You are on page 1of 25

Mechanics of Auxetic Materials

23
Hyeonho Cho, Dongsik Seo, and Do-Nyun Kim

Contents
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 734
2 Deformation Mechanisms of Auxetic Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 735
2.1 Re-entrant Mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 736
2.2 Rotating Unit Mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 738
2.3 Chiral Mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 740
2.4 Fibril/Nodule Mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 741
2.5 Miura-Folded Mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 743
2.6 Buckling-Induced Mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 744
2.7 Helical Auxetic Yarn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 745
2.8 Crumpled Mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 746
3 Relationship Between the Deformation Mechanism and the Material Properties . . . . . . . . . 747
4 Expected Properties of Auxetic Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 748
4.1 Synclastic Curvature in Bending . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 748
4.2 Variable Permeability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 749
4.3 High Shear Stiffness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 749
4.4 Enhanced Indentation Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 750
4.5 High Fracture Toughness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 751
4.6 Increased Energy Absorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 752
5 Potential Applications of Auxetic Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 753
6 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 755
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 755

Abstract
Poisson’s ratio is a mechanical property that represents the lateral behavior of
materials under an axial load. In contrast to typical natural materials with a
positive Poisson’s ratio, auxetic materials have a negative Poisson’s ratio (NPR)

H. Cho (*) · D. Seo · D.-N. Kim


Department of Mechanical and Aerospace Engineering, Seoul National University, Seoul, Republic
of Korea
e-mail: ttp85275@snu.ac.kr; a334krseods@snu.ac.kr; dnkim@snu.ac.kr

# Springer Nature Singapore Pte Ltd. 2019 733


C.-H. Hsueh et al. (eds.), Handbook of Mechanics of Materials,
https://doi.org/10.1007/978-981-10-6884-3_25
734 H. Cho et al.

characterized by unilateral shrinkage or expansion against axial compressive or


tensile loadings, respectively. Here, based on a comparison between conventional
and auxetic materials, a review of auxetic materials that exhibit various types of
deformation mechanisms and characteristics induced by a negative Poisson’s
ratio is provided. First, the deformation mechanisms in auxetic materials resulting
in lateral expansion under tensile loads are described. They are classified
according to their deformation mechanism or the structural motif enabling
the auxetic behavior including re-entrant structures, rotating unit structures,
chiral structures, fibril/nodule structures, buckling-induced structures, helical
yarn structures, Miura-folded structures, and crumpled structures. Then, the
expected properties of auxetic materials in several aspects are discussed. The
mechanical response of these materials can be drastically changed depending on
the amount of applied loads, and auxetic materials are expected to have unusual,
possibly enhanced geometrical and mechanical characteristics such as synclastic
curvature in bending, deformation-dependent permeability, high shear stiffness,
indentation resistance, and fracture toughness, and improved damping and sound
absorption properties. Finally, some representative potential applications of
auxetic materials are illustrated with a short discussion on the limitations and
outlook of auxetic materials.

Keywords
Auxetic materials · Negative Poisson’s ratio · Structural motif · Deformation
mechanism · Mechanical properties

1 Introduction

When materials are stretched in a certain direction, they usually become thinner
in directions orthogonal to the direction of the applied force (Fig. 1a). The degree
of this transverse deformation with respect to the deformation along the load-
ing direction is characterized by Poisson’s ratio, which is defined as the ratio
of the compressive transverse strain to the tensile axial strain. According to the

Fig. 1 Deformation of a rod under tension when its Poisson’s ratio is (a) positive (as in ordinary
materials) and (b) negative (as in auxetic materials)
23 Mechanics of Auxetic Materials 735

classical theory of elasticity, three-dimensional isotropic materials can have


Poisson’s ratios ranging from 1.0 to 0.5, while they can vary between 1.0 and
1.0 for two-dimensional isotropic materials [1]. Most natural materials have a
positive Poisson’s ratio. For example, metallic materials typically have a Poisson’s
ratio of 0.3, while the Poisson’s ratio of rubbery materials is close to 0.5 due to their
incompressibility.
Materials with a negative Poisson’s ratio are usually referred to as auxetic
materials, derived from the Greek word auxetikos meaning “that which tends to
increase” [2]. These materials, by definition, expand in all directions when a tensile
force is applied in one direction, as shown in Fig. 1b. Because auxetic behaviors
originate from the deformation mechanism of particular geometries and internal
structures in response to uniaxial loadings, a number of natural and synthetic
materials have been found and developed across the length scales. Compared with
conventional materials with a positive Poisson’s ratio, auxetic materials are expected
to have several interesting characteristics in their geometrical and mechanical prop-
erties, including synclastic curvature in bending [3, 4], variable permeability [5],
high shear stiffness [4, 6], enhanced indentation resistance [4, 7, 9], high fracture
toughness [3, 8, 10, 11], and damping and sound absorption [12–15]. These extraor-
dinary properties offer ample opportunities to use auxetic materials for various
applications including, but not limited to, biomedical materials [16], safety equip-
ment, shock absorbing materials [17], energy harvesting devices [18], sports equip-
ment [9], tunable filters [19], soft robotics [20], fashion textiles [5, 21], and
aerospace applications [6].
In nature, only a few auxetic materials have been found. The first experimental
study suggesting the existence of auxetic materials in nature was reported in 1882 for
iron pyrite monocrystals whose Poisson’s ratio was estimated to be 0.14 [22].
Other examples of natural auxetic materials are α-cristobalite, cow teat skin, pyro-
lytic graphite, polymorphic silicones, zeolites, silicates, and crystal cadmium [22].
Because only a limited number of auxetic materials are available in nature, despite
their potential to provide exceptional material properties for a broad range of
applications, a tremendous research effort has been devoted to the development of
artificial materials and structures with auxetic properties. In particular, many
researchers have investigated novel deformation mechanisms leading to auxetic
behavior with high controllability and low manufacturing costs. Representative
mechanisms are re-entrant structures [23–28], rotating rigid units [29–34], chiral
structures [35, 36], fibril/nodule structures [37, 38], Miura-folded structures [39],
buckling-induced structures [40–43], helical auxetic yarn [44], and crumpled struc-
tures [45, 46].

2 Deformation Mechanisms of Auxetic Materials

Since a pioneering study by Lakes in 1987 [3], various man-made auxetic structures
have been studied with structural motif inspiration observed in the microscopic
structures of natural and artificial materials with a negative Poisson’s ratio (NPR)
736 H. Cho et al.

[3, 23, 24, 37] or based purely on geometrical intuition from individual researchers
[36–38, 45, 46]. The degree of auxetic behavior and the related material properties
are highly dependent on the deformation mechanism. Therefore, depending on their
deformation mechanism or structural motif, which induces expansion in all direc-
tions when pulled in a certain direction, these auxetic structures are classified. For
example, some auxetic microstructures have a structural motif possessing initially
re-entrant parts that are gradually straightened up when loaded [3, 23–27]. In this
case, Poisson’s ratio may vary with the applied load as the re-entrance angle changes.
When all the re-entrant parts are fully straightened the auxetic behavior no longer
appears. Another case is the buckling-induced auxetic structure, where buckling
occurs under compressive loads. Unlike the case mentioned above, this structure
exhibits auxetic behavior only when subjected to a compressive force above a certain
critical value, but not when subjected to a tensile force [40–43]. Therefore, under-
standing and investigating the deformation mechanisms or structural motifs, which
differ according to each auxetic structure, can play a central role in the design and
use of auxetic materials for diverse potential applications in engineering and science.

