Sie sind auf Seite 1von 25

Accepted Manuscript

Molecular Modeling of Elastic Properties of Thermosetting Polymers Using a Dynamic


Deformation Approach

Natalia B. Shenogina, Mesfin Tsig, Soumya S. Patnaik, Sharmila M. Mukhopadhyay

PII: S0032-3861(13)00366-2
DOI: 10.1016/j.polymer.2013.04.034
Reference: JPOL 16154

To appear in: Polymer

Received Date: 14 February 2013


Revised Date: 13 April 2013
Accepted Date: 15 April 2013

Please cite this article as: Shenogina NB, Tsig M, Patnaik SS, Mukhopadhyay SM, Molecular Modeling
of Elastic Properties of Thermosetting Polymers Using a Dynamic Deformation Approach, Polymer
(2013), doi: 10.1016/j.polymer.2013.04.034.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

Molecular Modeling of Elastic Properties of

PT
Thermosetting Polymers Using a Dynamic

RI
Deformation Approach

SC
Natalia B. Shenogina1*, Mesfin Tsige2 *, Soumya S. Patnaik3, Sharmila M. Mukhopadhyay1

U
1Department of Mechanical and Materials Engineering, Wright State University, Dayton, OH,
AN
USA; 2Department of Polymer Science, University of Akron, Akron, OH, USA; 3Aerospace

Systems Directorate, Wright-Patterson Air Force Base, Dayton, OH, USA.


M

*AUTHORS EMAIL ADDRESS: Natalia.Shenogina@wright.edu and mtsige@uakron.edu


D
TE

KEYWORDS: molecular dynamics simulations, dynamic deformation approach, epoxy resin,

elastic constants.
EP

This paper employs fully atomistic molecular dynamics simulations to characterize relationships

between structural and elastic properties of thermosetting polymers both in glassy and rubbery
C

state. The polymer system investigated consists of epoxy resin DGEBA and hardener DETDA.
AC

An effective cross-linking procedure that enables generation of thermoset structures containing

up to 35000 atoms with realistic structural characteristics was used. A dynamic deformation

approach has been used that takes into consideration both potential energy and thermal motions

in the structure. Small uniaxial, volumetric and shear deformations were applied to the systems

1
ACCEPTED MANUSCRIPT

to obtain elastic moduli. A method to independently determine Poisson’s ratio was proposed that

reduces statistical errors and circumvents the time scale limitations of molecular dynamics

simulations. The influence of variables such as extent of curing and length of epoxy strands on

elastic response at various temperatures was explored. Expected trends in the dependence of the

PT
elastic constants on these practical process parameters were shown. The relationship between the

RI
four independently calculated elastic constants was seen to comply with those predicted by the

classical theory of linear elasticity in an isotropic medium, which provides confidence in the

SC
validity of these simulations. Moreover, the elastic properties obtained are also in good

agreement with experimental data reported in the literature. Close agreements between predicted

U
elastic constants and experimentally measured values underscore the ability of the approaches
AN
used in this study to provide realistic predictions of the mechanical response of thermosetting

polymers, both in glassy and rubbery states. These results show significant improvement over
M
earlier studies based on a static approach which takes into account the potential energy
D

contribution to the elastic response but ignores temperature effect.


TE

I. Introduction
EP

Thermosetting polymers are important materials in a variety of applications due to their high

thermal and structural stability compared to thermoplastic polymers. Properties of thermosets


C

depend to a large extent on chemical structure and cross-linking density, as well as on the
AC

processing conditions such as temperature and pressure. Computer simulations for property

prediction reduce experimental costs and help with accelerating the development and

applicability of these materials. Recently, molecular dynamics simulations have proven to be

powerful in understanding the behavior of state-of-the-art thermosetting polymers.

2
ACCEPTED MANUSCRIPT

Atomistic molecular dynamics simulations reported in the literature have provided great

insight into the elastic response of highly cross-linked polymer networks. Wu and Xu1 calculated

elastic moduli for diglycidyl ether of bisphenol A (DGEBA) cured with isophorone diamine

(IPD) using the static deformation approach. It was seen that use of the DREIDING force field

PT
resulted in unrealistically high elastic constants whereas the COMPASS force field yielded more

RI
reasonable but still high values compared to experimental measurements. Heine et al2 calculated

the elastic modulus of the united atom model of poly(dimethylsiloxane) (PDMS) networks as a

SC
function of strand length and found that to be in qualitative agreement with experimental data.

