Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Article

DOI: 10.1557/jmr.2020.157
Downloaded from https://www.cambridge.org/core. University of New England, on 05 Jul 2020 at 14:55:51, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2020.157

Dynamic mechanical properties of closed-cell aluminum


foams with uniform and graded densities
Ying Zhao1, Chenyang Ma1, Dabo Xin1,a), Ming Sun2
1
School of Civil Engineering, Northeast Forestry University, Harbin 150040, PR China
2
School of Civil Engineering, Harbin Institute of Technology, Harbin 150040, PR China
a)
Address all correspondence to this author. e-mail: xindabo@hit.edu.cn
Received: 26 November 2019; accepted: 1 June 2020

In this study, the quasi-static and dynamic mechanical behaviors and the energy absorption capacity of closed-
cell aluminum foams with uniform and graded densities were experimentally studied. The effects of density,
strain rate, and graded density on the mechanical performances of aluminum foams were quantitatively
evaluated. It was shown that the density had a significant effect on the quasi-static and dynamic compressive
stress of aluminum foams. Moreover, impact compression experiment results revealed that aluminum foam was
sensitive to the strain rate. As the strain rate increased, the plateau stress and energy absorption capacity
increased distinctly and the rate of deformation increased correspondingly. Finally, the investigation of aluminum
foams with uniform and graded densities to study their deformation and failure mechanisms, mechanical
characteristics, and energy absorption capacities showed that the GD 0.48-IV specimen exhibited superior impact
resistance. The present work can provide a valuable reference for the optimum design of aluminum foam against
impact loading.

Introduction [13] conducted experimental research on the dynamic com-


Closed-cell aluminum foam is a kind of cellular metallic mate- pressive behavior of closed-cell aluminum foam in a range of
rial that has extensive applications [1, 2]. Owing to its recycla- strain rates (400–2500 s−1) by using a Split-Hopkinson pres-
bility, high strength–weight ratio, and excellent energy sure bar (SHPB) experimental device. The test demonstrated
absorption and damping properties [3, 4, 5], closed-cell alumi- that the closed-cell aluminum foam was sensitive to the strain
num foam exhibits widespread applications in aerospace, auto- rate, and the strain rate sensitivity was more significant for a
mobile, and civil engineering structural applications; one of the relative density reaching 0.15. Raj et al. [14] reported the influ-
potential uses of closed-cell aluminum foam is in the core of ence of the quasi-static (0.001 s−1) and dynamic compressive
composite sandwich panels [6, 7, 8]. Closed-cell aluminum behaviors (750 s−1) of the closed-cell aluminum foam with a
foam has good performance and high application potential in range of relative density of 0.062–0.373. The study indicated
absorbing energy, and there is great research significance to that the plateau stress of aluminum foam increased with the
▪ Journal of Materials Research ▪ 2020 ▪ www.mrs.org/jmr

further explore its mechanical properties, energy-absorbing relative density and strain rate, and the energy absorption abil-
capacity, and deformation mechanisms. ity increased at a high strain rate.
Many researchers have investigated the mechanical proper- Other scholars hold the opposite viewpoint that the
ties of closed-cell aluminum foams under static and dynamic closed-cell aluminum foam is not sensitive to the strain rate.
loads at home and abroad in the past years, and many valuable Deshpande and Fleck [15] evaluated the compressive behavior
experimental results and conclusions had been obtained. of two cellular aluminum alloys (Alulight and Duocel) at strain
However, there are controversial views in academic circles con- rates reaching up to 5000 s−1 using an SHPB experiment. The
cerning whether closed-cell aluminum foam has a strain rate research indicated that the aluminum alloy foams had similar
effect. Some scholars believe that the dynamic response of mechanical behaviors under quasi-static and dynamic loading.
closed-cell aluminum foam is clearly dependent upon the The plateau stress and densification strain were insensitive to
strain rate effect [9, 10, 11, 12]. Dannemann and Lankford the strain rate but were dependent on the relative density of

© Materials Research Society 2020 cambridge.org/JMR 1


Article

the aluminum alloy foams. Ruan et al. [16] performed com- elasticity region, plateau region, and densification region. In
pressive tests on CYMAT aluminum foam with relative densi- the linear elasticity stage, the cell body showed elastic bending,
Downloaded from https://www.cambridge.org/core. University of New England, on 05 Jul 2020 at 14:55:51, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2020.157

