You are on page 1of 12

journal of the mechanical behavior of biomedical materials 15 (2012) 141–152

Available online at www.sciencedirect.com

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

Research Paper

Mechanical properties of luffa sponge

Jianhu Shena, Yi Min Xiea,n, Xiaodong Huanga, Shiwei Zhoua, Dong Ruanb
a
Centre for Innovative Structures and Materials, School of Civil, Environmental and Chemical Engineering, RMIT University, GPO Box 2476,
Melbourne 3001, Australia
b
Faculty of Engineering and Industrial Sciences, Swinburne University of Technology, John Street, Hawthorn, VIC 3122, Australia

art i cle i nfo ab st rac t

Article history: The paper presents the first scientific study of the stiffness, strength and energy absorption
Received 3 April 2012 characteristics of the luffa sponge with a view to using it as an alternative sustainable
Received in revised form engineering material for various practical applications. A series of compression tests on
25 June 2012 luffa sponge columns have been carried out. The stress–strain curves show a near constant
Accepted 6 July 2012 plateau stress over a long strain range, which is ideal for energy absorption applications. It
Available online 17 July 2012 is found that the luffa sponge material exhibits remarkable stiffness, strength and energy

Keywords: absorption capacities that are comparable to those of some metallic cellular materials in a

Luffa sponge similar density range. Empirical formulae have been developed for stiffness, strength,

Mechanical properties densification strain and specific energy absorption at the macroscopic level. A comparative

Energy absorption study shows that the luffa sponge material outperforms a variety of traditional engineering

Sustainability materials.
& 2012 Elsevier Ltd. All rights reserved.

1. Introduction to accommodate the natural environment to which they were


exposed, a large number of biological systems evolve periodic
Biological materials and structures have distinguished them cells with self-similar hierarchical microstructures. Similar to
from traditional human developed counterparts because of turtle’s shell (Damiens et al., 2012), jellyfish mesogloeas (Zhu
their unique characteristics (Meyers et al., 2011). The most et al., 2012), wood (Stanzl-Tschegg et al., 2011; Wegst, 2011),
attractive one is their long-term sustainability to the natural and E. aspergillum sponge (Mayer, 2011), those hierarchical
environment. The performance of the traditional man-made architectures are optimised or partially optimised and can
materials, notably various metals, metallic alloys, ceramics, achieve multi-functions with high toughness and efficiency.
plastics, as well as their composites, significantly surpass the As we approach the limit of non-renewable natural resources,
biological materials (Bonderer et al., 2008; Zhang et al., 2011). these properties are essential for the long-term sustainability
However, most of the man-made materials are not environ- of our habitat, and is becoming increasingly significant to
mentally friendly and little has been concerned with their human civilisations (Zhang et al., 2011).
sustainability (Zhang et al., 2011). By contrast, those biologi- Luffa sponge is one of such commercially viable and
cal materials are based on relatively weak base materials environmentally acceptable biological material derived from
such as minerals and proteins which are easily degradable, fruit of Luffa cylindrica (LC) plant and having recycling cap-
bio-compatible, pollution-free, recyclable and energy-effi- ability and triggered biodegradability (John and Thomas,
cient (Zhang et al., 2011). Furthermore, over millions of years 2008; Oboh and Aluyor, 2009). It is relatively stable in their

n
Corresponding author. Tel.: þ61 3 9925 3655.
E-mail address: mike.xie@rmit.edu.au (Y. Min Xie).

1751-6161/$ - see front matter & 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.jmbbm.2012.07.004
142 journal of the mechanical behavior of biomedical materials 15 (2012) 141 –152