2.1 Re-entrant Mechanism

The most common type of auxetic materials is re-entrant auxetic structures. Simply,
they are formed by truss structures that are composed of thin ribs and linking hinges.
In a 2D re-entrant structure, a unit cell of a polygonal truss, which consists of ribs, is
arranged repeatedly in the structure [25]. The unit cells of these structures are
commonly folded inside the unit cells and the re-entrant sides and vertices form an
important structural motif of the re-entrant structures. In addition, the re-entrant
vertex is joined with one or more neutral ribs that are members of one neighboring
cell, but do not belong to the original unit cell. For example, a re-entrant bowtie
honeycomb structure, which is illustrated in Fig. 2a, is a truss structure in whose unit
cell two facing vertices of a hexagonal structure are re-entrant inside the unit cell.
Re-entrant vertices are linked to the horizontal ribs of neighboring cells to deliver a
load in the lateral direction [23].
There is a close mechanical relationship between structural motifs and the
auxeticity of re-entrant structures. When a tensile load is applied to the re-entrant
structures, the re-entrant sides are unfolded with respect to the re-entrant vertex, and
the vertex moves outward from the unit cell. This lateral deformation of the re-
entrant vertex is transmitted to neighboring cells through neutral ribs connected to
the vertices and pushes nearby cells to induce vertical expansion. Typical examples
of the transformation mechanism in auxetic re-entrant structures are the bowtie
honeycomb structure, triangular re-entrant structure (Fig. 2b), and re-entrant star
structure (Fig. 2c). In the bowtie honeycomb structure, when tensile loads are
applied to horizontal ribs of a unit cell at the top and bottom, the tilted re-entrant
sides are aligned and stretched in the tensile direction. This lengthening behavior
causes the re-entrant vertices to move horizontally and push the neighboring cells
through the horizontal neutral ribs causing them to expand laterally. The triangular
23 Mechanics of Auxetic Materials 737

Fig. 2 Two-dimensional re-entrant auxetic structures: (a) bowtie honeycomb, (b) triangular re-
entrant, and (c) re-entrant star structure. (d) Unit cell of a 3D re-entrant auxetic structure [28]. (e) 3D
re-entrant auxetic structure [28] (Part (d and e) reprinted with permission from Ref. [28] # 2017
Elsevier)

re-entrant structure is a truss structure whose unit cell is an isosceles triangular


structure with a re-entrant base side. When a tensile load is applied to a triangular re-
entrant structure, loads are transferred from the two neutral ribs connected to the re-
entrant vertices so that the re-entrant sides are unfolded. As a result, the triangular
unit cell is widened in the lateral direction. Re-entrant star structures are truss
structures consisting of unit cells of three-, four-, or six-tip star shapes and neutral
ribs connecting their re-entrant vertices. The different star-shaped unit cells have
isotropy in three, four, and six directions, respectively, and when a tensile load is
applied through neutral ribs in a specific direction, not only the vertices jointed to the
loading ribs but also the other re-entrant vertices are unfolded to the same extent [26,
27]. As a result, the unit cell expands in all directions to push neighboring cells
outwards. The relationship between the NPR of the re-entrant auxetics and the
structural motif, or re-entrance, is clear in that the auxeticity is extinguished when
all re-entrant sides are unfolded to their maximum extent.
Three-dimensional re-entrant structures are designed by transforming the struc-
tural motif of a 2D structure into a 3D structure [24]. Further, there are structures in
which ribs and vertices are complexly interconnected, as well as simple applications
738 H. Cho et al.

at the level of simply extruding or three dimensionally stacking sheets of a 2D re-


entrant structure. In this complex example, as the unit cell of the 2D re-entrant
structure is a polygonal truss with re-entrant vertices, the unit cell of the 3D re-
entrant structure (Fig. 2d) is a polyhedral truss structure with vertices which are
dented inwards. Similar to the mechanism shown in the 2D structure, 3D re-entrant
structures also cause lateral deformation due to re-entrant sides unfolding with
respect to the tensile load. A typical 3D re-entrant structure is the 3D bowtie shape
structure, which inherits the geometry and structural motif of the 2D re-entrant
bowtie structure, as shown in Fig. 2e [28]. Another example is the 3D re-entrant
pyramid structure that is a 3D transformation example of the triangular re-entrant
structure. In addition to these structures, there are various other re-entrant structures.
By using a re-entrant polygonal geometry that can be tessellated, microscopic
structures with re-entrant motifs can be easily developed with unlimited design space.
There are advantages and disadvantages of the structure resulting from such
structural motifs. Because these structures have auxeticity not only for the tensile
load but also for the compressive load, these structures have a wide applicability in
comparison with a structure having a NPR only in one direction of the load.
Furthermore, re-entrant structures have high porosity, or low density, which is useful
from the standpoint of lightweight construction. However, re-entrant structures
consist of complicated connections of thin ribs, making them difficult to fabricate
with high accuracy and without defects. In the case of 3D re-entrant structures,
additive manufacturing is essential because of the complicated structure including
the internal cave. In addition, flexure and shear deformation of thin ribs are an
obstacle to intended auxetic behavior. In particular, when a large compressive load is
applied to a re-entrant structure, buckling, which is not considered in the kinematic
model, may occur in the thin ribs. Further, the thin ribs are vulnerable to fatigue
failure and may cause the overall structure to have low durability. Because of these
pros and cons, re-entrant structures are mainly applied to the core structures of
lightweight sandwich panels and analytical models for investigating the microscopic
structure of auxetic foam.

2.2 Rotating Unit Mechanism

Another representative auxetic mechanism is the rotating unit mechanism. Structures


with a rotating mechanism consist of rigid units connected by smooth hinges. Rigid
units are arranged according to a consistent rule, and their initial positions are
slightly tilted in a clockwise or counterclockwise direction, which is opposite to
the tilting direction of the nearby units.
A typical example with such a structural motif is a rotating square unit structure
[29]. Square rigid units are placed repeatedly in longitudinal and lateral directions.
Square units have slightly tilted initial positions and form rhombic voids. The initial
tilt of units induces a rotation of the unit with respect to the tensile load and directly
affects the lateral deformation. Because of the initial position of the units, applied
tensile loads are not collinear between the upper and lower hinges. By applying a
23 Mechanics of Auxetic Materials 739

