You are on page 1of 9

THE AERONAUTICAL JOURNAL MARCH 2003 175

Mechanical and electromagnetic behaviour of


auxetic honeycomb structures

F. Scarpa, F. C. Smith and B. Chambers


University of Sheffield, UK

G. Burriesci
University of Palermo, Italy

ABSTRACT 1.0 INTRODUCTION


In this work a combined analysis of the out-of-plane mechanical and Materials and microstructures showing a negative Poisson’s ratio
the electromagnetic properties of auxetic re-entrant honeycombs is behaviour are called auxetic, from the Greek auxetoe (‘which can
performed. Experimental and numerical simulations are carried out expand’). Figure 1 shows the performance of a cellular solid having
to evaluate the correlation between the anisotropicity of the trans- a negative Poisson’s ratio coefficient. Opposite to a ‘classical’ mate-
verse mechanical properties (shear and compressive modulus) and rial, when pulled in one direction it expands in the transverse one.
the permittivity tensor of auxetic honeycombs. The results are evalu- According to the theory of classical elasticity(1), this kind of behav-
ated to assess the feasibility of this kind of cellular solid for electro- iour is permissible for an homogeneous, isotropic and thermodynam-
magnetic absorption and window applications with high structural ically correct 3D solid, with a possible Poisson’s ratio values range
integrity performance. between –1 and 0⋅5. The upper limit corresponds to a material virtu-
ally incompressible, while the lower bound represents a solid with a
shear modulus tending to infinity(2). Materials with anisotropy have
Poisson’s ratio values outperforming the previous figures(2,3). In
NOMENCLATURE every case, a negative Poisson’s ratio value leads to a volume
t unit cell wall thickness increase of the solid subjected to a stress or strain field, without
b transversal gauge thickness of the unit cell involving any fundamental difference of the fundamental mecha-
Ec core material Young’s modulus nisms of deformation(2).
E3 honeycomb Young’s modulus in the three direction A material with auxetic features would offer some appealing char-
G13, G23 transverse shear moduli in the 13 and 23 planes acteristics to the structural designer. It can be demonstrated theoreti-
h,l unit cell walls lengths cally that indentation resistance and flexural bending strength of
plates and solids could be greatly enhanced compared to analogous
α cell wall aspect ratio (h/l) structures made of a more ‘classical’ material(2,4,5). In 1987 Roderick
β cell wall relative density (t/l) Lakes synthesised for the first time foam statistically isotropic with a
ε
=
dieletric permittivity tensor of the honeycomb Poisson's coefficient of –0.7(3). Since that time a consistent bulk of
εc complex core permittivity value research activity has been carried out to design high structural
ε11, ε22, ε33 honeycomb permittivities in the 1,2,3 directions integrity foams(2,10), long fibre composite laminates with through-
νc core material Poisson’s ratio the-thickness negative Poisson's ratio values(2), and PTFE and
ν12, ν21 in-plane Poisson’s ratios UHMWPE polymers with high negative Poisson’s ratio values(4,7).
θ internal cell angle Between the most significant examples of auxetic material one has
ρc core material density to consider honeycombs with re-entrant unit cell geometry(9). The
σbf brittle fracture stress of the core material linear elastic mechanical properties of these solids can be predicted
σ*3 compressive collapse stress of the honeycomb using the cellular material theory by Gibson and Ashby(8), both for

Paper No. 2774. Manuscript received 19 July 2002, revised version received 11 November 2002, accepted 2 January 2003.
176 T H E A ERONAUTICAL JOURNAL M ARCH 2003

(a) (b)

Figure 1. Auxetic honeycomb configuration (a) subjected to


monoaxial loading (b).