Fan and Yuen3 used a PCFF forcefield for EPON862/TETA (triethylenetetramine) structures

U
and calculated Young’s modulus higher than the experimental value. Tack and Ford4 used fully
AN
atomistic MD simulations of EPON862/DETDA structures of oligomeric mixtures using CFF91

and COMPASS force fields, and their calculated value of bulk modulus was also found to be
M
higher than the experimental value of a similar material (DGEBA/IPD). Clancy et al5 simulated
D

curing of DGEBA/DETDA systems up to 86%. While their calculated Young’s and shear moduli
TE

variations with temperature demonstrated monotonic decrease as expected, the dependence of

these moduli on degree of cure was less consistent, particularly for high degrees of cure. Li and
EP

Strachan6,7 studied EPON-862/DETDA systems containing up to 16000 atoms using DREIDING

force field with atomic charges obtained using electronegativity equalization method. An
C

increase in Young’s modulus with conversion degree was seen, but the trend of Poisson’s ratio
AC

was not clear, since the predicted values were extremely scattered, ranging from 0.2 to 0.5.

Bandyopadhyay et al8,9 modeled EPON862/DETDA systems containing up to 25000 atoms and

cured up to 76%. They applied different modes of deformation to the structures to directly obtain

two elastic constants and then used linear elastic formulae assuming isotropic materials to

3
ACCEPTED MANUSCRIPT

calculate the other two. While the Young’s and shear moduli were seen to increase, bulk

modulus was found to slightly decrease with conversion degree.

Although all of the above studies have made significant progress in predicting the mechanical

response of thermosetting polymer networks, many questions still remain unanswered. For

PT
example, the influence of degree of conversion and length of strands between cross-links on

RI
mechanical properties is still not fully understood, mainly due to difficulties in creating realistic

systems with conversion degree consistent with that typical in experiments. Moreover, small

SC
system sizes used in simulations lead to substantial amounts of scattering in mechanical

properties. Besides, the choice of forcefield is found to be critical in obtaining reasonable values

U
of elastic constants. It must also be noted that a complete range of mechanical studies are
AN
generally not performed using atomistic simulations. Typically, one or two modes of

deformation are simulated with the assumption that elastic constants obtained from those can
M
fully characterize the elastic response of the particular amorphous polymeric material. However,
D

the long simulation times needed to reproduce macroscopic mechanical behavior of thermosets
TE

has led to studies using larger deformations (beyond the elastic limit) and with limited statistical

sampling, causing inaccurate prediction of individual elastic constants.


EP

To answer some of the above mentioned questions and to also provide a systematic study of

available modeling approaches, we have recently investigated the dependence of thermo-


C

mechanical properties of thermosetting polymer DGEBA/DETDA on the extent of curing


AC

reaction, length of resin strands and size of simulation cell at two different temperatures10. We

used the effective cross-linking procedure developed by Accelrys11 that allows construction of

highly cross-linked polymer networks having structural characteristics close to those in real

systems. The resulting systems are characterized by high conversion degrees and are free of

4
ACCEPTED MANUSCRIPT

internal stresses and geometrical distortions. In our recent paper10 we found that while properties

such as density, coefficient of thermal expansion and glass transition temperature were found to

be in good agreement with experimental data available in the literature12-17, the values of elastic

constants, calculated using static deformation approach, showed notable deviation from values

PT
reported in experiments16,18-23.

RI
In the present work we are employing a dynamic deformation approach, which takes into

account both the potential energy contribution and the influence of thermal motions in the

SC
structure on its mechanical behavior. We explore the influence of temperature, extent of curing

and length of epoxy strands on elastic properties of thermosetting materials. To verify that the

U
acquired elastic properties meet assumptions of linear elasticity, we performed uniaxial,
AN
volumetric and shear deformation of the simulation cells and determined all four elastic

constants independently using small deformations. We also proposed a novel methodology of


M
Poisson’s ratio determination that reduces statistical errors and avoid time scale limitations of
D

molecular dynamics simulations.


