Sunteți pe pagina 1din 10

Composite Structures 184 (2018) 92–101

Contents lists available at ScienceDirect

Composite Structures
journal homepage: www.elsevier.com/locate/compstruct

Large deformation of an auxetic structure in tension: Experiments and finite MARK


element analysis

Jianjun Zhanga,b, Guoxing Lua, , Zhihua Wangb, Dong Ruana, Amer Alomaraha,
Yvonne Durandeta
a
Faculty of Science, Engineering and Technology, Swinburne University of Technology, Hawthorn, VIC 3122, Australia
b
Institute of Applied Mechanics and Biomedical Engineering, Taiyuan University of Technology, Taiyuan 030024, China

A R T I C L E I N F O A B S T R A C T

Keywords: The present paper reports on the post-yield behaviors of an auxetic structure, honeycomb with representative re-
Re-entrant hexagonal honeycomb entrant topology. Specimens were made of stainless steel and polymer, respectively. Quasi-static uniaxial tensile
Polymer and stainless steel tests were conducted in the two principal directions, followed by simulations using the commercial code –
Image correlation ABAQUS 6.11-2. The deformation, tensile stress-strain curves and Poisson’s ratio were of interest. A good
Poisson’s ratio
agreement was observed between the numerical simulations and the experimental results. Subsequently, the
Finite element analysis
effect of cell wall thickness and initial cell angle was studied by means of finite element analysis. An analytical
equation was also given for the yield stress of such materials under tension.

1. Introduction scales [20], especially for metal cellular structures. The selective laser
melting (SLM) or electron beam melting (EBM) techniques were em-
Over the past several decades, developments in structural en- ployed to fabricate the auxetic materials [21,22], and the direct laser
gineering design and technology in aircraft industry as well as auto- writing method (DLW) was also used [23,24]. One disadvantage of the
motive, sports, and leisure sectors have demanded novel materials to re-entrant honeycomb is, however, the difficulty in manufacturing them
meet higher engineering specifications [1]. Such materials are to pos- on a commercial scale [1]. Other topologies including chiral shape
sess a combination of high stiffness and strength with significant weight [25,26], star shape [27,28], arrowhead shape honeycombs [29] and
savings. Structural material with a negative Poisson’s ratio was ex- certain designed foams [30,31] leading to an auxetic effect might
plored [1,2], known as auxetic materials [3,4]. eventually yield a commercially acceptable structure in terms of man-
Lakes [5] first discovered this negative Poisson’s ratio effect in ufacturability and performance.
polyurethane (PU) foam with re-entrant structures and responded to a Materials with negative Poisson’s ratio demonstrate a series of
comment in Ref. [6] on this negative trait. The key to the auxetic be- particular characteristics over conventional ones, such as enhanced
havior was the negative value of Poisson’s ratio and this structural toughness and shear or indentation resistance [32], along with im-
material was foreseen with a wide range of applications: better artificial proved sound and vibration absorption [33]. They have been also
bones, improved sound and shock absorbers and enhanced bulletproof exploited for applications as fasteners [34], tougher composites [35],
vests; gaskets and seals were another evident area of applications [7]. medicine [36], tissue engineering [37] and others [33]. Up to now,
Subsequently, researchers managed to purposely fabricate a wide range studies have concentrated mainly on the small deformation elastic
of synthetic auxetic materials covering all major classes of materials, properties of negative Poisson’s ratio materials, in terms of the elastic
such as metals [8,9], polymers [10,11], textile [12,13], composites modulus and elastic Poisson’s ratio [19,38–41]. Chan and Evans
[14,15] and ceramics [16]. A methodology to convert conventional [10,42] fabricated auxetic foams and characterized their mechanical
foams into auxetic ones was first reported by Lakes [5] for polymer behaviors under compression and tension, respectively, with small de-
foams, which was followed by Choi and Lakes [8] and Friis et al. [9] to formation. Koudelka et al. [43] printed a three dimensional re-entrant
fabricate auxetic metallic foams. Fabrication processes such as multi- structure to test its stress variation under quasi-static compression.
phase auxetic fabrication [17,18] were then developed and modified by Unlike the unconverted foams, the re-entrant structured foams showed
some other researchers [19]. Recently, soft lithography has become an no significant plateau region up to a maximum compression at 50%
advantageous technology to make structural materials on different strain [44], similar to the outcomes of Ref. [10]. Auxetic effect and


Corresponding author.
E-mail address: glu@swin.edu.au (G. Lu).

