+ All Categories
Home > Documents > Multifidelity multiscale modeling of nanocomposites for ...

Multifidelity multiscale modeling of nanocomposites for ...

Date post: 20-Dec-2021
Category:
Upload: others
View: 4 times
Download: 0 times
Share this document with a friend
13
Contents lists available at ScienceDirect Composite Structures journal homepage: www.elsevier.com/locate/compstruct Multidelity multiscale modeling of nanocomposites for microstructure and macroscale analysis Ashwin Rai a , Aditi Chattopadhyay b, a Graduate Research Associate, School for Engineering of Matter, Transport, and Energy, Arizona State University, Tempe, AZ 85287, United States b RegentsProfessor & Ira. A Fulton Professor of Engineering, School for Engineering of Matter, Transport, and Energy, Arizona State University, Tempe, AZ 85287, United States ARTICLE INFO Keywords: Carbon nanotubes Multiscale modeling Surrogate modeling ABSTRACT A high-delity multiscale modeling framework that integrates information from atomistic simulations pertaining to polymer chain sliding and bond dissociation is utilized to study damage evolution and failure in carbon nanotube (CNT)-reinforced nanocomposites. The nanocomposite constituents (microber, polymer, and CNTs) are explicitly modeled at the microscale using representative unit cells (RUCs). The modeled constituents are subsequently employed in a multiscale framework to describe damage initiation and propagation in these sys- tems under transverse loading. Two CNT architectures, randomly dispersed and radially grown, are investigated. Damage initiation sites and damage evolution trends are studied, with results indicating that the presence of CNTs causes a unique stress state at the sub-microscale. This can lead to accelerated damage progression, which can be mitigated by architectural reconguration of the CNTs. Additionally, the Schapery potential theory is extended to develop an orthotropic nonlinear damage model that captures global behavior of the nanocomposite RUCs in a computationally ecient manner, and can be utilized as a numerical surrogate for structural scale nanocomposite analysis. 1. Introduction Structural nanocomposites in the recent literature have shown to oer many potential and promising applications for aerospace and civil technologies [1,2]. For instance, nanocomposites that use carbon na- notubes (CNTs) for nanoscale reinforcement in a polymer matrix, have exhibited signicant improvements in mechanical properties over conventional polymer based material systems [3,4]. In particular, carbon ber composites reinforced with a polymer matrix containing randomly dispersed CNTs have shown to provide marked improve- ments in mechanical strength, interlaminar fracture resistance, energy absorption, and thermomechanical properties [57] compared to tra- ditional carbon ber reinforced polymer (CFRP) composites. However, it has also been observed that the nanocomposite characteristics do not scale linearly with size, and there exist large disparities in microscale and structural scale properties, such as stiness and failure strengths [8]. Similarly, macroscale systems with CNTs embedded in their matrix, display only a marginal increase in fracture properties [9,10] with some studies indicating also reduced strength and fatigue life when compared to predicted values [11]. Such discrepancies are largely attributed to microscale architectural irregularities in the randomly dispersed nanollers, such as misalignment, agglomerations, and poor dispersion of the CNTs [12,13]. To mitigate these drawbacks, recent advances in nanotechnology techniques have been exploited to engineer novel CNT architectures, including nanoforests [1416], which utilize appropriate substrates for the growth of highly aligned dense mats of CNTs, fuzzy bers [17], which contain CNTs radially grown on the microbers, and CNT ropes [18], which harness the ultra-long strands of CNTs as a re- placement for microbers. Fuzzy ber architecture, in particular, has shown to provide increased in-plane strength, interlaminar shear strength, and fracture properties [19]]. However, the observed im- provements in properties due to the addition of CNTs is generally lower than the predicted theoretical values [20]. Possible causes for discrepancies between experimental results and theoretical predictions of the CNT nanocomposite properties may be the modeling approaches used for predicting material properties and be- havior. These approaches are often not appropriate for a complex het- erogeneous material system, such as the CNT/CFRP. Conventional methodologies predict structural scale properties using one of two ap- proaches: a top-down approach based on a bulk material analysis such as smeared material techniques [21,22], or a bottom-up approach that uses mean eld techniques, such as the Mori-Tanaka or the composite https://doi.org/10.1016/j.compstruct.2018.05.075 Received 16 December 2017; Received in revised form 28 April 2018; Accepted 17 May 2018 Corresponding author. E-mail address: [email protected] (A. Chattopadhyay). Composite Structures 200 (2018) 204–216 Available online 22 May 2018 0263-8223/ © 2018 Elsevier Ltd. All rights reserved. T
Transcript
Page 1: Multifidelity multiscale modeling of nanocomposites for ...

Contents lists available at ScienceDirect

Composite Structures

journal homepage: www.elsevier.com/locate/compstruct

Multifidelity multiscale modeling of nanocomposites for microstructure andmacroscale analysis

Ashwin Raia, Aditi Chattopadhyayb,⁎

aGraduate Research Associate, School for Engineering of Matter, Transport, and Energy, Arizona State University, Tempe, AZ 85287, United Statesb Regents’ Professor & Ira. A Fulton Professor of Engineering, School for Engineering of Matter, Transport, and Energy, Arizona State University, Tempe, AZ 85287, UnitedStates

A R T I C L E I N F O

Keywords:Carbon nanotubesMultiscale modelingSurrogate modeling

A B S T R A C T

A high-fidelity multiscale modeling framework that integrates information from atomistic simulations pertainingto polymer chain sliding and bond dissociation is utilized to study damage evolution and failure in carbonnanotube (CNT)-reinforced nanocomposites. The nanocomposite constituents (microfiber, polymer, and CNTs)are explicitly modeled at the microscale using representative unit cells (RUCs). The modeled constituents aresubsequently employed in a multiscale framework to describe damage initiation and propagation in these sys-tems under transverse loading. Two CNT architectures, randomly dispersed and radially grown, are investigated.Damage initiation sites and damage evolution trends are studied, with results indicating that the presence ofCNTs causes a unique stress state at the sub-microscale. This can lead to accelerated damage progression, whichcan be mitigated by architectural reconfiguration of the CNTs. Additionally, the Schapery potential theory isextended to develop an orthotropic nonlinear damage model that captures global behavior of the nanocompositeRUCs in a computationally efficient manner, and can be utilized as a numerical surrogate for structural scalenanocomposite analysis.

