You are on page 1of 51

Ref.

Ares(2019)4063597 - 26/06/2019

Project Reference: 721105

DELIVERABLE D1.2:

REPORT ON MATERIAL MODELS DESCRIBING


PROPERTY CALCULATIONS, THEIR INTERFACES
AND UNCERTAINITY QUANTIFICATION
Lead Beneficiaries UniTS
Contributing Beneficiaries LIST, e-Xstream, POLITO, GRANTA, INSA & UniTS
Deliverable Nature R (Report)1
Deliverable Identifier D1.2
Dissemination level PU2
Contractual date of Delivery 30-06-2019 (M30)

Project Title Multi-scale Composite Material Selection Platform with a


Seamless Integration of Materials Models and
Multidisciplinary Design Framework
Project website www.composelector.net
Contact (Project Coordinator) Dr Salim Belouettar salim.belouettar@list.lu
Call H2020-NMBP-23-2016
Topic Advancing the integration of Materials Modelling in Business
Processes to enhance effective industrial decision making
and increase competitiveness
EC Project Officer Dr. Anne F. de Baas

1R: Document, report (excluding the periodic and final reports), DEM: Demonstrator, pilot, prototype, plan designs, DEC: Websites, patents filing,
press & media actions, videos, etc. OTHER: Software, technical diagram, etc.
2PU = Public, fully open, e.g. web. CO = Confidential, restricted under conditions set out in Model Grant Agreement, CI = Classified, information as
referred to in Commission Decision 2001/844/EC.

Deliverable D1.2: Material models, interfaces and uncertainty quantification 1


Project Reference: 721105

AUTHORS

Lead Authors (Org) S. Pricl (UniTS), E. Laurini (UniTS), D. Marson (UniTS), M.


Fermeglia (UniTS)
Contributing Author(s) (Org) V. Reigner (e-Xstream), P. Asinari (POLITO), M. Petrolo
(POLITO), A. Madeo (INSA), & S. Belouettar (LIST)
Reviewers (Org) B. Patzak (CTU), P. de Luca (ESI) and D. Dykenman (Granta)
Validation (Org) S. Belouettar (LIST)

ACRONYMS
- 3D: Three-dimensional
- AEP: Aminoethylpiperazine
- APDMES: 3-Aminopropyldiethoxymethylsilane
- PEEK: Polyther ether ketone
- CAD: Computer-aided design
- CAE: Computer-aided engineering
- CFs: Carbon fibers
- CPDMES: 3-Chloropropyldimethylethoxysilane
- CPU: Central processing units
- DETA: Diethylenetriamine
- DGEBA: Diglycidyl ether of bisphenol A
- DPD: Dissipative particle dynamics
- EDA: Ethylenediamine
- EPON862: Trade name of diglycidyl ether of bisphenol F
- ESL: Equivalent single layer
- GPTMS: 3-Glycidoxypropyltrimethoxysilane
- IPD: Isophorone diamine
- KPI: Key performance indicator
- LW: Layer-wise
- MAE: Mean absolute error
- MC: Monte Carlo
- MD: Molecular dynamics
- MDO: Multidisciplinary design optimization
- MFH: Mean field homogenization
- NA: Not available
- NPs: Nanoparticles
- MODA: Material modeling data
- MTMS: Methyltrimethoxysilane
- MWCNTs: Multi-walled carbon nanotubes
- ODES (OCTEO): Octyltriethoxysilane
- PCE: Polynomial chaos expansion
- RMSE: Root-mean-square error
Deliverable D1.2: Material models, interfaces and uncertainty quantification 2
Project Reference: 721105

- SBR: Styrene-butadiene rubber


- TEPA: Tetraethylenepentamine
- TETA: Triethylenetetramine
- UQ: Uncertainity quantification

Disclaimer3:

3This document’s contents are not intended to replace consultation of any applicable legal sources or the necessary
advice of a legal expert, where appropriate. All information in this document is provided “as is” and no guarantee or
warranty is given that the information is fit for any particular purpose. The user, therefore, uses the information at its
sole risk and liability. For the avoidance of all doubts, the European Commission has no liability in respect of this
document, which is merely representing the authors’ view.

Deliverable D1.2: Material models, interfaces and uncertainty quantification 3


Project Reference: 721105

TABLE OF CONTENT
Authors .......................................................................................................................................................... 2
Acronyms ...................................................................................................................................................... 2
Table of Content............................................................................................................................................ 4
List of figures ................................................................................................................................................ 5
List of tables .................................................................................................................................................. 8
Deliverable description as reported in the DoA ........................................................................................ 9
Summary of the deliverable D1.2 ................................................................................................................ 9
CASE STUDY 1: Material and process selection of fuselage thermoplastic frame (Airbus) .............. 11
1.a. Atomistic and mesoscopic models (partner: UniTS) ................................................................... 11
1.b. Mean field homogenization models (partner: e-Xtream) ............................................................. 15
1.c. Enriched hyperelastic continuum models (main partner: INSA-Lyon) ...................................... 22
1.d. Finite-element (FE) models with enriched, variable kinematics (main partner: POLITO-MUL2
group) ....................................................................................................................................................... 24
CASE STUDY 2: Material and process selection of leaf-spring (Dow) ................................................. 27
1.f. Atomistic and mesoscopic models (partner: UniTS) .................................................................... 27
1.g. Mean field homogenization models (partner: e-Xstream) ........................................................... 34
1.h. Enriched hyperelastic continuum models (main partner: INSA-Lyon) ...................................... 35
1.i. LIST ..................................................................................................................................................... 35
CASE STUDY 3: Material and process selection for car tyre (Goodyear) ............................................ 41
1.j. Atomistic and mesoscopic models (partner: UniTS) .................................................................... 41
1.k. Full field homegenization models (partner: e-Xstream) .............................................................. 45
1.l. LIST ..................................................................................................................................................... 48
Embedding/wrapping methodology for performing material models sensitivity analysis and
uncertainty quantification (UQ) ................................................................................................................. 49
1.m Sensitivity analysis and uncertainty quantification of molecular dynamics data (partner:
POLITO-SMaLL group) ........................................................................................................................... 49
1.n. Global Sensitivity Analysis (partner: LIST) ................................................................................... 51

Deliverable D1.2: Material models, interfaces and uncertainty quantification 4


Project Reference: 721105

LIST OF FIGURES

Figure 1a. (A) Specific volume Vsp of pure PEEK and of a PEEK/MWCNTs nanocomposite (6 %wt) as a function of
temperature T. Only these two curves are shown for clarity. (B) Comparison of simulated (patterned bars) and
experimental values (filled bars) of Tg for pure PEEK and PEEK/MWCNTs nanocomposites as a function of
MWCNTs concentration (%wt). Experimental data from Sandler et al., Composites: Part A 33 (2002) 1033-1039;
Deng et al., Comp. Sci. Technol. 67 (2007) 2959-2964; Ogasawara et al., Comp. Sci. Technol. 71 (2011) 73-78. (C)
Simulated values of Tg for pure PEEK and PEEK/CFs nanocomposites as a function of MWCNTs concentration
(%wt). ......................................................................................................................................................................... 11
Figure 2a. (A) Thermal conductivity of random, and aligned MWCNTs in PEEK as a function of MWCNTs content (%wt).
(B) Thermal conductivity enhancement factor (EF = composite/matrix) for the systems in (A). Par = thermal
conductivity along the alignment of the MWCNTs; Perp = thermal conductivity perpendicular to the MWCNTs
alignment. .................................................................................................................................................................. 12
Figure 3a. (A) Comparison between the average value of the Young modulus E, calculated using data in Table 1
(patterned bars), and the corresponding experimental measured data (filled bars) for pure PEEK and
PEEK/MWCNTs nanocomposites as a function of MWCNTs concentration (%wt). Experimental data from Deng et
al., Comp. Sci. Technol. 71 (2011) 73-78. (B) Average Poisson ratio ν values, calculated using data in Table 1a for
pure PEEK and PEEK/MWCNTs nanocomposites as a function of MWCNTs concentration (%wt). .......................... 13
Figure 4a. (A) MD-predicted values of the Young modulus E for pure PEEK and PEEK/CFs nanocomposites as a function
of CFs concentration (%wt). (B) Young modulus enhancement factor EF (= E composite/Ematrix) for the systems in panel
A.................................................................................................................................................................................. 13
Figure 5a. (A) Behavior of the cohesive force (Fcoh) at the interface of a PEEK/MWCNT system in response to nanotube
pulling along the z (normal) direction. The value of the maximum in the Fcoh/displacement curve is estimated to
be equal to 1131 pN at the corresponding displacement of 0.0019 nm. (B) The corresponding profile of the
pullout force Fpo vs. displacement obtained by pulling the MWCNT along the x direction. The maximum value of
Fpo locates at a displacement of 0.2, and is equal to 120 pN. .................................................................................... 14
Figure 6a. Predicted values of (A) Storage modulus G’(ω) and (B) loss modulus G”(ω) as a function of the applied
oscillatory shear frequency  of PEEK/MWCNT nanocomposite systems as a function of the MWCNT content
(%wt). Symbol legend: circles, pure PEEK; squares, 3%; triangles, 6%; diamonds; 9%; hexagons, 12%; inverted
triangles, 15%. Corresponding error bars are also displayed for all cases (in many instances, the errors are too
small to be visible). Simulated data agree with the experimental evidence provided for PEEK/MWCNT
nanocomposites in a concentration range encompassing the values considered in this work (Bangarusampath et
al., Polymer 50 (2009) 5803-5811). For such real systems, the behavior of G’(ω) and G”(ω) is in excellent
qualitative agreement with the simulated data reported in Figure 6a; the percolation threshold in that case was
found to be equal to 1% wt, again in good agreement with the one predicted by the present combination of MD +
DPD simulations. ........................................................................................................................................................ 14
Figure 1b. Equilibrated MD simulation snapshot of the PEEK/MWCNTs used to estimate the values of the mechanical
quantities listed in Table 1b. The PEEK chains are shown as transparent spheres while MWCNTs are portrayed as
dark gray sticks. .......................................................................................................................................................... 16
Figure 2b. Comparison of the values of the Young modulus E for PEEK/MWCNTs (nano)composites with random
alignment of the MWCNTs in the PEEK matrix as function of the volume fraction V f as predicted by MD and MFH
simulations with and without calibration (data from Table 2b). ............................................................................... 17
Figure 3b. Representation of the elastic isotropic model used to simulate CF-reinforced PEEK/MWCNTs composites
with random alignment of MWCNTs within the polymer matrix. MWCNTs are shown as thin blue lines while the
CFs are shown as thick red cylinders. The PEEK matrix is not shown for clarity. ...................................................... 18
Figure 4b. Mean field model of the MWCNT inclusions (dark blue) surrounded by the coating (lighter blue). .............. 19
Figure 5b. Comparison of the values of the axial and transversal Young modulus component E for PEEK/MWCNTs
(nano)composites with perfect alignment of the MWCNTs in the PEEK matrix as function of the volume fraction V f
as predicted by MD and MFH simulations with calibration (data from Table 6b). ................................................... 20
Figure 6b. Representation of the elastic isotropic model used to simulate CF-reinforced PEEK/MWCNTs composites
with perfect alignment of MWCNTs within the polymer matrix. MWCNTs are shown as thin blue lines while the
CFs are shown as thick red cylinders. The PEEK matrix is not shown for clarity. ...................................................... 21
Figure 7b. Sketch of the model adopted in studying the effect of different MWCNTs orientation in PEEK/MWCNTs/CFs
(nano)composites using MFH algorithms. ................................................................................................................. 21

Deliverable D1.2: Material models, interfaces and uncertainty quantification 5


Project Reference: 721105

Figure 1c. Forming of the AIRBUS airframe (left) and of the DOW leaf-spring (right) through the developed enriched
continuum (second gradient) model.......................................................................................................................... 23
Figure 2c. Standardized enriched continuum model input/output definition for implementation in MuPIF platform. .. 24
Figure 1d. MUL2 FE model of the AIRBUS component. ..................................................................................................... 24
Figure 2d. Example of LW assembly of FE arrays maintaining the geometrical and material properties of each ply. ..... 25
Figure 3d. MUL2 and ABAQUS FE models of the AIRBUS component. ............................................................................. 26
Figure 4d. ABAQUS and MUL2 first buckling mode. .......................................................................................................... 26
Figure 3d. MUL2 in the COMPOSELECTOR workflows....................................................................................................... 26
Figure 1f. (A) Density as a function temperature for different degree of curing of neat epoxy. Error bars are omitted
for clarity. (B) Tg vs. degree of curing for neat epoxy. The numerical values at the different degree of curing (in
parenthesis) are Tg (K) = 379 (20%), 390 (40%), 409 (55%), 418 (70%), and 450 (85%). ........................................... 27
Figure 2f. (A) Young modulus as a function of temperature for different degree curing of neat epoxy resin. (B) T g vs.
degree of curing for neat epoxy resin obtained from the midpoints of the curves in panel A (patterned bars). The
corresponding Tg obtained from MD-predicted density vs. temperature curves (Figure 19) are also shown (filled
bars) for comparison. The numerical values at the different degree of curing (in parenthesis) are T g (K) = 379
(20%), 390 (40%), 409 (55%), 418 (70%), and 450 (85%)........................................................................................... 28
Figure 3f. (A) Stress-strain behavior and (B) Young modulus of pure epoxy and epoxy/MWCNTs nanocomposites as a
function of MWCNTs concentration (% wt) as predicted by atomistic MD simulations. .......................................... 28
Figure 4f. (A) Comparison between simulated and experimental values of the Young modulus E for pure epoxy and
epoxy/MWCNTs as a function of MWCNTs concentration. (B) Comparison between simulated and experimental
values of the tensile strength for epoxy/MWCNTs as a function of MWCNTs concentration. (C) Comparison of
failure strain for epoxy/MWCNTs (circles) and PEEK/MWCNTs (squares) nanocomposite systems as a function of
MWCNTs concentration (% wt). Experimental data from Tsuda et al, Composites: Part A 65 (2014) 1–9. ............. 29
Figure 5f. (A) DPD back-mapped MD simulated true stress-strain curves of pristine epoxy resin at 20% and 70% degree
of curing. (B) Comparison between the Young modulus values for neat epoxy resin at different degree of curing
obtained from atomistic MD (filled bars) and DPD back-mapped MD simulations (patterned bars). ...................... 29
Figure 6f. (A) Young modulus E for epoxy/MWCNTs composites at different degree of curing and at 7% wt nanofiller
concentration. (B) Young modulus enhancement factor EF for the systems in panel A. .......................................... 30
Figure 7f. Model developed to simulate interface properties in epoxy-resin/MWCNTs nanocomposites (see text for
details). ....................................................................................................................................................................... 30
Figure 8f. (A) Master curve for the cross-linking conversion showing that the resin structure (final or intermediate)
does not depend on the rate of crosslinking, but only on the degree of curing. (B) The indistinguishable RDF
curves confirm that the local structure is the same for all rcross values at the end of the simulation. rcross = critical
distance used in the crosslinking formation criterion. .............................................................................................. 32
Figure 9f. DPD representation of DGEBA (left) and TETA (right) molecules. .................................................................... 32
Figure 10f. Bead density profiles averaged over time for the spatial direction, x, during DPD crosslinking simulation.
Bead densities have homogenous average density distribution in the yz-plane as a function of x. Profiles in the
other two directions are similar. ................................................................................................................................ 32
Figure 11f. Time evolution of reacted and unreacted beads for fast (1) and slow (0.5) crosslinking reactions
(probabilities of secondary amine-epoxy reaction = 1 and 0.5, respectively). % are shown for comparison between
bead types. Crosslinking reaction is switched-on at t (DPD) = 4000. ........................................................................ 33
Figure 12f. Chemical structure of DGEBA monomers and of all different hardeners considered. ................................... 34
Figure 13f. Density at 25°C (left) and Tg (right) values for fully cross-linked (85%) DGEBA resins cured with different
hardeners. .................................................................................................................................................................. 34
Figure 14f. Young modulus (left), stress (middle), and strain (right) values for fully cross-linked (85%) DGEBA cured
with different hardeners. ........................................................................................................................................... 34
Figure 1-i: Multiscale simulation strategy: Hierarchical bottom modelling and simulation strategy. Fiber diameters are
of the order of 10 μm, while ply thicknesses are in the range 300 μm and the laminates are several mm in
thickness and above. .................................................................................................................................................. 36
Figure 2-i: Multiscale simulation strategy: Hierarchical bottom modelling and simulation strategy. MuPIF workflow for
the Leaf Spring user case............................................................................................................................................ 37
Figure 3i: Inputs data used at the micro-mechanics scale. EL: Axial Young modulus, ET: In-plane Young modulus, GTT:
In-plane shear modulus, GLT: Transverse shear modulus, nuTT: In-plane Poisson ratio and nuLT: Transverse
Poisson ratio ............................................................................................................................................................... 38
Figure 4i: Generated data at the macro scale (part level). ................................................................................................ 40
Figure 5i: Processing workflow. ......................................................................................................................................... 40
Deliverable D1.2: Material models, interfaces and uncertainty quantification 6
Project Reference: 721105

