
TE
P
CE

Graphical Abstract
AC

37
ACCEPTED MANUSCRIPT

Highlights
 Simulations based on meso-scale realistic cell structures enable us to distinguish the
effects of different physical factors on the strain-rate sensitivity of aluminium Alporas
foam.
 The rate dependence of cell-wall material is the main cause of the strain-rate hardening

T
of compressive strength (first peak stress).
 Micro inertia has a negligible contribution.

IP
 Shock enhances the loading stress, but has complicated effects on the supporting stress,
depending on the strain-rate sensitivity of the cell-wall material and the characteristics

R
of the cells contiguous to the sample end.

SC
NU
MA
D
P TE
CE
AC

38

You might also like