in-plane and out-of-plane properties. The theory assumes that the in-
plane properties are described assuming the cell walls behaving like
Euler-Bernoulli or Timoshenko beams under combined bending and
axial deformation. The geometric parameters of the unit cell play an Figure 2. Geometric parameters of a honeycomb unit cell.
important role into the definition of the honeycomb’s mechanical
properties. Further theoretical developments(11) underline the fact
that hinge mechanisms should also be taken into account to predict
the in-plane properties. A combined numerical and experimental
entrant unit cell configuration, with a subsequent comparison with
analysis of the monoaxial loading of auxetic honeycombs has
results given by the analysis of Zhang and Ashby(17). The transverse
demonstrated the validity of the theoretical framework(12).
shear moduli have been derived from simulation of shear tests ASTM
A possible interest on auxetic honeycombs lies upon the fact that
C273-00, with theoretical and experimental validations, together with
the flexural rigidity and buckling load of a sandwich composite with
unit cell implicit FEM models to verify the transverse shear modulus
a negative Poisson’s ratio core could be significantly increased
behaviour on the TW plane versus the gauge thickness.
compared to a similar structure with a regular hexagonal
The electromagnetic behaviour of a honeycomb composite is
honeycomb(5). This fact could affect positively the vibroacoustic
governed by the effective permittivity of the material (for non-
design of sandwich shells with lowered cut-of frequency to increase
magnetic composites)(15). Permittivity is a measure of the material's
the transmissibility loss factors in selected frequency ranges(6).
electrical polarizability. The effective relative permittivity in the
Moreover, an auxetic honeycomb panel assumes a sinclastic curva-
three spatial directions (the honeycomb is uniaxially anisotropic) has
ture under deformation, thus leading to an increased resistance to
been calculated using a finite difference time domain technique (FD-
buckling loads and easy manufacturing of curved sandwich struc-
TD)(16,23). FD-TD is used to predict the reflection coefficient of the
tures(2). Sandwich panels are still largely in use for aircraft radome
honeycomb for the three unit cell orientations. These data are then
constructions and satellite antennae, given their reduced modal
inverted to obtain the honeycomb’s effective permittivity(16).
densities at middle frequency ranges(28). Sandwich composites may
Homogenized permittivity data from the honeycomb have been vali-
also be employed in structural aerospace components that require
dated by comparison with measured data obtained using a rectan-
low radar scattering characteristics (so-called stealth)(22). Low radar
gular wave-guide technique(16).
cross section (RCS) performance is achieved by including within the
The data from the prediction methods have been then used to
composite a mechanism for absorbing radar waves(15,22). The use of
derive parametric curves of mechanical properties versus dielectric
honeycomb in radar applications requires the composite to be truly
permittivities with different unit cell geometry layouts of the honey-
multifunctional i.e. the composite is required to exhibit enhanced
comb. These graphs can be used at the design stage to select honey-
performance in two quite distinct domains — the mechanical and
comb configurations for combined mechanical and EM target perfor-
electromagnetic.
mance. Moreover, structural and electromagnetic engineers, thus
In this paper the authors consider a compared analysis between
allowing a true multifunctional design of the sandwich structure, can
mechanical and electromagnetic properties of auxetic honeycomb
use the graphs at the same time.
structures with different unit cell geometric parameters. The
mechanical characteristics considered in this work are the Young’s
modulus and compressive strength in the transversal direction, and
the two shear moduli in the X1X3 and X2X3 planes (Fig. 2) or, alterna- 2.0 DESCRIPTION OF MECHANICAL
tively, in the LW (ribbon) and TW (transverse-to-ribbon) PROPERTIES
planes(13,18). These mechanical characteristics are essential in order
to define the fundamental mechanical performance of a sandwich The in-plane and out-of-plane mechanical properties of auxetic and
structural element(5). general honeycombs can be adequately described using the cellular
The Young’s modulus and the compressive strength in the trans- material theory (CMT) by Gibson and Ashby(5). The theory is based
versal direction have been determined through a numerical simula- on the knowledge of the mechanical properties of the constituent
tion of the ASTM C393-00 standard for flatwise compression of phase of the core material, as well as geometric parameters of the
non-metallic honeycomb cores(13). The simulation has been unit cell layout, as shown in Fig. 2. The in-plane properties are
performed using an explicit finite element model set-up with the LS- modelled considering the cell walls behaving like Euler or Timo-
DYNA 3D code(14). The model has been then prepared for a re- shenko beams in combined flexular-axial deformation. The Euler-
SCARPA E T A L M ECHANICAL AND ELECTROMAGNETIC BEHAVIOUR OF AUXETIC HONEYCOMB STRUCTURES 177

Figure 3. In-plane Poisson’s ratio v12 versus internal cell angle and Figure 4. Comparison between theoretical, FEM and experimental
wall aspect ratio. values for the transverse Young’s modulus E3.

