Sunteți pe pagina 1din 14

Original

Paper

phys. stat. sol. (b) 242, No. 3, 681 – 694 (2005) / DOI 10.1002/pssb.200460386

Auxetic compliant flexible PU foams:


static and dynamic properties
F. Scarpa*, 1, P. Pastorino1, A. Garelli1, S. Patsias2, and M. Ruzzene3
1
Multidomain Cellular Solids Laboratory, Department of Mechanical Engineering,
University of Sheffield, S1 3JD Sheffield, UK
2
Department of Mechanical Engineering, University of Sheffield, S1 3JD Sheffield, UK
3
School of Aerospace Engineering, Georgia Institute of Technology, 270 Ferst Drive, Atlanta,
GA 30332-0150, USA

Received 29 April 2004, accepted 29 November 2004


Published online 15 February 2005

PACS 61.41.+e, 81.05.Lg


The paper describes the manufacturing and tensile testing of auxetic (negative Poisson’s ratio) thermo-
plastic polyurethane foams, both under constant strain rate and sinusoidal excitation. The foams are pro-
duced from conventional flexible polyurethane basis following a manufacturing route developed in previ-
ous works. The Poisson’s ratio behaviour over tensile strain has been analyzed using an Image Data pro-
cessing technique based on Edge Detection from digital images recorded during quasi-static tensile test.
The samples have been subjected to tensile and compressive tests at quasi-static and constant strain-rate
–1
values (up to 12 s ). Analogous tests have been performed over iso-volumetric foams samples, i.e., foams
subjected to the same volumetric compression of the auxetic ones, exhibiting a near zero Poisson’s ratio
behaviour. The auxetic and non-auxetic foams have been also tested under sinusoidal cycling load up to
10 Hz, with maximum pre-strain applied of 12%. The hysteresis of the cycling loading curve has been
measured to determine the damping hysteretic loss factor for the various foams. The measurements indi-
cate that auxetic foams have increased damping loss factor of 20% compared to the conventional foams.
The energy dissipation is particularly relevant in the tensile segment of the curve, with effects given by
the pre-strain level imposed on the samples.

© 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1 Introduction
Since 1987, when isotropic auxetic foam was manufactured for the first time [1], negative Poisson’s ratio
materials have created some interest for potential applications in structural integrity compliant structures
and sandwich components. By definition, an auxetic (or negative Poisson’s ratio) material expands in all
directions when pulled in only one, giving therefore a deformation kinematics opposite to the one of
“conventional” materials. This behaviour does not contradict the classical theory of elasticity: a homoge-
neous isotropic thermodynamically correct 3D solid has a potential Poisson’s ratio range between –1.0
and 0.5, while anisotropic solids can have also larger values in magnitude [2]. A negative Poisson’s ratio
coefficient for a material could lead to increase in indentation resistance [3], enhanced bending stiffness
in structural elements and shear resistance [4], optimal passive tuning of structural vibration [5] and
enhanced dielectric properties for microwave absorbers [6].

