molecular modeling of structural transformations in ionomer solutions and membranes

155
Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes by Mahdi Ghelichi Ghalacheh M.Sc., University of Tehran, 2011 Thesis Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy in the Department of Chemistry Faculty of Science © Mahdi Ghelichi Ghalacheh 2016 SIMON FRASER UNIVERSITY Summer 2016

Upload: others

Post on 11-Sep-2021

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

Molecular Modeling of Structural Transformations in Ionomer Solutions and

Membranes

by Mahdi Ghelichi Ghalacheh M.Sc., University of Tehran, 2011

Thesis Submitted in Partial Fulfillment of the

Requirements for the Degree of

Doctor of Philosophy

in the

Department of Chemistry

Faculty of Science

© Mahdi Ghelichi Ghalacheh 2016

SIMON FRASER UNIVERSITY

Summer 2016

Page 2: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

ii

Approval

Name: Mahdi Ghelichi Ghalacheh Degree: Doctor of Philosophy Title: Molecular Modeling of Structural Transformations in

Ionomer Solutions and Membranes Examining Committee: Chair: Dr. Krzysztof Starosta

Associate Professor

Michael Eikerling Senior Supervisor Professor

Steven Holdcroft Supervisor Professor

Barbara Frisken Supervisor Professor

Dr. Gary W. Leach Internal Examiner Associate Professor

Dr. Armand Soldera External Examiner Professor Department of Chemistry University of Sherbrooke

Date Defended/Approved: August 4, 2016

Page 3: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

III III

Abstract

Different functions are expected from the polymer electrolyte membranes used in

fuel cells. They work as a proton conduction medium, as a separator, and as an

electronic insulator. The current membrane materials of choice are perfluorosulfonic acid

(PFSA) ionomers such as Nafion. The two main challenges that PFSAs still face, after

three decades of extensive research, are a limited lifetime and a lack of basic structural

understanding.

To investigate the chemical degradation phenomena, we devised a kinetic model

of radical formation and attack to PFSA ionomers. Analytical relations are derived to

obtain the content of aggressive radicals as a function of iron ion content and hydrogen

peroxide. The mean-field type, coarse-grained ionomer model distinguishes ionomer

headgroups, side chains, and ionomer backbone. The model is used to study the impact

of different degradation mechanisms and ionomer chemistries on PEM degradation.

Application of the model to degradation data of various PFSAs highlights the important

role of radical attack to the ionomer headgroups.

The insufficient understanding of the membrane structure thwarts further forays

in degradation modeling. To this end, we undertook molecular dynamics simulations of

the conformation of single chain ionomers as a function of different structural

parameters. This study revealed the nonmonotonic effect of the side chain length and

density on the conformational behaviour and rigidity of ionomer backbones. We discuss

how the changes in these architectural parameters change the ionomer affinity to

counterions and the corresponding ion mobility.

Studying the aggregation of ionomer chains revealed their spontaneous

aggregation in dilute solution. We explored the effect of various parameters such as

ionomer hydrophobicity and side chain content on ionomer bundle formation.

Minimization of the surface free energy of hydrophobic backbones is the driving force of

ionomer aggregation, while the repulsion of anionic headgroups opposes the

aggregation. The results rationalize the experimental studies and highlight the role

ionomer bundles as the prevailing structural motif in PFSA materials.

Page 4: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

IV IV

Keywords: Ionomer degradation, Kinetic model, Single chain, Ionomer self-assembly, Molecular dynamics.

Page 5: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

V V

Dedication

To Sahar

Page 6: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

VI VI

Acknowledgements

I would like to thank:

My supervisor, Prof. Michael Eikerling, for letting me to work under his supervision, his

guidance throughout my study, and providing constant support during my PhD.

My supervisory committee Prof. Steven Holdcroft and Prof. Barbara Frisken for their

helpful discussions to improve my research.

My internal and external examiner Prof. Leach and Prof. Soldera for taking the time to

read my thesis.

Compute Canada and Westgrid systems for providing the computational resources.

Past and present members of the Eikerling research lab.

Page 7: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

VII VII

Table of Contents

Approval..................................................................................................................II

Abstract..................................................................................................................III

Dedication...............................................................................................................V

Acknowledgements................................................................................................VI

TableofContents...................................................................................................VII

ListofTables............................................................................................................X

ListofFigures..........................................................................................................XI

ListofAcronyms....................................................................................................XV

ListofSymbols.....................................................................................................XVII

Chapter1. Introduction................................................................................1

1.1. Fuelcells........................................................................................................1

1.2. Polymerelectrolytemembranematerials......................................................5

1.2.1. Perfluorinatedionomers...............................................................................6

1.2.2. Partiallyfluorinatedpolymers.......................................................................9

1.2.3. Non-fluorinatedhydrocarbons...................................................................10

1.3. Nafionmorphology......................................................................................13

1.4. Nafionsolutions..........................................................................................17

1.5. Nafiondurabilityanddegradation...............................................................18

1.6. Thesisscopeandoutline..............................................................................20

Chapter2. Theory.......................................................................................23

2.1. Moleculardynamicssimulation...................................................................23

2.2. Coarsegraining............................................................................................24

2.2.1. Bead–springmodel.....................................................................................26

2.3. Forcefield...................................................................................................27

2.3.1. Short-rangeinteractions.............................................................................27

2.3.2. Long-rangeelectrostaticinteractions.........................................................30

2.4. Periodicboundarycondition........................................................................31

Page 8: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

VIII VIII

2.5. Equationofmotionandthermostat.............................................................32

2.6. Integrationscheme......................................................................................34

2.7. Moleculardynamicscomputerprogram......................................................35

Chapter3. Coarse-grainedModelingofPFSAChemicalDegradation...........36

3.1. Abstract.......................................................................................................36

3.2. Introduction................................................................................................37

3.3. Modelingstrategyandassumptions............................................................39

3.4. Mechanismsofradicalformationandconsumption....................................39

3.5. Ionomerdegradation...................................................................................44

3.6. Developmentofcoarse-grainedkineticmodel.............................................47

3.6.1. Modelformulation......................................................................................47

3.6.2. Kineticparameters......................................................................................50

3.6.3. Parametricstudy.........................................................................................51

3.6.4. Applicationofthemodelandvalidation.....................................................56

3.7. Conclusion...................................................................................................60

Chapter4. MDSimulationStudyofSingleChainIonomers.........................62

4.1. Abstract.......................................................................................................62

4.2. Introduction................................................................................................63

4.3. Chainmodel................................................................................................66

4.4. Resultsanddiscussion.................................................................................69

4.4.1. Effectsofsidechainlengthanddensity......................................................69

4.5. Localizationofcounterions..........................................................................77

4.6. Effectofelectrostaticinteractionstrength...................................................79

4.7. Effectofcounterionvalency........................................................................80

4.8. Conclusion...................................................................................................81

Chapter5. MDSimulationStudyofIonomerAggregation...........................83

5.1. Abstract.......................................................................................................83

5.2. Introduction................................................................................................84

5.3. Ionomermodel............................................................................................86

5.3.1. Computationaldetails.................................................................................89

Page 9: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

IX IX

5.4. Results.........................................................................................................90

5.4.1. Calibratingthebackboneflexibility.............................................................90

5.4.2. Effectofionomerhydrophobicity...............................................................91

5.4.3. Effectofelectrostaticinteractionstrength.................................................96

5.4.4. Effectofsidechaincontent........................................................................99

5.4.5. Effectofcounterionvalence.....................................................................102

5.5. Discussion.................................................................................................105

5.6. Conclusions...............................................................................................107

Chapter6. Summaryandfuturework.......................................................109

6.1. Futurework...............................................................................................111

References....................................................................................................113

AppendixA. Solutionforcubicequation..................................................133

AppendixB. PythoncodecreatingtheLAMMPSdatafile.........................134

Page 10: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

X X

List of Tables

TABLE1.1.COMPARISONOFDIFFERENTTYPESOFFUELCELLTECHNOLOGIES.4,13–15..........................................................3TABLE3.1.SYSTEMOFREACTIONSANDTHEIRKINETICINFORMATIONCONSIDEREDINTHISSTUDY......................................40TABLE3.2.EXPERIMENTALCONDITIONANDTHEOBTAINEDKSVALUES.THEDEGRADATIONCONDITIONOFEACHEXPERIMENT

WASEMPLOYEDTOOBTAINTHE�OHCONTENTUTILIZINGTHEFIRSTBLOCKOFTHESTUDY.......................................58TABLE4.1.LISTOFSYSTEMPARAMETERSANDTHEIRBASELINEVALUESALONGWITHTHEEXPLOREDRANGE.........................68TABLE5.1.LISTOFSYSTEMPARAMETERSANDTHEIRBASELINEVALUESALONGWITHTHEEXPLOREDRANGE.........................88

Page 11: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

XI XI

List of Figures

FIGURE1.1.(A)GLOBALPOWERGENERATIONFROMDIFFERENTRESOURCESANDTHEFORECASTFORTHENEXT24YEARS(DATA

EXTRACTEDFROMIEEJ2014REPORT,REF[2]).(B)RAGONEPLOTCOMPARINGDIFFERENTENERGY

STORING/CONVERTINGDEVICES(REPRODUCEDFROMREF[3]).............................................................................2FIGURE1.2.REPRESENTATIONOFAPOLYMERELECTROLYTEFUELCELL(PEFC)................................................................4FIGURE1.3.CHEMICALSTRUCTUREOFNAFIONMEMBRANEMATERIAL..........................................................................7FIGURE1.4.NAFION117CONDUCTIVITYASAFUNCTIONOFMEMBRANEWATERCONTENT.REPRINTEDWITHPERMISSIONFROM

REF[33]...................................................................................................................................................7FIGURE1.5.CHEMICALSTRUCTUREOFAQUIVIONAND3MPFSAIONOMERS.................................................................9FIGURE1.6.CHEMICALARCHITECTUREOF(A)BAMMEMBRANESAND(B)RADIATION-GRAFTEDPOLY(VINYLIDENEFLUORIDE)-G-

PSSA(PVDF-G-PSSA).............................................................................................................................10FIGURE1.7.CHEMICALSTRUCTUREOF(A)POLYSTYRENESULFONICACID,PSSAAND(B)SULFONATEDPOLY(ETHERETHER

KETONE)S(S-PEEK)..................................................................................................................................11FIGURE1.8.SCHEMATICILLUSTRATIONOFTHEMICROSTRUCTURESOFNAFIONANDS-PEEKSHOWINGTHELESSPROMINENT

HYDROPHOBIC/HYDROPHILICSEPARATIONOFTHES-PEEKCOMPAREDTONAFION.63.............................................12FIGURE1.9.SCHEMATICREPRESENTATIONOFWATERDOMAINSANDIONICCLUSTERSINGIERKE’SMODEL.77.......................14FIGURE1.10.STRUCTURALEVOLUTIONMODELPROPOSEDBYGEBELATDIFFERENTWATERFRACTIONS[76].......................15FIGURE1.11.ILLUSTRATIONOFPARALLELWATER-CHANNELMODELFORNAFION,PROPOSEDBYSCHMIDT-ROHRETAL.79THE

CYLINDRICALWATERCHANNELSAREPRESENTEDINWHITE,THENAFIONCRYSTALLITESARESHOWNINBLACKRECTANGLES

ANDTHENON-CRYSTALLINEDOMAINSARESHOWNASLIGHTGREEN....................................................................16FIGURE1.12.(A)NAFIONDISPERSIONSTATEREPORTEDBYDIFFERENTGROUPS.81,89,92,93,71,94,95,96,84,85(B)POSSIBLE

CYLINDRICAL,LAMELLAR,ANDCAGE-LIKESUPERSTRUCTURESFORMEDBYIONOMERBUNDLES.97,79...........................18FIGURE1.13.(A)PRISTINENAFION112MEMBRANE(B)NAFION112AFTER100HOURSINSITUDEGRADATIONTESTING.

[ADOPTEDWITHPERMISSIONFROMREF[103],JOURNALOFPOWERSOURCES,A.C.FERNANDES,E.A.TICIANELLI,193

(2009)547–554,ELSEVIER].(C)SCHEMATICILLUSTRATIONOFTHECHEMICALDEGRADATIONENVIRONMENTIN

HYDRATEDNAFIONMEMBRANE...................................................................................................................19FIGURE2.1.SCHEMATICILLUSTRATIONOFCOARSE-GRAININGMETHODOLOGYFORAPOLYMERCHAIN.(A)THEATOMISTIC

MODEL.(B)THECOARSE-GRAINEDSCHEMEWHEREDIFFERENTATOMSAREGROUPEDINTODIFFERENTSUB-UNITS........26FIGURE2.2.SCHEMATICILLUSTRATIONOFTHEBEAD-SPRINGCHAINMODEL..................................................................26FIGURE2.3.LENNARD-JONESPOTENTIALSHOWINGTHEATTRACTIVERIJ

-6ANDTHEREPULSIVERIJ

-12TERMS.........................28

FIGURE2.4.INTERACTIONPOTENTIALBETWEENNON-BONDEDBEADSWHICHINTERACTVIAONLYTHEWCAPOTENTIALAND

BONDEDBEADSWHICHINTERACTVIABOTHWCAANDFENEPOTENTIAL.............................................................29

Page 12: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

XII XII

FIGURE2.5.PBCINTWODIMENSIONSWITHTHEPRIMARYCELLENCLOSEDBYITSIMAGECELLS.MOLECULESLEAVINGTHEANY

CELLWILLBEREPLENISHEDBYTHEIRIMAGEPARTICLESENTERINGTHECELL............................................................32FIGURE3.1.DEPENDENCEOFREACTIONRATESONIRONCONTENTSATTHEH2O2CONCENTRATIONOF1M........................41FIGURE3.2.CONTRIBUTIONOFEACHREACTIONTOTHEFORMATIONANDCONSUMPTIONOFTHEINVOLVEDSPECIESINASYSTEM

OFFENTON’SREAGENTS.............................................................................................................................42FIGURE3.3.CALCULATED�OHCONCENTRATIONFROMANALYTICALSOLUTIONSINAWIDERANGEOFH2O2CONCENTRATIONS

(1×10-4-10M),ATDIFFERENTLEVELSOFIRONCONTENT.................................................................................44FIGURE3.4.COARSE-GRAINEDREPRESENTATIONOFTHENAFIONSTRUCTUREBYHEADGROUP,TRUNKANDBACKBONE

SEGMENTSANDDEGRADATIONRATESASSOCIATEDWITHTHELOSSOFEACHSEGMENT............................................47FIGURE3.5.FITTINGTHEANALYTICALEQUATIONSOF3.31AND3.32TOTHEEXPERIMENTALDATAOF[107].....................51FIGURE3.6.(A)LOSSOFHEADGROUPSDURINGTHE75HOFSIMULATEDCHEMICALDEGRADATION.(B)CONTENTOF

HEADGROUPSCLEAVEDBYTHEINDIRECTMECHANISMOFBACKBONEDEGRADATION.(C)RATEOFBACKBONE-INDUCED

HEADGROUPLOSSINTHESTUDIEDRANGEOFKS...............................................................................................52FIGURE3.7.IMPACTOFCHEMICALSTABILIZATIONONTHEREDUCTIONOFCHEMICALDEGRADATION(ABSCISSA)ATVARIOUS

VALUESOFKS............................................................................................................................................54FIGURE3.8.INFLUENCEOFKSONTHEFORMATIONOFWEAKENDGROUPS,βAC,FORDIFFERENTTRUNKLENGTHS.LONGER

SIDECHAINSCANMOREEFFICIENTLYSUPPRESSTHEMAINCHAINSCISSIONPROCESSINTHEKINETICREGIMEOFLOWKS...55FIGURE3.9.CHANGESINTHEHEADGROUPCONTENTANDCORRESPONDINGDECAYINIECATDIFFERENTLENGTHSOFTHE

