You are on page 1of 16

The International Journal of Advanced Manufacturing Technology (2018) 99:2425–2440

https://doi.org/10.1007/s00170-018-2637-y

ORIGINAL ARTICLE

Tensile properties of an auxetic structure with re-entrant and chiral


features—a finite element study
Amer Alomarah 1,2 & Dong Ruan 1 & Syed Masood 1

Received: 15 March 2018 / Accepted: 22 August 2018 / Published online: 4 September 2018
# Springer-Verlag London Ltd., part of Springer Nature 2018

Abstract
A re-entrant chiral auxetic (RCA) structure is a new structure, which comprises three topologies: re-entrant, trichiral and anti-
trichiral. The aim of this study is to numerically investigate the influences of geometrical parameters of a RCA unit cell on the in-
plane tensile properties. Finite element (FE) models were developed using commercial software, ABAQUS 6.11–2, and verified
by experimental results. Based on the verified FE models, a parametric study was performed to examine the effects of geometrical
parameters on the stress–strain curve and Poisson’s ratio of a RCA structure. In order to reduce the variables studied, three
dimensionless groups of geometrical parameters were proposed. Results showed that the RCA structure predominantly deformed
by bending and stretching of walls as well as rotation of cylinders. The RCA structure displayed high stress, low strain and large
negative Poisson’s ratio when it was loaded in the X direction. On the other hand, the RCA structure exhibited low stress, high
strain and a transition from negative to positive Poisson’s ratio when it was loaded in the Y direction. The geometrical parameters
of the RCA structure had significant effects on its stress–strain curves and Poisson’s ratios. The findings of this study provide an
insight into the mechanical response of the RCA structure under uniaixal tensile loading.

Keywords Auxetic structure . In-plane tensile properties . Negative Poisson’s ratio . Finite element analysis . Re-entrant and chiral
features

1 Introduction conventional materials, auxetic materials exhibit a series of


superior properties including better indentation resistance,
Most engineering materials exhibit positive Poisson’s ratios shear modulus, fracture toughness, variable permeability and
[1], i.e. they contract laterally when subjected to an axial tensile energy absorption [7–10].
load and they expand laterally when subjected to an axial com- The most popular auxetic structures include chiral honey-
pressive load. In contrast, auxetic structures/materials (or meta- combs [11] and the re-entrant honeycomb [12, 13]. Chiral hon-
materials) expand laterally when they are stretched axially or eycombs compose straight ligaments (ribs) that tangentially
contract laterally when they are compressed axially, displaying connect to cylinders (nodes). Based on the coordinates or num-
negative Poisson’s ratios (NPR) [2, 3]. The first material with ber of ligaments connecting to each cylinder, chiral honey-
NPR was reported in 1985 [4], but wide acceptance and rec- combs are classified as trichiral honeycombs (3-coordinated)
ognition of this type of materials were from 1987 [5]. The term shown in Fig. 1a, b, tetrachiral (4-coordinated) honeycombs
“auxetic” was introduced in 1991 for the first time to describe shown in Fig. 1c, d and hexachiral (6-coordinated) honey-
the characteristics of these materials [6]. Compared with combs (not shown here) [14, 15]. Moreover, trichiral honey-
combs have two common configurations: trichiral configura-
tion shown in Fig. 1a where each adjacent cylinders are at-
tached to the connecting ligament/wall on opposite sides, and
* Dong Ruan
druan@swin.edu.au anti-trichiral configuration shown in Fig. 1b where each adja-
cent cylinders are attached to the connecting ligament/wall on
1 the same side. Tetrachiral honeycombs can also be classified in
Faculty of Science, Engineering and Technology, Swinburne
University of Technology, Hawthorn, VIC 3122, Australia a similar way and are shown in Fig. 1c, d. The deformation
2 mechanism of chiral honeycombs comes from wrapping and
Department of Mechanics, College of Engineering, Wasit University,
Al-Kut 52001, Iraq unwrapping of the ligaments around the rotated nodes as a
2426 Int J Adv Manuf Technol (2018) 99:2425–2440

Fig. 1 Sketches of unit cell of: a trichiral honeycomb; b anti-trichiral honeycomb; c tetrachiral honeycomb; d anti-tetrachiral honeycomb; e re-entrant
honeycomb; f RCA structure (ligaments for trichiral and anti-trichiral topologies are indicated by green and red colours, respectively)

response to a compressive or tensile loading, respectively. This made to propose novel auxetic structures by combining geo-
behaviour leads to auxeticity of these structures. Despite sym- metrical features of existing auxetic structures. For example,
metrical topology of chiral honeycombs, their auxetic effect is the re-entrant hexagonal honeycomb was modified into a
relatively small. Poisson’s ratio of the 6-coordinated chiral hon- cylinder-ligament chiral honeycomb [24]. Based on the hex-
eycomb was found to be approximately − 1 [16, 17]. On the agonal re-entrant honeycomb, two new structures (splined re-
other hand, re-entrant honeycomb has straight and inclined entrant and stiffened re-entrant honeycombs) were developed
walls as shown in Fig. 1e. Due to its simplicity, adaptability [25]. A new auxetic structure was put forward by embedding
and ability to produce large negative Poisson’s ratio (up to − 4) the configuration of a rhombic auxetic structure into a re-
compared with other auxetic structures, extensive studies entrant honeycomb [26]. These innovative efforts triggered
[18–23] have been conducted on re-entrant honeycomb. the development of the current RCA structure, which incor-
The attractive properties of auxetic structures have contrib- porated the chiral feature in the re-entrant honeycomb [27].
uted to the incentive to develop advanced auxetic structures The RCA structure has a lower density compared with the
with high performance. To date, many attempts have been corresponding re-entrant honeycomb that has identical