Bernoulli approximation is considered valid for low relative density higher values of β the following formula has to be implemented(27):
values (β < 0⋅2), while for increased wall thickness shear correction
factors proper of Timoshenko beam theory are required(5). The out-
 áCosí − â 
of-plane transverse Young's modulus is calculated scaling the phase è lim = Sin −1   . . . (5)
core material modulus by the area of the load-bearing section. The  2Cosí 
two transverse shear moduli are computed using the theorems of
minimum potential energy and minimum complementary energy. Figure 3 shows the behaviour of the Poisson’s ratio ν12 versus the
The theorems give upper and lower bound values of the Voigt and internal cell angle, for various wall aspect ratio values, and a relative
Reissner type(5,21), which coincide with the shear modulus G23. The density β of 0⋅12. While the negative internal cell angle dictates the
other transverse shear modulus G23 presents two different values for auxetic behaviour of the honeycomb, the absolute values of the
the upper and lower bounds, and a numerical technique is required to Poisson’s ratio are significantly affected also by the cell wall aspect
compute an unique value considering the actual shear stress distribu- ratio α. As an example, for α = 0⋅5 and θ = –5°, the theoretical value
tion over the gauge thickness. for ν12 would be –9⋅43. For the corresponding positive internal cell
angle (θ = 5°), the in-plane Poisson’s ratio value would be 6⋅63. The
effect of shear deformation for the cell walls leads to an increase of
2.1 In-plane Poisson’s ratio values the Poisson’s ratio magnitude. For the same cases, and considering a
From the CMT results, the in-plane linear elastic mechanical proper- honeycomb core made of NOMEX aramid paper, with a Poisson’s
ties obey to the classic orthotropic relation: ratio νc = 0⋅4(17), the ν12 values would be respectively –9⋅70 and 6⋅82,
with an average increase of the 2⋅8%. The shear deformation of the
E1 í 21 = E2 í 12 . . . (1)
cell walls is particularly evident for small internal cell angles, while
for larger θ values the effect is less relevant(10). All these results indi-
Considering both in-plane flexural and axial deformation behaviour, cate that, when loading in the 1 direction, the transverse displace-
the two in-plane Poisson’s ratio can be described as following(5): ments are larger for auxetic honeycombs. This feature is in agree-
ment with the theoretical increase in deformation that a solid
Cos 2è 1 + (1 ⋅ 4 + 1 ⋅ 5ν c )â 2 composed by a positive strain energy homogeneous material with
ν12 =
( )
Sinè (á + Sinè ) 1 + 2 ⋅ 4 + 1 ⋅ 5ν c + Cot 2 θ â 2
. . . (2) negative Poisson's ratio values would exhibit when subjected to
loading(2). Applications for this large deformation behaviour of re-
entrant structures have been sought in MEMS (micro electro-
mechanical systems) technologies and microsensor components(25),
Sinè (α + Sinè ) 1 + (1 ⋅ 4 + 1 ⋅ 5 í c )â 2
ν 21 = where displacement forces are an important scale factor.
Cos è2
( )
1 + 2 ⋅ 4 + 1 ⋅ 5í c + 2 á Cos 2 θ â 2
. . . (3)

2.2 Young’s modulus in the X3 direction and compressive


Inspecting Equations (2-3) it is evident that only for negative cell strength
angles (i.e., re-entrant cell honeycombs) the in-plane Poisson's ratio
values could be negative. The convex internal cell angle is limited in The Young’s modulus of generic honeycombs has been studied in
value by the need of avoiding the mutual contact of the internal previous works with analytical models and experimental results(5,17).
vertex during in-plane deformation(11): For commercial double-thickness honeycombs the Young’s modulus
in the loading direction X3 is given as follows(8):
á 
è lim = Sin −1   . . . (4) 1+ á
2 E3 = â Ec
Cosè (á + Sinè )
. . . (6)
Equation (4) is valid for low relative density honeycombs. For
178 T H E A ERONAUTICAL JOURNAL M ARCH 2003

solid under flatwise compression.


Zhang and Ashby have outlined the different failure mechanisms
that characterise the compressive strength of honeycombs. Cell wall
buckling and axial fracture represent the two fundamental failure
mechanisms, each of them depending in particular on the relative
density of the honeycomb material. Low values of β imply cell wall
buckling failure, while for higher relative densities the honeycomb
collapses by fracture. Zhang and Ashby have investigated the case of
unit cell wall aspect ratio. For general double thickness honeycombs,
the total critical buckling load can be expressed in the following
way(8):
KEc t 3  8 . . . (8)
Pcrit = 2 + 
1 − í c2 l  á

Where K is an end constraint factor, independent of the gauge thick-


ness, and varying between 3⋅29 and 5⋅73(17). The elastic collapse
stress based on Euler buckling can therefore be calculated as:
Pcrit KEc 3 á +4 KEc 3
ó *3 = = â = â óz
Figure 5. Geometric factor versus internal cell angle and wall 2l 2 cos è (α + sin θ) 1 − í c2 á cos è (á + sin è ) 1 − í c2 . . . (9)
aspect ratio.

