You are on page 1of 14

Journal of Materials Science & Technology 62 (2021) 11–24

Contents lists available at ScienceDirect

Journal of Materials Science & Technology


journal homepage: www.jmst.org

Invited Review

Fabrication, properties, and applications of open-cell aluminum


foams: A review
Tan Wan a,b , Yuan Liu a,b,∗ , Canxu Zhou a,b , Xiang Chen a,b , Yanxiang Li a,b
a
School of Materials Science and Engineering, Tsinghua University, Beijing 100084, China
b
Key Laboratory for Advanced Materials Processing Technology (Ministry of Education), Beijing 100084, China

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

Article history: Open-cell metallic foams or porous metals have a distinctive combination of excellent structural per-
Received 14 March 2020 formance and superior functional characteristics, such as their light weight, energy absorption, sound
Received in revised form 13 May 2020 absorption, heat dissipation, and electromagnetic shielding. As a primary representative of metallic
Accepted 20 May 2020
foams, aluminum foam has developed into a new engineering material with many unique applications
Available online 9 July 2020
in the fields of aerospace, automotive industry, petrochemical industry, building materials, and etc. This
paper summarizes the fabrication methods, properties, and applications of open-cell aluminum foams.
Keywords:
The current status and development trends are also introduced.
Open-cell aluminum foam
Fabrication method © 2020 Published by Elsevier Ltd on behalf of The editorial office of Journal of Materials Science &
Property Technology.
Application

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.1. Fabrication methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.2. Investment casting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.3. Metallic deposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.4. Infiltration process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.5. Powder-metallurgy method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.6. Additive manufacturing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.6.1. Direct printing route . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.6.2. Indirect printing route . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2. Structure control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.1. Effect of pattern design on cell structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2. Effect of processing parameters on cell structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3. Properties and applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
4. Outlooks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
5. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

∗ Corresponding author at: School of Materials Science and Engineering, Tsinghua University, Beijing 100084, China.
E-mail address: yuanliu@tsinghua.edu.cn (Y. Liu).

https://doi.org/10.1016/j.jmst.2020.05.039
1005-0302/© 2020 Published by Elsevier Ltd on behalf of The editorial office of Journal of Materials Science & Technology.
12 T. Wan et al. / Journal of Materials Science & Technology 62 (2021) 11–24

1. Introduction mer sponges [46] and space holders [47], which are applied in
investment casting, infiltration of space holders, and powder metal-
Many porous structures exist in nature, including honeycomb lurgy to prepare open-cell aluminum foams. Rapid developments
structures, such as wood and cork, and foam-like structures, such of additive manufacturing have allowed patterns with open cells
as bone, plant parenchyma, and sponge. These structures com- or even open-cell aluminum foams to be fabricated directly by
bine their light weight with excellent mechanical properties, which three-dimensional (3D) printing, which provides new ideas and
meet natural loading requirements [1]. Inspired by these naturally directions to update the technology to produce open-cell aluminum
formed porous structures, human-made porous metals were pre- foams [48–50]. The unique advantages of the replication process
pared to extend the applications of metals [2]. In general, porous and 3D printing in pattern design can achieve a stable design of the
metals can be categorized into two groups. The first group is called required structures and performance of the final foams, which is
metal or metallic foams, which is commonly characterized by important for aluminum foam manufacture.
closed-cell metal foam, and the other is called porous metals, which We summarize the preparation methods for open-cell alu-
is mostly characterized by open-cell structure. Aluminum [3], mag- minum foams with the pore size at millimeter level, which are
nesium [4], copper [5], titanium [6], iron [7], and other metals and termed investment casting, space holder infiltration, powder met-
their alloys have been formed into porous structures. Aluminum is allurgy, and additive manufacturing. The effects of pattern design
the most widely used non-ferrous metal in engineering because of and processing parameters on the foam cell structure are described.
many advantages such as its rich reserves, low density, good ductil- Some properties, applications, and development tendencies of
ity, and high corrosion resistance [8]. Therefore, porous aluminum open-cell aluminum foams are presented.
has become one of the most popular and widely studied porous
metals. Gas pores in aluminum matrices are functional phases that 1.1. Fabrication methods
provide porous aluminum with the special performance of energy
absorption [9], vibration damping [10], sound absorption [11], elec- In the replication process, the first step is pattern preparation,
tromagnetic shielding [12], and heat insulation or dissipation [13]. which is decisive in the final foam cell structures. Among vari-
Therefore, porous aluminum has the potential to be applied in ous types of patterns, polymer sponges and space holders are used
aerospace, automobiles, the petrochemical industry, and building most commonly. Based on the patterns, three common approaches
materials [14]. were developed to fabricate open-cell aluminum foams, namely
Porous aluminum can be divided into open- and closed-cell investment casting, space holder infiltration, and powder metal-
structures depending on the manufacturing process. Closed-cell lurgy. In addition, 3D printed patterns from resin and sand molds
porous aluminum (often also called closed-cell aluminum foam) are applied commonly during replication. We will introduce the
is used mainly as structural materials to bear loads and absorb above methods to prepare open-cell aluminum foams based on a
energy because of their excellent mechanical properties. Open-cell detailed replication process and additive manufacturing.
porous aluminum (also commonly known as open-cell aluminum
foam) can act as structural materials and serve as functional mate- 1.2. Investment casting
rials in the fields of high-temperature filtration, sound absorption,
and heat dissipation because of their permeability or open porosity Open-cell aluminum foams can be made by copying struc-
[14,15]. The abovementioned structural and functional character- tures of open porous polymer sponges, which is commonly termed
istics are directly related to the cell structures of porous aluminum, investment casting, as shown in Fig. 1. First, a polymer sponge with
such as their morphology, proportion, size, and cell distribution. To open cells is filled with a refractory slurry, such as a mixture of mul-
make the best use of the performance advantages of porous alu- lite, phenolic resin, and calcium carbonate [51] or a gypsum-type
minum, reliable processes must be developed to provide porous slurry [52]. After the slurry has been cured, the polymer sponge
aluminum with designable structures and performance [16]. is removed at a high temperature to obtain a structurally sta-
The main preparation methods of closed-cell aluminum foams ble preform with open voids, and then molten aluminum is cast
are the melt foaming method [17], gas injection method [18] into the interconnecting channels in the preform under pressure.
and powder compact melting technique [19]. In these methods, After cooling, the preform is removed by water scrubbing to form
cell generation is closely related to gas bubble production and an aluminum foam whose structure is like the polymer sponge.
control. In comparison, the manufacture of open-cell aluminum Hence, foam structures depend almost entirely on the initial poly-
foams has different principles, such as the most frequently used mer sponge when sufficient infiltration takes place, which means
replication process [20–23], which can prepare aluminum foams that porosity and pore size of the final foam is approximately equal
with structures that are determined by predesigned patterns, with to porosity and pore size of the applied polymer sponge. Pro-
advantages of a simple process, low manufacturing costs, and a high cess parameters like pattern and casting temperature can affect
degree of structure control. Pattern design and fabrication, metal the microstructure of the aluminum matrix, thereby influence the
filling, and pattern removal are three main steps in the replica- properties of cast aluminum foams. A superior pore homogeneity,
tion process [24], where pattern removal determines that cells in distribution, and high degree of pore opening are the outstand-
the final foams are usually open. The advantages of the replication ing advantages of this method, and a porosity above 90 % can be
process show promising applications of open-cell aluminum foams achieved easily. This method has some disadvantages in terms of
that are fabricated via the replication process. the complicated production steps and a relatively high cost. It is
Studies on open-cell aluminum foams can be traced to 1961 difficult to prepare aluminum foams with small pores and large
when Polonsky et al. first presented the applicability of aluminum dimensions by this method because of the difficult slurry grouting
foam replication using sintering coarse rock salt [25]. In 1966, and molten aluminum filling.
Kuchek [26] applied for the first patent on this method. Over
the past 60 years, extensive research has been carried out to 1.3. Metallic deposition
optimize cell structures [27,28], mechanical properties [29–33],
sound-absorption performance [34,35], electromagnetic shielding Another common method to fabricate open-cell metal foam
[36–38], and heat dissipation [13,39]. Open-cell aluminum foam via the employment of a polymer sponge is metallic deposition,
structures can be controlled by processing parameters [40,41] and which is not suitable for aluminum but has great potential in
pattern design [42–45]. Two pattern types exist, namely poly- preparing nickel foam and copper foam. The surface of a poly-
T. Wan et al. / Journal of Materials Science & Technology 62 (2021) 11–24 13

Fig. 1. (a) Typical polymer sponge, (b) Schematic of investment casting to produce open-cell aluminum foams [14].