BACKBONESEGMENT.................................................................................................................................56FIGURE3.10.COMPARISONOFTHEKINETICMODELTOEXPERIMENTALCHEMICALDEGRADATIONDATAOFNAFIONAND

AQUIVIONMEMBRANES[40,110,112]........................................................................................................57FIGURE3.11.COMPARISONBETWEENCALCULATEDIECLOSSANDEXPERIMENTALLYMEASUREDDATAFROMREF[107].......59FIGURE4.1.SCHEMATICILLUSTRATIONOFTHREEDIFFERENTLEVELSOFIONOMERMEMBRANES........................................64FIGURE4.2.SCHEMATICILLUSTRATIONOFCOMB-POLYELECTROLYTEARCHITECTURE.BACKBONEANDSIDECHAINHYDROPHOBIC

BEADSARESHOWNINBLUEANDGREEN,RESPECTIVELY.REDBEADSREPRESENTANIONICHEADGROUPSANDYELLOW

BEADSREPRESENTTHECOUNTERIONS............................................................................................................66FIGURE4.3.BACKBONERGFOR(A)NEUTRALCOMB-LIKEPOLYMERS,(B)COMB-POLYELECTROLYTECHAINS,AND(C)PURE

ELECTROSTATICCONTRIBUTIONTOTHEBACKBONERG.ALLTHEERRORBARSINTHISCHAPTERREPRESENTTHESTANDARD

DEVIATIONSOBTAINEDOVERTHEEQUILIBRIUMTRAJECTORIES............................................................................71FIGURE4.4.EQUILIBRATEDCOMB-POLYELECTROLYTECHAINSATDIFFERENTVALUESOFFANDLS.COLOR-CODINGISTHESAME

ASFIG.4.2..............................................................................................................................................73FIGURE4.5.EFFECTOFKθONBACKBONERG.COMB-POLYELECTROLYTE:(A)LS=3(B)F=1/2.NEUTRALCOUNTERPART:(C)LS

=3(D)F=1/2.ALLTHEPLOTSSHARETHELEGENDOF(A)................................................................................74

Page 13: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

XIII XIII

FIGURE4.6.LOCALPERSISTENCELENGTHFORTHECOMB-POLYELECTROLYTEANDTHENEUTRALBRANCHEDPOLYMERAT

DIFFERENTSIDECHAINFRACTIONSANDLENGTHS.............................................................................................76FIGURE4.7.RADIALDISTRIBUTIONFUNCTIONSBETWEENANIONICHEADGROUPSANDCOUNTERIONSFOR(A)F=1-1/5ANDLS

=3(B)F=1/2ANDLS=1-7(C)F=1/2,LS=3,ANDKθ=0-45KBT..............................................................78FIGURE4.8.RGVALUESASAFUNCTIONOFBJERRUMLENGTH(A)F=1/2ANDLS=1-7(B)F=1/2,LS=3,ANDKθ=0-45

KBT........................................................................................................................................................79FIGURE4.9.VALUESOFSQUAREROOTOFBACKBONERADIUSOFGYRATIONATDIFFERENTZC(A)F=1-1/5ANDLS=3(B)F=

1/2ANDLS=1-7.....................................................................................................................................81FIGURE5.1.INTHECHEMICALSTRUCTUREOFNAFIONIONOMERIN(A),NCORRESPONDSTO6CF2-CF2REPEATINGUNITS

BETWEENSIDECHAINS.COLORCODINGISTHESAMEASFIG4.2..........................................................................87FIGURE5.2.CALCULATEDPERSISTENCELENGTHOFSINGLELINEARCHAINMIMICKINGTHEIONOMERBACKBONE.AVALUEKθ=

25KBTREPRODUCESTHEPERSISTENCELENGTHOFAPTFECHAIN.THESNAPSHOTSSHOWTYPICALCHAIN

CONFORMATIONSINDIFFERENTREGIMESOFCHAINRIGIDITY.ALLTHEERRORBARSINTHISCHAPTERREPRESENTTHE

STANDARDDEVIATIONSOBTAINEDOVERTHEEQUILIBRIUMTRAJECTORIES............................................................91FIGURE5.3.(A)INITIALCONFIGURATIONOFCHAINSINTHESIMULATIONBOX;(B)TO(D)SHOWSNAPSHOTSOFTHESELF-

ASSEMBLEDSTRUCTUREATDIFFERENTVALUESOFεLJ=0.5,1.0,AND1.5KBT.ACLOSE-UPOFATYPICALBUNDLEIS

SHOWNBELOWTHEεLJ=1.0KBTBOX..........................................................................................................92FIGURE5.4.(A)CHANGEINAVERAGEAGGREGATESIZE,<K>,ASAFUNCTIONOFTHEVARIATIONINTHEINTERACTION

PARAMETERεLJBETWEENAPOLARBEADS.(B)CHANGEINTHEMINIMUM(RADIAL)ANDMAXIMUM(LONGITUDINAL)

BUNDLESIZEVS.εLJ.THEINSETSHOWSTHEASPECTRATIOOFIONOMERBUNDLES..................................................93FIGURE5.5.RADIALDISTRIBUTIONFUNCTIONSOF(A)MONOMER-MONOMER,GM-M(R),(B)HEADGROUP-HEADGROUP,GH-H(R),

(C)HEADGROUP-COUNTERION,GH-C(R),ATDIFFERENTVALUESOFεLJ.PEAKLOCATIONRATIOSARESHOWNIN(A)BY

CONSIDERINGTHEFIRSTPEAKASTHEREFERENCEPEAK......................................................................................95FIGURE5.6.(A)DEPENDENCEOFAVERAGEBUNDLESIZEONTHEVALUEOFλB.IMBEDDEDPLOTSSHOWTHEDISTRIBUTION

FUNCTIONOFTHEBUNDLENUMBERVS.KFORλB=1σ,6σ,AND12σ;(B)SHOWSTHEFRACTIONOFCOUNTERIONS

WITHIN2.5σOFTHEANIONICHEADGROUPS.................................................................................................97FIGURE5.7.MONOMER-MONOMERRADIALDISTRIBUTIONFUNCTIONSFORDIFFERENTVALUESOFTHEBJERRUMLENGTH.....98FIGURE5.8.SNAPSHOTSSHOWINGTHEIONOMERBUNDLESATDIFFERENTVALUESOFNS(A)NS=3,(B)NS=6,(C)NS=9,(D)

NS=15..................................................................................................................................................99FIGURE5.9.THEEFFECTOFNSON(A)EQUILIBRIUMAGGREGATENUMBER,<K>(IMBEDDEDPLOTSSHOWTHEDISTRIBUTION

FUNCTIONOFTHEBUNDLENUMBERVS.K),(B)COULOMBICENERGYOFANIONICHEADGROUPS,AND(C)MAXIMUMAND

MINIMUMBUNDLEGYRATIONRADIUS.........................................................................................................100FIGURE5.10.RADIALDISTRIBUTIONFUNCTIONAMONGANIONICHEADGROUPS,GH-H(R),FORTHEEXPLOREDVALUESOFNS. 101

Page 14: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

XIV XIV

FIGURE5.11.MORPHOLOGIESOFIONOMERSOLUTIONSFORSOLUTIONSCONTAINING(A)MONOVALENT,ZC=+1,(B)

DIVALENT,ZC=+2,AND(C)TRIVALENT,ZC=+3,COUNTERIONS......................................................................102FIGURE5.12.EFFECTOFCOUNTERIONVALENCEZCON(A)AVERAGEBUNDLESIZES,(B)FRACTIONOFLOCALIZEDCOUNTERIONS

AND(C)EFFECTIVELENGTHANDDIAMETEROFBUNDLES.................................................................................103FIGURE5.13.(A)MSDOFCOUNTERIONSWITHDIFFERENTVALENCESAND(B)CHANGESINMSDOFBACKBONEMONOMER

BEADSFORBUNDLESFORMEDINTHEPRESENCEOFCOUNTERIONWITHDIFFERENTVALENCEZC...............................104

Page 15: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

XV XV

List of Acronyms

Acronym Definition

PFSA Perfluorosulfonic acid

AFC Alkaline fuel cell

MCFC Molten carbonate fuel cell

PAFC Phosphoric acid fuel cell

PEM Polymer electrolyte membrane

PEFC Polymer electrolyte fuel cell

MD Molecular dynamics

SAXS Small Angle X-ray Scattering

DLS Dynamic light scattering

ESR Electron spin resonance

PSSA Poly(styrene) sulfonic acid

PVDF Poly(vinylidene fluoride)

PPP Poly(para-phenylene)

COOH Carboxylic acid groups

FER Fluoride emission rate

DFT Density functional theory

LSC Long side chain

SSC Short side chain

ADT Accelerated durability test

TFE Tetrafluorethylene

GDL Gas diffusion layer

CL Catalyst layer

SHE Standard hydrogen electrode

MEA Membrane electrode assembly

SANS Small Angle Neutron Scattering

s-PEEK Sulfonated polyether ether ketone

SSC Short Sidechain

PTFE Polytetrafluorethylene

IEC Ion exchange capacity

Page 16: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

XVI XVI

EV Equivalent weight

SOFC Solid oxide fuel cell

BAM Ballard Advanced Materials

XRD X-ray powder diffraction

NMR Nuclear magnetic resonance

AFM Atomic force microscopy

RDF Radial distribution functions

HOPG Highly ordered pyrolitic graphite

EELS Energy loss spectroscopy

DPD Dissipative particle dynamics

COM Center of mass

MSD Mean square displacement

PPPM Particle-particle particle-mesh

Page 17: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

XVII XVII

List of Symbols

Symbol Definition

E0eq Standard cell potential

F Faraday constant

ΔG Gibbs free energy change

T Temperature

ε0 Vacuum dielectric permittivity

ε Dielectric permittivity

λ Water content

ς Nondimensional Headgroup content

θact Nondimensional active trunk content

θ∗ Nondimensional inactive backbone content

βact Nondimensional active backbone content

β∗ Nondimensional active backbone content

Λs Apparent rate constant of side chain attack reaction

Λu Apparent rate constant of unzipping reaction

N Number of CF2 units in the backbone

N’ Number of CF2 units in the side chain

µi Nondimensional mass of ith species

m Membrane mass

Mi Molecular weight of ith species

V Membrane volume

φ Nondimensional fluorine release

Ls Side chain length

f Side chain fraction

Ns Number of pendant side chains

k Spring constant in FENE potential

R0 Maximum extension in FENE potential

rc Cut off distance in Lennard Jones (LJ) potential

εLJ LJ interaction parameter

Page 18: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

XVIII XVIII

λB Bjerrum length

ZC Counterion valence

kθ Bending rigidity

L Simulation box length

Rg Gyration radius

Lp Polymer persistence length

g(r) Radial distribution function

Page 19: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

1 1

Chapter 1. Introduction

1.1. Fuel cells

The vast increase in the global demand for power, limited resources of fossil

fuels such as coal, natural gas, and oil (Fig. 1.1a) as well as the environmental impact of

greenhouse gas emissions resulting from their use drives the research into and

development of clean energy conversion technologies.1,2 Notable technologies such as

supercapacitors, capacitors, fuel cells, and batteries all have the capability of storing

and/or converting energy in a clean and efficient way. Each technology has its own

benefits and drawbacks. The Ragone plot in Fig 1.1b shows the specific energy vs.

specific power of these different technologies.1,3 Among the technologies compared fuel

cells exhibit the highest specific energy and the lowest specific power. Stated differently,

fuel cells can supply energy for the longest time (range) compared to other energy

systems while providing a low rate of energy generation.4

A fuel cell is an electrochemical device in which the chemical energy stored in a

fuel and an oxidant is converted into electricity.1,5,6 Fuel cells benefit from a high

thermodynamic efficiency, low or zero emissions of environmentally unfriendly gases,

and quiet operation.

Historically, the first proposition of the operational principles of fuel cells is

attributed to Christian Friedrich Schönbein and William Grove in 1838 and 1839,

respectively.7 However, the first practical applications of fuel cells were in the Gemini

Page 20: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

2 2

space program in the early 1960s.8 Challenges such as limited performance, short

lifetime as well as high production costs prohibited the successful commercialization of

fuel cells at that time. These challenges have been tackled during the last three

decades.9,10

Figure 1.1. (a) Global power generation from different resources and the forecast for the next 24 years (Data extracted from IEEJ 2014 Report, Ref [2]). (b) Ragone plot comparing different energy storing/converting devices (reproduced from Ref [3]).

There are different ways to distinguish the fuel cells, but the most common

classification scheme is based on the electrolyte material. There are five major

categories of fuel cells based on different electrolytes.4,11,12 Fuel cells are categorized as:

(1) alkaline fuel cell (AFC), (2) phosphoric acid fuel cell (PAFC), (3) polymer electrolyte

fuel cell (PEFC), (4) molten carbonate fuel cell (MCFC), and (5) solid oxide fuel cell

(SOFC). Table 1.1 presents a brief description of each fuel cell type beside an overview

of their major attributes.

(b)(a)

Combus'onEngines

Page 21: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

3 3

Table 1.1. Comparison of different types of fuel cell technologies.4,13–15

Fuel Cell Type Electrolyte Operating Temperature

Fuel/ Oxidant

Application Efficiency

Alkaline fuel cell (AFC)

Potassium hydroxide (KOH)

60-90°C H2/ O2

Space applications

60 - 70%

Molten carbonate fuel cell (MCFC)

molten Li2CO3, K2CO3

620-660°C Natural or bio gas, H2/ O2, Air

Stationary power with cogeneration. Continuous power applications

65%

Solid oxide fuel cell (SOFC)

Yttria (Y2O2) stabilized zirconia (ZrO2 )

800-1000°C

Natural or bio gas, H2/ O2, Air

Stationary power with cogeneration, Continuous power applications

60 - 65%

Phosphonic acid fuel cell (PAFC)

concentrated H3PO4

160-220°C Natural or bio gas, H2/ O2, Air

Stationary power

55%

Polymer electrolyte fuel cell (PEFC)

H+ conducting membrane

40-90°C H2/ O2, Air

Portable, automotive, and stationary applications

50 - 68%

PEFCs offer numerous practical benefits over other fuel cell systems. The solid

polymer electrolyte reduces electrolyte leakage and the necessity of controlling the

corroding acids/bases. PEFCs can work at high power densities and comparatively low

temperatures (40 – 90 °C), allowing for rapid start-up under diverse conditions. Their

capability to consume several hydrogen-containing fuels (i.e., H2, methanol, ethanol)

and silent operation make them appealing for portable electronics as well as large-scale

immobile power generation.16–18

The major components of a single (H2/O2) PEFC are shown in Fig. 1.2. The fuel

cell consists of two electrodes, viz. anode and a cathode, and an electrolyte membrane

inserted between them.1,14,19 Each electrode contains a porous gas diffusion layer (GDL)

and a catalyst electrode layer (CL). The entire assembly of these components is called

the membrane-electrode-assembly (MEA).8,19

Page 22: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

4 4

Figure 1.2. Representation of a polymer electrolyte fuel cell (PEFC).

During operation, humidified hydrogen gas is supplied to the anode where it is

electro-catalytically oxidized to electrons and protons (Eq. 1.1). Since the membrane is

electrically insulative, liberated electrons are forced to flow through the outer circuit. The

hydrated protons migrate through the polymer electrolyte membrane (PEM) and react

with oxidant (i.e., O2) at the cathode to form water (Eq. 1.2).9,20 The only product of the

electrochemical reaction between H2 and O2 is water and the electrical energy is

generated (Eq. 1.3).