http://dx.doi.org/10.1016/j.compstruct.2017.09.076
Received 29 May 2017; Received in revised form 5 September 2017; Accepted 26 September 2017
Available online 28 September 2017
0263-8223/ © 2017 Elsevier Ltd. All rights reserved.
J. Zhang et al. Composite Structures 184 (2018) 92–101

compression behavior of 3D auxetic textile structure were investigated measurement of their dimensions by a digital caliper, which represents
by Zhou et al. [45] and they found that the auxetic composite behaved 12.5% of the total cell number in the stainless steel sample, or 14.3%
more like a damping material with a lower range of compression stress. for the polymer sample. Average dimensions of the printed samples are
Further, Zhang et al. [46] simulated the in-plane dynamic crushing given in Table 1.
behaviors of auxetic honeycombs and found that the plateau stress at From Fig. 1(a), it has H0 = 2(h0 − l0cosθ0) and L0 = 2l0sinθ0. For a
the proximal end and the energy absorption could be improved by in- repeatable cell in a re-entrant specimen, the thickness of the vertical
creasing the negative cell angle, the relative density, the impact velo- walls on both sides is t/2, the relative density (the ratio between the
city, and the matrix material strength. An effective numerical model volume occupied by the solid cell walls and the nominal volume of a
was developed to simulate auxetic composite sandwich panels sub- reputable cell, H0L0) of the specimen can thus be calculated as
jected to blast loading [47] and demonstrated that this proposed ρ∗ (2l 0 + h 0) t
auxetic panel was a promising protective structure against blast load- ρ0 = =
ρs 2l 0sinθ0 (h 0−l 0cosθ0) (1)
ings. In addition, Alderson et al. [48] simulated the responses of re-

entrant hexagonal and re-entrant trichiral honeycombs subject to out- where ρ and ρs are the densities of the re-entrant honeycomb and the
of-plane bending and found that they underwent synclastic (dome- solid cell wall material, respectively. Consequently, for stainless steel
shape) curvature deformation. All the previous studies have been con- auxetic structure, the relative density is 21%, while it is 8.59% and
cerned with small deformation behavior, mainly under compression. 9.91%, respectively, for polymer-X1 and polymer-X2. The slight differ-
Research on the large deformation post-yielding behaviors of auxetic ence in the relative density of polymer-X1 and polymer-X2 is due to a
structures under tension is limited, but in this case, their mechanical small difference in the value of angle θ0 (as shown in Table 1).
behaviors such as deformation and Poisson’s ratio may be of much The stainless steel samples (Fig. 1b and c) were 3D printed by using
difference. the selective laser melting machine ProX DMP 200 from 3D Systems,
In this paper, a simple but typical auxetic structure, re-entrant which has a maximum laser power of 300 W. Re-entrant stainless steel
honeycomb, is employed, and its large deformation tensile behavior is structures were built layer-by-layer using a laser beam of 1.07 µm
experimentally and numerically investigated. Polymer and stainless wavelength to melt the 17–4 PH powder with a particle size of
steel specimens were made by 3D printing. The deformation modes, 20–32 µm. The laser power and laser focal plane position were selected
stresses and Poisson’s ratios of the re-entrant honeycombs are studied. as 240 W and 0.6 mm, respectively. The travelling speed of laser beam
was set at 2500 mm/s. The total number of layers for each sample was
2. Experiments 180 with an approximate thickness of 40 µm for each layer. After
printing the re-entrant stainless steel samples, wire cutting was em-
2.1. Specimens ployed to cut off the samples from the printing substrate, using CNC
Wire cutting machine (model Alpha OC, Fanuc Robocut).
The re-entrant honeycombs were manufactured by using 3D Liquid epoxy resin containing reactive diluents was used to fabricate
printing technology with stainless steel and polymer, as shown in Fig. 1. the re-entrant polymer samples. Plastic/photopolymer auxetic samples
l 0 is the original length of the inclined cell walls and h 0 that of the (Fig. 1d and e) were created after hardening epoxy resin using stereo
vertical cell walls. θ0 is the initial angle between the inclined and ver- lithography (SL) process. In the ProJet 6000 HD 3D printer, SL process
tical cell walls. T and t are the specimen thickness and cell wall thick- uses visible or ultraviolet (UV) laser and scanning mechanism to se-
ness, respectively. Five cells of each sample were chosen randomly for lectively solidify liquid photopolymer in order to form a layer of the

Fig. 1. Photographs of samples: (a) sketch of a re-


entrant unit cell; (b) auxetic stainless steel sample
to be loaded in the X2 direction (5 × 8 cells); (c)
auxetic stainless steel sample to be loaded in the
X1 direction (5 × 8 cells); (d) auxetic polymer
sample to be loaded in the X2 direction (5 × 7
cells); (e) auxetic polymer sample to be loaded in
the X1 direction (5 × 7 cells); (f) details of the
fixtures holding one side of the sample.

(a) (b) (c)

(d) (e) (f)

93
J. Zhang et al. Composite Structures 184 (2018) 92–101

Table 1
Material and dimensions of the samples (Note: Letters X1 and X2 are the loading directions).