*
Corresponding author: e-mail: f.scarpa@shef.ac.uk, Fax: +441142227890

© 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


682 F. Scarpa et al.: Auxetic compliant flexible PU foams: static and dynamic properties

The dynamic behaviour of auxetic materials has been investigated by several authors [7– 12]. The
possible use of auxetic materials for viscoelastic damping applications has been examined in [10, 13],
where biphasic auxetic composite showed loss tangent exceeding lower Voigt limit and close to Hashin
upper bound. In a recent paper [17] one of the authors showed an increase in loss factor for an auxetic
PU foam manufactured using a modified route of the classical production layout. In this paper the au-
thors focus on specific auxetic foam manufactured following a particular process derived by the one
proposed in a seminal paper by Chan and Evans [16]. The foam exhibits remarkable resilience under
dynamic impact loading at high constant strain rates compared to the conventional foam used as a start-
ing base. The higher strain rate impact characteristics have been detected using a MTS tensile machine.
The rationale for these experiments is that foam cores in sandwich structures are used in packaging solu-
tions and energy absorption devices [17], and that the knowledge of dynamic loading behaviour is essen-
tial to determine the structural integrity of packages. All tests have been carried out also on specimens of
conventional foam used to manufacture the auxetic samples, in order to provide an indication on how the
manufacturing process affected the overall properties of the foam. To benchmark further the properties
of the auxetic solids, iso-volumetric foam samples have been manufactured and tested. The iso-
volumeric foams feature the overall volumetric compression ratio of the auxetic foams, with almost zero
values of the Poisson’s ratio. The tests indicated higher resilience of the auxetic foams under increasing
constant strain rate loading compared both conventional and iso-volumetric foams. In particular, the
comparison with the conventional foam highlighted a remarkable improvement in terms of crashworthi-
ness resistance. The auxetic foams showed also quasi-constant stress-strain loading pattern before densi-
fication, showing no significant sensitivity to different strain-rate values of loading.
The foam samples have been also tested under sinusoidal cyclic deformation with the same tensile test
machine for frequency up to 10 Hz, and different levels of pre-strain and amplitude. The hysteresis cy-
cles have been measured and equivalent storage and loss modulus of the foams been detected at room
temperature. The approach taken is similar to the one used by Gandhi and Wolons [18] to measure the
complex modulus of the pseudo-elastic properties of shape memory alloy Nitinol wires. This approach
has been deemed appropriate considering the significant dependence of the viscoelastic foam properties
versus the excitation frequency and different levels of pre-strain for fixed temperature conditions [11,
17].

2 Foams manufacturing
The specimens tested were obtained from conventional grey open-cells polyurethane foams with
30 –35 pores/inch and 0.0027 g/cm3 density supplied by McMaster-Carr Co., Chicago, IL. The conver-
sion process was obtained using a purpose made aluminium mould composed by cylinders backed by

Fig. 1 (online colour at: www.pss-b.com) Alumin-


ium mould used to manufacture the auxetic and iso-
volumetric foams.

© 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


Original
Paper

phys. stat. sol. (b) 242, No. 3 (2005) / www.pss-b.com 683

pistons to impose axial compression on the samples (Fig. 1). The conversion process applied to the
foams involved the imposition of a combined axial and radial compression [1] (with overall volumetric
compression ranging from 68% to 94%), followed by heating until the softening temperature is reached
[1, 2, 16, 17]. The base foam was supplied in squared blocks of 600 mm side by 50 mm thickness.
Cylindrical pieces of foams have been obtained using a sharp-edge tube penetrated inside the block on
conventional foam. The ends of the foam cylinders were then flattened using a cutter. Three different
tubes having 19 mm, 30 mm and 50 mm diameter respectively were used to impose different radial com-
pression values on the mould cylinders at the same time. The internal wall of the tube was then lubri-
cated with vegetable (olive) oil. It has to be pointed out that using distilled oils or oil-derivatives is
strongly not recommended because of their instability at high temperatures and unpleasant smells which
could follow [2, 16, 17]. The pre-cut foams were then inserted inside the tube with a wire in order to pull
the foam, rather than pushing it in the mould. In this way one could avoid the formation of unwanted
creases or torques effects. A conical inlet was found useful to avoid creases, especially for the specimens
with the larger initial diameter. The tube containing the foam was then positioned in the mould frame
and the piston pushed inside it. This operation was repeated turning the tube upside down in order to
slide the foam inside the tube, obtaining in this way uniform compression along the specimen. The lock-
ing of the piston in the desired position followed compression. The whole operation was then repeated
for the other specimens of the batch. The mould was then put into an industrial oven, with most of the
try-batches were made in order to find out the optimal time-temperature profile. The heating process was
followed by a room temperature cooling, which lasted about two hours on average. After this operation,
the specimens extracted were gently tensioned in order to relax the external surface. The stiffer ends of
the specimens were cut and the dimensions and weights were measured to calculate their densities. This
last operation should be made as soon as possible, since the specimens expand over a prolonged period
of time, leading to appreciable different density values if calculated after few days. Twelve try-batches
were made in order to identify for the specific base-foam the manufacturing process parameters, like
volumetric compression ratio, and find out the optimal time-temperature profiles. The general designa-
tion of the samples was built on a “XY” code, where “X” is the position of the specimen in the mould
and “Y” is the progressive letter assigned to every batch. Samples with higher uniformity of manufactur-
ing and mechanical properties belonged to the batches C and L (parameters shown in Table 1 and Fig. 2).