ties ranging from 0.05 to 0.2 using an MTS high-rate test and the stress–strain curves were almost linear. As compression
system at strain rates of 10−3–10−1 s−1. The plateau stress is continued, the stress–strain curves entered the yield platform
insensitive to the strain rate and is related to the relative density region, and the cell bodies began to collapse, the stress remained
by a power law. Peroni et al. [17] studied the strain rate sensi- constant, and the strain increased rapidly. Finally, the closed-cell
tivity of closed-cell aluminum foams with two different densi- aluminum foam was fully compacted, and the stress rose signifi-
ties (approximately 150 and 300 kg/m3) using specimens of cantly as the strain increased and entered the densification stage.
diameter 50 mm with lengths varying between 15 and The cell wall along the direction of compression began to wrin-
50 mm. Their investigation revealed that no particular strain kle, crack, and gradually compact with an increase in strain. The
rate effects were observed in the range of 100–300 s−1. above compressive deformation process of aluminum foam is
In general, there are two types of aluminum foam speci- shown in Fig. 1(b). The deformation of aluminum foam with
mens: uniform density aluminum foam (single density com- three different densities after compaction is shown in Fig. 1(c).
pose) and graded density aluminum foam (compose of As shown, the density of aluminum foam was lower, and the
different densities). Most researchers are concerned with the compression deformation was greater.
compressive behavior of uniform density aluminum foam, The UD 0.48-3 layer specimen was made by three layers of
and few studies have been conducted on the mechanical prop- closed-cell aluminum foam with the same density of 0.48 g/cm3
erties of graded density aluminum foam. Most studies on the fastened together with the epoxy resin. Figure 1(d) shows the
mechanical response of graded density aluminum foam have stress–strain curves of specimens with UD 0.48 and UD
been investigated via the static experiment and numerical simu- 0.48-3 layer under quasi-static. It demonstrates that epoxy
lation, while no work has reported their application in a dynamic resin placed in specimen has no effect distinctly on the
compressive test, especially conducting an SHPB experiment. mechanical properties and the deformation characteristics of
Some researchers [18, 19, 20] studied the blast-resistant perfor- aluminum foam.
mances of graded density aluminum foams by numerical simu- Figure 1(e) shows the quasi-static compression stress–strain
lation and compared them to those with uniform densities. The curves of aluminum foams with uniform and graded densities.
results showed that the density arrangement of graded density Compared with uniform density, the aluminum foams with
aluminum foam had a significant influence on its blast-resistant graded densities had shown no significant stress plateau
behavior, and the optimization design of graded density alumi- stage. At the initial stage of the compression, the yield stress
num foam was superior to that of uniform density aluminum of the density graded aluminum foams was notable lower
foam. Xia et al. [21] conducted a series of quasi-static compres- than that of the uniform density aluminum foams with the
sive tests on uniform density aluminum foams, linear gradient same average density. As the compressive progressed, the
density, and unordered gradient density and analyzed the com- graded density aluminum foams revealed a continuous growing
pressive strength and energy absorption capacity of the alumi- stress, and the compressive stress of the GD 0.48-I, GD 0.48-II,
num foams with different density distributions. and GD 0.48-IV specimens exceeded successively that of the
In this study, a series of the quasi-static and dynamic UD 0.48 specimen at a certain strain. This suggested that the
experiments were carried on to examine the mechanical prop- graded density aluminum foams were compacted from the
erties and the energy absorption capacity of closed-cell alumi- lower-density layer to the higher-density layer regardless of
num foams with uniform density, in order to validate whether the loading direction. The compressive stress of the GD
or not it has a strain rate effect. Then, closed-cell aluminum 0.48-IV specimen was higher than that of the other graded
foams with uniform density and graded density were discussed density specimens, and the total energy absorption of the GD
▪ Journal of Materials Research ▪ 2020 ▪ www.mrs.org/jmr

in detail to compare their deformation and failure mechanisms, 0.48-IV specimen is larger than the UD 0.48 specimen,
mechanical characteristics, and energy absorption capacities. which indicated that the GD 0.48-IV specimen was superior
This research can provide reliable data and guidance for the to the UD 0.48 specimen in resisting compressive loading.
optimum design of aluminum foam against impact loads. Plateau stress is a key factor in the evaluation of the energy
dissipation capability of aluminum foam. The densification
strain and plateau stress are determined by the energy absorp-
Results and Discussion
tion efficiency [22].
Quasi-static characteristics discussion The energy absorption efficiency η is defined as follows:
Figure 1(a) shows the quasi-static compression stress–strain 1a
curves for uniform density aluminum foams with different s(1)d1
h(1a ) = 0
, 0 ≤ 1a ≤ 1, (1)
densities. The diagram shows three different regions: linear sa

© Materials Research Society 2020 cambridge.org/JMR 2


Article
Downloaded from https://www.cambridge.org/core. University of New England, on 05 Jul 2020 at 14:55:51, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2020.157

Figure 1: (a) Stress–strain curves of uniform density aluminum foams with different densities. (b) Compressive deformation process of uniform density aluminum
foam (UD 0.48). (c) Uniform density aluminum foam with different densities after compaction. (d) Stress–strain curves of UD 0.48 and UD 0.48-3 layer specimens
under quasi-static. (e) Stress–strain curves of aluminum foams with uniform and graded densities under quasi-static.