Anode H2 g( )→ 2H + aq( )+ 2e- , (1.1)

Cathode

12O2 g( )+ 2H + aq( )+ 2e- → H2O aq( ) ,

(1.2)

Overall 2H2 g( )+O2 g( )→ 2H2O g( ) ,

(1.3)

The overall standard Gibbs reaction energy for Eq. 1.3 is ΔG = -237.0 KJ.mol-1

At equilibrium and at standard conditions (1 atm, 25°C), the cell potential (E0eq) relative

to the standard hydrogen electrode (SHE) is1,21

Eeq0 =

−ΔG2F

=1.23 V ,

(1.4)

Page 23: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

5 5

where F is the Faraday constant. The theoretical thermal efficiency of fuel cells can be

calculated by

εFuel cell = −

We

ΔH=ΔGΔH

= 83% ,

(1.5)

where ΔH is the reaction enthalpy (at standard conditions: ΔH = 286.0 KJ.mol-1) and We

is the electrical work. Theoretically, the imperfect efficiency of fuel cells is due to the heat

losses introduced by the -TΔS term in the ΔG (ΔG = ΔH - TΔS). Activation, ohmic and

mass transport losses in an operating fuel cell yield lower practical efficiency compared

to the theoretical prediction.1

Despite the promising prospects for the powering of cars, buses and stationary

power generators, the commercialization of PEFCs has been challenged mostly due to

the high cost of catalyst materials, durability constraints of PEMs, and the demand for

efficient and low-cost production and storage of hydrogen gas.

1.2. Polymer electrolyte membrane materials

A PEM is a thin film (<50 µm) made from an inert polymeric material that

contains acidic units.22 The acidic moiety is typically a strong acid such as sulfonic acid.

There are major requirements that have to be fulfilled for PEMs to be used in a fuel cell.

A PEM must be an electronic insulator and a good proton conductor. High proton

conductivity (in the range of 0.1 S.cm-1 at PEFC operating temperatures), high chemical

and thermal stability, robust mechanical properties, low electro-osmotic drag, low gas

permeability, and low cost are other basic requirements of a PEM.11,16,17,23–25 A wide

range of proton conducting polymers has been developed and studied as PEM materials

over the last two decades.4,6,20,22,25–28 These materials are usually classified as

perfluorinated ionomers, partially fluorinated ionomers, and non-fluorinated ionomers.

Page 24: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

6 6

1.2.1. Perfluorinated ionomers

The main class of PEMs are made of perfluorosulfonic acid (PFSA) ionomers.

These materials are copolymers of tetrafluorethylene (TFE) and perfluorinated vinyl

ether monomers with embedded sulfonyl fluoride units. The best known PEM material to

date is Nafion,28 invented by Walther Grot at DuPont in the late 1960s.29 The polymer

consists of a perfluorinated backbone with pendant perfluorinated vinyl ether side chains

terminated by sulfonic acid groups. The acidic units make up less than 15 mol % of the

material with respect to the amount of polymer repeating units of the backbone and play

a crucial role in water sorption and performance of PEFCs.30

Fig. 1.3 shows the chemical structure of Nafion. Nafion is synthetized through

free radical copolymerization of tetrafluoroethylene monomers and perfluorinated vinyl

ether monomers with pendant sulfonyl fluoride groups.22 The sulfonyl fluoride groups are

transformed into sulfonic acid by hydrolysis. The ion content of Nafion can be varied by

changing the ratio of the two monomers (i.e., varying x in Fig. 1.3). The equivalent

weight (EW in units of [g mol-1]) and the thickness are often used to describe

commercially available membranes. The EW is expressed as the weight of dry

membrane per mole of sulfonic acid units when the material is in the H+ form. The ion

content is also represented by the ion exchange capacity (IEC), defined as the reciprocal

of the EW (IEC = 1000/EW, with units of millimoles of ion exchange sites per gram of dry

polymer, [mmol.g-1]).1

Nafion exhibits exceptional mechanical strength and proton conductivity in the

range of 0.1-0.2 S cm-1 under fully hydrated conditions. Higher IEC results in a

membrane that can absorb more water, thus affording greater proton conductivity.

However, the high acid content affects the membrane mechanical strength as a PEM

with high IEC swells more excessively with water. For typical fuel cell applications,

Nafion PEMs have EWs of 900-1100 g mol-1.12,14,18,31

Page 25: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

7 7

Figure 1.3. Chemical structure of Nafion membrane material.

The performance of a PEM material is mostly assessed by its proton

conductivity, σ. In simple terms, a PEM consists of an interconnected network of water-

filled and proton-conducting pores.32 Therefore, σ depends on the pores’ connectivity,

their size and distribution as well as the proton conductivity in individual pores.1,8,9 The

connectivity and size of water-filled pores, and consequently the proton conductivity, are

functions of the membrane water content, λ, which is defined as the number of water

molecules per sulfonic acid group. Fig 1.4 shows the proton conductivity of Nafion 117 at

different levels of PEM water content.33 Nafion 117 represents a Nafion membrane with

EW of 1100 g.mol-1 and thickness of 0.007 in (175 µm). Higher water content results in

higher proton conductivity, mostly due to the formation of well-connected network of

pores that facilitates the efficient transport of protons through the PEM.

Figure 1.4. Nafion 117 conductivity as a function of membrane water content. Reprinted with permission from Ref [33].

The Teflon-like backbone, which is hydrophobic in nature, confers chemical

stability under both oxidizing and reducing conditions. The excellent chemical stability of

Nafion stems from backbone with the high C-F bond energy of 480 kJ/mol.20,29,34 The

hydrophilic sulfonic acid groups of the side chains can dissociate and donate their proton

O−CF2−CF−O−CF2−CF2−SO3-

CF3

−−− CF2− CF2 − CF2−CF−−−− x

Nafion x=6.6, EW=1100

Page 26: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

8 8

to the surrounding water phase, thus providing a medium for the transport of protons.

Nafion remains the most widely studied PEM in terms of conductivity and morphology.

The combination of excellent performance and chemical endurance renders Nafion the

technology standard to which new PEMs are compared.28

Although Nafion offers outstanding performance under normal conditions, it still

suffers from certain drawbacks: Nafion’s operating condition is usually limited to < 90 °C

and > 50% relative humidity (RH) due to its propensity for dehydration.25 The high cost of

manufacturing, 12% cost of a MEA with $250-500/m2, makes Nafion a poor choice for

automotive purposes.35 Recycling and disposal of fluorinated polymers like Nafion is

another concern. In ionic form, Nafion is challenging to process and cannot be simply

dissolved, melted, or extruded. For these reasons, there is an increasing demand to

develop alternative PEM materials at reduced cost and with improved features.36,37

In the mid-1980s, Ballard Power Systems presented several remarkable

enhancements in fuel cell performance utilizing PEMs from Dow Chemical. The Dow

materials are perfluorinated and similar in chemical structure to Nafion, but shorter

sidechains and a higher density of side chain grafting points along the backbone. It is

generally referred to as a short side chain (SSC) ionomer compared to Nafion which is

considered a long side chain (LSC) ionomer.38 PEMs formed from SSC ionomers exhibit

greater levels of crystallinity, lower electro-osmotic drag, and higher ion content, thereby

improving the PEM resistance to swelling and enhancing the conductivity. The Dow

membrane, however, was commercially abandoned due to its complex synthesis and

high cost. Lately, Solvay Solexis resumed the development of PEMs based on SSC

ionomers via a more feasible and cheaper synthetic method.39 The commercial name for

this PEM is Aquivion (previously known as Hyflon Ion). The side chain of Aquivion does

not contain the -O-CF2-CF(CF3)- segment next to the polymer backbone. Their proton

conductivity under well-hydrated conditions is 0.2 S cm-1 at PEFC operating

temperatures in the range of 40-90 °C.40,41 Another PFSA membrane is the 3M

membrane with longer side chain compared to Aquivion.42 However, this PEM is still in a

research and development stage and not much practical data are available.43,44 Fig. 1.5

shows the chemical structure of the Aquivion and 3M PFSA membranes.

Page 27: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

9 9

Figure 1.5. Chemical structure of Aquivion and 3M PFSA ionomers.

1.2.2. Partially fluorinated polymers

Due to the high cost and complexity of synthesizing the perfluorovinylether

co-monomers of PFSA membranes, partially fluorinated or nonfluorinated materials have

been continuously advanced in order to offer a better balance between cost and

performance. Partially fluorinated membranes can generally be prepared at lowered cost

while preserving several benefits of fully-fluorinated PEMs. Still, compared to

perfluorinated PEMs, these materials degrade faster and they exhibit lower

performance.4,26

In the mid to late 1990s Ballard Advanced Materials (BAM) developed three

generations of PEMs namely BAM1G, BAM2G and BAM3G.45,46 The first generation of

membranes was based on poly(phenylquinoxalene). Despite similar performance of

these PEMs to Nafion, their lifetime of about 350 hours fell short compared to the

required lifetime of 5000 hours for automotive purposes. BAM2G materials were based

on poly(phenylene oxide). These PEMs showed greater performance to the first

generation BAM, Nafion, and Dow PEMs, but underwent sever degradation after 500-

600 hours of operation. BAM3G was based on poly(trifluorostyrene) and showed higher

conductivity and a significantly improved lifetime of over 15000 hours.46 Varying the

ratios of trifluorostyrene monomers and/or manipulating the post-sulfonation conditions

can tune the ion content of these PEMs. The chemical structure of these PEMs is shown

in Fig. 1.6a. The fluorinated backbone mitigates radical-induced degradation while the

substituent (R) on the styrene groups acts as an internal plasticizer to lower brittleness

and enhance the mechanical properties.47–49

O−CF2−CF2−SO3-

−−−−CF2−CF2−CF2−CF−−−− y

O−CF2−CF2−CF2−CF2−SO3-

−−−−CF2−CF2−CF2−CF−−−− z

Aquivion y=5.5, EW=830

3M PFSA z=4.7, EW=850

Page 28: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

10 10

Figure 1.6. Chemical architecture of (a) BAM membranes and (b) radiation-grafted poly(vinylidene fluoride)-g-PSSA (PVDF-g-PSSA).

Radiation-grafted PEMs are another significant class of partially fluorinated

materials.50,51 These PEMs are prepared through radiation-induced grafting of a suitable

monomer such as styrene to a base fluoropolymer, followed by sulfonation of the

pendant chains. The base polymer is normally a commercially-available fluoropolymer

film such as poly(tetrafluoroethylene) (PTFE) and poly(vinylidene fluoride) (PVDF) as

shown in Fig. 1.6b. The grafting and sulfonation processes introduce ionic conductivity

while maintaining the mechanical integrity of the base polymer. Although the proton

conductivity of these PEMs has been reported to be comparable to other PFSA

membranes, the lack of structural control is an important problem in these materials.21

1.2.3. Non-fluorinated hydrocarbons

Hydrocarbon-based materials have been explored extensively over the last two

decades as alternative PEM materials.26,52 They offer several advantages over PFSA

membranes: they are less expensive, easy to synthetize, and provide greater control

over the macromolecular architecture.4,22,26,27 Polystyrene-based membranes exemplify

Page 29: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

11 11

an important class of non-flourinated membranes. One of the prominent examples in this

family is polystyrene sulfonic acid, PSSA. This material was the first PEM used in the

Gemini space programs. The chemical structure of this membrane is shown in Fig. 1.7a.

These membranes were inexpensive and easy to fabricate. However, they suffered from

a poor chemical stability and a short lifecycle because of the weaker C-H bond

compared to the C-F bond.53 Highly reactive radicals are supposed to deteriorate the

PEM by attacking the benzylic α-hydrogen atoms that can ultimately results in chain

scission.21

Figure 1.7. Chemical structure of (a) polystyrene sulfonic acid, PSSA and (b) sulfonated poly(ether ether ketone)s (s-PEEK).

Sulfonated polyaromatic PEMs are currently considered as one of the more

promising routes to high performance PEMs due to their appealing thermal and chemical

stability.47 Sulfonated poly(ether ether ketone)s (s-PEEKs) have been the focus of

numerous studies.52,54 The chemical structure of this polymer is shown in Fig. 1.7b. High

thermal and oxidative stability of these membrane materials is attributed to the fact that

the C-H bonds of the aromatic groups have higher bond energies compared to aliphatic

C-H bonds.47 The rigid and bulky backbone leads to stiff polymers with glass transition

temperatures (Tg) of more than 200 °C, appropriate for high temperature applications.

Increasing the operating temperature of fuel cells is desirable due to the enhanced

Page 30: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

12 12

tolerance of the anode catalyst to CO poisoning.11,20 However, these PEMs suffer from

their poor mechanical stability originating from excessive swelling in water when high

IEC PEMs are prepared to achieve high proton conductivity. The proton conductivity of

these hydrocarbon membranes drops more sharply with reducing water content than

Nafion. The brittleness of these membranes makes their handling challenging and can

cause mechanical breakdown during operation.27 Recent studies on s-PEEK membrane

focus on development of s-PEEK nanocomposite with the aim of improving the

mechanical stability.55 Silica oxide56, Zirconia oxide57, Graphene oxide58, and Titanium

oxide59 are notable examples of reinforcing additives to s-PEEK membranes. The

transport and mechanical features of these polymers are also found to be different from

those of Nafion and are explained by the different microstructure.60–62 In s-PEEK, the

backbone is less hydrophobic and the sulfonic acid groups are less acidic than those in

Nafion. Consequently, the phase separation into hydrophilic and hydrophobic domains is

less pronounced (Fig. 1.8). The aqueous channels in s-PEEK are tighter and more

branched compared to those in Nafion. The evidence of these less segregated domains

was captured through SAXS measurements.47,63

Figure 1.8. Schematic illustration of the microstructures of Nafion and s-PEEK showing the less prominent hydrophobic/hydrophilic separation of the s-PEEK compared to Nafion.63

Another recent trend in the development of high performance and durable

hydrocarbon membranes focuses on the synthesis of multiblock copolymer ionomers.

Multiblock copolymers with sulfonated hydrophilic blocks and neutral hydrophobic

segments exhibit phase separated and bicontinuous morphologies.64,65 High acid content

Page 31: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

13 13

in the hydrophilic domains guarantees high proton conductivity, even at low water

contents and high temperatures. High local density of sulfonic acid units results in

interconnection of water pores within the nano-sized hydrophilic areas under these harsh

conditions.24,66 The nonionic segments in the copolymer are responsible for providing

dimensional and mechanical integrity. Sulfonic acid groups are still introduced through

the post-sulfonation process, which results in less control over the level of copolymer

sulfonation and suffer from numerous side reactions.67 Direct sulfonation of multiblock

copolymers through incorporation of monomers bearing perfluorosulfonated moieties,

which are stronger acids than the –SO3H acids and are directly attached to aromatic

rings, is a vibrant topic of recent research.64,66–68

1.3. Nafion morphology

The molecular structure of Nafion facilitates the formation of interconnected and

proton-conducting water pathways under well-hydrated conditions.28 The substantial

difference in the solubility of the hydrophobic backbone and of the hydrophilic side chain

headgroups triggers spontaneous aggregation of polymer backbones that leads to a

phase-segregated morphology. Upon increasing the water content, a percolating

network of water-filled domains, lined by polar headgroups at the surface, forms. The

well-connected network of pores offers conductive pathways to transport protons

through the film, while the perfluorinated backbone provides the high structural integrity

of the PEM during hydration.32,69,70

Numerous experimental investigations based on small-angle X-ray (SAXS) and

small-angle neutron (SANS) scattering have been carried out to understand the

morphology of Nafion.71–74 A suite of morphological models have been suggested to fit

the same observed scattering profiles of Nafion membranes and to reproduce the

so-called “ionomer peak”.6,27,28,75,76 Nevertheless, an unanimously recognized structural

model for the solid-state structure of Nafion has not emerged because of the

co-existence of hydrophobic regions and hydrated ionic domains as well as the

Page 32: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

14 14

heterogeneous nature of the Nafion membrane due to randomly distributed sulfonic acid

groups under hydrated conditions.1

Gierke et al.77, based on SAXS data, proposed the first structural model of

Nafion. This model was based on the cluster-network understanding of Nafion.