Fig. 3 shows typical CaCl2 , NaCl granules and a schematic of


the infiltration process. At certain pressures, molten aluminum
fills the interstices between the space-holder particles with high-
temperature stability to produce a composite of aluminum and
space-holder particles. Then, an aluminum foam with open cells
can be obtained by immersing the composite into the water to
remove the space-holder particles. The most common space holder
that is used in the infiltration process is NaCl [57], which possesses
various advantages, such as a low cost, nontoxicity, easy dissolu-
tion in water, and a higher melting temperature than aluminum.
Shape and size of space-holder particles fundamentally decide cell
shape and size in the final foam, and contact points of these particles
determine cell interconnections of the resulting foams. Besides, the
foam porosity can be varied by adjusting the pattern density relying
Fig. 2. Schematic of electro-deposition technique to produce open-cell foams [14].
on the stacking density of these particles. Compared with invest-
ment casting, infiltration shows remarkable advantages, such as
mer sponge is directly coated with a thin metal layer under a low short processing time and a low cost. Cell structures can be con-
temperature through physical vapor deposition (PVD) or chemical trolled precisely by the size, shape, and volume fractions of the
vapor deposition (CVD), such as magnetron sputtering technique or space holders. However, the porosity is limited by the lower stack-
multi-arc ion plating technique [53]. Either the polymer sponge is ing density (< 65 %) of the space holders. The almost point-contact
pre-coated an electrically conductive substance by electroless plat- state between the space-holder particles may limit the dissolution
ing or cathode sputtering for the following plating of a thick metal removal of the space-holder particles. Therefore, it is difficult to
layer, mainly utilizing electrolytic deposition [54]. After completely obtain a larger void volume with as high as an 80 %–90 % poros-
decomposing the polymer by heating or chemical treatment, the ity like closed-cell aluminum foam. Some space-holder particles
polymer’s structure is also reproduced. Fig. 2 schematically illus- that remained in the aluminum matrix lead to lower porosity and
trates the method. Unlike the investment casting, the struts are lower opening degree [14]. The remaining space holders may cause
hollow. Similar to investment casting, the final foam structures corrosion problems during service [16].
depend partly on the initial polymer sponge; however, deposi-
tion conditions such as depositing time and temperature have a
1.5. Powder-metallurgy method
great effect on foam structures. Although the deposition method is
widely applied in producing foams of nickel [55] and copper [56], it
Space holders applied in the infiltration process can be mixed
is not suitable for aluminum with more negative electrode potential
with powdered aluminum, compacted and sintered in a powder-
than hydrogen in electro-deposition technique, and electroplating
metallurgy method. In most cases, with the small addition of
liquid is liable to cause environmental contamination and explo-
sintering aids and binders, aluminum powders are mixed with
sion. In addition, during the vapor deposition method, aluminum
space holders in appropriate proportions to be compacted to green
coating is oxidized easily, so the final foam is very brittle and shows
specimens. The aluminum powders can be sintered together or
poor performance.
melt at a temperature below the particle melting temperature.
Finally, aluminum foams with open cells can be obtained by
1.4. Infiltration process the removal of the space holders in the green specimens. Fig. 4
illustrates the process, which is termed the sintering–dissolution
The infiltration process is one of the earliest and most applied process. The frequently used space holders in the process are inor-
methods for open-cell aluminum foam preparation. It differs from ganic salt with higher melting points than aluminum, such as NaCl
gas pores or cells that are evolved from gas bubbles in liquid or [58] and CaCl2 [33]. However, low-melting-point urea [59] and
semi-solid aluminum matrices in traditional closed-cell aluminum sucrose [60] with weak corrosion can also serve as space hold-
foam. In the infiltration process, cells of open-cell aluminum foams ers, in which case the aluminum powders can only be sintered
are developed from as-called space holders in the solid state, which after leaching the space holders, and this technique is referred
commonly are water-soluble inorganic salts with high melting to as the dissolution and sintering process (DSP). In a few rare
points, such as NaCl, MgSO4 , and CaCl2 . Sucrose, urea, and other cases, aluminum power slurries that contain binders are coated on
organic particles with low melting points have been developed struts of polymer sponges, which can produce open-cell aluminum
into novel space holders, which, with inorganic salts, are used foams by burning the sponges and by subsequent powder sinter-
frequently in powder metallurgy to prepare open-cell aluminum ing [61]. Similar to infiltration, pore size, pore shape, and aluminum
foams. foam porosity can be tailored by space holder size, shape, and vol-
14 T. Wan et al. / Journal of Materials Science & Technology 62 (2021) 11–24

Fig. 3. Typical CaCl2 granules (a), typical NaCl granules (b), schematic of infiltrating process (c) to produce open-cell aluminum foams [14].

Fig. 4. (a) Typical sucrose granules, (b) Typical urea granules, (c) Schematic of sintering–dissolution process for manufacturing open-cell aluminum foams [58].

ume fraction. In addition, process parameters such as compaction structures of stainless steel, copper, titanium, and their alloys can be
pressure, sintering temperature can strongly affect the macro- or achieved by direct selective laser sintering (SLS), selective electron
microstructure of the foams. Space holders can be removed com- beam melting (SEBM), and selective laser melting (SLM) [62–64].
pletely by heat treatment when applying organic particles with low Because of the high reflectivity of aluminum and high thermal
melting points, which reduces matrix corrosion. Additional pro- conductivity of aluminum, this direct printing route is often com-
tection may be required to promote sintering because aluminum plicated and expensive [65], and only a few reports exist on the
powders are susceptible to oxidation, which together with their manufacture of aluminum foams by SLM [66–68]. Computer-aided
high manufacturing cost, makes powder metallurgy complicated, design (CAD) models of unit cells and cellular lattice structures
expensive, and unsuitable for large-scale production. were designed, and the final 3D aluminum foam was constructed
by scanning the aluminum powders layer by layer according to the
1.6. Additive manufacturing CAD models. Fig. 5 shows the CAD model and the final aluminum
lattice structure. Foam structures depend mainly on the unit cells
Improvements in computer technology have led to rapid devel- and periodic structures; moreover, processing parameters such as
opments in additive manufacturing. Using this method, objects can laser power, scanning speed, and spot size will also affect foam
be constructed by layerwise printing of powders such as metal, structures. The manufacturing process of printing metals requires
plastic or ceramics, which provides a new route for open-cell alu- expensive equipment and is time- and cost-consuming. Based on
minum foam preparation. Two common routes exist in additive this, the process is often applied to special-demand applications
manufacturing, namely direct and indirect printing. The direct such as biomedical implants and aerospace parts, and large-scale
printing route can construct foams by printing aluminum pow- industrial applications remain in their infancy.
ders directly, whereas the indirect printing route prepares foams
by combining printing patterns with traditional casting. These two 1.6.2. Indirect printing route
printing routes will be described separately. Difficulties in preparing open-cell aluminum foams by direct
printing occur; therefore, it is more common to prepare open-cell
1.6.1. Direct printing route aluminum foams via indirect printing routes, as shown in Fig. 6. In
Metal powder beds can be scanned selectively by energy this technique, additive manufacturing is used to print a pattern
sources, such as laser or electron beam spots, to form dense metal for a traditional casting, such as investment casting or lost foam
workpieces without additional post-processing, which provides casting; a pattern and not the final aluminum foam is obtained by
advantages of preparing and tailoring complex geometries for spe- 3D printing, and the pattern is usually made from sand [69], wax
cific needs that cannot be manufactured by other methods. Cellular [70] or a photosensitive resin [49]. Clearly, the printed pattern will
T. Wan et al. / Journal of Materials Science & Technology 62 (2021) 11–24 15

Fig. 5. (a) and (b) CAD models of unit cell and lattice structure, (c) final periodic cellular lattice structure of aluminum foam [67].