1. Introduction

Structural nanocomposites in the recent literature have shown tooffer many potential and promising applications for aerospace and civiltechnologies [1,2]. For instance, nanocomposites that use carbon na-notubes (CNTs) for nanoscale reinforcement in a polymer matrix, haveexhibited significant improvements in mechanical properties overconventional polymer based material systems [3,4]. In particular,carbon fiber composites reinforced with a polymer matrix containingrandomly dispersed CNTs have shown to provide marked improve-ments in mechanical strength, interlaminar fracture resistance, energyabsorption, and thermomechanical properties [5–7] compared to tra-ditional carbon fiber reinforced polymer (CFRP) composites. However,it has also been observed that the nanocomposite characteristics do notscale linearly with size, and there exist large disparities in microscaleand structural scale properties, such as stiffness and failure strengths[8]. Similarly, macroscale systems with CNTs embedded in their matrix,display only a marginal increase in fracture properties [9,10] with somestudies indicating also reduced strength and fatigue life when comparedto predicted values [11]. Such discrepancies are largely attributed tomicroscale architectural irregularities in the randomly dispersed

nanofillers, such as misalignment, agglomerations, and poor dispersionof the CNTs [12,13]. To mitigate these drawbacks, recent advances innanotechnology techniques have been exploited to engineer novel CNTarchitectures, including nanoforests [14–16], which utilize appropriatesubstrates for the growth of highly aligned dense mats of CNTs, fuzzyfibers [17], which contain CNTs radially grown on the microfibers, andCNT ropes [18], which harness the ultra-long strands of CNTs as a re-placement for microfibers. Fuzzy fiber architecture, in particular, hasshown to provide increased in-plane strength, interlaminar shearstrength, and fracture properties [19]]. However, the observed im-provements in properties due to the addition of CNTs is generally lowerthan the predicted theoretical values [20].

Possible causes for discrepancies between experimental results andtheoretical predictions of the CNT nanocomposite properties may be themodeling approaches used for predicting material properties and be-havior. These approaches are often not appropriate for a complex het-erogeneous material system, such as the CNT/CFRP. Conventionalmethodologies predict structural scale properties using one of two ap-proaches: a top-down approach based on a bulk material analysis suchas smeared material techniques [21,22], or a bottom-up approach thatuses mean field techniques, such as the Mori-Tanaka or the composite

https://doi.org/10.1016/j.compstruct.2018.05.075Received 16 December 2017; Received in revised form 28 April 2018; Accepted 17 May 2018

⁎ Corresponding author.E-mail address: [email protected] (A. Chattopadhyay).

Composite Structures 200 (2018) 204–216

Available online 22 May 20180263-8223/ © 2018 Elsevier Ltd. All rights reserved.

T

Page 2: Multifidelity multiscale modeling of nanocomposites for ...

cylinder assemblage techniques [23,24]. In the top-down approach, themacroscale bulk material analysis that is used cannot successfullycapture the unique stress states at the sub-microscale, particularlywhere the CNT has a dominant load transfer effect, and can thus lead tounreliable failure predictions. Additionally, such approaches requirelarge-scale experimental data to produce meaningful results. Mean fieldmicromechanical techniques, on the other hand, tend to overestimatemechanical properties due to the assumption of overly simplistic re-peating unit cells (RUCs), which in turn fail to account for the sto-chasticity in the microscale constituent properties and arrangement. Ineither case, the interaction of the CNTs and the polymer molecules atthe nanoscale is entirely ignored.

In the case of CNT nanocomposites, recent studies suggest that theunique stress state at the nanoscale may lead to a divergence in ob-served and predicted response at the higher length scales [25]. Sub-ramanian et al. developed a linear and post-linear multiscale approachfor CNT-enhanced epoxy polymers by combining molecular dynamics(MD) simulations and the continuum mechanics approach [26]. In theirstudy, the stress-strain response at the nano- and the sub-microscalewas seen to be significantly different from the average bulk stress-strainresponse of the nanocomposite, indicating local stress concentrations,which can then lead to accelerated local damage initiation as demon-strated in [27]. Recent atomistic scale studies also show that bondformation and bond breakage phenomena of the polymer moleculesmay have significant impact on the macroscale elastic and damagebehavior of CFRPs [25,26,28,29], a phenomenon not typically ac-counted for in conventional continuum modeling techniques.

Based on the above discussion, it is clear that a necessary first stepto achieve a systematic understanding of the behavior of nanocompo-site materials at the higher length scales requires comprehensive un-derstanding of the mechanics of load transfer between the nanofiller,the matrix, and the microfiber. One approach for such an investigationcould be through extensive experimental characterization as seen in[30]; however, the large number of physical variables and their com-plex interactions involved in the process makes for a daunting task. Analternative would be to develop new computational tools, which in-clude nanoscale mechanics to resolve atomistic interactions used in theinvestigation of the physical phenomena associated with the loadtransfer and damage mechanisms in CNT/CFRPs.

Numerous CFRP analyses with atomistic integration have beenperformed in the past. They include using a combination of molecularmechanics (MM) or molecular dynamics (MD) simulations and aver-aging techniques, such as elastic homogenization based micro-mechanics [20,31], statistical techniques such as Monte Carlo methods[32], or adopting a fully continuum mechanics finite element (FE) ap-proach [33]. While homogenization provides reasonably accurate re-sults for elastic analysis, it can also result in imprecise outcomes forinelastic and damage analyses due to the loss of spatial variability instress and strain fields [34]. Similarly, statistical techniques may re-quire an inordinately large number of simulations to achieve an accu-rate characterization of the complete nonlinear behavior spectrum. Forexample Monte Carlo simulations of amorphous polymer deformationhave to be severely limited in the spatial and temporal domain to obtainpractical results [32]. It has also been shown that investigating sto-chastic properties of composite materials with m uncertain variablesmay lead to a N m increase in computation time which can only bedecreased slightly using improved sampling schemes such as the LatinHypercube Sampling [35]. Finally, FE methods often fail to capture thefidelity at atomic scales due to the breakdown of the assumptions incontinuum mechanics. Zeng et al., provide a thorough comparison ofthe various methodologies along with the advantages and dis-advantages of each method [36].

In the literature there is considerable emphasis on deriving rigorousmathematical theories for describing the stress state of heterogeneousmedia, wherein the nanoscale mechanics is integrated using variousforms of homogenization, such as the variational asymptotic

homogenized micromechanics models [37,38]], eigen-deformationbased reduced order models [39,40], mathematical homogenized mi-cromechanics coupled with diffused damage models [41], and ad hocmathematical functions such as the boundary condition free micro-mechanics theory [42]. Recent studies have also used concurrent andsemi-concurrent coupled FE and MD simulations to capture nanoscalemechanics explicitly [43]. However, the nanoscale mechanics used inthese works utilize empirical force fields that are inadequate for cap-turing fundamental nonlinearities at the nanoscale, e.g., bond breakageof polymer chains [26]. Moreover, when reactive bond order forcefields are used in these methods, the implementation becomes com-putationally infeasible. It has also been shown that there are certainlocalized phenomena, e.g., stress recovery proceeding initial softeningand damage saturation, which can only be observed explicitly, bymodeling the nanoscale substructures using reactive bond order forcefields [27].