Name Material h0 (mm) l0 (mm) θ0 (degree) t (mm) T (mm) Number of Cells

Polymer – X1 VisiJet SL Flex 16.8 8.4 45 0.33 5 5×7


Polymer – X2 VisiJet SL Flex 16.8 8.4 40 0.33 5 5×7
Stainless steel – X1 17-4 PH 16.8 8.4 30 0.5 7.2 5×8
Stainless steel – X2 17-4 PH 16.8 8.4 30 0.5 7.2 5×8

sample. AUTOCAD 3D model was sliced to 40 layers and then the UV the subsequent analysis in this study, the exact reason for the difference
laser scanned according to the in-plane profile of the model, which was not explored further.
solidified the liquid resin. The resin was hardened, one slice at a time, The mechanical behavior of the cell wall material may be slightly
to a prescribed depth of 0.125 mm. The build platform moved down different from that of the standard samples. Therefore, some cell wall
0.125 mm after each layer was polymerized. The total thickness of 40 pieces were cut from the auxetic samples and then subjected to micro-
layers was T = 5 mm. The dimensional accuracy was within and nano-indentation tests. From the indentation tests it was found that
0.025–0.05 mm per 25.4 mm. the hardness of the auxetic samples was approximately 20% lower than
In both the manufacturing methods, in order to eliminate the effect that of the standard tensile test coupons. This may be because material
of building direction on the mechanical properties, all the samples were properties are sensitive to printing process such as the multiple passes,
built in the same orientation. The out-of-plane direction aligned with which affects the porosity, residual stresses etc. Therefore, in the finite
the building direction (i.e. the direction of the laser). element simulation a factor of 80% was applied to the flow stress in the
In order to capture the local displacement fields, Vic-2D image stress-strain curves measured from the tensile tests.
correlation system was employed, which required adequate and dis-
tinguishable stochastic spots on the specimens. Therefore, all the sam-
2.3. Experimental setup
ples were first coated with a layer of white paint-Dura Max (Dulux) and
then left for at least two hours till the paint had dried. Afterwards, the
Uniaxial tensile tests were conducted by using Zwick Roell machine
samples were sprayed with Spray Easy (British Paints) to have black
(Zwick/Z010) with a load cell of 1 kN. Polymer – X1 and Stainless steel
dots on the surface.
– X1 were tested in the X1 direction, while Polymer – X2 and Stainless
steel – X2 in the X2 direction. In each test, two sides of the specimen
were pin-joined, allowing the edge to rotate freely, and the other two
2.2. Properties of the parent materials
free. Small fixtures were fabricated from 17-4 PH stainless steel to allow
the lateral movement in the X2 direction (Fig. 1f). Meanwhile, grease
In order to obtain the properties of cell wall material, standard
was applied on the contact surfaces to reduce friction between the
samples to the ASTM standards E8/E8M-15a (for stainless steel) and
specimen and the fixture during the tests. One edge of the sample was
D638-14 (for polymer) were also printed using identical values of op-
attached to a fixed position while the other moved with the cross-head
erating parameters to those for the auxetic samples. For both stainless
at 3 mm/min. Each test was stopped when fracture occurred.
steel and polymer samples, their gauge length is 25.0 mm. The average
dimensions from several measurements are 6.34 mm in width and
5.28 mm in thickness for the stainless steel samples. For the polymer 3. Finite element analysis
samples, the average width and thickness are 5.99 mm and 3.35 mm,
respectively. Engineering stress-strain curves of the different samples Finite element (FE) models with the same dimensions as those in the
were obtained and are given in Fig. 2. Meanwhile values of the density tests were set up and corresponding numerical simulations were per-
of the stainless steel and polymer were measured as 7550 kg/m3 and formed by using ABAQUS 6.11-2. In the simulations, three-node
1164 kg/m3, respectively, by calculating the average ratio of mass of quadratic Timoshenko beam element (B22) was used to mesh the cell
the sample in the gauge length to the corresponding volume. Three walls of re-entrant honeycombs with the element size 0.24 mm (35
nominally identical samples were tested and the stress-strain curves of elements in each inclined wall), which was shown to produce con-
both the materials were repeatable except for the value of the fracture verged FE results. The stress-strain relationships obtained from ASTM
strain for the polymer. The difference in fracture strain might be due to standard tests were employed in the FE models as the material beha-
the manufacturing technique. Since fracture strain will not be used in vior. In regard of loads and confinements, a uniform displacement was

Fig. 2. Engineering stress-strain curves of the parent materials: (a) stainless steel; (b) polymer.

94
J. Zhang et al. Composite Structures 184 (2018) 92–101

Fig. 3. Deformed polymer samples in the tests: (a)-(d) for


loading in the X2 direction; (e)-(h) for loading in the X1
direction.

Fig. 4. Deformed stainless steel samples in the tests: (a)-(d)


for loading in the X2 direction; (e)-(h) for loading in the X1
direction.

Fig. 5. Deformed polymer specimens from FE simulations:


(a)-(d) for loading in the X2 direction; (e)-(h) for loading in
the X1 direction.

95
J. Zhang et al. Composite Structures 184 (2018) 92–101

Fig. 6. Deformed stainless steel specimens from FE simula-


tions: (a)-(d) for loading in the X2 direction; (e)-(h) for
loading in the X1 direction.

