You are on page 1of 10

View Online / Journal Homepage / Table of Contents for this issue

267 1

J. CHEM. SOC. FARADAY TRANS., 1995, 91(16), 2671-2680

Auxetic Two-dimensional Polymer Networks


An Example of Tailoring Geometry for Specific Mechanical Properties

Downloaded by Universidade do Minho (UMinho) on 02 October 2012


Published on 01 January 1995 on http://pubs.rsc.org | doi:10.1039/FT9959102671

Ken E. Evans,* Andrew Alderson? and Frances R. Christian


School of Engineering, University of Exeter, North Park Road, Exeter, UK EX4 4QF

The Poisson's ratios and Young's moduli of 2D molecular networks having conventional and re-entrant honeycomb forms have been modelled using molecular modelling. Three principle deformation mechanisms were
observed : bond hinging, flexure and stretching. Analytical models have also been developed that can be used to
describe each of these modes of deformation acting either independently or concurrently. A parametric fit of the
force constants in the concurrent analytical model calculations to the molecular model calculations yields good
agreement in the mechanical properties for all the structures studied. Specific trends in the force constants
required to fit the data are observed. Consequently, a force constants library can be compiled and has been
used to predict accurately the properties of more complex variants of the networks. This semi-analytical sub-unit
approach enables a more efficient use of computer-intensive molecular modelling programs.

1. Introduction
This paper considers one example of how the geometric
arrangement of a polymer network may be tailored to
achieve specific macroscopic mechanical properties. The
example considered here is the development of auxetic molecular structures ; polymer networks that exhibit a negative
Poisson's ratio. At the moment this work is in its early stages.
Some research has been published on the design of novel
polymeric networks that exhibit this unusual effect '-' and
some work has been progressed on the synthesis of such netw o r k ~ .Applications
~,~
for such materials have also been identified.7-9 As yet, no molecular structure has been modelled,
synthesised and characterized, but we expect that this will not
be long in coming. The development of a molecular network
with a negative Poisson's ratio is an excellent example of
understanding form to obtain a particular function. The first
examples of structures (as opposed to materials) with auxetic
functionality (the ability to expand laterally when stretched)
are to be found in the analysis of the mechanical properties of
Here, one further attribute other than form
honeycombs.".'
is required to demonstrate a function and that is the mode of
deformation. So, in the first analysis," the form, a hexagonal
honeycomb, was assumed to deform in a particular manner,
by flexure, leading to the mechanical property, a positive
Poisson's ratio. By modifying the form, using a re-entrant
honeycomb (see Fig. l), but maintaining the same mode of
deformation, a negative Poisson's ratio results.
Classical elasticity is not a scale-dependent formulation
and none of the simple analyses of structure that have been
used to predict negative Poisson's ratios contain any scaling
factors. One may therefore assume, naively, that geometric
structures and deformation mechanisms that apply at the
macrostructural level may also apply at the microstructural
and molecular level.
This assumption has been proved, experimentally, in the
microstructural case by the synthesis and testing of a novel
form of auxetic polyethylene.' 2*1 This paper considers the
further use of this approach at the molecular level. Of course,
at some length-scale classical mechanics will break down and
quantum mechanics take over. However, for the polymer networks considered here, classical mechanics is still an ade-

quate approximation, as embodied in many commonly used


molecular mechanics programs, where atomic forces and
long-range interactions are modelled by spatial and angulardependent, classical force constants. These issues have been
thoroughly covered e l s e ~ h e r e . ' ~ * ' ~
In the next section we introduce and develop the molecular
and analytical models used to describe the deformation of
polymer networks having conventional and re-entrant honey(a 1

'

f Current address : British Nuclear Fuels plc, Company Research

Laboratory, B5 16, Springfields Works, Preston, Lancashire, UK PR4


OXJ.

Fig. 1 Honeycomb cell geometry and co-ordinate system used in


the analytical models for (a) conventional and (b) reentrant molecular networks

View Online

2672

J. CHEM. SOC. FARADAY TRANS., 1995, VOL. 91

Downloaded by Universidade do Minho (UMinho) on 02 October 2012


Published on 01 January 1995 on http://pubs.rsc.org | doi:10.1039/FT9959102671

comb forms. The analytical models are based on the networks deforming by hinging, flexure and stretching of the
honeycomb cell walls, with each mode acting independently
of the others in the simplest (single-mode) models. A more
complex (multi-mode) analytical model where all three modes
act concurrently is then developed. Comparison between the
analytical and molecular model predictions of the mechanical
properties of the networks is then made and the results discussed accordingly.

2. Models
In this section we first describe the molecular model used to
predict the Poissons ratios and Youngs moduli of the molecular network structures and then the analytical models. In all
cases expressions for the Poissons ratios and Youngs moduli
are derived.
2.1 Molecular Model
Examples of the earliest conventional and re-entrant molecular network sub-units designed, are shown in Fig. 2(a) and
(b), respectively. The networks consist of branches of acetylene groups joined by benzene rings at the junctions. Each
benzene ring has three polyacetylene arms connected to it.
The connectivity of the polyacetylene arms to the benzene
rings determines the honeycomb geometry. When the arms
are connected to alternate benzene ring carbon atoms the
conventional honeycomb geometry is realised, whereas the
re-entrant structure is produced when the arms are connected
to three adjacent benzene ring carbon atoms. A naming convention has been adopted: (n,m)-flexyne refers to the conventional honeycomb structure, where n and M are the number
of acetylene links on the diagonal and vertical branches,
respectively. (n,m)-reflexyne refers, similarly, to the re-entrant
structure. Hence the sub-units shown in Fig. 2(a) and (b) are
(1,4)-flexyneand (1,4)-reflexyne,respectively.
The mechanical properties of these molecular networks
were determined using the POLYGRAF molecular modelling
programI6 (version 2.20) employing the DREIDING force
field,17 which is a well established package for predicting a
range of properties for polymeric materials. An infinite
system was approximated by periodically extending the
repeat unit in the x and y directions. The energy minimisation process (using the method of conjugate gradients)
described in ref. 2 was adopted in this work. For each structure, minimisations were performed for the undeformed structure and for uniaxial applied loads of kO.5 GPa in each of
the x and y directions. Previous calculations on a wider range
of stress increments have shown that this is sufficient to give
accurate results. Values of Poissons ratio ( v i j ) and Youngs
modulus (Ei) were then evaluated from the stress and strain
data with
vij

= -Ej/Ei

Ei = b i / E i

where bi is the stress applied in the i ( = x or y ) direction and


and c j are the true strains calculated using

E~

ci = ln(I/Z,)

(3)

where I (= X or Y) and I, (= X, and Yo) are the deformed


and undeformed repeat unit-cell lengths in the i direction.