In this paper, a novel multiscale methodology is presented, whichincludes nanoscale mechanics to resolve atomistic interactions, to in-vestigate the physical phenomena associated with the load transfer anddamage mechanisms in CNT/CFRPs. Each constituent of the CNT/CFRPmaterial system is explicitly modeled and assembled into a realisticmicroscale representative unit cells (RUCs) using the finite element (FE)method. The matrix constituent is modeled using a physical damageevolution law that uses the fundamental covalent bond dissociationinformation obtained from MD simulations of the polymer moleculesand chains, implemented using the reactive force field (ReaxFF) [44].Since the CNTs are modeled explicitly, a thorough post-linear in-vestigation of the interaction between the CNTs and the matrix con-stituent is also performed. This study provides insights into the loadtransfer mechanism, and the damage initiation and propagation phe-nomenon, in randomly dispersed CNT nanocomposites and radiallygrown fuzzy fiber nanocomposites, at the sub-micro and micro lengthscales. Additionally, a surrogate orthotropic material model based onthe Schapery potential theory [45], which reproduces the nanopolymerbehavior, is also developed. The material model is calibrated with theresponse obtained from the high-fidelity nanocomposite model, andapplied to a microscale 3D subcell-based sectional micromechanicsframework [46], to replicate the global behavior of the high-fidelityRUC. The calibrated Schapery/Subcell equations can be easily appliedas a user defined material model within laminated plate theories inmost finite element software packages for rapid damage and failureanalysis of CNT enhanced composite components such as skis, golfshafts, baseball bats, tennis rackets etc. The models presented in thispaper may also allow these material systems to be used in criticalstructural and multifunctional roles which is currently avoided due tothe uncertainty in material behavior predictions. This paper also setsthe procedure and workflow for similar investigations with differentepoxy matrix, carbon fiber, or CNTs and thus help composite designersin making optimum use of these material systems.

2. Constituent and model generation

This section details the methodology used to generate the explicitRUC of the CNT/CFRP material system. The three constituents of thenanocomposite, (i) microfiber (ii) polymer (iii) CNTs, are generatedindividually and then combined in a single model within the FE fra-mework using the commercial FE software, ABAQUS, thereby permit-ting the inclusion of deterministic or stochastic geometric and materialproperties. The algorithms and the methods for introducing stochasti-city for all three constituents are also briefly described.

MD simulations of DGEBF (Di-Glycidyl Ether of Bisphenol F) epoxyand DETA (Di-Ethylene Tri-Amine) hardener simulating the curingprocess are performed to yield a distribution of the most likely cross-linking degree for this system. Additionally, MD simulations are used todetermine the relationship between the cross-linking degree and ma-terial properties by recording the response of the MD unit cells under

A. Rai, A. Chattopadhyay Composite Structures 200 (2018) 204–216

205

Page 3: Multifidelity multiscale modeling of nanocomposites for ...

uniaxial loading. Hence, a most likely distribution of the polymer ma-trix properties can be obtained from these simulations. This informationis used in the FE model of the polymer constituent to simulate spatiallyvarying physically consistent matrix properties. As shown in Fig. 1a, thegeneration of the matrix model is scripted in such a way that it is di-vided spatially into multiple sections. Each section is then stochasticallyassigned a cross-linking degree along with its associated materialproperties, which is randomly sampled from the normal distributionsobtained from the MD simulations. The sampling is weighted accordingto the normal distribution; hence cross-linking values around theaverage are represented to a greater extent. Note that some colors maylook similar in Fig. 1a but this is only due to graphical limitations of thevisualizing software. This algorithm ensures material properties withstochastic spatial variations are applied to the polymer unit cells torepresent non-uniform curing at the microscale [25]. The details of theMD simulations can be found in Subramanian et al. [25].

The continuum damage mechanics (CDM) [47,48] based damageformulations, recently developed by the authors [27,44], are used tocompute the damage in the polymer matrix. The goal is to propagatethe fundamental molecular scale phenomenon such as the disassocia-tion of polymer chain covalent bonds and polymer chain sliding, cap-tured using MD simulations, to the continuum scale. The details of theCDM formulations for the polymer system can be found in Rai et al.[44,49] A brief description is presented here for completeness.

According to these formulations, Hooke’s law is derived to form acoherent constitutive equation that can be used to calculate the stress inthe damaged polymer system:

∼ = − − ∼≈σ LD D(1 )(1 ) :v

e∊∊ (1)

where ∼σ is the stress in the system, Dv is the volumetric damage term[48] that quantifies the difference in density between the damaged andpristine material generally associated with the effects of void growth,such that = −

≈LD1 ,ρ

ρ v0

is the polymer stiffness matrix, ∼e∊∊ is the elastic

strain and D is the damage in the polymer. The damage evolutionequation, which describes the temporal evolution of the damage termD, is formulated as a functional form of MD simulation results of thebond disassociation energy density variation under applied strain. Arecently developed hybrid MD simulation framework utilizing ReaxFFwas used to characterize the energy variations due to successive bondbreakages at the nanoscale for thermoset polymers under isothermalconditions and at operating temperatures below the glass transitiontemperature [25,26]. The continuum equation for the damage evolu-tion, shown in Eq. (2), reproduces this atomistic damage process and isdeveloped with parameters that vary according to the cross-linkingdegree, η.

= ⎡⎣⎢

− − −− − + − −

+ ⎤⎦⎥

D λ χη χ χ

η χ η

2

· sgn( )·(1 )| | | |

2(1 )| | (1 ) 11

(2)

where λ is the viscoplastic multiplier, sgn() is the signum function andχ is defined as:

⎜ ⎟= ⎛⎝

⎞⎠

−−

χ YY

2 1η

0

12(1 )

(3)

Fig. 1. Microstructure constituents.

A. Rai, A. Chattopadhyay Composite Structures 200 (2018) 204–216

206

Page 4: Multifidelity multiscale modeling of nanocomposites for ...

where Y is the elastic energy of the system and Y0 is a material para-meter associated with the maximum energy required to begin damagingthe material. Fig. 2 shows the variation of nominal rate of damage

=D λ ( 1) versus normalized elastic energy YY0.

The microfibers are generated using the hard-core model developedin [34] to recreate experimentally observed carbon fiber compositemicrostructures. Fiber positions are perturbed using a Monte Carloapproach until statistical measures indicate that the generated micro-structure can be classified as hard core. The detailed algorithm for theperturbation of the fiber positions to simulate realistic microstructuresis presented by Borkowski and Chattopadhyay [34]. Additionally, thematerial properties that are assigned to each fiber are sampled from aGaussian distribution to represent variation in fiber material. An ex-ample of a simulated microstructure can be seen in Fig. 1b.

Two separate CNT architectures are generated, (i) randomly dis-persed and (ii) radially grown. The CNT vertices are calculated using atransformation formula [27]. The randomly dispersed architecture iscreated by generating the vertex of the CNTs randomly between themicrofibers using the equations:

= −

= −=

x L ϕ θ

y L ϕ θz L ϕ

1 cos( )

1 sin( )CNT

2

CNT2

CNT (4)

where x y z, , are the coordinates of the vertices of the CNT in the globalcoordinate system of the FE model, = = −θ πα ϕ α2 , 2 1, and α is ran-domly generated with a value between 0 and 1. LCNT is the length of asegment of the CNT generated. Fig. 3 shows a schematic of the CNTgeneration algorithm. The schematic contains three segments and fourvertices for illustrative purposes; however, the generated CNT geome-tries contain seven segments with eight vertices per CNT, along with 3Dwaviness. Using Eq. (4), coordinates of each vertex of the CNT geo-metry, x y z, ,i i i, in the global FE coordinate system is obtained. Togenerate consistent CNT architectures, the CNT vertices must be ex-amined for viability, e.g., in the case of the randomly dispersed archi-tecture, vertex x y z, ,i i i is rejected if it is occupied by a microfiber, or if

it is outside the bounds of the RUC. In the case of the radially grownarchitecture, vertex x y z, ,1 1 1 for each CNT is constrained to remain onthe circumference of the microfiber while the other vertices are re-stricted by the same constraints as the randomly dispersed architecture.Fig. 1c and d show an example of the simulated randomly dispersedCNT architecture and the fuzzy fiber CNT architecture. The materialproperties and dimensions of the CNTs employed in this work havebeen obtained from Romanov et al. [50] and are reported in Table 1.Three-dimensional truss elements with 8 nodes per CNT are chosen tomodel the nanotubes since the CNTs provide structural reinforcement inthe matrix only and the stress variation through the CNTs is not ofinterest in this study.

