You are on page 1of 9

Acta Materialia 55 (2007) 2291–2299

www.actamat-journals.com

Density dependence of the compressive properties of porous copper


over a wide density range
Masataka Hakamada *, Yuuki Asao, Tetsumune Kuromura, Youqing Chen,
Hiromu Kusuda, Mamoru Mabuchi
Department of Energy Science and Technology, Graduate School of Energy Science, Kyoto University, Yoshidahonmachi, Sakyo, Kyoto 606-8501, Japan

Received 2 August 2006; received in revised form 14 November 2006; accepted 16 November 2006
Available online 19 January 2007

Abstract

The compressive properties of porous copper with relative densities, q/qs, of 0.22–0.96 were investigated. In the low relative density
range (q/qs < 0.5–0.6), porous copper showed a density exponent n of 2.3, where n represents the relative density dependence of yield
strength. In this range, the bending and buckling of cell walls and the formation of macroscopic deformation bands were observed. How-
ever, porous copper with a higher relative density (0.5–0.6 < q/qs < 0.9–1) had an n value of 1, where the dominant deformation mode
of cell walls was yielding, and no clear deformation band was observed. Also, in the highest relative density range (q/qs very close to 1),
the compressive properties degraded markedly with decreasing density, indicating that stress concentration around the minimal pores
occurred in this density range.
 2006 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Compression test; Foams; Porous material; Mechanical properties; Relative density

1. Introduction that is, the relative yield strength and relative density, are
used comprehensively to understand the relationship
Porous metals or metallic foams are emerging materials between the compressive strength and density of various
which possess many attractive properties, such as very low porous materials [1,10]. The relative yield strength (ry/
density, large compressive deformation at a constant flow rys) is the yield stress of a porous material (ry) divided
stress, large surface area, and thermal and chemical resis- by the yield stress of its cell strut (or wall) solid (rys). ry
tance [1–3]. They are promising candidates for structural is often replaced by the plateau stress (the flow stress of a
use [1–3], heat exchangers [2], filters [4], sound absorbers plateau region in the compressive stress–strain curve) of a
[5–7] and artificial bones [8,9], among others. In particular, porous material [1]. The relative density (q/qs) is the den-
one of the most promising applications of porous metals is sity of a porous material (q) divided by the density of its
in vehicles equipped with an impact-absorbing material, cell strut (or wall) solid (qs). The two variables are related
because they are very light and can absorb the impact by [1,10]
energy through their large compressive deformation at a  n
constant flow stress [1–3]. For such application, the defor- ry q
¼C ð1Þ
mation behavior of porous metals under compression must rys qs
be well understood.
where C is a constant for the geometric effect, and n is the
The compressive properties of porous metals mostly
density exponent, which is a constant giving the relative
depend on the density of the materials [1]. Two variables,
density dependence of the yield strength of a porous
*
Corresponding author. Tel.: +81 75 753 5421; fax: +81 75 753 5428.
material.
E-mail address: hakamada@g03.mbox.media.kyoto-u.ac.jp (M. Haka- To date, there are many studies on the deformation of
mada). porous metals based on compression tests [1–3,11–28].

1359-6454/$30.00  2006 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actamat.2006.11.024
2292 M. Hakamada et al. / Acta Materialia 55 (2007) 2291–2299