2.2 Single-mode Analytical Models


In the derivations that follow we assume unit thickness in the
z-direction (perpendicular to the x-y plane).
Flexure Model
The molecular networks described in Section 2.1 are based

Fig. 2 MolecuIar network sub-units: (a) (1,4)-flexyne, a positive


Poissons ratio molecular network; (b) (1,4)-reflexyne, a negative
Poissons ratio molecular network

on conventional and re-entrant honeycomb networks. Hence,


if we neglect the detailed molecular structure, it is possible to
derive expressions for the elastic properies of these networks
using conventional honeycomb theory., In this case the
network sub-units are represented by the conventional and
re-entrant honeycomb cells shown in Fig. l(a) and (b), respectively, and the deformation is due to flexure of the diagonal
honeycomb arms. The repeat unit-cell lengths in the x and y
directions are given by

x = 21 cos e
Y = 2(h + I sin 0)

(4)
(5)

where h and I are the lengths of the vertical and diagonal


arms, respectively. 8 is the honeycomb angle which is positive
for the conventional cell, and negative for the re-entrant cell,
see Fig. 1. The Poissons ratios and Youngs moduli associated with these networks have been developed elsewhere
and are quoted directly here:

2.2.1

VXY

= vy;

cos e x
sin 0 Y

= --

View Online

2673

J. CHEM. SOC. FARADAY TRANS., 1995, VOL. 91

For an x directed load we have

(7)
and

Downloaded by Universidade do Minho (UMinho) on 02 October 2012


Published on 01 January 1995 on http://pubs.rsc.org | doi:10.1039/FT9959102671

K , is the flexure force constant governing the flexure of the


ribs. For ribs of length I, thickness t, depth w and intrinsic
Youngs modulus E,, K , is found from standard beam
theory to be

K , = E, wt3/l
(9)
From eqn. (6) we note that positive Poissons ratios are realised owing to flexure when 8 is positive (conventional
honeycomb) whereas a negative value of 0 (re-entrant
honeycomb) yields negative Poissons ratios.
For an orthotropic material to have a symmetric stiffness
matrix and a positive definite strain energy for static equilibrium we require
vxy

Ey

= F,

sin 8/2
(23)
where F, is the force applied in the x direction which, for unit
thickness in the z direction, is given by

F, = 0, Y

(24)

From eqn. (21)-(24) we have


d0 = -(IY sin 8/2K,) dux
(25)
Given the possibility of non-linear elastic behaviour we use
the tangent Youngs modulus20
Ex = dadds,
(26)
From eqn. (14), (16), (25) and (26) the Youngs modulus for
unit thickness in the z direction is

= v y x Ex

and

I v x y I d (EJE,)

Similarly,

From eqn. (6)-(8) we have


vxyE, = vyxEx = K,/lz sin 8 cos 0

sin 0 Y
cos e

vyx = --

(12)

and

and

Ivxy I

= (EJEy)2

(13)

Hence the flexure model satisfies the requirements of a symmetric stiffness matrix [eqn. (lo)] and a positive strain energy
Ceqn. (1 113.
2.2.2 Hinging Model
We now consider the conventional and re-entrant honeycombs shown in Fig. 1 deforming by hinging of the honeycomb cell walls, i.e. by varying the honeycomb angle 0. From
eqn. (4) and (5) we have

dX

-21 sin 0 d8

(14)

dY = 21 cos 0 d8

(15)
The increment of true strain in the x direction is defined by

(16)
dE, = dX/X
with a similar expression for the y direction. In the case of an
x-directed load the Poissons ratio is defined by
-d~,,/d&,
(17)
which holds for non-linear as well as linear elastic behaviour.
Giving, from eqn. (4), (5), (16) and (17):
vXy =

cos 0 x
sin 8 Y
To derive the Youngs moduli due to hinging we introduce
a hinging force constant Kh defined in the usual manner by
vxy = --

M = Kha

(19)

where M is the moment applied to a diagonal honeycomb


arm and a is the angular displacement of the arm due to M.
M i s given by

M = IF

(20)

where F is the force applied perpendicular to the diagonal


arm of length 1. Hence for an infinitesimal increment in
applied force the change in angular displacement is
da

=1

dF/Kh

(21)

Note that the hinging-model Poissons ratio expressions


[eqn. (18) and (28)] are identical to the flexure model expressions [eqn. (6)]. Furthermore, the hinging-model Youngs
moduli [eqn. (27) and (2911 differ only by the nature of the
force constant from those for the flexure model [eqn. (7) and
(8), respectively]. Hence eqn. (10) and (11) will apply.

2.2.3 Stretching Model


The final single-mode analytical model we consider is deformation due to stretching of the arms of the honeycomb
network shown in Fig. 1, i.e. by varying 1 and h. From eqn.
(4) and (5) we have
axpi

=2

cos

aY/ah = 2

(30)
(3 1)

a Y / a l = 2 sin 8
(32)
The changes in the unit-cell lengths due to infinitesimal
increments ds, and ds, in the lengths 1 and h, respectively are
then
d x = (ax/azyds,

(33)

dY = (aY/dl) ds, + (aY/ah)ds,


(34)
Consider a load applied in the x direction. In this case the
vertical ribs (length h) remain at constant length since there is
no resultant component of the applied force along the length
of these ribs, i.e.
ds,

=0

(35)

Therefore, from eqn. (30)-(35) we have


cos 8 ds,

(36)

dY = 2 sin 8 ds,
and hence the Poissons ratio vXyis

(37)

dX

=2

VXY =

-sin 0 X
cos e Y

--

View Online

J. CHEM. SOC. FARADAY TRANS., 1995, VOL. 91

2674

For a y-directed load the force F,applied along the length of


the diagonal ribs is

The stretching force constant is defined by


F = K,s
(39)
where F is the force applied along the length of an honeycomb arm of length s. K, can be related to the intrinsic
material Youngs modulus and dimensions of the arm. For
example, consider the increment of extension ds of an arm of
length s, thickness t, depth w and Youngs modulus E, due to
an infinitesimal increase d F in the force applied along the
length of the arm. From Hookes law