where σa and ea are the stress and corresponding strain, respec- when compared to that of the UD 0.36 specimen and were
tively. The densification strain ed denotes when the energy enhanced by 38.41% and 136.64%, respectively. This indicates
absorption efficiency η has reached the maximum, i.e., that density has a significant impact on the quasi-static com-
pressive stress of closed-cell aluminum foams.
dh(1a )
| = 0, 0 ≤ 1d ≤ 1. (2) Along with an increase in density, the elastic modulus and
d1 1a =1d
plateau stress of the closed-cell aluminum foam increased, and
The plateau stress σpl is derived as follows: the densification strain decreased because the matrix materials
1d forming the main body resisted the loads during the compression
s(1)d1 process. The matrix materials of aluminum foam increased with
spl = 0
. (3)
1d
TABLE 1: Quasi-static compressive property of closed-cell aluminum foam (UD
According to the above formulas, the mechanical properties 0.36, UD 0.48, and UD 0.60).
of uniform density aluminum foams can be obtained by the Specimen Actual density Elastic modulus Plateau stress Densification
(g/cm3)
▪ Journal of Materials Research ▪ 2020 ▪ www.mrs.org/jmr

quasi-static tests, as shown in Table 1. The stress–strain and code (MPa) (MPa) strain
energy absorption efficiency–strain curves of uniform density UD 0.36-1 0.359 171.05 4.54 0.56
aluminum foams with different densities are plotted in Fig. 2. UD 0.36-2 0.356 175.42 4.45 0.58
UD 0.36-3 0.356 153.40 4.59 0.56
The plateau stress values were obtained when the standard
Average value 0.357 166.62 4.53 0.57
variance value was under 0.1. The quasi-static compressive data UD 0.48-1 0.481 202.12 6.35 0.54
of the different specimens were obtained: the average plateau UD 0.48-2 0.477 209.17 6.25 0.53
UD 0.48-3 0.474 211.81 6.22 0.55
stress of UD 0.36 was 4.53 MPa (standard variance of 0.058); Average value 0.477 207.70 6.27 0.54
that of UD 0.48 was 6.27 MPa (standard variance of 0.056); UD 0.60-1 0.600 360.49 10.65 0.51
and that of UD 0.60 was 10.72 MPa (standard variance of UD 0.60-2 0.594 374.34 10.79 0.53
UD 0.60-3 0.599 358.03 10.72 0.53
0.057). The experimental results showed that the average pla- Average value 0.598 364.29 10.72 0.52
teau stress of UD 0.48 and UD 0.60 increased dramatically

© Materials Research Society 2020 cambridge.org/JMR 3


Article

displacement on the contact surface between the specimen


and the transmission bar was u2. According to the one-
Downloaded from https://www.cambridge.org/core. University of New England, on 05 Jul 2020 at 14:55:51, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2020.157

dimensional stress wave propagation theory [14], u1 and u2


are represented as formulas (4) and (5), respectively, as follows:
t
u1 = C0 [1I (t) − 1R (t)]dt, (4)
0

and
t
u2 = C0 1T (t)dt, (5)
0

where C0 is the elastic longitudinal wave velocity; eI, eR, and eT


are the strain gauge measurements representing the incident,
reflection, and transmission pulses, respectively; and l0 is the
initial length of a specimen. According to formulas (4) and
(5), the strain of a specimen can be expressed as follows:
t
u1 − u2 C0
1S (t) = = [1I (t) − 1R (t) − 1T (t)]dt. (6)
l0 l0 0

The load on both ends of the specimen are F1 and F2:

F1 = AE[1I (t) + 1R (t)], (7)

F2 = AE1T (t). (8)

The mean stress of a specimen is

F1 + F2 EA
sS (t) = = [1I (t) + 1R (t) + 1T (t)], (9)
2AS 2AS

where A and E are defined as the cross-sectional area and the


elastic modulus of the bars, respectively, and AS is the cross-
sectional area of a specimen.
According to the stress uniformity assumption,

1I (t) + 1R (t) = 1T (t). (10)

Substituting formula (10) into formulas (6) and (9), we


obtain

2C0 t
1S (t) = 1R (t)dt, (11)
l0 0
Figure 2: Stress–strain and energy absorption efficiency–strain curves of alumi-
num foams with different densities: (a) UD 0.36, (b) UD 0.48, and (c) UD 0.60.
and
▪ Journal of Materials Research ▪ 2020 ▪ www.mrs.org/jmr

AE
an increase in density, and their ability to resist elastic deforma- sS (t) = 1T (t). (12)
AS
tion and yield strength increased. The pore size of the aluminum
foam decreased with an increase in density, the foam’s collapse The strain rate of a specimen is
space was reduced accordingly, the plastic platform stage became 2C0
1̇S (t) = 1R (t). (13)
shorter, and the densification stage began earlier. l0

According to the above formulas, the stress–strain curves of


Dynamic characteristics discussion closed-cell aluminum foam at different strain rates can be
In the experiment, the displacement on the contact surface obtained so that the dynamic mechanical behavior can be ana-
between the specimen and the incident bar was u1, and the lyzed by SHPB experiments.

© Materials Research Society 2020 cambridge.org/JMR 4


Article
Downloaded from https://www.cambridge.org/core. University of New England, on 05 Jul 2020 at 14:55:51, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2020.157

Figure 3: Strain rate–time curves of aluminum foams under the average strain rate of 500 and 1000 s−1: (a) uniform density specimens and (b) graded density
specimens.