intended lifetime but would biodegrade after disposal in supplier. The luffa fruits were harvested after they were fully
composting conditions. At the same time, the fruits of LC mature with their skin turning brown. The dried luffa fruits
have a netting-like fibrous vascular system. When they are were slightly squashed laterally to crack and remove the skin.
dried, the fibrous network structure serves like an open cell Then the two ends of luffa fruits were cut and the seeds were
foam material. It has the potential to be used as an alter- removed. The original luffa sponges were bleached using
native material for man-made cellular materials. The impor- liquid chlorine bleach (4%) for about 1 h to improve their
tance of biological materials such as the luffa sponge is appearance by making them whiter. After that, they were
growing as we search for sustainable solutions using new soaked in clean water for half an hour and then dried in the
materials. sun. The chemical composition of the luffa sponge depends
Only a limited amount of research has been conducted on on several factors, such as plant origin, weather condition,
the luffa sponge as a source of bio-fibres and bio-composites soil, pre-treatment, etc. A set of reference values for the
in the last ten years. Those researches indicated that it was a chemical composition of LC foam can be found in a previous
potential alternative material for packaging (Mazali and research (Siqueira et al., 2010).
Alves, 2005), water absorption (Bal and Bal, 2004; Demir The measured maximum diameter for each specimen of
et al., 2006), and waste water treatment (Laidani et al., 2001; dry luffa sponges tested was within the range of 55–86 mm.
Oboh et al., 2011). The luffa fibres were also used as reinfor- According to the standard of compressive tests and recent
cement fibre for other materials (Boynard and D’Almeida, research (Ashby et al., 2000) on other cellular material, the
2000; Ghali et al., 2009; Laranjeira et al., 2006; Paglicawan specimen size effect is negligible for foams when the dimen-
et al., 2005; Tanobe et al., 2005) and cell immobilisation for sion of specimen is sufficiently larger than the cell size, i.e., 7
biotechnology (Chen and Lin, 2005; Roble et al., 2002; Tavares times of the cell size for metallic foams. Thus the specimen
et al., 2008; Zampieri et al., 2006). At the same time, sponge for the luffa sponge should be large enough to eliminate the
gourd (LC), the origin of luffa sponge material, have not yet specimen size effect. The available maximum size along
had their potentialities fully explored. With regard to indus- radial direction for luffa sponge is whole section of the luffa
trial and technological development, the cost of fuel is on the sponge column. For most of the luffa sponge column, the
increase. Oil can be extracted from seeds for industrial use section area at two ends of the specimen is different, thus
(Bal and Bal, 2004). The oil extracted from LC is finding two thin layers of luffa sponge slices were cut at each end.
increasing use in the production of biodiesel which is now From a single luffa sponge, several cylindrical test specimens
gaining wide acceptance because of low CO2 emission and (approximately 50 mm high) and two measurement slices
other considerations (Ajiwe et al., 2005). (approximately 5 mm thick) from both ends were cut using a
However, there is a lack of scientific data on the mechanical bandsaw cutting machine. After the initial cutting, the speci-
properties of luffa sponge material because up to now its mens were milled further to make the two end surfaces
main practical use is a body scrub in the bathroom. Due to smooth and parallel. For calculation purpose, the cross
the lack of experimental proofs, their complex hierarchy sectional area was taken as the enclosed section (ignoring
microstructures and other common limitations of biological any internal voids). Because the cross section for most
materials, potential applications have not been implemented specimens was not a perfect circle, the actual cross sectional
in practice for luffa sponge material as well as luffa fibres. area was determined from photographs taken of each mea-
To this end, the mechanical properties of luffa sponge surement slice and processed using image software—
columns were tested and compared with other cellular Photoshop. The actual cross sectional area from the two
materials to check the performance of this light weight measurement slices was then averaged. An equivalent dia-
material. Uniaxial quasi-static compressive tests were con- meter was calculated from the average area. This equivalent
ducted at a strain rate of 103 s1 by using an Instron diameter was within the range of 42–81 mm. A total of 26
machine to study the mechanical properties of luffa sponge specimens were tested. Fig. 1(a) shows a typical cylindrical
material. Cylindrical specimens with different relative den- specimen for a compressive test. The volume of the luffa
sity were tested at a room temperature of 251 C and a sponge specimen is the product of this cross sectional area
humidity of approximately 40%. An energy efficiency method and the height of the specimen.
was adopted to obtain the values of the densification strain
and plateau stress, and thus the energy absorption capacity 2.2. Geometry features and standard variation of density
per unit volume. The experimental results were also dis-
cussed together with test results published by other research- After the specimens were cut, three different core topologies
ers for other celluar materials. were observed from our luffa sponge specimens as shown in
Fig. 2(a). The variation in specimen diameter and cross
sectional area is shown in Fig. 2(b).
2. Experiments The luffa sponge column is composed of luffa fibres. Those
fibres interconnect with each other and form networks with
2.1. Specimens micro-trusses. The length of those micro-trusses represents
the cell length of the luffa sponge material. A rough mea-
The luffa sponge used in our experiments was obtained from surement indicates that the average length of those micro-
pharmacies in Australia which was sold as a bath sponge. A trusses is in the scale of millimetre (1–5 mm). The major
brief treatment procedure for manufacturing these bath orientation of these luffa fibres in the luffa sponge columns
sponges from natural luffa fruits was provided by the exhibits a regular pattern. According to different orientation
journal of the mechanical behavior of biomedical materials 15 (2012) 141 –152 143

Core region Inner


surface

Interlayer

Outer
surface

Longitudinal
direction

Circumferential Radial
direction direction

Fig. 1 – A luffa cylindrical specimen for a compressive test: (a) Different regions. (b) Orientation of luffa fibres (left: inner
surface; middle: outer surface; right: cross section). (c) Microstructures of luffa fibres (left: luffa fibre in 1 mm; middle: cross
section of luffa fibre in 101 mm; right: cross section of luffa fibre in 102 mm).

of the luffa fibre, the luffa sponge cylinder can be divided into 350–650 kg/m3. A previous research also showed that
four regions, namely, the inner surface, the outer surface, the density of the luffa fibre is approximately 353 kg/m3,
interlayer and core as shown in Fig. 1(a). On the inner surface, with a Young’s modulus of 1332 MPa (735%) and tensile
the thickest luffa fibre grows along longitudinal directions. strength of 11.1 MPa (785%) (Paglicawan et al., 2005).
While on the outer surface, the thickest luffa fibre grows And the density for cell wall of luffa fibre is 820–
along circumferential directions. In the core region, the 920 kg/m3 (Siqueira et al., 2010). The microstructures of
strongest luffa fibre is along radial direction. In the interlayer cell wall of luffa fibre and their mechanical properties
between inner surface and outer surface, the fibre grows in remained unknown at present. For this reason, the empirical
all three directions as shown Fig. 1(b). formulae in this paper will be given on the macroscopic
It can be seen from Fig. 3, the variation of density of the level.
luffa sponge material is irrelevant to the core topology and Similar to other cellular materials, the mechanical proper-
the size of the luffa sponge columns. Luffa sponge is a ties of the luffa sponge material are closely related to its
material with hierarchical architectures at several length density rather than other geometric features mentioned
scales. At each hierarchical level, there is a corresponding above. Thus the specimens were grouped only according to
density. We observed three levels, namely, the luffa sponge their densities in order to obtain a reliable empirical formula
column (50 mm), the luffa fibre (1 mm) and the cell wall of to represent their mechanical performance. The effect of
luffa fibre (0.01 mm) as shown in Fig. 1. The density of luffa specimen size and core topology on its mechanical properties
sponge column is 25–65 kg/m3. The density for luffa fibre is was disregarded in the current paper.
144 journal of the mechanical behavior of biomedical materials 15 (2012) 141 –152