Fig. 6. Schematic of additive manufacturing route to make open-cell aluminum foams [71].

decide the formed foam structures. The indirect 3D printing route These deficiencies will limit the further application of open-
combines the advantages of high precision control of cell structures cell aluminum foams. The significant feature of the replication
by additive manufacturing and the superior material quality that is process is precise structural control of the final aluminum foams
provided by traditional casting. Nevertheless, compared with many through pattern design. Therefore, pattern design and the control
other replication methods, it is still costly and complicated, which of processing parameters are usually used to make up for these
is almost limited to the production of small parts. shortcomings.

2.1. Effect of pattern design on cell structures


2. Structure control
Investment casting is an effective method to improve the
Properties of open-cell aluminum foams are affected by matrix strength of open-cell aluminum foams by thickening the polymer
materials, and by cell structures, which include mainly the pore sponge struts. Wang et al. [72,73] impregnated polymer sponges
morphology, pore size, pore distribution, and porosity. Invest- in an organic liquid, so that the struts were coated with a layer of
ment casting and infiltration processes have become the most organic curing material. GmbH (Dresden, Germany) applied wax to
extensive preparation methods of open-cell aluminum foams for the struts for the same reason [16]. Alexander et al. [74] coated the
a relatively stable technology and low cost, and have been used strut surface with a polymer layer by spraying polymer granules on
for commercial production. For example, ERG (Oakland, CA, USA) them, and the relative density of the final aluminum foam reached
and Exxentis (Wettingen, Switzerland) have produced commercial 25 % eventually. Fig. 7 shows the aluminum foams before and after
open-cell aluminum foams by investment casting and infiltration, strut thickening treatment. The struts of aluminum foams were
respectively [16]. However, in investment casting, the fabricated thickened. Alan et al. [75] obtained reinforced aluminum foams
open-cell aluminum foams have a low mechanical strength and with a continuous density gradient by compressing the polymer
bearing capacity because of the high polymer sponge porosity (usu- sponges.
ally more than 90 %) and the narrow strut width. The porosity For infiltration, a replacement of the corrosive inorganic salt par-
of the open-cell aluminum foams that are fabricated by infiltra- ticles with non-corrosive or weakly corrosive particles can resolve
tion is usually less than 80 % for limited volume fractions of space or alleviate the corrosion problem. The porosity and pore connec-
holders, which results in a low lightweight extent. Inorganic salt tivity of open-cell aluminum foams can be improved by increasing
particles that are commonly used as space holders are highly cor- the stacking density of space holders and enlarging the contact area
rosive, which corrodes the matrix during leaching, especially where between space holders, respectively. Kan et al. [42] produced com-
pore structure connectivity is poor so that salt leaching is difficult. pressible ceramic soft balls to replace hard NaCl particles by mixing
The use of space holders in powder metallurgy leads to similar Al2 O3 powders, polyvinyl alcohol, water, and a small amount of
problems for infiltration for a lower stacking density (< 65 %) of bentonite and cellulose, which contributes to a higher space holder
the space holders. packing density and hence, a higher foam porosity, as shown in
16 T. Wan et al. / Journal of Materials Science & Technology 62 (2021) 11–24

Fig. 7. Open-pore AlZn11 foams with a pore density of 10 ppi (a) for untreated conditions, (b) after thickening treatment [74].

Fig. 8. (a) Soft ceramic balls after closed packing, (b) final aluminum foam [42].

Fig. 9. SEM images of aluminum foams manufactured by infiltration using preforms.


of 75 ␮m NaCl particles, Vf represents the density of the replicated Al foam: (a) unsintered (Vf = 0.4), and sintered at 755 ◦ C for 2 h (Vf = 0.37) (b), 9 h (Vf = 0.32) (c) and 25 h
(Vf = 0.25) (d) [43].
T. Wan et al. / Journal of Materials Science & Technology 62 (2021) 11–24 17