Figure 1j. Wetting-adhesion map for pristine SiO2 and SiO2 nanoparticles surface-modified with different silane-based
compatibilizers in a SBR matrix of fixed composition (26%S/74%B). Solid colored lines:  isocontours; red broken
lines, ΔWa), isocontours (see text for more details). ................................................................................................ 43
Figure 2j. Morphologies of the dispersion of SiO2 nanoparticles surface-modified with OCTEO (dg = 1 chain/nm2) as
obtained from TEM images (left) and from MD-parameterized DPD simulations (right) at two different
nanoparticle loading: 8%wt (top panel), and 20%wt (bottom panel). In the simulation images, the polymer matrix
is depicted as green (top) and gray (bottom) sticks while the modified SiO2 nanoparticles are shown in purple and
orange sphere representations, respectively. The area enclosed in the red boxes on the right images corresponds
to the size of the simulation boxes on the left. ......................................................................................................... 44
Figure 3j. Predicted values of the storage modulus G’(ω) and loss modulus G”(ω) for a SBR matrix loaded with OCTEO-
modified (dg = 1 chain/nm2) silica particles of PEEK/MWCNT nanocomposite systems as a function of the modified
SiO2 content (%wt). .................................................................................................................................................... 45
Figure 1k. Continuum model representation of the dispersion of 10%wt OCTEO –modified silica nanoparticles in the
SBR matrix as mapped from the corresponding DPD simulation. ............................................................................. 46
Figure 2k. Nearest neighbor distance distribution for the dispersion of 10%wt of OCTEO-modified silica nanoparticles
in the SBR matrix. Normalized size: 1 unit = 116 nm; radius of the silica particle = 0.043 (corresponding to a real
value of 5 nm, according to the experimentally employed nanoparticles.) .............................................................. 46
Figure 3k. Uniaxial (left), biaxial (middle), and planar (right) loading results for the dispersion of 10%wt of OCTEO-
modified silica nanoparticles in the SBR matrix. ........................................................................................................ 47
Figure 4k. Composite stress-strain (large deformation) curves for the dispersion of 10%wt of OCTEO-modified silica
nanoparticles in the SBR matrix. ................................................................................................................................ 47
Figure 5k. Effect of silica content on the SBR-composite deformation gradient-stress curves (1-direction) for quasi-
steady axial loading in the 1-direction. ...................................................................................................................... 48
Figure 1l. Multiscale simulation workflow for the tire model. .......................................................................................... 49
Figure 1m. (A) Workflow overview, I: Initial configuration, topology and result folder setup, II: energy minimization, III:
NVT equilibration, IV: NPT equilibration, V: 1st MD run: SDS adhesion, VI: 2nd MD run: SDS push, VII: constraints
checking and post-processing. (B) standalone MD simulation subsystem used in stages II and III. (C) continued MD
simulation subsystem used in stages IV, V and VI. .................................................................................................... 50

Deliverable D1.2: Material models, interfaces and uncertainty quantification 7


Project Reference: 721105

LIST OF TABLES

Table 1a. Values of the average elastic constant (C11,av) , average bulk modulus Kav, average Poisson ratio av, and
average Young modulus Eav for the pristine PEEK matrix and PEEK/MWCNTs nanocomposites as a function of the
nanofiller concentration as derived from unidirectional compression/tension and hydrostatic
compression/tension simulations, respectively. ....................................................................................................... 12
Table 1b. Values the Young modulus E, Poisson ratio (), and density d for pure PEEK and MWCNTs predicted by
atomistic molecular dynamics (MD) simulations. ...................................................................................................... 15
Table 2b. Values the Young modulus E for PEEK/MWCNTs (nano)composites with random alignment of the MWCNTs
in the polymer matrix at different filler volume fractions as predicted from atomistic molecular dynamics (MD)
simulations, native MFH and calibrated MFH using MD data (see text for details). The last column shows the error
(%) of the calibrated Young modulus MFH predictions with respect to the corresponding MD-based values. ....... 16
Table 3b. Values the components of the Young modulus, of the Poisson ratio, of the shear modulus and density for a
PEEK/MWCNTs/CFs (nano)composite characterized by V f of 7% of MWCNTs randomly distributed within the PEEK
matrix and a Vf of 40% in CFs as predicted by the MFH model. ................................................................................ 18
Table 4b. Values the axial and transversal component of the Young modulus E for PEEK/MWCNTs (nano)composites
characterized by a perfect alignment of the MWCNTs in the polymer matrix as obtained from atomistic MD
simulations. ................................................................................................................................................................ 19
Table 5b. Parameters of the three constituents of the material model developed to simulate the double-reinforced
PEEK/MWCNTs/CFs with MWCNTs aligned in PEEK matrix along the CFs orientation. ............................................ 19
Table 6b. Values the axial and transversal components of the Young modulus for PEEK/MWCNTs (nano)composites
with perfect alignment of the MWCNTs in the polymer matrix at different filler volume fractions as predicted
from MFH calibrated using MD data (see text for details). The pure MD data values (Table 4) are also shown for
comparison The last columns show the error (%) of the calibrated Young modulus MFH predictions with respect
to the corresponding MD-based values. .................................................................................................................... 20
Table 7b. Values the components of the Young modulus, of the Poisson ratio, of the shear modulus and density for a
PEEK/MWCNTs/CFs (nano)composite characterized by V f of 7% of MWCNTs aligned within the PEEK matrix and a
Vf of 40% in CFs as predicted by the MFH model. ..................................................................................................... 21
Table 8b. Values the components of the Young modulus, of the Poisson ratio, of the shear modulus and density for a
PEEK/MWCNTs/CFs (nano)composite characterized by V f of 7% of MWCNTs with different orientation within the
PEEK matrix and a Vf of 40% in CFs as predicted by the MFH model. ....................................................................... 22
Table 1d. ABAQUS and MUL2 first buckling loads (N). ...................................................................................................... 26
Table 1f. Interface parameters from DPD back-mapped atomistic MD simulations of fully cured (70%) epoxy-resins
loaded with 7%wt of MWCNTs of different radii (see text and Figure 7f) for more details. .................................... 31
Table 2f. Amounts of reacted and unreacted DPD beads. Bead D is a di-functional bead. Data are obtained from
simulations in which each crosslinks form equal probability (=1). ............................................................................ 33
Table 1g. Values the components of the Young modulus, of the Poisson ratio, of the shear modulus and density for a
fully cross-linked (85%) DGEBA resin cured with EDA and reinforced with glass fibers at a volume fraction V f of
60% as predicted by the MFH model. ........................................................................................................................ 35
Table 2g. Values the components of the Young modulus, of the Poisson ratio, of the shear modulus and density for a
fully cross-linked (85%) DGEBA resin cured with EDTA and reinforced with glass fibers at a volume fraction V f of
60% as predicted by the MFH model. ........................................................................................................................ 35
Table 1i: Generated data at the macro scale (part level). ................................................................................................. 39
Table 1k. Parameters of the hyperelastic (Ogden) model and elastic model adopted for in the full filed homogenization
simulations of SBR/modified silica nanocomposites. ................................................................................................ 46
Table 1l. Generated KPIs for the tire user case. ................................................................................................................ 49

Deliverable D1.2: Material models, interfaces and uncertainty quantification 8


Project Reference: 721105

DELIVERABLE DESCRIPTION AS REPORTED IN THE DOA


This report is based on the results of the work conducted in WP1 (Modeling workflows) during
months M3-M30 for the task
Task 1.2: Development and testing of the elementary material models, their post-processing and
uncertainties
Leader: UniTS
Involved partners: e-Xstream, INSA, LIST, POLITO
This task will mainly aim at developing all the model enhancements required for the supporting
software to provide the necessary functionalities. Every model inputs and outputs will be
rationalized and standardized in order to ease their coupling across the scales (T1.3), with process
models (T1.6) as well as later their integration in the modeling platform via dedicated APIs (T1.7).
Elementary material model testing will be performed following the targeted use cases so that later
couplings will rely on individual models validated on a set of common data. The sensitivity of
model predictions with respect to variations of model inputs will be performed in order to prevent
the workflow to be built over, or at least to master, any high model sensitivities. These sensitivity
studies will also aim at preparing the set-up of methodologies in order to propagate material
uncertainties across the workflow. The uncertainties will be propagated using stochastic models
whenever possible or Monte-Carlo methods, supported by model reduction order to limit
computational time.

SUMMARY OF THE DELIVERABLE D1.2


This deliverable concerns the development and testing of elementary material models, their post-
processing and uncertainty quantification for the three cases studies of COMPOSELECTOR, that is:
▪ CASE STUDY 1: Material and process selection of fuselage thermoplastic frame (end-user:
Airbus)
▪ CASE STUDY 2: Material and process selection of leaf-spring (end-user: Dow)
▪ CASE STUDY 3: Material and process selection of passenger tire platform with fixed carcass
(end-user: Goodyear)
The specifications and requirements for all three case studies are described in details in D6.1,
whereas the MODA documents describing the complete modelling scenarios with the goal to
explain all modeling capabilities within the Consortium and their logical interconnections are
reported in D1.1.
The present report describes the different models developed by the Consortium for each case
study, the relevant properties obtained from their post-processing, and the related uncertainties
quantifications. In brief:
▪ CASE STUDY 1 (end-user: Airbus)
o UniTS developed atomistic, mesoscopic, combination of atomistic/mesoscopic material
models for end-user systems (e.g., PEEK/MWCNTs and PEEK/CFs);
o e-Xstream developed Mean Field Homogenization (MFH) algorithms for the prediction of
the mechanical behavior of the end-user systems (e.g., PEEK/MWCNTs reinforced with CFs)
parameterized using lower level simulations;

Deliverable D1.2: Material models, interfaces and uncertainty quantification 9


Project Reference: 721105

o INSA-Lyon developed enriched hyperelastic continuum models for bridging mesoscale to


the final scale of the engineering airplane fuselage;
o POLITO-MUL2 group developed the finite element model of the macroscale continuum
model with component-wise characteristics to embed the layer-wise material properties
from the mesoscale;
o LIST contributed to the definition and coordination of the modelling tasks.
▪ CASE STUDY 2 (end-user: Dow)
o UniTS developed atomistic, mesoscopic, combination of atomistic/mesoscopic material
models for end-user systems (e.g., EPON862/TETA/MWCNTs and DGEBA-based resins
cross-linked with different hardeners);
o e-Xstream developed Mean Field Homogenization (MFH) algorithms for the prediction of
the mechanical behavior of the end-user systems (e.g., DGEBA-based resins reinforced with
glass fibers) parameterized using lower level simulations;
o INSA-Lyon developed enriched hyperelastic continuum models for bridging mesoscale to
the final scale of the engineering leaf spring;
o LIST developed and implemented an integrated simulation approach, including micro and
macro scale continuum models. LIST Extended or developed interfaces for linking the
homogenization models and tools developed to the two-scale Finite Element Analyses.
Performances defined as KPIS in the D6.1 are targeted. Strength and Stiffness are
estimated.
▪ CASE STUDY 3 (end-user: Goodyear)
o UniTS developed atomistic, mesoscopic, combination of atomistic/mesoscopic material
models for end-user systems (e.g., SBR + silica NPs decorated with different surface
modifiers)
o e-Xtream developed Full and Mean Field Homogenization algorithm (MFH) for the
prediction of the mechanical behavior of the end-user systems (e.g., SBR + silica NPs
decorated with different surface modifiers at different silica content) with and without
parameterization from lower level simulations;
o LIST developed a numerical methodology for the estimation of tyre rolling resistance and
breaking force. Coupled micromechanical and full three-dimensional finite element
modelling of a visco-hyperelastic inflated rubber tyre under the deformation of vertical
loading is performed.
▪ UNCERTAINITY QUANTIFICATION
o POLITO-SMaLL GROUP developed an innovative approach by embedding molecular
dynamics codes into well-established software intended for performing sensitivity analysis
and UQ for continuum models.
o LIST investigated the coupling accuracy by conducting global sensitivity analysis on
different multi-scale models. Data-driven model that approximates the input/output
behavior of the numerical model is built by using the Sparse Polynomial Chaos Expansion
(SPCE) methodology. Then, Sobol’ indices that quantify the sensitivity of the input factors
are calculated analytically from the data-driven model with a negligible additional
Deliverable D1.2: Material models, interfaces and uncertainty quantification 10
Project Reference: 721105

computational cost. Interaction effects among different parameters are also captured.