TE

II. Methodology
EP

A. Systems of interest and simulation details

We focused on a widely used resin-hardener system composed of DGEBA (diglycidyl ether of


C

bisphenol A) epoxy oligomers and aromatic amine hardener DETDA (diethylene toluene
AC

diamine). The molecular structures of the initial components used for the cross-linking reaction

are shown in Figure 1 and, we examined stochiometrically balanced compositions of reactants

permitting a theoretical conversion of 100%.

5
ACCEPTED MANUSCRIPT

PT
RI
Figure 1. Epoxy resin: DGEBA (diglycidyl ether of bisphenol A) with activated reactive sites

SC
(yellow); (b) Aromatic amine hardener: DETDA (diethylene toluene diamine). Reactive sites

(amine groups) are highlighted in yellow.

U
AN
Initially, the reactants were randomly distributed in a simulation box using the Amorphous

Cell module of the Materials Studio commercial package11, and all subsequent molecular
M
dynamics simulations were done with the Discover module of this software. Atomic interactions

are based on the Class II force field COMPASS24, which has been shown to provide accurate
D

predictions of thermo-mechanical properties of thermosetting polymers.1,4,10


TE

The cross-linking method developed by Accelrys11 was used to build highly cross-linked
EP

polymer networks. With this cross-linking method, it is possible to achieve high extents of

reaction typical of real systems, with no internal stresses and no geometrical distortions in the
C

structures. More details about this method are given in Ref. 10.
AC

To study the influence of the extent of curing reaction on the mechanical properties of

thermoset networks (DGEBA/DETDA epoxy resin), six conversion degrees ranging from 50% to

95% were selected for each system. These structures were then equilibrated at room temperature

and at elevated temperature, as discussed in detail in our earlier paper10, to determine the

mechanical properties of the obtained networks both in glassy and rubbery states. Elevated

6
ACCEPTED MANUSCRIPT

temperature was chosen to 480K which is above the range of glass transition temperatures (396-

430K) found for these structures in our previous study10. To examine the effect of epoxy chain

length on the mechanical properties, we constructed several systems using short epoxy oligomers

of one, two or four monomer resin molecules, referred hereafter as mono-, di-, and tetramers,

PT
respectively. The system based on epoxy monomers consisted of 512 epoxy monomers and 256

RI
cross-linkers and will be denoted hereafter by (512,256). Similarly, dimer and tetramer based

structures are denoted by (256,128) and (128,64), respectively. Note that similar system sizes

SC
were used for all cases.

For better prediction of the elastic response of polymer networks and to reduce statistical

U
scattering due to nanoscopically small simulation cells, data obtained from five topologically
AN
independent structures were averaged for each extent of reaction and epoxy chain length.

B. Choice of deformation approach


M
In an earlier paper10, we used the so-called static deformation approach25, where uniaxial and
D

shear deformations of a small magnitude are instantly applied to a simulation cell in different
TE

directions and energy minimization subsequently performed. This approach was introduced by

Theodoru and Suter to study small deformations of polymers at relatively low temperatures. Due
EP

to its computational efficiency, this methodology allows the analysis of a large number of

nanoscopically small volume elements to compensate for large scattering in mechanical


C

properties data and to partially circumvent size limitations of the molecular dynamics
AC

simulations method. However, this approach takes into account only potential energy

contribution to the mechanical response of the material, neglecting the contribution of thermal

motions. It is known that local motions in polymers occur even at temperatures a few degrees

above absolute zero and become significant in the rubbery state. For this reason, deformation in

7
ACCEPTED MANUSCRIPT

polymers should be considered an intrinsically dynamic event that involves thermal activation of

molecular rearrangements and, therefore, the effect of temperature on the mechanical properties

cannot be ignored. This assertion was confirmed in the course of our previous study, where we

found that static approach did not accurately predict the mechanical response, especially at

PT
elevated temperatures.

RI
In the present study, we modeled stress-strain behavior using a dynamic approach, taking into

account potential energy as well as entropic and vibrational contributions to the elastic response

SC
of thermosetting polymers. The deformation of a given structure was applied in stepwise fashion

at a rate of 108 [1/s], which is typical for MD simulations. At each step the structure was

U
deformed by 0.1% followed by energy minimization and equilibration at constant temperature

and volume for 10 ps.