2.3. Experimental procedures

Uniaxial compressive tests were conducted at very low strain


rates, namely, 103 s1, to obtain the mechanical properties
under quasi-static loading. An Instron machine was used to
conduct most of these tests as shown in Fig. 4. A Shimaszu
machine was also used for several trail tests. Both machines
were calibrated using the same load-cell to ensure the
consistency of the experiment results. The displacement
and load curves were recorded and the average compressive
force, densification strain and energy absorption can be
worked out using an energy efficiency method. A tempera-
ture and humidity metre was used to monitor the tempera-
ture and humidity around the test specimens during
the tests.
Like other natural sponges such as the sponge Euplectella
aspergillum (Johnsona et al., 2010), the mechanical properties
of luffa sponge are also influenced by the moisture. Thus the
Fig. 2 – Different core topologies and sizes for luffa sponge humidity during the test was monitored. The specimens were
specimens: (a) core topology (three different patterns are placed in the test room for at least 2 h before the compressive
shown, namely two links, three links and four links); (b) test to eliminate the difference in temperature and humidity.
variation of diameter of cross section (from 46 mm to The temperature in the test room was from 201 to 281, but
81 mm). during each test its variation was less than 11. The humidity
was between 30% and 44%, but during each test its variation
65
was less than 1.0%. Photographs were taken during the quasi-
static compressive test at every 30 s which corresponded to a
60 1.5 mm displacement interval for the test velocity used. It
Effect of outer diameter
55 should be noted that for the second specimen, a strong light
was used to illuminate the specimen to get better photo-
50
Density (Kg/m3)

graphs. The consequence of this action was that the tem-


45 perature increased from 21.5 1C to 28.5 1C and the humidity
dropped from 39% to 33% during this test. No obvious
40
variation of stress–strain curve was found for this test as
35 shown in Fig. 5(a).
30

25
3. Experimental results
20
55 60 65 70 75 80 85 90 3.1. Deformation features
Diameter of luffa sponge specimen (mm)
Typical force–displacement curves are given in Fig. 5(a) with
65 64.1 similar density. It shows clearly a fairly constant compressive
Core topology 2
60 Core topology 3 force over a long stroke, which represents an ideal energy
57.4
Core topology 4
55 54.1
50
Density (Kg/m3)

45

40

35 Humidity and Load cell


temperature
30 metre Top platen
28.5
25 27.3

20
0 5 10 15 20 25 30 35 40 45 50 55
Specimen number v0 Luffa specimen

Fig. 3 – Variation of density of luffa cylinders: (a) with Bottom platen


respect to equivalent diameter (no particular trend was
found); (b) with respect to core topology (no particular trend Fig. 4 – Experimental set-up of a quasi-static test of a luffa
was found). sponge column.
journal of the mechanical behavior of biomedical materials 15 (2012) 141 –152 145

Fig. 5 – Deformation features of luffa sponge columns crushed axially: (a) force–displacement curves of specimens of similar
density; (b) micro-truss level rotation and bending of luffa fibre. (Dotted lines represent the original fibre orientation, and the
dash dot lines in the right photograph represent rotated original fibres. It can be seen that the fibre has also bending
deformation.)

absorption feature. Generally, the deformation can be divided uniform in the plateau collapse region. Localised crushing
into three regions, namely, a rapid increasing elastic region, a band was observed which was similar to the static compres-
relative smooth plateau collapse region, and a densification sion behaviour of metallic foams described by Bastawros
region with sharp increase of force over displacement as et al. (2000) and Shen et al. (2010). At a smaller scale, luffa
shown in Fig. 5(a). The deformation process is shown in Fig. 6. fibres exhibited axial compression and tension as well as
Those photographs confirm that the overall compression micro-truss-level distortions and rotations which were simi-
strain of the specimen is axial along the longitudinal (load- lar to the open-cell metallic foams as shown in Fig. 5(b).
ing) direction rather than folding of the wall of the specimen. Nominal stress (defined as force over original cross sectional
The deformation is uniform in the elastic region. Similar to area) and nominal strain (defined as displacement over
metallic foams, the compressive deformation process for original thickness of the foam specimen) were calculated to
luffa sponge material under quasi-static loading was not get other mechanical properties.
146 journal of the mechanical behavior of biomedical materials 15 (2012) 141 –152

Fig. 6 – Typical deformation process of a luffa sponge material (Density: 52.3 kg/m3, Length: 51.3 mm). The deformation was
uniform until the deformation reach 3.5 mm, then localised crush band occurred at the middle part of the specimen.

3.2. Compressive strength where syf is the yield stress of the base material, rf is the
density of the base material, s0 and r0 are the compressive
Similar to other cellular materials (Gibson et al., 2010), the strength and density of the luffa sponge, A and B are two
compressive strength of the luffa sponge material is taken to constants determined by the topology and failure patterns of
be the initial peak stress if it exists. If there is no such peak cellular material. As mentioned previously, the base solid
stress, the stress at the intersection of two slopes is taken to material of luffa sponge and its mechanical properties
be the compressive strength. The two slopes are the slope for remained unknown at the present due to the lack of scientific
the initial loading and that for the stress plateau. It has been data on the mechanical properties of luffa sponge material.
found by other researchers (Ashby et al., 2000; Lu et al., 1999; The luffa sponge column specimen in our experiment was
McCullough et al., 1999; Miyoshi et al., 1999) that compressive composed of luffa fibre. Our preliminary tensile test results
strength of cellular materials obeys a power law with the on luffa fibres indicated that the density and tensile strength
relative density of luffa fibre varied greatly with different orientations in the
same luffa sponge and with different specimens for the same
!B orientation. A large quantity of tests should be conducted to
s0 r
¼A 0 ð1Þ get an accurate average values for mechanical properties of
syf rf
luffa fibres. Thus the empirical formulae are given in the
journal of the mechanical behavior of biomedical materials 15 (2012) 141 –152 147

Table 1 – Fitting parameters for empirical formulae of strength, densification strain and plateau stress.