According to this model, water sorption leads to the evolution of inverted micelles of

sulfonic acid of about 4 nm diameter, which are connected by narrow channels of ~1nm

diameter. The hydrophobic polymer surrounds these ionic clusters. Fig. 1.9 shows the

schematic illustration of this model.

Figure 1.9. Schematic representation of water domains and ionic clusters in Gierke’s model.77

Although simple, this model assisted the understanding of water fluxes and

proton transport of Nafion. Each cluster was anticipated to include about 70 sulfonic acid

units under fully hydrated conditions.77 The limitation of the cluster-network model was

its failure in describing the structural evolution of PEMs over a broad range of

humidification conditions, from the diluted ionomer solution to the dry membrane state.

Yeager and Steck78 proposed a three-phase model consisting of fluorocarbon

domains, an interfacial area, and ionic domains. Gebel et al.76 proposed one of the

widely accepted models. Using SAXS, they studied the structural evolution of Nafion as

a function of water content. Upon absorption of water, the isolated ionic clusters begin to

swell. As shown in Fig. 1.10, a percolation threshold is reached at a water volume

fraction Xv ~ 0.1, where Xv was defined as the volume of water to the volume of the

swollen membrane. Increasing Xv from 0.5 to 0.9 triggers a morphological inversion.6,76

At higher water contents, a network of ionomer aggregates dominates the solution

behaviour.

Page 33: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

15 15

Figure 1.10. Structural evolution model proposed by Gebel at different water fractions [76].

Schmidt-Rohr and Chen79 proposed a “parallel cylinder model” which

quantitatively replicated the experimental SAXS profiles of hydrated Nafion membranes.

As depicted in Fig. 1.11, this model considers parallel water channels coated with ionic

side chain groups, creating inverted-micelle cylinders with ~2.4 nm in diameter. A shell

of stiff hydrophobic backbone crystallites also stabilizes these packed water channels.

Stretched and cylindrical Nafion crystallites (~5 nm in cross-sections) are considered to

be parallel to the water channels, and form physical crosslinks to explain the mechanical

features of Nafion.79 The cylinder size is also larger than that in former models, which

justifies the exceptional transport properties of hydrated Nafion membranes.

Page 34: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

16 16

Figure 1.11. Illustration of parallel water-channel model for Nafion, proposed by Schmidt-Rohr et al.79 The cylindrical water channels are presented in white, the Nafion crystallites are shown in black rectangles and the non-crystalline domains are shown as light green.

The “parallel cylinder model” is seemingly supported by NMR studies on

stretched membranes. However, severe concerns regarding this model have recently27,75

been raised: a fixed number of parallel cylinders demands major structural

reorganization to adjust structural alterations introduced by water molecules, which is in

contrast to the fast equilibration of the membrane once water is introduced to it. Kreuer

and Portale75 suggested that the water structures is more likely to be locally flat and

narrow that allows the hydronium ions to electrostatically interact with several sulfonic

groups. This recent model can successfully reproduce the experimental SAXS data and

explain the changes in ionomer peak with change in PEM water content.27 Although

these studies provided us with useful information, they cannot unambiguously clarify the

microscopic and mesoscopic structure of the membrane.65,75,80

Polymerbackbone

Polymersidechain

Page 35: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

17 17

1.4. Nafion solutions

In addition to numerous studies focusing on the phase-separated morphology of

Nafion membranes, several studies have also been conducted to understand the state of

Nafion in the solution state in solvents containing alcohols and water.81–85,86–88

In accordance with the pioneering model of Aldebert et al.89 in Nafion solutions, SANS

and SAXS studies by the Grenoble group82,83 suggested that rod-like/bundle-like

structures prevail in dilute Nafion solutions. Fig. 1.12a shows various models of ionomer

bundles discussed in this thesis. Bundles have diameters of 12 nm, with the core made

up of aligned hydrophobic backbones and the side chains protruding out of the rod

surface. Bundle formation is driven by the minimization of surface energy of hydrophobic

ionomers. Electrostatic repulsion between the anionic charges of side chain headgroups

increases the energetic cost of ionomer aggregation and disfavours the formation of

ionomer bundles. The balance between these interactions controls the stable size of

ionomer bundles in solution.90

From an electron spin resonance (ESR) experiment, Pietek et al.84,85 proposed a

fringed-rod model of Nafion. According to this model, at low Nafion concentrations,

0.5-5 % w/w ionomer with respect to the solution, part of the Nafion chains are

incorporated in a bundle while other chains are dispersed in the surrounding solvent.

This structure is termed as the isolated Nafion bundle (Fig 1.12a). At higher Nafion

content, the chains outside of the bundles start incorporating into other bundles and

make a hexagonal structure, described as reverse micelles. With further increase in the

ionomer concentration, the incorporation of other floating chains continues and

eventually the entire polymer phase becomes connected.84,85,91 This also explains how

rod-like micelles in dispersions of the ionomer make a transition to reverse micelles in

the swollen membrane.

Based on dynamic light scattering (DLS) studies of Nafion solutions, Jiang et al.81

proposed a microgel model in accordance with the fringed-rod model. According to this

model, one or several of the Nafion chains form fringed ionomer bundles, which further

interconnect with each other through the incorporation of the chain ends to form cross-

Page 36: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

18 18

linked aggregates. Unlike the reverse micelles of the fringed-rod model, the microgel

model takes into account the long length of ionomer chains with respect to ionomer

bundles. One chain in the aggregate can join to two or more bundles acting as crosslink

sites.

Figure 1.12. (a) Nafion dispersion state reported by different groups.81,89,92,93,71,94,95,96,84,85 (b) Possible cylindrical, lamellar, and cage-like superstructures formed by ionomer bundles.97,79

At the microscopic level, theoretical and experimental studies indicate that

cylindrical ionomer bundles are the dominant structural motifs in PEMs. These bundles

define the electrostatic and elastic properties of the hydrated membrane. The bundles

can further self-assemble and form superstructures (Fig 1.12b). It can be postulated that

the superstructures of these bundles in different forms, such as parallel, random, cage-

like, and lamellar, can be responsible for different experimental observations and

intuitive structural models of hydrated membranes.70,98

1.5. Nafion durability and degradation

As discussed before, the mechanical, chemical, and thermal durability of the

membrane is a major concern in the commercialization of PEFCs.99,100 The membrane

should last for at least 5000 hours for operation in automobiles and 20,000 hours for

stationary applications. However, the membrane is one of the most vulnerable

Rubatatetal.2004Szajdzinska-Pieteketal.1994

Aldebertetal.1988 Jiangetal.2001

(a)

(b)

(a)

Page 37: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

19 19

components of a PEFC; it predominantly determines the lifetime of PEFCs in heavy-duty

bus applications.12,20,31 PEM are seen to become thinner and develop pinholes after a

few hundred to few thousand hours of operation, depending on the operating conditions.

Chemical degradation of PEMs is considered to contribute to the gradual performance

decay and membrane thinning. Mechanical degradation involving the formation of

cracks, tears, and punctures has been recognized as a major cause of abrupt

membrane failure.101,102 Physical cracks act as shortcuts for the reactant gases to reach

the opposite electrode and react on the catalyst surface. This would reduce the cell

voltage while the corresponding heat of reaction can melt the membrane in local

spots.48,103 Fig 1.13a-b shows the physical deterioration of a PEM after a degradation

test. Membrane is visibly thinner and has developed numerous cracks on its surface.

Figure 1.13. (a) Pristine Nafion 112 membrane (b) Nafion 112 after 100 hours in situ degradation testing. [Adopted with permission from Ref [103], Journal of Power Sources, A. C. Fernandes, E. A. Ticianelli, 193 (2009) 547–554, Elsevier]. (c) Schematic illustration of the chemical degradation environment in hydrated Nafion membrane.

Chemical degradation of PEMs is commonly believed to play the most significant

role in governing the lifetime of PEFCs. The PEM neighbours the oxidizing environment

of the anode and the reducing conditions of the cathode.104,105 Any crossover of gases

between anode and cathode can result in radical formation. Experimental findings

revealed that an extremely oxidizing medium containing •OH and •OOH and H2O2

species is generated during operation of PEFCs.100,101,106 These species can be the

(a) (b)

(c)

[F-aq][F-aq]

Page 38: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

20 20

result of direct chemical or electrochemical reactions. In fact, it is the attack by these

radicals that is considered the chief cause for chemical degradation of PEFC

membranes. They attack the ionomer structure causing irreversible damage.

PEM chemical degradation occurs primarily through radical attack on backbone

carboxylic acid groups (COOH) that are left over from the ionomer polymerization

process, or on the C-S or O-C bonds of the side chain. Formation of polymeric radicals

at backbone end or side chain termini indicate the susceptibility of Nafion structure to

radical attacks.107,108 Newer generations of Nafion have been chemically stabilized

through lowering the concentration of terminal COOH groups to negligible quantities.

Consequently, their backbone is resistant to radical attack while degradation strikes only

on the side chain.40,109,110 Chemical degradation of membrane yields fluoride ions ([F-aq]),

which are measured in the runoff water during degradation tests.20,101,104,105,110 The

fluoride emission rate (FER) is a standard parameter to evaluate the rate of the chemical

degradation of PFSA membranes. Experimental observations such as decrease in IEC,

change in pH, conductivity, and membrane thickness are all also used to quantify the

chemical degradation of Nafion membranes.108,111,112

1.6. Thesis scope and outline

PFSA membranes face a variety of challenges out of which the limited lifetime

and the lack of basic structural/morphological understanding are deemed the main

obstacles. In terms of durability studies, despite numerous indications of side chain-

initiated radical attack to PFSA ionomers,104,107,109,111 there is no analytical model in the

literature that takes into account this dominant mechanism of radical attack. Current

models113–115 solely focus on the backbone-initiated pathway of radical attack and as a

consequence they cannot rationalize the chemical degradation of chemically stabilized

PFSA materials. In current models, the lack of a proper understanding of radical

formation mechanisms has usually resulted in an overestimation of the radical contents

in PFSA materials. The objective of the chemical degradation model, presented in

Chapter 3, is therefore to devise the first-in-kind comprehensive model of radical

formation and ionomer attack in a coarse-grained analytical approach. The coarse-

Page 39: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

21 21

grained nature of the model allows the derived equations to be generalized to different

PFSA membranes with the typical comb-like structure of fuel cell ionomers. The model

successfully predicts the chemical degradation of different commercial ionomers and

underlines the common mechanism of radical attack to different PFSA materials. Effects

of different ionomer properties such as IEC and side chain length on the chemical

degradation of PFSA materials are also evaluated and discussed. Discrepancies are

debated within the context of morphological features of hydrated ionomers and the

possible changes in the morphology upon chemical degradation.

A major shortcoming of the chemical degradation model is the lack of

morphological details. This oversimplification is due to an insufficient understanding of

the self-organized structure of ionomer aggregates and the superstructures that they

could form. We believe that a proper knowledge of the conformational properties of

single ionomer chains and of the structure-forming processes during aggregation or

bundle formation of such chains could vastly enhance the current understanding of

morphological/structural properties of hydrated ionomer materials.

A few theoretical116,117 and experimental118 studies have focused on the

conformational properties of PFSA-type ionomer chains, which correspondingly define

their elastomechanical features. These approaches116-118 focused on the conformational

behaviour of short (oligomeric) PFSA chains with response to varying solvent conditions.

So far, the effects of architectural parameters, which can be tuned during the synthesis,

on the conformational properties of long (realistic) ionomer chains are not explored and

no thorough conformational picture is offered. In Chapter 4, we employ molecular

dynamics (MD) simulations to analyse the fine conformational behaviour of long comb-

like ionomer chains as a function of the key architectural parameters, including the line

density of pendant side chains along the backbone and their lengths.

As discussed in Section 1.3 and 1.4, numerous studies point toward the ionomer

bundles as the major structural motif in PEM materials. MD simulations are employed to

study the self-assembling behaviour of a group of ionomer chains in Chapter 5. The aim

is to capture the changes in the bundle size and geometry upon varying significant

structural parameters such as the strength of ionomer hydrophobicity and density of

pendant acidic side chains along the backbone.

Page 40: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

22 22

The structure of the thesis is as follows

Chapter 2 introduces the theoretical background of the molecular dynamics simulation

technique employed in this thesis.

Chapter 3 presents the coarse-grained model of ionomer degradation upon hydroxyl

radical formation and attack. We derive analytical relations for calculating radical

concentration, devise the model of side chain–initiated degradation, and compare the

model to experimental data.

Chapter 4 discusses the conformational behaviour of single chain ionomer

macromolecules through molecular dynamics simulations. We discuss how changes in

the length and density of pendant side chains can substantially change the size, shape

and rigidity of ionomer chains.

In chapter 5, we present the results of a coarse-grained MD study of ionomer

aggregation. We explore the ionomer bundle formation upon changes in hydrophobicity,

side chain density, Coulomb interaction strength, and counterion valency. The trends

rationalize experimental observations and underline the aggregated nature of ionomer

molecules in dilute solution.

Chapter 6 presents the conclusion of our study and provides an outlook on possible

future works.

Each chapter covers a through and precise literature review on the discussed topic

along with a detailed explanation of the developed model or the employed technique.

Page 41: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

23 23

Chapter 2. Theory

2.1. Molecular dynamics simulation

Molecular dynamics is a powerful computer simulation technique used for

molecular systems, from ideal gases and liquids to biomolecules and soft materials.

Numerical calculations can be carried out to predict the structure and properties of new

materials without having to manufacture them or to replicate experiments in order to

reveal the hidden details.120,121 Examples span from design of novel drug encapsulating

micelles to prediction of fine conformational features of DNA and proteins. The systems

studied by MD simulations are modeled as ensembles of interacting particles under

predefined internal and external settings.122 A basic MD simulation involves the following

key steps:

1. Define initial particle positions and velocities at time t = 0.

2. Calculate the total force on each particle caused by interactions with all other particles due to the imposed force fields included in the simulation.

3. Numerically solve the equations of motion for each particle using the known particle positions, forces, and velocities at a given time to obtain the positions of each particle at a next time t + Δt.

4. Starting from the new particle positions, repeat steps 2 and 3 to calculate the particle positions at a time t + 2Δt, then t + 3Δt, and so on in order to capture the evolution of the system.

5. Assess the equilibration of the simulation and carry out the simulation under the equilibrated condition followed by analysis of the simulation data.

Page 42: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

24 24

From the obtained particle trajectories numerous average system properties can

be calculated such as total energy of the system, average displacement of particles

during the simulation, particle densities, correlation functions, or particle mobilities. Due

to the rigorous control of the intramolecular motions, MD is specifically suited for the

study of soft condensed matter systems such as biomolecules, colloids, and polymers.123

For polymer systems, processes such as the self-organization of a group of polymer

chains can be simulated and various properties, which are otherwise inaccessible to

experiments, can be analyzed. The analyzed properties can be crudely grouped into four

classes

1. Structural characterization. Examples include radial distribution function, dynamic structure factor, and radius of gyration.

2. Equation of state. Examples of parameters representing the equation of state or phase of a material include free-energy functions and phase diagrams to static response functions such as thermal expansion coefficients.

3. Transport properties. Examples involve viscosity calculations, diffusivity, and thermal conductivity.

4. Non-equilibrium responses. Examples include plastic deformation and pattern formation.

Analyses of these relations can help understanding the relations between the structure,

effective properties and specific functions of the simulated many-particle system.124

2.2. Coarse graining

All-atomistic MD simulations, which are based on accurate MD potentials

obtained from quantum mechanical calculations, have become a powerful technique for

analyzing the complex physical properties of polymers such as their dynamics, viscosity,

and relaxation behavior.120,121,123,125 However, it is still not practical to perform all-

atomistic MD simulations of highly entangled polymers, due to the large equilibration and

relaxation times, long-range Coulomb interactions and great number of atoms. The all-

atomistic simulation for such systems, with micrometer size scale (10 - 100 µm) and

relaxation times on the scale of microseconds (~ 1 - 50 µs), would consist of millions of

atoms and require millions of time steps (~ 1×106 - 1×108) to run. This task is beyond the

Page 43: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

25 25

capability of the most sophisticated supercomputers available today. Thus, it is

necessary to reduce the complexity of the simulated system in order to perform

meaningful studies.