AN
Elastic moduli were then calculated as initial slopes of stress-strain curves obtained using
M
appropriate components of stress and strain tensors. More specifically, Young’s modulus was
D

obtained by applying tension and compression uniaxial strains individually at each coordinate
TE

direction of the simulation cell and calculated as , where and are diagonal elements

of the stress and strain tensors, respectively.


EP

Bulk modulus describes the material response to uniform pressure. In the present study it was

obtained by simultaneously applying equal compression or dilatation strains in all three


C

directions and determined as the initial slope of the curve representing the average of stress
AC

tensor diagonal components vs. volumetric deformation as

. (1)

8
ACCEPTED MANUSCRIPT

Similarly, shear modulus was obtained by applying shear deformation in each direction and

calculated as , where and are off-diagonal elements of the stress and strain

tensors. Poisson’s ratio was obtained in uniaxial deformation mode. The details of its

PT
determination are discussed in the next subsection.

Keeping in mind that nanoscopically small structures investigated using MD simulations are

RI
not perfectly isotropic and homogeneous, we performed uniaxial and shear deformations of

simulation cells in the three different directions and subsequently averaged the acquired data.

SC
It is also important to keep in mind that one of the characteristic features of amorphous

U
polymers is shallow energy landscape26. As a consequence of the ability to rearrange at the

AN
molecular level, even at low deformations, linear mechanical response of these materials can be

observed only for very small deformations, typically below 1%. At such low deformation, the
M
statistical scattering of the stresses computed from MD simulations can be significant.

Nevertheless, the slope of the stress-strain curve at infinitesimal deformation could be estimated
D

by taking into account that it smoothly changes at small deformations from compression to
TE

tension deformation mode. Hence, to a first approximation, the slope in the vicinity of zero

deformation can be calculated as an average over the slopes of tension and compression stress-
EP

strain curves measured within a few percent of deformation. In this work, we deformed
C

simulation cells up to 1.5% in uniaxial, volumetric and shear deformation modes, and elastic
AC

moduli were determined as an average over the slopes in tension and compression modes.

C. Poisson’s ratio calculation

Poisson’s ratio can be measured as the ratio of lateral to longitudinal strain in uniaxial tests.

However, due to the viscoelastic behavior of polymers in the course of uniaxial deformation,

lateral stresses approach equilibrium zero values over a finite period of time and lateral

9
ACCEPTED MANUSCRIPT

contractions are strain rate dependent. Keeping in mind that experimental and MD simulations

strain rates differ by 10-12 orders of magnitude one can expect simulated Poisson’s ratio values

to be lower than experimental values.

To address this issue and to obtain realistic values of Poisson’s ratio, we employed the

PT
following procedure to calculate this elastic constant. We applied stepwise uniaxial tension and

RI
simultaneous compression deformation in the transverse directions, corresponding to certain

values of Poisson’s ratio. Several Poisson’s ratios ranging from 0.0 to 0.5 were probed for each

SC
structure both at room and at elevated temperature and lateral stresses were monitored in the

course of deformation. At constant lateral conditions, when transverse shrinkage does not occur

U
during uniaxial tension (ν=0.0), positive lateral stresses are developed during tensile deformation
AN
yielding positive slope of the lateral stress-axial strain curve . Such a behavior in simulated
M
systems is predictable as typical experimental values of thermoset Poisson’s ratio fall in the

range of 0.33-0.40 at room temperature and rise to about 0.5 in the rubbery state19-22. In another
D

limiting case of incompressible material (ν=0.5), lateral stresses approach equilibrium zero
TE

values at elevated temperature (rubbery state), giving negligible slopes. At room


EP

temperature, these slopes are negative in the case of tension, denoting too-large lateral

contraction for glassy state. Probing up to six values of Poisson’s ratio reveals linear dependence
C
AC

of the slopes of the lateral stress-axial strain curves on the probed Poisson’s ratios.

10
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE

Figure 2. The ratio of lateral stress to axial strain , developed in the course of uniaxial
EP

deformation of monomer-based structure at room temperature and cured to 95%, at various

probed Poisson’s ratios. and are diagonal elements of the stress and strain tensors,
C

respectively. Indexes denote transverse (i) and longitudinal (k) directions. The red line represents
AC

a linear fit of the data.