Table 1 Relevant past research on auxetic structures

Structures Loading Methods used and properties investigated

Chiral honeycomb Compression • Poisson’s ratio and Young’s modulus (FEA in [14, 17, 28], experiments in [24]).
• Effects of geometrical parameters (α = L1/r ratio, β = t/r ratio) on Poisson’s ratio and
Young’s modulus (FEA in [14, 24]).
• Effects of geometrical parameters (α = L1/r ratio) on Poisson’s ratio, shear modulus
and deformation mode (analytical model in [15]).
Re-entrant honeycomb Compression • Mechanical performance of modified unit cell (experiment, FEA and analytical model in [26]).
• Young’s modulus and failure strength (experiments in [29]).
• Effects of geometrical parameters (H, L1 and inclination angle) on Poisson’s ratio
(Analytical model in [13]), on shear properties (experiments in [30]).
Tension • Effects of loading direction and geometrical parameters on deformation mode,
stress–strain curves and Poisson’s ratio (experiment and FEA in [31]).
• Tensile properties and Poisson’s ratio of modified unit cell (experiments and analytical model in [32]).
Rotating rigid units Tension • Effects of geometrical parameters on Poisson’s ratio (analytical model and FEA in [33, 34]).
• Young’s modulus, Poisson’s ratio and shear modulus (experiments and FEA in [35]).
RCA Tension • Effects of parent material and loading direction on the deformation mode,
stress–strain curve and Poisson’s ratio (experiments in [27]).
Int J Adv Manuf Technol (2018) 99:2425–2440 2427

Fig. 2 Manufactured aluminium


RCA samples [27] to be loaded in
the a Y direction; b X direction; c
quasi-static true stress–strain
curves to parent material
(aluminium alloy 6061)

400
350
300

True stress (MPa)


250
200
150
Test-1
100 Test-2
50 Test-3
Test-4
0
0 0.03 0.06 0.09 0.12 0.15 0.18 0.21
(c)
True strain

geometrical parameters such as unit cell wall length and incli- brief summary of auxetic structures studied previously is il-
nation angle [27]. The RCA structure possessed higher stress lustrated in Table 1.
and more evident auxetic effect compared with re-entrant hon- Mechanical properties of auxetic structures can be tuned by
eycombs when they were tensioned in the X direction [27]. A tailoring the geometrical parameters of their unit cells. Studies

Table 2 Material properties of aluminium (6061) used in the FEA [27]

Material Elastic modulus (GPa) Poisson’s ratio Density (g/cm3) Yield stress (MPa) Ultimate stress (MPa)

Aluminium (6061) 65 (GPa) 0.33 2.7 210 350

Fig. 3 a In-plane dimensions of a RCA sample and locations of selected direction; c loading and boundary conditions of FE model for a RCA
points to calculate the axial and lateral displacements; b loading and sample loaded in the Y direction
boundary conditions of FE model for a RCA sample loaded in the X
2428 Int J Adv Manuf Technol (2018) 99:2425–2440

Table 3 Parameters investigated in the current FEA

Studied parameters

Three dimensionless groups One parameter


α = (L1/r) β = (t1/r) λ = (H/L1) θ0
2.3 = 9/3.9 0.16 = 0.45/2.75 1.8 = 18/10 30°
3.27 = 9/2.75 0.18 = 0.5/2.75 2.0 = 18/9 40°
4.5 = 9/2 0.20 = 0.55/2.75 2.2 = 18/8.2 50°
6.0 = 9/1.5 0.22 = 0.6/2.75 2.4 = 18/7.5 60°
9.0 = 9/1 0.24 = 0.66/2.75 2.6 = 18/6.9 70°

have been conducted to investigate the influence of geometri-


cal parameters on the mechanical performance of auxetic
structures. The geometrical parameters (such as rib length Fig. 5 Experimental and FE simulated force–displacement curves of the
and thickness) of splined re-entrant and stiffened re-entrant aluminium RCA structure loaded in both the X and Y directions [27]
honeycombs were experimentally and numerically found
[25] to have a significant effect on the mechanical properties. it is time consuming and expensive to carry out a large number
Analytical and FE analysis confirmed that the modified unit of experiments to comprehensively investigate the effects of
cell (by adding narrow ribs) of re-entrant honeycomb led to a geometrical parameters. Moreover, sometimes it is almost im-
significant enhancement of Young’s modulus [36]. However, possible to obtain essential information from experiment such
as stress contour of a structure. Therefore, numerical simula-
tion is an efficient method and has been widely employed to
investigate the effects of various geometrical parameters on
the mechanical performance of structures [37–42].
The effect of the geometrical parameters such as rib length
and rib thickness of the unit cell of the new designs on the in-
plane properties was investigated. Finite element results
showed that the in-plane stiffness of the new designs was
improved significantly compared to the basic re-entrant hon-
eycomb. In this paper, finite element (FE) simulation of the
RCA structure is conducted. Based on the manufactured RCA
samples (aluminium alloy 6061), full-scale FE models using
ABAQUS 6.11–2 are developed and verified using previous
experimental results [27]. Subsequently, a parametric study is
performed to investigate the effects of geometrical parameters
on the mechanical properties and auxetic behaviour of the
RCA structure. Three dimensionless groups of geometrical

Fig. 4 Sketches of the simulated unit cells of the RCA structure with
various values of a α (β = 0.16, λ = 2, θ0 = 30°); b β (α = 3.27, λ = 2, Fig. 6 Comparison between a experimental and b FE simulated
θ0 = 30°); c λ (α = 3.27, β = 0.16, θ0 = 30°); d inclination angle θ0 (α = deformation patterns at 116 mm axial displacement when a RCA
3.27, β = 0.16, λ = 2) sample is loaded in the Y direction
Int J Adv Manuf Technol (2018) 99:2425–2440 2429

Fig. 7 FE simulated deformation


of the walls and cylinders of a
RCA structure when it is loaded
in the Y direction: a trichiral
topology; b anti-trichiral
topology; c manufacturing defects
(un-machined material) from the
waterjet cutting process

parameters (α = L1/r ratio, β = t/r ratio and λ = H/L1 ratio) and adjacent cylinders attach to the connecting green wall shown
one parameter (inclination angle, θ0) are examined. in Fig. 1f on the opposite sides (trichiral topology).