Mechanical parameter A(C) error B(D) error Correlation coefficient Standard derivation

Strength 0.38 0.24 0.74 0.095


Plateau stress 0.18 0.11 0.89287 0.0587
Densification strain 0.02172 5.38  104 0.48172 0.02467

Linear fit (y ¼ AþBx) was used for all experimental data. Data for strength and plateau stress were converted to logarithmic scales before fitting.
‘‘A error’’ is the standard error for A.

following format 0.50


Experimental data
B 0.45
s0 ¼ Ar0 ð2Þ Empirical formula, Eq. (2)

Compressive strength (MPa)


0.40
The unknown base material properties are included in the
coefficients A and B. By fitting to Eq. (2) using the quasi-static 0.35
data in our experiment, the parameters A and B are 2.25  103 0.30
and 1.28, respectively if the unit of stress is Pa and that of
0.25
density is kg/m3. The fitting parameters are listed in Table 1.
0.20
For truss material with compression and stretching deforma-
tion pattern, an exponent of 1 is expected for plastic yielding 0.15
and 2 for elastic buckling (Mostafa and Damiano, 2010). For 0.10
open cell foams with bending dominant deformation pattern,
0.05
an exponent of 1.5 is usually found for plastic collapse
strength and 2 for elastic collapse strength (Gibson et al., 0.00
20 25 30 35 40 45 50 55 60 65 70
2010). A lower exponent indicates fewer penalties for
Density (kg/m3)
strength and energy absorption with decreasing density for
designing very light weight cellular materials. The exponent, Fig. 7 – Compressive strength of luffa sponge material. The
1.28 from our test results, compares favourably with that for square symbols represent the experimental data, the error
open-cell foams. But it is larger than that for truss materials. bars show the derivation of experimental data from their
However, for solid material based metallic foams, the expo- empirical values.
nent within the range of 1.5–3 is found in other experimental
work (Gibson et al., 2010). The value from our experiment Young's Modulus
12
suggests that the collapse of the luffa sponge is determined Empirical formula
by the coupling of axial yielding and bending of the luffa
10
fibre. The experimental results and empirical prediction are
Young's moduls (MPa)

shown in Fig. 7. It should be noted that the derivation of luffa


sponge material is large as shown in Fig. 7 and discussed in 8

Section 2.1. The scattering effect may also contribute to the


low exponent. If the bending is the dominant deformation 6
mechanism for luffa fibres, more experimental data will
increase the value of this exponent. On the other hand, if 4
the axial stretching and compressing of luffa fibre is the
dominant deformation mechanism, more experimental data 2
will decrease the value of this exponent.
0
20 25 30 35 40 45 50 55 60
3.3. Elastic modulus Density (kg/m3)

Similar to metallic foams (Ashby et al., 2000), the slope of the Fig. 8 – Young’s modulus of luffa sponge material. The
initial loading portion of the curve is lower than that of the square symbols represent the experimental data, the error
unloading curve. It indicates that there is localised plastic bars show the derivation of experimental data from their
deformation in the specimen at stress levels well below that empirical values.
of the compressive strength of the luffa sponge material
reducing the slope of the loading curve. However, the reload- formula is
ing curves after unloading roughly is close to the initial
E ¼ 1:48  106 r0 1:16 ð3Þ
loading curves. As a result, measurement of Young’s modulus
was made from the slope of the initial loading curve in the where E is the Young’s modulus of luffa sponge; the unit of
current study. Thus, the slope between 25% and 75% of the stress is Pa, and that of density is kg m3.
compressive strength is taken as the elastic modulus of luffa The experimental results and empirical predictions are
sponge material. Following the same argument for the shown in Fig. 8. Similarly to the compressive strength, the
compressive strength, for Young’s modulus, the empirical exponent, 1.16, is compared favourably with those from
148 journal of the mechanical behavior of biomedical materials 15 (2012) 141 –152

Gibson et al. (2010), 1.5. However, for solid material based


1.0
metallic foam, this exponent is within the range of 1–2 found
in other experimental work (Gibson et al., 2010). It should be 0.9
noted that the standard derivation of luffa sponge material is 0.8
fairly large as discussed in Section 2.1. This may also
0.7
contribute to the low exponent.

Stress (MPa)
0.6
0.5
3.4. Unloading and reloading features of luffa sponge
material 0.4
0.3
For any energy absorption devices, the bounce-back beha-
0.2
viour is critical to some of their applications. The bounce-
back behaviour is dominated by the stored elastic strain 0.1
during collapse (Lu and Yu, 2003). This feature can be 0.0
captured by the unloading and reloading curves for a speci- 0.0 0.2 0.4 0.6 0.8 1.0
men at different collapse strains. Typical unload–reload Strain
characteristics at different compressive strains are presented
35
in Fig. 9(a). The relatively high gradient of unloading curves
Density=41.1 kg/m
indicates the stored elastic energy is relatively small for luffa
30
sponge material. The unloading and reloading curves do not