Figure 5 shows the nondimensional geometric factor σ−z of Equation


(9) versus the internal cell angle and different α values. For cell wall
aspect ratios equal or higher then one, the sensitivity of σ−z versus the
In the case of uniform wall thickness honeycombs, the same quantity internal cell angle is low. In particular for α = 1, the curve for θ
becomes(8): ranging from 0° to 45° can be approximated by a constant, as Zhang
2+á and Ashby have already pointed out(17). For cell wall aspect ratio
E3 = â Ec lower than 1, the re-entrant honeycomb configuration shows a steep
2Cosè (á + Sinè ) . . . (7)
increase in the collapse strength. For example, for α = 0⋅5, an
auxetic honeycomb with internal cell angle of –10° would have a
Equations (6-7) have been verified for commercial honeycomb wall buckling collapse strength 2⋅2 times higher than a correspon-
configurations, in particular with unit cell walls aspect ratio values(17). dent regular hexagonal honeycomb with the same relative density.
For more general honeycomb configurations, numerical simulations In order to validate Equation (9), a method based on explicit finite
of the flatwise compression test ASTM C365-00 have been element models has been developed to simulate the ASTM C365-00
performed using implicit finite element models with the commercial flatwise compression test for non-metallic honeycomb cores(18). The
FEM code ANSYS(19). The cell walls of the models have been numerical model represents an auxetic re-entrant honeycomb in
modelled using quadrilateral SHELL63 elements with Hermite inter- NOMEX aramid paper with unit cell wall aspect ratio and internal
polation functions. The model represented a whole double-thickness cell angle of –5º. The other geometric properties are listed in Table
honeycomb sample with the overall dimensions prescribed by the 1. The commercial finite element package LS DYNA(14) was
ASTM standard. Several models have been prepared with different selected to perform the numerical study, being the code particularly
unit cell layouts, with total number of elements ranging from 2,790 to suited to dynamic analyses including large deformations and consti-
5,400. The results have been validated against Equation (6) and avail- tutive non-linearity. Moreover, the code allows defining failure
able experimental results(17), and the outcome of the comparison is criteria, and includes contact routines useful in the simulation of the
shown in Fig. 4. As one can observe, the analytical and numerical compressive test. These features enable to simulate the behaviour of
predictions follow well the trend of the experimental results for the structure also in the phase of Eulerian buckling. The model is
various honeycomb configurations. However, there is a significant composed by 2,640 Belytschko-Lin-Tsai shell elements(14) that,
scattering of experimental results, with variation between the 20% having as reference a local co-ordinate system, avoid expensive
and 65% from the predicted values form theory and FEM simula- calculations of non-linear mechanics. Five integration points through
tions. A possible explanation for these discrepancies could be related the thickness were selected, in order to ensure an accurate represen-
to manufacturing imperfections of the samples. We observe that the tation of through-thickness stresses by enabling the simulation of the
closer comparisons between experimental and numerical results are partial and progressive failure of the cells sections. According to
for honeycombs with small positive internal cell angles, for which the Ref. 17, the core material was modelled as perfectly brittle, with a
overexpansion process is limited. However, this is only partially true, linear elastic Young modulus equal to 900MPa and collapse stress of
because for high overexpanded (OX) honeycomb (θ = 46º), the 80MPa (density ρ = 724kgm–3 and Poisson’s ratio ν = 0⋅4).
comparison is fairly good. The auxetic case (θ = –5º) in Ref. 17 gives Following the ASTM standard, the flatwise compression test was
a significant error compared to the numerical simulations. The simulated by containing the models between two flat rigid walls,
auxetic honeycomb is typically manufactured from an existing clas- moving towards each other at a relative speed of 0⋅5mms–1. Oppor-
sical honeycomb, applying an in-plane compression in order to tune master-slave contact algorithms were selected in order to obtain
induce a snapshot and create the re-entrant unit cell structure. The the load transmission between the rigid walls and the structure. To
particular manufacturing system could therefore lead to residual enforce the reality of the simulation, a friction contact model has
stresses that modify the effective bearing load area of the unit cell. been implemented between the structure and the base rigid wall. The
Other possible effects could include some degree of anisotropy for simulations were run on a SG UNIX machine with 14 336MHz
the core material, and the influence of the coating polymer and glue parallel processors and 3GB of RAM memory.
at the intersections of the cell walls. Table 1 shows the comparison between the results given by Equa-
The mesh template used to simulate the linear transverse Young’s tion (9), the explicit FEM simulation and the experimental test. The
modulus of the honeycomb has then been used to derive an explicit value of the wall aspect ratio for the experimental specimen was 1.
FEM model to simulate the compressive failure mode of the cellular The upper and lower bounds for the compressive strength depend on
SCARPA E T A L M ECHANICAL AND ELECTROMAGNETIC BEHAVIOUR OF AUXETIC HONEYCOMB STRUCTURES 179

Table 1
Comparison between theoretical, numerical and experimental values of collapse stress due to buckling

Sample Density θ Gauge b/l β σ*3 σ*3 σ*3


(kgm–3) b (mm) Equation (9) LSDYNA Experimental
(MPa) (MPa) (MPa)
Al-29-5ox 34⋅1 –5° 7⋅0 2⋅8 0⋅021 0⋅20 0⋅31 0⋅71 0⋅91

Table 2
Comparison between theoretical and numerical values of collapse stress due to brittle fracture