2.2 Materials and experimental tests


2 Experiments
Two aluminium RCA samples were fabricated from widely
2.1 The geometrical parameters of a RCA structure used aluminium alloy (6061) plate by an abrasive waterjet
machine WJM (MAXIEM 1515) and shown in Fig. 2a, b.
A unit cell of a RCA structure is shown in Fig. 1f. In this unit The stress–strain curves of aluminium alloy (6061) (parent
cell, L1, L2 and H are the original lengths of the walls. material of RCA samples) were measured in reference [27].
Inclination angle (θ0) is the original angle between the vertical Figure 2c shows the corresponding true stress–strain curves of
wall (H) and the inclined wall (L1). t1 is the in-plane thickness four repeated tests according to the ASTM standards E8/E8M
of walls and hollow cylinders. t2 (not shown) is the out-of- −15a. Both RCA samples had the same unit cell geometrical
plane thickness of the structure. Each cylinder has an outer parameters: r = 2.75 mm, H = 18 mm, L1 = 9 mm, L2 = 9 mm,
radius of (r) and it is tangentially attached by three walls. Two θ0 = 30°, t1 = 0.45 mm, t2 = 5 mm. Previous studies [27,
adjacent cylinders attach to the connecting red walls shown in 43–47] manifested that bulk properties could be achieved
Fig. 1f on the same side (anti-trichiral topology) while the two when a structure had a minimum of five unit cells in each
2430 Int J Adv Manuf Technol (2018) 99:2425–2440

Fig. 8 FE simulated deformation history of RCA structure under uniaxial tensile loading in the Y direction: a RCA sample; b unit cell at the centre of the
sample

direction. Therefore, the RCA samples were designed to have material properties for aluminium. Walls and cylinders of the
5 × 5 unit cells and their overall in-plane dimensions were RCA structures were simulated by standard shell element
190 mm × 124 mm. Uniaxial tensile tests were conducted (S4R) with four nodes. The element size, 0.45 mm, was de-
[27] using Zwick Roell machine (Zwick/Z010) equipped with termined from a mesh convergence analysis. Twelve points
a load cell of 10 kN. The bottom side of each sample was fixed (marked in blue in Fig. 3a) were selected in the middle area
(i.e., could not move along the loading direction) while the top and traced to calculate the axial and lateral displacements.
side moved at a constant speed of 0.05 mm/s with the machine When a RCA sample was simulated to be loaded in the X
crosshead. A lubricant was used to reduce the friction between and Y directions as shown in Fig. 3b, c, a uniform displace-
contacted surfaces. The fracture occurred at axial displace- ment was assigned to the nodes on the top edge of the models.
ment of 60 and 116 mm, respectively, when RCA samples Whilst nodes on the bottom edge of the models could move
were loaded in the X and Y directions. freely along the lateral direction, their other translational and
rotational degrees of freedom were constrained.
The unit cell of the RCA structure has six in-plane geomet-
3 Finite element model for uniaxial tensile rical parameters (H, L1, L2, r, t1 and θ0) as shown in Fig. 1f. It
loading has been found that mechanical properties of the RCA struc-
ture are not sensitive to the horizontal walls (L2), and there-
3.1 FE models fore, L2 remains constant. In order to reduce the variables
studied, dimensionless parameters are proposed. According
Finite element (FE) models were developed using ABAQUS/ to reference [48], dimensionless groups must be independent.
IMPLICIT software (Version 6.11–2). The elastic-plastic ma- Therefore, four geometrical parameters (H, L1, r and t1) can be
terial model was employed and material properties used in represented by three dimensionless groups (α = L1/r, β = t1/r
FEA are listed in Table 2. A typical true stress–strain curve and λ = H/ L1). For each loading direction, five values of α, β,
(Test 1) shown in Fig. 2c was input into the FE model as the λ and inclination angle (θ0) are selected to be studied as shown
Int J Adv Manuf Technol (2018) 99:2425–2440 2431

Fig. 9 FE simulated deformation history of RCA structure under uniaxial tensile loading in the X direction a RCA sample; b unit cell at the centre of the
sample

in Table 3. When the effect of any parameter is examined, the validated FE models to investigate the effects of loading di-
other parameters are fixed as those of the samples in rections and geometrical parameters on the mechanical perfor-
experiments (i.e. α = L1/r = 9/2.75 = 3.27, β = t1/r = 0.45/ mance of the RCA structure.
2.75 = 0.16, λ = H/L1 = 18/9 = 2, θ0 = 30°). For clarification,
Fig. 4 shows the sketches of simulated unit cells of
the RCA structure. 4 Results

3.2 Validation of the FE models 4.1 Deformation patterns

In order to validate the FE models, an FE model was setup Typical cylinders and walls of the RCA structure at 150 mm
which had the same dimensions of the manufactured sample axial displacement when it is loaded in the Y direction are
mentioned in Sect. 2.2. The simulated force–displacement shown in Fig. 7. It can be seen that cylinders having trichiral
curves are compared with the experimental results and shown topology (two adjacent cylinders connecting to a wall on the
in Fig. 5. The FE and experimental curves match very well opposite sides) rotate in the same direction leading to a full-
when the RCA is loaded in the Y direction. Moreover, when wave shape deformation of the inclined walls (L1) as shown in
the RCA is loaded in the X direction, FE and experimental Fig. 7a. In contrast, cylinders having anti-trichiral topology
curves are similar for small displacement (e.g. from 0 to (two adjacent cylinders connecting to a wall on the same
43 mm). Experimentally measured force fluctuates at 43, 48 sides) rotate in the opposite directions leading to a half-wave
and 52 mm, respectively. This was due to the unexpected shape deformation of the vertical (H) and horizontal (L2) walls
movement of fixtures employed to hold physical samples, as shown in Fig. 7b.
which was evidenced by videos taken in the experimental It is worth noting that in the current study, the FE models
tests. Moreover, typical experimental and FE simulated defor- for RCA were established with the perfect geometrical param-
mations of a RCA sample when it is loaded in the Y direction eters as designed. However, for the manufactured samples, the
are shown in Fig. 6. Very similar deformation patterns are geometrical parameters were slightly different from the de-
observed. Such comparisons manifest that the FE models are signed ones because of the imperfect manufacturing process.
valid. Further parametric study can be conducted using the Moreover, there was un-machined material left at the
2432 Int J Adv Manuf Technol (2018) 99:2425–2440
Int J Adv Manuf Technol (2018) 99:2425–2440 2433