Table 1 Radial and axial compressions for the various foam batches.

bath sample initial size [mm] imposed size [mm] compression ratio [%] heating law
diameter length diameter length radial axial volume
C 1C 30 200 19 85 36.7 57.5 83.0 time–
2C 30 180 19 85 36.7 52.8 81.1 temperature
3C 30 160 19 85 36.7 46.9 78.7 profile α
4C 30 140 19 85 36.7 39.3 75.6
5C 30 120 19 85 36.7 29.2 71.6
L 1L 48 180 19 80 62.0 55.6 93.6 time–
2L 48 160 19 80 62.0 50.0 92.8 temperature
3L 48 140 19 80 62.0 42.9 91.7 profile β
4L 48 120 19 80 62.0 33.3 90.4
5L 48 100 19 80 62.0 20.0 88.4
M 1M 19 220 19 40 0 81.8 81.8 time–
2M 19 210 19 40 0 81.0 81.0 temperature
3M 19 200 19 40 0 80.0 80.0 profile β
4M 19 200 19 50 0 75.0 75.0
5M 19 160 19 50 0 68.8 68.8

© 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


684 F. Scarpa et al.: Auxetic compliant flexible PU foams: static and dynamic properties

a) Time-temperature profile α

160

140

120
Temperature [ºC]

100

80

60

40

20

0
0 5 10 15 20 25
Time [min]

b) Time-temperature profile β

160

140

120
Temperature [ºC]

100

80

60

40

20

0
0 2 4 6 8 10 12 14 16 18 20
Time [min]

Fig. 2 Temperature profiles (a) α ; (b) β.

For comparison, specimens with no imposed radial compression were also tested; these latter manufac-
turing parameters were labelled as “M” batches. The tested specimens were designed with a “XYZ”
code, in which “Z” represents the progressive number of the “Y” batch. These particular specimens were
produced to verify if the peculiar mechanical properties of auxetic foams were derived mainly from den-
sity increase. As it will be clear in the following paragraphs, the density increase plays only a marginal
role.

3 Testing techniques
The densities of the specimens were obtained measuring their dimensions with a digital calliper and their
weights with an electronic balance. The tests were conducted at room temperature (20 °C) at average
70% of humidity. The humidity level was not controlled during testing. Since the specimens start to
expand after the manufacturing process, it has been necessary to measure the real density reached avoid-
ing letting more than one day between the density measure and the experimental test. In fact, differences
from the imposed volumetric ratio and the real volumetric ratio achieved were noticed to be up to 52%,
and, for the same specimen, the difference in density could be up to 30% after a week. The dimensions
of the specimens were acquired with a Kennedy digital calliper (sensitivity ±0.01 mm). Since the foams