Fig. 1 and Table 1 summarize the literature data for the ties of porous metals depend on their relative density, the
relationship between the relative density and the relative pore geometry and size distribution must be similar over
yield strength of various metals, ranging from high to a wide density range.
low porosity metals [11–28]. The slope of the double loga- The value of n in Eq. (1) may be governed by the defor-
rithmic plots for each material in Fig. 1 corresponds to the mation mode of individual cell struts or walls, such as
value of n in Eq. (1). Table 1 contains the values of n which yielding, bending and buckling [1,10,29]. In particular,
are calculated from the data in the literature. It can be seen the bending and buckling modes are peculiar to porous
from Fig. 1 that the scatter in relative yield strength is large materials consisting of cell struts or walls. Hence, it is
in a lower density range. A certain porous metal in the worth investigating how the individual cell struts or walls
lower density range exhibits a relative yield strength >10 are deformed when the relative density is varied in a wide
times higher than that of another porous metal, when the range. Another feature of the compression behavior of por-
two have identical relative density. The large scatter in rel- ous metals is the formation of deformation bands, i.e.,
ative yield strength is probably due to disunity in pore bands of deformed cells which are aligned vertically to
characteristics such as the shape and size, resulting in a the loading direction [11,30–32]. It is known that deforma-
large variation in the value of C in Eq. (1). It is of interest tion band formation, and thus the compressive properties
to note from Table 1 that the values of n are in the range of porous materials, are affected by pore characteristics
1.5–3 for porous metals with low relative density [11–23]; such as pore shape [29,33–37] and structural defects [38–
however, for porous metals with higher relative density, 44] (missing cell walls, elongated pores and significant cell
there is large variation in the n value from 1.0 to 6.3 [24– wall curvature, among others). Therefore, for a proper
28]. Therefore, it appears that there is a ‘critical relative understanding of the relative density dependence of struc-
density’ which divides the compressive properties of porous tural deformation characteristics (the deformation mode
metals, affecting the values of n. However, the literature of individual cell struts or walls and the formation of a
data in Fig. 1, which are scattered by the disunity of the deformation band), it is required that pore shape is united
pore characteristics, may give uncertain information for and that the structural defects are sufficiently limited.
the relative density dependence of the compressive Fine control of pore characteristics can be attained by
strength. To know precisely how the compressive proper- spacer methods in which metal matrices containing spacer
particles are processed, after which the spacer particles are
removed [8,18,19,23,45–53]. In the present paper, porous
100
copper specimens with various relative densities ranging
from 0.22 to 0.96 are fabricated via the spacer method,
and the compressive properties of porous copper are exam-
ined to elucidate the dependence of the compressive defor-
10-1 mation behavior on the densities of porous metals. In
addition, deformation characteristics such as the deforma-
tion mode of individual cell walls and the formation of a
deformation band are discussed at low and high densities.
Relative yield strength, σ y /σ ys

10-2 2. Experimental
Al-Si-Mg(SiC) [11]
Al-Si-Cu [12] 2.1. Specimen preparation
Zn-Cu [12]
Al [13]
10-3
Al-Mg [13] Pure copper powder with an average powder size of
Al 7075 [13]
Zn [14]
1.5 lm, purchased from Fukuda Metal Foil & Powder
Al(SiC) [15] Co. Ltd, was used as a starting material. Ammonium
Al-Si [16]
Al-Cu [17] hydrogen carbonate (NH4HCO3) particles, purchased from
Al [18]
Al [19]
Nacalai Tesque, were sieved and the particles with average
10-4 Al [20] sizes of 300–425 lm were prepared as spacer particles. The
Al-Si-Mg [21]
Al (spherical pore) [23] copper powder and spacer particles were thoroughly mixed
Al (angular pore) [23] in an agate mortar. The mixing ratios were varied according
Cu (5 μm) [24]
Cu (45 μm) [24] to the desired relative densities. The mixed powder was
Fe [25]
Cu-Sn [25] uniaxially pressed at a pressure of 100 MPa into green com-
10-5 Fe [26] pacts. The compact was then sintered in a tubular furnace
Be [27]
under an Ar + 5 vol.% H2 atmosphere. During sintering,
10-1 100 the NH4HCO3 spacer particles decomposed to gaseous spe-
Relative density, ρ /ρs
cies (ammonia, water and carbon dioxide), leaving pores
Fig. 1. Relationship between relative density and relative yield strength of behind. To avoid the burst of green compact due to the
various porous metals in the literature [11–21,23–27]. rapid expansion of the generated gas, the temperature was
Table 1
Density and yield stress of various porous metals
Material Structure Density q (g cm3) Relative density q/qs Yield stress ry (MPa) Relative yield strength ry/rys Relative density dependence n Reference
Al–7Si–0.3Mg (SiC) Closed-cell 0.081–0.38 0.03–0.14 0.063–2.9 1.6 · 104–7.5 · 103 2.5 [11]
Al–6Si–4Cu Closed-cell 0.2–0.8 0.07–0.29 5.9–23.7 4.2 · 102–1.7 · 101 1.89 [12]