ƒFig. 10 FE simulated stress–strain curves of the RCA structure loaded in the X direction than that in the Y direction. When a RCA
the X (right figures) and Y (left figures) directions with various values of: a, sample is loaded in the Y direction, the stress–strain curves
b α (L/r); c, d β (t/r); e, f λ (H/L1); g, h initial inclination angle (θ0)
broadly have three stages: the elastic stage, plateau stage and
densification stage. The third (i.e. densification) stage when
connections of cylinders-walls (indicated by the yellow the stress increases sharply is more evident at larger values of
dashed lines in Fig. 7c) due to limitation of the waterjet cutting α and θ0, which is similar to conventional cellular materials
process. Manufacturing defects might affect the deformation under compression [49]. However, when a RCA sample is
of the manufactured samples. loaded in the X direction, the stress increases monolithically
Figure 8a shows the simulated deformation history of the with strain with no evident plateau region.
RCA sample when it is loaded in the Y direction. Figure 8b
displays corresponding the axial and lateral dimensions of the 4.3 Poisson’s ratios
original and deformed unit cell that is allocated in the middle
area of the RCA sample. In general, the RCA sample expands In order to investigate the auxetic behaviour of the RCA struc-
uniformly in both the axial and lateral directions under uniax- ture, the instantaneous displacements in both directions are
ial tensile loading. When the RCA sample is loaded, the ten- calculated based on the displacements of the 12 points men-
sile force causes stretching and bending of the vertical walls tioned in Sect. 3.1 and shown in Fig. 3. The average displace-
(H) and the inclined walls (L1) simultaneously. Moreover, the ment between the points in the top and bottom rows (1 and 9;
tensile force generates torques on the cylinders leading to their 2 and 10; 3 and 11; 4 and 12) is taken as the axial displace-
in-plane rotation. This rotation introduces additional bending ment, while the average displacement between those in the left
to the walls. As the tension progresses, when the axial dis- and right rows (1 and 4; 5 and 6; 7 and 8; 9 and 12) is
placement is approximately 125 mm, the inclined walls (L1) taken as the lateral displacement. True strains and Poisson’s
become horizontal (shown in Fig. 8a) indicating the maximum ratio are then calculated accordingly by the equations given in
lateral displacement. With further tension, the structure ex- reference [27].
pands axially and shrink laterally, similar to a conventional The effects of the examined parameters on the Poisson’s
hexagonal honeycomb. Comparing with honeycombs, RCA ratio of the RCA structure are shown in Fig. 11. In general,
can experience much larger axial deformation. For example, at Poisson’s ratio of the RCA structure changes from negative to
180 mm axial displacement, the axial displacement of a unit positive when a RCA sample is tensioned in the Y direction as
cell is 2.4 times of its original dimension. shown in Fig. 11a, c, e, g. This indicates the expansion and
On the other hand, when the RCA sample is loaded in the X then the shrinkage in the lateral direction, which echoes the
direction, similar deformation pattern is observed, except no previous observation in Sect. 4.1. In the case of loading in the
stretching of the “H” walls but stretching of the “L2” walls. X direction, the auxetic feature is present during the entire
The RCA sample expands in both directions due to the rota- tension process. Poisson’s ratio first decreases gradually then
tion of cylinders and the deflection of the inclined walls (L1). increases sharply with the axial strain as shown in Fig. 11b, d,
When the original inclined walls (L1) become vertical (shown f, h. The minimum value (i.e. the maximum absolute value) of
in Fig. 9), they experience stretching without bending, similar Poisson’s ratio is corresponding to the moment when the in-
to the “L2” walls, and the lateral displacement reaches the clined walls (L1) become vertical as shown in Fig. 9.
maximum. Moreover, the RCA sample loaded in the X direc- When a RCA sample is loaded in the Y direction, the effect
tion undergoes smaller axial but larger lateral displacements of α is shown in Fig. 11a. The slope of Poisson’s ratio vs. axial
compared with that RCA sample loaded in the Y direction. strain curves increases with α. Larger values of α (i.e. longer
L1 or smaller r) result in greater slope of Poisson’s ratio.
4.2 Stress–strain curves Figure 11a shows that the RCA displays auxetic behaviour
up to approximately 0.45 axial strain and Poisson’s ratio
Nominal stress (σ) is calculated by dividing the total nodal changes from − 1.4 to 3.5 when α = 9 (the largest value of α
reaction forces at the bottom of the FE models by the cross- studied). The RCA also demonstrates auxetic behaviour up to
sectional areas. While engineering strain (ɛ) is the ratio of the approximately 0.8 axial strain and Poisson’s ratio changes
instantaneous change in displacement to the original length in from − 0.75 to 0.4 when α = 2.3 (the smallest value of α
the axial direction. Figure 10 presents the effects of the exam- studied). On the other hand, when the RCA is loaded in the
ined parameters on the engineering stress–strain curves. X direction as shown in Fig. 11b, a similar influence of α is
Generally, the RCA structure shows smooth stress–strain firstly observed (i.e. the larger the α, the larger the slope of
curves without any fluctuation during the tensile testing. Poisson’s ratio) until the intersection point of Poisson’s ratio
Geometrical parameters affect the stress–strain curves signif- curves at approximately 0.27 axial strain. However, the
icantly. Details will be discussed in Sect. 5.1. Moreover, stress Poisson’s ratio decreases to the minimum (in the range from
is higher and strain is smaller when a RCA sample is loaded in − 3.25 to − 3.5) at approximately 0.34 axial strain. Beyond
2434 Int J Adv Manuf Technol (2018) 99:2425–2440
Int J Adv Manuf Technol (2018) 99:2425–2440 2435