Fig. 3 Rotating unit auxetic structures: (a) square, (b) rectangular, (c) trans-rectangular, (d) bi-
square, (e) triangular, (f) isosceles triangular, (g) bi-triangular, and (h) hexa-triangular rotating unit
structures

tensile force, a torque is applied to the unit, and units rotate in the clockwise or
counterclockwise direction opposite to their neighboring units. Because units are
rigid and hardly deformed compared to hinges, local rotation of units causes lateral
movement to the hinges on the flank side and lateral expansion. Local rotation
creates the auxetic behavior in which the hinges connected to the left and right
units move outward of the unit and expand in all directions. This rotation occurs until
the tilted square is fully rotated and aligned with the tensile load, and a NPR can no
longer be shown in the fully rotated state.
The rotating square unit structure is the simplest 2D structure with a rotating
motif, which can be used to construct various 2D unit structures. For example,
scaling and tilting these basic square unit structures result in rectangle [31] and
parallelogram unit structures [30], respectively. It is also possible to replace the
rotating unit with two or more different types of unit. Representative examples of
these heterogeneous rotating unit structures are the trans-rectangle (Fig. 3c) and the
bi-square unit structures (Fig. 3d). In addition to quadrangular unit structures with
four hinge points in one unit, 2D rotating structures using a triangular unit with three
hinge points have been developed. Similar to the quadrangular type, the equilateral
triangle unit is the simplest one, and rotating unit structures with arbitrary triangles
or heterogeneous triangles have been developed from the basic geometry (Fig. 3e–h)
[31]. A hexagon can be tessellated in two dimensions, like a quadrangle and triangle,
but it cannot be used to make a rotating unit structure by itself alone. A hexagonal
unit is used with other types of units to create a 2D heterogeneous rotating unit
structure. Typical examples of this type are the hexagon-triangle rotating unit
structure and the hexagon-triangle-rhombi rotating unit structure.
740 H. Cho et al.

Similar to re-entrant structures, 3D rotating unit structures can be designed using


the structural motif of a 2D rotating unit structure. A typical example is a 3D cubic
unit structure [34] that is derived from the geometry of a 2D rotating square unit
structure. Like the 2D rotating square unit structure, tilted cubic units are arranged
regularly and connected by joints. Therefore, when a tensile load is applied to the 3D
rotating cubic unit structure, the tilted unit rotates in a certain direction and lateral
deformation occurs.
In addition to the structures developed so far, a structure having a rotating motif
can be implemented in a wide variety of forms. In particular, heterogeneous rotating
unit structures composed of more than two kinds of units have fundamentally
unlimited design variations, like re-entrant structures. Moreover, there are many
other advantages of rotating unit structures as well as a large design space. Unlike
most other auxetic structures that contain thin ribbons or fibrils, rotating unit
structures do not have slender parts in their geometry; they have very low porosity
compared to structures with other structural motifs. This low porosity enables
subtractive manufacturing and avoids the problems of additive manufacturing
including the limits in selecting mother material, the manufacturing time, and the
accuracy. In particular, they are easily fabricated by patterning slits or ellipses on a
2D sheet, which can give auxeticity to thin-walled structures such as shells [33].
These slit or ellipse patterned structures are used in applications that deal with
compressive loads because they exhibit auxetic behavior stably under compressive
loads. However, because units are rigid, they do not deform significantly, and most
deformation occurs as the hinge bends. This deformation causes stress concentration
in the hinge area and deteriorates the durability of the structure. In addition, because
of the low porosity, there is a limit to how much the effective density can be lowered,
which causes difficulty in engineering lighter structures.

2.3 Chiral Mechanism

A chiral motif with chiral arrangement of unit circles and ribs also causes auxetic
behavior. The unit circles of chiral structures are uniformly arranged in tri-, tetra-, or
hexa-tessellation, and the ribs chirally wrap around the circular units and link one
circle unit to another. When a normal load is applied to an arranged chiral structure,
the load is transferred to the circular units through chiral ribs, and the transferred load
is deflected from the center of circle which generates a rotational torque. The circle
unit then rotates in a certain direction and pulls or pushes the adjacent circular units
through ribs connected in a different direction to the load direction [35].
Two-dimensional chiral structures are composed of circular units and ribs,
and they are limited to five different types [36], unlike the other types of auxetic
structures with unlimited variation. Chiral structures are largely classified as tri-
chirals, tetrachirals, and hexachirals according to the unit circle arrangement
rule, and chiral tessellations and anti-chiral tessellations according to the rotation
direction of the auxetic circle unit. The arrangement rules of unit circles follow basic
2D polygon tessellation. That is, trichirals, tetrachirals, and hexachirals are deter-
mined by three types of tessellation which are triangles, squares, and hexagons,
23 Mechanics of Auxetic Materials 741

Fig. 4 Chiral auxetic structures: (a) tetrachiral structures, (b) anti-tetrachiral structures, (c) trichiral
structures, (d) anti-trichiral structures, and (e) hexachiral structures

respectively. The orientation of ribs surrounding the unit circle depends on the
rotation direction of a unit circle caused by a load. Thus, whole circular units of
the chiral tessellated structures rotate in the same direction, yet the circular units of
the anti-chiral tessellated structures rotate in the opposite direction to their neigh-
boring units. A combination of these categories can create a total of five auxetic
chiral structures: trichiral structures, anti-trichiral structures, tetrachiral structures,
anti-tetrachiral structures, and hexachiral structures, as shown in Fig. 4a–e. Note that
an anti-hexachiral structure does not exist because it cannot be arranged to rotate the
three units of the triangle in different directions.
The chiral mechanism consists of slender structures which have similar structural
advantages and disadvantages of the re-entrant mechanism. The high porosity of
the slender structures allows the structure to be lightweight, but has a detrimental
effect on the durability and stability of the chiral structures. Although the structure
also has NPR behavior under compressive loads, it is vulnerable to local buckling
if compressive loads between circular units are transmitted through thin ribs. Further,
owing to geometrical arrangement, chiral structures have limited structural variation
and a narrow design space, unlike the re-entrant mechanism and the rotating
mechanism.

2.4 Fibril/Nodule Mechanism

Structures with a fibril/nodule mechanism, as the name implies, consist of fibrils


transferring tensile load and rigid nodule units placed between the fibrils. The fibrils
are connected to the nodule units, and if there is no applied force, nodule units are
742 H. Cho et al.

Fig. 5 Typical shape of fibril-nodule structures: (a) single fibril-type structural model for liquid
crystalline polymer (bundle type) [38], (b) Multi-fibril structures with rectangular nodules, and (c)
circular nodules (network type) (Part (a) reprinted with permission from Ref. [38] # 1998 American
Chemical Society)

intertwined with fibrils. However, when a tensile load is applied to a fibril-nodule


structure, a tensile load straightens the fibrils. In this process, the fibrils push the
solid nodule units in the perpendicular direction to the load, and increase the distance
to the adjacent fibrils and nodule units. This tensile force on the fibril leads to
expansion in that direction, resulting in a NPR.
Various fibril/nodule structures are broadly categorized into two types depending
on how the fibrils and nodules are connected: bundle type and network type. The
bundle-type fibril/nodule structures are made up of single fibril/nodule chains in
bundles, as shown in Fig. 5a. Each single chain has an anisotropic nodule attached to
the fibril in a row. When the fibril is stretched, the fibril intertwined with the nodule
unfolds to increase the effective radius of the fibril/nodule chain. This causes single
chains to push each other, and the whole bundle indirectly expands in the perpen-
dicular direction to the applied load. On the other hand, the network-type fibril/
nodule structures, in which one nodule unit is connected to several fibrils, is a nodule
network with linking fibrils rather than a bundle of fibril/nodule chains, as shown in
Fig. 5b, c. When a tensile load is applied to the network fibril/nodule structure,
intertwined fibrils are unfolded between nodules, dropping the overlapping nodules.
Because of the pushing nodules, the tensile load directly causes expansion in all
directions.
The fibril/nodule mechanism is one of the auxetic mechanisms developed in
the early days of auxetic material research, and it is a mechanism suitable for
describing the auxetic polymer materials due to its fiber-like nature. Liquid crystal-
line polymer is an auxetic polymer material whose auxetic behavior can be under-
stood as a fibril/nodule mechanism. Liquid crystalline polymer has the bundle-type
fibril/nodule structure that forms crystalline in stress-free states, but show auxeticity
with decrystallized microscopic structure when a tensile load is applied [38].
Another example is polytetrafluoroethylene foam, which is a kind of auxetic foam
material. The internal structure of polytetrafluoroethylene is composed of a network
of nodules connected by fibrils [37]. Because the fibril/nodule mechanism includes
fibrils that have no shear resistance, fibril/nodule structures cannot bear compression,
under which microscopic fibrils warp. Therefore, the fibril-nodule mechanism only
derives a NPR when stretched, and the auxeticity vanishes under compressive loads.
23 Mechanics of Auxetic Materials 743