© 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


Original
Paper

phys. stat. sol. (b) 242, No. 3 (2005) / www.pss-b.com 685

tested were compliant, the acquisition process involved extreme handling care. The procedure used con-
sisted on closing the ends of the calliper at very low speed, meanwhile moving the calliper in a direction
perpendicular to the dimension measured; as soon as the specimen was pinched by the calliper the meas-
ure was reported and mediate with others acquired in different positions. A weight measure using an
electronic balance (sensitivity ±0.01 g) followed, and then the values obtained were used to calculate the
density. It has to be noticed that even if the different batches were made using the same parameters, the
densities of corresponding specimens were not the exactly same, due to a non perfect repetitiveness of
the manufacturing process.
Poisson’s ratio measures were based on image data acquired with a SONY DCR-TRV50E digital
camera and processed with the software MATLAB v6.1. The images were acquired during quasi-static
tests performed with a MTS 858 servo-hydraulic machine. Both static test and the high strain rate test
were performed using the MTS servo-hydraulic machine. The Poisson ratio measures were carried out
cutting each specimen in 40 mm length and gluing it to the end clamp of the machine with Loctite® prod-
uct, then stretched until breaking or ungluing at 0.1 mm/s (e = 0.0025 s -1 ).
Being the image acquisition affected by perspective effect it has been necessary to take a calibration
photo for each test, in order to have the possibility to convert the measures from pixels to millimetres.
After calibration, a set of photos was acquired to measure longitudinal and transverse strain during test-
ing. This operation has been performed using a MATLAB routine, which calculates for every photo ten
values of diameters and ten values of length along the specimen, and gives as output the mean value and
the standard deviation. Once these data are known it is possible to plot a Poisson’s ratio versus longitu-
dinal strain graph. It must be reminded that the specimens, being glued to the end clamps, suffered from
end effect. This problem has been solved painting in white the extreme thirds of the dark-grey specimens
and thus considering only the central third in the calculation of the lengths to avoid Saint-Venant effects.
The perspective effect in the image capture could have been avoided if the images were recorded through
a telescope placed at a long distance from the specimens. This is was not available at the time and hence
the procedure suggested in [20] was implemented. A ruler was placed at a set distance from the camera
iris and the scaling factor was obtained by converting the pixels in the image to mm taking into account
the distance of the camera to the specimen and the nominal focal length.
The MTS 858 servo-hydraulic system used includes the load unit with cross-head-mounted actuator, a
hydraulic power unit and a test controller with at least two channels of control (force and displacement).
The load unit is fatigue-rated at 25 kN, and can be operated at frequencies up to 30 Hz, with an integral
crosshead-mounted linear actuator with an attached manifold. The MTS machine has been used to carry
out tensile test (Poisson’s ratio measures) on batches C4, L2 and M1, quasi-static compression tests on
batches C5, L1 and M2, high strain-rate tests on batches C8, L5 and M3 and for testing conventional
foam specimens.
Quasi-static compression tests were carried out on batches C5, L1 and M1. A gap of 40 mm was left
between the clamp ends of the machine, and samples cut in 40 mm length were placed between them,
lubricating the contact surfaces to minimize friction. The speed of the head was 0.1 mm/s, indicating a
strain-rate of 0.0025 s–1, with maximum compressive strain reached at 75%. Data were acquired in dis-
placement-crossing mode, acquiring displacement and force every 0.1 mm. With the knowledge of the
initial sizes of the specimens it was then possible to plot a strain-stress curve for every case. The sinusoi-
dal excitation tests were carried out on batches M4, L3 and C6, with pre-deformation levels up to
10 mm, and amplitudes increasing at steps of 1 mm up to 5 mm. The tests were carried out over samples
coming from several manufactured batches to avoid past loading histories and residual stress conditions.
The data (force and displacements) were acquired at 0.200 ms of sampling for frequencies up to 10 Hz.
The tests were carried out at average temperature of 20 °C and relative humidity of 70%. High-strain rate
tests were carried out on batches C8, L5 and M3 using the MTS servo-hydraulic machine. This type of
test strongly modifies the mechanical characteristics of the specimens. For this reason every specimen
should not be tested more than once. Since the maximum speed reachable with the machine was
200 mm/s, it has been necessary to decrease the size of the specimens to obtain strain-rates up to 12 s–1.

© 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


686 F. Scarpa et al.: Auxetic compliant flexible PU foams: static and dynamic properties

Final volumetric ratio versus imposed volumetric ratio


14

12

10
Final volumetric ratio

C
8
L
M
6
Best fit

0
0 5 10 15
Imposed volumetric ratio

Fig. 3 Variation of effective volumetric ratio vs. imposed one.

A collateral advantage of this procedure was the testing at different strain-rates on the same batch sam-
ple, because it was possible to cut three specimens from each one, and then testing a new specimen for
each strain-rate. However, reducing the length of the specimen decreased also the densification strain of
the material. Strain-rates of respectively 8, 10 and 12 s–1 were reached cutting the specimens in 15 mm
length and moving the head of the machine at 120, 150 and 180 mm/s. The maximum strain reached in
this way was 80%. As for the quasi static-tests, the end clamps of the machine were positioned leaving a
15 mm gap. The specimens were then inserted lubricating the contact surfaces, and the test performed
acquiring displacement and force every 0.163 ms.