Figure 2 represents such a dependence, which characterizes lateral stresses developed in the

structure during uniaxial deformation at various probed Poisson’s ratios. Such dependences were

plotted for each extent of reaction, both at room and at elevated temperature. Each point on the

11
ACCEPTED MANUSCRIPT

plot represents an average over five structures and all three directions in tension and compression

simulations to reduce statistical errors and take into account anisotropic effects. To estimate the

Poisson’s ratio at which lateral stresses are not developed during deformation of a given

structure, a linear fit to the data was used (Figure 2) to interpolate the Poisson’s ratio value at the

PT
intersection of the fit with the horizontal axis. This Poisson’s ratio value is identified as true

RI
Poisson’s ratio of the given structure. This method was applied for each extent of reaction to

determine the Poisson’s ratios in the glassy state (at 298K) and rubbery state (at 480K). For a

SC
given structure at a given temperature Poisson’s ratio found using this approach was then used in

the direct measurement of Young's modulus, i.e., by keeping the Poisson ratio of the structure to

U
this value during uniaxial deformation.
AN
D. Comparing Simulation Results with the Predictions of Linear Elasticity Theory

The theory of linear elasticity27 has been successfully used to describe the mechanical response
M
of materials to infinitesimal deformations. The most common experimental means of polymer
D

characterization are uniaxial tension and shear deformations, while bulk (volumetric) response
TE

and Poisson’s ratio experimental data are limited due to difficulties associated with making

precise measurements of very small deformations. Nevertheless, in experiments, having any two
EP

elastic constants available from direct measurements of an amorphous polymer at small

deformation is considered sufficient to fully characterize the mechanical properties of the


C

system, since the remaining constants are often obtained by using the theory of linear elasticity
AC

with the assumption that structures are homogeneous and isotropic.

However, nanoscopically small simulation cells of thermosetting polymers used in molecular

dynamics simulations are less homogeneous and less isotropic than macroscopic samples used in

experiments, as each simulation cell is characterized by a unique distribution of matter and

12
ACCEPTED MANUSCRIPT

therefore generates unique mechanical properties, causing significant scattering in the data.

Furthermore, different deformation modes such as uniaxial, volumetric and shear strains may

result in significantly different molecular rearrangements in the structure. Therefore, starting

with some deformation level within the range typically employed in MD simulations of

PT
amorphous polymers, elastic constant values obtained using two different deformation modes

RI
could be different beyond statistical errors28. It is thus very important to verify the compatibility

of the results obtained using the different deformation modes.

SC
There is a set of experimental papers28-32 in which the authors examined the experimental

limitations of deriving elastic constants from properties measured directly. Similar to these

U
studies, we conducted independent simulations of uniaxial, volumetric and shear deformations to
AN
acquire all four elastic constants. We then used the theory of linear elasticity to calculate elastic

moduli from any two constants obtained by means of direct simulations. Finally, the calculated
M
moduli were compared with the moduli obtained from direct simulations. The similarities and
D

differences between moduli will be discussed at the end of the next section.
TE

To the best of our knowledge, this is a first reported comprehensive study of the mechanical

response of thermosetting polymers which compares elastic constants acquired by direct


EP

simulations with the corresponding material functions computed using the theory of linear

elasticity.
C

III. Results and discussions


AC

A. The role of deformation approach

Figure 3 shows the elastic constants of monomer-based structures obtained from direct

deformation simulations using both static (reported in Ref. 10) and dynamic approaches.

13
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D

Figure 3. Elastic moduli at 298K (blue closed symbols) and 480K (black open symbols) as a
TE

function of the extent of the reaction obtained from static (squares) and dynamic (circles)
EP

simulations: (a) Young’s modulus; (b) bulk modulus; (c) shear modulus; (d) Poisson’s ratio. Red

lines represent experimental values at room temperature: (a) Young’s modulus for
C

GDEBA/DETDA structure18; (c) Shear modulus for similar epoxy structure23; (d) Poisson’s ratio
AC

for similar epoxy structures (lower border of shaded area –Ref.20; upper border of shaded area –

Ref. 21, 22).