Since traditional meshing methodologies generate highly complexmeshes for these models, the embedded mesh technique [50] is usedhere to embed the microfiber and CNT meshes within the polymer meshalong with periodic boundary conditions. The embedded elementtechnique allows the constraining of the nodal translational degrees offreedom of a group of elements that lie embedded in a group of hostelements. In this technique, the embedded nodal degrees of freedom areappropriately interpolated from the values of the nodal degrees offreedom of the host elements. Usage details of the embedded methodfor CNT matrix models can be found in Rai et al. [44]. The CNT/CFRPmodel generation process is completely automated using Python scriptsthat construct the corresponding finite element (FE) models inABAQUS, with the continuum damage model for the matrix included asa user material subroutine. The final meshed RUC is shown in Fig. 4.

3. Low-fidelity micromechanical model

The high-fidelity model of the CNT/CFRP material system describedin the previous section provides a direct numerical simulation (DNS) ofthe nanocomposite and hence allows for a thorough study of local in-teractions between the three main constituents at the micro and sub-

Fig. 2. Damage evolution for different cross-linking degrees [44].

Fig. 3. Schematic of the CNT generation algorithm.

Table 1Table of CNT Properties [50].

CNT Length 0.5μmCNT Diameter 9 nm

Type Single WalledElastic Modulus 475 GPaPoisson’s ratio 0.35

Fig. 4. Final assembled model.

A. Rai, A. Chattopadhyay Composite Structures 200 (2018) 204–216

207

Page 5: Multifidelity multiscale modeling of nanocomposites for ...

microscale. Such studies are important to realize microstructure kine-matics, which assists in material development studies and fundamentalunderstanding of the mechanics of heterogeneous media. However,performing multiscale studies that involve macroscale elements, con-currently with high-fidelity models of the microscale at each finite

element integration point, can be computationally prohibitive. Since anefficient multiscale structural analysis would require only the RUC re-sponse at the microscale, without the local sub-microscale information,it is advantageous to reproduce the global microscale RUC behaviorwithout running the high-fidelity DNS. Hence, a low-fidelity ortho-tropic damage model based on the Schapery potential theory is devel-oped in this section, which can be calibrated with the DNS of thepolymer or the CNT-polymer mixture applied in a conventional 3Dsubcell micromechanical model to rapidly evaluate the microscale RUCresponse of the composite or nanocomposite material. Fig. 5 shows thedifference between the high fidelity and low-fidelity models schemati-cally. It can be seen that the low-fidelity model approximates the ma-terial behavior using a significantly reduced computational spatial do-main and the corresponding analytical equations permit theemployment of faster numerical tools. However, this comes with thedisadvantage of a lowered understanding of spatial variation in thestress and strain fields along the material constituents.

The progressive damage theory based on the work potential model[45] considers the total strain energy U to be the sum of the elasticstrain energy density We and the dissipated strain energy density Wdsuch that:

= +U W We d (5)

The dissipated strain energy Wd describes the material or geometricirreversible processes occurring at the micro or nanoscale that causes

Fig. 5. Schematic comparison of the high fidelity and low-fidelity models.

Fig. 6. Unitcell discretization in the subcell method [53].

Fig. 7. Comparison of stress-strain response.

A. Rai, A. Chattopadhyay Composite Structures 200 (2018) 204–216

208

Page 6: Multifidelity multiscale modeling of nanocomposites for ...

nonlinearity in the material response. Wd can be described using a set ofinternal state variables Si, which account for all the irreversible pro-cesses occurring within the system. For isotropic materials, accountingfor a single damage source such as polymer matrix microdamage, thedissipated strain energy can be represented using a single internal statevariable S [51]. However, damage in orthotropic material systems canbe directionally dependent and hence, will require a separate internal

state variable for each strain element, ∊ ∊ ∊ ∊ ∊ ∊, , , , ,11 22 33 12 23 31. Thedissipated strain energy can then be expressed as:

∑==

=

W Sdi

i

i1

6

(6)

=S δ si i i (7)

Fig. 8. Randomly dispersed CNT architecture under loading.

A. Rai, A. Chattopadhyay Composite Structures 200 (2018) 204–216

209

Page 7: Multifidelity multiscale modeling of nanocomposites for ...

where si is a state variable that will be defined explicitly and δi is anactivation function used to activate the appropriate internal statevariable at the existence of strain in that direction. This can be ex-pressed as:

= ⎧⎨⎩

∊ >∊ =

δ1 when | | 0 and0 when | | 01

11

11 (8)

=⎧⎨⎩

∊ >∊ =

δ1 when | | 0 and0 when | | 06

12

12

It has been shown that the total strain energy density is stationarywith respect to the changes in internal state variables associated withdamage and structural processes [45,52], such that:

∂∂

=US

0i (9)

Substituting Eq. (5) and (6), in Eq. (9), the following expression can

be obtained:

∂∂

= −WS

1e

i (10)

Using the chain rule, Eq. (10) can be written as:

∂∂

= ∂∂

∂∂

WS

Ws

sS

·e

i

e

i

i

i (11)

The second term in Eq. (11) can be obtained by differentiating Eq.(7) with respect to Si, which is found to be:

∂∂

=sS δ

1i

i i (12)

which can be substituted in Eq. (11) and the result of which can besubstituted in Eq. (10) to obtain:

∂∂

= −Ws

δe

ii (13)

Fig. 9. Magnified images of the CNTs in the randomly dispersed CNT architecture.

A. Rai, A. Chattopadhyay Composite Structures 200 (2018) 204–216

210

Page 8: Multifidelity multiscale modeling of nanocomposites for ...

Eq. (13) represents a system of equations that describes the evolu-tion of the state variables on the application of any form of strain en-ergy. The elastic strain energy as a function of strain for a general or-thotropic system is:

= ∊ + ∊ + ∊ + ∊ + ∊

+ ∊ + ∊ ∊ + ∊ ∊ + ∊ ∊+ ∊ ∊ + ∊ ∊ + ∊ ∊

W C C C C C

C C C CC C C

(

)

e12 11 11

222 22

233 33

244 23

255 13

2

66 122

12 11 22 13 11 33 21 22 11

23 22 33 31 33 11 32 33 22 (14)

where = − = −Ci j1 6, 1 6 are elements of the stiffness matrix of the materialsystem. These system of equations can be reduced to a single equationfor isotropic materials, where = ⋯= =S S S S

1 2 6 6 such that =W Sd , and δ

Fig. 10. Radially grown CNT architecture under loading.