Downloaded by Universidade do Minho (UMinho) on 02 October 2012


Published on 01 January 1995 on http://pubs.rsc.org | doi:10.1039/FT9959102671

K , = E,wt/s
(41)
For honeycombs consisting of homogeneous or uniform
material then s = I (and ds = ds,) in the case of the diagonal
arms at angle 8. However, as will be shown, in the stretching
model the molecular structure is more important than it was
in the hinging and flexure models. It becomes necessary to
consider which parts of the honeycomb network undergo
stretching. Molecular model calculations (see Section 2.1)
were performed for (2,2)-flexyne for stresses of -2.0, 0 and
+2.5 GPa applied along the y direction. It was found that ca.
62.5% of the total extension of the vertical arms was due to
extension of the bonds connecting the acetylene arms to the
benzene rings. A further ca. 19% of the extension was due to
distortion of the benzene rings themselves, with only ca. 19%
of the extension due to deformation of the acetylene
branches. In other words, ca. 80% of the extension of the
honeycomb arms was found to be related to the benzene ring
junctions (benzene ring and connecting bond). Hence, the
length s in eqn. (41) may not be directly related to either I or
h since the molecular structure introduces inhomogeneity
into the material forming the honeycomb networks on this
length-scale. As a first approximation then, we neglect acetylene branch elongation and consider all the extension in the
stretching model to be due to the benzene ring and connecting bond. Hence, since there are equal numbers of benzene
ring junctions for the vertical and diagonal arms (two) in all
cases, then s is the same in both arms and, therefore, from
eqn. (41) the stretching force constant for both types of arm is
equal.
In the case of an x-directed load, the force applied along
one of the diagonal honeycomb arms is given by

and we have
s = s,
(43)
where s, is the active length (i.e. benzene ring junction) of the
diagonal arm. From eqn. (39), (42) and (43) we have

(44)

Hence the increment of strain in the x direction is, from eqn.

(W, (301, (33) and (44)


ds,

= ( Y COS

8/K,X) do,

(45)

and from eqn. (26) and (45) the Youngs modulus in this case
is

Now consider a load applied in the y direction. In this case


the vertical ribs also extend, i.e.
dSh # 0

(48)

and that along the length of the vertical ribs (Fh) is


Fh

(49)

= Xo,

Therefore, from eqn. (39), (48) and (49) we have


ds, = (X sin 8/2K,) do,

(50)

dsh = (X/K$ do,


(51)
Hence, from eqn. (30)-(34), (50) and (51) the changes in the
unit-cell lengths due to an infinitesimal incremental change
do, in a y-directed load are

Therefore, from eqn. (39) and (40)

ds, = (Y cos 8/2K,) do,

F,= (X sin 8/2)o,

(47)

cos 8 sin 8/KJ do,

(52)

dY = [X(2 + sin 8)/K,] do,

(53)

dX

= (X

and the Poissons ratio vyx is, therefore,


Vyx =

cos 8 sin 8 Y
(2 + sin e)

(54)

The increment of strain in the loading direction is in this case


ds,

= [X(2

+ sin

8)/K,y3 do,

(55)

and hence the Youngs modulus is


E,

K,
y
(2 + sin 8) x

From eqn. (38) and (54) we notice that the stretching model
predicts negative Poissons ratios when 8 is positive, whereas
positive Poissons ratios are realised for negative values of 8.
In other words, the stretching model predicts the opposite
sign of Poissons ratio to the flexure and hinging models. We
also note from eqn. (38), (46),(54) and (56)
vXyE , = v,,E , = -K , sin 8/[cos 8(2

+ sin

8)]

(57)

8)1/2X]/(cos 8Y) > I vXy I

(58)

and
(EJEy)l = [(2

+ sin

and hence the conditions for a symmetric stiffness matrix


[eqn. (lo)] and a positive definite strain energy [eqn. (ll)] are
satisfied by the stretching model.

2.3 Concurrent Analytical Model

In a real molecular structure all three modes of deformation


may operate. Hence a multi-mode analytical model for the
networks where hinging, flexure and stretching act concurrently is developed.
In the derivations that follow we once again assume unitthickness in the z direction.
The total incremental change dX in the unit-cell length X
due to an incremental change in the applied load for the concurrent model is simply given by the sum of the incremental
changes due to each of the individual modes of deformation,
i.e.

dX = dXh + dX,

+ dX,

(59)
where dX,, dX, and dX, are the incremental changes in X in
the hinging, stretching and flexure models, respectively. Similarly, the change in Y is given by
dY = dYh + dY, + dY,
(60)
Consider the case of an x-directed load. We have for the
flexure model

dX,

= (YZ

sin 8/K,) do,

dY, = -(YZ2 sin 8 cos B/K,) do,

(61)
(62)

View Online
2675

J. CHEM. SOC. FARADAY TRANS., 1995, VOL. 91

In the case of the hinging model we find from eqn. (14), (15)
and (25)
dxh

= (Y1

Sin 8/Kh) da,

dY, = -(Y1

(63)

Sin 8 COS 8/Kh) da,

(64)

For the stretching model eqn. (36), (37) and (44)yield


dX,

= ( Y COS

8 / K s )do,

(65)

dY, = (Y sin 8 cos O/KJ da,

(66)

Substituting eqn. (61), (63) and (65) into (59) gives the total
change in X :

Downloaded by Universidade do Minho (UMinho) on 02 October 2012


Published on 01 January 1995 on http://pubs.rsc.org | doi:10.1039/FT9959102671

dX = Y[(sin e/Khf) + (cos 8/K,)] do,

(67)

where Khf is a function of the hinging and flexure force constants, defined by
Khfl = 12(K,

+ K , )

(68)

Similarly, from eqn. (60), (62), (64) and (66) we have


dY

Y sin 6 cos 0(Ks- - Khfl) da,

(69)

Eqn. (16) (and the de, equivalent), (17), (26), (68) and (69) give
a Poissons ratio v,, of
vxy

- 11 X
sin 8 cos O[(Ks/Khf)
[(K$KM)sin2 8 + 60s 01 Y

(70)

and a Youngs modulus E x of

Eqn. (70) illustrates that the sign of vxy is now dependent on


the geometry (0) and the force constants ratio ( K J K h f ) ,e.g.
negative and positive values of v,, are realised for 8 positive
with KJK,, < 1 and >1, respectively. That is, the sign and
magnitude of the Poissons ratio is both geometry- and
mechanism-dependent .
Similarly, for a y-directed load we have

Eqn. (70)-(73) yield


Vxy

Ey

= VyxEx
-

K , sin 8 cos 8[(:(K$Khf) - 11


[(K$Khf)sin 8 COS el
x [(K,/K&OS
8 + 2 + sin

and, therefore

(74)