2.5 Miura-Folded Mechanism

The Miura-folded mechanism easily assigns auxeticity by simply folding a 2D shell


structure. In the Miura-folded structure, the concavities and convexities appear
repeatedly along a zigzag fold line, as shown in Fig. 6. To create a Miura-folded
structure, a symmetrical parallelogram grid is engraved repeatedly on a 2D shell
structure and then the shell is folded regularly along the grid lines. The interior angle
of the parallelogram grid is a key geometric parameter that determines the shape and
the characteristics of a Miura-folded structure. In the maximally folded state, the
Miura-folded structure has a very small thickness in a certain direction and facets
overlap each other in the other direction. When a tensile load is applied in this state,
the fold lines of the Miura-folded structure unfold and expand the overlapping facets
along the loading direction [39].
By adjusting the angle of the parallelogram grid, which is a key parameter of this
structure, geometry including the bottom shape and the height of the structures, and
mechanical properties including NPR can be controlled. This can be used to create
two different Miura-folded structures with the same bottom shape and top fold lines
but different heights. These different structures can be alternately stacked to form a
3D cellular Miura-folded structure.
The Miura-folded structure has the advantage of achieving a very wide range
of NPRs. In addition, unlike other auxetic mechanisms that require additive or

Fig. 6 A unit cell of a Miura-folded structure. a, b, γ and dihedral fold angle θ are the geometric
parameters that determine a parallelogram facet. The other parameters of the outer dimension H, S,
V, L and angles φ, ξ, ψ are defined to describe the folded state usefully [39] (Reprinted with
permission from Ref. [39] # 2013 PNAS)
744 H. Cho et al.

subtractive manufacturing processes, the Miura-folded mechanism simply has


auxeticity by folding a 2D shell. Therefore, it is useful for applications, such as
fluid transport, because the entire structure is connected continuously and the
material cannot move through the structure. However, the Miura-folded structure
is a thin shell with a fold line, so it is less rigid than other auxetic structures. In
addition, its durability is low because deformation stress is concentrated on thin fold
lines that behave like hinges [39].

2.6 Buckling-Induced Mechanism

The buckling-induced mechanism is a unique mechanism, quite unlike the other


auxetic mechanisms described above. Because auxeticity is created by buckling,
buckling-induced auxetic materials have a NPR only when the applied compressive
force is greater than a critical value. In addition, unlike other auxetic mechanisms
exhibiting initially auxetic behavior, auxetic material with the buckling-induced
mechanism does not exhibit auxeticity in its initial shape or at small strain ranges.
The buckling-induced mechanism is represented by simple structures in which
circular patterns are arranged at regular intervals in vertical and horizontal directions
on a 2D sheet (Fig. 7a) [40]. Despite their simple geometry, these structures exhibit
complicated behavior with three phases depending on load magnitude.
A buckling-induced auxetic structure behaves differently depending on the level
of strain due to compressive load. At small strain ranges, owing to the symmetry of
the placed circles, a buckling-induced auxetic structure undergoes linear elastic
deformation, which is similar to that of conventional structures. In this linear
deformation phase, or pre-buckling phase, no special lateral deformation appears.
However, as the load increases and beyond the critical strain point, it shifts to the
buckling phase where nonlinearity appears. The thin part between the circle patterns
is buckled and bent symmetrically by the applied compressive load. As a result, the
circle patterns become ellipse or dumbbell shape and alternate in longitudinal and
lateral directions. In this process, the entire structure shrinks not only in the com-
pressive load direction but also in the lateral direction, resulting in a NPR. In the
post-buckling phase, the auxeticity remains until the load is further increased and the
compression is progressed. In this inner contact phase, the structure no longer has
auxetic behavior, like the linear deformation phase.
The simple structural motif of the buckling-induced mechanism can be easily
incorporated into 3D structures from 2D circle-patterned sheets. As a typical exam-
ple, a circle-patterned cylinder is a structure in which circular patterns are arranged in
the direction of the axis and tangent to the surface of a thick cylinder, as shown in
Fig. 7c [41]. Similar to the 2D circle-patterned structures, the nonlinear buckling
behavior is only shown with respect to a compressive load in the axial direction of
the cylinder. An example of applying a 2D circle-patterned structure to a 3D
structure in a more sophisticated manner is a 3D buckyball structure, as shown in
Fig. 7b [42]. Like the 2D structure, it has a simple uniformly patterned structure and
the nonlinear buckling behavior is shown above the critical point, after which the
23 Mechanics of Auxetic Materials 745

Fig. 7 Buckling-induced auxetic structures: (a) 2D circle-patterned sheet and its deformation [40],
(b) 3D buckyball structure and its deformation [42], (c) circle-patterned cylinder [41] (Part (a)
reprinted with permission from Ref. [40] # 2008 Elsevier. Part (b) reprinted with permission from
Ref. [42] # 2012 PNAS. Part (c) reprinted with permission from Ref. [41] # 2016 Elsevier)

surface of the sphere is locally rotated. In this process, the structures show NPR
behavior. In addition, a number of buckyball structures may be regularly stacked to
form a 3D bulk structure [43]. When a compressive load is applied, the buckyball
structures inside the 3D bulk structure shrink under load, resulting in a NPR bulk
structure.
A buckling-induced structure has the advantages of controlling the physical
properties and performing motion actuating. However, as buckling is induced by
compression, auxeticity does not appear with tensile loads. Furthermore, because it
is very sensitive to loading and boundary conditions, when applying the buckling-
induced mechanism, devices that can set and control the load and boundary condi-
tions accurately should be used.