A. Rai, A. Chattopadhyay Composite Structures 200 (2018) 204–216

211

Page 9: Multifidelity multiscale modeling of nanocomposites for ...

is always 1 to avoid triviality. Such a case leads to the original equa-tions for isotropic materials as derived in previous works [35]. Since Cijis dictated by the state variables si, it can be functionally expressed as

= ⋯C f s s( , , )ij 1 6 . These functions can be chosen appropriately to modelthe selected material system, for example, Pineda et al. [51] used apolynomial function to describe the dependence of the elastic constantswith the state variables. In this work, a second order polynomialfunction is used such that:

= + + + + + + + +

+ + + +

C C C s C s C s C s C s C s C s C s

C s C s C s C s

ij ij ij ij ij ij ij ij ij ij

ij ij ij ij

1 21

32

43

54

65

76

812 9

22

1032 11

42 12

52 13

62

(15)

where Cijk are constants. The independent constants can be selectively

calibrated depending on the application of the model and can also bedramatically reduced using symmetry conditions if they exist. As can beseen from Eq. (15), any strain in a certain direction may producechanges in the material properties in the non-dominant directions, theamount of which is governed by Eq. (13). For example, application of

any loading in the transverse (22) direction will activate the statevariable s2, which then produces a change in all the material constants.This methodology simulates the reduced ability of the material to resistloads in non-dominant directions as a consequence of damage in thedominant direction.

In this work, the low-fidelity Schapery damage model is used toreplicate the CNT-epoxy matrix response, which is calibrated using thehigh-fidelity explicit model of the CNT-epoxy system, and applied as thematrix phase in a subcell based micromechanical framework [46]. Sucha methodology allows the investigation of the effects of the matrixphase as well as the fiber phase, and permits the application of classicalmacroscale failure theories to perform a coherent computationally ef-ficient multiscale study that may involve the structural scale. It is to benoted that since the matrix phase response is calibrated with the high-fidelity atomistically informed model, the nanoscale influences areimplicitly accounted for in the surrogate low-fidelity Schapery model ofthe CNT-epoxy system. However, the local sub-microscale informationfrom the high-fidelity model is lost and only the global RUC response of

Fig. 11. Magnified images of the CNTs in the radially grown CNT architecture.

A. Rai, A. Chattopadhyay Composite Structures 200 (2018) 204–216

212

Page 10: Multifidelity multiscale modeling of nanocomposites for ...

the CNT-epoxy system is bridged. The micromechanics framework usedin this work is the 3D sectional micromechanical model developed byZhu et al. [53,54], which is based on the method of cells approach [55]and can be used to compute the effective nonlinear deformation re-sponse of CFRP composites based on the individual constituent re-sponse. The sectional model accounts for in-plane deformations, out of

plane normal deformations and transverse shear deformations, allowingfor 3D characterization and analysis. In this model, a quarter of therepeating fiber-matrix unit cell, which utilizes a square fiber packingarrangement, perfect fiber-matrix interfacial bonding (assumed), andperiodic boundary conditions, is divided into eight rectangular subcells. Symmetry of the unit cells permits the use of a quarter section.The discretization of the unit cell is shown in Fig. 6. Appropriate con-tinuity conditions are applied between the subcells, and volume aver-aging of the appropriate quantities in all the subcells provide the ef-fective stress and strain increments for the unit cell. Using such ananalysis, an accurate study of the microscale and macroscale behaviorcan be performed since the constituent response is computed in-dependently.

4. Results and discussion

4.1. Microstructure analysis

As a case study for microstructure analysis, the microscale RUCswith random and radially grown CNT architectures and 0.1% weightfraction of CNTs are generated using the high-fidelity model generationalgorithm detailed in Section 2. These RUCs are loaded transverse to thefiber, corresponding to the positive x-axis in Fig. 4, and under quasi-static conditions up to complete failure or until the end of the simula-tion.

Fig. 12. Calibration of Schapery model with experimental polymer response.

Table 2Calibration parameters for the polymer Schapery model.

C111 2.383E9 C12

1 1.021E9 C661 1.362E9

C112 10151 C12

2 101025 C662 634

C113 91632 C12

3 101025 C663 634

C114 91632 C12

4 245 C664 1249

C115 1 C12

5 1 C665 1

C116 1 C12

6 1 C666 1

C117 1 C12

7 1 C667 1

C118 1.05 C12

8 0.012 C668 0.0045

C119 1.95 C12

9 0.012 C669 0.0045

C1110 1.95 C12

10 0.002 C6610 0.0045

C1111 0.001 C12

11 0.001 C6611 0.001

C1112 0.001 C12

12 0.001 C6612 0.001

C1113 0.001 C12

13 0.001 C6613 0.001

Fig. 13. Comparison of experimental CFRP response under transverse loading[57] and polymer subcell/Schapery model.

Fig. 14. Comparison of nanocomposite CFRP response from direct numericalsimulation and nanopolymer subcell/Schapery model.

Fig. 15. Stochastic nanocomposite CFRP response under transverse loading.

A. Rai, A. Chattopadhyay Composite Structures 200 (2018) 204–216

213

Page 11: Multifidelity multiscale modeling of nanocomposites for ...

Fig. 7 compares the stress-strain response of the CNT/CFRP modelswith radially grown and randomly dispersed CNT architectures, re-spectively. It is seen that the randomly dispersed CNT architecturedisplays slightly higher stiffness compared to the radially grown ar-chitecture. However, the radially grown architecture shows delayeddamage initiation and slower propagation of damage compared to therandomly dispersed configuration. This observation is consistent withexisting experimental literature [56], where radially grown CNTs dis-played better fracture properties than randomly dispersed CNT nano-composites and traditional composites. An attempt to understand thisphenomenon is undertaken by investigating the damage trends, localstress hot spots, and the sub-microscale stress state at the interactingregion between the CNT and the matrix.

Fig. 8 depicts the progression of damage in the polymer and thestate of stress in the nanotubes in a nanocomposite model with ran-domly dispersed CNTs at various stages of loading. Fig. 8b, d, and fillustrate the axial stress state of the CNTs, where the blue end of thespectrum shows zero stress conditions and red colors of the spectrumshows tension. All negative stresses are visualized in black color; henceall CNTs visualized as black are in compression. Since the maximumcompression stresses are significantly lower than maximum tensilestresses, all negative stresses are chosen to be visualized as black forclarity. The stress state is measured along the axial direction of the tubein the corresponding CNT local coordinate system, which allows thevisualization of the local stress state of each CNT under a globaltransverse load. Additionally, Fig. 9a–c show a magnified region of theCNTs where critical damage is observed. The CNT system is slightlyrotated out-of-plane to emphasize the depth effect. Fig. 8b shows that atstrains below the elastic limit, local regions that are relatively matrixrich display higher local stresses corresponding to stress concentrationzones. Furthermore, since they are randomly oriented, the CNTs thatare most favorably oriented in the loading direction carry the highestload; yet, note that the majority of the CNTs are loaded and activated,leading to an almost uniform stress distribution among the CNTs withstress gradients occurring in matrix rich zones. Additionally, some CNTsare also noted to be in compression due to the Poisson’s effect duringextension.