3. Results
3.1 Molecular Model Data
Molecular model calculations were performed on a total of
12 structures with n = 1 or 2, 2 < rn d 8 and 8 = +3W. For
each structure, energy minimisations were performed for the
undeformed (zero stress) configuration and with uniaxial
stresses of k0.5 GPa applied in each of the x and y directions. Previous calculations on a wider range of stress
increments have shown this to give accurate results. The
Poissons ratio and Youngs modulus data were then calculated following the procedure outlined in Section 2.1. In most
cases different values of each elastic constant are realised for
tension and compression. This has been discussed elsewhere,
where strain-dependent effects have been considered.
However, in this paper we are concerned only with the elastic
moduli of the undeformed state (although the molecular and
analytical models can be used to calculate strain-dependent
mechanical properties) and so the average of the values due
to tension and compression will be compared to the analytical model calculations.
In Table 1 we present the average values of the elastic
moduli calculated from the molecular model for each structure, with an error estimated from the difference between
tension and compression. We also show the calculated undeformed unit-cell lengths for each structure. For reasons to
become apparent later, the data due to loading in the x direction for the (2,rn)-reflexyne structures were not averaged (i.e.
tension and compression in the x direction were considered
separately for these structures). The v,, E , us. vyx E x data,
evaluated from the average elastic moduli in Table 1, are
plotted in Fig. 3. With the exception of the data point corresponding to compressive loading in the x-direction for (2,5)reflexyne, the data are scattered about the equality line within
the calculated uncertainties. Hence the molecular model data
satisfy the requirement of a symmetric stiffness matrix [eqn.
(lo)]. Furthermore, the data are grouped into four clusters
with each cluster characterised by the value of n and 8 [i.e. (n,
8) = (1, +30), (1, -30), (2, +30) and (2, -30)]. This will
be discussed later.
It is readily shown that the data in Table 1 also satisfy the
condition of a positive definite strain energy for static equilibrium [eqn. (1l)].

el
3.2 Comparison of Analytical Models with Molecular Model

and

I v,, I -= (EX/E,)

(75)
The expressions for the elastic moduli involve the unit-cell
dimensions X and Y which are given by eqn. (4) and (9,
respectively, and are dependent on the honeycomb angle 8
and the cell-wall lengths h and 1. In order to relate X and Y
to the molecular networks it is necessary to determine h and 1
in terms of the molecular dimensions. We define A and B to
be the acetylene and benzene junction lengths defined in
Fig. 2(a). A is simply the length of one acetylene link. B/2 is
the sum of the radius of the benzene ring and :he length of
that part of the connecting bond not associated with an
acetylene link [i.e. where most of the stretching is found to
occur, see Section 2.2.3 and Fig. 2(a)]. Hence
l=B+nA

(76)

h=B+rnA

(77)

The concurrent analytical model Poissons ratio expressions


can be rearranged to give the force constants ratio K J K ,
required to fit the analytical model to the molecular model
data. From eqn. (70) and (72) we have
Ks/Khf

[(vxy/tan

and

exy/x)

-k

l1/C1 - xy tan e(y/x)l (79)

[(2 + sin 8)v,, + (Y/x)sin 8 cos 81


(80)
[(Y/X)sin e cos e - cosz ev,,]
For each structure four values of KJK,, , required to fit the
concurrent analytical model Poisson7s ratio data to the
molecular model data, were evaluated using eqn. (79) and
(SO), corresponding to tensile and compressive uniaxial
loading in each of the x and y directions. With the notable
exception of tensile loading in the x direction for the (2,m)reflexyne structures, the values of K J K , were found to be
KJKhf

View Online

J. CHEM. SOC. FARADAY TRANS., 1995, VOL. 91

2676

TaMe 1 Average Poisson's ratios (v,, and vyx)and Young's moduli [ E x and E, (GPa)] calculated from the molecular model calculations.
Undeformed unit-cell lengths Xoand Yo(A) are also presented
structure

vxY

VYX

(12)-flexyne
( 1,4)-flexyne
(1,6)-flexyne
(2,2)-flexyne
(2,4)-flexyne
(2,6)-flex yne
( 1,4)-reflexyne
(1,5)-reflexyne
( 1,6)-reflexyne
(2,5)-reflexyne

0.46 [6]
0.32 [3]
0.24 [3]
0.85 [2]
0.60[4]
0.47 [2]
-0.29 [8]
-0.29 [S]
-0.22 [S]
-0.69"
- 1.04'
-0.72"
-0.88'
-0.50"
-0.Mb

0.696 16)
0.9 [l]
0.99 [4]
0.84 [S]
1.11 [S]
1.33 [4]
-0.29 [2]
-0.386 [2]
-0.42 [3]
-0.53 [4]

Downloaded by Universidade do Minho (UMinho) on 02 October 2012


Published on 01 January 1995 on http://pubs.rsc.org | doi:10.1039/FT9959102671

(2,6)-reflexyne
(2,8)-reflexyne

-0.70 [3]
-0.902 [2]

EX

EY

XO

YO

75 c41
56.2 [4]
45 c13
30 C5l
23 c21
19 c41
124 [63
95 c11
84 c41
57.4"
44.7'
35.8"
35.8*
31.4"
31.5'

120 [lo]
160 [lo]
220 [40]
30 c31
42 c41
55 c41
110 [lo]
116 [6]
140 [lo]
21 c41

11.4709
11.4735
11.4671
15.6314
15.5975
15.6375
11.7017
11.7102
11.7151
15.8858

24.6685
34.2045
43.7892
27.0193
36.6377
46.1 108
21.3666
26.1698
30.9691
23.8009

31 c31

15.9271

28.6706

40 ClOl

15.9098

38.1895

Numbers in square brackets are uncertainties in least significant figures. In the case of the (2,m)-reflexyne structures under loading in the x
direction actual values calculated due to compression' and tensionb are quoted (see text).

dependent on the number of acetylene links in the diagonal


branches (n)and on the sign of the honeycomb angle (O), i.e.
conventional or re-entrant honeycomb, but independent of
the number of acetylene links in the vertical branches (m).For
the (2,mkreflexyne structures the values of K J K , obtained
due to tensile loading in the x direction were found to be
significantly different to the values obtained from the other
loading conditions.
The mean values of KJK, for each particular combination
of n and 8 are given in Table 2. From the molecular model
calculations 8 was found to be ca. + 30 and cu. - 30" for the
conventional and re-entrant honeycomb geometries, respectively, hence these are the values used in Table 2. With the
exception of the data for the (2,m)-reflexyne structures under
tensile loading in the x direction, the errors associated with
each KJK, value in Table 2 were taken to be the standard
deviation in the mean of the 12 values for each (n, 0) combination since this dominates over the errors inherent in the
molecular modelling package. In the case of the data for the
(2,m)-reflexyne structures under tension in the x direction,
the uncertainty associated with KJK, was calculated by
adding the fractional errors in K , and K, (see later) in quad80

"

-60
-60

-40

-20

I
0

20

40

60

80

vwEV IGPa
F'ig. 3 v,, E , 0s. vyxEx from the molecular model calculations.
Uncertainties are calculated from the deviation of the elastic moduli
in tension and compression from the average value. 0,
(1,m)-flexyne;
A, (2,m)-flexyne, 0,
(2,mbreflexyne (tension in x direction); V,(2,m)reflexyne (compression in x direction) and 0 ,(1,m)-reflexyne.