ƒFig. 11 FE simulated Poisson’s ratio curves of the RCA structure loaded in when the RCA structure is loaded in the X and Y, respectively.
the X (right figures) and Y (left figures) directions with various values of: a, Afterward, the inclined walls (L1) deform plastically when the
b α (L/r); c, d β (t/r); e, f λ (H/L1); g, h initial inclination angle (θ0)
actual stress in walls is beyond the yield stress of the parent
material (aluminium alloy 6061). The inclined walls deform
easily due to relatively long arm of plastic bending moment
this point, Poisson’s ratio increases dramatically with axial (APBM) when φ and ψ are larger than 45° as shown in
strain. Fig. 12a. The APBM continues to increase to its maximum
The effect of β on the Poisson’s ratio is shown in Fig. 11c, value when the inclined walls (L1) become horizontal.
d. It can be seen that Poisson’s ratio is insensitive to β (in the Subsequently, the APBM decreases, and a larger tensile force
range studied) when RCA samples are loaded in both the X is required, resulting in a steep rising of stress (in some cases
and Y directions. sharp increases).
The effect of λ on the Poisson’s ratio is shown in the Conversely, when the RCA structure is loaded in the X
Fig. 11e, f. When RCA samples are loaded in the Y direction, direction, the relatively short APBM when φ is larger than
the various values of λ do not affect the Poison’s ratio up to 45° as shown in Fig. 12b requires larger tensile force for the
0.25 axial strain as shown in Fig. 11e. Beyond this axial strain, plastic deformation of the inclined walls (L1). With the prog-
with the increase of λ, the slope of the Poisson’s ratio vs. axial ress of loading, the APBM continues to decrease (i.e. larger
strain curve increases. The largest value of λ (λ = 2.6) results force is required) until it becomes zero when the inclined walls
in the largest slope, and the Poison’s ratio increases from − 0.9 (L1) become vertical (as shown in Fig. 12b). Eventually, the
to 2.25, while the smallest value of λ (λ = 1.8) results in the remarkable increases of the stress as shown on right side of
smallest slope, and the Poison’s ratio increases from − 0.9 to Fig. 10b, d, f, h is an evidence of the stretching of the inclined
0.6. When RCA samples are loaded in the X direction, the walls (L1) simultaneously with the (L2) walls.
effect of λ is shown in Fig. 11f. The Poisson’s ratio remains The FE models enable the observation of the local stress
negative. When λ is the smallest, λ = 1.8, the minimum concentration in unit cells as well. From Fig. 12, it can be seen
Poison’s ratio of the RCA structure is − 4, whose absolute that the maximum local stress occurs in the interconnecting
value is the largest. When λ is the largest, λ = 2.6, the mini- areas where the inclined (L1), horizontal (L2) and vertical (H)
mum Poison’s ratio is approximately − 2, whose absolute val- walls join cylinders.
ue is the smallest.
The effect of initial inclination angle (θ0) is shown in 5.1.2 Effects of parameters (α, β, λ and θ0)
Fig. 11g, h, indicating that θ0 has a dominant effect on the
auxetic behaviour of the RCA structure. When the RCA sam- Figure 10a, b shows that stress level increases when α in-
ple is loaded in the Y direction, the maximum axial strain at creases. In our previous experimental work [27], it has been
which the auxetic feature is present, increases with the de- stated that the inclined walls (L1) are subjected to two opposite
crease of θ0. At a small value of θ0 (θ0 = 30°), the auxetic pairs of bending moments. The first pair results from the ten-
feature remains up to 0.75 axial strain, and Poisson’s ratio sile force applied, while the second pair results from the in-
increases monolithically from − 0.9 to 0.6, while at a large plane rotation of cylinders. Therefore, as the sum of the two
value of θ0 (θ0 = 70°), the auxetic feature remains only up to pairs of the bending moments, the inclined walls deform sym-
0.2 axial strain, and Poisson’s ratios increase from − 0.3 to 5. metrically in full-wave shape as shown in Fig. 7. The smaller
When the RCA sample is loaded in the X direction, similar the α (i.e. larger the cylinder’s radius), the larger the bending
influence of θ0 is observed. moment resulting from the rotation, the smaller the net bend-
ing moment resulting from the sum of two pairs, and hence the
lower the force/stress. Additionally, Fig. 13 displays the de-
5 Discussions formation patterns of the inclined walls (L1) at 150 mm axial
displacement when the RCA structure are loaded in the Y
5.1 Stress–strain curves direction. It can be seen that at the largest value of α (α = 9),
the full-wave deformation is more evident, while it is not
5.1.1 Effect of loading direction apparent at the smallest value of α (α = 2.3). It is known that
a full-wave deformation is a higher energy dissipation mode
In the case of loading in the Y direction, the unit cells initially [49] than other deformation modes.
deform elastically via bending of the inclined walls (L1) and Figure 10c, d demonstrates that β has a significant effect on
stretching of the vertical walls (H). The accompanying change the stress–strain curves. It is clear that the stresses in both
in the inclination angle is expected when the inclined walls loading directions increase as β increases. At a larger values
(L1) are deflected. Therefore, a new inclination angles φ and of β (i.e. thicker walls and cylinders), the resistance to the
ψ are used instead of the original angle θ0 as shown in Fig. 12 plastic collapse (plastic bending and stretching) increases.
2436 Int J Adv Manuf Technol (2018) 99:2425–2440

To clarify the effect of λ shown in Fig. 10e, f, the highest APBM. In the case of loading in the X direction, Fig. 10h
and lowest values of λ are used as examples. Since the shows that θ0 has a significant influence on the stress–strain
cylinder’s radius is constant when λ is varied, the bend- curves and a similar effect on the axial strain to that observed
ing moment resulting from cylinder rotation is constant, in the case of loading in the Y direction. Figure 14b shows the
while the bending moment resulting from the tensile effect of θ0 on the deformed RCA structure at 8 mm axial
force is different due to various lengths of APBM. For displacement. When θ0 is larger, APBM is smaller and re-
the same applied force on the RCA structure that has quired force/stress is higher.
λ = 1.8, the resulting bending moment on the inclined
walls (L1) is much larger than that for λ = 2.6 due to a longer 5.2 Poisson’s ratio
APBM. Therefore, the longer the L1 (i.e. the lowest λ), the
easier the deformation. RCA structure combines trichiral, anti-trichiral and re-entrant
Figure 14a shows the effect of θ0 on the deformation pat- topologies. Therefore, its Poisson’s ratio is related to those
tern of the RCA structure when it is loaded in the Y direction topologies.
at 36 mm axial displacement. At a large value of θ0 (θ0 = 70°), The effect of α shown in Fig. 11a, b can be explained as
the inclined walls (L1) become horizontal at a small displace- follows. When RCA is loaded in the Y direction, at large
ment. Afterward, the unit cells become slender and deform in values of α (long L1 or small r), the chiral effect is less sig-
a similar manner to a conventional honeycomb. In other nificant and the RCA behaves similar to a re-entrant honey-
words, beyond the point when the inclined walls (L1) comb. The Poisson’s ratio of RCA is similar to that of a re-
become horizontal, a larger force is required for further entrant honeycomb, which transits from negative to positive
deformation. The sharp increase in stress observed in [27, 31]. In contrast, at small values of α (short L1 or large r),
Fig. 10g is attributed to the remarkable decrease of the the chiral topology plays a dominant role, which results in