CASE STUDY 1: MATERIAL AND PROCESS SELECTION OF FUSELAGE


THERMOPLASTIC FRAME (AIRBUS)
This user-case is focused on material and process selection for the replacement of conventional
airplane fuselage frame structure materials – currently Aluminum or thermoset composites – with
thermoplastic-based polymer composites, ultimately achieving a weight reduction of about 20%
while preserving outstanding mechanical properties. All relevant details and Key Performance
Indicators (KPIs) are described in details in Deliverable D6.1. In what follows a summary of all
activities performed for this user-case is reported, organized by the level of the simulation models
employed.

1.a. Atomistic and mesoscopic models (partner: UniTS)


In the framework of this case study, UniTS developed computational recipes based on different
material models to estimate sets of physical quantities (atomic and/or mesoscopic model results)
liked to user case KPIs and/or specifically requested by the end-user as a property. All
computational procedures are described in details in Laurini et al., Multimodel approach for
accurate determination of industry-driven properties for polymer nanocomposite materials,
Journal of Computational Science 26 (2018) 28-38 (https://doi.org/10.1016/j.jocs.2018.03.002). All
simulations performed together with the input data and the relevant results are stored in the
Granta database (GRANTI MI) system and can be retrieved from it. Below a synthesis of the main
results obtained from UniTS for this user-case is reported.
1. Material model(s): Atomistic (MD); Model system(s): a) PEEK/MWCNTs and b) PEEK/CFs;
Material property(ies): Tg and density as function of % of MWCNTs and temperature; KPI:
weight reduction/no weight increase.
Figure 1a shows some examples of the results obtained for model systems a) and b), respectively
and their comparison with experimental data, where available. As seen in this Figure, the
predicted glass transition temperature of the thermoplastic PEEK (Tg,sim (PEEK)= 414 ± 18 K, Tg,exp
(PEEK) = 416 K) is unaffected by the dispersion of either MWCNTs or CFs up to nanofiller weight
fraction of 15%, in agreement with experimental observations.

Figure 1a. (A) Specific volume Vsp of pure PEEK and of a PEEK/MWCNTs nanocomposite (6 %wt) as a function of
temperature T. Only these two curves are shown for clarity. (B) Comparison of simulated (patterned bars) and
experimental values (filled bars) of T g for pure PEEK and PEEK/MWCNTs nanocomposites as a function of MWCNTs
concentration (%wt). Experimental data from Sandler et al., Composites: Part A 33 (2002) 1033-1039; Deng et al.,
Comp. Sci. Technol. 67 (2007) 2959-2964; Ogasawara et al., Comp. Sci. Technol. 71 (2011) 73-78. (C) Simulated values
of Tg for pure PEEK and PEEK/CFs nanocomposites as a function of MWCNTs concentration (%wt).

Deliverable D1.2: Material models, interfaces and uncertainty quantification 11


Project Reference: 721105

2. Material model(s): Atomistic (MD) + mesoscopic (DPD); Model system(s): PEEK/MWCNTs;


Material property(ies): thermal conductivity as a function of % of MWCNTs KPI: no specific KPI
but requested by the end-user as a property.
Figure 2a shows some examples of the results obtained for the model system, from which it is
evident that the thermal conductivity component parallel to the alignment of the MWCNTs
undergoes the maximum enhancement as a function of filler content.

Figure 2a. (A) Thermal conductivity of random, and aligned MWCNTs in PEEK as a function of MWCNTs content (%wt).
(B) Thermal conductivity enhancement factor (EF = composite/matrix) for the systems in (A). Par = thermal conductivity
along the alignment of the MWCNTs; Perp = thermal conductivity perpendicular to the MWCNTs alignment.

3. Material model(s): Atomistic (MD); Model system(s): PEEK/MWCNTs; Material property(ies):


mechanical properties as function of % of MWCNTs; KPI: no change (worsening) in mechanical
properties.
Table 1a lists the principal mechanical properties for the model system, expressed as averages
from the values obtained by two different computational methodologies developed (i.e.,
unidirectional compression/tension and hydrostatic compression/tension simulations,
respectively). Figure 3a is a graphical representation of a subset of data from Table 1a. In
particular, Figure 3a(A) shows a comparison between the average calculated values of the Young
modulus (Table 1a) and corresponding experimental data (Deng et al., Comp. Sci. Technol. 67
(2007) 2959-2964). Simulations correctly capture an exponential increase of the Young modulus
with increasing nanofiller concentration in the composition interval considered. For these models,
the mean absolute error is 0.45 and the root-mean-square error is 0.70.

MWCNTs (%wt) C11,av (GPa) Kav (GPa)  av (-) Eav (GPa)


0 7.368 ± 0.592 5.542 ± 0.299 0.386 3.807
3 7.393 ± 0.477 5.872 ± 0.288 0.379 4.263
6 8.426 ± 0.501 6.079 ± 0.273 0.368 4.815
9 10.02 ± 0.599 7.283 ± 0.256 0.371 5.637
12 11.96 ± 0.548 8.892 ± 0.301 0.381 6.349
15 18.14 ± 0.621 13.85 ± 0.354 0.392 8.972

Table 1a. Values of the average elastic constant (C11,av) , average bulk modulus Kav, average Poisson ratio av, and
average Young modulus Eav for the pristine PEEK matrix and PEEK/MWCNTs nanocomposites as a function of the
nanofiller concentration as derived from unidirectional compression/tension and hydrostatic compression/tension
simulations, respectively.

Deliverable D1.2: Material models, interfaces and uncertainty quantification 12


Project Reference: 721105

Figure 3a. (A) Comparison between the average value of the Young modulus E, calculated using data in Table 1
(patterned bars), and the corresponding experimental measured data (filled bars) for pure PEEK and PEEK/MWCNTs
nanocomposites as a function of MWCNTs concentration (%wt). Experimental data from Deng et al., Comp. Sci.
Technol. 71 (2011) 73-78. (B) Average Poisson ratio ν values, calculated using data in Table 1a for pure PEEK and
PEEK/MWCNTs nanocomposites as a function of MWCNTs concentration (%wt).

4. Material model(s): Atomistic (MD); Model system(s): PEEK/CFs; Material property(ies):


mechanical properties as function of % of CFs; KPI: no change (worsening) in mechanical
properties.
Figure 4a shows results obtained for user-case systems PEKEK/CFs As seen in this Figure, the
predicted Young modulus of the PEEK matrix progressively increases with increasing CFs content
(Figure 4a(A)), reaching an increment of almost 50% in correspondence of the highest CFs weight
fraction considered (15%, Figure 4a(B)).

Figure 4a. (A) MD-predicted values of the Young modulus E for pure PEEK and PEEK/CFs nanocomposites as a function
of CFs concentration (%wt). (B) Young modulus enhancement factor EF (= Ecomposite/Ematrix) for the systems in panel A.

5. Material model(s): Atomistic (MD); Model system(s): PEEK/MWCNTs; Material property(ies):


interfacial polymer/filler cohesive and shear strength; KPI: no specific KPI but by the end-user
as a property correlated to material strength and processability.
Figure 5a(A) shows the behavior of the cohesive force Fcoh at the polymer/nanofiller interface as
the MWCNT is displaced away from the PEEK chains along the normal direction. The
corresponding profile of the pull-out (aka shear) force Fpo vs. displacement is reported in Figure
5a(B). Overall, these experiments support the idea that cohesive forces play a major role at
determining the interface properties in PEEK/MWCNTs nanocomposite systems.

Deliverable D1.2: Material models, interfaces and uncertainty quantification 13


Project Reference: 721105

Figure 5a. (A) Behavior of the cohesive force (Fcoh) at the interface of a PEEK/MWCNT system in response to nanotube
pulling along the z (normal) direction. The value of the maximum in the Fcoh/displacement curve is estimated to be
equal to 1131 pN at the corresponding displacement of 0.0019 nm. (B) The corresponding profile of the pullout force
Fpo vs. displacement obtained by pulling the MWCNT along the x direction. The maximum value of F po locates at a
displacement of 0.2, and is equal to 120 pN.

6. Material model(s): Atomistic (MD) + mesoscopic (DPD); Model system(s): PEEK/MWCNTs;


Material property(ies): viscoelastic storage and loss moduli; KPI: no specific KPI but by the end
user as properties correlated to processability.
As an example of the predicted viscoelastic behavior of the model system Figure 6a shows that,
while the simulated data for the pristine PEEK matrix follow linear viscoelastic scaling laws, for
MWCNT content of 3%wt onward this terminal behavior disappears, and both moduli tend toward
a plateau-like regime, indicative of a transition from liquid-like to solid-like behavior. According to
these data, a MWCNT concentration of 3%wt can constitute the so-called rheological percolation
threshold, i.e., the concentration at which a continuous, three-dimensional network of nanotubes
originates. Above this threshold, filler-filler interactions prevail over polymer-filler interactions,
and a 3D filler network is fully established.

Figure 6a. Predicted values of (A) Storage modulus G’(ω) and (B) loss modulus G”(ω) as a function of the applied
oscillatory shear frequency  of PEEK/MWCNT nanocomposite systems as a function of the MWCNT content (%wt).
Symbol legend: circles, pure PEEK; squares, 3%; triangles, 6%; diamonds; 9%; hexagons, 12%; inverted triangles, 15%.
Corresponding error bars are also displayed for all cases (in many instances, the errors are too small to be visible).
Simulated data agree with the experimental evidence provided for PEEK/MWCNT nanocomposites in a concentration
range encompassing the values considered in this work (Bangarusampath et al., Polymer 50 (2009) 5803-5811). For
such real systems, the behavior of G’(ω) and G”(ω) is in excellent qualitative agreement with the simulated data

Deliverable D1.2: Material models, interfaces and uncertainty quantification 14


Project Reference: 721105

reported in Figure 6a; the percolation threshold in that case was found to be equal to 1% wt, again in good agreement
with the one predicted by the present combination of MD + DPD simulations.

1.b. Mean field homogenization models (partner: e-Xtream)


Within the AIRBUS user-case, e-Xstream initially developed algorithms based on mean field theory
to compute the mechanical properties of the PEEK/MWCNT composites further reinforced with
CFs under different alignments of MWCNTs in the polymeric matrix (random, parallel and a set of
specifically defined orientations) and different compositions of both reinforcements (volume
fraction of 0, 2, 5, 7, 9 and 12% for MWCNTs and 40 and 60% for CFs, respectively). All simulations
performed together with the input data and the relevant results are stored in the Granta database
system (GRANTA MI) and can be retrieved from it. Below a synthesis of the main results obtained
from e-Xstream for this user-case is reported.
1. Material model(s): Mean field homogenization (MFH); Model system(s): a) PEEK/MWCNTs
reinforced with CFs with random orientation of MWCNTs in the PEEK matrix; Material
property(ies): axial and in-plane Young modulus, transverse and in-plane shear modulus,
transverse and in-plane Poisson ratio; KPI: weight reduction/no weight increase.
It is well known that the mean field theory is not able to accurately capture the effects of nano-
inclusion in a polymeric matrix (the so-called nano-effect). This is because the underlying
hypotheses in composite materials mean field theory neglect inclusions size while considering only
the shape and the content of the fillers. For this reason, in order to obtain accurate macroscopic
prediction of e.g., the mechanical properties of polymer-based (nano)composites lower level (e.g.,
atomistic or mesoscopic) simulation models must be adopted. However, the efforts (both in terms
of CPU and human time) required to get these results using atomistic (e.g., molecular dynamics) or
mesoscopic (e.g., dissipative particle dynamics) simulation is substantially larger than those
involved in MFH simulations. One way to achieve a compromise between time and accuracy in
accounting for effects exerted at the nano-level by the inclusions in polymer (nano)composite
systems on their macroscopic mechanical properties is to incorporate a limited and targeted set of
parameters, inferred from lower models (atomistic/mesoscopic) simulations, in the MFH algorithm
with the specific purpose of MFH model calibration.
The first user-case system considered – i.e., polymer (nano)composite system based on a random
dispersion of MWCNTs in a PEEK matrix further reinforced with CFs, illustrates the adoption of the
procedure described above. The first step of the computational pathway consisted in the
estimation, via atomistic molecular dynamics simulations performed by UniTS, of the values for
the Young modulus E, Poisson ratio , and density d for the pure PEEK matrix and MWCNTs,
respectively, as shown in Table 1b and Figure 1b.

Parameter PEEK MWCNTs


E 4.24 GPa 1.10 TPa
 0.38 0.34
d 1.321 g/cm3 1.760 g/cm3

Table 1b. Values the Young modulus E, Poisson ratio (), and density d for pure PEEK and MWCNTs predicted by

Deliverable D1.2: Material models, interfaces and uncertainty quantification 15


Project Reference: 721105

atomistic molecular dynamics (MD) simulations.

Figure 1b. Equilibrated MD simulation snapshot of the PEEK/MWCNTs used to estimate the values of the mechanical
quantities listed in Table 1b. The PEEK chains are shown as transparent spheres while MWCNTs are portrayed as dark
gray sticks.

For the construction of the MFH algorithm, both polymeric matrix and the MWCNT models were
defined as elastic isotropic. Specifically, the microstructure was defined as randomly distributed
ellipsoid inclusions with an aspect ratio of 2, as used in the corresponding MD simulations (Figure
1b).
Table 2b reports the results of the application of the MFH algorithm in terms of the predicted
Young modulus for the PEEK-MWCNT nanocomposites with random MWCNT alignment in the
PEEK matrix as a function (nano)filler volume fraction, and its comparison with the relevant
prediction from atomistic MD simulations.

MWCNTs volume Predicted E value Predicted E value Predicted E value (MD-calibrated Error
fraction (%) (MD) (MFH) (GPa) MFH) (GPa) (%)
0 4.24 4.24 4.24 -0.12

2 5.00 4.43 4.83 +3.24


5 5.72 4.73 5.74 -0.34
7 6.48 4.95 6.38 +1.66
9 7.10 5.17 7.03 +0.88
12 8.27 5.52 8.07 +2.39

Table 2b. Values the Young modulus E for PEEK/MWCNTs (nano)composites with random alignment of the MWCNTs
in the polymer matrix at different filler volume fractions as predicted from atomistic molecular dynamics (MD)
simulations, native MFH and calibrated MFH using MD data (see text for details). The last column shows the error (%)
of the calibrated Young modulus MFH predictions with respect to the corresponding MD-based values.

As seen from this table, for all MWCNT volume fraction considered the MFH always
underestimates the stiffness values of the PEEK-based composites with respect to those predicted
by atomistic MD simulations which, on the contrary, are in good agreement with the
corresponding experimental data reported in the preceding section. In order to improve the MFH
predictions, the model was then calibrated in order to match the value of the Young modulus
computed by MD simulations at the volume fraction of 5%. The calibration was performed by

Deliverable D1.2: Material models, interfaces and uncertainty quantification 16


Project Reference: 721105

parametrically varying the value of the MWCNT aspect ratio as in input parameter of the MFH
model, until the optimal value of 13.5 was found. Using this calibrated MFH model, the predicted
values of the Young modulus for all composite systems are very close to the ones calculated by
atomistic MD simulations, as shown in the last two columns in Table 2b.
Figure 2b reports a graphical comparison between the Young modulus prediction results obtained
from MD simulations and with the two (native and calibrated) MFH models. From this Figure we
can easily notice that the evolution of E as a function of the inclusion volume fraction is indeed
very similar to the one computed by MD simulations.