rature since the smaller sample of K$K, values involved in


this case (three values instead of 12) make a statistical treatment inappropriate.
In all structures K J K , > 1 in Table 2, indicating that
hinging and/or flexure dominate over stretching. Stretching is
most important for the (1,m)-reflexyne structures since they
have the lowest values of K J K , .
The concurrent analytical model Poisson's ratio calculations for the structures modelled in the molecular model
calculations are shown in Table 3. The calculations used the
library of mean K J K , values in Table 2 and the actual unitcell lengths determined for the undeformed structures in the
molecular model calculations (see Table 1). Also shown in
Table 3 are the hinging, flexure and stretching analytical
model Poisson's ratio calculations, employing the molecular
model unit-cell lengths. As noted in Section 2.2.2 the hinging
and flexure model Poisson's ratio expressions are identical
and so these models yield the same numerical values.
It is clear from Tables 1 and 3 that the stretching model
consistently predicts the incorrect sign of Poisson's ratio. The
hinging and flexure models predict the correct sign but consistently overestimate the magnitude of the Poisson's ratio
(by as much as a factor of three in some cases), indicating
that some other mechanism (e.g. stretching) is necessary to
reproduce the molecular model data. The concurrent analytical model us. molecular model Poisson's ratios for loading in
the x and y directions are shown in Fig. 4(a) and (b), respectively. The excellent scatter of the concurrent model calculations about the equality line in Fig. +a) and (b) confirms
that the deformation of these structures is principally due to
a combination of bond hinging, flexure and stretching. It also
indicates that the concurrent model and associated force constants library may be used to predict the Poisson's ratios of
other (n,m)-flexyne and (n,m)-reflexyne structures, and other
related structures built from similar benzene ring and acetylene link subunits (see later).
The values of K , in the concurrent model can be found by
fitting the Young's moduli expressions [eqn. (71) and (7311 to
the molecular model Young's moduli in Table 1, using the
corresponding K J K M values from the fit to the Poisson's
ratio data. K, can then simply be evaluated from the values
of K$K, and K, thus obtained. This was done for each
loading condition for all the structures considered in the
molecular model calculations. Following the same procedure
used for the KJK, data, the mean values of K, and K , for
each (no)combination were evaluated and are also presented
in Table 2.

View Online

J. CHEM. SOC. FARADAY TRANS., 1995, VOL. 91

2677

Table 2 Mean force constants and B / A [see text and Fig. %a)] data library for concurrent analytical model evaluated from parametric fits to
molecular model Poisson's ratio and Young's modulus data for (n,m)-flexyne and (n,m)-reflexyne

1
1
2
2"
2b

+ 30
- 30
+ 30

5.6 C1.21
2.83 C0.57)
20.7 [6.8]
8.9 [1.9]
37.7 C9.41

- 30
- 30

358 [56]
281 [57]
320 [llo]
173 [58]
697 [41]

65.1 C9.11
101 [20]
15.6 C1.91
19.1 C4.81
19.1 C4.8)

1.776 [0.017]
1.776 [0.017]
1.776 [0.017]
1.776 [0.017]
1.776 [0.017]

" Values for loads applied in y direction or compressive loads in x direction. Values for tensile load applied in x direction. Numbers in square
brackets are the estimated standard deviation in the mean value.

Table 3 Analytical model Poisson's ratio calculations for (n,m)-flexyne and (n,m)-reflexyne structures
Downloaded by Universidade do Minho (UMinho) on 02 October 2012
Published on 01 January 1995 on http://pubs.rsc.org | doi:10.1039/FT9959102671

concurrent
structure
(1,2)-flexyne
(1,4)-flexyne
(1,6)-flexyne
(2,2)-flexyne
(2,4)-flexyne
(2,6)-flexyne
( 1,4)-reflexyne
( 1,5)-reflexyne
(1,6)-reflexyne
(2,5)-reflexyne
(2,6)-reflexy ne

(2,8)-reflexyne

VXY

0.43 [ 5 ]
0.31 [4]
0.24 [3]
0.83 [5]
0.61 [4]
0.49 [3]
-0.30 [6]
-0.24 [5]
-0.21 [4]
-0.77 [6]"
-1.04 [31b
-0.64 [5]"
-0.87 [2]'
-0.48 [4]"
-0.65 [21b

stretching

hingindflexure
vYx

VXY

VYX

VXY

VYX

0.66 [8]
0.9 [l]
1.2 [2]
0.83 [5]
1.13 [7]
1.42 [9]
-0.33 [7]
-0.41 [9]
-0.5 [l]
-0.58 [5]

-0.27

-0.19
-0.15
-0.33
-0.25
-0.20
0.32
0.26
0.22
0.39

-0.41
-0.57
-0.73
-0.33
-0.45
-0.57
0.35
0.43
0.5 1
0.29

0.81
0.58
0.45
1.oo
0.74
0.59
-0.95
-0.78
-0.66
- 1.16

1.24
1.72
2.20
1.oo
1.36
1.70
- 1.05
- 1.29
- 1.53
-0.87

-0.69 [6]

0.32

0.35

-0.96

- 1.04

-0.92 [8]

0.24

0.46

-0.72

- 1.39

Concurrent model calculations employ K J K , library of Table 2. Molecular model Y / X values were employed in all analytical model calculations. " Compression in x direction; tension in x direction. Numbers in square brackets are the estimated uncertainty in the least significant
figures.

'

Considering, first, the data for (1, m)-flexyne, (1,m)-reflexyne


and (2,m)-flexyne we see that, within the calculated standard
deviation from the mean values, K, remains constant at
K, x 320 N m-'. However, the value of K,, varies greatly
between these three (no) combinations, leading to the trends
observed in the KJK, data. Hence, in these cases changing
the geometry primarily affects the hinging/flexure force coefficient K, . This will be discussed later.
In the case of the (2,m)-reflexyne data, K , showed similar
variation about a mean value for all loading conditions, as
observed for the other structures. Hence in Table 2 we show
the mean K, value (with associated standard deviation) for
all loading conditions. K , , however, showed a significant
increase when a tensile load was applied in the x direction
compared with the other loading conditions, for the (2,m)reflexyne networks. Hence two values of K, are given in
Table 2 for (no)= (2, -30"); these being the mean of the K,
values for a tensile load in the x direction, and the mean of
the K, values for all other loading conditions.
In Table 4 we present the concurrent analytical model
Young's modulus calculations for the networks considered in
this paper, using eqn. (71) and (73) employing the Y/Xratio
from the molecular model calculations and KJK, and K,
values from Table 2. These are plotted against the molecular
model calculations (Table 1) in Fig. 5(a) and (b)for loading in
the x and y directions, respectively. Once again, excellent
agreement is achieved in both cases. One might expect reasonable agreement since the force constants used in the concurrent model calculations were the mean of those derived
from a fit to the molecular model data for each structure.
However, the fact that good agreement is achieved in all the
Poisson's ratio and Young's modulus data for uniaxial
loading in both principal directions indicates that the concur-