M. Hakamada et al. / Acta Materialia 55 (2007) 2291–2299


Zn–4Cu Closed-cell 1.0–2.0 0.14–0.27 5.8–21.2 2.8 · 102–1.0 · 101 2.01 [12]
Al Open-cell 0.17–0.37 0.06–0.14 0.4–2.1 6.0 · 103–4.2 · 102 2.4 [13]
Al–7Mg Open-cell 0.14–0.56 0.05–0.21 0.7–13.2 3.1 · 103–5.7 · 102 2.1 [13]
Al 7075 Open-cell 0.26–0.56 0.1–0.21 1.5–13.5 4.3 · 103–3.9 · 102 2.6 [13]
Zn Open-cell 0.34–0.55 0.05–0.08 0.2–1.6 1.0 · 103–8.0 · 103 2.8 [14]
Al (SiC) Closed-cell 0.27–0.54 0.1–0.2 0.64–2.57 2.4 · 103–6.6 · 103 1.5 [15]
Al–Si Closed-cell 0.81–1.03 0.3–0.38 16.3–29 6.5 · 102–1.2 · 101 2.5 [16]
Al–4Cu Closed-cell 0.25–0.84 0.09–0.31 1.7–18.5 6.8 · 103–7.4 · 102 2.2 [17]
Al Closed-cell 0.27–0.625 0.1–0.25 1.7–11.4 3.4 · 102–2.3 · 101 1.9 [18]
Al Open-cell 2.14–4.35 0.19–0.27 0.79–1.6 1.6 · 102–3.2 · 102 1.63 [19]
Al Closed-cell 0.49–0.82 0.19–0.31 2.2–6.9 1.3 · 102–4.1 · 102 2.5 [20]
Al–7Si–0.45Mg Closed-cell 0.21–2.67 0.08–1.0 1.57–226 6.9 · 102–1.0 1.3 [21]
Ni–Cr Open-cell 0.98–7.68 0.022–0.032 0.11–0.91 1.2 · 102–3.6 · 103 2.5 [22]
Al (spherical pore) Open-cell 0.2–0.7 0.075–0.26 0.11–1.6 2.1 · 103–3.2 ·102 2.3 [23]
Al (angular pore) Open-cell 0.28–0.66 0.1–0.24 0.17–1.17 3.3 · 103–2.3 ·102 2.3 [23]
Cu (5 lm) Closed-cell 3.9–8.4 0.43–0.94 30–129 2.0 · 101–7.2 ·101 1.8 [24]
Cu (45 lm) Closed-cell 3.8–5.9 0.43–0.65 28–45 1.9 · 101–3.0 · 101 1.0 [24]
Fe Closed-cell 5.2–6.9 0.67–0.89 43–168 2.2 · 101–8.4 · 101 4.9 [25]
Cu–Sn Closed-cell 5.4–7.3 0.61–0.83 53–129 2.7 · 101–6.5 · 101 2.9 [25]
Fe Closed-cell 5.4–7.9 0.69–1.0 180–450 4.0 · 101–1.0 2.4 [26]
Be Closed-cell 1.36–1.76 0.74–0.95 87–514 1.5 · 101–8.6 · 101 6.3 [27]
Steel (without annealing)a Closed-cell 4.2–5.0 0.53–0.64 1.8 [28]
Steel (with annealing)a Closed-cell 3.3–4.7 0.46–0.60 2.5 [28]
a
The data for yield stress and relative yield strength are not included here because the scaling of stress–strain curves is unusual in this reference.