Figure 2b. Comparison of the values of the Young modulus E for PEEK/MWCNTs (nano)composites with random
alignment of the MWCNTs in the PEEK matrix as function of the volume fraction Vf as predicted by MD and MFH
simulations with and without calibration (data from Table 2b).

The main advantage of the calibrated-MFH methodology is that it relies on the provision of a
model able to predict the mechanical properties of a complex nanocomposite material at a very
low computational cost. Indeed, the CPU time to obtain a result using MFH models is around 1
second while more than several hours are required to get the same results using atomistic
molecular models (MD).
Therefore, in order to achieve the best results in terms of both time and accuracy for the
prediction of the mechanical properties of the end-user initial system (PEEK/MWCNTs composite)
it is suggested that 1 point (i.e., one system at a given volume fraction of inclusion) is calculated by
atomistic MD models and this point is then used to calibrate the MFH model. Next, the mechanical
properties of the full system set can be quickly and reliably calculated using MFH-based models.
This is a general conclusion, which, as will be seen in what is reported below, holds true for all
systems considered in this user-case and in the other user cases as well.
With the calibrated MFH model for the PEEK-MWCNTs (nano)composite at hand, the mechanical
properties of the complete end-user system – i.e., CFs-reinforced PEEK/MWCNTs with random
alignment of MWCNTs in the PEEK matrix were calculated. To the purpose, the material model
adopted considered the matrix and the CF as elastic isotropic models, as shown in Figure 3b.

Deliverable D1.2: Material models, interfaces and uncertainty quantification 17


Project Reference: 721105

Figure 3b. Representation of the elastic isotropic model used to simulate CF-reinforced PEEK/MWCNTs composites
with random alignment of MWCNTs within the polymer matrix. MWCNTs are shown as thin blue lines while the CFs
are shown as thick red cylinders. The PEEK matrix is not shown for clarity.

The input parameters required to define the elastic isotropic model of the matrix (PEEK/MWCNTs)
were taken from Table 2b, while those corresponding to the CFs were extracted from the e-
Xstream Digimat material database, and are as follows: E = 230 GPa;  = 0.20; d = 1.8 g/cm3.
As in the conceived model the CFs are taken to be aligned (Figure 3b), the mechanical model for
this system is transversely isotropic in the direction given by the CFs themselves. As an example of
the results obtained for this double-reinforced PEEK (nano)composite, Table 3b lists the values of
the mechanical properties for the system characterized by a Vf of 7% of MWCNTs randomly
distributed in the PEEK matrix and a Vf of 40% of CFs.

Property Value Property Value


Axial Young modulus 97.6 GPa Transverse Poisson ratio 0.28
In-plane Young 19.9 GPa In-plane shear modulus 6.83 GPa
modulus
In-plane Poisson ratio 0.46 Transverse shear modulus 7.54 GPa
Density 1.516 g/cm3

Table 3b. Values the components of the Young modulus, of the Poisson ratio, of the shear modulus and density for a
PEEK/MWCNTs/CFs (nano)composite characterized by Vf of 7% of MWCNTs randomly distributed within the PEEK
matrix and a Vf of 40% in CFs as predicted by the MFH model.
2. Material model(s): Mean field homogenization (MFH); Model system(s): a) PEEK/MWCNTs
reinforced with CFs with aligned orientation of MWCNTs in the PEEK matrix; Material
property(ies): axial and in-plane Young modulus, transverse and in-plane shear modulus,
transverse and in-plane Poisson ratio; KPI: weight reduction/no weight increase.
In order to study the effect of different orientations of the MWCNTs on the mechanical properties
of double-reinforced PEEK/MWCNTs/CFs (nano)composites, new models have been developed.
The first system considered was the one featuring all MWCNTs aligned in the same direction of the
CFs. As for the randomly distributed MWCNTs case, initially atomistic MD simulations were
performed by UniTS to compute the Young modulus of the PEEK matrix loaded with different V f of
aligned MWCNTs, as listed in Table 4b.

Deliverable D1.2: Material models, interfaces and uncertainty quantification 18


Project Reference: 721105

MWCNTs volume fraction (%) MD-predicted axial Young MD-predicted transverse Young
modulus (GPa) modulus (GPa)
2 5.42 4.61
5 6.23 5.32
7 7.06 6.05
9 7.37 6.56
12 8.99 7.87

Table 4b. Values the axial and transversal component of the Young modulus E for PEEK/MWCNTs (nano)composites
characterized by a perfect alignment of the MWCNTs in the polymer matrix as obtained from atomistic MD
simulations.

Next, again in analogy with the case of the randomly distributed PEEK/MWCNTs systems, a mean
field model was developed in order to fit the MD data. In this specific case, the mean field model
was constituted by three material components: the PEEK matrix, the MWCNTs represented again
as ellipsoidal inclusions, and a coating around each inclusion, as shown in Figure 4b.

Figure 4b. Mean field model of the MWCNT inclusions (dark blue) surrounded by the coating (lighter blue).

The input parameters of the corresponding MFH algorithm adopted for the are listed in Table 5b.

Parameter Value Parameter Value Parameter Value


Young 4.24 Young modulus 1.10 TPa Young modulus 15.0 GPa
modulus (MWCNT) (COATING)
(PEEK)
Poisson ratio 0.38 Poisson ratio 0.34 Poisson ratio 0.38
(PEEK) (MWCNT) (COATING)
Density (PEEK) 1.321 g/cm3 Density (MWCNT) 1.760 g/cm3 Density 1.321 g/cm3
(COATING)

Table 5b. Parameters of the three constituents of the material model developed to simulate the double-reinforced
PEEK/MWCNTs/CFs with MWCNTs aligned in PEEK matrix along the CFs orientation.

The aspect ratio of the inclusion was set to 2.1 and the relative thickness of the inclusion coating
was taken equal to 50%. The coating Young modulus and the inclusion aspect ratio values were
parametrically determined in order to calibrate the model to fit the axial and transversal Young
modulus values predicted by atomistic molecular dynamics simulations for the PEEK/MWCNTs
system with a Vf of 7% (Table 4b).
Deliverable D1.2: Material models, interfaces and uncertainty quantification 19
Project Reference: 721105

Table 6b shows the results obtained for the calibrated MFH-based prediction of the mechanical
properties of all PEEK/MWCNTs (nano)composites with perfect MWCNTs alignment in the
polymeric matrix. The corresponding values predicted by atomistic MD simulations (Table 4b) are
also shown for comparison, along with the error (%) of the calibrated MFH predictions with
respect to the corresponding MD-based values.

MWCNTs Calibrated MFH- Calibrated MFH- Error


MD-predicted MD-predicted Error axial
volume predicted axial predicted transversal
axial Young transversal Young Young modulus
fraction Young modulus transversal Young Young modulus
modulus (GPa) modulus (GPa) (%)
(%) (GPa) modulus (GPa) (%)
2 4.95 5.42 4.68 4.61 -9 2
5 6.09 6.23 5.40 5.32 -2 1
7 7.05 7.06 5.99 6.05 0 -1
9 7.95 7.37 6.56 6.56 8 0
12 9.83 8.99 7.74 7.87 9 -2

Table 6b. Values the axial and transversal components of the Young modulus for PEEK/MWCNTs (nano)composites
with perfect alignment of the MWCNTs in the polymer matrix at different filler volume fractions as predicted from
MFH calibrated using MD data (see text for details). The pure MD data values (Table 4) are also shown for comparison
The last columns show the error (%) of the calibrated Young modulus MFH predictions with respect to the
corresponding MD-based values.

It must be noted that, as well evidenced by Figure 5b, the relative error in the MFH-based
prediction is larger for the axial component of the Young modulus than for the transversal one.

Figure 5b. Comparison of the values of the axial and transversal Young modulus component E for PEEK/MWCNTs
(nano)composites with perfect alignment of the MWCNTs in the PEEK matrix as function of the volume fraction V f as
predicted by MD and MFH simulations with calibration (data from Table 6b).

With the calibrated MFH model for the PEEK/MWCNTs/CFs (nano)composite at hand, the
mechanical properties of the complete end-user system – i.e., CFs-reinforced PEEK/MWCNTs with
perfect alignment of MWCNTs in the PEEK matrix were calculated. To the purpose, the material
model adopted is transversely isotropic in the direction imposed by the MWCNTs, as shown in
Figure 6b.

Deliverable D1.2: Material models, interfaces and uncertainty quantification 20


Project Reference: 721105

Figure 6b. Representation of the elastic isotropic model used to simulate CF-reinforced PEEK/MWCNTs composites
with perfect alignment of MWCNTs within the polymer matrix. MWCNTs are shown as thin blue lines while the CFs are
shown as thick red cylinders. The PEEK matrix is not shown for clarity.

An example of the results obtained form the application of the calibrated-MFH model is shown in
Table 7b for the selected case of the PEEK/MWCNTs/CFs (nano)composite characterized by Vf of
7% of MWCNTs aligned within the PEEK matrix and a Vf of 40% in CFs.

Property Value Property Value


Axial Young modulus 7.05 GPa Transverse Poisson ratio 0.37
In-plane Young 5.99 GPa In-plane shear modulus 2.14 GPa
modulus
In-plane Poisson ratio 0.40 Transverse shear modulus 2.25 GPa
Density 1.352 g/cm3

Table 7b. Values the components of the Young modulus, of the Poisson ratio, of the shear modulus and density for a
PEEK/MWCNTs/CFs (nano)composite characterized by V f of 7% of MWCNTs aligned within the PEEK matrix and a Vf of
40% in CFs as predicted by the MFH model.

Finally, in order to investigate the impact of MWCNTs different alignments with respect to the CFs
axis, a further MFH model has been developed. The model considers the MWCNTs and the CFs to
be coplanar with a fixed angle between the two inclusion axes, as sketched in Figure 7b.

Figure 7b. Sketch of the model adopted in studying the effect of different MWCNTs orientation in PEEK/MWCNTs/CFs
Deliverable D1.2: Material models, interfaces and uncertainty quantification 21
Project Reference: 721105

(nano)composites using MFH algorithms.

Using the MFH-based results obtained for the system characterized by an aligned dispersion of
MWCNTS in the PEEK matrix, the mechanical properties of CF-reinforced PEEK/MWCNTs
(nano)composites have been predicted as a function of different orientation angles (10°, 20°, and
30°). As an example, Table 8b lists the values of the components of the Young modulus E, Poisson
ratio , shear modulus G and density d calculated for the system with Vf of 7% in MWCNTs and of
40% in CFs.

0° 10° 20° 30°


E11 96.84 GPa 96.82 GPa 96.75 GPa 96.63 GPa
E22 13.24 GPa 13.27 GPa 13.63 GPa 13.54 GPa
E33 13.24 GPa 13.24 GPa 13.29 GPa 13.35 GPa
12 0.295 0.296 0.298 0.300
21 0.0404 0.0406 0.0413 0.0420
13 0.295 0.296 0.293 0.292
31 0.0404 0.0406 0.0403 0.0403
23 0.519 0.519 0.518 0.517
32 0.519 0.519 0.515 0.510
G11 5.036 GPa 5.075 GPa 5.175 GPa 5.290 GPa
G22 5.036 GPa 5.075 GPa 5.006 GPa 4.973 GPa
G33 4.360 GPa 4.366 GPa 4.384 GPa 4.413 GPa
d 1.577 g/cm3 1.577 g/cm3 1.577 g/cm3 1.577 g/cm3

Table 8b. Values the components of the Young modulus, of the Poisson ratio, of the shear modulus and density for a
PEEK/MWCNTs/CFs (nano)composite characterized by V f of 7% of MWCNTs with different orientation within the PEEK
matrix and a Vf of 40% in CFs as predicted by the MFH model.

As seen from this Table, the orientation effect of MWCNTs up to relative angles of 30° on the
mechanical properties as predicted by MFH is very small. However, the model correctly captures
the progressive transformation of the system from transversely isotropic (angle between MWCNTs
and CFs = 0°, for which E22 = E33, 12 = 13, 21 = 31, 23 = 32, and G12 = G13, first column in Table
8b), to ortotropic, characterized by different values of all components of each mechanical
quantity.

1.c. Enriched hyperelastic continuum models (main partner: INSA-Lyon)


The work related to Task 1.2 performed at INSA-Lyon was focused on the development of
materials model enhancement for bridging the mesoscale to the final scale of the engineering
pieces considered in the targeted application cases. In particular, the materials which were studied
for the AIRBUS (and DOW, see next section) application cases are constituted by yarns which are
themselves composed by hundreds or thousands of fibers. The yarns then confer one part of the
final mechanical properties to the designed engineering pieces, the other part being provided by
suitably perfused resins. New enriched hyperelastic continuum models (second gradient models)
have been developed with the aim of finally providing a bridge between the mesoscale of the
Deliverable D1.2: Material models, interfaces and uncertainty quantification 22
Project Reference: 721105

yarns and the final scale of the engineering piece (bridging is described in detail in the deliverable
D1.4 associated to Task 1.4). The proposed enriched continuum model contains classical
information relative to the yarn elongation and shear stiffness and is enhanced by second gradient
terms which also allow to consider bending stiffness of the yarns at the scale of the engineering
piece.
For the specific AIRBUS user-case (but also for the DOW user-case, see section 1.h), the proposed
enriched models were used to perform the forming simulations of AIRBUS airframe using the code
COMSOL Multiphysics® (see Figure 1c).

Figure 1c. Forming of the AIRBUS airframe (left) and of the DOW leaf-spring (right) through the developed enriched
continuum (second gradient) model.

The enriched simulation outputs have been implemented via dedicated API in the MuPIF platform.
The connection of COMSOL® to MuPIF was not existing and was built ex-novo for integrating the
developed enriched models to the global workflow. In order to connect the enriched continuum
models implemented in COMSOL® with the MuPIF platform (Python language), it has been
necessary to go through one of the software that could be linked with COMSOL® through the
LiveLink connection (i.e., Excel or Matlab). In particular, LiveLink for Excel allows to take advantage
of the capabilities and structured simplicity offered by Excel to extend the COMSOL® modeling
capacity. Parameters and variables that are defined and modeled in COMSOL® are instantly
available in Microsoft Excel and automatically synchronized with the physics model. Moreover,
Excel can be connected with python with the Win32 Python package and can be used to activate
any functionality of COMSOL through LiveLink.
The enriched continuum model inputs and outputs have been rationalized and standardized for
implementation in the MuPIF, as shown in Figure 2c.