The contours displayed in Fig. 8a, c and e, depict the damage statein the matrix, with blue representing zero damage and red indicatingcompletely damaged/failed elements. Fig. 8c describes the beginning ofdamage in the structure, initiating at the local stress concentrationzones. The authors have previously demonstrated that local volumeconcentration difference between the CNTs and surrounding matrix innanopolymer can lead to local stress concentration zones, which ac-celerates damage progression due to the stiffness difference betweenthe CNT and the polymer matrix [44]. A similar phenomenon is ob-served in the current model of the nanocomposite, with stress con-centration regions occurring at matrix rich zones that contain relativelylarge volume concentration gradients of CNTs. Fig. 8e shows the finaldamage state of the nanocomposite wherein failure is caused due tomatrix cracks. It can be inferred from Fig. 8f that the matrix cracks leadto CNT pullout in the cracked region. Additionally, the stress state inthe CNTs around the crack zone is reduced due to energy dissipation.The CNT pullout also leads to further increase in volume concentrationgradients of CNTs, which subsequently leads to further acceleration ofthe rate of damage. Hence, the randomly dispersed CNT architecturedoes not take advantage of the unique mechanical properties of theCNTs, thereby leading to a rapid decay of the material after damagesaturation.

Fig. 10 illustrates the progression of damage in the polymer and thestate of stress in the CNTs in a nanocomposite model with radiallygrown CNTs at various stages of loading. Fig. 11a–c show a magnifiedregion of the CNTs where critical damage is observed. The CNT systemis slightly rotated out-of-plane to emphasize the depth effect. Fig. 10bdepicts the state of stress in the CNTs within the elastic limit. It is ob-served that fewer CNTs are activated due to the directionality effects;

however, the activated CNTs are under higher stresses than in therandomly dispersed architecture. The CNTs transverse to the loadingdirection are under high compression stresses due to the Poisson’s ef-fect. The high stresses on the activated CNTs may lead to failure orbuckling of the CNTs. In contrast, the CNTs in the randomly dispersedarchitecture are relatively less susceptible to failure, since more CNTscontribute to the load sharing process and thus, the average stress oneach CNT is lower. Hence, the quality of CNTs play a significantly largerrole in the radially grown architecture. It is also observed that regionsthat display the largest changes in volume concentration of CNTs leadto stress concentration zones. Regions with thin volumes of matrixbetween the microfibers, surrounded by CNTs, show significant stressconcentrations. Fig. 10c demonstrates that the damage initiates andpropagates around these stress concentration zones. However, Fig. 10dshows that the CNTs surrounding the damage path continue to beloaded even after significant polymer damage. This is indicative of CNTbridging, which is possible due to the directionality of the CNTs in theradially grown architecture. CNT bridging slows the crack growth ratein the nanocomposite, thus changing the damage profile compared tothe randomly dispersed architecture.

Fig. 10e shows the damage state of the nanocomposite near failure.Like the randomly dispersed architecture, the nanocomposite fails dueto matrix cracks; however, it is observed that the damaged regions arestrictly contained around the CNTs corresponding to the stress con-centration zones. Unlike the radially grown architecture, the randomlydispersed architecture displayed significant volumetric damage. Theradially grown architecture displayed a concentrated damage areaaround the circumference of the microfibers. This damage mechanismleads to the conclusion that the cracks originate around the CNT matrixinterphase zones, which is supported by previous research [26,27,44].Hence, the architecture of the CNTs can be used to control the direc-tionality of originating damage and its progression. Fig. 10f also showsthat the CNTs around the crack path are deactivated and in a stress-freecondition, indicating CNT pullout.

4.2. Micromechanical analysis

To perform the low-fidelity equivalent of the DNS model, the matrixphase is represented using the equations derived in Section 3. Both thepolymer matrix as well as the nanopolymer matrix can be approxi-mated. As mentioned in Section 3, this model can be calibrated withexperimental or simulated response of standard characterization tests.Fig. 12 shows the calibration of the Schapery model with the responseof a uniaxial quasi-static tensile test of a flat dogbone specimen madefrom Epon E863 Resin and Epi-Cure 3290 hardener (100/27 weightratio). Table 2 lists the constants for the calibrated Schapery model. Thenumber of constants is reduced due to the isotropic nature of thepolymer as well as the intended application of this model, which is theuniaxial transverse loading of the CFRP composite. The calibratedmodel can be used to represent the matrix constitutive equations in thesubcell micromechanics to simulate global fiber-matrix response. Acomparison of the subcell CFRP unit cell response with the calibrateddamage model, and quasistatic CFRP tests under transverse loading[57] can be seen in Fig. 13. The Schapery model determines the damagestate in the matrix phase and the overall unit cell failure is determinedby the modified Hashin’s criteria [35]. It can be seen that with thecombination of an independently calibrated Schapery model and thesubcell micromechanics theory, the global response of the CFRP can bepredicted.

For the analysis of the CNT/CFRP system, the Schapery model mustbe re-calibrated to the nanopolymer response. High-fidelity nanopo-lymer models can be generated using the framework described inSection 2, without the incorporation of the microfibers. Details of thegeneration of such high-fidelity models of the nanopolymer system andits response can be found in Reference [44]. The Schapery model canthen be used to represent the CNT-Epoxy system and applied as the

A. Rai, A. Chattopadhyay Composite Structures 200 (2018) 204–216

214

Page 12: Multifidelity multiscale modeling of nanocomposites for ...

matrix phase in the subcell micromechanics. It should be noted that thecomputation of the low-fidelity Schapery/subcell model was carried outusing a FORTRAN77 program that ran on a desktop computer runningLinux with 4 GB RAM and an Intel E6600 dual-core processor. Eachsimulation took less than 2 seconds of computational time for comple-tion. Contrarily, the high-fidelity FE simulations were carried out usingthe High Performance Computing resources at Arizona State University,specifically the Ocotillo cluster. The simulations were run on a partialnode which contains an Intel Xeon E5-2660 Processor with up to 64 GBof RAM. With the available hardware and 2 CPU ABAQUS tokens, thesimulations required about 6–8 hours of computing. Fig. 14 displays thecomparison of the high-fidelity dispersed CNT/CFRP model responseand the subcell/Schapery model response under uniaxial transverseloading. The subcell/Schapery model captures all the essential char-acteristics of the unit cell response as obtained from the high-fidelitymodel, while utilizing significantly decreased computational time.However, the local sub-microscale information, which can be computedusing the high-fidelity models is lost using the low-fidelity subcell/Schapery methodology but the significant computational efficiency al-lows parametric studies of the CNT/CFRP system while considering thevarious causes of uncertainties, such as variation in volume fraction andpolymer curing.