One sophisticated method to increase the scope of molecular-level simulations is

to group atoms or monomers into bigger sub-units (Fig. 2.1).126 This coarse-graining is

due to the presence of numerous repeating units in polymer chains. Polymers display

time and length scaling for their dynamics and static properties.127,128 In fact, polymers

with dissimilar topologies respond universally to temperature and chain length

modifications. For instance, for the entangled systems the polymer viscosity, µ0, follows

µ0 ~ N3.4 where N is the degree of polymerization or the number of repeating units in the

polymer backbone.128,127 The benefit of coarse-graining lies in decreasing the number of

simulated groups and reducing the number of degrees of freedom in the simulated

system. This simplification allows improvement in two ways:

(1) increase the size of the simulated system: more chains or longer chains can be

studied through MD; (2) increase the simulation time to allow more thorough

equilibration and sampling. During the past two decades, a variety of coarse-grained

models have been proposed for polymers. One of the most successful approaches in

simulating polymer systems is the bead-spring model that will be introduced next.129

Page 44: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

26 26

Figure 2.1. Schematic illustration of coarse-graining methodology for a polymer chain. (a) the atomistic model. (b) the coarse-grained scheme where different atoms are grouped into different sub-units.

2.2.1. Bead–spring model

The bead-spring model is one of the most successful coarse-graining

approaches used for polymers.126,130,131 In this model, the molecular chain is replaced by

a string of beads, each bead representing one or a few monomer units.130,131 Fig. 2.2

shows the configurations of linear chains represented by this model. In this model, the

springs are represented by an anharmonic potential with two parameters, the spring

constant, k, and the maximum extensibility of an individual spring, R0.

Figure 2.2. Schematic illustration of the bead-spring chain model.

This model has been introduced in MD simulations of a single polymer chain

immersed in a solvent.131–134 It has received wide recognition as an efficient simulation

tool following polymer melt simulations by Kremer and Grest.126,130 This thesis presents

simulations using the bead-spring model.

(a)

(b)

Springs

Beads

Page 45: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

27 27

2.3. Force field

The monomers in the bead-spring model are usually represented by spherical

particles that interact via different forms of potentials. These potentials can be

short-range in nature such as the Lennard-Jones potential or long-range such as the

Coulomb potential in the case of charged beads.

2.3.1. Short-range interactions

The Lennard-Jones (LJ) potential is a widely used potential energy function for

systems controlled by dispersion forces.136,137 For two particles separated by a distance

r (σ), the commonly employed form of the Lennard-Jones potential is given by

U

LJ(r

ij) = 4ε

LJ

σrij

⎝⎜⎜

⎠⎟⎟

12

−σrij

⎝⎜⎜

⎠⎟⎟

6⎡

⎢⎢

⎥⎥, r

ij< r

c

(2.1)

where rc is a parameter defining the range of the interaction, and εLJ is a parameter

determining the strength of the dispersion force. The attractive term of the LJ potential,

(σ/rij)6, characterizes the dispersion force between the particles, and prevails at larger

separations. Dispersion forces are the result of induced dipoles caused by correlations in

electron density fluctuations of two atoms or molecules (Fig. 2.3). They are proportional

to the polarizability of the atoms or molecules and also depend on ionization

energies.122,136 The repulsive term, (σ/rij)12, accounts for Pauli repulsion due to the

overlap of electron orbitals.120,122,137 This component dominates at very short range and

prevents unphysical overlap between atoms or particles. The rij-12 dependency is not

based on any quantum mechanical derivation, but provides a reasonable agreement

with the empirically-derived potential.

Page 46: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

28 28

Figure 2.3. Lennard-Jones potential showing the attractive rij-6 and the repulsive

rij-12 terms.

The LJ potential well has a minimum at rij = 21/6 (σ) with ULJ(rij) = -εLJ. The LJ potential is

truncated and shifted at the specific cutoff distance of rc of the form

ULJ (rij ) =4εLJ

σrij

⎝⎜⎜

⎠⎟⎟

12

−σrij

⎝⎜⎜

⎠⎟⎟

6

−σrc

⎝⎜⎜

⎠⎟⎟

12

+σrc

⎝⎜⎜

⎠⎟⎟

6⎡

⎢⎢

⎥⎥, rij < rc

0 , rij > rc

⎪⎪

⎪⎪

(2.2)

In one common approach the cutoff distance is chosen at the minimum of ULJ(rij) and

then the potential is shifted such that it vanishes at this distance. This choice of cutoff,

rc = 21/6 (σ), results in purely repulsive part of LJ potential that is also known as the

Weeks-Chandlers-Andersen (WCA) potential.138 To account for attractive interactions,

larger cutoff distances such as rc = 2.5 (σ) are required.

For the chemically bonded beads along the polymer chain a combination of potentials

must be used: the WCA potential (Eq. (2.2)) to account for excluded volume interaction,

and the attractive Finite Extensible Non-Linear Elastic (FENE) potential (Eq. (2.3)) to

keep the consecutive beads along the chain bonded together. The anharmonic FENE

potential is the most common bonding potential in the polymer literature139

UFENE(r

ij) =

−12kR

0ln 1−

rij

R0

⎝⎜⎜

⎠⎟⎟

2⎛

⎜⎜

⎟⎟, r

ij< R

0

∞ , rij> R

0

⎨⎪⎪

⎩⎪⎪

(2.3)

4εLJ

σrij

⎝⎜⎜

⎠⎟⎟

12

−4εLJ

σrij

⎝⎜⎜

⎠⎟⎟

6Poten&

alene

rgy(U

LJ)

Page 47: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

29 29

where k is the spring constant and R0 is the maximum extension value. Superposition of

the FENE and WCA potentials (as shown in Fig. 2.4), yields the anharmonic spring

interaction between connected beads with an equilibrium bond length of b0 and a final

bond length of R0. Consequently, bond crossing in these systems is energetically

prohibited and chain entanglement is inherently obtained.140 (refer to the range of

parameters in FENE)

Figure 2.4. Interaction potential between non-bonded beads which interact via only the WCA potential and bonded beads which interact via both WCA and FENE potential.

There are numerous potentials to model the imposed rigidity of polymer chains. In our

study, the flexibility of the bead-spring chain is modeled with the following bond angle

potential of the form

Uθ rij( ) = kθ 1− cos θ −θ0( )( ) ,

(2.4)

where is the bending rigidity and θ0 the equilibrium angle that is set to 180º for linear

segments along the polymer backbone in the entire thesis.

R0

Poten&

alene

rgy(U

LJ)

r/Å

Page 48: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

30 30

2.3.2. Long-range electrostatic interactions

Since this thesis deals with charged polymers, it is necessary to introduce the

electrostatic method employed in the MD simulations. The pairwise long-range

interactions are the computationally most expensive part of molecular simulations. The

number of calculations for each atom scales with r2 where r is the spherical cut-off

radius.136,141 For a system of N particles with charges qi at positions ri in a cubic

simulation cell of length L with periodic boundary conditions (PBC), the long-range

Coulomb energy is calculated by

UCoul rij( ) = kBTλB

qiq jrij +mLi≠ j

N

∑m=1

m

∑,

(2.5)

where rij = ri - rj, qi is the partial charge of a given particle, λB = e02/4πε0εskBT is the

Bjerrum length, e0 is the unit charge and ε0 and εs are the permittivity of vacuum and the

relative dielectric constant of the solvent. The sum over m takes into account the

periodic images of the charged bead (m is an integer identifying a particular image cell;

m=(0;0;0) specifies the ‘‘real’’ cell). The 1/r nature of this potential and the fact that it is

only conditionally convergent, make it impossible to calculate Eq. 2.5. Also, employing a

cutoff to a long-ranged interaction has some drawbacks; a short cutoff would cause a

huge error whereas a long cutoff would be computationally expensive. The other

inherent challenge of this potential is that it varies steeply at small distances and decays

gradually at longer distances. Ewald142 developed a method to solve the problem of

long-ranged interactions in periodic systems; he split the potential into two parts

using142,143

1r=f r( )r

+1− f r( )r ,

(2.6)

f(r) should be chosen such that the two parts of Eq. 2.6 can be solved separately:

• f(r)/r should have a very small value beyond a cutoff distance. This allows the summation to be safely terminated after the cutoff.

Page 49: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

31 31

• (1-f(r))/r must vary gradually over the entire range.

The choice for the f(r) is f(r) = erfc(αr) where erfc(r) is the complementary error

function.122,137,141,142 The Ewald splitting parameter, α, is an inverse length that

harmonies the relative weights among the real and reciprocal spaces. Once

decomposed in this way, the summation converges faster and the complexity of the

computation decreases to O(N2). Ewald summation works well for small system sizes

with approximately hundreds of particles, but is still computationally too expensive for

systems involving thousands to millions of charged particles. The most time-consuming

part of this summation is the Fourier transformation. Replacing it by a faster scheme

accelerates the calculation. This idea is employed by more sophisticated family of

mesh-based techniques such as the Particle-Mesh Ewald (PME) method and

Particle-Particle Particle-Mesh (P3M) method, which lower the computation time to scale

with N.logN. This decrease in computational complexity is captured by substituting the

Fourier transform by a fast Fourier transform (FFT).144 This scheme entails discretizing

the space and charge that is done by mapping the charges onto a mesh. By doing that,

these approaches gain speed at the cost of losing resolution and the demonstration of a

point charge is not precise anymore, but it can be inferred as the accurate illustration of

a “cloud” of charge with the size of the grid spacing α. P3M is the most convenient

mesh-based technique and is broadly used to compute electrostatic interactions in

polymer systems.144

2.4. Periodic boundary condition

Molecular dynamics is classically applied to finite systems containing thousands

of particles. To approximate a large (infinite) system and to report the proper bulk

properties without the surface effects that are inherent in finite systems, periodic

boundary conditions or PBCs are applied to the system.122,137 In this scenario, particles

are generated in a volume V that is called the primary cell. The bulk is supposed to be

formed of the primary cell bordered by its replicas in order to simulate a macroscopic

sample. The image cells have the same size and shape as the primary cell. Cells are

Page 50: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

32 32

divided by open boundaries so particles can spontaneously enter or leave any of the

cells.121

Figure 2.5. PBC in two dimensions with the primary cell enclosed by its image cells. Molecules leaving the any cell will be replenished by their image particles entering the cell.

When particles leave the cell, their images concurrently enter the cell through the

opposite side. Fig. 2.5 illustrates the periodic boundary condition in two dimensions with

the primary cell surrounded by eight of its image cells.

2.5. Equation of motion and thermostat

Having detailed the potential energy function U(rij) that is composed of bonding,

bending, short-range, and electrostatic potentials, the next step is to calculate the atomic

forces on each particle.120,121 This can be done through stepwise time integration of

Newton’s equations of motion for a set of N particles

fi = −

∂∂riUi r1,r2 ,...rN( ) ,

(2.7)

Page 51: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

33 33

midVidt

= fi r1,r2 ,...rN( ) ,

(2.8)

here fi is the total force acting on particle i and Ui is the total potential energy, from which

the force is derived, Vi is the velocity vector of the particle i, and mi is the mass of

individual particles. By solving Newton’s equations, the total energy of the simulated

system is preserved and the time averages obtained during the simulations are

equivalent to averages in the micro-canonical ensemble with constant numbers of

particles N, volume V, and energy E (NVE). Most often this ensemble is not practical

when comparing with experiments that are carried out at constant temperature with

conserved N, V, and T (NVT).141,143

All the MD simulation results presented in this thesis were performed for NVT

ensemble, using Langevin thermostat.126,130 One way to simulate this ensemble is by

replacing Newton’s equations of motion with the Langevin equation

midV!"i t( )dt

= −∇Ui

!"!−Γdri!t( )

dt+ Fi!"!t( ) ,

(2.9)

Two additional forces are included in Eq. (2.9) to introduce the kinetic energy of the

system: a frictional force proportional to the velocity (Γ being the friction coefficient of the

particle) and a stochastic force Fi. The strength of the noise is related to Γ via the

fluctuation dissipation theorem, that is136,138,139

Fi!"!t( ) = 0 , (2.10)

Fi!"!t( ).Fi!"!t +τ( ) = 6kBTΓδijδ τ( ) ,

(2.11)

where <...> denotes an ensemble average and δij is the Kronecker delta.125 Through the

combination of frictional and random forces, the Langevin thermostat reflects the

coupling of particles with a virtual heat bath. A Langevin thermostat has two major

advantages that make it suitable for simulating polymeric materials for which long

simulation time is essential.122,123,125 The first advantage is that it accepts larger time

steps than other thermostats without causing instabilities. The other is that by coupling

Page 52: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

34 34

the system to a heat bath, the accumulation of errors is diminished, improving the

system’s stability over long simulation runs.

2.6. Integration scheme

Newton’s equations of motion are integrated according to the velocity-Verlet

integration scheme.143 The position and velocity of each particle with mass m in the

system are updated from time t to time t +Δt using an intermediate half time-step at

t + 1/2 Δt

r t +Δt( ) = r t( )+V t( )Δt + 12 a t( )Δt

2 , (2.12)

and

V t + 1

2Δt

⎝⎜

⎠⎟=V t( )+ 12 a t( )Δt ,

(2.13)

Using the updated positions and velocities from Eq. (2.12) and Eq. (2.13), new forces

f (t+Δt) are calculated. The new velocities are computed as

V t +Δt( ) =V t + 1

2Δt

⎝⎜

⎠⎟+12a t +Δt( )Δt

,

(2.14)

where a (t+Δt) is calculated by

a t +Δt( ) = −∇U r t +Δt( )( ) /m,

(2.15)

This integrator is fast, simple and stable and, importantly, it satisfies the requirements of

time-reversibility.120,122

Page 53: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

35 35

2.7. Molecular dynamics computer program

All simulations were carried out using the LAMMPS (Large-scale

Atomic/Molecular Massively Parallel Simulator) code.145 This is an open source

simulation code written in C++ by Steven Plimpton and coworkers at Sandia National

Laboratories. Visualizations were carried out by VMD (visual molecular dynamics).146

Python codes are written to analyze and post-process the simulation trajectories.

Page 54: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

36 36

Chapter 3. Coarse-grained Modeling of PFSA Chemical Degradation

3.1. Abstract

In this chapter we present a kinetic model of chemical degradation in PFSA-type

ionomers. The chapter is a compact version of the modelling work and analysis

discussed in detail in Ref [147].1 This model accounts for different possible pathways of

radical formation along with mechanisms of ionomer degradation through radical attack.

Simplifications in the set of model equations lead to the development of practically useful

analytical solutions for the concentration of hydroxyl radicals as a function of initial

concentrations of iron ions and hydrogen peroxide molecules. The coarse-grained

ionomer degradation model distinguishes units that correspond to ionomer headgroups,

trunk segments of ionomer sidechains, and backbone segments between two

sidechains. A set of differential equations is formulated to describe changes in

concentrations of these units. The model is used to study the impact of different

degradation mechanisms and ionomer chemistries on fluorine loss and rationalize the

variation in ion exchange capacity. Comparison of the model with experimental

degradation data for Nafion and Aquivion membranes allows rate constants of

degradation processes to be determined.

1 This chapter reproduces in revised form material from Ref [147], with permission from the

American Chemical Society.

Page 55: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

37 37

3.2. Introduction

The susceptibility of PFSA PEMs toward oxidative degradation has been

attributed to traces of partially hydrogenated groups, such as –CHF2 or –CF2COOH, that

are formed during ionomer synthesis, or to –CF=CF2 groups formed in the termination

step of the synthesis105,106,110,148. Reaction of these groups with oxidative free radicals

such as hydroxyl radicals (�OH) initiates a degradation cascade that leads to the

unzipping of the ionomer backbone. Reactions of hydrogen peroxide (H2O2), generated

in the electrodes149,150, and iron impurities (Fe2+/Fe3+) present in the membrane151, are

considered as the primary sources of �OH in PEFCs114,152.