Sample Density θ Gauge b/l β σ*3 /σ


σbf σbf
σ*3 /σ
[kgm-3] b (mm) Equation (10) LSDYNA

Al-29-5ox 340⋅1 –5° 7⋅0 2⋅8 0⋅20 0⋅047 0⋅046

the end-conditions (simply supported or fully clamped) of the unit 2.3 Transverse shear moduli G13 and G23
cell walls. One has to note that for unitary values of α the value of
The overall bending stiffness of sandwich beams and plates is
the compressive strength given by Equation (9) is the same predicted
affected by the shear core resistance(5), while the face sheets
by Zhang and Ashby method(17). The numerical simulation gives a
contribute to increment the inertia moment of the structure cross
compressive stress value closer to the experimental one, although the section. A generic honeycomb material is a special orthotropic solid
predictedcollapse level still results lower than in experimental tests. with different shear moduli in the X1X3 and X2X3 planes(5). Accord-
This is probably due to the selected failure criteria, which is very ingto the CM Theory(8), the transverse shear modulus G13 is given
conservative and assumes that the resistance of the elements drops to by:
zero when the collapse stress is reached. A better understanding of
the failure mechanism could contribute to obtain more realistic Cosè
G13 = Gc β . . . (11)
predictions. Another factor could also be the high value for the á + Sinè
contact friction coefficient which has been introduced to enforce the
quasi-clamped conditions of the cell walls over the basement. In Fig. Also in this case, auxetic configurations lead to an increase of the
5 the stress contour maps for the above model are reported at shear modulus for some combinations of unit cell shapes. Figure 7
selected stages of the test simulation, with the progressive order shows that for small internal cell angles and reduced wall aspect
going from top-down, left-right. In particular, the maximum prin- ratio values, an auxetic honeycomb would have a steep increase in
cipal stress at the middle section is represented. the shear modulus value compared to regular hexagonal configura-
From the first figure of the sequence it is possible to observe that tion (α = 1⋅0, θ = 30º). For higher a values, the value of the shear
failure first occurs at the external segment of double thickness, modulus would be substantially constant. The values plotted in Fig.
where the stress level reduces after failure. During the test sequence, 7 come from Equation (11) and FEM simulations of the ASTM
stress concentrate in two in-plane sections close to the free edges, C273-00 shear test for non-metallic honeycomb materials(13). Also in
where the buckles due to the Eulerian load develop more signifi- this case, the FEM models represent honeycomb samples with the
cantly. With the increase of load and deformation, the stress inten- minimum overall dimensions allowed by the test standard. As in the
sity accumulates in the cell segments with double thickness. The flatwise-compressive test case, implicit quadrilateral four nodes
described stress distribution is regular and extended to the entire shells have been used to model the cell walls, with meshes having a
specimen, with exception of the external cells, especially along the total number of elements varying between 2,400 and 4,800 elements.
re-entrant sides. These arrays of cells, being partially unconstrained, The general agreement between the results is very good. The simula-
are freer to deform and tend to concentrate the stress at the median tions have been performed for honeycomb samples having a gauge
and external regions of the specimen. It is worth noticing that in the thickness ratio of 8. However, one has to bear in mind that the gauge
case of general honeycomb, previous experiences(27) showed a less thickness does affect (theoretically) the value of the transverse shear
uniform distribution, characterised by stresses concentration along modulus G13. On the contrary, for the shear modulus G23 the height
the median line, in the direction of the double thickness segments of the honeycomb affects the shear stress distribution over the cell
(see Fig. 6 and Table 1). walls(20,21). The shear modulus value G23 is contained between an
For thick brittle honeycombs the collapse stress is given by brittle upper limit (Reissner bound) and a lower value (Voigt bound)(21). A
failure(17), which is directly proportional to density of the honey- semi empirical formula to relate the two bounds into a unique value
comb and the brittle fracture stress of the core material σbf: is the following(21):
ñ
ó *3 =   ⋅ ó bf
 ñc 
. . . (10) G23 = G23
low
+
K
bl
(
up
G23 − G23
low
) w
. . . (12)
Equation (10) leads to some interesting features for auxetic honey- 0 ⋅ 787 è ≥ 0 0
combs, because their density increases for wall aspect ratios up to 2 for where K = 
1 ⋅ 342 è < 0
0
the same relative density β(11). Equation (10) has been validated also
with the same explicit FEM model developed for the buckling compres-
sive load, but this time with a relative density β of 0⋅2. Table 2 shows The values of K have been calculated through a nonlinear least
the comparison between the results from Equation (10) and the numer- square fitting algorithm from 2,400 data sets coming from FEM
ical simulation. As one can observe, the convergence is very good. homogenisation process of unit cells with different layouts. The
180 T H E A ERONAUTICAL JOURNAL M ARCH 2003

Figure 6 (c-b). Flatwise compression deformation of an auxetic honeycomb at different time steps (left-right, top-down).
SCARPA E T A L M ECHANICAL AND ELECTROMAGNETIC BEHAVIOUR OF AUXETIC HONEYCOMB STRUCTURES 181

Figure 7. Nondimensional transverse shear modulus G13 versus


internal cell angles and wall aspect ratios. ● - theoretical ◊- implicit Figure 8. Nondimensional transverse shear modulus G23 versus
FEM. internal cell angle and wall aspect ratio.

value of K has been recalculated for auxetic honeycombs because the


one available in literature(20) underestimated to steep increase of the
upper limit G23up for high negative internal cell angle configurations,
as can be observed in Fig. 8.
On the same Figure, the values from Equation (12) have been
compared with experimental results available(17). Figure 9 shows that
all results show a good convergence trend, in particular for the auxetic
configuration and the positive internal 50º cell angle. Also in this case
there is a significant scattering of the experimental results. However,
opposite to the flatwise compression test case (Fig. 4), there is good
agreement between Equation (12) and experimental results for the
extreme cases (auxetic honeycomb and OX honeycomb).