Loading modulus (MPa)


overlap with each other. Due to the steep gradient the 25
dissipated energy in the hysteresis loop is very small com-
pared to the area under the entire stress–strain curve. Thus 20
luffa sponge can be used in those situations where a once-off,
15 Young's Modulus from non-cyclic loading
non-reversible energy absorption is required. The unloading 6.45MPa
curve is not linear and is steeper at the beginning of the
10
unloading. The reloading curve, on the other hand, is almost
linear. This phenomenon indicates the hysteresis of the luffa 5
sponge material at a strain rate of 103 s1. The hysteresis is
more prominent at larger compressive strain. For a given 0
temperature and humidity, it is observed that the unloading 0.0 0.2 0.4 0.6 0.8 1.0
and reloading curves intersect at exact unloading point as Unloading strain
shown in Fig. 9(a). The slope of reloading curve, reloading Fig. 9 – Unloading and reloading features of luffa sponge
modulus, varies with the strain level at the unloading point material: (a) stress–strain curves with 9 unloading points
as shown in Fig. 9(b). The reloading modulus is very similar to and the total unloading strain is about 5% (an enlarged view
the Young’s modulus measured in Section 3.3 at the plateau of unloading features in the elastic deformation region is
collapse region ranging from the first peak stress to the provided); (b) loading modulus in each of the unloading
densification strain. circle versus loading strain at the unloading point.

3.5. Energy efficiency method to determine densification


strain, plateau stress and specific energy absorption Then, the plateau stress spl is calculated as
R ed
sðeÞde
The compressive strength, s0, corresponds to the collapse of spl ¼ 0 ð6Þ
ed
the weakest layer of cells which was crushed first in com-
pression. Hence, this stress is not applicable to representing An example is shown in Fig. 10.
the energy dissipation of foams in compression. It is the The densification strain physically corresponds to the start
plateau stress, spl, which is closely related to the energy point from which the stress starts to rise sharply. From this
dissipation capacity of foams and its value is obtained using point onwards, the luffa sponge can still dissipate energy by
an energy efficiency method (Li et al., 2006). plastic deformation, but its dissipation efficiency will start to
The energy dissipation efficiency, Ed, of the foam is defined decrease. Based on the present experimental data, we can
as (Li et al., 2006) propose that the densification strain is linearly related to the
R ed density of the luffa sponge. Hence
sðeÞde
Ed ðea Þ ¼ 0 ; 0rea r1 ð4Þ ed ¼ CDr0 ð7Þ
sa
By fitting to Eq. (7) using the quasi-static data in our
and the densification strain, ed, is defined as the maximum
experiment, the parameters C and D can be obtained as
value of ei which satisfies the condition of the maximum
0.68 and 1.62  103, respectively. The unit of density is
efficiency
kg/m3. The fitting parameters are listed in Table 1. The
dEd ðea Þ empirical prediction and experimental data are shown in
9ea ¼ ei ¼ 0; 0rei r1 ð5Þ
de Fig. 11.
journal of the mechanical behavior of biomedical materials 15 (2012) 141 –152 149

Strain
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
1.0 0.60
Compressive Stress
0.9 0.54
Energy Efficiency
0.8 0.48
Compressive stress (MPa)
0.7 0.42
Plateau stress

Energy efficiency
0.6 0.31MPa 0.36

0.5 0.30
Compression strength
0.27MPa
0.4 0.24

0.3 0.18
Densification strain
0.2 0.57
0.12

0.1 0.06

0.0 0.00
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Strain

Fig. 10 – Illustration of energy efficiency method.

0.7 0.30
Energy absorption per unit volume (N/mm2)

Experimental data
Empirical formula
0.6
0.25
0.5
Densification strain

0.20
0.4

0.3
0.15

0.2 Experimental data for luffa sponge material


Empirical formula 0.10
0.1 Aluminium foam

0.0 0.05
0.06 0.08 0.10 0.12 0.14 0.16 0.18 20 25 30 35 40 45 50 55 60
Relative density Density (kg/m3)

Fig. 11 – Densification strain for luffa sponge specimens. Fig. 12 – Energy dissipation per unit initial volume versus
relative density. The square symbols represent the
experimental data, the error bars show the derivation of
By fitting to Eq. (2) using the quasi-static data for plateau experimental data from their empirical values.
stress in our experiment, the parameters A and B can be
obtained as 8.65  103 and 1.22, respectively if the unit of
stress is Pa and that of density is kg/m3. The fitting para- the density range in our experiment can be written as
meters are listed in Table 1. They are slight different from
w ¼ 8:65  103 ð0:68r0 1:22 1:62  103 r0 2:22 Þ ð8Þ
these for compressive strength. Similar to the compressive
strength of the luffa sponge, the exponent, 1.22, is compared Fig. 12 shows that the energy dissipation capacity per unit
favourably with those from Gibson et al. (2010), 1.5. However, volume relatively linearly increases with relative density in
for solid material based metallic foam, this figure should be the density range tested.
within the range of 1.5–3 found in other experimental work
(Gibson et al., 2010). Similar to strength of luffa sponge
material, the scattering effect may also contribute to the 4. Comparison with other materials
low exponent.
The energy dissipation capacity of luffa sponge material As a biological material, the experimental data in this paper
can be characterised by the energy absorption per unit initial about luffa sponge material may significantly expand the
volume, w, during the compression process before densifica- density range of previously studied natural cellular materials
tion occurs. This value is equal to the area under the (Gibson et al., 2010) with respect to the strength, as shown in
stress–strain curve. Thus the energy dissipation capacity for Fig. 13. This special feature provides an opportunity to widen
150 journal of the mechanical behavior of biomedical materials 15 (2012) 141 –152

1000 Natural cellular materials


such as trabecular bones, palms, and wood 8000

Energy dissipation per unit mass (J/kg)


Aporas® aluminium foam
Luffa sponge Cymate aluminium foam
100 (results from present study) 7000 Polymer foams
Strength (MPa)