 For the API to work, Livelink for Excel must be installed and the reference to COMSOLCOM must be
activated (inside VBA for applications from the screen in Tools/References).

Deliverable D1.2: Material models, interfaces and uncertainty quantification 23


Project Reference: 721105

Figure 2c. Standardized enriched continuum model input/output definition for implementation in MuPIF platform.
Further technical details about the setting-up of enriched continuum models, as well as for their use in
the targeted application cases, are reported in Deliverable D1.4.

1.d. Finite-element (FE) models with enriched, variable kinematics (main partner:
POLITO-MUL2 group)
Handling of material models within the framework of the MUL2 code for the macroscale
continuum model of the AIRBUS component via the finite element (FE) approach, as exemplified in
Figure 1d.

Figure 1d. MUL2 FE model of the AIRBUS component.

FE models in MUL2 are component-wise (CW) and based on 1D and 2D elements with enriched,
variable kinematics. Although 1D or 2D, the constitutive material laws are 3D, i.e., in the case of
static analysis of structures composed of orthotropic materials, nine elastic coefficients are
necessary to characterize each ply. The use of such elements expands the modeling approaches of
commercial codes for enhanced geometrical and material modelizations. Concerning the AIRBUS
component, two approaches are of interest, namely, the Equivalent Single Layer (ESL) and Layer-
Wise (LW) ones. The former homogenizes the material properties of each ply and is the approach
commonly implemented in commercial codes. LW keeps the material properties of each ply
constant and homogenization acts at the interface only, as shown in Figure 2d. LW leads to higher
computational costs but improves the detection of transverse stresses; that is, ESL is a low-
medium fidelity approach whereas LW is medium-high.

Deliverable D1.2: Material models, interfaces and uncertainty quantification 24


Project Reference: 721105

Figure 2d. Example of LW assembly of FE arrays maintaining the geometrical and material properties of each ply.

Depending on the workflow architecture (Figure 3d), MUL2 has interfaces with the material
models provided by the COMPOSELECTOR partners.
MUL2 has interfaces with the material models provided by the COMPOSELECTOR partners. The
inputs for the MUL2 can come from homogenization model or from a database or user inputs. The
actual input source is defined by the particular simulation workflow.
In the MUL2 code, the 3D displacement field reduces to an axial displacement function 𝒖𝝉 (𝑦) and
a cross-section expansion function 𝐹τ (𝑥, 𝑧) along the x-z plane6. The generalized displacement field
is as follows:
𝒖(𝑥, 𝑦, 𝑧) = 𝐹𝜏 (𝑦, 𝑧)𝒖𝜏 (𝑥)
where the index τ corresponds to the expansion terms of the structural theory. Repeated indices
denote summation as per Einstein notation which means τ is summed from 1 to M expansion
functions. The choice of 𝐹τ defines the adopted structural theory and, in COMPOSELECTOR, MUL2
adopts Lagrange polynomials as expansion function (LE). The component-wise (CW) modeling
approach, an extension of LE models, allows modeling of multi-component structures through a
compact formulation and the assembly technique is in Fig. 2d. The standard finite element
approach discretizes the structure along the y-direction,
𝒖 = 𝐹𝜏 (𝑦, 𝑧)𝑁𝑖 (𝑥)𝒖𝜏𝑖
where 𝑁𝑖 is the ith shape function of order p, 𝒖𝜏𝑖 is the generic finite element nodal vector, and i is
summed from 1 to p+1. Using the principle of virtual work, the invariant form of the structural
stiffness matrix is formulated as:

𝛿𝐿𝑖𝑛𝑡 = 𝛿𝒖𝑗𝑠 ∫ 𝑁𝑗 (𝑥)𝐹𝑠 (𝑦, 𝑧)𝑫𝑇 𝑪𝑫 𝑁𝑖 (𝑥)𝐹𝜏 (𝑦, 𝑧)𝒖𝜏𝑖 𝑑𝑉 = 𝛿𝒖𝑗𝑠 𝒌𝑖𝑗𝜏𝑠 𝒖𝜏𝑖 1
𝑉

where 𝑫 is the differential operator matrix of size 6 x 3, 𝑪 is the constitutive material matrix and
𝒌𝑖𝑗𝜏𝑠 is the fundamental nucleus of the stiffness matrix of size 3 x 3, whose formal expression
remains the same irrespective of the choice of shape function and expansion function. As
mentioned above, the C matrix terms can stem from homogenizations, databases or user inputs
accordingly to the given workflow.
The verification of the MUL2 FE model was carried out by comparing the buckling loads provided
by the ABAQUS model built by AIRBUS, see Fig. 3d.

Deliverable D1.2: Material models, interfaces and uncertainty quantification 25


Project Reference: 721105

Figure 3d. MUL2 and ABAQUS FE models of the AIRBUS component.

For verification purposes, the structure was clamped at one end and kept free at the other and a
linearized buckling analysis was considered. Table 1d shows the numerical values of the first
buckling loads.

Figure 4d. ABAQUS and MUL2 first buckling mode.


Table 1d. ABAQUS and MUL2 first buckling loads (N).

Mode ABQ MUL2

1 6.15 6.18

The interface implementation exploits the MuPIF capabilities and GRANTA database. The MUL2
material modeling requires the nine elastic coefficients and the fiber orientation.

Figure 3d. MUL2 in the COMPOSELECTOR workflows.

Deliverable D1.2: Material models, interfaces and uncertainty quantification 26


Project Reference: 721105

CASE STUDY 2: MATERIAL AND PROCESS SELECTION OF LEAF-SPRING (DOW)


This user-case is focused on material and process selection for large production (30,000-50,000
parts/year) of car leaf-springs based on chemically diverse thermosetting resins loaded with
carbon-based nanofillers in place of the currently adopted (and substantially heavier) glass fibers,
while preserving optimal stiffness and mechanical strength for optimal passenger comfort. All
relevant details and KPIs are described in details in Deliverable D6.1. In what follows a summary of
all activities performed for this user-case is reported, organized by the level of the simulation
models employed.

1.f. Atomistic and mesoscopic models (partner: UniTS)


In the framework of this case study, UniTS developed computational recipes based on different
material models to estimate sets of physical quantities (atomic and/or mesoscopic model results)
linked to user case KPIs and/or specifically requested by the end-user as a property. A publication
reporting all computational details is in preparation. All simulations performed together with the
input data and the relevant results are stored in the Granta database system (GRANTA MI) and can
be retrieved from it. Below a synthesis of the main results obtained from UniTS for this user-case
is reported.
1. Material model(s): Atomistic (MD); Model system(s): EPON862+TETA; Material property(ies):
glass transition temperature Tg and density as function of degree of curing; KPI: Tg - no specific
KPI but requested as properties correlated to processability and mechanical performance; d –
weight reduction.
Figure 1f shows the MD predicted behavior of density and Tg for a pure thermosetting resin
composed by EPON862 (matrix) and TETA (hardener), as a function of curing degree. As seen from
this figure, a non-linear (i.e., exponential) increase in both properties, and in particular in Tg, is
anticipated with increasing degree of curing. It has to be remarked here that these data –
fundamental for the material processing fine tuning - are extremely difficult (if not impossible) to
determine experimentally and, even if so, the relevant costs (both in economical and resources
terms) will be exceedingly high.

Figure 1f. (A) Density as a function temperature for different degree of curing of neat epoxy. Error bars are omitted
for clarity. (B) Tg vs. degree of curing for neat epoxy. The numerical values at the different degree of curing (in
parenthesis) are Tg (K) = 379 (20%), 390 (40%), 409 (55%), 418 (70%), and 450 (85%).

2. Material model(s): Atomistic (MD); Model system(s): EPON862+TETA; Material property(ies):


mechanical properties as a function of degree of curing; KPI: no specific KPI but requested as
properties correlated to processability and mechanical performance.
Figure 2f(A) shows behavior of the Young modulus E as a function of temperature predicted by MD

Deliverable D1.2: Material models, interfaces and uncertainty quantification 27


Project Reference: 721105

simulations for a pure thermosetting resin composed by EPON862 (matrix) and TETA (hardener),
as a function of curing degree. As seen from this figure, there is a remarkable increase in E values
as the curing process time increases, with consequent increase of the relevant glass transition
temperature Tg (Figure 2f(B)), calculated from the midpoints of the corresponding E(T) curves in
panel A.

Figure 2f. (A) Young modulus as a function of temperature for different degree curing of neat epoxy resin. (B) T g vs.
degree of curing for neat epoxy resin obtained from the midpoints of the curves in panel A (patterned bars). The
corresponding Tg obtained from MD-predicted density vs. temperature curves (Figure 19) are also shown (filled bars)
for comparison. The numerical values at the different degree of curing (in parenthesis) are Tg (K) = 379 (20%), 390
(40%), 409 (55%), 418 (70%), and 450 (85%).

3. Material model(s): Atomistic (MD); Model system(s): EPON862+TETA+MWCNTs; Material


property(ies): stress-strain as a function of % of MWCNTs for fully cured (70%) resins; KPI:
optimal stiffness and mechanical strength.
Figure 3f(A) illustrates the MD-simulated stress-strain curves for the fully cured (70%) epoxy resin
per se and in the presence of increasing amounts of MWCNTs dispersed within the polymer
network. The model is able to capture the correct stress-strain behavior of thermoset-based
composites in that, while the pristine resin will likely undergo rupture only at very large applied
strains, this feature progressively decreases with increasing MWCNTs content. However, the
corresponding Young modulus values (Figure 3f(B)), calculated as the slope of the relevant curves
in Figure 21A, are greatly enhanced in the presence of increasing filler concentration.

Figure 3f. (A) Stress-strain behavior and (B) Young modulus of pure epoxy and epoxy/MWCNTs nanocomposites as a
function of MWCNTs concentration (% wt) as predicted by atomistic MD simulations.

4. Material model(s): Atomistic (MD); Model system(s): EPON862+TETA+MWCNTs; Material


property(ies): mechanical properties as a function of % of MWCNTs for fully cured (70%)
resins; KPI: optimal stiffness and mechanical strength.
Figure 4f shows the comparison of the Young modulus, tensile strength, and failure strain values
for the fully cured (70%) epoxy resin per se and in the presence of increasing amounts of MWCNTs
dispersed within the polymer network determined directly by MD-simulations and obtained with
experimental techniques. Notwithstanding the challenge in predicting such complex properties for
Deliverable D1.2: Material models, interfaces and uncertainty quantification 28
Project Reference: 721105

very complex systems such as those considered here at the atomistic level, the results are very
good. For this model, the mean absolute error (MAE) for E is 2.4 and the corresponding root-
mean-square error (RMSE) is 3.1; the MAE and RMSE for the tensile strength are 25 and 38,
respectively, while the MAE and RMSE for the failure strain are 0.078 and 0.091, respectively.

Figure 4f. (A) Comparison between simulated and experimental values of the Young modulus E for pure epoxy and
epoxy/MWCNTs as a function of MWCNTs concentration. (B) Comparison between simulated and experimental values
of the tensile strength for epoxy/MWCNTs as a function of MWCNTs concentration. (C) Comparison of failure strain
for epoxy/MWCNTs (circles) and PEEK/MWCNTs (squares) nanocomposite systems as a function of MWCNTs
concentration (% wt). Experimental data from Tsuda et al, Composites: Part A 65 (2014) 1–9.

5. Material model(s): Atomistic (MD) + mesoscopic (DPD); Model system(s): EPON862+TETA;


Material property(ies): mechanical properties as a function curing; KPI: optimal stiffness and
mechanical strength.
Figure 5f(A) shows an example of true stress-strain curves obtained using DPD back-mapped MD
simulations of pure epoxy resins at two different degrees of curing (20% and 70%, respectively). A
comparison of the Young modulus values obtained from data in panel A with those calculated via
direct MD simulations without DPD back-mapping (i.e., data in Figure 2f(A)) is shown in Figure
5f(B). As seen from this image, the results are very close one another for all degrees of curing
considered. Since the DPD back-mapped MD simulations are substantially more computationally
expensive than direct MD simulations, and the two techniques yield comparable results, the
former procedure is less convenient in the framework of business decision-driven material
selection and optimization of COMPOSELECTOR.

Figure 5f. (A) DPD back-mapped MD simulated true stress-strain curves of pristine epoxy resin at 20% and 70% degree
of curing. (B) Comparison between the Young modulus values for neat epoxy resin at different degree of curing
obtained from atomistic MD (filled bars) and DPD back-mapped MD simulations (patterned bars).

6. Material model(s): Atomistic (MD) + mesoscopic (DPD); Model system(s):


EPON862+TETA+MWCNTs; Material property(ies): mechanical properties as a function curing
at fixed content of MWCNTs (7%wt); KPI: optimal stiffness and mechanical strength.
Figure 6f(A) shows the results obtained from DPD simulations parameterized via atomistic MD
simulations for the Young modulus E of an epoxy-resin matrix loaded with a fixed content of
Deliverable D1.2: Material models, interfaces and uncertainty quantification 29
Project Reference: 721105

MWCNTs (7%wt) as a function of the resin degree of curing. The specific MWCNTs content was
selected in the perspective of achieving good mechanical performances with the least filler
loading. The results clearly shown that E exponentially increases with increasing curing degree,
and that a fully cured resin (85% curing) loaded with this specific amount of MWCNTs is
characterized by a Young modulus enhancement factor EF (defined in Figure 4b) of 7 (Figure 6f(B)).

Figure 6f. (A) Young modulus E for epoxy/MWCNTs composites at different degree of curing and at 7% wt nanofiller
concentration. (B) Young modulus enhancement factor EF for the systems in panel A.

7. Material model(s): Atomistic (MD) + mesoscopic (DPD); Model system(s):


EPON862+TETA+MWCNTs; Material property(ies): (nano)composite interface properties for a
fully cured (70%) epoxy-resin at fixed content of MWCNTs (7%wt) as a function of MWCNTs
radius; KPI: optimal stiffness and mechanical strength.
Among the critical factors that determines the ultimate, macroscopic mechanical performance of a
polymer-based (nano)composite, interface properties are of outmost importance. At the same
time, the experimental characterization and determination of interfacial properties is extremely
time and resource consuming. Thus, computational techniques constitute a strong alternative to
wet-laboratory investigations. In order to simulate interface properties of epoxy-resin/MWCNTs
nanocomposites a model has been developed as shown in Figure 7f.

Figure 7f. Model developed to simulate interface properties in epoxy-resin/MWCNTs nanocomposites (see text for
details).