3.0 EFFECTIVE RELATIVE PERMITTIVITY


The honeycomb unit cell geometry and its construction material
govern both the mechanical and electromagnetic properties of the
macro-structure. The radar characteristics of a honeycomb composite
are governed by the material’s tensor effective relative permittivity =ε.
Honeycombs used to provide protection for sensors (radomes) require
the permittivity to be loss free, i.e. purely real(16). At relatively low Figure 9. Theoretical and experimental nondimensional transverse
electromagnetic frequencies — frequencies at which the radome's shear modulus G23 versus internal cell angle for α = 1.
thickness is less than one-quarter wavelength — it is necessary to
minimise the honeycomb’s permittivity to reduce the impact of the
window on the electromagnetic signal. However at higher frequencies,
the window’s permittivity and thickness can be chosen to exploit reso-
nance phenomena that cause the window to be highly transparent(15).
For electromagnetic wave absorbing applications, e.g. radar stealth, a  å11 0 0
 
power loss mechanism must be introduced into the honeycomb during å=0 å22 0 . . . (13)
the manufacturing process. The loss mechanism causes the permit-
tivity to become complex. A description of the complex permittivity  0 0 å33 
of a conventional honeycomb is given in Ref. 16. To optimise for
power absorption, it is necessary to control, though not necessarily In a conventional honeycomb, α and θ are fixed and the effective
minimise, the real and imaginary components of permittivity (this can permittivity is governed only by β (a measure of the amount of mate-
be done through controlling α, β and θ). rial in the unit cell) and εc (the permittivity of the material used to
The use of an auxetic unit cell provides no intrinsic benefits to the construct the unit cell). Provided β and εc are not large, the values of
radar signature designer. In an auxetic honeycomb, however, there ε 11, =ε 22 and =ε 33 are linearly related to β and εc(15). The effect on a
=

are three independent design variables (α, θ and β), compared to one conventional honeycomb of increasing β or εc is simultaneously to
(β) in a conventional honeycomb; thus the radar signature designer increase =ε 11, =ε 22 and =ε 33; =ε 11, =ε 22 and =ε 33 cannot be varied indepen-
has much greater freedom to choose an optimised permittivity, and dently. In a re-entrant honeycomb, θ and α are not fixed. From the
therefore an optimised radar signature. The effective relative permit- electromagnetic design viewpoint, control of θ and α facilitates a
tivity of a honeycomb is Ref. 23: much greater control over the honeycomb’s effective permittivity
182 T H E A ERONAUTICAL JOURNAL M ARCH 2003

Figure 10. Out-of plane honeycomb mechanical properties versus Figure 12. Honeycomb dielectric permittivities versus nondimensional
in-plane Poisson’s ratio. transverse Young’s modulus.

materials complex permittivities have been measured separately, and


the values used within the numerical models. The comparison
between the homogenised data and the equivalent permittivities of the
samples gave very good results(15,23). In particular, for the auxetic cast
epoxy resin samples, the experimental findings confirmed the strong
dielectric anisotropy of the cellular solid(23).

4.0 COMMENTS
The results from literature and tests carried out by the present
workers on the rectangular waveguide outline the validity of the
numerical simulations about the compressive and shear mechanical
properties of the honeycombs, as well as the finite difference-time
domain technique developed to predict the dielectric properties of
the cellular solids. Honeycomb materials are used in absorber and
radome applications(22) and, more general, in airframe constructions.
The possibility of increasing the degree of anisotropy both from the
mechanical and dielectric point of view using auxetic honeycombs
Figure 11. Honeycomb dielectric permittivities versus in-plane allows the aerospace structural and electromagnetic analyst to intro-
Poisson’s ratio. duce more degrees of freedom into the integrated design of aero-
space components.
Figure 10 shows the behaviour of the out-of-plane mechanical prop-
erties of general honeycombs (Young’s modulus in the X3 direction
than is possible in a conventional honeycomb. and transverse shear moduli G13 and G23) with wall aspect ratio α =
A FD-TD -based method utilising periodic boundary conditions 2⋅4 and relative density β = 0.12 versus the honeycomb Poisson's coef-
has been developed to predict the effective permittivity of a dielectric ficient ν21. The normalised gauge thickness is 3. For increasing nega-
honeycomb(15,23). The reflection coefficient is predicted for the case tive Poisson’s ratio values the transverse shear modulus G23 increases,
with a minimum value around ν21 = 0, corresponsing to a rectangular
of an infinite planar layer of dielectric honeycomb subject to
unit cell honeycomb(8). The transverse shear modulus G23 tends to
planewave illumination from three mutually orthogonal directions.
increase once again for positive Poisson’s ratio values. It is worth
The reflection data are then inverted to obtain data on the honey- noticing that the magnitude of the transverse shear modulus values G23
comb’s effective permittivity. The data inversion process is based on remain higher in the negative Poisson's ratio field. This behaviour is
a conventional planewave transmission line model of the honeycomb. less marked for high gauge thickness, because the shear modulus
The inversion process homogenises the unit cell, replacing the inho- approaches the lower bound, which decreases for negative internal cell
mogeneous anisotropic material with an equivalent homogeneous angles(21). The transverse shear modulus G13 shows a slight increment
material. Using the homogenous model, the unit cell parameters can for negative Poisson’s ratio values, but these variations are negligible
be chosen to yield an appropriate permittivity characteristic for a for wall aspect ratios higher than 1⋅5(21). The Young's modulus shows
given radar absorption or radar window problem. Moreover, by large increase in magnitude for negative Poisson's ratio values, with
combining the electromagnetic and mechanical homogenised models, increments up to the 230% compared to cell layouts with correspon-
global optimisation for all engineering functions can take place. dent positive internal cell angles(21). However, the increase of trans-
The results from the FD-TD technique have been validated against verse Young’s modulus is accompanied by a correspondent increase in
experimental data gathered on NOMEX and cast epoxy resin samples honeycomb density(11). In any case, all linear elastic transverse
tested in a rectangular waveguide facility(16,23). The effective core mechanical properties scale linearly with the relative density β.
SCARPA E T A L M ECHANICAL AND ELECTROMAGNETIC BEHAVIOUR OF AUXETIC HONEYCOMB STRUCTURES 183