2293
2294 M. Hakamada et al. / Acta Materialia 55 (2007) 2291–2299

raised very slowly to 473 K, taking several hours. After the 5 mm min1 for all tests. A 0.2% proof stress was adopted
slow heating, the temperature was kept at 1173 K for 2 h, as the yield stress of the porous copper. For the measure-
followed by furnace cooling. As a result, cylindrical copper ment of Young’s modulus, the specimens were unloaded
samples with relative densities of 0.22–0.96 were fabricated. just after yielding and then loaded again to avoid the effect
The maximum density was achieved when no spacer parti- of specimen face misalignment. The modulus was then cal-
cles were used, while the minimum density was achieved culated according to the nominal compressive stress–strain
when the mixing volume ratio of Cu powder to spacer par- curves obtained after the second loading [16,54].
ticle was 0.15. The dimensions of the specimens were
8.5 mm diameter and 10 mm high. The microstructure 2.3. Microstructural observation
of the specimens with relative densities of 0.23 and 0.70
observed with an optical microscope is shown in Fig. 2. For microstructural observation, the specimens with the
The fabricated specimens had a closed-cell structure. The relative densities of 0.22 and 0.78 were fabricated via the
attained pore size was 240 lm, a little smaller than the spacer method. The specimens had plane outer surfaces.
sizes of spacer particles, probably owing to specimen Both specimens were observed before deformation and at
shrinkage during heat treatment. Nevertheless, the pore a nominal compressive strain of 0.2 to investigate their
sizes and shapes were uniform throughout the specimen, compressive deformation behavior.
owing to proper sieving of the spacer particles. Few struc-
tural defects such as missing cell walls, significant cell wall 3. Results
curvature and coupled cells were found in the fabricated
specimens. 3.1. Stress–strain curve

2.2. Compression test Fig. 3 shows the nominal stress–strain curves of copper
specimens with various relative densities under compres-
Compression tests on the copper specimens with relative sion. The unloading and the second loading curves for
densities of 0.22, 0.27, 0.34, 0.40, 0.44, 0.46, 0.50, 0.55, Young’s modulus measurement are omitted in Fig. 3 for
0.59, 0.67, 0.76, 0.83, 0.89 and 0.96 were carried out at clarity. The compression tests revealed that an increase in
room temperature using a universal testing machine (AG-
50kNG, Shimadzu Corp.). The crosshead velocity was 80
0.46 0.44

0.50 0.40 0.34


Nominal stress, σ / MPa

60
0.27

40

0.22
20

0
0 0.1 0.2 0.3 0.4 0.5 0.6
Nominal strain, ε

0.96 0.89 0.83 0.76


300

0.67
Nominal stress, σ / MPa

200
0.59

0.55
100

0
0 0.1 0.2 0.3 0.4
Nominal strain, ε

Fig. 3. Compressive stress–strain curves for copper specimens with


Fig. 2. Optical microscopic images of porous copper fabricated via spacer various relative densities of (a) 0.22–0.50 and (b) 0.55–0.96. The values
method: (a) relative density of 0.23; (b) relative density of 0.70. attached to curves indicate relative densities.
M. Hakamada et al. / Acta Materialia 55 (2007) 2291–2299 2295

the relative density of specimens causes an increase in flow pores act as defects or sources of stress concentration in
stress. The curves of specimens with lower relative densities the specimens with relative densities close to 1. As a result,
showed three distinct deformation regions, that is, the elas- it is noted that the deformation can be divided into three
tic region at the initial stage of the deformation, the plateau regions: Region I in the lower relative density range,
region where the specimens deform under a nearly constant Region II in the higher relative density range and Region
flow stress (plateau stress), and the densification region III in the highest relative density range close to 1, as shown
where the flow stress increases steeply at higher strain. in Fig. 4.
The densification strain, at which the steep increase in flow
stress initiates, was larger for the specimens with lower den- 3.3. Young’s modulus
sity [28]. These characteristics are the same as reported so
far [1]. However, the curves of specimens with higher rela- Fig. 5 shows the relationship between Young’s modulus
tive densities did not show these clear distinctions; in other and the relative density for porous copper. The relationship
words, the plateau and the densification regions seem to between the Young’s modulus (E) and the relative density
have emerged simultaneously after the elastic region. (q/qs) of porous materials is given by [1]
 n0
3.2. Compressive yield stress q
E ¼ C0 ð2Þ
qs
Fig. 4 shows the relationship between the compressive where C 0 and n 0 are constants. The slope of the plots, which
yield stress and the relative density for porous copper fab- corresponds to n 0 in Eq. (2), was 2.0 at relative densities
ricated via the spacer method. In Fig. 4, the plateau stress <0.5, while the slope was 1.2 at relative densities > 0.5.
was not substituted for the yield stress ry in Eq. (1), The Young’s modulus of the specimen with a relative den-
because the curves of specimens with higher relative densi- sity of 1 was much larger than the others. It should be
ties did not have clear plateau regions, and the plateau noted that the deformation can be divided into three re-
stress was ambiguous. The axes being scaled logarithmi- gions from the viewpoint of not only the yield stress, but
cally, the slope of the plots in Fig. 4 is the density exponent also Young’s modulus.
n in Eq. (1). The value of n at relative densities <0.5–0.6
was 2.3. Similar n values of 2–3 have been reported in 4. Discussion
many works [11,12,16–18,20,22,23]. However, n was 1.3
at higher relative densities. Also, the specimens with the 4.1. Deformation mode of cell wall
relative densities of 0.89, 0.96 and 1.0 showed an acute
drop in compressive yield stress with decreasing density. The results of the compression tests showed that the
The specimens with relative densities of 0.89 and 0.96 are deformation can be divided into three regions, depending
likely to be regarded as bulk materials containing a small on the relative density. The specimens in Region III, which
number of pores. Thus, it is suggested that the minimal is a relative density range close to 1, can be regarded as