4 Experimental results
4.1 Density
Knowing the density of the base foam and the volumetric compression ratio imposed it is possible to
calculate for every specimen the final volumetric ratio as the ratio Rv between the final density and base
foam density.
Using the data acquired it has been possible to estimate the mean value and the standard deviation for
each set of manufacturing parameters. It is worth noticing that the final volumetric ratio versus imposed
volumetric ratio curve (Fig. 3) related to batches C and L shows a quasi-linear trend. The line fitting the
experimental results has the following equation:

Rv = 0.6 ◊ Rv 0 + 1.4 (1)

where Rv0 is the volumetric ratio imposed during the manufacturing process.
However, specimens made only with axial compression (i.e., batches belonging to the M series) do not
show this specific behaviour, having fitting line with equation:

Rv = 0.23 ◊ Rv 0 + 2.0 . (2)

4.2 Poisson’s ratio


The initial length and diameter of each specimen have been measured on the images representing the
unloaded specimen. Via the Image Data detection technique the variations of length and diameter have

© 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


Original
Paper

phys. stat. sol. (b) 242, No. 3 (2005) / www.pss-b.com 687

Poisson's ratio specimens C4


0.05

0.00

-0.05
Poisson's ratio

1C4
2C4
-0.10 3C4
4C4
5C4
-0.15

-0.20

-0.25
0.00 0.20 0.40 0.60 0.80 1.00 1.20 1.40 1.60
Tensile strain

Fig. 4 Poisson’s ratio vs. tensile strain for samples belonging to batch C4.

been detected to calculate the longitudinal and the transverse strains and the Poisson’s ratio defined
as:
ey
n xy = - . (3)
ex

Figures 4 – 6 show the behaviour of the Poisson’s ratio vs. tensile strain for the different foam batches.
Specimens C4 and L2 (which received both axial and radial compression) show a quasi-parabolic be-
haviour in most cases, while specimens M1 show a positive trend. All of them present auxetic behav-
iour in a wide range of tensile strain. It interesting to notice that the minimum values measured for
the L2 specimens are reached at lower strain values than the C4 specimens. Since the specimens
M1 received only axial compression during the manufacturing process, auxetic behaviour was not ex-
pected. However, this did not happen, probably because the basic foam (which has positive Poisson’s
ratio) induces a small amount of radial compression during the axial compression phase. For these
batches, no specific correlations between final volumetric compression and Poisson’s ratio have been
found. It has to be noticed that the Poisson’s ratio for the conventional foam specimens measured with
the same technique showed an almost constant value of 0.4 between 10% and 50% of tensile strain load-
ing.

Poisson's ratio specimens L2


0.10

0.05

0.00
Poisson's ratio

-0.05 1L2
2L2
-0.10 3L2
4L2
-0.15 5L2

-0.20

-0.25

-0.30
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90
Tensile strain

Fig. 5 Poisson’s ratio vs. tensile strain for samples belonging to batch L2.

© 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


688 F. Scarpa et al.: Auxetic compliant flexible PU foams: static and dynamic properties

Poisson's ratio specimens M1


0.00

-0.02

-0.04
Poisson's ratio

1M1
2M1
-0.06 3M1
4M1
5M1
-0.08

-0.10

-0.12
0.00 0.20 0.40 0.60 0.80 1.00 1.20
Tensile strain

Fig. 6 Poisson’s ratio vs. tensile strain for samples belonging to batch M1.

4.3 Quasi-static tensile tests


The aim of the quasi-static tests was to identify a strain-stress curve for each specimen and study the
variation due to the different manufacturing parameters, with comparison with the conventional foam. It
is worth of notice that all manufactured foams have mechanical characteristics at least one order of mag-
nitude higher than the base foam. For example, at 70% strain the base foam exhibits a stress of 23 kPa,
while specimens like 1L1 and 1C5 show 1.3 MPa and 547 kPa respectively. It is also noticeable that the
mechanical strengths of the specimens tested increase with the volumetric compression reached after the
manufacturing process in most cases. Another important consideration relates to the fact that the foams
tested, unlike conventional foams, do not present the classical elastic-plateau region, but feature an
almost continuous linear slope of the compressive stress-strain pattern, like other auxetic foams de-
scribed in literature [4, 15, 16]. This means that not only the manufacturing process leads to an increase
of the mechanical characteristics, but also changes the overall stress-strain pattern of the foam. Further-
more, as already mentioned in [1, 17], the elastic modulus of the foam which received both axial and
radial compression (batches C and L) decreases with the final volumetric compression ratio reached after
the manufacturing process. Figures 7 and 8 show the comparison of the stress-strain patterns vs. the
conventional foam for L1 and C1 batches respectively.