Figure 3 shows the strain rate–time curves of aluminum in Fig. 4(a) and under the average strain rate of 1000 s−1 shown
foam specimens with uniform density and graded density, in Fig. 4(b). From Figs. 4(a) and 4(b), the compressive stress of
which indicated that the average strain rate of specimens stud- uniform density closed-cell aluminum foams increased with an
ied in this paper were 500 and 1000 s−1, respectively. increase in density under the same strain rate.
The SHPB experiment primarily follows two fundamental When the average strain rate was 500 s−1, the average pla-
assumptions: one-dimensional stress wave propagation and teau stress of the UD 0.36 specimens was 5.23 MPa (standard
the stress uniformity assumption. Ravichandran and Subhash variance of 0.022); that of the UD 0.48 specimens was
[23] maintained that the stress uniformity was achieved when 7.01 MPa (standard variance of 0.082); and that of the UD
the stress waves propagated three times back and forth in the 0.60 specimens was 11.22 MPa (standard variance of 0.040).
specimen. The total time to reach stress uniformity (Δt) and The average plateau stresses of the UD 0.48 and UD 0.60 spec-
wave velocity (c) are represented as formulas (14) and (15), imens were improved by 34.03% and 114.53%, respectively,
respectively, as follows: compared to that of the UD 0.36 specimen.
When the average strain rate was 1000 s−1, the average pla-
Dt = 6L/c, (14)
teau stress of the UD 0.36 specimens was 13.16 MPa (standard
variance of 0.062); that of the UD 0.48 specimens was

c= E/r, (15) 16.72 MPa (standard variance of 0.060); and that of the UD
0.60 specimens was 26.57 MPa (standard variance of 0.053).
where L is the initial length of a specimen and E and ρ are the
The average plateau stresses of the UD 0.48 and UD 0.60 spec-
elastic modulus and the density of specimen, respectively.
imens were improved by 26.86% and 101.59%, respectively,
When the average strain rate was 500 s−1, the ratio of the
compared to that of the UD 0.36 specimen.
corresponding stress at time Δt to the plateau stress of the
The above analysis of the experimental data indicates that
UD 0.36 specimen was 9.18%; that of the UD 0.48 specimen
density has a significant influence on the plateau stress of
was 14.98%; and that of the UD 0.60 specimen was 8.29%.
closed-cell aluminum foam under impact loading. Generally,
When the average strain rate was 1000 s−1, the ratio of the cor-
the variation of plateau stress (σpl ) with relative density obeys
responding stress at time Δt to the plateau stress of the UD 0.36
▪ Journal of Materials Research ▪ 2020 ▪ www.mrs.org/jmr

the power law relationship [1]:


specimen was 14.74%; that of the UD 0.48 specimen was
8.37%; and that of the UD 0.60 specimen was 14.64%. Thus,  B
r
the ratio of the corresponding stress at time Δt to the plateau spl = A , (16)
rs
stress of the specimens studied in this paper was not more
than 15%, which proved that the stress uniformity assumption where A is the strengthening coefficient, ρ is the density of the
was satisfied during impact loading. aluminum foam, ρs is the density of the base material, ρ/ρs is
defined as the relative density, and B is the exponent to relative
Effect of density density.
The stress–strain curves of aluminum foams with three differ- Figure 4(c) shows the plateau stress changes as a function
ent densities under the average strain rate of 500 s−1 are shown of relative density at different strain rates. The experimental

© Materials Research Society 2020 cambridge.org/JMR 5


Article
Downloaded from https://www.cambridge.org/core. University of New England, on 05 Jul 2020 at 14:55:51, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2020.157

▪ Journal of Materials Research ▪ 2020 ▪ www.mrs.org/jmr

Figure 4: Stress–strain curves of aluminum foams with three different densi-


Figure 5: Stress–strain curves of specimens under different strain rates: (a) UD
ties under the average strain rate of (a) 500 s−1 and (b) 1000 s−1. (c) Plateau
0.36, (b) UD 0.48, and (c) UD 0.60.
stress as a function of relative density at different strain rates.

data for plateau stress as a function of relative density can be Effect of strain rate
fitted using the power law relationship of Eq. (14). The values The stress–strain curves of specimens UD 0.36, UD 0.48, and
of “A” and “B” were 169.32 and 1.85, respectively, at a strain UD 0.60 were analyzed and compared under different strain
rate of 0.001 s−1; 124.41 and 1.61, respectively, at a strain rate rates as shown in Fig. 5. The plateau stress increased signifi-
of 500 s−1; and 247.19 and 1.50, respectively, at a strain rate cantly when the average strain rate was doubled with specimens
of 1000 s−1. The value of “B” in this study was identified of the same density. As the average strain rate increased,
with the reported value (1.5–3.0) for aluminum foam [1]. the rate of deformation of the specimens increased

© Materials Research Society 2020 cambridge.org/JMR 6


Article

correspondingly. For the UD 0.36 specimen, the final strain was for the same specimen. This indicated that the deformation
0.12 under the average strain rate of 500 s−1 and 0.18 under the and failure modes of the same specimen were consistent
Downloaded from https://www.cambridge.org/core. University of New England, on 05 Jul 2020 at 14:55:51, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2020.157