Region III Region III

Region I Region II 100 Region I Region II

50
Compressive yield stress, σy / MPa

100
Young's modulus, E / GPa

1
50 1.2

1.3
10 1
1
2.0
10
5
2.3
5
1

1 1
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Relative density, ρ/ρs Relative density, ρ /ρs

Fig. 4. Relationship between relative density and yield strength of copper Fig. 5. Relationship between relative density and Young’s modulus of
specimens fabricated via spacer method. copper specimens fabricated via spacer method.
2296 M. Hakamada et al. / Acta Materialia 55 (2007) 2291–2299

bulk materials including a small amount of pores, as


explained in the previous section. In this section, the differ-
ences in the deformation mode of individual cell walls
between Regions I and II are discussed.
Figs. 6 and 7 show the observed cell wall deformation in
the specimen with a relative density of 0.22 (in Region I),
and Fig. 8 shows the observed cell wall deformation in
the specimen with the relative density of 0.70 (in Region
II). In Figs. 6–8, the cell walls of the specimens before
deformation are shown in (a), and the same portions com-
pressed at a nominal strain of 0.20 are shown in (b). The
loading direction is vertical in Figs. 6–8. In the specimen
in Region I, bending of cell walls causing plastic hinges
at the nodes (Fig. 6) and buckling of cell walls (Fig. 7) were
observed. Bastawros et al. [30] showed three processes in
the deformation of a porous metal: localized straining at
cell nodes, discrete bands of concentrated strain and,
finally, complete plastic collapse, and they demonstrated
with X-ray computed tomography analysis that buckling
and bending occur in each cell, resulting in localized defor-
mation. The deformation in Region I in Figs. 6 and 7 is in
agreement with their results. However, as shown in Fig. 8,
few bent or buckled cell walls were found, and most of the
cell walls yielded and became thick in the specimen in
Region II.
The exponent n can be theoretically obtained by analyz-
Fig. 7. Cell wall buckling in porous copper with relative density 0.22.
ing the deformation modes of a cubic closed-cell model
[1,10]. When the cell walls initially deform by bending,
buckling or yielding, the theoretical value of n is 2, 3 or

Fig. 6. Cell wall bending in porous copper with relative density 0.22. Fig. 8. Cell wall yielding in porous copper with relative density 0.70.
M. Hakamada et al. / Acta Materialia 55 (2007) 2291–2299 2297