Since the lifetime of PFSA PEMs is in the range of thousand of hours,

accelerated tests are commonly employed in PEM degradation studies153–155.

Accelerated durability tests (ADT) are performed in situ or ex situ. In the first case, the

membrane operates in a fuel cell as part of a membrane electrode assembly (MEA), and

degradation is studied either under on-line operation or by post-mortem analysis of the

PEM after pulling it out from the MEA. Ex situ tests, including the widespread Fenton’s

reagents test, which is a mixed solution of H2O2 and iron ions, exposes the membrane to

reaction conditions that mimics or exaggerates the aggressive environment encountered

in an operating fuel cell.156

Until recently, backbone unzipping was considered the major cause of chemical

degradation of Nafion and other PFSA ionomers110,114,115,157–159. Chemical stabilization of

Nafion reduces the concentration of COOH end groups and therefore minimizes the

impact of backbone unzipping. However, complete elimination of COOH units does not

fully eliminate degradation110,158. Recent theoretical160,161 and experimental107,111 studies

show that the loss of sidechain units of stabilized Nafion commence near the sulfonate

group. These findings suggest an alternative pathway, in addition to backbone

degradation, responsible for chemical degradation of Nafion and other PFSA ionomers

such as Aquivion and 3M PFSA.

Page 56: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

38 38

To date, few modeling works have attempted to simulate chemical degradation of

PFSA ionomers, particularly Nafion. The kinetic model by Gubler et al.114 employs an

extensive set of chemical equations to describe processes involving radical formation

and consumption. In their work, the unzipping of Nafion backbones under the attack of

�OH was considered as the sole degradation pathway. They assumed a large content of

COOH units (1.17 M), which is comparable to the sulfonate headgroup content. The 1D

model by Shah et al.115 accounts for the generation of H2O2 in electrodes, the formation

of radicals in the PEM, and the attack of radicals to Nafion ionomer. The set of reactions

considered in the radical balance was however incomplete. This resulted in

unreasonably high concentrations of radicals, which led to an overestimation of the rates

of ionomer degradation. They also focused on backbone degradation only; a high

concentration of weak end groups was employed to explain this simplification. The

kinetic model proposed by Xie and Hayden113 does not account for radical formation

pathways; it considered backbone-initiated and sidechain-initiated chemical degradation

of Nafion. It assumed that sidechain-initiated degradation leads to backbone cleavage

and the formation of COOH end groups. This model correlates the relative COOH

content to the fluorine loss of the membrane; contribution of lost fluorine from Nafion

sidechains is totally neglected. None of the models listed so far is suitable for stabilized

PFSAs with low initial content of COOH end groups.

In the following, we present a kinetic model of chemical degradation in the

presence of Fenton’s reagents. Section 3.3 presents the modeling strategy and

assumptions made. In Section 3.4, we present an analytical model to calculate the

balance of radicals under conditions that correspond to a test with Fenton’s reagents. In

Section 3.5, we provide a review of degradation mechanisms in PFSA ionomers, which

we utilize in Section 3.6 for the development of a coarse-grained model of PFSA

degradation. As part of this section, we validate the model and compare it to

experimental data. In the last section, we discuss the results and draw our conclusions.

Page 57: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

39 39

3.3. Modeling strategy and assumptions

The kinetic model of radical formation and attack to Nafion is zero-dimensional. It

assumes a uniform composition of the PEM and a homogenous distribution of Fenton’s

reagents throughout the membrane. Specific morphological properties of Nafion and

structural changes caused by and influencing chemical degradation are not taken into

account. Furthermore we consider isothermal conditions.

The time scale of formation and equilibration of radicals is much smaller than the

time scale of degradation phenomena in the PEM. Aggressive radicals such as hydroxyl

(�OH) and hydroperoxyl (�OOH) can be formed on the time scale of seconds or below162

while it takes hours or days for significant changes in membrane structure and

properties, caused by radical-induced degradation, to occur. We use this separation of

time scales to organize the model into two separate blocks, viz. one for the radical

balance and another for ionomer degradation. In the first block, we list the reactions

responsible for the generation and the consumption of oxidative radicals in the presence

of Fe2+/Fe3+ and H2O2. We formulate a set of equations that accounts for the currently

known mechanisms of radical formation and consumption. We prune this inventory by

discarding the reactions with negligible rates. This simplification allows analytical

relations for the radical concentrations as a function of initial iron content and hydrogen

peroxide concentration to be obtained. In the degradation block, we evaluate a coarse-

grained structural model of Nafion; we then derive the set of differential equations that

govern changes in the content of each segment.

3.4. Mechanisms of radical formation and consumption

Two reaction pathways are usually considered for a system of Fenton’s reagents:

a radical pathway, which proceeds via �OH and �OOH radicals that is favored under

acidic conditions and a non-radical pathway, which involves the generation of ferryl ions

and is preferred at high pH values.163,164 The radical-mediated pathway is considered in

Page 58: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

40 40

this study due to the superacidic nature of the PFSA ionomer. Table 3.1 lists the relevant

reactions.

Table 3.1. System of reactions and their kinetic information considered in this study.

# i

Reaction Reaction rate ri

Rate constant at room

temperature (M-1 s-1), ki

Activation energy

(kJ mol-1)

Rate constants

at 80°C (M-1 s-1)

1 H2O2 +Fe2++H+→ OH· +H2O + Fe3+ k1[H2O2][Fe2+]

65 [165] 35.4 [165]

602

2 H2O2 + Fe3+ → OOH· + Fe2++ H+ k2[H2O2][Fe3+]

7×10-4 [166] 126 [166]

2

3 Fe2+ + ·OH +H+→ Fe3+ + H2O k3 [Fe2+][.OH] 2.3×108 [167] 9 [168] 4×108

4 Fe2+ + OOH· +H+→ Fe3+ + H2O2 k4 [Fe2+][.OOH] 1.2×106

[169,170] 42 [169] 1.7×107

5 Fe3+ + OOH· → Fe2+ + H+ + O2 k5 [Fe3+][.OOH] 2×104 [170] 33 [171] 1.6×105

6 H2O2→ 2OH· k6 [H2O2]

1×10-22 [172] 200 [172]

2.9×10-17

7 OH· + H2O2 → OOH· + H2O k7 [.OH][H2O2] 2.7×107 [173] 14 [174] 6.5×107

8 OOH· + H2O2 → OH· +H2O + O2 k8 [.OOH][H2O2] 3 [175] 30 [142] 22

9 2OOH· → H2O2 + O2 k9 [.OOH]2 8.6×105 [177] 20.6 [178]

9.3×105

10 OH· + OH· → H2O2 k10 [.OH]2 5.4×10+9 [178] 7.9 [178] 5.6×10+9

11 OOH· + OH· → H2O + O2 k11 [.OOH][.OH] 6.8×10+9 [178] 14.2

[178] 7.2×10+9

The time-dependent concentrations of Fe2+, Fe3+, �OH, and �OOH are

determined by the following set of equations, based on the elementary reactions in Table

3.1,

Fe2+!"

#$+ Fe

3+!"

#$= Fe2+!

"#$0 ,

(3.1)

d Fe2+!" #$dt

= −r1 + r2 − r3 − r4 + r5,

(3.2)

Page 59: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

41 41

d •OH[ ]dt

= r1 − r3 + 2r6 − r7 + r8 − 2r10 − r11,

(3.3)

d •OOH[ ]dt

= r2 − r4 − r5 + r7 − r8 − 2r9 − r11,

(3.4)

where ri is the rate of reaction i in Table 3.1. The numerical solution of this set of

equations yields the rates of reactions 1-11 for iron content in the range of 1-100 ppm

and 1 M of H2O2 at 80 °C. This range of iron content and concentration of H2O2

reproduce the normal levels of iron contamination and generated H2O2 in the membrane

under both in situ and ex situ conditions.104,105,179,180 These steady state rates are shown

in Fig. 3.1. As can be seen, the rate of reaction 6 is independent of the iron content,

while rates of the other reactions increase with [Fe2+]0.

Figure 3.1. Dependence of reaction rates on iron contents at the H2O2 concentration of 1 M.

1 10 10010-1710-1610-1510-1410-1310-1210-1110-1010-910-810-710-610-510-410-310-2

Rea

ctio

n ra

te (M

s-1

)

[Fe2+]0 (ppm)

r1

r2

r3

r4

r5

r6

r7

r8

r9

r10

r11

Page 60: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

42 42

Considering the reaction rates in Fig. 3.1, we can evaluate which contributions to the

formation and consumption of �OH, �OOH, Fe2+, and Fe3+ will dominate. For instance,

Fe2+ is generated in reactions 1, 3, and 4 and the relative contribution of reaction 4 is

r4 (%) =

r4r1 + r3 + r4

(3.5)

Fig. 3.2 shows the relative contribution of each elementary reaction to the formation of

�OH, �OOH, Fe2+, and Fe3+ for [Fe2+]0 ranging from 1 to 100 ppm and [H2O2] = 1 M.

Figure 3.2. Contribution of each reaction to the formation and consumption of the involved species in a system of Fenton’s reagents.

Fe2+ is transformed to Fe3+ primarily by reaction 1 while it is regenerated via

reactions 2 and 5. Reaction 2 dominates the regeneration of Fe2+ at low [Fe2+]0 while

reaction 5 prevails at high [Fe2+]0. It can be seen that reactions 1 and 8 make the major

contributions to the generation of �OH with reaction 8 prevailing at low [Fe2+]0 and

reaction 1 prevailing at high[ Fe2+]0. Reaction 7 is the dominant reaction among reactions

3, 7, 10, and 11 in the consumption of �OH. The same reaction plays an important role

Fe2+

[1]

[3]

[4]

Fe3+

[2]

[5]

98.2-64.7 (%)

0.1-0.05 (%)

1.8-35.3 (%)

95.3-39.3 (%)

4.7-60.7 (%)

OH

0.1-0.05 (%) [3]

[7]

[10]

[11]

100 (%)

0.1-0.05 (%)

0.1-0.05 (%)

OOH

0.05-33.6 (%) [4]

[5]

[8]

[9]

0.05-57.9 (%)

97.4-4.7 (%)

2.5-3.8 (%) [11]

0.1-0.05 (%)

[1] [6]

[8]

[2]

[7]

2.6-93.1 (%)

0.1-0.05 (%)

97.4-6.9 (%)

2.5-36.2 (%)

97.5-63.8 (%)

.

.

Page 61: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

43 43

in the formation of �OOH in competition with reaction 2. Reactions 4, 5, and 8 control the

consumption of �OOH with reaction 8 prevailing at low [Fe2+]0.

At higher concentration of ferrous ions, reactions 5 and 4 play significant roles as sinks

of �OOH. Reactions 3, 6, 10, and 11 do not have any significant impact on radical

formation or transformation in the considered range of solution compositions.

In Fig. 3.2, reactions with < 0.1 (%) contribution to the formation or consumption

of the ensuing species are omitted from the reaction lists. Eliminating these insignificant

reactions simplifies the set of equations that needs to be solved in order to obtain the

steady-state concentrations of radicals and iron ions. From the remaining reaction set,

we derive analytical expressions for the steady-state concentrations of Fe2+, Fe3+, �OH,

and �OOH. The ratio of [Fe2+]/[Fe3+] is described by Eq. 3.6,

Fe2+!" #$Fe3+!" #$

=k2 H2O2[ ]+ k5 •OOH[ ]k1 H2O2[ ]+ k4 •OOH[ ]

(3.6)

Using [Fe3+] = [Fe2+]0 - [Fe2+], we obtain

Fe2+!" #$=

k2 H2O2[ ]+ k5 •OOH[ ]k1 + k2( )H2O2 + k4 + k5( ) •OOH[ ]

%

&''

(

)** Fe

2+!" #$0

(3.7)

The steady-state concentration of [�OH] is obtained as a function of [Fe2+] and [�OOH],

•OH[ ] =

k1 Fe2+!" #$+ k8 •OOH[ ]

k7

(3.8)

We then obtain a cubic equation for [�OOH], as shown in Appendix A. Using the solution

for [�OOH], expressions for [�OH] and [Fe2+] can be found from Eqs. 3.7 and 3.8. We

can thus express all concentrations of interest as analytical functions of [Fe2+]0 and

[H2O2]. The steady state concentration of [�OH] is shown in Fig. 3.3 over the considered

range of [Fe2+]0 and [H2O2].

Page 62: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

44 44

Figure 3.3. Calculated �OH concentration from analytical solutions in a wide range of H2O2 concentrations (1×10-4-10 M), at different levels of iron content.

Fig. 3.3 shows that [�OH] increases monotonically with [Fe2+]0 and [H2O2]. The

dependence on [H2O2] is pronounced at low [Fe2+]0. However, as [Fe2+]0 increases,

[�OH] becomes less sensitive to [H2O2]. [�OH] is mostly controlled by [Fe2+]0 in the entire

range of [H2O2]. Different regimes of ex situ and in situ conditions are demonstrated in

Fig. 3.3. The main difference between these regimes of degradation is the H2O2 content.

The �OH content possesses a comparable value in these two regimes of degradation

conditions. Quantification of �OH content and direct observation of this radical is not

possible, whether ex situ or in situ, due to its short lifetime. A way to assess the

presence of �OH is to transform it into longer-lived spin adducts and analyze them by

ESR methods, which is not a quantitative analysis.109,179,181,182 The plot in Fig. 3.3 can be

used to estimate the amount of �OH radicals under given degradation conditions.

3.5. Ionomer degradation

As stated in the previous section, we consider the hydroxyl radical to be the only

radical that can attack the PFSA polymer. This assumption is supported by the work

presented in Ref [182]. The authors of that article conducted an electron spin resonance

(ESR) analysis of a fuel cell (FC) inserted in an ESR resonator, in which Ce(III) was

10-4 10-3 10-2 10-1 100 101

10-12

10-11

10-10

in situ

100 ppm80 ppm

20 ppm

10 ppm

40 ppm

5 ppm

OH

radi

cal s

tead

y-st

ate

conc

entra

tion

(M)

H2O

2 (M)

1 ppm

ex situ

Page 63: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

45 45

added to Nafion as a radical scavenger. Results were compared to similar experiments

on uncontaminated Nafion. The most abundant radical was �OH in the case of the

untreated Nafion, while �OOH was the most abundant radical species in Ce(III)-treated

Nafion. Moreover, the carbon-centered radical (CCR) generated by membrane

fragmentation due to radical attack is absent in the case of Ce(III)-treated Nafion, while it

is present in the absence of Ce(III). Radical fragments confirming the presence of

hydrogen radicals (�H) are also reported in the same work. However, we do not address

the possible degradation impact of �H since the considered system of Fenton’s reagents

lacks H2, a precursory species for the formation of �H. According to Gubler et al.114, the

rate of hydrogen abstraction by �OH is more than 7 orders of magnitude faster than by

�OOH, due to the different bond dissociation energies of O-H in water and hydrogen

peroxide. The polymerization process of Nafion generates non-fully fluorinated end

groups (e.g. COOH, CF2, etc.) at the terminus of backbone chains. These defective

groups are the initiation points of the unzipping mechanism110,183 described by reactions

3.9 - 3.13,

RF -CF2COOH + •OH→RF -CF2CO2 • + H2O (3.9)

RF -CF2CO2 • →RF -CF2 • + CO2 (3.10)

RF -CF2 •+ •OH→RF -CF2OH (3.11)

RF−CF

2OH→ R

F−COF+HF

(3.12) RF -COF + H2O→RF -COOH + HF

(3.13)

Recently, advanced synthetic strategies and pre-treatment methods have

dramatically reduced the concentration of weak end groups in the Nafion backbone110.