rent model contains the essential features of the deformation


of these structures. (The Young's modulus expressions for the
single-mode analytical models cannot be fitted to the molecular model data using the same value of the appropriate force
constant for loading in both the x and y directions.) Furthermore, the agreement in the data indicates that more complex
networks can be assembled and their properties predicted to
reasonable accuracy using the concurrent model and the
associated force constants library (Table 2), see later.
The force constants trends also explain why the molecular
Table 4 Concurrent analytical model Young's modulus calculations
for (n,m)-flexyne and (n,m)-reflexyne, employing the K$KM and K ,
library of Table 2
structure
(1,2)-flexyne
( 1,4)-flexyne

(1,6)-flexyne
(2,2)-flexyne
(2,4)-flex yne
(2,6)-flex yne
(1,4)-reflexyne
(1,5)-reflexyne
(1,6)-reflexyne
(2,5)-reflexyne
(2,6)-reflexyne
(2,8)-reflexyne

EJGPa

Ey/GPa

77 c11
56 c11
43.6 [0.8]
31 c21
23 c11
18 c11
110 [lo]
86 c91
73 C81
39 171"
46 [8Ib
32 [6]"
38 [7Ib
24 [4]"
29 [5Ib

119 [2]
166 [3]
212 [4]
31 C2I
42 c31
53 c31
120 [lo]
140 [20]
170 [20]
29 c51
35 C61
47 C81

Molecular model Y / X values were used in the calculations for all


structures. " Compression in x direction; tension in x direction.
Numbers in square brackets are the estimated uncertainty in each
value.

'

View Online

2678

J. CHEM. SOC. FARADAY TRANS., 1995, VOL. 91

and E, remain the same as for the simpler structures, i.e. eqn.
(70) and (71), respectively. However, the unit-cell lengths ratio
to be employed in these equations for the stabilised structures
is now

1.2

0.8

Y - (2[(B/A)+ rn]/[(B/A)+ n]
-

0.4

;0.0
-

._

-0.8

Downloaded by Universidade do Minho (UMinho) on 02 October 2012


Published on 01 January 1995 on http://pubs.rsc.org | doi:10.1039/FT9959102671

-1.2

1.6

- 1.6 - 1 . 2

-0.8 -0.4 0.0

0.4 0.8

1.2

vxy( m oI ecu I ar m odeI I ing)


1.6

1.2

0.8

.-

0.4

0.0

s
a
v

+ sin 01

(81)

For the case of loading in the y direction it is easily shown


that

s2 -0.4

cos

Si
-0.4

-0.8

-0.8 -0.4 0 . 0 0 . 4 0.8


vyx(molecular modelling)

1.2

1.6

Fig. 4 (a) Concurrent analytical model vxy us. molecular model vxy
for (n,m)-flexyne and (n,m)-reflexyne structures. (b) Concurrent analytical model vyx us. molecular model vyx for (n,rn)-flexyne and (n,m)reflexyne.

model vxyE , us. vyx E x data are grouped into four sets according to their (n,8) combination (Fig. 3). From eqn. (74) we see
that ~ , , E , ( = v , ~ E is
~ )dependent on K,/K,, K, and 8 only,
and is independent of Y / X . Therefore, the concurrent analytical model predicts the value of v x y E y to be constant for a
given (n,O) combination, irrespective of m, as observed in the
molecular model.
3.3 Model Predictions for More Complex Networks
Attempts to synthesise networks of the type so far discussed
indicate that one or more benzene rings need to be incorporated into the vertical branches in order for the networks to
remain in a stable configuration.2'*22The force constants
trends established in this paper indicate that the mechanical
properties of these more complex networks can be modelled
to a reasonable degree of accuracy by the concurrent analytical model. For example, we consider here the addition of an
extra benzene ring, midway along the vertical arms [referred
to here as stabilised (n,m)-flexyne and stabilised (n,m)reflexyne, where m is now the number of acetylene links
between a junction benzene ring and the non-junction
benzene ring in the vertical branches].
Incorporating an extra benzene ring into the vertical polyacetylene branches introduces an extra component of stretching into the vertical branches. Following the procedure
outlined in Section 2.3, the elastic constants of stabilised networks can be derived. It is found that the expressions for vXy

vyx

E,

sin 8 cos 8[(K$KM)- 13 Y


[ ( K , / K ~ ) c oes ~+ 4 + sin2 el

Ks

Y
-

+ + sin2 81 x

[ ( K ~ / K ~ ) c8o s 4
~

(82)
(83)

where Y / X is again given by eqn. (81). When n = rn the


expression for the unit-cell lengths ratio [eqn. (8 l)] becomes
trivial, depending only on 8. In order to predict the elastic
constants for the more complex networks when n is not equal
to rn, however, it is necessary to know the value of B / A to be
used in eqn. (81). This can be found by returning to the
simpler structures modelled earlier. For each structure modelled in Table 1, B / A was calculated by substituting the undeformed unit-cell lengths, given in Table 1, from the molecular
model calculations into eqn. (78). n, rn and 8 are known for
each structure, enabling B / A to be determined. Unlike the
force constants, no discernible trend in B / A with n or 8 was
observed, the value remaining approximately constant for all
structures. A mean value of B/A = 1.776 _+ 0.017 was found
for the simpler structures modelled in this paper, with the
calculated uncertainty being the estimated standard deviation
in the mean value. This would be the value of B / A to use in
the calculations for the stabilised structures when n is not
equal to m. However, in this paper we will only consider the
case of stabilised structures with n = rn, which leads to a cancellation of the B / A terms in eqn. (81).
Table 5 contains the unit-cell lengths ratio, Poisson's ratio
and Young's modulus calculations from the concurrent analytical model for some stabilised structures. No uncertainty in
Y / X is quoted since there is no dependence on B / A in these
cases. The corresponding molecular model data are also
shown in Table 5. It is important to note that the concurrent
analytical model calculations use the force constants established from the earlier comparison between the concurrent
analytical model and molecular model calculations on the
simpler (n,rn)-flexyne and (n,m)-reflexyne structures (Table 2).
There has been no fitting of the concurrent model data to the
molecular model data for the stabilised structures in Table 5.
The values of Y / X from the molecular model are predicted
to two decimal places by the concurrent analytical model.
This level of agreement is easily acceptable when considering
the errors associated with the force constants to be employed
in the concurrent analytical model (Table 2). Excellent agreement is found in the elastic moduli for each structure, confirming that the concurrent analytical model and its associated force constants library can be used to model more
complex molecular networks accurately. This enables more
efficient and effective use of the computer intensive molecular
modelling package for the modelling of molecular networks
with specific mechanical properties.
In order to test the validity of using a fixed value of B / A to
calculate the unit-cell lengths ratio from the concurrent analytical model when n # rn, the undeformed structure for stabilised (2,l)-reflexyne was determined using the molecular
model program. A value of Y / X = 1.112 was found for this
structure. Employing the mean value of B / A = 1.776
(established from the earlier work on the simple structuressee above) in the concurrent analytical model yields