Fig. 8. Goodall et al. [76] mixed salt, flour, and water to produce of aluminum melts in a porous pattern. The powder-metallurgy
a paste that could be shaped by hand, with salt as the main body. method involves the mixing, compacting, and metallurgy of
Thus, compressible spheres that were made from the paste could aluminum powders and space holders. Accordingly, processing
be used as space holders to be stacked into a precursor with a parameters vary in the different procedures in the casting and
larger contact area between the space holders and a higher pack- powder-metallurgy methods. For instance, processing parameters
ing density. In general, the sintering and pressing of space holders in casting include the infiltration pressure and melt and pattern
are two common ways to densify the pattern with a stable surface temperatures, whereas those in the powder-metallurgy method
bonding between the space holders [43,77–80]. Both can increase include the ratio of aluminum powders to space holders, sinter-
the contact areas between the space holders to a certain extent, ing pressure, temperature, and time. These parameters will affect
which increases the compactness of the pattern. Goodall et al. [43] the macro- or microstructure, and determine the performance of
densified NaCl preforms by sintering and cold isostatic pressing to the final open-cell aluminum foams.
produce open-cell aluminum foams with porosities in the range of Poor wetting of the pattern materials by molten aluminum does
64 % and 90 %. Fig. 9 shows the aluminum foams that were obtained not allow the melt to flow into the pattern interstices sponta-
from unsintered and sintered preforms of 75 ␮m NaCl particles neously. Therefore, a certain pressure-assisted infiltration process
under different conditions. Apichart et al. [78] pressed porous salt is usually required for the first ingress of the melt into the porous
particles with an average pore diameter of 1.4 mm and 2.0 mm pattern [86–88]. The pressure at which the metal melt can infiltrate
at 32 MPa, 39 MPa, and 50 MPa to be sintered at 730 ◦ C for 4 h a large fraction of the interstices is termed the ẗhreshold pressure¨.
to form patterns. Aluminum foams with porosities of ∼70 % and Hilden et al. [89] calculated this threshold pressure P by numerical
90 % could be obtained. Goodall et al. [76,77] pointed out that the simulation:
sintering of salt particles greater than 150 ␮m was dominated by 10LV
an evaporation–condensation mechanism, which can change the P= (1)
R
neck morphology but lead to a limited increase in pattern density,
even for long sintering times. For millimeter-sized salt particles, where LV is the surface tension of the metal liquid and R is the
the risk of crushing during mechanical pressing is high. Therefore, radius of the space holders. The magnitude of the surface tension
the above two methods, namely sintering and pressing, are unsuit- of an aluminum liquid is usually 1 J m−2 . Therefore, one atmosphere
able for the preparation of aluminum foams with a pore size greater will cause infiltration of metal melt into a preform of space holders
than 1 mm. To overcome the limitations of a large size, Erardo et al. that have a radius of 100 ␮m or more [26]. In addition, Cordovilla
[41] increased the porosity of the aluminum foam by ∼9% to 10 % et al. [90] provided an alternative empirical relationship based on
via pattern vibration to achieve a higher stacking density. Ma et al. experimental data:
[27] covered the space holders with a layer of resin to adjust the F
passage between two cells, and aluminum foams with porosities of P [MPa] = 16.15 − 0.09 (2)
(1 − F) D
73 %–85.8 % and pore sizes of 1.19–2.85 mm were prepared under
high-pressure infiltration. where F and D is the pattern density and a characteristic diameter
Many studies have mentioned that compared with metal foams of the solid on average. This relation indicates that the thresh-
with irregular cells, spherical cell metal foams have uniform cells old pressure depends on the size of the channels to be infiltrated
of regular shape and size, which makes their production and and the volume fraction of these channels. In the infiltration pro-
performance repeatable and reliable, and favors the design and pre- cess based on space holders, not all but a large fraction of the
diction of foam structure and properties [28,33,47,59,78,81–83]. open space can be infiltrated under the threshold pressure, and
The preparation of open-cell aluminum foams with spherical cells porosity of the final foam can be roughly predicted by the stacking
is inseparable from the selection or preparation of spherical space density of these space holders. Furthermore, insufficient infiltra-
holders. Years ago, it was difficult to obtain spherical NaCl par- tion may take place when a pressure lower than the threshold
ticles in the market because the NaCl particle shape was mostly pressure is applied, while a much high pressure can force the
irregular, and the melting point of easily obtainable urea with a melt into small channels between neighboring particles, which
spherical shape was much lower than that of aluminum. It was dif- will cause difficulty to remove the pattern. Thus, control of the
ficult to prepare open-cell aluminum foams with spherical cells infiltration pressure is crucial to produce the metal foam pos-
through infiltration. Nevertheless, Jiang et al. [28,81] prepared sessing a uniform distribution of pores without leaving many
spherical cell aluminum foams by the sintering–dissolving process pattern materials uneasy to remove in it. Metal foam fabricated by
using commercial 1-mm-diameter spherical urea. Goodall et al. investment casting based on a polymer sponge features equivalent
[82] produced spherical NaCl particles by melting, dispersing, and structures possessed by the original polymer foam at a pressure
solidifying irregular NaCl particles, which were used to fabricate equal to or higher than the threshold pressure. Hence, the poros-
open-cell aluminum foams via infiltration, as shown in Fig. 10. Jin- ity of the final foam is almost the same as the applied polymer
napat et al. [47,78,84] mixed NaCl, flour, and water with heating sponge. Actually, it is hard to predict the ultimate foam’s porosity
and stirring at a high speed in vegetable oil. The filtered particles precisely in the melt infiltration approaches. Nevertheless, accu-
were sintered into porous spheres, which were used to prepare rate measurement of the pore structure like porosity is easy and
open-cell aluminum foams with spherical cells through infiltration convenient.
and sintering–dissolution. Nowadays, spherical CaCl2 particles are Despois et al. [40] investigated the effect of infiltration pressure
easily obtained in the market and have been used to replace NaCl on the structure of aluminum foams in infiltration experimentally.
particles in the infiltration [85] and sintering–dissolving [6] pro- For preforms of 400-␮m space holders that were stacked to a 75
cesses to prepare open-cell metal foams because of their low cost % packing density, aluminum foams with a porosity close to 85 %
and easy availability. can be obtained at 0.1 MPa, as shown in Fig. 11. As the infiltration
pressure increased, the aluminum foam porosity decreased, and
2.2. Effect of processing parameters on cell structures finger-like protrusions appeared on the cell wall. These finger-like
protrusions resulted from a replication of the narrow interstices
The investment casting and infiltration processes to prepare of adjacent salts. Chen et al. [91] provided a calculation model to
open-cell aluminum foams can be summarized as a casting method predict the degree of pore opening of aluminum foams. The degree
according to the liquid state of aluminum, which involves the flow of pore opening increased with a decrease of pressure and particle
18 T. Wan et al. / Journal of Materials Science & Technology 62 (2021) 11–24

Fig. 10. SEM images of aluminum foams with spherical pores of mean diameter 400 ␮m with volume fractions solid, Vs: (a) 0.31, (b) 0.25 [82].

3. Properties and applications

Excellent integration of structural and functional properties


provides wide prospects for the application of open-cell aluminum
foams [97], whose performance is dependent on the matrix and cell
structure properties. Open-cell aluminum foams have an inherited
elasticity, corrosion or temperature resistance from the aluminum
matrix, and possess characteristics of a low density, a high spe-
cific surface area, and excellent energy absorbing properties from
the cells of the foams [98]. The most important mechanical and
functional properties are reviewed.
The mechanical properties of open-cell aluminum foams are
affected by the matrix, but the effects from the cell structures
have attracted the most attention. Because of the cell structure,
the mechanical properties rely mainly on the relative density, and
they depend on the cell roundness, cell shape, cell connectivity, and
diameter distribution. Unlike closed-cell aluminum foam, there is
no effect of air capture in open-cell aluminum foams because of the
interconnected cells. Thus, the mechanical properties of open-cell
aluminum foams are not affected significantly by impact velocity
Fig. 11. Influence of infiltration pressure on density Vs (䊉) of aluminum foam. The
porosity of the respective preform Vp is also indicated (). All preforms were fabri- [16]. During open-cell aluminum foam compression, the linear elas-
cated with a primitive porosity close to 25 % [39]. tic region is dominated by cell wall bending, whereas the plateau
region is affected by plastic hinges. When the cells have almost
collapsed completely, opposite cell walls contact and further strain
size, whereas it decreased with a decrease of wetting angle between compresses the matrix, which leads to a rapid increase in stress
the melt and the particle. The effect of infiltration pressure was ver- and the commencement of a densification stage [15]. Gibson and
ified experimentally. Berchem et al. [92] gave similar conclusions Ashby proposed a simple cubic element beam model to describe the
through analytical calculation. mechanical properties of aluminum foams based on porous lattice
Fischer et al. [93] studied the effects of mold and pouring materials [99]:
temperature on the open-cell aluminum foam structure that was Yf f
 a
prepared by investment casting. A decrease in mold temperature ≈C (3)
Ys s
and an increase in pouring temperature helped to improve the fill-  
ing capacity of the aluminum melt, which reduced the number of f
εD = 1 − 1.4 (4)
eutectic silicon particles and increased the strut thickness. Paul s
et al. [94] analyzed the influence of heat treatment on the structure where Y is the Young’s modulus or strength and  is the density.
of open-cell aluminum foams that were synthesized by investment Subscripts f and s represent the foam and matrix, respectively. εD
casting and found that solution annealing (W) and solution anneal- is the densification strain. The values of C and a are 1 and 2 to
ing plus aging (T6, T7) had almost no effect on the coarse Si and predict Young’s modulus, respectively, whereas they are 0.3 and
intermetallic phase morphology. 1.5, respectively, to predict the plastic collapse strength. The model
Surace et al. [95] tested the effects of composition, pressure, shows that the main mechanical properties of the aluminum foams
and sintering time on the structure of open-cell aluminum foams depend strongly on the relative density, which is suitable for most
prepared by sintering and dissolution through the design of experi- metal foams according to most experimental results [100–104].
ment (DOE) and ANOVA statistical analysis. The Al fraction was the Michailidis et al. [60] prepared the use of open-cell aluminum
most important parameter that affected the relative density of the foams with sucrose as space holders. The compression yield stress
final foams, whereas the sintering time and pressure showed little decreased with an increase in porosity and the platform region
effect. Hossein et al. [96] avoided Al and NaCl powder segregation became flatter and longer because there was more opportunity for
under high compaction pressures by adding a small amount of glyc- cell walls to deform and collapse. These findings agreed well with
erin and acetone during sintering and dissolution, which prevented the Gibson–Ashby model. Experiments showed that cell size had
partial spalling of the final aluminum foam and improved the pore little effect on the stress–strain curve of aluminum foams, as shown
structure homogeneity. in Fig. 12. Jiang et al. [83] obtained similar conclusions. Hasan et al.
T. Wan et al. / Journal of Materials Science & Technology 62 (2021) 11–24 19

Fig. 12. (a) Compressive stress–strain curves of aluminum foams with various porosities (b) reliance of yield stress on relative density of the aluminum foam (c) Compressive
stress–strain curves of aluminum foams with various pore sizes [60].