The model has three components: the filler, the interface between the filler and the matrix, and
Deliverable D1.2: Material models, interfaces and uncertainty quantification 30
Project Reference: 721105

the matrix itself (Figure 7f, top left). Three main quantities are next defined: rparticle, which
represents the (nano)inclusion radius, tinterface, which is the interface thickness between the
(nano)particle and the matrix, and rmatrix = rparticle + tinterface (Figure 7f, bottom left).
Mesoscopic models of fully cured (70%) epoxy-resins loaded with 7%wt of MWCNTs of different
radii (18, 20 and 22 Å, respectively) were built and subjected to DPD simulations. Next, the
corresponding (nano)composite morphologies were back-mapped to atomistic models, which
were then subjected to atomistic molecular dynamics simulations. The results obtained are listed
in Table 1f.

rMWCNT (Å) E (GPa) rpeak (Å) tabsorption (Å) rmatrix (Å) tinterface (Å) Einterface (GPa)
18 19.95 23.60 6.16 29.36 11.76 39.10
20 18.30 25.01 6.08 31.09 11.09 23.97
22 17.20 26.02 6.31 32.33 10.33 11.52

Table 1f. Interface parameters from DPD back-mapped atomistic MD simulations of fully cured (70%) epoxy-resins
loaded with 7%wt of MWCNTs of different radii (see text and Figure 7f) for more details.

These results show very important information, which cannot be attained with real experiments.
Moreover, from the results in Table 7f it is readily seen that the MWCNT/matrix interface play a
dominant role in determining the macroscopic mechanical properties (e.g., the Young modulus E)
of the corresponding (nano)composites. Indeed, while the change in the overall E value changes
only slightly as a function of the MWCNT radius (Table 7f, first two columns), the latter has a major
influence on the interface Young modulus Einterface (last column, Table 7f), which pitfalls from 39.10
GPa to 11.52 GPa for a very small change in rMWCNT (22 Å and 18 Å, respectively). Also, quite
notably, Einterface is always considerably larger than the corresponding value of E; therefore,
neglecting the existence, influence and properties of the interface between a filler and the matrix
may results in exceedingly large error in the evaluation of the corresponding mechanical behavior
using e.g., continuum models.
These simulations also allowed to characterize other interface parameters for this systems, which
again can be of major help in modeling these (nano)composite with higher level models. These
further results can be obtained by analyzing the corresponding radial density distribution curves;
the left panel of Figure 7f shows the radial density distribution curves for the system with r MWCNT =
20 Å as an example. This, for instance, from the remaining values in Table 7f it can be seen that,
contrarily to Einterface, interface thickness tinterface is not drastically affected by the MWCNTs radius
variation. At the same time, the distance at which the density of the polymer surrounding the filler
reaches the value of the bulk matrix (rpeak) is decidedly more sensible to the MWCNTs dimensions,
ranging from 23.60 Å to 26.02 Å for r MWCNT equal to 18 and 22 Å, respectively.
8. Material model(s): Atomistic (MD) + mesoscopic (DPD); Model system(s): DGEBA+TETA;
Material property(ies): structural properties as a function of degree of curing ; KPI: no specific
KPI but by the end-user as a property.
The results from the first structural analysis for the pure epoxy-resin as a function of the curing
extent are shown in Figure 8f. In particular, the master curve for the cross-linking conversion in
Figure 8f(A) reveals that the resin structure does not depend on the rate of cross-linking but only
on the degree of curing. This is further supported by the corresponding radial distribution

Deliverable D1.2: Material models, interfaces and uncertainty quantification 31


Project Reference: 721105

functions shown in panel B of Figure 8f, which confirm that the resin local structure is the same,
independently of the values of the critical distance used in simulation for the crosslinking
formation criterion.

Figure 8f. (A) Master curve for the cross-linking conversion showing that the resin structure (final or intermediate)
does not depend on the rate of crosslinking, but only on the degree of curing. (B) The indistinguishable RDF curves
confirm that the local structure is the same for all rcross values at the end of the simulation. rcross = critical distance used
in the crosslinking formation criterion.

9. Material model(s): Atomistic (MD) + mesoscopic (DPD); Model system(s): DGEBA+TETA;


Material property(ies): structural analysis and curing kinetics; KPI: no specific KPI but
requested as properties correlated to processability.
The last step in this user-case was the formulation of a simulation procedure to mimic curing
kinetics and possibly derive some structural information along the curing process. To understand
the results obtained from the mesoscopic simulations performed with the parameters obtained
from simulations performed at a lower level (MD), Figure 9f shows the adopted DPD
representation of the resin components.

Figure 9f. DPD representation of DGEBA (left) and TETA (right) molecules.

Figure 10f illustrates the structural results of the DPD simulations during curing kinetics in terms of
density distribution of the different DPD bead types in the yz-plane along the x-direction.
Correspondingly, Table 2f lists the amount of reacted and unreacted DPD bead types.

Figure 10f. Bead density profiles averaged over time for the spatial direction, x, during DPD crosslinking simulation.

Deliverable D1.2: Material models, interfaces and uncertainty quantification 32


Project Reference: 721105

Bead densities have homogenous average density distribution in the yz-plane as a function of x. Profiles in the other
two directions are similar.
Bead type Reacted once Reacted twice Unreacted Total conversion
A 12267 - 2904 0.808
D 1721 4393 23 0.966
E 1956 - 1088 0.643

Table 2f. Amounts of reacted and unreacted DPD beads. Bead D is a di-functional bead. Data are obtained from
simulations in which each crosslinks form equal probability (=1).

Figure 11f finally show the DPD simulations results of curing kinetics for the DGEBA+TETA system
for two crosslinking conditions (i.e., fast (=1) and slow (=0.5)). The representation in % term is
adopted for comparison between bead types.

Figure 11f. Time evolution of reacted and unreacted beads for fast (1) and slow (0.5) crosslinking reactions
(probabilities of secondary amine-epoxy reaction = 1 and 0.5, respectively). % are shown for comparison between
bead types. Crosslinking reaction is switched-on at t (DPD) = 4000.

10. Material model(s): Atomistic (MD); Model system(s): DGEBA + different hardeners; Material
property(ies): density and glass transition temperature for fully cured (85%) DGEBA resins as
function of the hardener chemistry; KPI: no specific KPI but requested as properties correlated
to processability.
Figure 12f shows the chemical structures of the DDEBA resin monomer and of all different
hardeners considered, while Figure 13f presents the results obtained from atomistic MD
simulations for the behavior of density d and glass transition temperature T g for fully cross-linked
(85%) DGEBA resins cured with all different hardeners. From these results it is well evident that
the type of hardener has no influence on the system density (Figure 13f(A)) while it has a higher
effect on the process-related parameter Tg, the DGEBA/AEP and the DGEBA/IPD having the lowest
and highest Tg values, respectively (Figure 13f(B)).

Deliverable D1.2: Material models, interfaces and uncertainty quantification 33


Project Reference: 721105

HN
NH2
O O O O H2N N
NH2
DGEBA EDA AEP

H H
N N NH2
H2N NH2 H2N N
H

DETA TETA

H H H 2N
N N
H2N N NH2
H
TEPA NH2
IPD

Figure 12f. Chemical structure of DGEBA monomers and of all different hardeners considered.

Figure 13f. Density at 25°C (left) and Tg (right) values for fully cross-linked (85%) DGEBA resins cured with different
hardeners.

11. Material model(s): Atomistic (MD); Model system(s): DGEBA + different hardeners; Material
property(ies): mechanical properties for fully cured (85%) DGEBA resins as function of the
hardener chemistry; KPI: optimal stiffness and mechanical strength.
Figure 14f shows the results obtained from atomistic MD simulations for the mechanical
properties (i.e., Young modulus, stress and strain) of fully cross-linked DGEBA resins cured with
different hardeners (Figure 12f). From these results it can be seen that the different hardener type
has only limited influence on the mechanical behavior of the corresponding cured resin, the strain
being the parameter most sensitive to the hardener chemistry.

Figure 14f. Young modulus (left), stress (middle), and strain (right) values for fully cross-linked (85%) DGEBA cured
with different hardeners.

1.g. Mean field homogenization models (partner: e-Xstream)


In the framework of this case study, e-Xstream has developed a mean field model to predict the

Deliverable D1.2: Material models, interfaces and uncertainty quantification 34


Project Reference: 721105

mechanical behavior of two fully cross-linked (85%) DGEBA resins cured with two different
hardeners, i.e., EDA and EDTA (Figure 12f) and reinforced with glass fibers at a volume fraction V f
of 60%.
The input parameters for this MFH model relative to the different resins, i.e., the Young modulus,
Poisson ratio and density were taken from the relevant results produced by UniTS (see previous
section); the mechanical properties of the glass fibers were extracted from the e-Xstream Digimat
material database, and are as follows: E = 80 GPa;  = 0.22; d = 2.58 g/cm3.
In the construction of the MFH material model, all glass fibers were considered to be perfectly
aligned; accordingly, the corresponding material law is transversely isotropic.
Tables 1g and 2g list the values of the components of the Young modulus E, Poisson ratio , shear
modulus G and global system density for the two DGEBA-EDA and DGEBA-DETA resins,
respectively.

Property Value Property Value


Axial Young modulus 49.0 GPa Transverse Poisson ratio 0.26
In-plane Young 8.86 GPa In-plane shear modulus 3.05 GPa
modulus
In-plane Poisson ratio 0.45 Transverse shear modulus 3.47 GPa
Density 1.988 g/cm3

Table 1g. Values the components of the Young modulus, of the Poisson ratio, of the shear modulus and density for a
fully cross-linked (85%) DGEBA resin cured with EDA and reinforced with glass fibers at a volume fraction V f of 60% as
predicted by the MFH model.

Property Value Property Value


Axial Young modulus 48.97 GPa Transverse Poisson ratio 0.26
In-plane Young 8.24 GPa In-plane shear modulus 2.83 GPa
modulus
In-plane Poisson ratio 0.45 Transverse shear modulus 3.23 GPa
Density 1.988 g/cm3

Table 2g. Values the components of the Young modulus, of the Poisson ratio, of the shear modulus and density for a
fully cross-linked (85%) DGEBA resin cured with EDTA and reinforced with glass fibers at a volume fraction V f of 60% as
predicted by the MFH model.

1.h. Enriched hyperelastic continuum models (main partner: INSA-Lyon)


The enriched hyperelastic continuum model developed by INSA-Lyon for this user-case is
discussed together with the one proposed for the Airbus user-case (see Figure 1c in section 1.c).

1.i. LIST
A hierarchical, bottom-up approach was developed to carry out virtual tests of composite
materials and leaf spring structure. The overall multiscale simulation scheme is depicted in Figure

Deliverable D1.2: Material models, interfaces and uncertainty quantification 35


Project Reference: 721105

1-I and takes advantage of the fact that composite structures are made up of laminates which in
turn are obtained by stacking individual plies with different fiber orientation. This leads to three
different entities (ply, laminate, and component) whose mechanical behaviour is characterized by
three different length scales, namely fiber diameter, ply and laminate thickness, respectively. This
separation of length scales is useful to carry out multiscale modeling by computing the properties
of one entity (e.g., individual plies) at the relevant length scale, homogenizing the results into the
material relation, and passing this information to the simulations at the next length scale to
determine the mechanical behavior of the larger entity (e.g., laminate). Thus, multiscale modeling
is carried out through the transfer of information between different length scales rather than by
coupling different simulation techniques.

UD Laminate
[+45/-45/0/90]s
Molecular Scale
C

O O O O Macroscale:
Continuum Solid
Mechanics
UD Composite
Part Scale
(Ply Scale)

Figure 1-i: Multiscale simulation strategy: Hierarchical bottom modelling and simulation strategy. Fiber diameters are
of the order of 10 μm, while ply thicknesses are in the range 300 μm and the laminates are several mm in thickness
and above.

The estimation of the KPIs (tensile strength, bending stiffness and elongation to break) is carried
out in three successive steps within the framework of the finite element (FE) method. In the first
one, micromechanics is used to predict the ply properties. The properties of the constituents
(fiber, matrix and interfaces), together with the volume fraction and spatial distribution of the
fibers within an individual ply are provided by the industrial end-user (Dow chemical, NL). The
fiber properties (stiffness, strength, coefficients of thermal expansion) are provided by the fiber
manufacturer. Starting from the homogenized ply properties and information about the laminate
lay‐up as well as the interplay behavior, computational mesomechanics is then used to
determine the homogenized behavior of laminates. These results are finally used within the
framework of computational mechanics to obtain estimate the technical KPIs of the leaf-spring
component.

Deliverable D1.2: Material models, interfaces and uncertainty quantification 36


Project Reference: 721105

Unit-Cell Mesh

Properties Database
(Granta Mi)

Performance Estimation and Post-


processing

Post-processing
for KPIs

Figure 2-i: Multiscale simulation strategy: Hierarchical bottom modelling and simulation strategy. MuPIF workflow for
the Leaf Spring user case.

Both glass and carbon fiber options will be evaluated. Epoxy Resins are used for this application.
Four (4) formulations with well-known ingredients (molecular structure) and well-known
properties will be used:
1. Epoxy I: Longer Cure; Higher mechanical properties (Aromatic Amines); DER 330/331 epoxy
 + DETDA (Ethacure 100)
2. Epoxy II: Medium Cure, Medium mechanical properties (Cycloaliphatic Amines); DER
330/331 epoxy  + IPDA 30 min cure time
3. Epoxy III: Faster Cure: 15 min cure time; Lower mechanical properties (Aliphatic Amines);
DER 330/331 epoxy  + TETA/DETA - typical 60 min  + cure time
4. Hardener: is mainly used to cure the epoxy resin, which causes a chemical reaction without
changing its own composition. The curing time mainly depends on the hardener and epoxy
mixing ratio.
Formulations baseline: 1:1 (resin–hardener ratio) with possibility to go up down 10% at both sides.