Luffa sponge from present study


6000
10
5000
1 4000

3000
0.1
Other biological materials 2000
such as tomato, potato, and apple
0.01 1000
10 100 1000
Density (kg/m3) 0
0 50 100 150 200 250 300
Fig. 13 – Comparison between luffa sponge and other Density (kg/m3)
natural materials (Gibson et al., 2010).
1000
900 Luffa sponge (density: 44 ~ 47 kg/m )
the applications of natural cellular materials in general and
Nickel microlattice (density: 43 kg/m )
the luffa sponge in particular, especially when the light 800

Compressive stress (kPa)


weight is a key design requirement. 700
It is worth noting that human beings have been able to
600
discover and develop many materials, notably various metals,
500
metallic alloys, ceramics, plastics, as well as their compo-
sites, with performances significantly surpassing biological 400
materials (Bonderer et al., 2008; Zhang et al., 2011). When 300
light weight of the material is considered, various metallic/
200
ceramic/polymeric foams, honeycombs, and microlattices,
100
are designed and fabricated using those materials as base
materials. However, most of these man-made materials are 0
not environmentally friendly and have not been designed 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
with genuine concerns over their long-term sustainability Strain
(Zhang et al., 2011). An efficient solution is to find an
Fig. 14 – Comparison of energy absorption capacity per unit
alternative biological counterpart for those materials in use.
mass between luffa sponge and: (a) other man-made foams;
For this purpose, a comparative study has been carried out
(b) Ni–P microlattice (Schaedler et al., 2011).
below to show the merit of the luffa sponge material besides
its sustainable features.
As a cellular material, the luffa sponge is extremely light as
shown in Fig. 13. At less than 50 kg/m3, its density is about microstructures of luffa sponge material, we have also
one fifth of that of a typical aluminium foam (Alporass). For a compared the performance of the luffa sponge with other
similar density (25–65 kg/m3), very few materials currently cellular materials with similar density mentioned above.
exist such as Cymat aluminium foam (Z70 kg/m3), Ni–P Cymat foam has a strength of 40 kPa at a density of 70 kg m3
microlattices (Z0.9 kg/m3), silica aerogels (Z1.0 kg/m3), car- (Ashby et al., 2000). Aerogels exhibit a continuous increasing
bon nanotube aerogels (Z4.0 kg/m3), and polymer foams of stress over strain under compression which is not compar-
(Z8.0 kg/m3). However, our research results indicate that, able to luffa sponge material. Expended Polystyrene foams
when the luffa column is compressed longitudinally, the have a strength of 200 kPa at a density of 45 kg m3 (Avalle
amount of energy it can absorb per unit mass is comparable et al., 2001). The polyamide reinforced with modified poly-
to that of the aluminium foam (Idris et al., 2009; Shen et al., phenylene and polystyrene foam has a strength of 800 kPa at
2010) and most of the polymer foams (Avalle et al., 2001), as a density of 45 kg m3 (Avalle et al., 2001). As an important
shown in Fig. 14(a). It should be noted that Alporass alumi- example, the comparison with a recently invented metallic
nium foam has superior specific energy absorption than microlattice of a similar density (Schaedler et al., 2011) shows
other metallic foams when their density is lower than 300 kg that the average plateau stress of the luffa sponge (about
m3 (Ashby et al., 2000). The good specific energy absorption 350 kPa) is significantly higher than that of the Ni–P micro-
of luffa sponge is attributed partially to its light base material lattice (about 120 kPa), as shown in Fig. 14(b). It should be
as well as a higher densification strain. Due to the high noted that the strength decreases with density for any kind of
strength-to-weight ratio of cellular materials, luffa sponge cellular materials. These comparisons indicate that the luffa
can be used as a good packaging material and an excellent sponge material has better compressive strength than most
energy dissipation material. It should be seen that the of other available cellular material of similar density such as
average plateau stress of aluminium foam is 4–5 times higher Cymat aluminium foam, expended Polystyrene foams and
than the luffa sponge material. To show the efficiency of the Ni–P microlattices.
journal of the mechanical behavior of biomedical materials 15 (2012) 141 –152 151