2.7 Helical Auxetic Yarn

Helical auxetic yarn is a unique auxetic material that consists of two types of threads
[44]. The core thread, one of the constituents, is a thick but soft thread that has
straight shape in stress-free states. On the other hand, the wrap thread, the other
constituent, is thin but has a high stiffness. This wrap thread helically coils around
and attaches to the core thread in the initial state, and the two threads do not move
relative to each other. In the initial state, the effective diameter of the entire helical
yarn is defined as the diameter of the core thread plus twice the diameter of the wrap
thread (Fig. 8a). When a tensile load is applied to the helical yarn, a dramatic shape
746 H. Cho et al.

a
Stiffer Wrap
c

Elastomeric Core

b
Stiffer Wrap

Elastomeric Core

Fig. 8 Helical auxetic yarn structure: (a) Undeformed [44] and (b) deformed configurations of
helical auxetic yarn [44]. (c) Auxetic fabric with helical yarn [44] (Part (a–c) reprinted with
permission from Ref. [44] # 2009 Elsevier)

change occurs due to the difference in stiffness values between the two threads.
Because the wrap thread is stiffer than the core thread, the helically coiled wrap
thread straightens in the tensile load direction. In this process, the soft core thread is
pushed by the wrap thread and twisted along the attached side of the wrap thread, as
shown in Fig. 8b. As a result, the core thread is now helically coiled around the wrap
thread as opposed to the condition in the stress-free state. In this state, the effective
diameter of the deformed shape is defined as the diameter of the wrap thread plus
twice the diameter of the core thread. Because the core thread is considerably thicker
than the wrap thread, the effective diameter increases due to the tension load. As
a result, an entire fabric composed of helical auxetic yarns has a NPR, and extends in
a perpendicular direction to the applied load.
Unlike other auxetic structures, helical auxetic yarns have the advantage that they
can be easily fabricated to surround a free-form surface, such as a body. However,
the structures are composed of core and wrap threads that have no resistance to shear
forces, so they do not exhibit effective auxetic behavior when subjected to compres-
sive loads. In addition, as the helical yarns have indirect auxetic behavior, the lateral
expansion due to a tensile load is not large and does not produce a high NPR value
compared to other mechanisms.

2.8 Crumpled Mechanism

A crumpled mechanism uses random and isotropic crumples as a structural motif.


Crumpled structures have microscopic crumples due to omnidirectional compres-
sions or chemical imperfections inside the structure. These crumples are randomly
oriented, as shown in Fig. 9, so that when a tensile load is applied to the crumpled
structure the crumples unfold and expand in all directions. In addition, when thin foil
structures are compressed in all directions, natural random crumples are generated
23 Mechanics of Auxetic Materials 747

Fig. 9 A typical example of crumpled structures: (a) 3D [46] and (b) 2D crumpled aluminum foil
[46] (Part (a) and (b) reprinted with permission from Ref. [46] # 2013 Elsevier)

by membrane buckling. Even after the compression load is removed, the crumples
are retained by plastic deformation and the desired crumpled mechanism can be
represented. In addition to the method of compressing the mother material, an
imperfection in the mother material can be used to harness crumples. For example,
a carbon imperfection in hexagonal tessellations of graphene is induced to create a
2D crumpled graphene. A carbon imperfection results in bending to minimize the
internal energy, and harnessing crumples in load-free conditions [45].
Crumpled auxetics, like the Miura-folded structure, do not pattern on a 2D sheet,
so fluid does not flow through the sheet. That is, unlike other auxetic mechanisms,
crumpled structures can be easily applied to engineering problems dealing with
fluids, such as fluid transportation. In addition, because of the use of random
crumples without structural symmetry, unlike other auxetic materials that undergo
cutting or additive manufacturing processes, they only undergo compression and
imperfection creation. This makes it easy to obtain large quantities of high-quality
auxetic materials compared to other structures that require high precision in the
manufacturing process. Moreover, the structures can control the mechanical prop-
erties of the entire material through compression and imperfection control in the
process without difficulty [46].

3 Relationship Between the Deformation Mechanism and


the Material Properties

As described above, much research on the structural behavior of auxetic material and
materials that are likely to become auxetic has been conducted. However, there was
not as much information on the relationship between the mechanical properties of
auxetic materials and their deformation mechanism. In 2012, Elipe et al. [47]
performed a comparative study of two- and three-dimensional auxetic geometries
748 H. Cho et al.

using simulation tools. They found some relevant design properties of different
auxetic geometries such as the Poisson’s ratio, the maximum volume, the area
reduction, the equivalent Young’s modulus, and the density. In this study, they
used computer-aided design tools to first design several auxetic structures with
different geometries by replicating unit cells. These structures were then analyzed
by using the finite element method (FEM) to derive the relevant properties. From
these results, the relationship between some geometric parameters and material
properties could be found including the Poisson’s ratio versus the maximum area
(for 2D structures) or volume (3D structures) reduction, the equivalent Young’s
modulus for different 2D/3D auxetic structures, and the Young’s modulus versus the
density with respect to bulk properties. The characteristics of auxetic materials vary
significantly depending on the geometry and shape of the structures. Although only
few properties were considered, the results clearly demonstrate that auxetic materials
offer a versatile way of achieving the specific target properties by employing an
appropriate auxetic mechanism [47].

4 Expected Properties of Auxetic Materials

Because of the NPR effect of auxetic materials, there are various expected geomet-
rical and mechanical properties hardly observed in natural, nonauxetic materials.
The properties include, but are not limited to, synclastic curvature in bending in
contrast to anticlastic curvature appeared in conventional materials, variable perme-
ability, high shear stiffness, enhanced indentation resistance, high fracture tough-
ness, and improved damping and sound absorption properties. Proper understanding
of geometric, static, and dynamic properties of auxetic materials is therefore essen-
tial to fully utilize these extraordinary properties for specific target applications.
In this section, we describe the fundamental reasons why auxetic materials are
expected to have such properties by comparing them with the properties of conven-
tional, nonauxetic materials.

4.1 Synclastic Curvature in Bending

When out-of-plane bending is applied to nonauxetic materials, the surface displays


anticlastic curvature, whereby the material is curved from a vertical position to the
opposite direction [3], as shown in Fig. 10a. This behavior leads to a requirement
of additional forces for conventional materials when constructing convex shapes.
However, the surface of auxetic materials has synclastic curvature if deformations
occur in the same direction [4], as shown in Fig. 10b. Compared to the saddle shape
with anticlastic curvature, the dome-like shape with synclastic curvature is
expected to be useful because it can be used to construct complex structures
through a relatively small number of machining steps and low manufacturing
cost [4].
23 Mechanics of Auxetic Materials 749

Fig. 10 Curvature in bending: (a) anticlastic curvature of conventional, nonauxetic materials and
(b) synclastic curvature of auxetic materials

Fig. 11 An illustration of
variable permeability in a
triangular re-entrant structure

4.2 Variable Permeability

In a porous material used for filtration, such as filters, controlling permeability is one
of the key factors. Porous auxetic materials are excellent for varying permeability
thanks to the NPR effect. These materials increase the size of their pores when a
tensile load is applied in a certain direction. For example, a triangular re-entrant
structure improves permeability by adjusting the size of its pores as each unit cell
unfolds in all directions, as shown in Fig. 11. Because of auxetic behaviors, this
variable permeability can be utilized from macroscale to nanoscale materials [5].

4.3 High Shear Stiffness

Shear modulus (G) is a property measured to determine the deformation that occurs
when a force is applied parallel to one side of an object and the other side is fixed. If
the value of shear modulus is large, the material is rigid and has high shear
resistance. Thus, shear modulus is an important factor to select when designing
and manufacturing any structure. In isotropic auxetic materials, shear resistance is
expected to be more beneficial than in nonauxetic materials. This feature can be
750 H. Cho et al.

explained by the following equations consisting of the Young’s modulus (E), the
shear modulus (G), the bulk modulus (K ), and the Poisson’s ratio (v) [4].