View Online

2679

J. CHEM. SOC. FARADAY TRANS., 1995, VOL. 91

Downloaded by Universidade do Minho (UMinho) on 02 October 2012


Published on 01 January 1995 on http://pubs.rsc.org | doi:10.1039/FT9959102671

20

40

60

80

100

120

140

Edmolecular modelling)/GPa

m
Q)

160
0

m
m

80

40

l
u
i

40

80

120

160

200

240

280

Ej(molecular modelling)/GPa
Fig. 5 (a) Concurrent analytical model E , us. molecular model E x
data for (n,rn)-flexyne and (n,rn)-reflexyne structures. (b) Concurrent
analytical model E y us, molecular model E y data for (n,rn)-flexyne and
(n,rn)-reflexyne structures.

Y/X = 1.120 f 0.003. Hence the level of agreement in Y / X


for n # rn is of the same order as that for n = rn (see Table 5).

4. Discussion
We have modelled 2D molecular networks based on conventional and re-entrant honeycomb cells, using molecular modelling and analytical modelling techniques. The mechanical
properties of these molecular networks are determined by

their geometry and the behaviour of the sub-unit. The


network can be tailored by modifying the sub-unit, or by
altering the connectivity of the network. Most interestingly,
the force constants required to fit the concurrent analytical
model data to the molecular model data were found to be
directly related to the particular geometry of the network (see
Table 2). This enabled accurate predictions to be made of the
mechanical properties of the more complex stabilised structures using a concurrent analytical model, without recourse
to a full molecular modelling computation. We now consider
why these force constants trends are observed.
So far we have used the analytical model Young's modulus
expressions for unit-thickness in the z direction. Hence one
possible source of deviation in the force constants is that the
value of the unit-cell length perpendicular to the x-y plane,
i.e. Z, may be different for different (n,@ combinations.
However, the molecular model calculations yield essentially
planar structures [slight deviation from planarity being
observed only for (2,5)-reflexyne], having similar values of 2
for all 12 simple structures, with a mean value of
Z = 3.572 & 0.015 A. Hence the observed force constants
trends cannot be attributed to changes in 2.
In the case of the (n,rn)-flexyne structures, increasing n from
1 to 2 reduces the mean value of the hingindflexure force
constant, K,, from 65.1 & 9.1 to 15.6 & 1.9 N m-' (see
Table 2), i.e. by a factor of 4.2 & 0.8. For (n,rn)-reflexyne K ,
decreases from 101 & 20 to 19.1 & 4.8 N m-l as n increases
from 1 to 2, i.e. a reduction in K, by a factor of 5.3 & 1.7,
which is consistent with the value obtained from the (n,rn)flexyne data. From the definition of K,, [eqn. (68)] we see
that K, is inversely proportional to the square of the length
of the diagonal honeycomb arm (i.e. K, proportional to l-').
However, for the molecular structures considered here, this
length should be modified to be the length of the diagonal
acetylene branches since flexure of the benzene rings is
unlikely to occur and hinging occurs about the benzene ring
(rather than hinging of the benzene ring itself). Hence doubling the length of the diagonal acetylene branch by increasing
n from 1 to 2 should yield a decrease in K, by a factor of 4
(= 2*). In fact a factor of greater than 4 would be expected
since K , is also inversely proportional to the length of the
diagonal acetylene branches [see eqn. (9) and (68)].
The effect of varying n on the value of K , appears to be
negligible, within the accuracy of the data of Table 2. K, =
358 f 56 and 320 & 110 N m-' for (1,rn)-flexyne and (2,rn)flexyne, respectively, and hence K , remains constant as n
increases from 1 to 2 within the calculated uncertainties. A
similar comparison cannot be made for the (n,rn)-reflexyne
structures due to the anisotropy observed in K , for (2,rn)reflexyne under tensile and compressive loading in the x
direction.

Table 5 Concurrent analytical model and molecular model undeformed unit-cell lengths ratio, Poisson's ratio and Young's modulus data for
stabilised (n,rn)-flexyne and (n,rn)-reflexyne structures
model

structure

y/x

concurrent
analytical

stabilised (1,l)-flexyne
stabilised (1,l)-reflexyne
stabilised (2,2)-flexyne
stabilised (2,2)-reflexyne

2.8868
1.7321
2.8868
1.7321

0.32 [4]
-0.31 [7]
0.50 [3]
-0.66 [6]"
-0.90 [2]b

molecular

stabilised (1,l)-flexyne
stabilised (1,l)-reflexyne
stabilised (2,2)-flexyne
stabilised (2,2)-reflexyne

2.8896
1.7377
2.8933
1.7335

0.34 [l]
-0.36 [l]
0.49 [4]
- 0.85"
-0.82'

VXY

EJGPa

E,/GPa

0.7 [l]
-0.22 [ S ]
1.3 [l]
-0.54 [6]

58 c11
110 [lo]
19 c11
34 [6]"
40 [7ib

122 [6]
80 [lo]
47 c41
27 [GI

0.74
-0.24
1.26
-0.67

59 c21
120 [20]
19 c31
31"
43'

123 [4]
90 [lo]
49 c71
27 P I

vYx

[l]
[l]

[S]
[S]

" Compression in x direction; ' tension in x direction, Numbers in square brackets are estimated uncertainty in least significant figures.