For a composite system manufactured with IM7 fiber and Di-Glycidyl Ether of Bisphenol F (DGEBF)-based resin system using stan-dard manufacturing processes, it is found that the fiber volume fractionshows a Gaussian distribution with an average of 63.9% and a standarddeviation of 2.21% [35]. Additionally, fundamental variation in cross-linking in the polymer can be calculated from MD simulations for theDGEBF resin system and is found to be distributed normally. Theaverage and standard deviation of the cross-linking variation is calcu-lated to be 56.02% and 4.11%, respectively [25]. These two randomvariables, approximated using normal distributions, can be randomlysampled and applied as input to the subcell model to compute thedistribution of nanocomposite CFRP response under uncertainty.Fig. 15 exhibits the result of 1000 simulations of the subcell model withrandomly sampled cross-linking and volume fraction information. TheCNT-epoxy matrix is represented using the previously calibratedSchapery model and computes the degradation of the matrix phase. Thefigure shows a spectrum of response for the dispersed CNT/CFRPsystem and it can be observed that although the elastic response of thematerial does not present large variabilities, the damage and failureresponse is associated with significant stochasticity.

5. Conclusions

In this paper, a methodology for direct numerical simulations andlow-fidelity surrogate nanocomposite modeling and analysis were de-tailed. These tools were used to generate nanocomposites with CNTs asnanofillers for deterministic and stochastic studies with various levelsof detail. For microstructural analysis, the high-fidelity direct numericalsimulation is used to understand the phenomenon of damage in dis-persed and radially grown CNT/CFRP systems. Since the high-fidelitymodel is coupled with a previously developed multiscale damage for-mulation that utilizes atomistic information of the polymer chain mo-tion and covalent bond dissociation, an accurate investigation at thesub microscale could be performed. As a case study, the elastic anddamage state of nanocomposites with 0.1% weight fraction CNTs inrandomly dispersed and radially grown configurations were studiedunder quasi-static loading and under loading transverse to the fiberaxis. The study provided the following critical insights: (i) due to therandomness of the orientation of the CNTs in the randomly dispersedarchitecture, the load sharing between CNTs is less efficient, resultingin more CNTs being activated. However, in the radially grown archi-tecture, fewer CNTs were activated, but load sharing was more efficientleading to higher stresses on the activated CNTs; (ii) the higher stresseson the activated CNTs in the radially grown architecture can lead to

failure of the CNTs and hence, the quality of CNTs plays an importantrole in this architecture; (iii) the damage zones occurred at regions withlarge local volume concentration gradients of the CNTs. Thus, the CNTarchitectures may be engineered to direct damage for greater nano-composite mechanical performance. Furthermore, the work potentialtheory was further extended to account for orthotropy and damagehistory, and was applied as a surrogate constitutive model for thepolymer and CNT-epoxy matrix that was calibrated with the high-fi-delity models. By using the Schapery model as a surrogate in combi-nation with 3D subcell-based micromechanics techniques, the CNT/CFRP unit cell response under uniaxial transverse loading could bereproduced with significantly increased computational efficiency. Itwas shown that such models can be easily adapted for probabilistic andparametric studies.

Acknowledgment

This research is supported by the Office of Naval Research (ONR),Grant No.: N00014-17-1-2037 and N00014-17-1-2029. The programmanager for both grants is Mr. William Nickerson.

References

[1] Thostenson ET, Ren Z, Chou T-W. Advances in the science and technology of carbonnanotubes and their composites: a review. Compos Sci Technol 2001;61:1899–912.

[2] Yu X, Kwon E. A carbon nanotube/cement composite with piezoresistive properties.Smart Mater Struct 2009;18:055010.

[3] Balazs AC, Emrick T, Russell TP. Nanoparticle polymer composites: where two smallworlds meet. Science 2006;314:1107–10.

[4] Godara A, Mezzo L, Luizi F, Warrier A, Lomov SV, Van Vuure A, Gorbatikh L,Moldenaers P, Verpoest I. Influence of carbon nanotube reinforcement on the pro-cessing and the mechanical behaviour of carbon fiber/epoxy composites. Carbon2009;47:2914–23.

[5] Green KJ, Dean DR, Vaidya UK, Nyairo E. Multiscale fiber reinforced compositesbased on a carbon nanofiber/epoxy nanophased polymer matrix: synthesis, me-chanical, and thermomechanical behavior. Compos Part A 2009;40:1470–5.

[6] Inam F, Wong DW, Kuwata M, Peijs T. Multiscale hybrid micro-nanocompositesbased on carbon nanotubes and carbon fibers. J Nanomater 2010;2010:9.

[7] Cho J, Daniel I, Dikin D. Effects of block copolymer dispersant and nanotube lengthon reinforcement of carbon/epoxy composites. Compos Part A 2008;39:1844–50.

[8] Sochi EJ. Challenges for insertion of structural nanomaterials in aerospace appli-cations. In 15th European Conference on Composite Materials.

[9] Gojny FH, Wichmann M, Köpke U, Fiedler B, Schulte K. Carbon nanotube-reinforcedepoxy-composites: enhanced stiffness and fracture toughness at low nanotubecontent. Compos Sci Technol 2004;64:2363–71.

[10] Qiu J, Zhang C, Wang B, Liang R. Carbon nanotube integrated multifunctionalmultiscale composites. Nanotechnology 2007;18:275708.

[11] Ren X, Burton J, Seidel GD, Lafdi K. Computational multiscale modeling andcharacterization of piezoresistivity in fuzzy fiber reinforced polymer composites. IntJ Solids Struct 2015;54:121–34.

[12] Wicks SS, de Villoria RG, Wardle BL. Interlaminar and intralaminar reinforcementof composite laminates with aligned carbon nanotubes. Compos Sci Technol2010;70:20–8.

[13] Savvas D, Stefanou G, Papadopoulos V, Papadrakakis M. Effect of waviness andorientation of carbon nanotubes on random apparent material properties and rvesize of cnt reinforced composites. Compos Struct 2016;152:870–82.

[14] Zhang Y, Lau S, Huang L, Tay B. Carbon nanotubes grown on cobalt-containingamorphous carbon composite films. Diamond Rel Mater 2006;15:171–5.

[15] Gürkan İ, Cebeci H. An approach to identify complex cnt reinforcement effect on theinterlaminar shear strength of prepreg composites by taguchi method. ComposStruct 2016;141:172–8.

[16] Ürk D, Demir E, Bulut O, Çakroğlu D, Cebeci FÇ, Öveçoğlu ML, Cebeci H.Understanding the polymer type and cnt orientation effect on the dynamic me-chanical properties of high volume fraction cnt polymer nanocomposites. ComposStruct 2016;155:255–62.

[17] Garća EJ, Hart AJ, Wardle BL. Long carbon nanotubes grown on the surface of fibersfor hybrid composites. AIAA J 2008;46:1405–12.

[18] Chou T-W, Gao L, Thostenson ET, Zhang Z, Byun J-H. An assessment of the scienceand technology of carbon nanotube-based fibers and composites. Compos SciTechnol 2010;70:1–19.

[19] Li R, Lachman N, Florin P, Wagner HD, Wardle BL. Hierarchical carbon nanotubecarbon fiber unidirectional composites with preserved tensile and interfacialproperties. Compos Sci Technol 2015;117:139–45.