Fig. 12 Loading of the inclined walls (L1), deformation patterns and stress (GPa) concentration of walls at 50 and 100% axial strains when the RCA
structure is loaded in the a Y direction; b X direction
Int J Adv Manuf Technol (2018) 99:2425–2440 2437

Fig. 13 Deformation patterns and stress contours of the inclined walls (L1) at a α = 2.3; b α = 6

reducing the slope of Poisson’s ratio and hence extending the has small negative Poisson’s ratio when it is loaded in the X
period of auxetic feature of the tensioned RCA. direction [31, 50]. In the case of RCA structure, the rotation of
On the other hand, when the RCA structure is loaded in the the cylinders introduces additional auxetic effect to that com-
X direction, Poisson’s ratio remains negative in the entire ten- ing from the deflection of the inclined walls (L1). From
sile process. It has been stated that the re-entrant honeycomb Fig. 11b, it can be seen that the smaller the α (i.e. the larger

Fig. 14 The deformation patterns of a RCA unit cell as a function of initial cell angle (θ0) when it is loaded in the a Y direction at 36 mm axial
displacement; b X direction at 8 mm axial displacement
2438 Int J Adv Manuf Technol (2018) 99:2425–2440

the r), the minimum the value (i.e. the maximum absolute the Poisson’s ration were also similar for the RCA structure
value) of Poisson’s ratio obtained. loaded in the Y direction, i.e. Poisson’s ratio increased mono-
The effect of λ observed in Fig. 11e, f can be attributed to tonically with axial strain and its slope increased with α, λ and
the ratio between the vertical (H) and the inclined (L1) walls, θ0. On the other hand, when the RCA structure was loaded in
while the chiral effect is similar for all values of λ. At small the X direction, Poisson’s ratio first decreased and then in-
values of λ (short H or long L1), the re-entrant effect is creased. The inflexion point was independent of α, in-
more significant due to the relative long length of the creased with λ and decreased with θ0. Poisson’s ratio
inclined walls (L1) compared with the vertical walls (H). was independent of β irrespecitve of whether the RCA
Therefore, the period of the auxetic effect (i.e. up to the structure was loaded in either the X or Y direction. This
point when inclined walls (L 1) become horizontal or study provided useful information for the optimal design
vertical) extends. For example, when λ = 1.8 and 2.6, the axial of unit cell of RCA structure. Further analytical models
strains are 0.78 and 0.53, respectively, as shown in Fig. 11e are expected to be developed to theoretically interpret the
when the inclined walls become horizontal (i.e. RCA with influence of geometrical parameters on the mechanical perfor-
negative Poisson’s ratio). mance of the RCA structure.
The effect of the inclination angle (θ0) on Poisson’s ratio is
shown in Fig. 11g, h. The increase of the inclination angle Acknowledgments The first author would like to thank the Republic of
Iraq, Ministry of Higher Education and Scientific Research, and the
leads to a decrease in the axial strain and an increase in
Higher Committee of Education Development in Iraq (HCED) for the
Poisson’s ratio. The position of the deflected inclined walls financial support of a postgraduate research award.
(L1) corresponding to the axial strain as shown in Fig. 14 is the
main reason for this influence. In the case of loading in Publisher’s Note Springer Nature remains neutral with regard to jurisdic-
the Y direction, the larger the initial angle, the sooner tional claims in published maps and institutional affiliations.
the inclined walls change to horizontal (i.e. ψ = 90°)
after a small axial strain as shown in Fig. 14a (i.e.
Poisson’s ratio transits from negative to positive at a References
small axial strain). Then, the unit cells turn into con-
1. Greaves G, Greer A, Lakes R, Rouxel T (2011) Poisson’s ratio and
ventional hexagonal cells earlier as compared with other
modern materials. Nat Mater 10(11):823–837. https://doi.org/10.
initial angles. On the other hand, in the case of loading in the X 1038/nmat3134
direction, the larger the initial angle, the sooner the inclined 2. Bezazi A, Scarpa F (2009) Tensile fatigue of conventional and
walls change to vertical (i.e. φ = 90°) after small axial strain as negative Poisson’s ratio open cell PU foams. Int J Fract 31(3):
shown in Fig. 14b. 488–494. https://doi.org/10.1016/j.ijfatigue.2008.05.005
3. Li D, Dong L, Lakes R (2013) The properties of copper foams with
negative Poisson's ratio via resonant ultrasound spectroscopy. Phys
Status Solidi B 250(10):1983–1987. https://doi.org/10.1002/pssb.
6 Conclusions 201384229
4. Almgren RF (1985) An isotropic three-dimensional structure with
This paper numerically investigates the effects of loading di- Poisson’s ratio =−1. J Elast 15(4):427–430. https://doi.org/10.1007/
BF00042531
rection and geometrical parameters of the re-entrant chiral 5. Lakes R (1987) Foam structures with a negative Poisson’s ratio. Sci
auxetic (RCA) structure on its tensile properties, i.e. stress- 235(4792):1038–1040. https://doi.org/10.1126/science.235.4792.
strain curves and Poisson’s ratio. The parameters studied are 1038
α = L1/r, λ = H/L1, β = t/r and initial inclination angle θ0. 6. Evans K (1991) Auxetic polymers: a new range of materials.
Results showed that the RCA structure predominantly de- Endeavour 15(4):170–174. https://doi.org/10.1016/0160-9327(91)
90123-S
formed by bending and stretching of walls as well as rotation 7. Choi J, Lakes R (1996) Fracture toughness of re-entrant foam ma-
of cylinders. The inclined walls (L1) deformed into a full- terials with a negative Poisson’s ratio: experiment and analysis. Int J
wave shape, while the vertical (H) and horizontal (L2) walls Fract 80(1):73–83. https://doi.org/10.1007/BF00036481
deformed into a half-wave shape. It was also found that the 8. Spadoni A, Ruzzene M, Scarpa F (2005) Global and local linear
buckling behavior of a chiral cellular structure. Phys Status Solidi B
load direction had a significant effect on the tensile properties.
242(3):695–709. https://doi.org/10.1002/pssb.200460387
High stress level, low strain and relatively large negative 9. Javadi A, Faramarzi A, Farmani R (2012) Design and optimization
Poisson’s ratio were observed when the RCA structure was of microstructure of auxetic materials. Eng Comput 29(3):260–276.
loaded in the X direction. While low stress, high strain and https://doi.org/10.1108/02644401211212398
transition of Poisson’s ratio from negative to positive were 10. Photiou D, Sarris E, Constantinides G (2016) On the conical inden-
tation response of elastic auxetic materials: effects of Poisson's ra-
observed when the RCA structure was loaded in the Y direc-
tio, contact friction and cone angle. Int J Solids Struct 81:33–42.
tion. The effects of α, β, λ and θ0 on the stress–strain curves https://doi.org/10.1016/j.ijsolstr.2015.10.020
were found to be similar—the larger the parameter, the greater 11. Rossiter J, Takashima K, Scarpa F, Walters P, Mukai T (2014)
the stress and the slope of stress. The effects of α, λ and θ0 on Shape memory polymer hexachiral auxetic structures with tunable
Int J Adv Manuf Technol (2018) 99:2425–2440 2439