Fig. 7. Engineering stress-strain curves: polymer sample loaded in the X2 direction (a) and in the X1 direction (b); and stainless steel sample loaded in the X2 direction (c) and in the X1
direction (d).

applied as loading to the corresponding nodes of the models whilst the 4. Experimental and finite element analysis results
supporting nodes were constrained in the loading direction, but free in
the other perpendicular direction. Failure (fracture) of materials was 4.1. Deformation pattern
not considered in simulations.
For each test, a load-displacement curve was obtained directly from
the machine. The load can be converted to the nominal engineering
stress by taking its ratio to the nominal cross-sectional area of the
specimen, and the displacement divided by the initial length of the

96
J. Zhang et al. Composite Structures 184 (2018) 92–101

Fig. 8. Points traced for determining the longitudinal (blue


points) and lateral (orange points) engineering strain during
the tests. Note that this figure is similar to Fig. 3(a)-(d).

Fig. 9. Total engineering strain in the two directions against the cross-head displacement (a); and the averaged value (b).

specimen gives the engineering strain. As expected, when the polymer data with respect to the characteristics of curves and magnitudes of
auxetic sample was stretched in the X2 direction, lateral expansion stress. In Fig. 7(c), in the test some cell walls of the stainless steel
occurred, as shown in Fig. 3(a) and (b). This expansion indicates a honeycomb fractured at approximately 0.15 of the engineering strain,
negative Poisson’s ratio, until the initially inclined cell walls became resulting in a sudden decrease in the stress at a small strain. Further-
horizontal (Fig. 3c). Afterwards, cells shrank in the X1 direction as they more, in terms of the shape of the curves for the loading in the X2
were further stretched in the X2 direction (Fig. 3d), which indicated direction, there exists a plateau stage after an initial elastic stage, fol-
that the sample then possessed a positive Poisson’s ratio during this lowed by a final stage with a rapid increase in stress, which is similar to
stage, similar to conventional materials. When loaded in the X1 direc- cellular materials under compression [49–51]. It should be noted that,
tion, as exhibited in Fig. 3(e)-(h), the polymer sample kept expanding in from Fig. 7(a), the test result is a slightly higher than that of the si-
both the X1 and X2 directions until the initially inclined cell walls be- mulation. This may be explained by recalling the comparison between
came vertical and finally broke. The characteristic of a negative Pois- the deforming profiles in Figs. 3(c) and 5(c). The cells deform non-
son’s ratio was present during the complete tensile process. Similarly, uniformly in the test resulting from the horizontal friction on the con-
the deformation patterns of the stainless steel samples are shown in tacting surfaces at the top and bottom ends, while they deform uni-
Fig. 4(a)-(d) (X2 direction loading) and Fig. 4(e)-(f) (X1 direction formly in simulation. This shows that the slight transverse restraint
loading). A difference from the polymer sample is that the sample in leads to a small increase of stress.
Fig. 4(d) fractured at some cells before the cells started to contract. In
addition, considerable frictional effect (Fig. 3c, h, 4 b and g) was ob-
served even though grease was applied to minimise friction in the tests. 4.3. Poisson’s ratio
The overall specimen became barrel shaped and the cells deformed
inhomogeneously, especially those within the bend of two columns in Poisson’s ratio is an important parameter of a material, which
Fig. 3(h). Corresponding to the experimental observations, Figs. 5 and 6 generally represents the negative of the ratio of the lateral extension to
show the FE deformation patterns (with free lateral movements, or no the longitudinal extension (extension in the loading direction).
friction). The cells expanded homogeneously. Poisson’s ratio calculated from the engineering strain reflects the
macroscopic and averaged value. To observe the progressive deforma-
tion of auxetic honeycombs in both the X1 and X2 directions, a series of
4.2. Tensile stress versus strain curves images (frames) were captured by the Digital Image Correlation (DIC)
System during the testing process. Consequently, the successive incre-
As mentioned before, the engineering stress-strain curves can be ment of the total extension between frames was captured and the
obtained from the load-displacement curves of the specimens and they Poisson’s ratio could then be calculated at each increment. From these
are shown in Fig. 7, for re-entrant honeycombs made of polymer and images, local displacement field and hence engineering strain field)
stainless steel. Generally, the experimental results agree with the FE were calculated. Here, eight pairs of points were selected and

97
J. Zhang et al. Composite Structures 184 (2018) 92–101

Fig. 10. Poisson’s ratio from the tests and finite element simulations: polymer samples loaded in the X2 (a) and X1 (b) directions; stainless steel samples loaded in the X2 (c) and X1 (d)
directions.