Fig. 14. Relationship between effective thermal conductivity and porosity (Ks = 205
W (m K)−1 , 33 ◦ C) [13].
Fig. 13. Effect of cell size on compressive mechanical properties of aluminum foams
with porosities of 65 % [33].

[33] found that the plateau stress grew with an increase of cell size where D represents the densification strain and Es is the Young’s
when the porosity was consistent. According to the experimental modulus of the matrix.
results, the plateau stress improved more than three times as the In addition to the mechanical properties, the functional char-
pore diameter changed from 1–1.5 mm to 3.5–5 mm, as shown in acteristics of open-cell aluminum foams are related to connected
Fig. 13, which was explained by the inconsistent cell wall thickness. cells and large specific surface areas have attracted much attention,
The mechanical properties of aluminum foams are influenced by such as the superior performance of heat dissipation [105–108].
the cell shape to some extent. For aluminum foams with irregular By combining large specific surface areas with the good thermal
cells, the stress concentration is prone to occur at sharp cell cor- conductivity of aluminum, open-cell aluminum foams are used
ners, which causes jagged fluctuations in the stress–strain curve often as heat exchangers in the field of forced convection heat
and unstable mechanical properties. Hence, compared with alu- transfer [105], such as heat exchangers for aerospace apparatus
minum foams with irregular cells, the mechanical properties of and heat sinks for electronic devices with high power and regen-
aluminum foams with spherical cells are more predictive and easily erators for heat engines. The thermal conductivity of aluminum
accountable. Hasan et al. [33] compared the mechanical properties foams is a key parameter for the above applications, which can
of open-cell aluminum foams that were prepared by using spheri- be measured experimentally by a steady method or a transient
cal urea and polygonal urea in the sintering and dissolving process. technique. The thermal conductivity of aluminum foams is depen-
The mechanical properties of the spherical-celled aluminum foams dent on the pore density, decreases as the pore density increases
were superior to those of the irregular-celled aluminum foams, [107,109], and is not influenced significantly by the pore size
such as an improved platform stress and yield stress, which agreed [107,109,110].
well with the theoretical predictions. Jiang et al. [28] reached Abuserwal et al. [13] tested the effective thermal conductivity
similar conclusions. Goodall et al. [82] contrasted the mechani- of open-cell aluminum foams via a steady state method, and found
cal properties of open-cell aluminum foams that were prepared by that the effective thermal conductivity increased with a decrease
using spherical and angular NaCl in the infiltration process. They in porosity and was hardly influenced by the pore size, as shown
found that the influence of cell shape on the mechanical proper- in Fig. 14. Similar conclusions were given by Solórzano et al. [111]
ties could be ignored if the aspect ratio of the angular particles was who used the transient plane source method to study the thermal
approximately the same, whereas the mechanical properties of the conductivity of open-cell aluminum foams.
spherical and angular pore foam differed when the aspect ratio of Theoretical models based on an analytical and empirical
the angular particles differed significantly. approach were developed to predict the thermal conductivity
The long plastic platform before the densification stage is of open-cell aluminum foams. An empirical scaling relationship
attributed to the energy absorption characteristics of aluminum between the effective thermal conductivity and porosity was pro-
foams, which is one of the most important properties of aluminum vided by Abuserwal et al. [13]:
foams, and gives them bright application prospects in the field
of collision and explosion protection. Gibson and Ashby provided 0.16
Kf = Ks (1 − ε)2.15(1−ε) (6)
an expression for the maximum absorbed energy per unit volume
Wmax :

Wmax D D
 2/3 where ε is the porosity, and the indices f and s represent the
= [1 − 3.1 ] (5) aluminum foam and matrix, respectively. Fig. 15 shows that the
Es ES Es prediction agrees well with a large porosity range. Ashby et al. [15]
20 T. Wan et al. / Journal of Materials Science & Technology 62 (2021) 11–24

Fig. 16. Electromagnetic interference shielding effectiveness of an open-cell alu-


minum foam: Comparison between experimental measurements and analytical data
Fig. 15. Predicted effective thermal conductivity by theoretical model versus poros- [38].
ity [13].

tions agreed well with the experiments. Wang et al. [115] used
provided a simple model to predict the thermal conductivity of the point matching to solve the sound wave propagation equation and
metal foam: obtained an optimal sound-absorbing cell size of the order of 0.1
ıf
  k mm. Li et al. [34] explored the sound-absorption properties of
f
= (7) open-cell aluminum foams with spherical cells, and found that the
ıs s
sound-absorption coefficient of the aluminum foam increased with
where k ranges from 1.65 to 1.8, and the indices f and s represent an increase of pore opening density and a decrease of pore opening
the aluminum foam and matrix, respectively. diameter from 0.3 to 0.4 mm.
Barari et al. [39] studied experimentally the thermal perfor- As commonly understood, metals have excellent electrical
mance of open-cell aluminum foams as thermal regenerators and conductivity and can absorb and reflect electromagnetic inter-
found that the pressure drop increased with velocity, whereas ference. Analogously, compared with polymers, metal foams can
the heat transfer decreased with an increase in Reynolds num- be furtherly developed for electromagnetic interference shielding
ber. Hwang et al. [112] investigated experimentally the convective because of their conductivity. Although not as powerful as dense
heat transfer of open-cell aluminum foams and showed that the metal, open-cell metal foams provide a high added value owing to
volumetric heat transfer coefficient decreased as the pore density their other good performance of low density, energy absorption,
increased at a certain Reynolds number. Hamid et al. [113] reported and allowing for mass, light and heat transmission with respect
that the heat transfer performance of open-cell metal foams as heat to traditional shielding materials. Giuseppina et al. [38] utilized a
sinks showed a positive relationship with the relative density and plasma model to describe the electromagnetic behavior SE of an
a negative relationship with cell size by numerical analysis. open-cell aluminum foam with thickness d:
No channels exist for material circulation because of indepen-
dent cells in closed-cell aluminum foams, whereas many channels SE Analytical = R + A + M = f(ωp , ) (10)
are derived from connected cells for open-cell aluminum foams,
and these channels can be used for sound wave propagation and (no + ns )2
R= (11)
loss [114]. Hence, open-cell aluminum foams can be used as excel- 4n0 ns
lent sound-absorbing materials. When sound waves are incident  
(n0 − ns )2
on cells of open-cell aluminum foams, three attenuation mecha- M = 1− exp (i2kd) (12)
nisms exist: (i) vibration is generated by sound waves and causes (n0 + ns )2
air flow inside the cells, and friction exists between flowing air A = exp (ikd) (13)
and the cell wall, which causes a portion of the sound energy to
be converted into thermal energy and attenuates sound waves. where R, A, and M represent the reflection, the absorption and
(ii) High-frequency sound waves accelerate air movement inside the multiple reflections contributions to SE; ωp and  represent
the cells, which results in stronger heat exchange between the air the plasma frequency and damping factor respectively. The results
and the cell walls, which weakens the sound waves. The connected illustrated in Fig. 16 indicate a good agreement between experi-
cells form a Helmholtz resonant sound-absorption structure, which mental and analytical measurements.
creates strong friction between the intensely vibrating air and the Aluminum foams are environmentally friendly and weather-
connected cells, and results in more sound energy consumption. resistant composite materials that are fire-resistant, flame-
The effect of the Helmholtz resonant absorption is apparent when retardant, and recyclable. Compared with closed-cell aluminum
a cavity is introduced into the back of the sound-absorbing mate- foams, open-cell aluminum foams are applied when light weight,
rial. Han et al. [11] discovered that the connected pores of open-cell load-bearing, energy absorption or anti-collision are required. They
aluminum foams play a role in sound-absorption performance. Lu are also used as high-temperature filters, for heat dissipation and
et al. [35] studied the sound-absorption performance of open-cell sound absorption from interconnected cells.
aluminum foams and proposed an improved theoretical model to In the field of space debris shields, a layer of open-cell aluminum
calculate the sound-absorption coefficient ˛ for predictions: foam has been embedded in the multi-layer defensive structure by
the European space agency to form significant multiple shocks to
4Rn /0 c0
˛=  2   (8) the projectile, and melt impacting projectiles, which reduces the
1 + Rn /0 c0 + Mn /0 c0 damage of the projectile to the spacecraft. Fig. 17 shows the debris
cloud evolution when a defensive structure of a layer of open-cell
Zn = Rn + iM n (9)
aluminum foams is applied. Fig. 18 shows the energy absorber that
where 0 is the density of air, c0 is the speed of sound in air, and is applied with open-cell aluminum foams used by NASA to protect
Zn is the acoustic impedance of the acoustic system. The predic- core components in space.
T. Wan et al. / Journal of Materials Science & Technology 62 (2021) 11–24 21

Fig. 17. Debris cloud evolution to apply open-cell aluminum foam of 10 ppi and 40 mm thickness [116].