Fig. 7 (online colour at: www.pss-b.com) Quasi-static stress-strain curves for specimens L1 and conven-
tional foam.

© 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


Original
Paper

phys. stat. sol. (b) 242, No. 3 (2005) / www.pss-b.com 689

Fig. 8 (online colour at: www.pss-b.com) Quasi-static stress-strain curves for specimens C1 and con-
ventional foam.

4.4 Harmonic loading


Using the complex modulus approach [18], the constitutive relation between tensile stress σ and strain ε
for a damping material subjected to steady-state sinusoidal (harmonic) excitation is given by:
s = (G ¢ + iG ¢¢ ) e , (4)
where G′ and G″ are the storage and loss modulus respectively. Imposing a harmonic strain input
e = e 0 cos w t and corresponding measured stress one obtains following [18]:
2p w
1 w
G¢ = Ú s (t ) cos w t ◊ dt , (5)
e0 p 0

2p w
1 w
G ¢¢ =
e0 p Ú
0
s (t )sin w t ◊ dt . (6)

The typical hysteresis-loop cycle diagram of the damping material will have an internal area equal to
pG ¢¢e 02 . Equations (5 – 6) have been used to extract the frequency and strain-dependent storage and loss
1L3 (ν=-0.26)
3.50

3.00
Storage modulus [MPa]

2.50

p5a2
2.00
p5a5
p10a2
1.50
p10a5

1.00

0.50

0.00
0 1 2 3 4 5 6 7 8 9 10
Frequency [Hz]

Fig. 9 Storage modulus vs. frequency for samples 1L3 at different levels of pre-deformation and peak
amplitude.

© 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


690 F. Scarpa et al.: Auxetic compliant flexible PU foams: static and dynamic properties

Conventional PU foam
0.09

0.08

0.07
Storag e modulus [MPa]

0.06

0.05 p10a2
0.04 p10a5

0.03

0.02

0.01

0.00
0 1 2 3 4 5 6 7 8 9 10
Frequency [Hz]

Fig. 10 Storage modulus vs. frequency for samples of base foam at different levels of pre-deformation
and peak amplitude.

modulus of the foam by numerical integration (trapezoidal rule) of the stress-strain time histories of the
measured curves.
Figure 9 shows the behaviour of specimens 1L3 (Poisson’s ratio of –0.26 at 40% of tensile strain)
versus frequency for various levels of pre-deformation and amplitude. A pre-deformation of 5 mm and
oscillation amplitude of 2 mm gives rise to average storage modulus values around 500 kPa significantly
constant over the frequency range considered. Increasing the corresponding amplitude to 5 mm leads to
an increase of the storage modulus in magnitude of 70% on average, always with a negligible variation
over frequency. Increasing the pre-deformation level to 10 mm and amplitude 2 mm leads to a remark-
able increase in storage modulus value, with a magnitude of 2.8 MPa at 6 Hz. Also in this case, it ap-
pears only a slight dependence over the frequency domain considered. It is worth noticing that increasing
the amplitude to 5 mm for the latter case, the storage modulus is slightly decreased, as opposed to the
previous case, probably for relaxation of the tested sample. The significant increase in storage modulus
compared with the conventional foam is evident from Fig. 10, where the storage modulus of the conven-
tional foam for pre-deformation of 10 mm and amplitudes of 2 mm and 5 mm respectively are plotted.
The maximum values recorded are, for these cases, 70 kPa. Figure 11 shows the comparison in terms of
loss factor (ratio between G″ and G′) for the conventional and auxetic foam 1L3. The base foam exhibits
higher loss factor, with less regular dependence over frequency, while the auxetic one shows loss factors

Conventional PU foam and 1L3 (ν=-0.26)


0.40

0.35

0.30

0.25
Loss factor

conv p10a2
conv p10a5
0.20
1L p10a2
0.15 1L p10a5

0.10

0.05

0.00
0 1 2 3 4 5 6 7 8 9 10
Frequency [Hz]

Fig. 11 Comparison between loss factors for auxetic (1L3) and conventional foam samples at 10 mm of
pre-deformation and different peak amplitudes vs. frequency.