stiffness. Smart Mater Struct 23(4):45007. https://doi.org/10.1088/ Status Solidi B Res 251(2):367–374. https://doi.org/10.1002/pssb.
0964-1726/23/4/045007 201384256
12. Evans K, Alderson A, Christian F (1995) Auxetic two-dimensional 29. Yang L, Harrysson O, Cormier D, West H, Gong H, Stucker B
polymer networks: an example of tailoring geometry for specific (2015) Additive manufacturing of metal cellular structures: design
mechanical properties. J Chem Soc Faraday Trans 91:2671–2680. and fabrication. JOM 67(3):608–615. https://doi.org/10.1007/
https://doi.org/10.1039/ft9959102671 s11837-015-1322-y
13. Wan H, Ohtaki H, Kotosaka S, Hu G (2004) A study of negative 30. Yang L, Harrysson O, West H, Cormier D (2015) Shear properties
Poisson’s ratios in auxetic honeycombs based on a large deflection of the re-entrant auxetic structure made via electron beam melting.
model. Eur J Mech A Solids 23(1):95–106. https://doi.org/10.1016/ Solid Free Fabr Symp 1394–409
j.euromechsol.2003.10.006 31. Zhang J, Lu G, Wang Z, Ruan D, Alomarah A, Durandet Y (2017)
14. Alderson A, Alderson K, Attard D, Evans K, Gatt R, Grima J, Large deformation of an auxetic structure in tension: experiments
Miller W, Ravirala N, Smith C, Zied K (2010) Elastic constants and finite element analysis. Compos Struct 184:92–101. https://doi.
of 3-, 4- and 6-connected chiral and anti-chiral honeycombs subject org/10.1016/j.compstruct.2017.09.076
to uniaxial in-plane loading. Compos Sci Technol 70(7):1042– 32. Subramani P, Rana S, Ghiassi B, Fangueiro R, Oliveira D,
1048. https://doi.org/10.1016/j.compscitech.2009.07.009 Lourenco P, Xavier J (2016) Development and characterization of
15. Spadoni A, Ruzzene M (2012) Elasto-static micropolar behavior of novel auxetic structures based on re-entrant hexagon design pro-
a chiral auxetic lattice. J Mech Phys Solids 60(1):156–171. https:// duced from braided composites. Compos Part B 93:132–142.
doi.org/10.1016/j.jmps.2011.09.012 https://doi.org/10.1016/j.compositesb.2016.02.058
16. Prall D, Lakes R (1997) Properties of a chiral honeycomb with a 33. Attard D, Manicaro E, Gatt R, Grima J (2009) On the properties of
poisson’s ratio of −1. (longer version). Int J Mech Sci 39(3):305– auxetic rotating stretching squares. Phys Phys Status Solidi B
314. https://doi.org/10.1016/S0020-7403(96)00025-2 246(9):2045–2054. https://doi.org/10.1002/pssb.200982035
17. Álvarez Elipe JC, Díaz Lantada A (2012) Comparative study of 34. Grima J, Farrugia P, Gatt R, Attard D (2008) On the auxetic prop-
auxetic geometries by means of computer-aided design and engi- erties of rotating rhombi and parallelograms: a preliminary investi-
neering. Smart Mater Struct 21:105004. https://doi.org/10.1088/ gation. Phys Status Solidi B Basic Res 245(3):521–529. https://doi.
0964-1726/21/10/105004 org/10.1002/pssb.200777705
18. Hou Y, Neville R, Scarpa F, Remillat C, Gu B, Ruzzene M (2014) 35. Slann A, White W, Scarpa F, Boba K, Farrow I (2015) Cellular
Graded conventional-auxetic Kirigami sandwich structures: flat- plates with auxetic rectangular perforations. Phys Status Solidi B
wise compression and edgewise loading. Compos Part B 59:33– 252(7):1533–1539. https://doi.org/10.1002/pssb.201451740
42. https://doi.org/10.1016/j.compositesb.2013.10.084
36. Lu Z, Li X, Yang Z, Xie F (2016) Novel structure with negative
19. Levy O, Krylov S, Goldfarb I (2006) Design considerations for
Poisson’s ratio and enhanced Young’s modulus. Compos Struct
negative Poisson ratio structures under large deflection for MEMS
138:243–252. https://doi.org/10.1016/j.compstruct.2015.11.036
applications. Smart Mater Struct 15(5):1459–1466. https://doi.org/
37. Besant T, Davies G, Hitchings D (2001) Finite element modelling
10.1088/0964-1726/15/5/035
of low velocity impact of composite sandwich panels. Compos A
20. Alderson A, Rasburn J, Ameer-Beg S, Mullarkey P, Perrie W,
Appl Sci Manuf 32(9):1189–1196. https://doi.org/10.1016/S1359-
Evans K (2000) An Auxetic filter: a Tuneable filter displaying
835X(01)00084-7
enhanced size selectivity or Defouling properties. Ind Eng Chem
38. Hou X, Deng Z, Zhang K (2016) Dynamic crushing strength anal-
Res 39:654–665. https://doi.org/10.1021/ie990572w