It should be noted that the luffa sponge material is a As mentioned in Section 2.1, the luffa sponge used in our
cellular material with structural hierarchy as indicated in experiments was obtained from pharmacies in Australia
Fig. 1. As mentioned previously, the luffa sponge column which was sold as a bath sponge. There were two treatments
specimen is composed of luffa fibres. According to other in manufacturing the bath sponge from natural luffa fruits
research on the microstructures of luffa sponge material which may influence the mechanical properties of luffa
(Boynard and D’Almeida, 2000; Ghali et al., 2009; Laranjeira sponge, namely, bleaching and lateral squashing. Similar to
et al., 2006; Paglicawan et al., 2005; Tanobe et al., 2005), the other cellulose based natural fibres such as cotton, bleach
luffa fibres are hollow with many micro-tunnels. Those would result in slight loss of mass so as to its mechanical
microstructures have also been proven by our own investiga- strength of untreated luffa sponge. The lateral squashing
tions. Similar to other natural cellular material with similar would result in damage of the original luffa fibres. The
components, the cell walls of those micro-tunnels may be experimental data and empirical formula from our experi-
fibrous themselves and consist of oriented cellulose nanofi- ment will underestimate the strength of untreated luffa
brils in a hemicellulose and lignin matrix. Those hierarchical sponge material.
levels required further investigation. However, a rough esti-
mation can be made for contributions of topology at different
hierarchical levels to the strength, plateau stress and energy 5. Conclusions
absorption capacities of luffa sponge material using the
available material properties from others research In this study, a series of compressive tests were conducted to
(Paglicawan et al., 2005). The density is approximately examine the stiffness, strength and energy absorption char-
353 kg/m3 with a Young’s modulus of 1332 MPa (735%) and acteristics of the luffa sponge material under quasi-static
tensile strength of 11.1 MPa (785%). The obtained relative compressive load. The Young’s modulus, compressive
plateau stress versus relative density was compared with strength, densification strain, plateau stress and energy
other polymeric materials (Gibson and Ashby, 1982) as shown absorption capacity of the luffa sponge material have been
in Fig. 15. The experimental data and empirical prediction obtained for the first time. A set of empirical formulae have
exhibit a similar trend with that of other polymeric materials. been proposed to predict the mechanical properties of luffa
It indicates that the relative strength are not superior to sponge material on the macroscopic scale. It should be noted
others cellular material if only one level of structural hier- that there is a limitation of our current work, i.e., the luffa
archy is considered. The prediction formulae for other hier- sponge material used in our experiment was bleached and
archical levels are not available because there is a lack of the laterally squashed initially. The effect of such pre-treatments
mechanical properties of the base composite materials as on the mechanical properties of the luffa sponge is not
well as its specific composite patterns of those base ingre- available and is under further investigation. From our inves-
dients. It is worth further research effort to understand and tigation, the following conclusions can be drawn.
bio-mimic the luffa sponge material with superior strength
and energy absorption capacities. We also find that the (1) The stress–strain curves show a near constant plateau
severely crushed luffa sponge column (up to a nominal final stress over a long strain range, which is ideal for energy
compressive strain of 95%) is able to recover to its original absorption application. It could be used as an alternative
length and shape (up to 98%) after it is submerged in water packaging material.
and then dried. We are currently conducting further investi- (2) The exponent from our experiments suggests that the
gation on this very interesting phenomenon. deformation of the luffa sponge is determined by the
coupling of axial compression/tension and bending of
luffa fibres.
(3) It is found that the luffa sponge material exhibits remark-
able stiffness, strength and energy absorption capabilities
that are comparable to those of some metallic cellular
materials such as aluminium foams and Ni–P microlat-
tices. The strength of luffa sponge is better than most of
other available cellular materials in the similar density
range such as expended Polystyrene foams and Ni–P
microlattices. As an ultra-light cellular material, it has
great potential to be used as an environmentally friendly
engineering material.

r e f e r e nc e

Fig. 15 – Comparison of relative plateau stress between luffa Ajiwe, V., Ndukwe, G., Anyadiegwu, I., 2005. Vegetable diesel fuels
sponge and other polymeric materials (Gibson and Ashby, from Luffa cylindrica oil, its methylester and ester-diesel
1982). blends. Chemistry Class Journal 2, 1–4.
152 journal of the mechanical behavior of biomedical materials 15 (2012) 141 –152