Figure 11 shows the behaviour of the relative permittivities in the x, REFERENCES


y and z directions for the same honeycomb of Fig. 10, versus the
1 TIMOSHENKO, S.P. History of Strength of Materials, 1953, McGraw-
Poisson’s ratio ν21. The relative permittivity of the core material is εc = Hill, New York.
4⋅0. The Poisson’s ratio coefficients ν21 are calculated by means of the 2 EVANS, K.E. Auxetic polymers: a new range of materials, Endeavour,
CMT theory, while the relative permittivities are computed with the New Series, 1991, 15, (4), pp 170-174.
FT-TD technique. From the graph it is noticeable the almost linear 3 LAKES, R. Foam structures with a negative Poisson’s ratio, Science,
dependency of the dielectric properties versus the Poisson’s ratio, with 1987, 235, pp 1038-1040.
increase permittivities in the x and y directions for negative ν21 values. 4 EVANS, K.E. Microstructural modelling of auxetic microporoues poly-
The εz permittivity is more sensitive to negative Poisson’s ratio values mers, J Mat Sci, 1995, 30, pp 3319-3332.
compared the other permittivities in xy plane. The sensitivity increase 5 SCARPA, F. and TOMLINSON, G. Investigation on negative Poisson’s ratio
honeycombs, 1998, Fourth seminar on experimental techniques and
of εz is matched by a correspondent increase in the transverse Young’s design in composites, FOUND, M.S. (Ed), 1-2 September1998, Sheffield.
modulus E3 in the negative Poisson’s ratio domain. Figure 12 shows 6 SCARPA, F. and TOMLINSON, G. Theoretical characteristics of the vibra-
for the same honeycomb the behaviour of the dielectric properties tion of sandwich plates with in-plane negative Poisson’s ratio values, J
versus the ratio of the transverse Young's modulus with the core mate- Sound Vib, 2000, 230, (1), pp 45-67.
rial one. For high values of E3/Ec there is a correspondent increase of 7 ANDERSON, K.L. and EVANS, K.E. The fabrication of microporous poly-
the permittivity values, with a strong effect for the transverse εz ethylene having a negative Poisson’s ratio, Polymer, 1992, 33, (20), pp
permittivity. For each unit cell configuration the transverse permit- 4435-4438.
tivity εz is increased by an average 15% compared to the in-plane 8 GIBSON, L.J. and ASHBY, M.F. Cellular Solids-Structure and Properties,
Second edition, 1997, Cambridge Press, UK.
values εx and εy. This increase is substantial for high negative
9 CLARKE, J.F., DUCKETT, R.A., HINE, P.J., HUTCHINSON, I.J. and WARD,
Poisson’s ratio values, with a rise up to 43%. These results suggest I.M. Negative Poisson’s ratios in angle-ply laminates: theory and exper-
that the mechanical anisotropy associated to auxetic honeycombs is iment, Composites, 1994, 25, (9), pp 863-868.
accompanied by a dielectric properties anisotropy, very much accentu- 10 WARREN, W.E and KRAYNIK, A.M. Foam mechanics: the linear elastic
ated for extreme negative Poisson’s ratio values in the plane X1X2, response of two-dimensional spatially periodic cellular materials,
where high transverse Young’s modulus values are present. Mechanics Mater, 1987, 6, pp 27-31.
A re-entrant auxetic honeycomb is a particular subset of a general 11 MASTERS, I.G. and EVANS, K.E. Models for the elastic deformation of
honeycomb structure having orthotropic symmetry. It is well honeycombs, Comp Struct, 1996, 34, (4), pp 403-422.
12 SCARPA, F., PANAYIOTOU, P. and TOMLINSON, G. Numerical and experi-
known(8) that solids having this kind of symmetry possess other
mental uniaxial loading on in-plane auxetic honeycombs, J Strain
Poisson’s ratio values. The CMT theory states that the out-of-plane Analysis, 2000, 35, (5), pp 383-388.
Poisson’s ratio coefficients are not affected by the presence of a 13 ASTM Standard E 1091-98, Standard specification for nonmetallic
negative internal cell angle. Moreover, the Poisson’s ratios and are honeycomb core for use in shelter panels.
proportional to the Poisson’s ratio of the core material, while the 14 HALLQUIST, J.O. LS-DYNA theoretical manual, 1998, Livermore Soft-