E
G¼ (1)
2ð 1 þ v Þ
E
K¼ (2)
3ð1  2vÞ

As can be seen in Eqs. (1) and (2), if the Poisson’s ratio (v) approaches 0.5, the
value of the shear modulus is reduced but the bulk modulus is greatly increased. This
means that the material easily undergoes shear deformation, but the shape of material
does not change much; the material is incompressible like rubber. However, when
the Poisson’s ratio is close to 0.5, the shear modulus (G) and the Young’s modulus
(E) have almost the same value and a higher value than the bulk modulus (K ). Thus,
in this case, it is easy to compress the material but hard to shear it. Further, as the
Poisson’s ratio (v) approaches 1, the shear modulus (G) becomes infinite which is
similar to the hardness becoming infinity [6] and the shear resistance gets signifi-
cantly larger [48]. In other words, it is hard to shear but easy to undergo volumetric
deformation. Therefore, by adjusting the Poisson’s ratio (v) it is possible to design a
structure efficiently according to its purpose.

4.4 Enhanced Indentation Resistance

When an impact is applied to nonauxetic materials in a certain region, these materials


tend to spread from the region to relieve the pressure, as shown in Fig. 12a. Thus,
this behavior causes a decrease in density at the impact region. On the contrary, in the
same case, isotropic auxetic materials move to regions where the impact is applied.
This leads to an increase in indentation resistance by having regions of higher
density [4], as shown in Fig. 12b. This phenomenon can also be confirmed by the
relationship between the Poisson’s ratio (v), the Young’s modulus (E), and the

Fig. 12 An illustration of the


resistance to indentation for
(a) nonauxetic and (b) auxetic
materials
23 Mechanics of Auxetic Materials 751

material hardness (H ) associated with the indentation resistance. The relationship


between material hardness, Young’s modulus, and Poisson’s ratio is given by
 
E γ
H/ , (3)
1  v2

where γ is 1 or 2/3 for uniform pressure distribution or Hertzian indentation,


respectively.
As mentioned earlier, Poisson’s ratios for 3D isotropic materials range from 1.0
to 0.5. When the Poisson’s ratio approaches 1.0, the hardness becomes infinite and
this leads to an increase in indentation resistance [4]. However, Poisson’s ratios for
the 2D isotropic materials can change between 1.0 and 1.0, which means that the
upper limit of the ratios is different from the 3D materials. Therefore, in 2D materials
the hardness can be infinite even for positive Poisson’s ratio. This behavior is not
presented in anisotropic materials because the above equation is only applicable for
isotropic materials.
Many studies have been carried out on various kinds of synthetic auxetic mate-
rials to investigate hardness properties. For instance, Alderson et al. found experi-
mentally that auxetic ultra-high molecular weight polyethylene (UHMWPE) is twice
as hard as conventional UHMWPE. This behavior specifically occurs at low loads
(10–100 N) and in the radial direction. At low loads, because UHMWPE undergoes
elastic deformation, the NPR effect is significant. Further, although indentation
resistance increases only in the radial direction, there is no reduction to the hardness
in the axial direction. Therefore, anisotropic as well as isotropic auxetic UHMWPE
can be beneficial to use when considering hardness [7].

4.5 High Fracture Toughness

Fracture resistance is better for auxetic materials than for conventional materials [8].
Auxetic materials lead to low crack propagation and require more energy to expand
materials. Thus, many studies have been conducted to find the difference in fracture
toughness between auxetic and nonauxetic materials.
Donoghue et al. demonstrated experimentally that auxetic materials need less
energy for crack propagation than conventional materials [10]. Further, Maiti
showed that the fracture toughness of a conventional material is determined by the
ratio of its normalized density [11]. From the critical tensile stress equation, Lakes
indicated that toughness can be changed as the Poisson’s ratio varies, and also if the
Poisson’s ratio approaches 1 the material becomes very tough [3]. The critical
tensile stress equation is as follows:
 1=2
πET
σ¼ , (4)
2r ð1  v2 Þ

where E is the Young’s modulus, T is surface tension, and r is the plane circular crack
radius.
752 H. Cho et al.

Later, the fracture toughness of re-entrant foam and conventional foam was
analytically derived and a comparison between the fracture toughness of the two
foams was conducted by Choi and Lakes [8]. The fracture toughness ( K IC ) of
conventional foam can be expressed by
 
K IC ρ
pffiffiffiffi ¼ 0:19 , (5)
σ f πl ρs

where ρs is the density of the solid made from the foam, ρ is the density of the foam,
σ f is the fracture strength of the cell rib, and l is the cell rib length.
The fracture toughness (K TIC ) of the re-entrant foam is given by
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  
K TIC 1 þ sin ð0:5π  φÞ ρ
pffiffiffiffi ¼ 0:10 , (6)
σ f πl 1 þ cos 2φ ρs

where φ is the rib angle which determines cell shapes.


By combining the above two equations, Eqs. (5) and (6), the proportion of the
fracture toughness between the re-entrant foam and the conventional foam can be
derived.
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
K TIC 1 þ sin ð0:5π  φÞ
¼ 0:53 (7)
K IC 1 þ cos 2φ

From Eq. (7), it can be seen that density is not needed when comparing re-entrant
foam with conventional foam. Although these equations provide the basic insight
into foam-type structures, the expressions are based on ideal models rather than
actual ones, and so it is difficult to predict the behavior of real cell shapes from these
analytical expressions.

4.6 Increased Energy Absorption

Compared to conventional materials, auxetic materials exhibit superior features in


terms of damping and sound absorption. NPR foams are better at absorbing sound
than conventional materials [12]. The ability to absorb sound depends on the size
of the pore of an auxetic foam, which is excellent for NPR foams with smaller pores
at frequencies above 630 Hz. In addition, NPR covered foams have better acoustic
absorption capacity than NPR uncovered foams at the frequencies between 250
and 1000 Hz [13]. Alderson et al. demonstrated experimentally that the highest
ultrasonic attenuation coefficient of auxetic UHMWPE was 1.5 times higher than the
coefficient of the nonauxetic microporous UHMWPE and 3 times higher than foams
that were processed conventionally [14]. Furthermore, auxetic foams have excellent
resilience under dynamic impact loading, which is not noticeable in conventional
foams [15].
23 Mechanics of Auxetic Materials 753

5 Potential Applications of Auxetic Materials

Auxetic materials, whose geometrical and mechanical properties have many novel
characteristics due to the NPR effect, can be applied in various engineering and
scientific fields, and many potential applications have been suggested. In this
section, several applications with their basic mechanisms are described.
In biomedical engineering, there are many possible applications for using auxetic
materials, such as arterial prostheses, implants, stents, scaffolds, dilators, sutures,
ligament/muscle anchors, bandages, and prosthetic linings [16]. One of the well-
known smart biomedical devices is the auxetic bandage. When the auxetic bandage
is attached to an infected wound, the bandage expands and the pores extend. This
mechanism, derived from shape memory ability and the NPR effect, increases
porosity and breathability and allows the wound to heal faster.
A filter using the variable permeability property is another example of an appli-
cation. Because particulates trapped in the pores reduce the efficiency of filter
systems, pore opening is important for effective filtration. If the filter is fabricated
with the characteristics of an auxetic material, the size of its pores is extended in all
directions as a tensile load is applied in one direction. Therefore, by controlling pore
size, when compared to a conventional filter, the filtration efficiency and functional
performance of the auxetic filter structure can be improved [19].
Another example is a press-fit fastener, which uses the auxetic material property
to expand in the transverse direction when a tensile load is applied longitudinally.
This characteristic makes it more difficult to remove the fastener [49, 50], and so the
holding capacity of this fastener is increased with tensile load. Similar effects occur
in fiber-reinforced composites, which enhance the pull-out resistance as the fiber
expands [50]. When a fiber with a NPR is pulled along one axis, it will become wider
perpendicular to the applied force, which can prevent loosening between the matrix
and the fiber. The behavior of this reinforcing fiber is described in Fig. 13.
In addition, by using auxetic materials, a soft metamaterial robot with only one
actuator can be designed to move through a travel channel. This robot is made with