Fig. 18. Energy absorber formed with open-cell aluminum foams [117].

Fig. 19. Applications of open-cell aluminum foams manufactured by ERG (a)–(c) heat exchangers, (d) EMI shielding, (e) fluid control baffle, (f) cryogenic tank [117].

Open-cell aluminum foams that are manufactured by ERG mechanical features with additional functional properties. The key
(Oakland, CA, USA) via investment casting are applied as heat to tailor properties of open-cell aluminum foams is the precise con-
exchangers, insulation devices, and electromagnetic shielding trol over the pore structures, such as porosity, pore shape, pore
boxes on satellites and space station-related devices by NASA, as connectivity, pore size, and pore size distribution, which can be
illustrated in Fig. 19. Open-cell aluminum foams that are produced perfectly achieved by a sacrificial pattern assisted replication pro-
by Exxentis (Wettingen, Switzerland) through infiltration are used cess.
as mufflers, heat exchangers, and vacuum-generating molds, as Based on the performance requirements of structural or
shown in Fig. 20. lightweight, sound absorber, thermal management, and elec-
In the acoustic engineering field of absorption and noise reduc- tromagnetic shielding applications, it is necessary to develop
tion, Z̈HB aluminum foam acoustic panelsḧave been applied as applicable theoretical or numerical models allowing for clari-
sound-absorbing materials. Fig. 21 shows projects in which open- fying the relations between pore structures and corresponding
cell aluminum foams are applied on sound barriers. behaviors. These pore structures predesigned can be furtherly
realized accurately by pattern design during the replication pro-
cess.
4. Outlooks
For the additive manufacturing, 3D printing of patterns like gyp-
sum or inorganic salt allows for combining the ease of pattern
As a structural-functional integrated material, open-cell alu-
removal with the freedom production of complicated structures,
minum foam opens up new potentialities to combine the
22 T. Wan et al. / Journal of Materials Science & Technology 62 (2021) 11–24

Fig. 20. Application of open-cell aluminum foams fabricated by Exxentis (a)–(b) vacuum-forming molds, (c) heat exchanger, (d)–(e) pneumatic silencers, (f) weapon silencer
[118].

Fig. 21. Application of open-cell aluminum foams as sound barriers [119].

which shows the highest control over the pore structure but intro- sive inorganic salt particles with non-corrosive or weakly corrosive
duces relatively high cost and sophisticated procedures. particles can resolve or alleviate the corrosion trouble.
As for open-cell aluminum foams fabricated based on poly- Flexible employment of varied 3D printed patterns, polymer
mer sponges, highly interconnected network structures allow a sponges, and space holders contributes to the manufacture of open-
high mass transfer efficiency with low flow resistance, which are cell aluminum foams with desirable structures and properties to
popularly applied in heat exchange fields. However, traditionally meet the demands of engineering applications. In order to furtherly
manufactured polymer sponges exhibit stochastic and irregular broaden structural and functional applications in some specific sit-
structures, thus causing difficulty in developing models to describe uations, a suitable choice of patterns and appropriate models to
the properties of the final foams accurately. In the future, polymer predict the behaviors of the final foams should be taken into con-
sponges with regular and periodic structures made via technical sideration. The emphasis and difficulty of current research lie in
innovation may solve the problem. Additionally, special atten- designing suitable structures of open-cell aluminum foams to sat-
tion should be paid to simplifying the preparation procedure and isfy performance in some specific and multiple target applications.
improving the strength of final foams, mainly in investment casting. CAD-associated numerical simulation and analytical solutions may
The space holder method including the infiltration process and be used to evaluate the mechanical and physical behaviors of the
powder-metallurgy method allows for controlling the final pore predesigned structures.
structure directly through the replication of space holder parti-
cles in the most straightforward and lowest-cost way. However, 5. Conclusion
nearly rigid point contact between the hard and brittle particles
will lead to a poor pore interconnection, no matter in infiltration Because of their permeability or open porosity, open-cell
casting or powder metallurgy. Thus adverse effects on functional aluminum foams are superior in terms of their structural and
characteristics and the problem of pattern removal appear, which functional performance. As the main method to prepare open-
can be avoided to some extend by improving the compressibility cell aluminum foams, the replication process reflects the unique
of the used particles, and porosity can be adjusted simultaneously advantage of pattern design to obtain tailored cell structures to
in a wide range in this way. Moreover, a replacement of the corro- meet special requirements and applications. Cell structures can
be controlled through pattern design and processing parameters.
T. Wan et al. / Journal of Materials Science & Technology 62 (2021) 11–24 23