average strain rate of 1000 s−1. The final strain of the UD 0.48 under the average strain rate of 500 and 1000 s−1.
specimen was 0.12 under the average strain rate of 500 s−1 and Under the average strain rate of 1000 s−1, the aluminum
0.16 under the average strain rate of 1000 s−1. The final strain foams with uniform density and graded density were
of the UD 0.60 specimen was 0.11 under the average strain rate completely compacted. Figure 6(c) shows the before and after
of 500 s−1 and 0.15 under the average strain rate of 1000 s−1. test photos of the specimens under the average strain rate of
In comparison with the quasi-static loading, the plateau 500 s−1. The layer 1 bottom side of the specimen is defined
stress of the UD 0.36 specimens increased by 15.45% and as the impact wave propagation direction. For the UD 0.48
190.95% under the average strain rates of 500 and 1000 s−1, specimen, the cell wall of the specimen was squeezed and
respectively; that of the UD 0.48 specimens increased by began to wrinkle along the direction of compression and gen-
11.80% and 166.67% under the average strain rates of 500 erated cracks which expanded to a partial collapse and then
and 1000 s−1, respectively; and that of the UD 0.60 specimens triggered a greater area of collapse. The closed-cell aluminum
increased by 4.66% and 147.85% under the average strain rates foam was completely compacted under a higher strain rate.
of 500 and 1000 s−1, respectively. The results show that The initiation and propagation of cracks along the entire height
closed-cell aluminum foam has an obvious strain rate effect, of the UD 0.48 specimen can be seen in the fracture specimen.
and the lower the density, the more the strain rate effect. As shown in Fig. 6(c), at layer 1 of the GD 0.48-I specimen,
The strain rate effect under dynamic compressive is the cell wall buckled and fractured, and part of the cell body
described by the Cowper–Symonds model as follows: collapsed. At layer 3 of the GD 0.48-II specimen, most of the
 P1 cell bodies were compacted, and the cracks extended to the
sdyn 1̇ middle layer. At layer 1 and layer 3 of the GD 0.48-III speci-
=1+ , (17)
sstatic C men, the cell bodies were completely compacted, and the
where σdyn/σstatic is the dynamic hardening coefficient, 1̇ is the local cell wall began to wrinkle at layer 2. At layer 2 of the
strain rate, and C and P are the material parameters. GD 0.48-IV specimen, the cell wall bent and fractured, part
For the UD 0.36 specimen, the Cowper–Symonds parame- of the cell body collapsed, and layer 1 and layer 3 had no obvi-
ters can be obtained by the fitting data C = 831.31 and P = 0.27, ous deformation.
for the UD 0.48 specimen by the fitting data C = 880.07 and P The uniform density aluminum foams were compacted
= 0.27, and for the UD 0.60 specimen by the fitting data C = layer by layer along the direction of impact loading. However,
934.49 and P = 0.21. owing to the defect of the pore cell structure, the first collapse
of the cell body may occur at the interface of the aluminum
foam and the protected structure.
Effect of graded density The graded density aluminum foams were compacted from
The stress–strain curves of aluminum foams with uniform den- the lower-density layer to the higher-density layer regardless of
sity and graded density under the average strain rate of 500 and the impact wave propagation direction. After the lower-density
1000 s−1 are compared in Figs. 6(a) and 6(b), respectively. layer was completely compacted, the stress of the graded den-
Unlike the uniform density aluminum foams, those with sity aluminum foams grew further, but not the plateau stress
graded density displayed no obvious plateau region, except region as in the uniform density aluminum foams. Therefore,
for the GD 0.48-III specimen. At the early stage of the com- density graded aluminum foams can adapt to more changeable
pression process, the initial yield stress of the density graded situations because of the variations in the plateau stress.
aluminum foams was lower than that of the uniform density
▪ Journal of Materials Research ▪ 2020 ▪ www.mrs.org/jmr

aluminum foams with the same average density. As compres-


Energy absorption characteristics
sion continued, the compressive stress of the GD 0.48-IV speci-
men exceeded that of the UD 0.48 specimen at a certain strain. The closed-cell aluminum foam is a porous metallic material
The compressive stresses of the GD 0.48-I, GD 0.48-II, and GD with a good capacity for dissipating energy and cushioning.
0.48-III specimens were lower than the plateau stress of the UD The energy absorption of aluminum foam can be expressed
0.48 specimen. The GD 0.48-IV specimens had the highest as the energy absorbed per unit volume during the compres-
compressive strength, and the GD 0.48-III specimens had the sion process [24] as follows:
lowest compressive strength compared to the other graded den- 1 
n

sity specimens in this study. It is observed that the compressive W= s(1)d1 = si 1i , (18)
0 i=1
stress increases significantly as the average strain rate increases
but has not changed the configuration of stress–strain curves where σ and e are the compressive stress and strain,

© Materials Research Society 2020 cambridge.org/JMR 7


Article
Downloaded from https://www.cambridge.org/core. University of New England, on 05 Jul 2020 at 14:55:51, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2020.157

Figure 6: Stress–strain curves of aluminum foams with uniform and graded densities under the average strain rate of (a) 500 s−1 and (b) 1000 s−1. (c) Before and
after tests photos of specimens: UD 0.48, GD 0.48-I, GD 0.48-II, GD 0.48-III, and GD 0.48-IV under the average strain rate of 500 s−1.