ysis of Auxetic honeycombs. Acta Mech Solida Sin 29(5):490–501.
21. Li S, Hassanin H, Attallah M, Adkins N, Essa K (2016) The devel-
https://doi.org/10.1016/S0894-9166(16)30267-1
opment of TiNi-based negative Poisson’s ratio structure using se-
lective laser melting. Acta Mater 105:75–83. https://doi.org/10. 39. Alkhader M, Vural M (2008) Mechanical response of cellular
1016/j.actamat.2015.12.017 solids: role of cellular topology and microstructural irregularity.
22. Grima J, Attard D, Ellul B, Gatt R (2011) An improved analytical Int J Eng Sci 46(10):1035–1051. https://doi.org/10.1016/j.
model for the elastic constants of auxetic and conventional hexag- ijengsci.2008.03.012
onal honeycombs. Cell Polym 30(6):287–310 40. Deqiang S, Weihong Z, Yanbin W (2010) Mean out-of-plane dy-
23. Scarpa F, Panayiotou P, Tomlinson G (2000) Numerical and exper- namic plateau stresses of hexagonal honeycomb cores under impact
imental uniaxial loading on in-plane auxetic honeycombs. J Strain loadings. Compos Struct 92(11):2609–2621. https://doi.org/10.
Anal Eng Des 35(5):383–388. https://doi.org/10.1243/ 1016/j.compstruct.2010.03.016
0309324001514152 41. Asprone D, Auricchio F, Menna C, Morganti S, Prota A, Reali A
24. Alderson A, Alderson K, Chirima G, Ravirala N, Zied K (2010) (2013) Statistical finite element analysis of the buckling behavior of
The in-plane linear elastic constants and out-of-plane bending of 3- honeycomb structures. Compos Struct 105:240–255. https://doi.
coordinated ligament and cylinder-ligament honeycombs. Compos org/10.1016/j.compstruct.2013.05.014
Sci Technol 70(7):1034–1041. https://doi.org/10.1016/j. 42. Xu S, Ruan D, Beynon J (2014) Finite element analysis of the
compscitech.2009.07.010 dynamic behavior of aluminum honeycombs. Int J Comput
25. Zied K, Osman M, Elmahdy T (2015) Enhancement of the in-plane Methods 11:1344001. https://doi.org/10.1142/
stiffness of the hexagonal re-entrant auxetic honeycomb cores. Phys S0219876213440015
Status Solidi B 252(12):2685–2692. https://doi.org/10.1002/pssb. 43. Onck P, Andrews E, Gibson L (2001) Size effects in ductile cellular
201552164 solids. Part I: modeling. Int J Mech Sci 43(3):681–699. https://doi.
26. Fu M, Chen Y, Hu L (2017) A novel auxetic honeycomb with org/10.1016/S0020-7403(00)00042-4
enhanced in-plane stiffness and buckling strength. Compos Struct 44. Andrews E, Gioux G, Onck P, Gibson L (2001) Size effects in
160:574–585. https://doi.org/10.1016/j.compstruct.2016.10.090 ductile cellular solids. Part II: experimental results. Int J Mech Sci
27. Alomarah A, Ruan D, Masood S, Sbarski I, Faisal B (2018) An 43(3):701–713. https://doi.org/10.1016/S0020-7403(00)00043-6
investigation of in-plane plastic properties of re-entrant chiral 45. Dai G, Zhang W (2008) Size effects of basic cell in static analysis of
auxetic structure. Int J Adv Manuf Technol 96:2013–2029. https:// sandwich beams. Int J Solids Struct 45(9):2512–2533. https://doi.
doi.org/10.1007/s00170-018-1605-x org/10.1016/j.ijsolstr.2007.12.007
28. Pozniak A, Wojciechowski K (2014) Poisson’s ratio of rectangular 46. Yang L, Harrysson O, West H, Cormier D (2015)
anti-chiral structures with size dispersion of circular nodes. Phys Mechanical properties of 3D re-entrant honeycomb auxetic
2440 Int J Adv Manuf Technol (2018) 99:2425–2440

structures realized via additive manufacturing. Int J Solids Struct 49. Gibson LJ, Ashby MF (1997) Cellular solids: structure and proper-
69–70:475–490. https://doi.org/10.1016/j.ijsolstr.2015.05.005 ties, 2nd edn. Cambridge University Press, Cambridge. https://doi.
47. Cai K, Luo J, Ling Y, Wan J, Qin Q (2016) Effects of size and org/10.1017/CBO9781139878326
surface on the auxetic behaviour of monolayer graphene kirigami. 50. Alomarah A, Zhang J, Ruan D, Masood S, Lu G (2017)
Sci Rep 6:35157. https://doi.org/10.1038/srep35157 Mechanical properties of the 2D re-entrant honeycomb made
48. Lu G, Yu TX (2003) Energy Absorption of Structures and via direct metal printing. IOP Conference Series: IOP Conf
Materials. Woodhead Publishing Limited, Cambridge, UK. Ser Mater Sci Eng 229(1):012038. https://doi.org/10.1088/
https://doi.org/10.1533/9781855738584.backmatter 1757-899X/229/1/012038

You might also like