Young’s, bulk and shear moduli show monotonic increase with the degree of cure both in static

and dynamic simulations, reflecting the expected increase in material stiffness. The slopes of

dynamic curves, however, were found to be notably lower than that of static curves, as can be

14
ACCEPTED MANUSCRIPT

seen in the figure. The values of all three elastic moduli at both temperatures, obtained using

dynamic simulations, show excellent improvement compared to those found using static

simulations, a clear justification that one must fully account for dynamic effects. The Young’s

modulus of the dynamic simulations is in very good agreement with experimental data at room

PT
temperature18. Though no experimental data exist for the other moduli of the DGEBA/DETDA

RI
thermoset, the values obtained using dynamic simulations are within the range for common

thermosetting polymers19-23.

SC
The simulation results for elastic moduli can be understood if we recall that epoxy polymers

demonstrate nonlinear behavior even at very low deformations. In Figure 4, a representative

U
stress-strain curve obtained from tensile simulations is shown. In this seemingly stepwise curve,
AN
regions of increasing of internal stresses alternate with regions of stress relaxation. The

relaxation in the stress is due to thermally activated molecular relaxations and results in
M
decreasing Young’s modulus values. In contrast to this, the deformation level in static
D

simulations was less than 0.1%, where molecular relaxations are less intensive, causing high
TE

stress derivative and unrealistically large elastic moduli values. Similar behavior can be observed

at the initial 0.3% deformation portion of the stress-strain curve. Moreover, extremely high
EP

deformation rates used in molecular dynamics simulations result in less intensive molecular

relaxations in polymeric systems and may increase the elastic moduli values.
C
AC

15
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE

Figure 4. Stress-strain curve obtained at uniaxial tension dynamic deformation of monomeric

structure at 108 [1/s] deformation rate.


EP

Poisson’s ratios (Figure 3d) obtained using dynamic simulations show different behavior from
C

previously obtained static simulation results (also shown in the figure) both qualitatively and
AC

quantitatively. While the results of the static simulations show no dependence of Poisson’s ratio

on degree of cure, the dynamic results show a monotonic decrease in Poisson’s ratio with degree

of cure. The static approach gives unrealistically low values of Poisson’s ratio and will not be

discussed further. The values of Poisson’s ratio at high temperature are higher than that at room

temperature, displaying the behavior typical in rubbery and glassy states. The Poisson’s ratio

16
ACCEPTED MANUSCRIPT

value at high degree of cure at room temperature is found to be within the range of experimental

values for similar materials. 20-22

B. Relationships between elastic constants

Figures 5 shows a comparison of the elastic constants obtained from direct simulations and by

PT
applying linear elasticity theory both at room and elevated temperatures. Almost all properties

RI
are found in excellent mutual agreement. However, both Poisson’s ratio and bulk modulus

calculated using direct Young’s and shear simulation results, represented as and ,

SC
respectively, are characterized by significant scattering of the data (Figure 5(b, d)). A close

examination of the following equations

U
(2) AN
(3)
M
used to calculate bulk modulus and Poisson’s ratio, respectively, reveals that these calculated
D

values are sensitive to the ratio of Young's modulus to shear modulus. In equation (2), when the
TE

Young’s modulus is about three times higher than shear modulus, the uncertainty in the

calculated bulk modulus value grows dramatically and may grow to several orders of magnitude.
EP

In addition, since G is small, especially at elevated temperature, the ratio of E and G values in

the calculation of Poisson’s ratio using equation (3) causes significant scattering of the data.
C

Note that a good agreement between all four elastic constants determined by different
AC

deformation modes is expected only under certain conditions, since molecular mechanisms

contributing to different deformation modes are not the same. For instance, uniform compression

involves local motions of molecular segments, while shear deformation involves both local and

extensive molecular rearrangements. Such diversity in mechanical responses can cause

significant variation in the values of elastic constants obtained by different deformation modes if

17
ACCEPTED MANUSCRIPT

large strains are imposed. Mutual agreement between all four elastic constants in our simulations

confirms that slopes of the stress-strain curves at infinitesimal deformations are calculated

correctly and the acquired results were within elastic limits.