© 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


Original
Paper

phys. stat. sol. (b) 242, No. 3 (2005) / www.pss-b.com 691

1M4 (ν=-0.067)
0.50
0.45
0.40
Storage modulus [MPa]

0.35
0.30 p5a2
p5a5
0.25
p10a2
0.20 p10a5
0.15
0.10
0.05
0.00
0 2 4 6 8 10
Frequency [Hz]

Fig. 12 Storage modulus vs. frequency for iso-volumetric sample (1M4) at different levels of pre-
deformation and peak amplitude.

substantial constant at 0.15 for amplitudes of 2 mm, while the increase of peak amplitude to 5 mm leads
to a slight decrease of the 14% of the loss factors.
It is worth noticing the stability of the damping values over the frequency range considered. All meas-
urements were performed on samples pre-deformed for 10 mm. Iso-volumetric specimens (such 1M4
with marginal negative Poisson’s ratio of –0.067 at 40% of tensile strain) showed improved storage
modulus values compared to the conventional base foam, although lower than the auxetic version. Fig-
ure 12 shows that for pre-deformation of 10 mm and peak amplitude of 2 mm, an average (and marginal
frequency dependent) storage modulus of 200 kPa is reached, while the conventional foam exhibits a
maximum value almost three times lower. It is worth noticing how the increase of the peak amplitude to
5 mm leads to a two-fold increase of the related storage modulus, always with apparent slight frequency
dependence. The behaviour of the loss factors, compared to the ones of the conventional foam, is sub-
stantially similar to the one of the related auxetic foam, with values around 0.13 for peak amplitude of
2 mm and pre-deformation of 10 mm, and similar decrease of the loss factor for increased peak ampli-
tude value (Fig. 13).

Conventional PU foam and 1M4 (ν=-0.067)


0.40

0.35

0.30

0.25
Loss factor

conv p10a2
conv p10a5
0.20
1M p10a2
0.15 1M p10a5

0.10

0.05

0.00
0 2 4 6 8 10
Frequency [Hz]

Fig. 13 Comparison between loss factors for iso-volumetric (1M4) and conventional foam sample at
10 mm of pre-deformation and different peak amplitudes vs. frequency.

© 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


692 F. Scarpa et al.: Auxetic compliant flexible PU foams: static and dynamic properties

Fig. 14 (online colour at: www.pss-b.com) Constant strain-rate stress-strain curves for auxetic specimen
1L5.

4.5 Constant strain-rate tests


A constant strain-rate value is given by the following relation:
v
e = , (7)
l0
where v is the imposed velocity on the sample, and l0 the initial length of the specimen considered. Due
to internal hardware limitations, the tensile machine employed has not permitted to reach strain-rate
values over 12 s–1. As for the case of quasi-static tests, converted foams show mechanical stiffness at
least one order of magnitude larger than the one of the conventional foam. For example, at 70% of com-
pressive strain foams 1L5 showed stress values ranging from 5.8 MPa to 7 MPa at strain rates of 8 and
12 s–1 respectively (Fig. 14). The specimens with marked negative Poisson’s ratio behaviour showed a
more significant dependence of the stress-strain patterns over the strain rate value, while the iso-
volumetric samples (batch M3) give different stress-strain curve values around densification, being vir-
tually independent of strain rate up to 30% strain (Fig. 15). The conventional foam exhibits at different
constant strain rates the usual stress-strain pattern of flexible polymeric foams [8], but with stress values
of two orders of magnitude lower (Fig. 16).

Fig. 15 (online colour at: www.pss-b.com) Constant strain-rate stress-strain curve for iso-volumetric
specimen 1M3.

© 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


Original
Paper

phys. stat. sol. (b) 242, No. 3 (2005) / www.pss-b.com 693

Fig. 16 (online colour at: www.pss-b.com) Constant strain-rate stress-strain curve for conventional base
foam.