1, respectively (see Appendix). As shown in Fig. 4, the


experimental value of n in Region I was 2.3, which is
between 2 for the bending mode of cell walls and 3 for
the buckling mode. Thus, it is likely that cell walls of com-
pressed specimens in Region I predominantly deform by
bending and buckling. The expectation of cell wall defor-
mation mode from the n value of 2.3 is in agreement with
the observation shown in Figs. 6 and 7.
However, the experimental value of n in Region II was
1.3, as shown in Fig. 4. This value is a little larger than
1, which is the theoretical value for the yielding mode of
cell walls. The approximation given by
q t
/ ð3Þ
qs l
must hold for the validity of Eq. (1), where t is the cell wall
thickness, and l is the cell size [1]; however, because the
relationship between q/qs and t/l can be more properly rep-
resented as
q t  t 2  t 3
¼3 3 þ ð4Þ
qs l l l
the approximation in Eq. (3) does not hold when t/l (and
hence q/qs) is large. In Region II, q/qs is too large to fulfill
the approximation represented by Eq. (3), indicating that Fig. 9. Surface of compressed porous copper (e = 0.20) with relative
the theoretical figure in Region II deviates from 1 to some density 0.22. Deformation bands are indicated by surrounding dashed
lines.
extent. Consequently, it is suggested that cell walls of com-
pressed specimens in Region II deform predominantly by
yielding.
Therefore, it is conclusively demonstrated from the
results in Fig. 4 and the observations in Figs. 6–8 that
low-relative-density (high-porosity) porous metals in
Region I deform by the bending and buckling of their cell
walls and that high-relative-density (low-porosity) porous
metals in Region II deform by the yielding of their cell
walls.

4.2. Deformation band

The emergence of the plateau region in the compressive


stress–strain curve for a porous material is closely related
to the formation of deformation bands [11,30–32]. Hence,
the simultaneous occurrence of the plateau and densifica-
tion regions in the stress–strain curves for the high-rela-
tive-density specimens, as shown in Fig. 3, is probably
because the specimens with higher relative densities tend
to deform uniformly, and clear deformation bands do not
form inside the specimens. Therefore, the compressive
behaviors of the specimens in Regions I and II may differ
not only in the deformation mode of individual cell walls,
Fig. 10. Surface of compressed porous copper (e = 0.20) with relative
but also in the macroscopic deformation behavior, namely, density 0.70.
the presence or absence of deformation band formation.
Fig. 9 shows the surface of the deformed porous copper
specimen with the relative density of 0.22 (in Region I), and presence of deformation bands in the porous copper spec-
Fig. 10 shows that with the relative density of 0.70 (in imen with the relative density of 0.22, as indicated by white
Region II), where both specimens were compressed at a dashed lines in Fig. 9. This is in agreement with the results
nominal strain of 0.20. Minute observation revealed the in the previous works [11,30–32]. In the specimen with a
2298 M. Hakamada et al. / Acta Materialia 55 (2007) 2291–2299