F2 head displacement in Fig. 9(a) and their averaged value is shown in


Fig. 9(b). As expected, the strain in the loading direction increased
2 linearly with the cross-head displacement. However, strain in the other
A direction increased initially and reached a peak, then it decreased
gradually. This demonstrates the expansion and then shrinking of the
auxetic honeycombs in uniaxial tensile tests, as observed before.
For plastic deformation, Poisson’s ratio was calculated as the ratio of
M strain increments in the two directions[52,53], instead of using the
X2 A
B total engineering strain. It represents the instantaneous incremental
X1 deformation features of the re-entrant honeycombs. Values of the
F2 0 Poisson’s ratio such calculated from the tests and finite element analysis
2 are plotted in Fig. 10. In all the plots, the Poisson’s ratio changes with
the total strain in a non-linear way, with both negative and positive
values. For auxetic polymer samples loaded in the X2 direction, Pois-
F2
X2 son’s ratio from both the test and simulation increases gradually from
2
about −2.5 to about 2.0, reaching zero at an approximate strain of 0.6.
2 This indicates that when a polymer re-entrant honeycomb is subjected
X1 B
to tension in the longitudinal direction, it expands firstly and then
F2 M
(a) (b) contracts along the lateral direction. This agrees with the observed
deformation patterns during the tensile tests (Fig. 3). It also shows that
Fig. 11. A single re-entrant cell loaded in the X2 direction (a); and the free body diagram
as the deformation progresses the re-entrant honeycomb behaves from
of the cell wall AB (b).
an auxetic material to a conventional one with positive Poisson’s ratio.
Under the load in the X1 direction (Fig. 10b), Poison’s ratio was always
monitored (four in the horizontal direction as marked in orange, and negative and it decreased to as low as −2.8 at a longitudinal strain of
the other four in the vertical direction, in blue) (Fig. 8). Using the first 0.38. However, around this point the lowest values from the test and
frame as the reference, the vertical relative displacement between the simulation are very different. The reason can be found by referring to
corresponding points divided by the initial distance gave the en- Fig. 3(h). The two columns of cells did not move freely due to friction
gineering strain in the loading direction. Similarly, the lateral en- on both the loading and fixed edges, resulting in reduced lateral ex-
gineering strain was calculated. The lateral strains and longitudinal pansion. Similarly, for the stainless steel sample in Fig. 10(c), the
strains measured at the various locations are plotted against the cross-

98
J. Zhang et al. Composite Structures 184 (2018) 92–101

Fig. 12. Tensile stress and theoretical yield stress of polymer sample (a); and stainless steel sample (b), loaded in both the X1 and X2 directions.

5. Discussions

5.1. Yield stress

It would be interesting to give some theoretical consideration on the


yield stress of a sample. For reasonably thick cell walls, this global
yielding is governed by the plastic collapse of a cell. Consider a single
cell under tensile force F2 in the X2 direction (Fig. 11). When plastic
collapse occurs, plastic hinges form at all the corners. Assuming the
bending moment at the hinges reaches the fully plastic bending moment
Mp = 1/4σys Tt 2 , where σys is the yield stress of the parent material of cell
walls, from the equilibrium of cell wall AB, the force F2 is
F2 = 2σ2 l 0 Tsinθ0 . The corresponding nominal stress is hence the ratio of
F2 and the cross-sectional area:
2
σ2 1 t 1
= ⎛ ⎞⎜ ⎟

σys 2 ⎝ l 0 ⎠ sin2 θ0 (2)


Fig. 13. Effect of cell wall thickness (t) on engineering stress-strain curves of the re-
entrant honeycombs loaded in the X2 direction (θ0 = 40°). Similarly, the tensile yield stress in the X1-direction, σ1, is given as
2
σ1 1 t 1
= ⎛ ⎞⎜ ⎟

σys 2 ⎝ l 0 ⎠ cosθ0 (h 0 / l 0−cosθ0 ) (3)

Analytical predictions from Eqs. (2) and (3) are compared with the
experimental results for polymer and stainless steel samples loaded in
both the X1 and X2 directions in Fig. 12. As in the finite element si-
mulations, the yield stress of the parent materials is taken as 80% of
that from the coupon tests. The comparison indicates that Eqs. (2) and
(3) could predict the initial yield stress of stainless steel samples for
both the directions (Fig. 12b), but it is not the case for the polymer
samples. This is because, compared with conventional manufacturing
processes, the stereo lithography (SL) process method lacks dimensional
accuracy due to parameters involving layer thickness, hatch over-cure
and hatch spacing [54].

5.2. Effect of geometrical parameters

Combining Eqs. (1) and (2), the tensile yield stress in the X2-di-
Fig. 14. Effect of the cell angle on the engineering stress-strain curves of re-entrant rection, σ2, can be re-written as
honeycombs loaded in the X2 direction (ρ0 = 7.5%).
2
σ2 ρ∗
= 2 ⎛⎜ ⎞⎟
2
( h0
l0
−cosθ0 )
Poisson’s ratio was negative firstly, about −4.0, and then it increased σys h 2
gradually. The magnitude from the test becomes lower than that from ⎝ ρs ⎠ (2 + 0
)
l0 (4)
the simulation after a longitudinal engineering strain of 0.1, due to the
lateral constraint from friction and fracture of some cell walls in the It indicates that, in addition to the cell dimensions l0 and h0, the yield
test. For the stainless steel sample loaded in the X1 direction (Fig. 10d), stress is also related to both the relative density (ρ0) and initial angle of
the trend is the same as that in Fig. 10(b). cells (θ0).
Influence of the relative density on the tensile stress was studied by
using the FE models in Section 3. This was achieved by changing the

99
J. Zhang et al. Composite Structures 184 (2018) 92–101

Fig. 15. Deformed (ε = 0.7) patterns for different values of


friction coefficient of the polymer samples: (a) μ = 0.1; (b)
μ = 0.3 ; (c) μ = 0.5 ; (d) μ = 0.8 . The original specimen is
also shown in dotted lines.