coefficients and are virtually zero(8). The present authors have not ware Technology Corp.
investigated these results from a numerical or experimental point of 15 BETTERMANN, J. and WENTZEL, H. Design features and test of low
view, but note that the conclusions of the CMT theory regarding the observable structures for large aircrafts, 1993, Third international
conference on electromagnetics in aerospace applications, 14-17
out-of-plane Poisson's ratio values satisfy the reciprocity require-
September 1993, Torino, Italy.
ments for solids with symmetric orthotropy. To the knowledge of the 16 SMITH,. F.C. Effective permittivity of dielectric honeycombs, Proceed-
presente workers, no negative out-of-plane Poisson’s ratio value has ings of Institute of Electrical Engineers, Microwave antennas,
been detected in auxetic re-entrant honeycomb structures. Propagat, 1999, 146, (1), pp 55-59.
17 ZHANG, J. and ASHBY, M.F. The out-of-plane properties of honeycombs,
Int J Mech Sci, 1992, 34, (6), pp 475-489.
18 ASTM Standard C365-00, Standard test method for flatwise compres-
CONCLUSIONS sion properties.
19 ANSYS Rel. 5.6 User’s Manual, 2000, Swanson.
Unit cell re-entrant auxetic honeycombs have been analysed from 20 GREDIAC, M. A. finite element study of the transverse shear in honey-
the mechanical and dielectric point of view to assess their use in comb cores, Int J Solids Struct, 1993, 30, pp 1777-1788.
sandwich constructions for electromagnetic shield applications. By 21 SCARPA, F. and TOMLIN, P.J. On the transverse shear modulus of nega-
means of their geometric unit cell layout these honeycombs offer tive Poisson’s ratio honeycomb structures, Fatique Fractures Eng Mat
increased transverse Young’s modulus compared to analogous Structures, 2000, 23, pp 717-720.
regular hexagonal honeycombs with the same relative density. The 22 JOHNSON, R. Radar-absorbing material: a passive role in an active
transverse shear moduli show increased sensitivity versus in-plane scenario, 2001, http://www.randf.com/ramapriaas.html.
negative Poisson’s ratio values. Relative permittivities in the three 23 SMITH, F.C., SCARPA, F. and CHAMBERS, B. The electromagnetic proper-
ties of re-entrant dielectric honeycombs, IEEE Microw, Guided Wave
main directions show also a global increase versus negative Pois- Lett, 2000, 10, (11), pp 451-453.
son's ratio values. This trend is particularly significant for the rela- 24 CLARKSON, B.L. and RANKY, M.F. Modal density of honeycomb plates,
tive transverse permittivity in the z direction, with increased dielec- J Sound Vib, 1983, 91, pp 103-118.
tric anisotropy. Regular hexagonal honeycombs do not allow for this 25 LARSEN, U.D., SIGMUND, O. and BOUWSTRA, S. Design and fabrication
distinct variation of the anisotropy of the composite. These proper- of compliant mechanisms and structures with negative Poisson’s ratio, J
ties of auxetic honeycombs provide more degrees of freedom to Microelectronic Syst, 1997, 6, (2), pp 99-106.
design globally optimised low observable aerospace components. 26 SCARPA, F., REMILLAT, C., LANDI, F.P. and TOMLINSON, G. Damping
They also allow radomes and other electromagnetic structures to be modelisation of auxetic foams, 2000, Smart structures and materials,
TUPPER HYDE (Ed), pp 336-343, 6-8 March 2000, Newport Beach, CA.
more closely tailored to a specific application. 27 SCARPA, F., BURRIESCI, G., SMITH, F.C. and CHAMBERS, B. Comporta-
mento meccanico e dielettrico di strutture a nido d’ape auxetiche, 2001,
XXX AIAS conference, pp 621-630, 11-15 September 2001, Alghero,
ACKNOWLEDGEMENTS Italy.
28 FERGUSON, N.S. and CLARKSON, B.L. The modal density of honeycomb
This research has been partially funded through the EPSRC projects shells, ASME J Vib Acoustics, 1986, 108, pp 399-404.
GR/K/89368/01 and GR/R/98108. The authors would like to thank
the company Ove Arup & Partners for the permission of use of the
LS-DYNA 3D software. The authors would like also to thank the
anonymous referees for their useful comments.

You might also like