[20] Kundalwalal S, Kumar S, Wardle B. Multiscale modeling of microscale fiber re-inforced composites with nano-engineered interphases. arXiv preprint arXiv:1509.05140; 2015.

[21] Hashin Z. Failure criteria for unidirectional fiber composites. J Appl Mech1980;47:329–34.

[22] Tsai SW, Wu EM. A general theory of strength for anisotropic materials. J Compos

A. Rai, A. Chattopadhyay Composite Structures 200 (2018) 204–216

215

Page 13: Multifidelity multiscale modeling of nanocomposites for ...

Mater 1971;5:58–80.[23] Liu H, Brinson LC. Reinforcing efficiency of nanoparticles: a simple comparison for

polymer nanocomposites. Compos Sci Technol 2008;68:1502–12.[24] Seidel GD, Lagoudas DC. Micromechanical analysis of the effective elastic proper-

ties of carbon nanotube reinforced composites. Mech Mater 2006;38:884–907.[25] Subramanian N, Rai A, Chattopadhyay A. Atomistically informed stochastic mul-

tiscale model to predict the behavior of carbon nanotube-enhanced nanocompo-sites. Carbon 2015;94:661–72.

[26] Subramanian N, Koo B, Rai A, Chattopadhyay A. A multiscale damage initiationmodel for cnt-ehanced epoxy polymers. In 20th International Conference onComposite Materials. Copenhagen, Denmark. pp. 4410–8.

[27] Rai A, Subramanian N, Chattopadhyay A. Investigation of piezo-resistivity in cntnano-composites under damage. In SPIE Smart Structures and Materials+Nondestructive Evaluation and Health Monitoring, International Society for Opticsand Photonics.

[28] Yang M, Koutsos V, Zaiser M. Interactions between polymers and carbon nanotubes:a molecular dynamics study. J Phys Chem B 2005;109:10009–14.

[29] Yang S, Yu S, Ryu J, Cho J-M, Kyoung W, Han D-S, Cho M. Nonlinear multiscalemodeling approach to characterize elastoplastic behavior of cnt/polymer nano-composites considering the interphase and interfacial imperfection. Int J Plast2013;41:124–46.

[30] Gibson RF. A review of recent research on mechanics of multifunctional compositematerials and structures. Compos Struct 2010;92:2793–810.

[31] Papadopoulos V, Tavlaki M. The impact of interfacial properties on the macroscopicperformance of carbon nanotube composites. A fe 2-based multiscale study. ComposStruct 2016;136:582–92.

[32] Chui C, Boyce MC. Monte carlo modeling of amorphous polymer deformation:evolution of stress with strain. Macromolecules 1999;32:3795–808.

[33] Fisher F, Bradshaw R, Brinson L. Effects of nanotube waviness on the modulus ofnanotube-reinforced polymers. Appl Phys Lett 2002;80:4647–9.

[34] Borkowski L, Liu K, Chattopadhyay A. From ordered to disordered: the effect ofmicrostructure on composite mechanical performance. Comput Mater Continua2013;37:161–93.

[35] Johnston JP. Stochastic multiscale modeling and statistical characterization ofcomplex polymer matrix composites [Ph.D. thesis]. Arizona State University; 2016.

[36] Zeng Q, Yu A, Lu G. Multiscale modeling and simulation of polymer nanocompo-sites. Prog Polym Sci 2008;33:191–269.

[37] Yu W, Tang T. A variational asymptotic micromechanics model for predictingthermoelastic properties of heterogeneous materials. Int J Solids Struct2007;44:7510–25.

[38] Oskay C, Fish J. Fatigue life prediction using 2-scale temporal asymptotic homo-genization. Int J Numer Meth Eng 2004;61:329–59.

[39] Oskay C, Fish J. Eigendeformation-based reduced order homogenization for failureanalysis of heterogeneous materials. Comput Methods Appl Mech Eng2007;196:1216–43.

[40] Bogdanor MJ, Oskay C. Prediction of progressive damage and strength of im7/977-3 composites using the eigendeformation-based homogenization approach: staticloading. J Compos Mater 2016;0021998316650982.

[41] Murari V, Upadhyay C. Micromechanics based ply level material degradation modelfor unidirectional composites. Compos Struct 2012;94:671–80.

[42] Peng B, Yu W. A new micromechanics theory for homogenization and dehomo-genization of heterogeneous materials. American Society of Composites-30thTechnical Conference; 2015.

[43] Talebi H, Silani M, Bordas SP, Kerfriden P, Rabczuk T. A computational library formultiscale modeling of material failure. Comput Mech 2014;53:1047–71.

[44] Rai A, Subramanian N, Koo B, Chattopadhyay A. Multiscale damage analysis of cntnanocomposite using a continuum damage mechanics approach. J Compos Mater2016.

[45] Schapery R. A theory of mechanical behavior of elastic media with growing damageand other changes in structure. J Mech Phys Solids 1990;38:215–53.

[46] Zhu L, Chattopadhyay A, Goldberg R. A 3d micromechanics model for strain ratedependent inelastic polymer matrix composites. In 47th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference 14th AIAA/ASME/AHS Adaptive Structures Conference 7th. p. 1689.

[47] Lemaitre J. A continuous damage mechanics model for ductile fracture. J Eng MaterTechnol 1985;107:83–9.

[48] Chaboche J, Boudifa M, Saanouni K. A cdm approach of ductile damage with plasticcompressibility. Int J Fract 2006;137:51–75.

[49] Rai A, Subramanian N, Chattopadhyay A. Investigation of damage mechanisms incnt nanocomposites using multiscale analysis. Int J Solids Struct 2017.

[50] Romanov VS, Lomov SV, Verpoest I, Gorbatikh L. Modelling evidence of stressconcentration mitigation at the micro-scale in polymer composites by the additionof carbon nanotubes. Carbon 2015;82:184–94.

[51] Pineda EJ, Waas AM, Bednarcyk BA, Collier CS, Yarrington PW. Progressive damageand failure modeling in notched laminated fiber reinforced composites. Int J Fract2009;158:125–43.

[52] Schapery R. Mechanical characterization and analysis of inelastic composite lami-nates with growing damage. Mech Compos Mater Struct 1989:1–9.

[53] Zhu L. Multiscale high strain rate models for polymer matrix composites [Ph.D.thesis]. Arizona State University; 2006.

[54] Zhu L, Chattopadhyay A, Goldberg RK. A 3d micromechanics model for strain ratedependent inelastic polymer matrix composites. In Proc., 47th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics & Materials Conference.

[55] Aboudi J. Micromechanical analysis of composites by the method of cells. ApplMech Rev 1989;42:193–221.

[56] Wicks SS, Wang W, Williams MR, Wardle BL. Multi-scale interlaminar fracturemechanisms in woven composite laminates reinforced with aligned carbon nano-tubes. Compos Sci Technol 2014;100:128–35.

[57] Gilat A, Goldberg RK, Roberts GD. Experimental study of strain-rate-dependentbehavior of carbon/epoxy composite. Compos Sci Technol 2002;62:1469–76.

A. Rai, A. Chattopadhyay Composite Structures 200 (2018) 204–216

216


Recommended