PT
RI
U SC
AN
M
D
TE
EP

Figure 5. Elastic moduli at 298K (closed symbols) and 480K (open symbols) as a function of the

extent of the reaction calculated using direct dynamic simulations (squares) and using linear
C

elasticity theory. (a) Young’s modulus: E(B,G) – red circles; E(B,ν) – green triangles; E(G,ν) –
AC

blue stars. (b) Bulk modulus: B(E,G) – red circles; B(E,ν) – green triangles; B(G,ν) – blue stars.

(c) Shear modulus: G(B,E) – red circles; G(B,ν) – green triangles; G(E,ν) – blue stars. (d)

Poisson’s ratio: ν (B,E) – red circles; ν (B,G) – green triangles; ν (E,G) – blue stars.

C. The role of epoxy chain length

18
ACCEPTED MANUSCRIPT

Figure 6 represents elastic constants of monomer-, dimer- and tetramer-based structures

obtained from direct deformation simulations using the dynamic approach. The dynamically

obtained Young’s, bulk and shear moduli of all oligomer-based structures show monotonic

increase with the degree of cure, while Poisson’s ratio values demonstrate notable decrease with

PT
extent of reaction, particularly at high temperature. It can be seen that elastic properties of

RI
oligomer-based structures reveal the same trend: structures with shorter distance between cross-

links tend to show more pronounced dependence on degree of cure as can be observed from the

SC
slope of the fitted lines.

U
AN
M
D
TE
C EP
AC

Figure 6. Elastic moduli at 298K (closed symbols) and 480K (open symbols) as a function of the

extent of the reaction for the atomic structures built using monomers (black squares), dimers (red

19
ACCEPTED MANUSCRIPT

circles) and tetramers(green triangles) of the epoxy resin: (a) Young’s modulus; (b) bulk

modulus; (c) shear modulus; (d) Poisson’s ratio. Solid lines represent linear fits to the data.

Conclusions

PT
This study uses a proven cross-linking procedure on large polymer systems containing up to

35000 atoms, to generate stress-free thermoset networks with high degree of cure. A dynamic

RI
deformation approach has been used to simulate the elastic response of the generated structures

SC
and predict their mechanical properties. The dependence of elastic constants on process

parameters such as temperature, degree of conversion and length of resin strands has been

U
investigated in these simulations. Young’s, shear and bulk moduli, as well as Poisson’s ratio,

AN
were obtained directly at small deformations from atomistic simulations using uniaxial,

volumetric and shear deformation modes. A novel algorithm to calculate the Poisson’s ratio was
M
proposed and found to successfully reduce statistical errors and circumvent the time-scale

limitations of molecular dynamics simulations. Finally, all simulation results of individual


D

parameters were shown to compare favorably with values calculated using linear elasticity
TE

theory.

The dynamic deformation approach has provided realistic values for mechanical properties,
EP

both in glassy and rubbery states. Values of Young’s, shear and bulk moduli and Poisson’s ratio
C

at high extents of curing reaction were found to be in very good agreement with experimental
AC

data of actual cured polymers and show excellent improvement compared to elastic constants

calculated using the static deformation approach. This finding supports that thermal motions

have significant influence on the mechanical response of highly cross-linked polymers, both in

glassy and rubbery states. Further insight into the role of extent of reaction and length of resin

strands on elastic properties of thermosets was attained showing realistic mechanical response.

20
ACCEPTED MANUSCRIPT

To the best of our knowledge, this is the first reported paper that predicts all four elastic

coefficients of a thermosetting polymer by direct simulation and then successfully compares

them with the corresponding values computed using the theory of linear elasticity.

The approaches used in this study exhibit significant promise in their ability to predict the

PT
mechanical behavior of highly cross-linked polymeric materials.

RI
ACKNOWLEDGMENT

SC
This work was primarily supported by the Low Density Materials Program of the Air Force

U
Office of Scientific Research Grant Number: FA9550-09-1-0358. The authors gratefully

AN
acknowledge Dr. Charles Lee (AFOSR) for valuable discussions, Wright State University for

partial salary support, and the Air Force Research Laboratory DoD Supercomputing Resource
M
Center High Performance Computing for computer time.
D

REFERENCES
TE

(1) Wu, C.; Xu, W. Polymer 2006, 47, 6004.