View Online

Downloaded by Universidade do Minho (UMinho) on 02 October 2012


Published on 01 January 1995 on http://pubs.rsc.org | doi:10.1039/FT9959102671

2680

J. CHEM. SOC. FARADAY TRANS., 1995, VOL. 91

Considering now the effect of changing 8 from positive to


negative, we see that for n = 1 an increase in K, from
65.1 f 9.1 to 101 k 20 N m-l is observed for 0 = +30 and
-3o", respectively, which is an increase by a factor of
1.55 k 0.38. A corresponding decrease in K, (from K , = 358
56 to 281 & 57 N m-') by a factor of 0.78 f 0.20 is also
observed. In the case of the n = 2 structures K, increases
from 15.6 & 1.9 to 19.1 +_ 4.8 N m-' as 8 changes from +30
to -3O", i.e. K , increases by a factor of 1.22 f 0.34 which
agrees within the calculated uncertainties with the value
obtained above for the n = 1 structures. Again the anisotropy
in K, for the (2,m)-reflexyne structures renders a comparison
of the K, values meaningless in this case.
The increase in K, as 8 changes from +30 to -30" indicates that hinging or flexure becomes more difficult to
achieve for the re-entrant structure. It is reasonable to
assume that the increase in K, is due to hinging since flexure
occurs along the acetylene arms whereas hinging occurs at
the benzene junction and it is the connectivity at the benzene
junction that is being investigated here. The reason for the
increase in K, is not clear. Naively, we may expect (n,m)flexyne structures to hinge more easily (and hence have a
lower K, value, as observed) than (n,m)-reflexyne structures
since there is an intermediate benzene ring carbon atom
about which to hinge in the conventional honeycomb
geometry [see Fig. 2(a) and (b)].
The slight decrease observed in K, as 8 changes from +30
to -30" for n = 1 structures may be due to the influence of
non-bonded interactions. It is known, for example, that the
value of the stretching force constant of the C-C bond is
reduced by a factor of ca. 5 for diamond owing to the presence of non-bonded interaction^.^^ It is clear from Fig. 2(a)
and (b) that the bonds connecting the polyacetylene branches
to the junction benzene rings are substantially more crowded
(leading to an increased number of non-bonded interactions)
for the (n,m)-reflexyne structures than for the (n,m)-flexyne
structures. These bonds have been found to make the largest
contribution to the stretching mode of deformation (see
earlier) and hence a reduction in K, would be expected for
(n,m)-reflexyne structures.
In the case of (2,m)-reflexyne, however, the value of K, was
found to be dependent on the loading condition (see Table 2).
This is probably due to the concurrent model not containing
all the deformation mechanisms in this case. It appears that
when n = 2 the benzene rings at the re-entrant junctions
become close enough to each other to interact in such a way
as to introduce other modes of deformation in the molecular
model calculations. This is consistent with the observation of
the slight deviation from planarity of the (2,5)-reflexynestructure.

5. Conclusions
The mechanical properties of 2D molecular networks, having
either positive or negative Poisson's ratios, have been calculated using molecular modelling and analytical modelling
techniques. The molecular networks are based on conventional and re-entrant honeycomb sub-units and consist of
polyacetylene branches connected by benzene rings at the
junctions. An analytical model employing flexure, hinging
and stretching modes of deformation acting concurrently has
been found to produce excellent agreement with the molecular model Poisson's ratio and Young's modulus data when a
parametric fit of the force constants governing these deformation mechanisms is performed. The force constants required
to fit the concurrent analytical model to the molecular model
are found to follow certain trends, being dependent on the
number of acetylene links in the diagonal branches (n)and on

whether the honeycomb angle (0) is positive or negative. The


force constants trends enable a library of force constants to
be compiled for different (n,0)combinations, which can then
be used to predict the properties of other molecular networks
of the type considered here. The concurrent analytical model
can be extended to cover more complex variants of these networks. We developed the concurrent analytical model to
describe networks having an extra benzene ring in the vertical branches of the sub-units. The mechanical properties of
these stabilised structures predicted from the concurrent analytical model and the associated force constants library
(established from the earlier work on the simpler structures)
were found to be in good agreement with the molecular
model calculations. In principle the properties of even more
complex networks (2D and 3D) can be modelled, as a first
approximation, using this concurrent analytical model
approach. This would enable more efficient use of molecular
modelling programs, providing first-order estimates of
properties having first obtained the properties of the appropriate sub-units. Hence, better selection of the network structures to be modelled using full molecular modelling
techniques should be possible.
The authors acknowledge the support of the Engineering and
Physical Sciences Research Council of the United Kingdom
and of Oxford Materials Ltd. K.E.E. currently holds an
EPSRC Advanced Fellowship.

References
1 K. E. Evans, M. A. Nkansah, I. J. Hutchinson and S. C. Rogers,
Nature (London), 1991,353, 124.
2 M. A. Nkansah, K. E. Evans and I. J. Hutchinson, Mod. Sim.
Mat. Sci. Eng., 1994,2, 337.
3 R. H. Baughman and D. S. Galvao, Nature (London), 1993, 365,
735.
4 K. W. Wojciechowski and A. C. Branka, Phys. Rev. A , 1989,40,
7222.
5 J . S. Moore and J. Zhang, Angew. Chem. Int. Ed. Engl., 1992,31,
922.
6 J . S. Moore, Nature (London), 1993,365,690.
7 K. E. Evans, Endeavour, 1991,15,170.
8 R. Lakes, A h . Muter., 1993,5,293.
9 A. W. Lipsett and A. I. Beltzer, J. Acoust. SOC.Am., 1988, 84,
2179.
10 L. J. Gibson and M. F. Ashby, in Cellular Solids: Structure and
Properties, Pergamon Press, Oxford, 1988, ch. 4.
11 K. E. Evans, Compos. Struct., 1990, 17,95.
12 K. L. Alderson and K. E. Evans, Polymer, 1992,33,4435.
13 K. E. Evans and K. L. Alderson, J. Muter. Sci. Lett., 1992, 11,
1721.
14 U. Burkert and N. L. Allinger, in Molecular Mechanics, American Chemical Society, Washington, 1982, ch. 1.
15 N. W. Ashcroft and N. D. Mermin, in Solid State Physics, Holt,
Rinehart and Winston, New York, 1976, p. 443.
16 POLYGRAF, Molecular Simulations Inc., 199 South Los
Robles Avenue, Suite 540, Pasadena, California 91 101, USA.
17 S. L. Mayo, B. D. Olefson and W. A. Goddard 111, J. Phys.
Chem., 1990,94,8897.
18 R. J . Roark and W. C. Young, in Formulas for Stress and Strain,
McGraw-Hill, New York, 5th edn., 1976.
19 B. M. Lempriere, Am. Inst. Aeronaut. Astronaut. J., 1968, 6,
2226.
20 J. E. Shigley, in Applied Mechanics of Materials, McGraw-Hill,
New York, 1976.
21 D. Leigh, UMIST, personal communication.
22 J. S. Moore, University of Illinois, personal communication.
23 M. O'Keeffe and B. G. Hyde, in Structure and Bonding in Crystals, ed. M. O'Keeffe and A. Navrotsky, Academic Press, 1981,
vol. I, p, 247.

Paper 5/013251; Received 3rd March, 1995

You might also like