(a) (b) (c) (d)

Other definitions such as element size, loading and boundary conditions


were the same with those in the Section 3. The stress-strain relationship
of polymer in the Section 2.2 was employed in the FE models.
Fig. 15 shows the deformed samples at ε = 0.7 (corresponding to
the displacement 72.5 mm). The deformation type at μ = 0.3 is similar
to that in Fig. 3(b). Meanwhile, for μ = 0.8, no lateral movement of
cells at both the proximal and distal ends was observed, as if the lateral
movement was fully constrained. Fig. 16 illustrates the force-displace-
ment curves for different values of friction coefficient. A large value of
friction coefficient brings forward the final stage. However, when the
friction is small (μ ≤ 0.3) the curves are almost the same with excep-
tion of slight fluctuations. For the friction coefficient lager than 0.3, the
curves are all similar to that of the fully constrained case (μ = 0.8).

6. Conclusions
Fig. 16. Force-displacement curves with different values of friction coefficient.
As typical auxetic structures, several stainless steel and polymer re-
entrant honeycomb samples were printed using the 3D printing tech-
cell wall thickness, t, in Eq. (1). Four values of the cell wall thickness nology. Quasi-static tensile tests were then performed in both the X1
were considered: t = 0.15 mm (ρ0 = 4.5%), t = 0.20 mm (ρ0 = 6.0%), and X2 directions, followed by the finite element simulation. The de-
t = 0.25 mm (ρ0 = 7.5%), and t = 0.30 mm (ρ0 = 9.0%). FE results are formation, tensile stress and Poisson’s ratio of re-entrant honeycombs
plotted in Fig. 13. As indicated in Eq. (4), a higher value of relative were studied in the tests and simulations. A good agreement between
density leads to a higher value of the tensile stress plateau, as well as the experimental and FE results has been obtained.
the initial slope of the stress-strain curves. It has been observed that when loaded in the X2 direction the re-
The initial angle of cell, θ0, also affects the tensile stress. FE models entrant honeycombs had a large strain of 1.3 and the value of Poisson’s
with θ0 = 30°, 35°, 40°, 45° were established and the cell dimensions, l0 ratio changed during the deformation. The honeycombs initially ex-
and h0, were fixed at the same values as those in the tests. The relative hibited an auxetic feature with negative Poisson’s ratio almost as low as
density of all these FE models was kept the same at 7.5% by changing −4. Beyond a certain point of deformation, they behaved as conven-
the value of thickness t. Therefore, a larger value of the initial angle tional honeycombs with a positive value of Poisson’s ratio. In the X1-
corresponds to a thicker cell wall. Fig. 14 presents effect of the cell direction, the auxetic samples kept expanding until the inclined cell
angle on the tensile stress in the X2-loading direction. A larger initial walls became horizontal. The Poisson’s ratio was negative for the whole
angle leads to a higher tensile stress plateau and the elastic slopes of process. In terms of the tensile stress, all the stress-strain curves had a
stress-strain curves, but it decreases the value of the maximum strain plateau stage, similar to compression of conventional cellular materials.
when the stress increases rapidly, similar to the locking strain in cel- An analytical formula based on the plastic collapse of the cell walls
lular materials. For a single cell in Fig. 1(a), the initial height of this cell has been given, which could lead to a reasonable estimate of the yield
is 2(h0 − l0cosθ0) and the maximum extension is 2l0(1 + cosθ0), the stress for stainless steel honeycombs, but not so for the polymer sam-
maximum critical strain, εmax, can be expressed as ples. A higher relative density led to a higher stress plateau and the
1 + cosθ0 elastic slopes of stress-strain curves, as expected. For a given density, a
εmax = larger initial cell angle also enhanced the plateau stress, but reduced the
h 0 / l 0−cosθ0 (5)
maximum strain in the final stage of the stress-strain curves.
It shows that, for a special case of h0 = 2l0, the corresponding max-
imum strain is related only to the initial angle of cells: a larger one
reduces the critical strain of the final stage of curves. Acknowledgements

5.3. Effect of friction The authors wish to thank Dr Andrew Ang for his help with the
nano-indentation tests. They acknowledge, with thanks, the financial
This section explores effect of possible lateral constraint due to support from the Australian Research Council (Grant No.
friction. One loading case shown in Fig. 1(d) was considered. Details of DP160102612), National Natural Science Foundation of China (Grant
the fixture used in the experiment were reproduced in the FE model and Nos. 51578361 and 11572214), Shanxi Scholarship Council of China
different values of the friction coefficient, μ, were defined for the con- (2013-046) and the State Scholarship Fund of China Scholarship
tacting surfaces (General Contact). In the simulation, three-node Council (CSC).
quadratic Timoshenko beam element (B32) was used for meshing.