Chemical modification of Nafion, Aquivion40 and 3M44 ionomer membranes has

increased the durability, as deduced from the reduced fluorine loss upon testing them

with Fenton’s reagents. However, chemical stabilization does not completely stifle

radical-mediated degradation of PEMs. For instance, in the work of Curtin et al.110, the

stabilization of Nafion has reduced the emission of fluoride by 56% after 54 hours of a

test conducted with Fenton’s reagents. Treating Nafion with �OH revealed a non-zero

intercept when plotting the fluoride emission rate (FER) as a function of COOH content

in the membrane158. The residual value of the FER corresponds to a degradation

Page 64: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

46 46

mechanism that cannot be eliminated by chemical stabilization of the backbone. The

loss of side chain units during the treatment of stabilized Nafion membranes with

Fenton’s reagents was observed.107 These observations highlight the importance of

degradation that stems from the sidechain of Nafion.

The authors of Refs. 184 and 160 used density functional theory (DFT)

calculations to study the cleavage of C-S bond by reaction with �OH. Yu et al.160

predicted the release of tetrafluoroethylene (CF2=CF2) and HSO4- groups upon �OH

attack, which results in the simultaneous loss of SCF2 (CF2 unit next to the sulfonate

group) and α-OCF2 (α-ether unit, ether bond nearby the sulfonate group) from Nafion

sidechains. Using 19F-NMR, Ghassemzadeh et al.107 also reported the loss of SCF2 and

α-OCF2 units. From the observation of -OCF2CF2SO3-, they concluded that the α-ether

unit in the ionomer sidechain is the primary point of �OH attack111. However, the 19F-

NMR technique cannot detect HSO4-, the signature degradation fragment of the C-S

cleavage mechanism. Thus, the dominance of C-S cleavage or α-ether loss is not

resolved yet. Both experimental results and theoretical findings suggest that irrespective

of the primary site of attack, either being the C-S bond or the α-ether site, the whole

OCF2CF2SO3- segment is lost, whether in a single-step or multi-step reaction.

Coms105 proposed that, after cleavage of a C-S bond by �OH, weak COOH

groups would be formed in the sidechain. Cipollini101 suggested that the C-S bond could

be hydrolyzed to detach the sulfonate unit and convert the neighboring CF2 group into a

COOH group. The mechanism proposed by Ghassemzadeh et al.108 also suggested the

formation of a COOH unit following the loss of OCF2CF2SO3-. The COOH group thus

formed is prone to further attack by �OH. From this point on, degradation processes

could propagate along the trunk, in the same manner as for backbone unzipping. At a

sidechain grafting point degradation proceeds further by splitting the backbone into two

chains both terminated by COOH groups. This step is knows as backbone scission.

Thereafter, main chain unzipping can thus occur on both backbone fragments.

Page 65: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

47 47

3.6. Development of coarse-grained kinetic model

3.6.1. Model formulation

Based on experimental findings discussed in the previous section, we adopt a

coarse-grained description for PFSA-type ionomers, which is applicable to Nafion and its

derivatives. The ionomer is divided into three units, as shown in Fig. 3.4, comprising

headgroup segments (S), trunk units (T), and backbone segments between two

branching points (B). Each unit is either in an inactive or an active state, the former

referring to a state immune to �OH attack contrary to the latter. Headgroups are always

active (Sact). Trunk segments are activated upon loss of the attached headgroup

segment (Tact). The inactive form (T*) denotes segments in undamaged sidechains.

Correspondingly, we distinguish units Bact and B* for the backbone.

Figure 3.4. Coarse-grained representation of the Nafion structure by headgroup, trunk and backbone segments and degradation rates associated with the loss of each segment.

The initial concentrations of the three segments are equal, [S]0 = [T]0 = [B]0. We

introduce dimensionless concentrations,

ς =

Sact"# $%S"# $%0

(3.14)

θact =

Tact"# $%S"# $%0

and θ* =T *"#

$%

S"# $%0

(3.15)

O | CF2 | CF2 | SO3

- Head group

Trunk

Backbone

(Tact, T*)

(Bact, B*)

(Sact)

rB

rT

rS

--CF2-CF2-CF2-CF2-CF2-CF2 -CF2-CF2-CF2-CF2-CF2-CF2--

| O | CF2 | CF(CF3) |

Page 66: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

48 48

βact =

Bact"# $%S"# $%0

and β* =B*"#

$%

S"# $%0

(3.16)

The apparent rate constant of headgroup loss can be represented by Λs = ks [�OH], as

suggested by DFT studies and experimental work.111,160 In Eq. 3.17 the rate of

headgroup loss can be represented by

rs = ks •OH!" #$ς

(3.17)

The apparent rate constant of the unzipping process in trunks and backbone units,

Λu (s-1), can be written as

Λu = ku •OH"# $% (3.18)

where ku is the rate constant of the unzipping process. In order to determine the rate for

the loss of trunk and backbone units, we assume that the loss of these units involves N´

and N consecutive pseudo first-order reactions with identical rate constants ku. N´ and N

can be identified with the number of CF2 units in trunk and backbone segments,

respectively. Deranleau185 showed that the general solution of the integrated rate

equations for such a process is given by

rT =

ku •OH!" #$θactN '

(3.19)

rB =

ku •OH!" #$βactN

(3.20)

The following set of ordinary differential equations describes the dynamics of these units,

dςdt= −ks •OH#$ %&ς −

ku •OH#$ %&N

βactς

(3.21)

dθactdt

= ks •OH"# $%ς −ku •OH"# $%N '

θact −ku •OH"# $%N

βactθact

(3.22)

dθ*

dt= −ks •OH#$ %&ς −

ku •OH#$ %&N

βactθ*

(3.23)

Page 67: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

49 49

dβ*

dt= −2

ku •OH#$ %&N '

θact −ku •OH#$ %&N

βact

(3.24)

dβactdt

= 2ku •OH"# $%N '

θact

(3.25)

Sidechain segments S and T can be lost through direct attack by �OH radical on their

active form or, indirectly, through the unzipping of the backbone. Both mechanisms are

accounted for in Eqs. 3.21 and 3.22.

The primary variables of the ionomer system, ς, θact, θ*, βact, and β*, defined in

Eqs. 3.14 to 3.16, are not measurable. However, changes in these variables can be

related to the fluorine emission rate and to the ion exchange capacity (IEC), which can

be monitored in experiment. We define dimensionless masses,

µi =

miMiV S!" #$0

(3.26)

where mi is the mass of species i, Mi is the corresponding molecular weight and V is the

volume of membrane. The dimensionless mass of the released fluorine is obtained from

the following equation

µF−= NSrS + NTrT + NBrB( )dt

0

t

(3.27)

where NS, NT, and NB are numbers of fluorine atoms liberated during the loss of

headgroup, trunk, and backbone units, respectively. The dimensionless mass of the dry

membrane is

µPEMdry =

1IEC( )0MF−

−MSrS +MTrT +MBrB( )

0

t

∫ dt

MF−

(3.28)

where MF- is the molecular weight of the released fluoride, 19 g mol-1, and (IEC)0 is the

initial ion exchange capacity of the membrane, respectively. The mass of released

fluorine, normalized to the dry mass of the membrane is given by

Page 68: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

50 50

φ =

µF−

µPEMdry

(3.29)

and the IEC is given by

IEC = ς

µPEMdry M

F−

(3.30)

3.6.2. Kinetic parameters

In this part, we discuss the extraction of the kinetic parameters from experimental

data. Ghassemzadeh and Holdcroft107 used quantitative 19F-NMR to track the

concentration of fluorine according to the functional group it belonged to. These authors

did not find any significant main chain degradation. In the absence of backbone

degradation, analytical expressions for the dimensionless concentrations of released

headgroup and trunk units can be obtained

ςloss =1− exp(−ks •OH#$ %&t) (3.31)

θloss =1

kuN '

− ks

ks exp(−ku •OH#$ %&N '

t)−kuN 'exp(−ks •OH#$ %&t)

#

$''

%

&((+1

)

*

++++

,

-

.

.

.

.

(3.32)

The degradation condition of 10 ppm Fe2+ and 20 vol% H2O2 at 80 °C yields the �OH

content of 2.4×10-11 M. By comparing the solutions given in Eqs. 3.31 to 3.32 with

experimental data we find ks= 4.4×104 M-1 s-1 for headgroup dissociation and ku= 8.3×105

M-1 s-1 for trunk loss, with N´ = 2. Fig. 3.5 shows the experimental data and the

calculated concentrations.

Page 69: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

51 51

Figure 3.5. Fitting the analytical equations of 3.31 and 3.32 to the experimental data of [107].

Experimental data and model exhibit a significant discrepancy at short

degradation time. During the initial period of the study (0-24 h), the rate of trunk

degradation seems to be suppressed. The model cannot reproduce this behavior. The

concentration of active trunk segments will be lower than values obtained in our

calculations. We attribute this difference to morphological effects. For headgroups, it is

reasonable to assume that under given experimental conditions, all of them are equally

exposed to the aqueous electrolyte, which contains Fenton reagents and �OH radicals.

However, the majority of activated trunk and backbone segments will not be surrounded

by electrolyte inhibiting their reactions with �OH. We speculate that only structural

reorganization,90 triggered by chemical degradation, exposes these units to the

electrolyte. Nevertheless, the value for the rate of the unzipping process in the trunk is in

reasonable agreement with the unzipping rate constant of (7.3±0.2) ×105 M-1 s-1 reported

for model compounds186.

3.6.3. Parametric study

In this section, we investigate the sensitivity of our model to its key parameters.

The main kinetic parameter is ks. The number of repeating units in the trunk, N´, and in

the backbone, N, are two structural parameters characterizing the specific PFSA

αCF2 SCF

2

CF3 CF(s) βCF

2

model (kS= 4.4x104 M-1 s-1)

model (ku= 8.3x105 M-1 s-1)

0 10 20 30 40 50

0

2

4

6

8

10

12

14

16

Fluo

rine

Loss

(%)

Time (h)

Page 70: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

52 52

modeled. All the parametric studies were carried out for 75 hours of simulated

degradation, assuming degradation conditions with 20 ppm Fe2+ and 30% H2O2 at 80 °C.

The studied membrane mimics a stabilized Nafion with (IEC)0 = 0.91 mmol g-1, and

βact,0 = 0. N and N´ are considered to be 14 and 2, respectively. Deviations of βact, N or

N´ from these reference values will be explicitly stated.

The impact of ks on the dimensionless headgroup content, ς, is shown in

Fig. 3.6a. At values of ks in the range of 102 M-1 s-1 to 103 M-1 s-1, the loss of headgroup

content after 75 hours is insignificant. However, almost complete loss of the Nafion

headgroups occurs after only 5 h with ks = 107 M-1 s-1 and after 30 h with ks=106 M-1 s-1.

All the sidechain segments are lost at high rates of headgroup dissociation; in this case,

since all the sidechain-induced backbone cleavages have occurred, chemical

degradation proceeds only through backbone degradation.

Figure 3.6. (a) Loss of headgroups during the 75 h of simulated chemical degradation. (b) Content of headgroups cleaved by the indirect mechanism of backbone degradation. (c) Rate of backbone-induced headgroup loss in the studied range of ks.

102 103 104 105 106 1070.00

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

102 103 104 105 106 1070.0

1.0x10-7

2.0x10-7

3.0x10-7

4.0x10-7

5.0x10-7

0 10 20 30 40 50 60 70

0.0

0.2

0.4

0.6

0.8

1.0

Cle

aved

ς fr

om

indi

rect

bac

kbon

e de

grad

atio

n

(c)(b)

Rat

e of

indi

rect

he

ad g

roup

loss

(M s-1

)

ks (M-1 s-1)

(a)

ς

Time (h)

ks=102 M-1.s-1

ks=103 M-1.s-1

ks=104 M-1.s-1

ks=105 M-1.s-1

ks=106 M-1.s-1

ks=107 M-1.s-1

Page 71: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

53 53

In addition to the direct loss of headgroups, backbone unzipping induces the

detachment of sidechains. The portion of headgroups lost due to this indirect mechanism

after 75 h of chemical degradation is shown in Fig. 3.6b. As can be seen in Fig. 3.6c, the

kinetics of indirect headgroup loss reaches a maximum at ks ~105 M-1 s-1. The existence

of such a maximum is expected: for the indirect loss to be significant, sufficient

backbone cleavage must occur, in which case the direct loss is actually dominant; when

the direct loss is very slow, the indirect loss is even more so. Indirect loss has a minor

influence on ς in both limits of small and large ks.

We have also explored the impact of stabilization. In non-stabilized membranes

we consider βact = 0.01 as the upper limit, based on the reported molecular weight of

200,000 g.mol-1 for Nafion 112110. Considering the equivalent weight of 1100 g mol-1 for

Nafion 112, approximately 200 sidechains per polymer are necessary to reach this

average molecular weight. Assuming both backbone termini to be defective, we thus

obtain the upper limit of the initial COOH content: it is a factor 100 times lower than the

sidechain content meaning βact,0 = 0.01. This value is consistent with the data reported

for pristine PFSA membrane,187 which exhibit an initial COOH content of around

0.01 mmol g-1 and a sidechain content of 1.1 mmol g-1.

Fig. 3.7 shows the reduction of φ when decreasing βact. The y-axis can be

interpreted as the efficiency of stabilization that can be obtained through replacing a

non-stabilized PEM by its stabilized counterpart. This reduction in the fluorine loss

strongly depends on ks. At ks << 104 M-1 s-1 stabilization significantly lowers the chemical

degradation while at regimes of swift headgroup loss, ks ≥ 105 M-1 s-1, stabilized and

non-stabilized membranes exhibit similar fluorine release rates.

Page 72: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

54 54

Figure 3.7. Impact of chemical stabilization on the reduction of chemical degradation (abscissa) at various values of ks.

Fig. 3.8 shows the effect of different sidechain lengths on the magnitude of βact

after 75 h of simulated degradation for the considered range of ks. In the entire range of

ks, ionomers with longer sidechains experience a lower formation of end groups.

Increasing the sidechain length by 5 CF2 units at ks = 102 M-1 s-1 results in a 45%

reduction of end groups content while a 27% reduction is seen at ks = 107 M-1 s-1.

Increasing the sidechain length reduces the amount of βact and delays the formation of

weak end groups. This decrease in βact is due to the lowered rate of trunk dissociation,

which controls the main chain scission process. Thus, increasing the sidechain length

can suppress main chain scission; the effect is more pronounced at lower ks.

102 103 104 105 106 107

0

20

40

60

80

100

φ (stabilized)

φ (non-stabilized)-φ(stabilized)

ks (M-1s-1)

Page 73: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

55 55

Figure 3.8. Influence of ks on the formation of weak end groups, βac, for different trunk lengths. Longer sidechains can more efficiently suppress the main chain scission process in the kinetic regime of low ks.

Fig. 3.9 shows the influence of backbone segment length, N, on the headgroup

content ς and corresponding changes in IEC. The legend shows the values of the IEC

that correspond to different values of N. Longer backbone segments lower the rate of

backbone degradation, following Eq. 3.20. Consequently, fewer headgroups are lost

from indirect backbone degradation for longer backbone segments. The decline in IEC

as a result of chemical degradation is less pronounced for Nafion with longer backbone

segments or lower initial IEC. Membranes with lower initial IEC are more durable in

terms of maintaining the sulfonate content during degradation and as a result are

expected to demonstrate greater durability. Rodgers et al.188 have recently observed that

Nafion with 1100 EW is demonstrating greater chemical stability compared to 950 EW

Nafion.

102 103 104 105 106 10710-3

10-2

10-1

100

45%

β act

ks (M-1.s-1)

N’=1 N’=2 N’=3 N’=4 N’=5 N’=6

27%

Sidechain length

Page 74: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

56 56

Figure 3.9. Changes in the headgroup content and corresponding decay in IEC at different lengths of the backbone segment.