EP

(2) Heine, D. R.; Grest, G. S.; Lorenz, C. D.; Tsige, M.; Stevens, M. J. Macromolecules 2004,
C

37, 3857.
AC

(3) Fan, H. B.; Yuen, M. M. F. Polymer 2007, 48, 2174.

(4) Tack, J. L.; Ford, D. M. J. Mol. Graphics Modell. 2008, 26, 1269.

(5) Clancy, T. C.; Frankland, S. J. V.; Hinkley, J. A.; Gates, T. S. Polymer 2009, 50, 2736.

(6) Li, C.; Strachan, A. Polymer 2010, 51, 6058.

21
ACCEPTED MANUSCRIPT

(7) Li, C.; Strachan, A. Polymer 2011, 52, 2920.

(8) Bandyopadhyay, A.; Valavala, P. K.; Clancy, T. C.; Wise, K. E.; Odegard, G. M. Polymer

PT
2011, 52, 2445.

(9) Bandyopadhyay, A.; Odegard, G. M. Modelling Simul. Mater. Sci. Eng. 2012, 20, 045018.

RI
(10) Shenogina, N.B.; Tsige, M.; Patnaik, S.S.; Mukhopadhyay, S.M. Macromolecules 2012,

SC
45, 5307.

U
(11) Accelrys Software inc.: Materials Studio. http://accelrys.com/products/materials-studio/

(accessed Dec 19, 2012). AN


(12) Gao, J. G.; Li, Y. F.; Zhao, M.; Liu, G. D. J. Appl. Polym. Sci. 2000, 78, 794.
M
(13) Jansen, B. J. P.; Tamminga, K. Y.; Meijer, H. E. H.; Lemstra, P. J. Polymer 1999, 40,
D

5601.
TE

(14) Ratna, D.; Manoj, N. R.; Varley, R.; Raman, R. K. S.; Simon, G. P. Polym. Int. 2003, 52,

1403.
EP

(15) Ratna, D.; Varley, R.; Singh, R. K.; Simon, G. P. J. Mater. Sci. 2003, 38, 147.
C

(16) Shen, L.; Wang, L.; Liu, T. X.; He, C. Macromol. Mater. Eng. 2006, 291, 1358.
AC

(17) Liu, W.; Varley, R. J.; Simon, G. P. Polymer 2006, 47, 2091.

(18) Qi, B.; Zhang, Q. X.; Bannister, M.; Mai, Y. W. Compos. Struct. 2006, 75, 514.

(19) Kalantar, J.; Drzal, L. T. J. Mater. Sci. 1990, 25, 4186.

22
ACCEPTED MANUSCRIPT

(20) Kalantar, J.; Drzal, L. T. J. Mater. Sci. 1990, 25, 4194.

(21) O'Brien, D. J.; Sottos, N. R.; White, S. R. Exp. Mech. 2007, 47, 237.

PT
(22) Tcharkhtchi, A.; Faivre, S.; Roy, L. E.; Trotignon, J. P.; Verdu, J. J. Mater. Sci. 1996, 31,

2687.

RI
(23) Drzal, L.T. Mater. Sci. Eng., A 1990, 126, 289.

SC
(24) Sun, H. J. Phys. Chem. B 1998, 102, 7338.

U
(25) Theodorou, D. N.; Suter, U. W. Macromolecules 1986, 19, 139.
AN
(26) Wales, D. J. Energy Landscapes; Cambridge University Press: Cambridge, 2003.
M
(27) Landau, L.D.; Lifshitz, E. M. Theory of Elasticity; Butterworth Heinemann: Oxford, 1986.
D

(28) Sane, S.B.; Knauss, W. G. Mech. Time-Depend. Mater. 2001, 5, 325.


TE

(29) Sane, S.B.; Knauss, W. G. Mech. Time-Depend. Mater. 2001, 5, 293.


EP

(30) Yee, A.F.; Takemori, M.T. J. Polym. Sci. 1982, 20, 205.

(31) Lu, H.; Zhang, X.; Knauss, W. G. Polym. Eng. Sci. 1997, 37, 1053.
C

(32) Deng, T.H.; Knauss, W. G. Mech. Time-Depend. Mater. 1997, 1, 33.


AC

23

Das könnte Ihnen auch gefallen