Fig. 13 Pull-out resistance of


a reinforcing fiber when a
tensile load is applied
754 H. Cho et al.

two passive clutch components and a linear actuator. The two passive clutch
components, connected to a linear actuator, possess opposite signs of Poisson’s
ratio. Thus, the two components are composed of auxetic and normal materials,
and they have different mechanical behaviors. When the actuator is expanded in an
axial direction, the auxetic material is compressed but the normal material is
expanded transversely. In this case, the limiting friction of the auxetic material is
relatively reduced when compared to that of the normal material. Therefore, the
normal material does not move but the auxetic material slides along the channel. On
the contrary, if the actuator is compressed, the limiting friction is decreased in the
normal material. This case has the opposite effect and so the normal material slips.
With this simple mechanism controlling limiting friction, linear unidirectional
motion is realized with only one actuator, and the difficulty of synchronization can
also be solved [20].
Manufacturing clothes using auxetic fabrics could be a potential candidate for a
solution to solve the problem of children growing quickly. Parents usually buy
clothes that are much larger than their child’s size, taking into account the fact that
they will grow. As a result, children wear loose clothes, which can cause serious
injuries when playing. However, thanks to the behavior of auxetic fabrics, which
expand in all directions when a tensile force is applied in one direction, auxetic
fabric with a foldable structure could provide long-lasting clothing for children.
This could relieve the discomfort and burden for both children and their parents
[5].
Konakovic et al. introduced a computational method to approximate a non-
developable surface with a double-curvature from flat rigid materials [21]. Although
thin flat materials, which are inextensible but bend freely, are able to approximate
developable surfaces, which are single curved surfaces, they cannot approximate
nondevelopable surfaces. If rotating rigid units with triangular patterns are inserted
into these flat materials, the material gets much more freedom and can be stretched.
However, it is hard to intuitively predict flat domains that can construct target
surfaces. Thus, by using a conformal map and numerical optimization, the flat
domain that is the easiest to approximate a target surface is obtained. After deciding
the cuts and orientations of a specific cutting pattern, a 2D region can be lifted onto
target surfaces. These approximations provide new possibilities in material design
and have the potential to be applied in many areas.
Furthermore, auxetic material can be used for composite sandwich panels that are
designed to resist high strain rate loads. A panel with multiple layers of the same
auxetic unit cells and metal facets enhances its resistance to blast damage while
remaining lightweight. By using this benefit, the panel can be applied to armored
vehicles to increase their ballistic and impact resistance. Further, composite struc-
tures are used to prevent underwater explosions in ship hulls and improve crash
absorption in automobile parts [18]. In addition, to protect soldiers on the battlefield
thick protective clothing is required, which is normally stiff, heavy, and rigid.
However, if protective clothing with auxetic tiles is produced, it can provide a
similar level of protection but is lighter and thinner. In addition, body armor and
combat jackets with auxetic material improves energy absorption and reduces wear
23 Mechanics of Auxetic Materials 755

resistance. These textiles also lead to changes in high volume and drapability due to
the properties of auxetic materials [9].
Auxetic materials with synclastic curvature have been applied to the field of
aerospace, especially in aircraft nose cones and wing panels. Helmets and kneepads
are also manufactured with auxetic material, because it can better cover the body [6].
In addition to these applications, there are other potential applicable fields, including
efficient piezoelectric sensors, sports equipment, and sound-absorbing materials
[17].

6 Concluding Remarks

Because of the NPR effect, there are many unique, enhanced properties expected for
auxetic materials. But, at the same time, certain physical properties can be deterio-
rated when incorporating the deformation mechanism of auxetic materials. For
example, most auxetic structures have hinges and pores that often make these
structures weaker and more flexible than conventional nonauxetic solids due to the
nature of their geometry and deformation mechanism. Hence, the use of these
materials for large load-bearing structures can be limited. Also, a significant amount
of stresses can be concentrated near the hinge regions, rendering the auxetic structure
vulnerable to cyclic fatigue loadings. Therefore, when designing and using auxetic
materials, extreme care must be taken to consider the weakened properties due to the
effect of negative Poisson’s ratio as well.
Nevertheless, auxetic materials have been successfully applied to various fields of
engineering applications and offer a great possibility to develop materials and
structures with extraordinary mechanical properties. In order to further broaden the
application of auxetic materials, more in-depth understanding and thorough inves-
tigation on the deformation mechanism and its link to the mechanical properties are
essential. In particular, combining auxetic materials with conventional, nonauxetic
materials would enlarge the design space of mechanical properties that have not been
reached by using solely conventional materials or auxetic materials.

References
1. Wan H, Ohtaki H, Kotosaka S, Hu G. A study of negative Poisson’s ratios in auxetic
honeycombs based on a large deflection model. Eur J Mech – A/Solids. 2004;23:95–106.
2. Prawoto Y. Seeing auxetic materials from the mechanics point of view: a structural review on
the negative Poisson’s ratio. Comput Mater Sci. 2012;58:140–53.
3. Lakes R. Foam structures with a negative Poisson's ratio. Science. 1987;235:1038–41.
4. Evans KE, Alderson A. Auxetic materials: functional materials and structures from lateral
thinking! Adv Mater. 2000;12:617–28.
5. Wang Z, Hu H. Auxetic materials and their potential applications in textiles. Text Res J.
2014;84:1600–11.
6. Underhill RS. Defense applications of Auxetic materials. Adv Mater. 2014;1:7–13.
7. Alderson KL, Pickles AP, Neale PJ, Evans KE. Auxetic polyethylene: the effect of a negative
Poisson’s ratio on hardness. Acta Metall Mater. 1994;42:2261–6.
756 H. Cho et al.