Deliverable D1.2: Material models, interfaces and uncertainty quantification 37


Project Reference: 721105

Matrix (Resin) Fibre

Epoxy I Epoxy II Epoxy III Carbon I Carbon II Carbon II Glass I Glass II Glass III
EL (GPa) 4.00 3.50 2.50 241.00 231.00 230.00 72.00 73.10 85.00
ET (GPa) 4.00 3.50 2.50 30.00 28.00 15.00 72.00 73.10 85.00
GLT (GPa) 1.48 1.30 0.93 25.00 24.00 15.00 28.00 29.96 35.42
GTT (GPa) 1.48 1.30 0.93 12.00 10.70 7.00 28.00 29.96 35.42
nuLT 0.35 0.35 0.35 0.27 0.26 0.20 0.29 0.22 0.20
nuTT 0.35 0.35 0.35 0.25 0.31 0.07 0.29 0.22 0.20

Figure 3i: Inputs data used at the micro-mechanics scale. EL: Axial Young modulus, ET: In-plane Young modulus, GTT:
In-plane shear modulus, GLT: Transverse shear modulus, nuTT: In-plane Poisson ratio and nuLT: Transverse Poisson
ratio

UD's
Stiffness Max Princ Stress Density
Name (kN/mm) (MPa) (g/cm3) Mass (Kg)
Epoxy_1_Glass_1_vf_20 2,756 314,000 1,506 2,377
Epoxy_1_Glass_2_vf_20 2,787 313,000 1,504 2,374
Epoxy_1_Glass_3_vf_20 3,121 315,100 1,500 2,368
Epoxy_1_Carbon_1_vf_20 7,265 396,900 1,388 2,191
Epoxy_1_Carbon_2_vf_20 7,047 394,000 1,372 2,166
Epoxy_1_Carbon_3_vf_20 7,438 424,800 1,370 2,163
Epoxy_2_Glass_1_vf_20 2,665 315,700 1,498 2,365
Epoxy_2_Glass_2_vf_20 2,696 314,900 1,496 2,362
Epoxy_2_Glass_3_vf_20 3,023 318,100 1,492 2,355
Epoxy_2_Carbon_1_vf_20 6,972 411,000 1,380 2,178
Epoxy_2_Carbon_2_vf_20 6,764 406,200 1,364 2,153
Epoxy_2_Carbon_3_vf_20 7,085 430,800 1,362 2,150
Epoxy_3_Glass_1_vf_20 2,460 325,800 1,490 2,352
Epoxy_3_Glass_2_vf_20 2,489 326,900 1,488 2,349
Epoxy_3_Glass_3_vf_20 2,795 335,900 1,484 2,343
Epoxy_3_Carbon_1_vf_20 6,241 458,800 1,372 2,166
Epoxy_3_Carbon_2_vf_20 6,063 451,900 1,356 2,141
Epoxy_3_Carbon_3_vf_20 6,251 468,800 1,354 2,137

Deliverable D1.2: Material models, interfaces and uncertainty quantification 38


Project Reference: 721105

Epoxy_1_Glass_1_vf_30 3,764 315,500 1,614 2,548


Epoxy_1_Glass_2_vf_30 3,808 314,700 1,611 2,543
Epoxy_1_Glass_3_vf_30 4,295 320,700 1,605 2,534
Epoxy_1_Carbon_1_vf_30 10,110 432,700 1,437 2,268
Epoxy_1_Carbon_2_vf_30 9,802 430,000 1,413 2,231
Epoxy_1_Carbon_3_vf_30 10,230 482,500 1,410 2,226
Epoxy_2_Glass_1_vf_30 3,664 318,600 1,607 2,537
Epoxy_2_Glass_2_vf_30 3,708 319,500 1,604 2,532
Epoxy_2_Glass_3_vf_30 4,183 327,400 1,598 2,523
Epoxy_2_Carbon_1_vf_30 9,693 448,300 1,430 2,257
Epoxy_2_Carbon_2_vf_30 9,406 442,800 1,406 2,219
Epoxy_2_Carbon_3_vf_30 9,749 486,800 1,403 2,215
Epoxy_3_Glass_1_vf_30 3,424 337,000 1,600 2,526
Epoxy_3_Glass_2_vf_30 3,465 338,800 1,597 2,521
Epoxy_3_Glass_3_vf_30 3,904 350,400 1,591 2,512
Epoxy_3_Carbon_1_vf_30 8,640 504,600 1,423 2,246
Epoxy_3_Carbon_2_vf_30 8,401 496,600 1,399 2,208
Epoxy_3_Carbon_3_vf_30 8,586 525,500 1,396 2,204
Epoxy_1_Glass_1_vf_50 5,837 314,500 1,830 2,889
Epoxy_1_Glass_2_vf_50 5,909 315,500 1,825 2,881
Epoxy_1_Glass_3_vf_50 6,720 323,200 1,815 2,865
Epoxy_1_Carbon_1_vf_50 16,100 493,900 1,535 2,423
Epoxy_1_Carbon_2_vf_50 15,590 495,200 1,495 2,360
Epoxy_1_Carbon_3_vf_50 15,930 628,700 1,490 2,352
Epoxy_2_Glass_1_vf_50 5,726 320,000 1,825 2,881
Epoxy_2_Glass_2_vf_50 5,796 321,700 1,820 2,873
Epoxy_2_Glass_3_vf_50 6,586 330,500 1,810 2,857
Epoxy_2_Carbon_1_vf_50 15,480 496,600 1,530 2,415
Epoxy_2_Carbon_2_vf_50 15,010 496,200 1,490 2,352
Epoxy_2_Carbon_3_vf_50 15,260 612,300 1,485 2,344
Epoxy_3_Glass_1_vf_50 5,433 339,800 1,820 2,873
Epoxy_3_Glass_2_vf_50 5,498 342,400 1,815 2,865
Epoxy_3_Glass_3_vf_50 6,228 354,800 1,805 2,849
Epoxy_3_Carbon_1_vf_50 13,870 536,300 1,525 2,407
Epoxy_3_Carbon_2_vf_50 13,480 529,500 1,485 2,344
Epoxy_3_Carbon_3_vf_50 13,560 605,700 1,480 2,336
Table 1i: Generated data at the macro scale (part level).

Deliverable D1.2: Material models, interfaces and uncertainty quantification 39


Project Reference: 721105

Figure 4i: Generated data at the macro scale (part level).

The complete simulation of processing is a daunting task, as it has to consider mold filling, resin
flow through the stack of fiber fabrics as well as within each fiber bundle, generation and transfer
of heat due to the chemical reactions generated during curing, and the effect of the consolidation
pressure. Moreover, these problems are coupled because the heat generated by chemical
reactions accelerates the cross‐linking of the polymer network, dramatically increasing the resin
viscosity until the gel point is reached while the consolidation pressure deforms the fiber fabric
and changes its permeability.

Fibre Textile Preforming

Resin

Part
Infusion Curing
performance

Figure 5i: Processing workflow.

Deliverable D1.2: Material models, interfaces and uncertainty quantification 40


Project Reference: 721105

CASE STUDY 3: MATERIAL AND PROCESS SELECTION FOR CAR TYRE (GOODYEAR)
This user-case is focused on material and process selection for the formulation/optimization of
nano-rubber-based materials for automotive tire applications, with the ultimate aim of improving
tire rolling resistance, manufacturability, and durability. All relevant details and Key Performance
Indicators (KPIs) are described in details in Deliverable D6.1. In what follows a summary of all
activities performed for this user-case is reported, organized by the level of the simulation models
employed.

1.j. Atomistic and mesoscopic models (partner: UniTS)


In the framework of this case study, UniTS developed computational recipes based on different
material models to estimate sets of physical quantities (atomic and/or mesoscopic model results)
linked to user case KPIs and/or specifically requested by the end-user as a property. Two
publications reporting all computational details are in preparation. All simulations performed
together with the input data and the relevant results are stored in the Granta database system
(GRANTA MI) and can be retrieved from it. Below a synthesis of the main results obtained from
UniTS for this user-case is reported.
1. Material model(s): Atomistic (MD); Model system(s): SBR + SiO2 nanoparticles surface-
modified with different silane-based compatibilizers; Material property(ies): nanoparticle
wettability and work of adhesion maps; KPI(s): rolling resistance and manufacturability.
In producing polymer-based nanocomposites, one of the major obstacles is achieving a good
dispersion of the nanoparticles within a given polymeric matrix. This is due to several reasons,
among which one of the most preeminent is the chemico-physical incompatibility between the
filler and the polymer. To minimize this, a third component - often termed compatibilizer - is
usually added to the blend, which simultaneously interacts with both the nanofiller and the
polymer matrix, ultimately facilitating the blending process. Nanoparticle surface chemical
modification with a suitable compatibilizer molecule is by far the most followed route in polymer-
based nanotechnology. Choosing the right compatibilizer chemistry and the optimal compatibilizer
grafting density (i.e., the number of compatibilizer molecules per nm2 of nanoparticle surface,
another critical parameter in nanocomposite design) is all but a trivial process.
The specific user-case consists in styrene-butadiene rubber (SBR) reinforced with silica (SiO2)
nanoparticles. Dispersing pristine SiO2 in a SBR matrix is not possible, due to the opposite chemical
nature of the two blend components (SBR: organic, apolar; SiO2: inorganic, polar). This chemical
incompatibility translates into very poor wetting of the nanoparticles by the polymer and very low
adhesion energy between the two blend components. This, in turn, ultimately results in extensive
nanoparticle aggregation and quite modest system mechanical performance.
Car tire industry has developed, along the years and exclusively by trial and error methods, a
number of compatibilizers – substantially based on silane-based small molecules – to achieve good
SiO2 dispersion within SBR matrices. To provide the end-user with a reliable, fast and inexpensive
simulation-based procedure for this fundamental goal – i.e., choice of an ideal
compatibilizer/optimization of a compatibilizer already in use – resulting in a drastic decrease of
time and resources, we developed a computational model based on fundamental relationships of
wetting and adhesion theory. Specifically, for wetting, the underlying theory relies on the main
concept of contact angle , defined as shown in Equation 1j

Deliverable D1.2: Material models, interfaces and uncertainty quantification 41


Project Reference: 721105

𝑝 𝑝
√𝛾𝑃𝑑 +𝛾𝐹𝑑 √𝛾𝑃 +𝛾𝐹 𝑊
𝑐𝑜𝑠𝜃 = −1 + 2 +2 = −1 + 2 𝑊𝑃𝐹 1j
𝛾𝐹 𝛾𝐹 𝐹𝐹

in which 𝛾𝑓 is the surface tension of the filler, 𝛾𝑃𝑑 and 𝛾𝐹𝑑 are the dispersive components of the
surface tension for the polymer and the filler, 𝛾𝑃𝑝 and 𝛾𝐹𝑝 are the polar components of the surface
tension for the polymer and the filler, and 𝑊𝑃𝐹 and 𝑊𝐹𝐹 are the polymer-filler and filler-filler work
of adhesion values, respectively. As well known, the lower the value of the contact angle the
higher the filler wettability by the polymer and the better the filler dispersion within the polymeric
matrix (ideally, cos should be equal to 1, that is 0° is the condition of perfect wettability).
From the definition of  in Equation 1j, two conditions are immediately derived, as shown by
Equation 2j:
𝑊 𝑊𝑃𝐹
−1 + 2 𝑊𝑃𝐹 <1
𝐹𝐹 𝑊𝐹𝐹
𝑐𝑜𝑠𝜃 = { 𝑊𝑃𝐹 2j
1 ≥1
𝑊𝐹𝐹

From equation 2j, it is readily seen that, for 𝑊𝑃𝐹 /𝑊𝐹𝐹 < 1, nanoparticle-nanoparticle attraction is
higher than nanoparticle/polymer attraction; as a consequence, the nanoparticles will tend to
agglomerate in order to lower interfacial energy. On the contrary, for 𝑊𝑃𝐹 /𝑊𝐹𝐹 ≥ 1 nanoparticle-
polymer attraction is equal to or larger than nanoparticle-nanoparticle attraction; therefore, the
nanoparticles will be endowed with good initial dispersion within the polymer matrix.
As mentioned above, the driving force for flocculation/(re)agglomeration of nanoparticles in
polymeric matrices is determined by the change in potential energy when nanoparticle-
nanoparticle and polymer-polymer interfaces are created from two filler-polymer interfaces. This
potential energy change can be effectively captured by the change in the corresponding adhesion
energy (aka work of adhesion Δ𝑊𝑎 ), as shown by equation 3j:
𝑊𝑎 = 𝑊𝐹𝐹 + 𝑊𝑃𝑃 − 2𝑊𝑃𝐹 3j
in which 𝑊𝐹𝐹 , 𝑊𝑃𝑃 , and 𝑊𝑃𝐹 are the filler-filler, polymer-polymer, and polymer-filler work of
adhesion values, respectively.
It is simple to show that Equation 3j can be recast in the terms appearing in Equation 1j as shown
by Equation 4j:

Δ𝑊𝑎 = 2(𝛾𝐹𝑑 + 𝛾𝐹𝑝 ) + 2(𝛾𝑃𝑑 + 𝛾𝑃𝑝 ) − 2 (√𝛾𝑃𝑑 𝛾𝐹𝑑 + √𝛾𝑃𝑝 𝛾𝐹𝑝 ) 4j

From 4j it is simply concluded that the larger Δ𝑊𝑎 the larger the extent of
flocculation/(re)agglomeration of the nanoparticles within the polymeric matrix.
With these Equations at hand, a computational recipe for the estimation of the surface energy and
surface energy components of filler and polymer was developed based on atomistic MD
simulations. This was applied to the user-case, for the specific development of wetting-adhesion
maps for SBR of a specific composition (26% styrene/74% butadiene) loaded with pristine and
surface-modified SiO2 with different silane-based compatibilizers at both constant and different
grafting density.
Figure 1j summarized the results achieved for a selected set of SiO 2 surface-modifiers in the form
of the corresponding wetting-adhesion map.

Deliverable D1.2: Material models, interfaces and uncertainty quantification 42


Project Reference: 721105

Figure 1j. Wetting-adhesion map for pristine SiO2 and SiO2 nanoparticles surface-modified with different silane-based
compatibilizers in a SBR matrix of fixed composition (26%S/74%B). Solid colored lines:  isocontours; red broken lines,
Δ𝑊𝑎 ), isocontours (see text for more details).

The analysis of Figure 1j is straightforward: to achieve a good nanoparticle dispersion within the
polymeric matrix the points on this graph (i.e., the surface energy of the filler plotted vs. its
dispersive energy component) must locate in the areas enclosed by the isocontours corresponding
to the lowest values for both  and Δ𝑊𝑎 . Accordingly, from this Figure it is clearly seen that,
among all systems considered, 5%wt SiO2 nanoparticles modified with ODTES (aka OCTEO) with a
grafting density (gd) of 1 ODTES chain/nm2 can be best dispersed in the SBR matrix, as the
corresponding value follows in the areas enclosed by the lowest isocontour values of both  and
Δ𝑊𝑎 (filled black triangle in Figure 1j). By changing silane chemistry while maintaining dg and
nanoparticle loading constant, the dispersion of the relevant nanoparticles in the SBR matrix
becomes gradually worse in the order CPDMES (black filled diamond), APDMES (black filled
square), MTMS (black filled dots), and unmodified SiO2 (unfilled dot). It is also particularly
important to observe that the computational methodology is able to capture a subtler effect, that
is the influence of the grafting density for a given compatibilizer chemistry. Thus, as seen again in
Figure 1j, a remarkable improvement is predicted in the dispersion state of SiO 2 modified by
MTMS by increasing the corresponding silane dg (black filled dots) with respect to unmodified
silica nanoparticles. Also, for this specific compatibilizer the method predicts an interesting feature,
that is, the dispersibility of the corresponding system as a function of the compatibilizer dg
increases up to 0.3 silane chains/nm2 while a further increase in dg (0.4 silane chains/nm2) leads to
a worsening in the dispersion properties of the resulting nanocomposite.
Since the developed method is based on atomistic MD simulations, it can be applied to predict the
initial dispersion of any type of nanoparticle modified by any compatibilizer at each desired dg in
any polymer matrix. As such, it constitutes the first example of versatile in silico material selection
based entirely on computational material models.
2. Material model(s): Atomistic (MD) + mesoscopic (DPD); Model system(s): SBR + SiO2
nanoparticles surface-modified with different silane-based compatibilizers; Material
property(ies): dispersion morphology; KPI(s): rolling resistance and manufacturability.
The morphology of the dispersion of silane-modified SiO2 nanoparticles within a SBR matrix of
fixed composition (26%S/74%B) at two user-defined nanoparticle loadings (8% and 20% by weight)

Deliverable D1.2: Material models, interfaces and uncertainty quantification 43


Project Reference: 721105

and at different dg were predicted using mesoscopic models parameterized via atomistic MD
simulations.
Figure 2j shows, by way of example, the predicted morphologies of the dispersion of the OCTEO-
surface-modified SiO2 (dg = 1chain/nm2) at two different filler loadings (8%wt and 20%wt,
respectively), and their comparison with the corresponding experimental transmission electron
microscopy (TEM) images.