100
J. Zhang et al. Composite Structures 184 (2018) 92–101

References [28] Theocaris P, Stavroulakis G, Panagiotopoulos P. Negative Poisson’s ratios in com-


posites with star-shaped inclusions: a numerical homogenization approach. Arch
Appl Mech 1997;67:274–86.
[1] Alderson A, Alderson KL. Auxetic materials. Proc Inst Mech Eng Part G-J Aerosp Eng [29] Larsen UD, Signund O, Bouwsta S. Design and fabrication of compliant micro-
2007;221:565–75. mechanisms and structures with negative Poisson’s ratio. J Microelectromech Syst
[2] Bhullar SK. Three decades of auxetic polymers: a review. e-Polymers 1997;6:99–106.
2015;15:205–15. [30] Babaee S, Shim J, Weaver JC, Chen ER, Patel N, Bertoldi K. 3D Soft metamaterials
[3] Evans KE, Nkansah MA, Hutchinson IJ, Rogers SC. Molecular network design. with negative Poisson’s ratio. Adv Mater 2013;25:5044–9.
Nature 1991;353:124. [31] Alderson A, Alderson KL, McDonald SA, Mottershead B, Nazare S, Withers PJ, et al.
[4] Evans KE. Auxetic polymers – a new range of materials. Endeavour 1991;15:170–4. Piezomorphic materials. Macromol Mater Eng 2013;298:318–27.
[5] Lakes R. Foam structures with a negative Poisson’s ratio. Science [32] Greaves GN, Greer A, Lakes R, Rouxel T. Poisson’s ratio and modern materials. Nat
1987;235:1038–40. Mater 2011;10:823–37.
[6] Burns S. Negative Poisson’s ratio materials. Science 1987;238:551. [33] Jiang JW, Park HS. Negative poisson’s ratio in single-layer black phosphorus. Nat.
[7] Cherfas J. Stretching the point. Science 1990;247:630. Commun. 2014;5.
[8] Choi J, Lakes R. Non-linear properties of metallic cellular materials with a negative [34] Choi J, Lakes R. Design of a fastener based on negative Poisson’s ratio foam. Cell
Poisson’s ratio. J Mater Sci 1992;27:5375–81. Polym 1991;10:205–12.
[9] Friis E, Lakes R, Park J. Negative Poisson’s ratio polymeric and metallic foams. J [35] Sun Y, Pugno N. Hierarchical fibers with a negative Poisson’s ratio for tougher
Mater Sci 1988;23:4406–14. composites. Materials 2013;6:699–712.
[10] Chan N, Evans K. The mechanical properties of conventional and auxetic foams. [36] Scarpa F. Auxetic materials for bioprostheses. IEEE Signal Process Mag
Part I: compression and tension. J Cell Plast 1999;35:130–65. 2008;25:125–6.
[11] Scarpa F, Pastorino P, Garelli A, Patsias S, Ruzzene M. Auxetic compliant flexible [37] Park YJ, Kim JK. The effect of negative Poisson’s ratio polyurethane scaffolds for
PU foams: static and dynamic properties. Phys Status Solidi B-Basic Solid State Phys articular cartilage tissue engineering applications. Adv Mater Sci Eng
2005;242:681–94. 2013;2013:1–5.
[12] Wang ZY, Hu H. Auxetic materials and their potential applications in textiles. Text [38] Yang W, Li ZM, Shi W, Xie BH, Yang MB. Review on auxetic materials. J Mater Sci
Res J 2014;84:1600–11. 2004;39:3269–79.
[13] Reda H, Rahali Y, Ganghoffer JF, Lakiss H. Wave propagation in 3D viscoelastic [39] Alderson A, Evans K. Microstructural modelling of auxetic microporous polymers. J
auxetic and textile materials by homogenized continuum micropolar models. Mater Sci 1995;30:3319–32.
Compos Struct 2016;141:328–45. [40] Prall D, Lakes R. Properties of a chiral honeycomb with a Poisson’s ratio of -1. Int J
[14] Zhang G, Ghita O, Evans KE. The fabrication and mechanical properties of a novel Mech Sci 1997;39:305–14.
3-component auxetic structure for composites. Compos Sci Technol [41] Hou X, Hu H, Silberschmidt V. A novel concept to develop composite structures
2015;117:257–67. with isotropic negative Poisson’s ratio: effects of random inclusions. Compos Sci
[15] Miller W, Ren Z, Smith C, Evans K. A negative Poisson’s ratio carbon fibre composite Technol 2012;72:1848–54.
using a negative Poisson’s ratio yarn reinforcement. Compos Sci Technol [42] Chan N, Evans K. The mechanical properties of conventional and auxetic foams.
2012;72:761–6. Part II: shear. J Cell Plast 1999;35:166–83.