respectively, and W is the area of the compressive curve under a percentage increase of energy absorption compared to quasi-
certain strain. static loading did not show monotonous relations.
The energy absorption capabilities of specimens with dif- Figure 7(b) shows the energy absorption of uniform density
ferent densities under different average strain rates are shown and graded density specimens under the average strain rate of
in Fig. 7(a). The limit in the integral for obtaining energy 500 s−1. The energy absorption of the UD 0.48 specimen was
▪ Journal of Materials Research ▪ 2020 ▪ www.mrs.org/jmr

absorption is the area of stress–strain curves when the strain approximately equal to that of the GD 0.48-IV specimen, and
is 0.1 from Fig. 5. Under the average strain rates of 500 and the energy dissipation capacity was the best compared to the
1000 s−1, compared to quasi-static loading, the energy absorp- other specimens. The energy absorption of the GD 0.48-I,
tion of the UD 0.36 specimen increased by 31.82% and GD 0.48-II, and GD 0.48-III specimens decreased by 22.67%,
215.91%, respectively; that of the UD 0.48 specimen increased 32%, and 41.33%, respectively, in comparison to that of the
by 48.08% and 228.85%, respectively; and that of the UD 0.60 GD 0.48-IV specimen. Thus, the energy absorption capacity
specimen increased by 25.77% and 165.98%, respectively. The of the GD 0.48-III specimen was the worst. Based on the anal-
absorption energy of closed-cell aluminum foams with the ysis of the "Effect of graded density" section, density graded
same density exhibits a greater energy absorption capacity at aluminum foams can adapt to more changeable situations
higher strain rates. Under the same strain rate, the energy because of the variations in the plateau stress, which showed
absorption was increased as density increases, but the that the GD 0.48-IV specimen exhibited superior impact

© Materials Research Society 2020 cambridge.org/JMR 8


Article

was superior to that of the UD 0.48 specimen, and the energy


dissipation capacity was the best compared to the other speci-
Downloaded from https://www.cambridge.org/core. University of New England, on 05 Jul 2020 at 14:55:51, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2020.157

mens. The energy absorption of the UD 0.48, GD 0.48-I, GD


0.48-II, and GD 0.48-III specimens decreased by 8%, 20.4%,
28%, and 38%, respectively, in comparison to that of the GD
0.48-IV specimen. Therefore, the GD 0.48-IV specimen
would provide the better energy absorption performance and
impact resistance at a higher strain rate than that of the UD
0.48 specimen.

Conclusions
In this study, the compressive mechanical properties and
energy absorption capacity of closed-cell aluminum foams
with uniform and graded densities were experimentally studied
under quasi-static and impact loading. The effects of density,
strain rate, and graded density on the dynamic characteristics
of closed-cell aluminum foams were also experimentally inves-
tigated, and the strain rate effect constitutive model of
closed-cell aluminum foam under impact loading was fitted
by the Cowper–Symonds exponential function. The main con-
clusions of this work are summarized as follows:

(i) The density of closed-cell aluminum foams has a


significant impact on the compressive stress under
quasi-static loading. As the density increases, the elastic
modulus and plateau stress of the closed-cell aluminum
foam increases and the densification strain decreases.
(ii) The density of closed-cell aluminum foam has a
significant influence on the plateau stress under impact
loading. The relation curve between plateau stress and
relative density was fitted by using the power law function
under dynamic compression.
(iii) Closed-cell aluminum foam has an obvious strain rate
effect, and the plateau stress increases significantly and the
rate of deformation increases correspondingly as the
strain rate increases. The closed-cell aluminum foam
exhibits a greater energy absorption capacity at a higher
strain rate.
(iv) Compared with uniform density, the aluminum foam with
▪ Journal of Materials Research ▪ 2020 ▪ www.mrs.org/jmr

graded densities has shown no significant stress plateau


Figure 7: (a) Energy absorption of specimens with different densities under stage, which indicated that density graded aluminum
different average strain rates. Energy absorption of uniform density and graded foam could adapt to more situations owing to its changing
density specimens under the average strain rate of (b) 500 s−1 and (c)
1000 s−1.
plateau stress. Uniform density aluminum foams are
compacted layer by layer along the loading direction. The
graded density aluminum foams are compacted from a
resistance, although energy absorption of the GD 0.48-IV and lower-density layer to a higher-density layer regardless of
UD 0.48 specimen are the same. the loading direction. The GD 0.48-IV specimen is
Figure 7(c) shows the energy absorption of uniform density superior to other specimens studied in this paper in
and graded density specimens under the average strain rate of resisting quasi-static compressive loading and impact
1000 s−1. The energy absorption of the GD 0.48-IV specimen loading.

© Materials Research Society 2020 cambridge.org/JMR 9


Article
Downloaded from https://www.cambridge.org/core. University of New England, on 05 Jul 2020 at 14:55:51, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2020.157

Figure 8: Configuration of graded density aluminum foam specimens: (a) GD 0.48-I, (b) GD 0.48-II, (c) GD 0.48-III, and (d) GD 0.48-IV.