Figure 2j. Morphologies of the dispersion of SiO2 nanoparticles surface-modified with OCTEO (dg = 1 chain/nm2) as
obtained from TEM images (left) and from MD-parameterized DPD simulations (right) at two different nanoparticle
loading: 8%wt (top panel), and 20%wt (bottom panel). In the simulation images, the polymer matrix is depicted as
green (top) and gray (bottom) sticks while the modified SiO2 nanoparticles are shown in purple and orange sphere
representations, respectively. The area enclosed in the red boxes on the right images corresponds to the size of the
simulation boxes on the left.

As seen from this Figure, the different experimental aggregation states and morphologies of the
systems considered are well captured by the corresponding simulations at both nanoparticle
loading considered.
3. Material model(s): Atomistic (MD) + mesoscopic (DPD); Model system(s): SBR + SiO2
nanoparticles surface-modified with different silane-based compatibilizers; Material
property(ies): viscoelastic storage and loss moduli; KPI(s): rolling resistance and
manufacturability.
As an example of the predicted viscoelastic properties of the user-case systems, Figure 3j shows
the behavior of the elastic and loss moduli as a function of the applied oscillatory frequency for
the SBR matrix loaded with different OCTEO-modified SiO2 (dg = 1 chain/nm2). As seen from this
Figure, the pure SBR matrix is characterized by a viscoelastic behavior prototypical of polymer
systems, in that at lower oscillatory strains the loss modulus G” is larger than the elastic
component G’ until a cross-over is observed, after which the elastic behavior prevails. Loading the
polymer matrix with 8%wt silica-based nanofillers still exhibits the main features of a polymeric
solution although the cross-over frequency is shifter to lower values. Finally, at the higher filler
Deliverable D1.2: Material models, interfaces and uncertainty quantification 44
Project Reference: 721105

loading considered (20%wt), the elastic response of the system is always larger than the viscous
one, and both are almost parallel and independent of the applied oscillation frequency. This is
indicative of a transition from liquid-like to solid-like behavior. According to these data, the
modified SBR/SiO2 nanocomposite at 20% silica loading is above the so-called rheological
percolation threshold, i.e., the concentration at which a continuous, three-dimensional network of
silica nanoparticles originates. Thus, for this system, filler-filler interactions prevail over polymer-
filler interactions, and a 3D filler network is fully established, as verified by the morphological
simulations reported in the previous section.

Figure 3j. Predicted values of the storage modulus G’(ω) and loss modulus G”(ω) for a SBR matrix loaded with OCTEO-
modified (dg = 1 chain/nm2) silica particles of PEEK/MWCNT nanocomposite systems as a function of the modified SiO2
content (%wt).

1.k. Full field homegenization models (partner: e-Xstream)


In the framework of this case study, with the purpose of predicting the effect of modified silica
nanoparticle dispersions in SBR matrices on the mechanical behavior of the corresponding
nanocomposites, e-Xstream developed a full field homogenization model using the results of the
relevant mesoscopic simulations provided by UniTS. For the specific purposes, UniTS performed
DPD simulations with the following specifics:
SBR composition = 30S/70B;
silica surface modifier = OCTEO;
silica surface modifier grafting density = 0.5 chain/nm2;
modified silica loading in the SBR matrix = 10% and 20% by weight;
simulation cell geometry and dimension = cubic, 116 nm.
To develop the relevant continuum model, the morphology predicted at the DPD level were
mapped onto the corresponding representative geometry, as shown in Figure 1k.

Deliverable D1.2: Material models, interfaces and uncertainty quantification 45


Project Reference: 721105

Figure 1k. Continuum model representation of the dispersion of 10%wt OCTEO –modified silica nanoparticles in the
SBR matrix as mapped from the corresponding DPD simulation.

The dispersion of the silica particles can be represented by the distribution of the distance
between the nearest neighbors, as shown in Figure 2k for the 10% SBR modified-silica loaded
system. As seen from this Figure, this diagram features a maximum around 0.08.

Figure 2k. Nearest neighbor distance distribution for the dispersion of 10%wt of OCTEO-modified silica nanoparticles
in the SBR matrix. Normalized size: 1 unit = 116 nm; radius of the silica particle = 0.043 (corresponding to a real value
of 5 nm, according to the experimentally employed nanoparticles.)

In order to perform a full field homogenization simulation, the hyperplastic material model (aka
the Ogden model) was adopted for the rubber matrix and an elastic model was adopted for the
silica nanofiller. The relevant parameters are shown in Table 1k.

Parameters Values
Moduli = 0.0874; 1.18×10-7; 0.47 (MPa)
Hyperlastic Ogden (SBR)
Exponents = 3.25; 11.28; 0.74
Density (SBR) 1.1 g/cm3
Young modulus (SiO2) 201 GPa
Poisson ratio (SiO2) 0.3
Density (SiO2) 1.0 g/cm3

Table 1k. Parameters of the hyperelastic (Ogden) model and elastic model adopted for in the full filed homogenization
Deliverable D1.2: Material models, interfaces and uncertainty quantification 46
Project Reference: 721105

simulations of SBR/modified silica nanocomposites.

A full field homogenization was performed in order to obtain the stress-strain behavior at the
composite level. Three different mechanical loadings were used:
1. Pure loading: a quasi-steady macroscopic uniaxial state in the 1-direction;
2. Biaxial loading: a quasi-steady macroscopic biaxial state in the 1- and 2-direction;
3. Planar loading: a quasi-steady macroscopic biaxial state in the 1-direction and 0 strain state in
the 2-direction.
Finally, a volumetric mean was carried out on the simulation results for the three loadings (Figure
3k), to compute the macroscopic (i.e., composite level) stress-strain (large deformation) curves
along the 1-direction, as shown in Figure 4k.

Figure 3k. Uniaxial (left), biaxial (middle), and planar (right) loading results for the dispersion of 10%wt of OCTEO-
modified silica nanoparticles in the SBR matrix.

Figure 4k. Composite stress-strain (large deformation) curves for the dispersion of 10%wt of OCTEO-modified silica
nanoparticles in the SBR matrix.

A mean field homogenization (MFH) model was also developed to compute the macroscopic
mechanical response of a SBR loaded with modified silica nanoparticles. For this model, the
material law for the two phases (polymer and filler) are taken to be identical to those employed in
the full field model (Table 1k). At variance, in the MFH model the composite microstructure is not
mapped from the corresponding DPD simulations but is constituted by a random distribution of
spherical inclusion within the rubber matrix. With this MFH model, the effect of a silica content
ranging from 40 to 70% by volume was computed for a quasi-steady axial loading in the 1-
direction. The relevant results are shown in Figure 5k.

Deliverable D1.2: Material models, interfaces and uncertainty quantification 47


Project Reference: 721105

Figure 5k. Effect of silica content on the SBR-composite deformation gradient-stress curves (1-direction) for quasi-
steady axial loading in the 1-direction.

It has to be noted that the two methods (full field homogenization and mean field homogenization)
involve very different computation times, in that while several CPU hours are required by the
former, only ≈1 s is needed to compute the composite mechanical behavior using latter model.

1.l. LIST
One objective of this user case is to predict the footprint and the breaking force at the tire level
(macro-scale) using the data generated from the low scales (atomistic and micro scales). To this
end, a simplified finite element model of the tire is developed. The stress-strain curves generated
by e-Xstream are used as inputs for this macro-scale tire model. We developed an inverse
identification approach to estimate the parameters of the Ogden hyperelastic constitutive law
from the stress-strain curves computed by e-Xstream’s Digimat FE/MF. The Odgen hyperelastic
laws were generated for three volume fractions of SiO2 in SBR (20%, 30% and 40%). The
commercial finite element code ABAQUS is used to solve the macro-scale problem with the
appropriate loading and boundary conditions. The two KPIs are then estimated by post-processing
the output data from ABAQUS. Figure 1l shows the multiscale modeling strategy used here to
estimate the two KPIs requested at the tire scale. Table 1l gives both KPIs computed for the three
SiO2 volume contents.

MR: Hyperelastic MR

Breaking Force &


Rolling Resistance

29

Deliverable D1.2: Material models, interfaces and uncertainty quantification 48


Project Reference: 721105

Figure 1l. Multiscale simulation workflow for the tire model.

SiO2 %vol in SBR 20 30 40


Footprint (mm2) 3473.11 3158.64 2804.82
Breaking force (N) 1.72 1.68 1.54
Table 1l. Generated KPIs for the tire user case.

EMBEDDING/WRAPPING METHODOLOGY FOR PERFORMING MATERIAL MODELS


SENSITIVITY ANALYSIS AND UNCERTAINTY QUANTIFICATION (UQ)
Within the COMPOSELECTOR project, Task 1.2 is also devoted to explore automatic tools for
investigating the sensitivity of model predictions with respect to variations of model inputs, for all
model types. These sensitivity studies are essential (i) for preparing the set-up of methodologies in
order to propagate material uncertainties across the workflow and also (ii) for assessing the
suitability of coupling/linking methodologies (see Task 1.3 and the related Deliverable D1.3).

1.m Sensitivity analysis and uncertainty quantification of molecular dynamics data


(partner: POLITO-SMaLL group)
The sensitivity analysis and the uncertainty quantification (UQ) has been applied by POLITO-SMaLL
group to molecular dynamics (MD) simulations. To date, there are no native features in MD
software for performing automatic sensitivity analysis and UQ. Hence an innovative methodology
has been developed within COMPOSELECTOR by embedding MD codes into well-established
software intended for performing sensitivity analysis and UQ for continuum models. This
embedding/wrapping methodology is general, but, for the sake of concreteness, in what follows
two software will be employed in a proof-of-principle study.
As an example of software for automatic sensitivity analysis and UQ, the ModeFRONTIER platform
was selected. ModeFRONTIER is a Multidisciplinary Design Optimization (MDO) platform
developed by ESTECO spa, a COMPOSELECTOR project partner. Like other schedulers, the
interface is a graphical block diagramming tool, where the user is able to build the simulation
workflow by using blocks and links and to define the input and output variable to manage.
Because it is designed natively for automotive and aerospace market, the program offers many
tools capable to manage CAD and CAE programs, but thanks the possibility to use script program
(bash, Matlab, Python, etc..) and automatic handling and extracting data from files, it is possible to
extend its applicability field. In order to prove the feasibility of the newly proposed
embedding/wrapping methodology, a workflow able to completely govern an MD simulation,
using only bash scripts and a python script for the post-processing, was built. First, a sub-system
that only runs the MD simulation locally or remotely (Figure 1l(B) and 1l(C)) was built and studied.
This allows the multiple re-use of the sub-system in the workflow and defines the phases of the
simulation from the energy minimization to the post-processing (Figure 1l(A)). The result is a quite
complex diagram that allows a deep control. An alternative would be to use a script that wraps all
the simulation phases in only one node that receives input and gives output variables: this would
lead to a simpler workflow, but with less flexibility.
Deliverable D1.2: Material models, interfaces and uncertainty quantification 49
Project Reference: 721105

Figure 1m. (A) Workflow overview, I: Initial configuration, topology and result folder setup, II: energy minimization, III:
NVT equilibration, IV: NPT equilibration, V: 1st MD run: SDS adhesion, VI: 2nd MD run: SDS push, VII: constraints
checking and post-processing. (B) standalone MD simulation subsystem used in stages II and III. (C) continued MD
simulation subsystem used in stages IV, V and VI.

Once the workflow is built, exploded or condensed, all the set of tools given by modeFRONTIER
can be used; one of such capabilities is the Multi-Objective Robust Design Optimization (MORDO),
where it is possible to perform a UQ study to measure the reliability of a certain design (set of
variables). The use of this tool is quite simple and it consists of three steps:
• It starts with the definition of the input variables and their probability distribution;
• The second step is to decide the sampling algorithm between Monte Carlo (MC) or Latin
Hypercube;
• Then a decision must be taken on how to measure numerically the mean and the standard
deviation. The default option is the Polynomial Chaos Expansion (PCE), which show faster
convergence and better accuracy respect the classic MC ; moreover with this technique is
possible to produce a surrogate model and compute the cumulative distribution function. As a
drawback, it can be numerically expensive if many random variables must be managed. In the
latter case, the alternative is to use the Adaptive Sparse PCE, which focus only on the most
significant terms or disabling it and use only the MC (brute force method);
• Finally, modeFRONTIER has also the Design Space environment where is possible to analyze
the results. For example, by means of the Distribution Fitting tool is possible to identify the
output distribution at a glance by plotting the histogram, the cumulative distribution, quantile-
quantile plot, and Kolmogorov Smirnov test.

Deliverable D1.2: Material models, interfaces and uncertainty quantification 50


Project Reference: 721105

Even though the previous description refers only to a possible implementation of the developed
embedding/wrapping methodology, it is quite interesting because it demonstrates that it is
feasible to embed a standard MD software into UQ software, which were originally intended for
continuum models only. This is an original outcome of the COMPOSELECTOR project, which may
have a positive impact on the molecular modeling community who do not have access to
automatic UQ tools so far.

1.n. Global Sensitivity Analysis (partner: LIST)


Global Sensitivity Analysis is performed to explore the quantitative influence of the parameters on
multi-scale modeling accuracy. Sobol’ indices involving Monte-Carlo methods and metamodeling
methods are used. We build up a data-driven model that approximates the input/output behavior
of the original numerical model. Then, Sobol’ indices can be calculated analytically from the data-
driven model with a negligible additional computational cost. To “train” the data-driven model,
the Sparse Polynomial Chaos Expansion methodology, which uses orthogonal polynomials to
approximate the response surface. An algorithm is proposed to approximate the data-driven
model with SPCE. This algorithm is based on Bayesian model averaging. This strategy can be
referred as a data-driven approach since the settings of input parameters are explored by using a
large number of numerical experiments instead of intuition or personal experience.

Deliverable D1.2: Material models, interfaces and uncertainty quantification 51

You might also like