Ashby, M.F., Evans, A.G., Fleck, N.A., Gibson, L.J., Hutchinson, J.W., Mazali, I.O., Alves, O.L., 2005. Morphosynthesis: high fidelity
Wadley, H.N.G., 2000. Metal Foams: A Design Guide. Butter- inorganic replica of the fibrous network of loofa sponge (Luffa
worth-Heinemann, Warrendale. cylindrica). Anais da Academia Brasileira de Ciências 77, 25–31.
Avalle, M., Belingardi, G., Montanini, R., 2001. Characterization of McCullough, K.Y.G., Fleck, N.A., Ashby, M.F., 1999. Uniaxial
polymeric structural foams under compressive impact loading stress–strain behaviour of aluminium alloy foams. Acta Bio-
by means of energy-absorption diagram. International Journal materialia 47, 2323–2330.
of Impact Engineering 25, 455–472. Meyers, M.A., Chen, P.-Y., Lopez, M.I., Seki, Y., Lin, A.Y.M., 2011.
Bal, K.E., Bal, Y., 2004. Gross morphology and absorption capacity Biological materials: a materials science approach. Journal of
of cell-fibers from the fibrous vascular system of loofah (Luffa the Mechanical Behavior of Biomedical Materials 4, 626–657.
cylindrica). Textile Research Journal 74, 241–247. Miyoshi, T., Itoh, M., Mukai, T., Kanahashi, H., Kohzu, H., Tanabe,
Bastawros, A.F., Bart-Simth, H., Evans, A.G., 2000. Experimental S., Higashi, K., 1999. Enhancement of energy absorption in a
analysis of deformation mechanisms in a closed-cell alumi- closed-cell aluminum by the modification of cellular struc-
nium alloy foam. Journal of the Mechanics and Physics of tures. Scripta Materialia 41, 1055–1060.
Solids 48, 301–322. Mostafa, S.A.E., Damiano, P., 2010. Multiscale structural design of
Bonderer, L.J., Studart, A.R., Gauckler, L.J., 2008. Bioinspired columns made of regular octet-truss lattice material. Interna-
design and assembly of platelet reinforced polymer films. tional Journal of Solids and Structures 47, 1764–1774.
Science 319, 1069–1073. Oboh, I.O., Aluyor, E.O., 2009. Luffa cylindrica—an emerging cash
Boynard, C.A., D’Almeida, J.R.M., 2000. Morphological character- crop. African Journal of Agricultural Research 4, 684–688.
ization and mechanical behavior of sponge gourd (Luffa Oboh, I.O., Aluyor, E.O., Audu, T.O.K., 2011. Application of luffa
cylindrica)-polyster composite materials. Polymer-Plastics cylindrica in natural form as biosorbent to removal of divalent
Technology and Engineering 39, 489–499. metals from aqueous solutions—kinetic and equilibrium
Chen, J.P., Lin, T.C., 2005. Loofa sponge as a scaffold for culture of study. In: Einschlag, F.S.G. (Ed.), Waste Water—Treatment
rat hepatocytes. Biotechnology Progress 21, 315–319. and Reutilization. InTech, pp. 195–212.
Damiens, R., Rhee, H., Hwang, Y., Park, S.J., Hammi, Y., Lim, H., Paglicawan, M.A., Cabillon, M.S., Cerbito, R.P., Santos, E.O., 2005.
Horstemeyer, M.F., 2012. Compressive behavior of a turtle’s Loofah fiber as reinforcement material for composite. Philip-
shell: experiment, modeling, and simulation. Journal of the pine Journal of Science 134, 113–120.
Mechanical Behavior of Biomedical Materials 6, 106–112. Roble, N., Ogbonna, J., Tanaka, H., 2002. A novel circulating loop
Demir, H., Atikler, U., Balköse, D., Tihminlioglu, F., 2006. The effect bioreactor with cells immobilized in loofa (Luffa cylindrica)
of fiber surface treatments on the tensile and water sorption sponge for the bioconversion of raw cassava starch to ethanol.
properties of polypropylene-luffa fiber composites. Compo- Applied Microbiology and Biotechnology 60, 671–678.
sites Part A 37, 447–456. Schaedler, T.A., Jacobsen, A.J., Torrents, A., Sorensen, A.E., Lian, J.,
Ghali, L., Msahli, S., Zidi, M., Sakli, F., 2009. Effect of pre-treatment Greer, J.R., Valdevit, L., Carter, W.B., 2011. Ultralight metallic
of luffa fibres on the structural properties. Materials Letters microlattices. Science 334, 962–965.
63, 61–63. Shen, J., Lu, G., Ruan, D., 2010. Compressive behaviour of closed-
Gibson, L.J., Ashby, M.F., 1982. The mechanics of three-dimen- cell aluminium foams at high strain rates. Composites Part B-
sional cellular materials. Proceedings of the Royal Society of Engineering 41, 678–685.
London Series A 382, 43–59. Siqueira, G., Bras, J., Dufresne, A., 2010. Luffa cylindrica as a
Gibson, L.J., Ashby, M.F., Harley, B.A., 2010. Cellular Materials in lignocellulosic source of fiber, microfibrillated cellulose, and
Nature and Medicine. Cambridge University Press, Cambridge. cellulose nanocrystals. BioResources 5, 727–740.
Idris, M.I., Vodenitcharova, T., Hoffman, M., 2009. Mechanical Stanzl-Tschegg, S.E., Keunecke, D., Tschegg, E.K., 2011. Fracture
behaviour and energy absorption of closed-cell aluminium tolerance of reaction wood (yew and spruce wood in the TR
foam panels in uniaxial compression. Materials Science and crack propagation system). Journal of the Mechanical Beha-
Engineering A 517, 37–45. vior of Biomedical Materials 4, 688–698.
John, M.J., Thomas, S., 2008. Biofibres and biocomposites. Carbo- Tanobe, V.O.A., Sydenstricker, T.H.D., Munaro, M., Amico, S.C.,
hydrate Polymers 71, 343–364. 2005. A comprehensive characterization of chemically treated
Johnsona, M., Walterb, S.L., Flinnc, B.D., Mayerc, G., 2010. Influ- Brazilian sponge-gourds (Luffa cylindrica). Polymer Testing 24,
ence of moisture on the mechanical behavior of a natural 474–482.
composite. Acta Biomaterialia 6, 2181–2188. Tavares, J., Israel, N., Rui, O., Wilton, S., Valderi, D., 2008. Nitrifica-
Laidani, Y., Hanini, S., Henini, G., 2001. Use of fiber luffa cylindrica tion in a submerged attached growth bioreactor using Luffa
for waters traitement charged in copper. Study of the possi- cylindrica as solid substrate. African Journal of Biotechnology
bility of its regeneration by desorption chemical. Energy 7, 2702–2706.
Procedia 6, 381–388. Wegst, U.G.K., 2011. Bending efficiency through property gradi-
Laranjeira, E., Carvalho, L.H.d., Silva, S.M.d.L., D’Almeida, J.R.M., ents in bamboo, palm, and wood-based composites. Journal of
2006. Influence of fiber orientation on the mechanical proper- the Mechanical Behavior of Biomedical Materials 4, 744–755.
ties of polyester/jute composites. Journal of Reinforced Plas- Zampieri, A., Mabande, G.T.P., Selvam, T., Schwieger, W., Rudolph,
tics and Composites 25, 1269–1278. A., Hermann, R., Sieber, H., Greil, P., 2006. Biotemplating of
Li, Q.M., Magkiriadis, I., Harrigan, J.J., 2006. Compressive strain at Luffa cylindrica sponges to self-supporting hierarchical zeolite
the onset of the densification of cellular solids. Journal of macrostructures for bio-inspired structured catalytic reactors.
Cellular Plastics 42, 371–392. Materials Science and Engineering C 26, 130–135.
Lu, G., Lu, G.Q., Xiao, Z.M., 1999. Mechanical properties of porous Zhang, Z., Zhang, Y., Gao, H., 2011. On optimal hierarchy of load-
materials. Journal of Porous Materials 6, 359–368. bearing biological materials. Proceedings of the Royal Society
Lu, G., Yu, T.X., 2003. Energy Absorption of Structures and B 278, 519–525.
Materials. Woodhead Publishing Ltd, Cambridge. Zhu, J., Wang, X., He, C., Wang, H., 2012. Mechanical properties,
Mayer, G., 2011. New toughening concepts for ceramic compo- anisotropic swelling behaviours and structures of jellyfish
sites from rigid natural materials. Journal of the Mechanical mesogloea. Journal of the Mechanical Behavior of Biomedical
Behavior of Biomedical Materials 4, 670–681. Materials 6, 63–73.

You might also like