Materials and Experiments test machine [Fig. 9(a)] with a constant loading rate of 2
mm/min (strain rate 0.001 s−1) at room temperature. Three
Specimens
specimens were provided for each quasi-static test to reduce
To compare the influences of density on the mechanical perfor- experimental error. The complete stress–strain curves of uniform
mance of closed-cell aluminum foams, aluminum foam speci- density aluminum foam from the compressive tests were used to
mens with uniform density and graded density were made. study the mechanical properties of the materials such as the elas-
Uniform density aluminum foams with densities of 0.36 g/cm3 tic modulus, plateau stress, and densification strain.
(the cell size is 3–5 mm), 0.48 g/cm3 (the cell size is 3–4 mm),
and 0.60 g/cm3 (the cell size is 2–3 mm) were prepared in this
study. The specimens were cut from large aluminum foam blocks
Dynamic test
with three different densities into cylinders with diameters of
50 mm and a height of 50 mm for quasi-static testing and diam- To perform a comparative analysis of the differences between
eters of 50 mm and a height of 30 mm for dynamic testing with the dynamic and quasi-static mechanical behaviors of alumi-
the adoption of a wire electrical-discharge machining technology. num foams with uniform density and the differences in the
The specimen dimensions were more than five times the cell size dynamic mechanical properties of uniform density and graded
to ignore scale effect, and the height-to-diameter ratio of each density aluminum foam under the same average density, a
▪ Journal of Materials Research ▪ 2020 ▪ www.mrs.org/jmr

specimen was between 0.5:1 and 1:1 to reduce inertia effect number of tests were performed using an SHPB device with
[25, 26]. The graded density specimens were made by three layers different strain rates at room temperature. The face of a speci-
of closed-cell aluminum foam with different densities fastened men connected to the incident bar was defined as the wave
together with epoxy as shown in Fig. 8, and the layer 1 bottom front, and the face connected to the transmission bar was
side of the specimen is in contact with incident bar. The average defined as the wave back. Three specimens were provided for
density of all the graded specimens was 0.48 g/cm3. each SHPB test to obtain more accurate data.
The SHPB technique was first presented by Hopkinson in
1941 to measure the pulse waveform of impact load.
Quasi-static test Researchers [27, 28, 29] then proposed an improved SHPB
The uniaxial compressive tests of the uniform density alumi- apparatus that obtained stress versus strain curves of materials
num foams were performed using an Instron 5505 universal at a high strain rate. A typical SHPB equipment [Fig. 9(b)]

© Materials Research Society 2020 cambridge.org/JMR 10


Article
Downloaded from https://www.cambridge.org/core. University of New England, on 05 Jul 2020 at 14:55:51, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2020.157

Figure 9: (a) Quasi-static compres-


sive test. (b) Schematic of SHPB
equipment. (c) SHPB test setup.

mainly includes a bullet launcher, striker bar (bullet), incident Acknowledgments


bar, transmission bar, and data acquisition system. The speci-
This research was supported by the National Key Research
men is sandwiched between the incident bar and the transmis-
and Development Program of China (Grant Nos.
sion bar. An air compression engine is used to launch a bullet
2018YFC0809605 and 2018YFC0809600), the Fundamental
at the impact incident bar to generate a compression wave. The
Research Funds for the Central Universities (2572018BJ04),
compressive wave propagates along the incident bar and
and the Scientific and Technological Innovation Talent
reaches the specimen. A part of the wave is reflected back,
Program of Harbin (2017RAQXJ014).
and another part is transmitted through the specimen and
into the transmission bar as a transmitted wave.
In this study, the SHPB experimental facility was adapted
References
according to the size and property of each specimen as
1. L.J. Gibson and M.F. Ashby: Cellular Solids Structures and
▪ Journal of Materials Research ▪ 2020 ▪ www.mrs.org/jmr

shown in Fig. 9(c). The diameter of the striker bar (bullet),


Properties, 2nd ed. (Cambridge University Press, Cambridge, UK,
the incident bar, and the transmission bar was 100 mm. The
striker bar was 900 mm long, the incident bar and the trans- 1997).

mission bar were 3300 mm long, and the absorption bar was 2. M.F. Ashby, A.G. Evans, N.A. Fleck, L.J. Gibson,

2700 mm long. The material of all the bars was aluminum J.W. Hutchinson, and H.N.G. Wadley: Metal Foams: A Design
with a density of 2.7 g/cm3. To improve the signal pulse ampli- Guide (Butterworth-Heinemann, USA, 2000).
tude of a transmitted wave, a semiconductor strain gauge was 3. D.P. Mondal, M.D. Goel, and S. Das: Compressive deformation
used to acquire the weak transmission signal on the transmitted and energy absorption characteristics of closed cell aluminum-fly ash
bar. The sensitivity of the semiconductor strain gauge could be particle composite foam. Mater. Sci. Eng. A 507, 102–109 (2009).
increased approximately 50-fold compared with that of a stan- 4. J.A. Reglero, E. Solórzano, M.A. Rodríguez-Pérez, J.A. de Saja,
dard strain gauge. and E. Porras: Design and testing of an energy absorber

© Materials Research Society 2020 cambridge.org/JMR 11


Article

prototype based on aluminum foams. Mater. Des. 31, 3568–3573 17. M. Peroni, G. Solomos, and V. Pizzinato: Impact behaviour
(2010). testing of aluminium foam. Int. J. Impact Eng. 53, 74–83 (2013).
Downloaded from https://www.cambridge.org/core. University of New England, on 05 Jul 2020 at 14:55:51, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2020.157