relative density of 0.70, however, clear deformation bands Acknowledgement


were not found, and local deformation was observed par-
tially around the pores. Zhang and Wang [24] investigated The authors thank Dr. Y. Yamada (National Institute
the deformation of porous copper with medium to high of Advanced and Industrial Science and Technology, Ja-
densities, and they showed that spherical pores changed pan) for advice on the spacer method.
to take on an ellipsoidal shape, but that no local deforma-
tion band was found. It is therefore likely that the deforma- Appendix
tion is relatively homogeneous in the specimen with a high
relative density, resulting in no clear deformation bands. According to the analysis by Gibson and Ashby [1], the
The specimens in Region I and those in Region II dif- cell deformation in a closed-cell porous material consists of
fered from the viewpoints of both the microscopic (defor- the stretching of cell faces (membranes) and the bending of
mation mode of cell walls) and macroscopic (presence or cell edges, rather than merely the bending of cell walls with
absence of local deformation bands) deformation behav- uniform thickness. The former mechanism of cell deforma-
iors. Therefore, the microscopic and macroscopic deforma- tion holds for a porous metal produced by a liquid-state
tion behaviors may have strong relationships. It is thus process, because the cell face is much thinner than the cell
suggested that the bending and buckling of cell walls cause edge, owing to drainage during the liquid-state process [55].
intense local deformation, compared with the yielding of However, in the present work, because porous copper was
cell walls, resulting in the enhancement of the formation fabricated by a solid-state process, the thickness of the cell
of deformation bands. wall was more uniform than that fabricated by a liquid-
state process. Hence, it is reasonable that the cell deforma-
5. Conclusions tion for the porous copper in the present work does not
contain the stretching of cell faces, but consists of bending,
Porous copper specimens with various relative densities buckling and/or yielding of a cell wall.
of 0.22–0.96 were fabricated via the spacer method, and The theoretical values of n in Eq. (1) for cell wall bend-
their compressive properties were examined. The conclu- ing and cell wall buckling are given in the study by Gibson
sions are as follows. [10], where deformation mechanisms for bending and
buckling are denoted as ‘plastic yielding’ (at the nodes)
(1) An increase in the relative density of the porous cop- and ‘elastic buckling’, respectively, in Table 1 in the litera-
per caused an increase in flow stress and a decrease in ture [10]. According to the analysis, the value of n for cell
densification strain under compression. wall bending is 2 and that for cell wall buckling is 3. The
(2) The yield strength and Young’s modulus of the por- value of n for cell wall yielding is derived from the follow-
ous copper depended on its relative density in three ing discussion.
ways. In the lower relative density range (Region I, Here, the cubic closed-cell model with cell wall thickness
q/qs < 0.5–0.6), the relative density dependence n of t and cell size l, which is found in Fig. 6 in Ref. [10], is
yield stress was 2.3, while the value was 1 in the assumed. If the cell wall is deformed by yielding where
higher relative density range (Region II, 0.5– the cell wall thickens with strain, the critical applied load
0.6 < q/qs < 0.9). In the highest relative density range for plastic deformation (Fyield) is proportional to ltrys,
(Region III, q/qs is very close to 1), the relative den- because the cross-sectional area of the cell wall is lt.
sity strongly affected the yield strength and Young’s Therefore, considering the proportional relationship
modulus, probably owing to stress concentration between t/l and q/qs (Eq. (3)), the critical stress for cell wall
around the minimal pores. yielding (ryield) is given by
(3) Microstructural observation revealed that in the
F yield t q
lower relative density range (Region I), the cell walls ryield / / rys / rys ðA:1Þ
dominantly deform by bending and buckling, but the l2 l qs
dominant cell wall deformation mode in the higher Therefore, the value of n in Eq. (A.1) is 1 for cell wall
relative density range (Region II) is yielding. These yielding.
findings were in agreement with the results of the den-
sity exponent obtained by compressive tests.
References
(4) Deformation bands were observed in the compressed
specimen in the lower relative density range (Region [1] Gibson LJ, Ashby MF. Cellular solids: structure and proper-
I), but not in the compressed specimen in the higher ties. Cambridge: Cambridge University Press; 1997.
relative density range (Region II). Thus, the speci- [2] Evans AG, Hutchinson JW, Ashby MF. Prog Mater Sci 1999;43:171.
mens in Region I and those in Region II differed from [3] Banhart J. Prog Mater Sci 2001;46:559.
[4] Despois JF, Mortensen A. Acta Mater 2005;53:1381.
the viewpoints of both the microscopic (deformation
[5] Lu TJ, Hess A, Ashby MF. J Appl Phys 1999;85:7528.
mode of cell walls) and macroscopic (presence or [6] Lu TJ, Chen F, He D. J Acoust Soc Am 2000;108:1697.
absence of local deformation bands) deformation [7] Hakamada M, Kuromura T, Chen Y, Kusuda H, Mabuchi M. Appl
behavior. Phys Lett 2006;88:254106.
M. Hakamada et al. / Acta Materialia 55 (2007) 2291–2299 2299