Besides applications in which a light weight, energy absorption, [41] E.M. Elizondo Luna, F. Barari, R. Woolley, R. Goodall,J.Vis. Exp. (2014),
and vibration damping are required, open-cell aluminum foams e52268.
[42] K.S. Chou, M.A. Song, Scr. Mater. 46 (2002) 379–382.
function well in fields of sound absorption, heat dissipation, and [43] R. Goodall, J.F. Despois, A. Marmottant, L. Salvo, A. Mortensen, Scr. Mater. 54
electromagnetic shielding. For open-cell aluminum foams based (2006) 2069–2073.
on space holders, more effective and cheaper ways should be used [44] Y. Xue, X. Wang, W. Wang, X. Zhong, F. Han, Mater. Sci. Eng. A 722 (2018)
255–262.
to improve the porosity, remove the space holders, and prevent [45] A.H. Brothers, D.C. Dunand, Adv. Eng. Mater. 8 (2006) 805–809.
corrosion, whereas technical innovation must be focused on simpli- [46] Y. Yamada, K. Shimojima, Y. Sakaguchi, M. Mabuchi, M. Nakamura, T.
fying the preparation procedure and improving the strength in the Asahina, T. Mukai, H. Kanahashi, K. Higashi, Mater. Sci. Eng. A 280 (2000)
225–228.
polymer sponge method. The design of suitable foam structures is
[47] A. Jinnapat, A. Kennedy, Metals 2 (2012) 122–135.
important to broaden the application in specific and multiple target [48] Y. Xue, F. Han, Mater. Lett. 254 (2019) 99–102.
fields. [49] Y. Huang, Y. Xue, X. Wang, F. Han, Mater. Sci. Eng. A 696 (2017) 520–528.
[50] H. Wang, Y. Fu, M. Su, H. Hao,Mater. Des. 167 (2019), 107631.
[51] Y. Yosida, C. Hayashi, Proc. Conf. Casting Science and Technology,
Acknowledgements September, 1990, p. 103.
[52] Y. Yamada, K. Shimojima, Y. Sakaguchi, M. Mabuchi, M. Nakamura, T.
Asahina, T. Mukai, H. Kanahashi, K. Higashi, Adv. Eng. Mater. 2 (2000)
This work was financially supported by the National Natural 184–187.
Science Foundation of China (No. 51771101). [53] W.C. Sun, Device and Preparation Method of Aluminum Foam Roll Blanket,
Chinese patent, No. CN1515700, 2004. (in Chinese).
[54] W.C. Sun, X.J. Wu, Metallic Functional Materials. 10 (2003) 19–21 (in
References Chinese).
[55] Y. Yamada, C. Wen, K. Shimojima, M. Mabuchi, M. Nakamura, T. Asahina, T.
[1] X. Zheng, H. Lee, T.H. Weisgraber, M. Shusteff, J. DeOtte, E.B. Duoss, J.D. Aizawa, K. Higashi, Mater. Trans. 41 (2000) 1136–1138.
Kuntz, M.M. Biener, Q. Ge, J.A. Jackson, S.O. Kucheyev, N.X. Fang, C.M. [56] J.H. Kim, R.H. Kim, H.S. Kwon, Electrochem. commun. 10 (2008) 1148–1151.
Spadaccini, Science 344 (2014) 1373–1377. [57] C. San Marchi, J.F. Despois, A. Mortensen, Microstr. Inves. Analys. 4 (2000)
[2] J. Banhart, Adv. Eng. Mater. 15 (2013) 82–111. 34–39.
[3] X.N. Liu, Y.X. Li, X. Chen, Y. Liu, X.L. Fan, J. Mater. Sci. 45 (2010) 6481–6493. [58] Y.Y. Zhao, D.X. Sun, Scr. Mater. 44 (2001) 105–110.
[4] Y. Liu, Y. Li, H. Zhang, J. Wan, Rare Metal Mat. Eng. 34 (2005) 1128–1130. [59] H. Bafti, A. Habibolahzadeh, Mater. Des. 31 (2010) 4122–4129.
[5] L. Chen, H. Zhang, Y. Liu, Y. Li, Acta Metall. Sin. 48 (2012) 329–333 (in [60] N. Michailidis, F. Stergioudi, A. Tsouknidas, E. Pavlidou, Mater. Sci. Eng. A
Chinese). 528 (2011) 1662–1667.
[6] B. Xie, Y.Z. Fan, T.Z. Mu, B. Deng, Mater. Sci. Eng. A 708 (2017) 419–423. [61] R. Kumar, H. Jain, S. Sriram, A. Chaudhary, A. Khare, V.A.N. Ch, D.P.
[7] M.H. Golabgir, R. Ebrahimi-Kahrizsangi, O. Torabi, H. Tajizadegan, A. Mondal,Mater. Chem. Phys. 240 (2020), 122274.
Jamshidi, Adv. Powder Technol. 25 (2014) 960–967. [62] P. Heinl, L. Muller, C. Korner, R.F. Singer, F.A. Muller, Acta Biomater. 4 (2008)
[8] A. Inoue, H. Kimura, Mater. Sci. Eng. A 286 (2000) 1–10. 1536–1544.
[9] F. Xu, X. Zhang, H. Zhang, Eng. Struct. 171 (2018) 309–325. [63] S. McKown, Y. Shen, W.K. Brookes, C.J. Sutcliffe, W.J. Cantwell, G.S. Langdon,
[10] Y. Mu, G. Yao, H. Luo, Mater. Des. 31 (2010) 610–612. G.N. Nurick, M.D. Theobald, Int. J. Impact Eng. 35 (2008) 795–810.
[11] F. Han, G. Seiffert, Y. Zhao, B. Gibbs, J. Phys. D-Appl. Phys. 36 (2003) 294–302. [64] M. Agarwala, D. Bourell, J. Beaman, H. Marcus, J. Barlow, Rapid Prototyping J.
[12] Z. Xu, H. Hao, J. Alloys Compd. 617 (2014) 207–213. 1 (1995) 26–36.
[13] A.F. Abuserwal, E.M. Elizondo Luna, R. Goodall, R. Woolley, Int. J. Heat Mass [65] N.A. Meisel, C.B. Williams, A. Druschitz, the 23rd Annual International Solid
Transf. 108 (2017) 1439–1448. Freeform Fabrication Symposium - An Additive Manufacturing Conference,
[14] J. Banhart, Prog. Mater. Sci. 46 (2001) 559–632. Austin, TX, U.S., August 6-8, 2012.
[15] M.F. Ashby, A. Evans, N.A. Fleck, L.J. Gibson, J.W. Hutchinson, H.N.G. Wadley, [66] C. Yan, L. Hao, A. Hussein, S.L. Bubb, P. Young, D. Raymont, J. Mater, Process.
Metal Foams: A Design Guide, first ed., Butterworth-Heinemann, Boston, Technol. 214 (2014) 856–864.
2000, pp. 1–5. [67] C. Qiu, S. Yue, N.J.E. Adkins, M. Ward, H. Hassanin, P.D. Lee, P.J. Withers, M.M.
[16] F. Garcia-Moreno, Materials 9 (2016) 85. Attallah, Mater. Sci. Eng. A 628 (2015) 188–197.
[17] Y. Cheng, Y. Li, X. Chen, T. Shi, Z. Liu, N. Wang, Metall. Mater. Trans. B 48 [68] I. Maskery, N.T. Aboulkhair, A.O. Aremu, C.J. Tuck, I.A. Ashcroft, R.D.
(2017) 754–762. Wildman, R.J.M. Hague, Mater. Sci. Eng. A 670 (2016) 264–274.
[18] N. Wang, X. Chen, Y. Li, Z. Liu, Z. Zhao, Y. Cheng, Y. Liu, H. Zhang, Colloid Surf. [69] D. Snelling, Q. Li, N. Meisel, C.B. Williams, R.C. Batra, A.P. Druschitz, Adv. Eng.
A-Physicochem. Eng. Asp. 527 (2017) 123–131. Mater. 17 (2015) 923–932.
[19] X. Ding, Y. Liu, X. Chen, H. Zhang, Y. Li, Mater. Lett. 216 (2018) 38–41. [70] A. Lyons, S. Krishnan, J. Mullins, M. Hodes, D. Hernon, 20th Annual
[20] S.C. Han, K. Kang, Mater. Today 31 (2019) 31–38. International Solid Freeform Fabrication Symposium, Austin, Texas, U.S.,
[21] P. Schüler, S.F. Fischer, A. Bührig-Polaczek, C. Fleck, Mater. Sci. Eng. A 587 August 3-5, 2009.
(2013) 250–261. [71] S. Chiras, D.R. Mumm, A.G. Evans, N. Wicks, J.W. Hutchinson, K. Dharmasena,
[22] X. Yang, Q. Hu, J. Du, H. Song, T. Zou, J. Sha, C. He, N. Zhao, Int. J. Fatigue 121 H.N.G. Wadley, S. Fichter, Int. J. Solids Struct. 39 (2002) 4093–4115.
(2019) 272–280. [72] L.C. Wang, F. Wang, Acta Metall. Sin. (Eng. Lett.)14 (2009) 27–32.
[23] K. Lietaert, J. van Deursen, T. Lapauw, L. Weber, A. Mortensen, J. [73] L. Wang, H. Li, F. Wang, J. Ren, China Foundry 2 (2005) 56–59.
Vleugels,Mater. Charact. 157 (2019), 109895. [74] A.M. Matz, B.S. Mocker, D.W. Müller, N. Jost, G. Eggeler, Adv. Mater. Sci. Eng.
[24] Y. Conde, J.F. Despois, R. Goodall, A. Marmottant, L. Salvo, C. San Marchi, A. 2014 (2014) 1–9.
Mortensen, Adv. Eng. Mater. 8 (2006) 795–803. [75] A.H. Brothers, D.C. Dunand, Mater. Sci. Eng. A 489 (2008) 439–443.
[25] L. Polonsky, S. Lipson, H. Markus, Mod. Cast. 39 (1961) 57–71. [76] R. Goodall, A. Mortensen, Adv. Eng. Mater. 9 (2007) 951–954.
[26] H.A. Kuchek, Method of Making Porous Metallic Article, US Patent, No. [77] R. Goodall, J.F. Despois, A. Mortensen, J. Eur. Ceram. Soc. 26 (2006)
3236706, 1966. 3487–3497.
[27] L.Q. Ma, Z.L. Song, D.P. He, Scr. Mater. 41 (1999) 785–789. [78] A. Jinnapat, A. Kennedy, Metals 1 (2011) 49–64.
[28] B. Jiang, N. Zhao, C. Shi, J. Li, Scr. Mater. 53 (2005) 781–785. [79] S. Yu, J. Liu, M. Wei, Y. Luo, X. Zhu, Y. Liu, Mater. Des. 30 (2009) 87–90.
[29] C. San Marchi, A. Mortensen, Acta Mater. 49 (2001) 3959–3969. [80] L. Stanev, B. Drenchev, A. Yotov, R. Lazarova, J. Mater. Sci. Technol. 22 (2014)
[30] J.F. Despois, Y. Conde, C. San Marchi, A. Mortensen, Adv. Eng. Mater. 6 (2004) 44–53.
444–447. [81] B. Jiang, N.Q. Zhao, C.S. Shi, X.W. Du, J.J. Li, H.C. Man, Mater. Lett. 59 (2005)
[31] T.G. Nieh, K. Higashi, J. Wadsworth, Mater. Sci. Eng. A 283 (2000) 105–110. 3333–3336.
[32] S.F. Fischer, Mater. Lett. 184 (2016) 208–210. [82] R. Goodall, A. Marmottant, L. Salvo, A. Mortensen, Mater. Sci. Eng. A 465
[33] H. Bafti, A. Habibolahzadeh, Mater. Des. 52 (2013) 404–411. (2007) 124–135.
[34] Y. Li, X. Wang, X. Wang, Y. Ren, F. Han, C. Wen,J. Appl. Phys. 110 (2011), [83] B. Jiang, Z. Wang, N. Zhao, Scr. Mater. 56 (2007) 169–172.
113525. [84] A. Jinnapat, A.R. Kennedy, J. Alloys Compd. 499 (2010) 43–47.
[35] T.J. Lu, F. Chen, D. He, J. Acoust. Soc. Am. 108 (2000) 1697–1709. [85] J.A. Liu, F. Gao, Y.Q. Rao, C.L. Wu, Y. Liu, Mater. Trans. 55 (2014) 1906–1908.
[36] S. Chen, M. Bourham, A. Rabiei, Procedia Mater. Sci. 4 (2014) 293–298. [86] C. Gaillard, J.F. Despois, A. Mortensen, Mater. Sci. Eng. A 374 (2004) 250–262.
[37] L. Catarinucci, O. Losito, L. Tarricone, F. Pagliara, IEEE International [87] M. Bahraini, L. Weber, J. Narciso, A. Mortensen, J. Mater. Sci. 40 (2005)
Symposium on Electromagnetic Compatibility, Portland or U.S.A, August, 2487–2491.
2006, pp. 14–18. [88] V. Michaud, A. Mortensen, Compos. Pt. A-Appl. Sci. Manuf. 32 (2001)
[38] G. Monti, L. Catarinucci, L. Tarricone, 3rd European Conference on Antennas 981–996.
and Propagation, Berlin, Germany, March 23-27, 2009. [89] J. Hilden, K. Trumble, J. Colloid Interface Sci. 267 (2003) 463–474.
[39] F. Barari, E.M.E. Luna, R. Goodall, R. Woolley, J. Mater. Res. 28 (2013) [90] C. Garcia-Cordovilla, E. Louis, J. Narciso, Acta Mater. 47 (1999) 4461–4479.
2474–2482. [91] F. Chen, A.W. Zhang, D.P. He, J. Mater. Sci. 36 (2001) 669–672.
[40] J.F. Despois, A. Marmottant, L. Salvo, A. Mortensen, Mater. Sci. Eng. A 462 [92] K. Berchem, U. Mohr, W. Bleck, Mater. Sci. Eng. A 323 (2002) 52–57.
(2007) 68–75.
24 T. Wan et al. / Journal of Materials Science & Technology 62 (2021) 11–24