5. A. Aldoshan and S. Khanna: Effect of relative density on the 18. J. Li, G. Ma, H. Zhou, and X. Du: Energy absorption analysis of
dynamic compressive behavior of carbon nanotube reinforced density graded aluminium foam. Int. J. Protect. Struct. 2, 333–350
aluminum foam. Mater. Sci. Eng. A 689, 17–24 (2017). (2011).
6. J. Banhart and H.W. Seeliger: Aluminium foam sandwich panels: 19. J.D. Li, H.Y. Zhou, and G.W. Ma: Numerical simulation of blast
Manufacture, metallurgy and applications. Adv. Eng. Mater. 10, mitigation cladding with gradient metallic foam core. Appl. Mech.
793–802 (2008). Mater. 82, 461–466 (2011).
7. K. Mohan, T.H. Yip, S. Idapalapati, and Z. Chen: Impact 20. Q. Chang, Y. Li-jun, and Y. Shu: Simulation and optimization of
response of aluminum foam core sandwich structures. Mater. Sci. blast-resistant performance of graded aluminum foam sandwich
Eng. A 529, 94–101 (2011). structures. J. Vib. Shock. 32, 70–75 (2013).
8. I. Elnasri and H. Zhao: Impact perforation of sandwich panels 21. Y. Xia, C.Q. Wu, Z.X. Liu, and Y.M. Yuan: Protective effect
with aluminum foam core: A numerical and analytical study. Int of graded density aluminium foam on RC slab under blast load-
J. Impact Eng. 96, 50–60 (2016). ing – An experimental study. Constr. Build. Mater. 111, 209–222
9. T. Mukai, H. Kanahashi, T. Miyoshi, M. Mabuchi, T.G. Nieh, (2016).
and K. Higashi: Experimental study of energy absorption in a 22. Z.H. Wang, J.H. Shen, G.X. Lu, and L.M. Zhao: Compressive
close-celled aluminum foam under dynamic loading. Scr. Mater. behavior of closed-cell aluminum alloy foams at medium strain
40, 921–927 (1999). rates. Mater. Sci. Eng. A 528, 2326–2330 (2011).
10. H. Zhao, I. Elnasri, and S. Abdennadher: An experimental study 23. G. Ravichandran and G. Subhash: Critical appraisal of limiting
on the behaviour under impact loading of metallic cellular mate- strain rates for compression testing of ceramics in a Split-
rials. Int. J. Mech. Sci. 47, 757–774 (2005). Hopkinson pressure bar. J. Am. Ceram. Soc. 77, 263–267 (1994).
11. I. Irausquín, J.L. Pérez-Castellanos, V. Miranda, and 24. H.W. Song, Z.J. Fan, and G. Yu: Partition energy absorption of
F. Teixeira-Dias: Evaluation of the effect of the strain rate on the axially crushed aluminum foam-filled hat sections. Int. J. Solids
compressive response of a closed-cell aluminium foam using the Struct. 42, 2575–2600 (2005).
split Hopkinson pressure bar test. Mater. Des. 47, 698–705 (2013). 25. W. Chen, B. Zhang, and M.J. Forrestal : A split Hopkinson
12. P.F. Wang, S.L. Xu, Z.B. Li, J.L. Yang, H. Zheng, and S.S. Hu: bar technique for low impedance materials. Exp. Mech. 39, 81–85
Temperature effects on the mechanical behavior of aluminum foam (1999).
under dynamic loading. Mater. Sci. Eng. A 599, 174–179 (2014). 26. M.I. Idris, T. Vodenitcharova, and M. Hoffiman: Mechanical
13. K.A. Dannemann and J. Lankford, Jr.: High strain rate com- behaviour and energy absorption of closed-cell aluminium foam
pression of closed-cell aluminium foams. Mater. Sci. Eng. A 293, panels in uniaxial compression. Mater. Sci. Eng. A 517, 37–45
157–164 (2000). (2009).
14. R.E. Raj, V. Parameswaran, and B.S.S. Daniel: Comparison of 27. M.A. Kariem, D. Ruan, J.H. Beynon, and D.A. Prabowo: Mini
quasi-static and dynamic compression behavior of closed-cell round-Robin test on the Split-Hopkinson pressure bar. J. Test.
aluminum foam. Mater. Sci. Eng. A 526, 11–15 (2009). Eval. 46, 457–468 (2018).
15. V.S. Deshpande and N.A. Fleck: High strain rate compressive 28. B.A. Gama, S.L. Lopatnikov, and J.W. Gillespie, Jr.: Hopkinson
behaviour of aluminium alloy foams. Int. J. Impact Eng. 24, 277– bar experimental technique: A critical review. Appl. Mech. Rev. 57,
298 (2000). 223–250 (2004).
16. D. Ruan, G. Lu, F.L. Chen, and E. Siores: Compressive behaviour 29. S.S. Hu, L.L. Wang, L. Song, and L. Zhang: Review of the
of aluminium foams at low and medium strain rates. Compos. development of Hopkinson pressure bar technique in China.
▪ Journal of Materials Research ▪ 2020 ▪ www.mrs.org/jmr

Struct. 57, 331–336 (2002). Explos. Shock Waves 34, 641–657 (2014).

© Materials Research Society 2020 cambridge.org/JMR 12

You might also like