5 Conclusions
In this work a comparative analysis of some static and dynamic properties of auxetic compliant flexible
polyurethane foams has been performed. As a general remark, negative Poisson’s ratio foams show a
different stress-strain pattern in compressive loading compared to conventional polymeric thermoplastic
foams. The viscoelastic properties (storage and loss modulus) evaluated using the hysteresis cycle mod-
els show a significant increase in storage modulus values and, within the frequency range considered,
small variations with the harmonic angular excitation. More significant appears the dependence over
levels of pre-deformation and peak amplitude, a common feature with most polymeric open cell solids.
Constant strain-rate testing shows enhanced stress-strain characteristics, with stress values in general two
order of magnitude higher for auxetic foams compared to their base conventional counterpart.
To benchmark further the characteristics of the compliant foam produced, a set of samples having the
same overall volumetric compression ratio used to manufacture the auxetic foams has been produced,
resulting in specimens with virtually zero Poisson’s ratio. The comparative results carried out on the
latter specimen seem to confirm that the enhanced mechanical characteristics in compression of auxetic
foams are not given only by increased density resulting from the manufacturing process. While the loss
factors of the iso-volumetric foams were comparable with the auxetic ones, their related storage moduli
were significantly lower. The stress-strain curves for constant strain-rate values showed also stress val-
ues of one order of magnitude lower than the ones of their auxetic counterpart.

Acknowledgements The work has been partially funded through the EPSRC Grant GR/R93737. The authors
would like also to thank the Referees for their valuable suggestions.

References
[1] R. S. Lakes, Science 253, 1038 (1987).
[2] K. E. Evans, Endeav. New Ser. 15(4), 170 (1991).
[3] R. S. Lakes and K. Elms, J. Compos. Mater. 27, 1193 (1993).
[4] R. S. Lakes, ASME J. Mech. Design 115, 696 (1993).
[5] F. Scarpa and G. Tomlinson, J. Sound Vib. 230(1), 45 (2000).
[6] F. C. Smith and F. Scarpa, IEEE Microw. Guid. Wave Lett. 10(11), 451 (2000).
[7] A. W. Lipsett and A. I. Beltzer, J. Acoust. Soc. Am. 84, 2179 (1988).
[8] C. P. Chen and R. S. Lakes, Cell. Polymers 8, 343 (1989).
[9] B. Howell, P. Prendergast, and L. Hansen, Acoustic behaviour of negative Poisson’s ratio materials, DTRC-
SME-91/01 (David Taylor Research Centre, Annapolis, MD, 1991).

© 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


694 F. Scarpa et al.: Auxetic compliant flexible PU foams: static and dynamic properties

[10] C. P. Chen and R. S. Lakes, J. Eng. Mater. Technol. 118, 285 (1996).
[11] C. P. Chen and R. S. Lakes, J. Mater. Sci. 28, 4288 (1993).
[12] E. O. Martz, R. S. Lakes, and J. B. Park, Cell. Polymers 15(5), 349 (1996).
[13] F. Scarpa, L. G. Ciffo, and J. R. Yates, Smart Mater. Struct. 13, 49 (2004).
[14] L. J. Gibson and M. F. Ashby, Cellular Solids: structure and properties, 2nd edition (Cambridge University
Press, Cambridge, 1997).
[15] J. B. Choi and R. S. Lakes, Int. J. Mech. Sci. 1, 51 (1995).
[16] N. Chan and K. E. Evans, J. Mater. Sci. 32, 5945 (1997).
[17] F. Scarpa, J. Yates, and L. G. Ciffo, J. Mech. Eng. Sci. 216, 1 (2002).
[18] F. Gandhi and D. Wolons, Smart Mater. Struct. 8, 49 (1999).
[19] J. B. Choi and R. S. Lakes, J. Mater. Sci. 27, 4678 (1992).
[20] A. Blake and M. Isard, Active Contours – The application of techniques from graphics, vision, control theory
and statistics to visual tracking of shapes in motion (Springer-Verlag, London, 1998).

© 2005 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

S-ar putea să vă placă și