8. Choi JB, Lakes RS. Fracture toughness of re-entrant foam materials with a negative Poisson’s
ratio: experiment and analysis. Int J Fract. 1996;80:73–83.
9. Liu Q. Literature review: materials with negative Poisson’s ratios and potential applications to
aerospace and defence. Defence Science and Technology Organisation Victoria (Australia) Air
Vehicles Div; 2006.
10. Donoghue JP, Alderson KL, Evans KE. The fracture toughness of composite laminates with
a negative Poisson’s ratio. Phys Status Solidi B. 2009;246:2011–7.
11. Maiti SK, Ashby MF, Gibson LJ. Fracture toughness of brittle cellular solids. Scr Metall.
1984;18:213–7.
12. Chekkal I, Bianchi M, Remillat C, Becot FX, Jaouen L, Scarpa F. Vibro-acoustic properties of
auxetic open cell foam: model and experimental results. Acta Acustica United Acustica.
2010;96:266–74.
13. Howell B, Prendergast P, Hansen L. Examination of acoustic behavior of negative poisson’s
ratio materials. Appl Acoust. 1994;43:141–8.
14. Alderson KL, Webber RS, Mohammed UF, Murphy E, Evans KE. An experimental study of
ultrasonic attenuation in microporous polyethylene. Appl Acoust. 1997;50:23–33.
15. Scarpa F, Ciffo LG, Yates JR. Dynamic properties of high structural integrity auxetic open cell
foam. Smart Mater Struct. 2003;13:49.
16. Bhullar SK, Lala NL, Ramkrishna S. Smart biomaterials - a review. Review on advanced. Mater
Sci. 2015;40:303–14.
17. Novak N, Vesenjak M, Ren Z. Auxetic cellular materials - a review. Strojniški Vestnik – J Mech
Eng. 2016;62:485–93.
18. Imbalzano G, Tran P, Ngo TD, Lee PV. Three-dimensional modelling of auxetic sandwich
panels for localised impact resistance. J Sandw Struct Mater. 2017;19:291–316.
19. Alderson A, Rasburn J, Ameer-Beg S, Mullarkey PG, Perrie W, Evans KE. An auxetic filter:
a tuneable filter displaying enhanced size selectivity or defouling properties. Ind Eng Chem Res.
2000;39:654–65.
20. Mark AG, Palagi S, Qiu T, Fischer P. Auxetic metamaterial simplifies soft robot design. In
Robotics and Automation (ICRA), International Conference on Robotics and Automation.
IEEE. 2016.
21. Konaković M, Crane K, Deng B, Bouaziz S, Piker D, Pauly M. Beyond developable: compu-
tational design and fabrication with auxetic materials. ACM Trans Graph (TOG).
2016;35:79–89.
22. Carneiro VH, Meireles J, Puga H. Auxetic materials - a review. Mater Sci – Poland.
2013;31:561–71.
23. Masters IG, Evans KE. Models for the elastic deformation of honeycombs. Compos Struct.
1996;35:403–22.
24. Yang L, Harrysson O, West H, Cormier D. Mechanical properties of 3D re-entrant honeycomb
auxetic structures realized via additive manufacturing. Int J Solids Struct. 2015;69:475–90.
25. Larsen UD, Signund O, Bouwsta S. Design and fabrication of compliant micromechanisms and
structures with negative Poisson's ratio. J Microelectromech Syst. 1997;6:99–106.
26. Theocaris PS, Stavroulakis GE, Panagiotopoulos PD. Negative Poisson's ratios in composites
with star-shaped inclusions: a numerical homogenization approach. Arch Appl Mech.
1997;67:274–86.
27. Wang ZP, Poh LH, Dirrenberger J, Zhu Y, Forest S. Isogeometric shape optimization of
smoothed petal auxetic structures via computational periodic homogenization. Comput
Methods Appl Mech Eng. 2017;323:250–71.
28. Wang XT, Wang B, Li XW, Ma L. Mechanical properties of 3D re-entrant auxetic cellular
structures. Int J Mech Sci. 2017;131:396–407.
29. Grima JN, Evans KE. Auxetic behavior from rotating squares. J Mater Sci Lett.
2000;19:1563–5.
30. Grima JN, Farrugia PS, Gatt R, Attard D. On the auxetic properties of rotating rhombi and
parallelograms: a preliminary investigation. Phys Status Solidi B. 2008;245:521–9.
23 Mechanics of Auxetic Materials 757

31. Grima JN, Gatt R, Alderson A, Evans KE. On the auxetic properties of rotating rectangles’ with
different connectivity. J Phys Soc Jpn. 2005;74:2866–7.
32. Grima JN, Chetcuti E, Manicaro E, Attard D, Camilleri M, Gatt R, Evans KE. On the auxetic
properties of generic rotating rigid triangles. In Proceedings of the Royal Society of London A.
2011.
33. Shan S, Kang SH, Zhao Z, Fang L, Bertoldi K. Design of planar isotropic negative Poisson’s
ratio structures. Extreme Mech Lett. 2015;4:96–102.
34. Kim J, Shin D, Yoo DS, Ki K. Regularly configured structures with polygonal prisms for three-
dimensional auxetic behaviour. In Proceedings of the Royal Society of London A. 2017.
35. Prall D, Lakes RS. Properties of a chiral honeycomb with a Poisson’s ratio of 1. Int J Mech
Sci. 1997;39:305–14.
36. Grima JN, Gatt R, Farrugia PS. On the properties of auxetic meta-tetrachiral structures. Phys
Status Solidi B. 2008;245:511–20.
37. Evans KE, Caddock BD. Microporous materials with negative Poisson’s ratios II. Mechanisms
and interpretation. J Phys D Appl Phys. 1989;22:1877–83.
38. He C, Liu P, Griffin AC. Toward negative Poisson ratio polymers through molecular design.
Macromolecules. 1998;31:3145–7.
39. Schenk M, Guest SD. Geometry of Miura-folded metamaterials. Proc Natl Acad Sci.
2013;110:3276–81.
40. Bertoldi K, Boyce MC, Deschanel S, Prange SM, Mullin T. Mechanics of deformation-triggered
pattern transformations and superelastic behavior in periodic elastomeric structures. J Mech
Phys Solids. 2008;56:2642–68.
41. Javid F, Liu J, Shim J, Weaver JC, Shanian A, Bertoldi K. Mechanics of instability-induced
pattern transformations in elastomeric porous cylinders. J Mech Phys Solids. 2016;96:1–17.
42. Shim J, Perdigou C, Chen ER, Bertoldi K, Reis PM. Buckling-induced encapsulation of
structured elastic shells under pressure. Proc Natl Acad Sci. 2012;109:5978–83.
43. Babaee S, Shim J, Weaver JC, Chen ER, Patel N, Bertoldi K. 3D soft metamaterials with
negative Poisson's ratio. Adv Mater. 2013;25:5044–9.
44. Miller W, Hook PB, Smith CW, Wang X, Evans KE. The manufacture and characterisation
of a novel, low modulus, negative Poisson’s ratio composite. Compos Sci Technol.
2009;69:651–5.
45. Grima JN, Winczewski S, Mizzi L, Grech MC, Cauchi R, Gatt R, Rybicki J. Tailoring graphene
to achieve negative Poisson's ratio properties. Adv Mater. 2015;27:1455–9.
46. Bouaziz O, Masse JP, Allain S, Orgéas L, Latil P. Compression of crumpled aluminum thin foils
and comparison with other cellular materials. Mater Sci Eng A. 2013;570:1–7.
47. JCA E, Lantada AD. Comparative study of auxetic geometries by means of computer-aided
design and engineering. Smart Mater Struct. 2012;21:104993–5004.
48. Saxena KK, Das R, Calius EP. Three decades of auxetics research - materials with negative
Poisson’s ratio: a review. Adv Eng Mater. 2016;18:1847–70.
49. Choi JB, Lakes RS. Non-linear properties of polymer cellular materials with a negative
Poisson's ratio. J Mater Sci. 1992;27:4678–84.
50. Evans KE. Auxetic polymers: a new range of materials. Endeavour. 1991;15:170–4.

You might also like