[8] Wen CE, Yamada Y, Shimojima K, Chino Y, Hosokawa H, Mabuchi [31] Bart-Smith H, Bastawros AF, Mumm DR, Evans AG, Sypeck DJ,
M. J Mater Res 2002;17:2633. Wadley HNG. Acta Mater 1998;46:3583.
[9] Takemoto M, Fujibayashi S, Neo M, Suzuki J, Kokubo T, [32] Markaki AE, Clyne TW. Acta Mater 2001;49:1677.
Nakamura T. Biomaterials 2005;26:6014. [33] Miyoshi T, Itoh M, Mukai T, Kanahashi H, Kohzu H, Tanabe S,
[10] Gibson LJ. J Biomech 1985;18:317. et al. Scr Mater 1999;41:1055.
[11] Simone AE, Gibson LJ. Acta Mater 1998;46:3109. [34] Nieh TG, Higashi K, Wadsworth J. Mater Sci Eng A 2000;283:105.
[12] Banhart J, Baumeister J. J Mater Sci 1998;33:1431. [35] Nieh TG, Kinney JH, Wadsworth J, Ladd AJC. Scripta Mater
[13] Thornton PH, Magee CL. Metall Trans A 1975;6:1253. 1998;38:1487.
[14] Thornton PH, Magee CL. Metall Trans A 1975;6:1801. [36] Yamada Y, Shimojima K, Mabuchi M, Nakamura M, Asahina T,
[15] Prakash O, Sang H, Embury JD. Mater Sci Eng A 1995;199:195. Mukai T, et al. Philos Mag Lett 2000;80:215.
[16] Sugimura Y, Meyer J, He HY, Bart-Smith H, Grenstedt J, Evans AG. [37] Tantikom K, Aizawa T, Mukai T. Int J Solids Struct 2005;42:2199.
Acta Mater 1997;45:5245. [38] Jeon I, Asahina T. Acta Mater 2005;53:3415.
[17] Kunze HD, Baumeister J, Banhart J, Weber M. Powder Metall Int [39] Zhu HX, Windle AH. Acta Mater 2002;50:1041.
1993;25:182. [40] Gradinger R, Rammerstorfer FG. Acta Mater 1999;47:143.
[18] Wen CE, Yamada Y, Asahina T, Kato K, Sonoda T, Watazu A, [41] Chen C, Lu TJ, Fleck NA. J Mech Phys Solids 1999;47:2235.
et al. Mater Trans 2004;45:327. [42] Simone AE, Gibson LJ. Acta Mater 1998;46:3929.
[19] San Marchi C, Mortensen A. Acta Mater 2001;49:3959. [43] Silva MJ, Gibson LJ. Int J Mech Sci 1997;39:549.
[20] Kennedy AR. J Mater Sci 2004;39:3085. [44] Han F, Cheng H, Wang J, Wang Q. Scripta Mater 2004;50:13.
[21] Yang CC, Nakae H. ISIJ Int 2000;40:1283. [45] Polonsky L, Lipson S, Markus H. Modern Cast 1961;39:57.
[22] Choe H, Dunand DC. Mater Sci Eng A 2004;384:184. [46] Zhao YY, Sun DX. Scripta Mater 2001;44:105.
[23] Goodall R, Marmottant A, Despois JF, Salvo L, Mortensen A. [47] Wen CE, Mabuchi M, Yamada Y, Shimojima K, Chino Y, Asahina
Replicated microcellular aluminum with spherical pores. In: Nakajima T. Scripta Mater 2001;45:1147.
H, Kanetake N, editors. JIMIC-4 porous metals and metal foaming [48] Chou KS, Song MA. Scripta Mater 2002;46:379.
technology. Sendai: The Japan Institute of Metals; 2005. p. 207–10. [49] Wen CE, Mabuchi M, Yamada Y, Shimojima K, Chino Y,
[24] Zhang E, Wang B. Int J Mech Sci 2005;47:744. Hosokawa H, et al. J Mater Sci Lett 2003;22:1407.
[25] Wang B, Klepaczko JR, Lu G, Kong LX. J Mater Process Technol [50] Zhao YY, Fung T, Zhang LP, Zhang FL. Scripta Mater 2005;52:
2001;113:574. 295.
[26] da Silva MG, Ramesh KT. Int J Plast 1997;13:587. [51] Brothers AH, Scheunemann R, DeFouw JD, Dunand DC. Scripta
[27] Beasley D, Turner GI, Edwards KL. J Mater Sci 1975;10:436. Mater 2005;52:335.
[28] Park C, Nutt SR. Mater Sci Eng A 2000;288:111. [52] Hakamada M, Yamada Y, Nomura T, Chen Y, Kusuda H, Mabuchi
[29] Yamada Y, Wen CE, Shimojima K, Hosokawa H, Chino Y, Mabuchi M. Mater Trans 2005;46:2624.
M. Mater Trans 2002;43:1298. [53] Bakan HI. Scripta Mater 2006;55:203.
[30] Bastawros AF, Bart-Smith H, Evans AG. J Mech Phys Solids [54] McCullough KYG, Fleck NA, Ashby MF. Acta Mater 1999;47:2323.
2000;48:301. [55] Simone AE, Gibson LJ. Acta Mater 1998;46:2139.

You might also like