[93] S.F. Fischer, P. Schüler, C. Fleck, A. Bührig-Polaczek, Acta Mater. 61 (2013) [109] J.W. Paek, B.H. Kang, S.Y. Kim, J.M. Hyun, Int. J. Thermophys. 21 (2000)
5152–5161. 453–464.
[94] P. Schüler, R. Frank, D. Uebel, S.F. Fischer, A. Bührig-Polaczek, C. Fleck, Acta [110] R. Dyga, S. Witczak, the 20th International Congress of Chemical and Process
Mater. 109 (2016) 32–45. Engineering CHISA 2012, Prague, Czech Republic, August 25-29, 2012.
[95] R. Surace, L.A.C. De Filippis, A.D. Ludovico, G. Boghetich, Mater. Des. 30 [111] E. Solórzano, M.A. Rodriguez-Perez, J.A. de Saja, Adv. Eng. Mater. 10 (2008)
(2009) 1878–1885. 596–602.
[96] H. Gilani, S. Jafari, R. Gholami, A. Habibolahzadeh, M. Mirshahi, Met. [112] J.J. Hwang, G.J. Hwang, R.H. Yeh, C.H. Chao, J. Heat Transfer 124 (2002)
Mater.-Int. 18 (2012) 327–333. 120–129.
[97] J. Banhart, Adv. Eng. Mater. 2 (2000) 188–191. [113] H.R. Seyf, M. Layeghi,J. Heat Transfer 132 (2010), 071401.
[98] S. Prabhu, V.K.B. Raja, R. Nikhil, Appl. Mech. Mater. 766 (2015) 511–517. [114] J.P. Arenas, M.J. Crocker, Sound Vib. 44 (2010) 12–18.
[99] L. Gibson, M.F. Ashby, Cellular Solids: Structures and Properties, second ed., [115] X. Wang, T.J. Lu, J. Acoust. Soc. Am. 106 (1999) 756–765.
Cambridge University Press, London, 1997, pp. 175–234. [116] R. Destefanis, F. Schäfer, M. Lambert, M. Faraud, E. Schneider, Int. J. Impact
[100] J.T. Beals, M.S. Thompson, J. Mater. Sci. 32 (1997) 3595–3600. Eng. 29 (2003) 215–226.
[101] I.W. Hall, M. Guden, C.J. Yu, Scr. Mater. 43 (2000) 515–521. [117] ERG. Available online: http://ergaerospace.com/photo-gallery/ (Accessed on
[102] O.B. Olurin, N.A. Fleck, M.F. Ashby, Mater. Sci. Eng. A 291 (2000) 136–146. 11 December 2019).
[103] A.R. Kennedy, J. Mater. Sci. 39 (2004) 3085–3088. [118] Exxentis. Available online: http://www.exxentis.co.uk/porous-aluminium/
[104] U. Ramamurty, A. Paul, Acta Mater. 52 (2004) 869–876. applications/ (Accessed on 11 December 2019).
[105] T.J. Lu, H.A. Stone, M.F. Ashby, Acta Mater. 46 (1998) 3619–3635. [119] Shanghai Zhonghui Foam Aluminum Co., Ltd. Available online:
[106] T.J. Lu, C. Chen, Acta Mater. 47 (1999) 1469–1485. http://www.zhonghuiaf.com/02300054/php/pic.php?menuid=3&lang=0/
[107] E. Solórzano, J.A. Reglero, M.A. Rodríguez-Pérez, D. Lehmhus, M. Wichmann, (Accessed on 11 December 2019).
J.A. De Saja, Int. J. Heat Mass Transf. 51 (2008) 6259–6267.
[108] T. Fiedler, E. Solórzano, A. Garcia-Moreno, F. Öchsner, I. Belova, G. Murch,
Materialwiss. Werkstofftech. 40 (2009) 139–143.

You might also like