[16] Huang X, Blackburn S. Developing a new processing route to manufacture honey- [43] Embala A, Folije N. Corrosion behavior and the weak-magnetic-field effect of alu-
comb ceramics with negative Poisson’s ratio. Key Eng Mater 2001;206:201–4. minum packaging paper. Materiali in Tehnologije 2016;50:165–73.
[17] Grima JN, Attard D, Gatt R, Cassar RN. A novel process for the manufacture of [44] Allen T, Shepherd J, Hewage T, Senior T, Foster L, Alderson A. Low-kinetic energy
auxetic foams and for their re-conversion to conventional form. Adv Eng Mater impact response of auxetic and conventional open-cell polyurethane foams. Phys
2009;11:533–5. Status Solidi B-Basic Solid State Phys 2015;252:1631–9.
[18] Bianchi M, Frontoni S, Scarpa F, Smith C. Density change during the manufacturing [45] Zhou L, Jiang L, Hu H. Auxetic composites made of 3D textile structure and poly-
process of PU–PE open cell auxetic foams. Phys Status Solidi B-Basic Solid State urethane foam. Phys Status Solidi B-Basic Solid State Phys 2016;253:1233.
Phys 2011;248:30–8. [46] Zhang XC, Ding HM, An LQ, Wang XL. Numerical investigation on dynamic
[19] Critchley R, Corni I, Wharton JA, Walsh FC, Wood RJ, Stokes KR. A review of the crushing behavior of auxetic honeycombs with various cell-wall angles. Adv Mech
manufacture, mechanical properties and potential applications of auxetic foams. Eng 2015;7:1–12.
Phys Status Solidi B-Basic Solid State Phys 2013;250:1963–82. [47] Imbalzano G, Tran P, Ngo TD, Lee PV. A numerical study of auxetic composite
[20] Xu B, Arias F, Brittain ST, Zhao X-M, Grzybowski B, Torquato S, et al. Making panels under blast loadings. Compos Struct 2016;135:339–52.
negative Poisson’s ratio microstructures by soft lithography. Adv Mater [48] Alderson A, Alderson K, Chirima G, Ravirala N, Zied K. The in-plane linear elastic
1999;11:1186–9. constants and out-of-plane bending of 3-coordinated ligament and cylinder-liga-
[21] Li S, Hassanin H, Attallah MM, Adkins NJ, Essa K. The development of TiNi-based ment honeycombs. Compos Sci Technol 2010;70:1034–41.
negative Poisson’s ratio structure using selective laser melting. Acta Mater [49] Gibson LJ, Ashby MF. Cellular solids: structure and properties. Cambridge
2016;105:75–83. University Press; 1999.
[22] Yang L, Harrysson O, West H, Cormier D. Modeling of uniaxial compression in a 3D [50] Li ZQ, Zhang JJ, Fan JH, Wang ZH, Zhao LM. On crushing response of the three-
periodic re-entrant lattice structure. J Mater Sci 2013;48:1413–22. dimensional closed-cell foam based on Voronoi model. Mech Mater 2014;68:85–94.
[23] Jang D, Meza LR, Greer F, Greer JR. Fabrication and deformation of three-dimen- [51] Ruan D, Lu G, Wang B, Yu T. In-plane dynamic crushing of honeycombs-a finite
sional hollow ceramic nanostructures. Nat Mater 2013;12:893–8. element study. Int J Impact Eng 2003;28:161–82.
[24] Walia S, Shah CM, Gutruf P, Nili H, Chowdhury DR, Withayachumnankul W, et al. [52] Wang DA, Pan J. A non-quadratic yield function for polymeric foams. Int J Plast
Flexible metasurfaces and metamaterials: a review of materials and fabrication 2006;22:434–58.
processes at micro-and nano-scales. Appl Phys Rev 2015;2:011303. [53] Deshpande V, Fleck N. Isotropic constitutive models for metallic foams. J Mech
[25] Sigmund O, Torquato S, Aksay IA. On the design of 1–3 piezocomposites using Phys Solids 2000;48:1253–83.
topology optimization. J Mater Res 1998;13:1038–48. [54] Park WS, Lee SH, Cho HS, Leu MC. Evaluating the contributions of process para-
[26] Liu X, Hu G. Elastic metamaterials making use of chirality: a review. Strojniški meters in SLA using artificial neural network. Intelligent Manufacturing Systems.
vestnik-J Mech Eng 2016;62:403–18. 1997;IMS’97:189–94.
[27] Grima JN, Gatt R, Alderson A, Evans K. On the potential of connected stars as
auxetic systems. Mol Simul 2005;31:925–35.

101

S-ar putea să vă placă și