3.6.4. Application of the model and validation

We use this model to evaluate the degradation of Nafion and of Aquivion

membranes. Degradation of Nafion 112 with (IEC)0 of 0.91 mmol g-1 and Aquivion, E870,

with (IEC)0 of 1.15 mmol g-1 are the subject of experimental studies. Considering the

chemistry of Aquivion membranes, their IEC corresponds to N = 14. βact is set to 0.01 for

all non-stabilized membranes. The rate constant of headgroup dissociation, ks, is the

free parameter. The results are shown in Fig. 3.10. Obtained ks values and degradation

conditions for each experiment are reported in Table 3.2.

0 10 20 30 40 50 60 70

0.80

0.85

0.90

0.95

1.00

0 10 20 30 40 50 600.6

0.7

0.8

0.9

1.0

1.1

1.2

1.3

1.4

ς

Time (h)

(IEC)0=1.43 mmol g-1 N=6

(IEC)0=1.11 mmol g-1 N=10

(IEC)0=0.91 mmol g-1 N=14

(IEC)0=0.77 mmol g-1 N=18

(IEC)0=0.66 mmol g-1 N=22

IEC

(mm

ol/g

)Time (h)

Page 75: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

57 57

Figure 3.10. Comparison of the kinetic model to experimental chemical degradation data of Nafion and Aquivion membranes [40,110,112].

19F NMR spectra of Aquivion membranes treated with �OH showed that loss of

SCF2 and the only O-CF2 group does not occur concurrently.112 Thus, unlike Nafion, both

SCF2 and O-CF2 cannot reside in a single grain. The headgroup of Aquivion only

consists of the sulfonate group and the rest of its sidechain forms the trunk. Within the

context of this model, Aquivion and Nafion membranes differ in terms of the number of

fluorine atoms released from their headgroup, 4 F atoms for Nafion and 2 F atoms for

Aquivion. The trunk length of both membranes will also be the same, N´ = 2. Within this

context, 3M will have the same headgroup as Aquivion with N´ = 4. Since experimental

data regarding the fluorine loss of 3M membranes are not available in the literature, the

model is only applied to Nafion and Aquivion membranes.

0 5 10 15 20 250

10

20

30

0 10 20 30 40 500

1

2

3

4

5

6

0 5 10 15 20 250

5

10

15

20

25

30

35

0 5 10 15 20 25 300

5

10

15

20

25

30

35

φ (m

g F- /g

pol

ymer

)

Time (h)

Ghassemzadeh et al. model

(d)

(b)

(c) Aquivion

φ (m

g F- /g

pol

ymer

)

Time (h)

Curtin et al. non-stabilized Curtin et al. stabilized model model

Nafion(a)

φ (m

g F- /g

pol

ymer

)

Time (h)

Ghassemzadeh et al. model

φ (m

g F- /g

pol

ymer

)

Time (h)

Merlo et al. model

Page 76: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

58 58

Table 3.2. Experimental condition and the obtained ks values. The degradation condition of each experiment was employed to obtain the �OH content utilizing the first block of the study.

Experimental condition Membrane ks (M-1 s-1) Reference 10 ppm Fe2+ and 20% H2O2 N211 4.4×104 Ghassemzadeh et al.107

2 ×10-4 (M) Fe2+ and 20% H2O2 N112 6.4×104 Ghassemzadeh et al.112

20 ppm Fe2+ and 30% H2O2 N112 1.5×103 Curtin et al.110

(DuPont’s data)

2 ×10-4 (M) Fe2+ and 20% H2O2 E870 8.8×104 Ghassemzadeh et al.112

36 ppm Fe2+ and 15% H2O2 E870 3.5×104 Merlo et al.40 (Solvay data)

We obtain ks = 1.5×103 M-1 s-1 from the experimental study of Curtin et al.110,

shown in Fig. 3.10b by open squares. The open diamond symbols in Fig. 3.10b refer to

the same experiment carried out with stabilized Nafion, in which the weak end group

content is reduced by 60% as compared to the original membrane (square symbols).

Thus, in order to obtain the grey curve in Fig. 3.10b, βact was set to 0.004 (40% of the

initial 0.01 content). Good agreement is observed between the model prediction and the

experimental observations. However, discrepancies between model results and

experimental data are seen for the last point reported in the time series. We conjecture

that this systematic deviation arises due to morphological effects in PFSA membranes.

Chemical degradation triggers the transformation of ionomer bundles82 and pore network

properties. Our analysis suggests that this coupling, which is not accounted for in our

model, slows down chemical degradation. Morphological features of PFSA membrane

as well as the interplay of chemical degradation and phase-separated structure will be

investigated in our future work.

Changes in IEC at the degradation condition of Ref 107 (10 ppm Fe2+ and 20%

H2O2) are simulated using ks = 4.4×104 M-1 s-1, as found from the analysis of the fluorine

Page 77: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

59 59

loss in Fig. 3.5. Fig. 3.11 shows a good agreement of the predicted IEC change with

experimental data. Fluorine loss and IEC change can thus be consistently reproduced.

Figure 3.11. Comparison between calculated IEC loss and experimentally measured data from Ref [107].

Values of ks for Nafion headgroup dissociation obtained from application of the

model to the experimental data are smaller than < 105 M-1 s-1. The obtained rate

constants for Nafion are close to each other, except the ks obtained from Ref 110. The

low ks from this study can be related to the lower detected fluorine release due to the

partial evaporation of F- and also the employed technique that is sensitive only to F-. The

liquid 19F-NMR technique employed in Ref 112 can measure the total fluorine content in

the exhaust water (F- as well as fluorine atoms in degradation fragments); they thus give

a more accurate measure of degradation.

Employing a kinetic approach based on spin trapping electron spin resonance

(ESR), Dreizler and Roduner186 obtained the rate constant smaller than 105 M-1 s-1 for

CF3-SO3- and (3.7±0.1) ×106 M-1 s-1 for CF3CF2OCF2CF2SO3

-. The extracted ks values

from Nafion and Aquivion fall within the range of the former model compound. The two ks

values obtained for Aquivion are in a matching range and also close to the ks values for

Nafion membranes. This similarity between the ks values obtained for these two

membranes can also suggest that �OH radicals attack the same site in these PFSAs.

0 10 20 30 40 500.76

0.78

0.80

0.82

0.84

0.86

0.88

0.90

0.92

IEC

(mm

ol/g

)

Time (h)

Ghassemzadeh et al. model

Page 78: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

60 60

3.7. Conclusion

In this study, we have developed a model of chemical degradation for PFSA-type

ionomers. In the first part, an analytical expression was derived that provides the radical

concentration as a function of iron content and hydrogen peroxide concentration. The

plot of �OH content over relevant ranges of H2O2 and iron content can be used to

quantify radical content at different conditions, encompassing both ex situ and in situ

evaluation. We have shown that between the level of iron content and H2O2

concentration, it is the level of iron contamination that mostly controls the �OH content.

Without addressing the morphological features of the ionomer, the kinetic model of

ionomer degradation distinguishes kinetic mechanisms that affect the amounts of

headgroup, trunk and backbone units of the ionomer. The model consistently describes

fluorine loss and IEC changes in ionomer membranes. The loss of headgroups has a

major impact on backbone scission and subsequent formation of new defective

backbone end groups. The efficiency of chemical stabilization therefore strongly

depends on the kinetics of headgroup loss. Increasing the trunk length lowers backbone

scission, which slows down chemical degradation. Increasing the length of backbone

segments lowers the rate of indirect loss of sidechains. Application of the model to

chemical degradation data of Nafion and Aquivion membranes revealed similarity of

hydroxyl radical attack to the headgroups of these membranes. The observed

discrepancy between the model prediction and simulation results highlights the

importance of morphological features on the transient degradation of ionomers.

The discussed results have different implications in terms of improving the

lifetime and durability of PEMs. The similarity of radical attacks to the PFSA materials

highlights the susceptibility of sulfonic acid moieties to OH radical attack. Thus, one

viable strategy in improving membrane chemical durability, besides adding radical

scavenging additives and lowering the iron content, can be replacing sulfonic acid

groups with more

�OH-resistance acidic units. Degradation studies focused on exploring the effect of

different ionomer properties such as side chain length and density on the membrane

chemical degradation would be beneficial in validating a series of model outcomes.

Page 79: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

61 61

DFT studies aiming at obtaining the activation barrier (Ea) of �OH attack to different

ionomer groups can help us in extending the model predictions to different temperatures.

The current model can be integrated with in situ models of PEM chemical degradation.

These are the models with more focus on the formation and transport of H2O2/radical

species into the PEM, normally combined with crude and simplistic membrane

degradation mechanisms. A potential future work can be the coupling of the degradation

model with a morphological model, which can possibly improve the quality of fittings and

result in more accurate predictions of membrane degradation. Absence of a coherent

morphological model was our motivation in undertaking a bottom-up approach in

understanding the basic structural forming processes of ionomer materials.

Page 80: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

62 62

Chapter 4. MD Simulation Study of Single Chain Ionomers

4.1. Abstract

In this chapter, we present the coarse-grained MD study of single ionomer chains

based on the research presented in Ref [189].1 The model polymer is comprised of a

hydrophobic backbone chain with grafted side chains that terminate in anionic

headgroups. The comb-polymer is modeled at a coarse-grained level with implicit

treatment of the solvent. The computational study rationalizes conformational properties

of the backbone chain and localization of counterions as functions of the side chain

length, the grafting density of side chains, the backbone stiffness, and the counterion

valence. The main interplay that determines the ionomer properties unfolds between

electrostatic interactions among charged groups, hydrophobic backbone interactions,

and steric effects induced by the pendant side chains. Depending on the branching

density, we have found two opposing effects of side chain length on the backbone

gyration radius and the local persistence length. A variation in the comb-polyelectrolyte

architecture is shown to have nontrivial effects on the localization of mobile counterions.

Changes in Bjerrum length and counterion valence are also shown to alter the strength

of Coulomb interactions and emphasize the role of excluded-volume effects on

controlling the backbone conformational behavior. The simulation results are in

qualitative agreement with existing experimental and theoretical studies. The

comprehensive conformational picture provides a fundamental framework for future

studies of comb-polyelectrolyte systems. 1 This chapter reproduces in revised form material from Ref [189], with permission from American

Chemical Society.

Page 81: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

63 63

4.2. Introduction

Polyelectrolytes are a class of polymers with a high fraction of periodically

attached ionizable groups.190,191 Polyelectrolyte solutions exhibit intriguing physical

properties and they find a wide range of applications as rheology modifiers and colloidal

stabilizers, adsorbent agents, biocompatible coatings, and biosensors.190–194

The structure of polyelectrolyte chains in aqueous solution is dictated by a subtle

interplay of electrostatic interactions between ionic species and hydrophobic interactions

of the polymer backbone with the solvent.193,195–199

In polyelectrolytes, the ionizable groups can be located within the backbone

(linear polyelectrolytes or ionenes) or grafted to it. Structural properties of linear

polyelectrolytes have been explored extensively in theoretical studies193,195,197,198 and

computer simulations.134,192,193,196,199–204 Graft polyelectrolytes, on the other hand, have

received generally less attention in studies of basic ionomer properties. In graft

polyelectrolytes, ionizable groups are either located along pendant side chains or at the

termini of the side chains. Many synthetic and biological polymers belong to the latter

group. One of the notable class of polyelectrolytes with comb-like or grafted architecture

is perfluorinated ionomers.1,22,27,91 Ionomers contain a smaller fraction of ionizable

groups compared to polyelectrolytes. Ionomer applications span from paint

suspensions91,205 to electrochemical energy applications25, 26,30,31,206–209. Ionomers are

synthetized both in acid and basic forms.25,68,210–212 Acid-type ionomers are the key

material in polymer electrolyte fuel cells (PEFCs).1,24,27,31 The basic forms are receiving

escalating recent attention as anion exchange membranes (AEM) in alkaline membrane

fuel cells.25,213

The variation in length and density of grafted side chains has been employed in

both PFSA and hydrocarbon ionomers to tune membrane properties such as water

sorption, mechanical and chemical durability, and ionic conductivity.40,214,215,26,27,68,216 The

Page 82: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

64 64

valency of counterion exerts a pronounced effect on mechanical and water sorption

properties of the ionomer membrane.217,218 A recently developed poroelectroelastic

model of water sorption in ionomer membranes relates these changes to the alteration of

the elastic properties of the polymer walls formed by self-aggregated ionomer

backbones.70 A coherent understanding of structure, properties and performance of

ionomer membranes is complicated by the multiple scales involved. The three main

levels (Fig. 4.1) are: (i) the molecular level comprising ionomer chains, ionomer bundles,

anionic charges and surrounding cloud of protons, (ii) the pore level at which elastic

properties of pore walls as well as proton distribution and mobility in the pore can be

defined, and (iii) the hydrated membrane level at which sorption characteristics, proton

conductivity and elasto-mechanical properties can be defined.76,219

Figure 4.1. Schematic illustration of three different levels of ionomer membranes.

Understanding basic structure and properties at the single chain level is of

paramount importance. From a practical standpoint, it determines the ionomer solution

behavior and defines how the structure and properties will evolve at the two higher

levels. However, single chain studies are scarce; mostly because of the practically

motivated focus on membrane-level investigations; but also due to spontaneous

aggregation of hydrophobic chains that complicates single chain investigations.

Recently, Minko and coworkers reported atomic force microscopy (AFM) imaging

of single-chain adsorbed Nafion chains on mica and highly ordered pyrolitic graphite

(HOPG) substrates.118 A compact globular structure was reported for adsorbed ionomer

chains.118 Theoretically, conformational features of single chains of Nafion-type ionomers

have been studied mainly through atomistic simulations. In the early 2000’s, Vishnyakov

et al.116,220 studied the effect of aqueous and alcoholic solvents on the conformation of

Effective properties (proton conductivity, water transport, stability��

hydrophobic phase

hydrophilic phase

Primary chemical structure •  backbones •  side chains •  acid groups

Molecular interactions (polymer/ion/solvent), persistence length

Self-organization into aggregates and dissociation

Secondary structure •  aggregates •  array of side chains •  water structure

“Rescaled” interactions (fluctuating sidechains, mobile protons, water)

Heterogeneous PEM •  random phase separation •  connectivity •  swelling SingleChain Porelevel Membranelevel

Effective properties (proton conductivity, water transport, stability��

hydrophobic phase

hydrophilic phase

Primary chemical structure •  backbones •  side chains •  acid groups

Molecular interactions (polymer/ion/solvent), persistence length

Self-organization into aggregates and dissociation

Secondary structure •  aggregates •  array of side chains •  water structure

“Rescaled” interactions (fluctuating sidechains, mobile protons, water)

Heterogeneous PEM •  random phase separation •  connectivity •  swelling

Page 83: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

65 65

Nafion oligomers. Recently, DFT combined with electron energy-loss spectroscopy

(EELS) was employed to study the effect of different side chain chemistries on the

conformational features of single PFSA backbone chains.221 Computational limitations

associated with the incorporation of atomistic details along with ill-defined simulations of

inadequately short oligomeric chains have hindered drawing definite conclusions

regarding the interplay of side chain length and density on the conformational features of

comb-like ionomer molecules.

Coarse-grained MD simulations, in which chemical details are foregone for the

sake of increasing the time and length scale of simulations, have been able to capture

structural changes in charged polymers. Changes in the radius of gyration of the

backbone and in the persistence length of single linear polyelectrolyte chains have been

explored in relation to variations in chain rigidity222,223, charge fraction10,225,47,49,50, and

dielectric properties of the solvent.10,202,19,20,49 For branched polyelectrolytes with charged

side chains, conformational properties of the hydrophobic backbone have been studied

as a function of the length and density of the side chains.224,226 In neutral polymers with

pendant side chains (bottle-brush or comb-polymers) it is known that steric hindrances

due to side chain congestion increase the radius of gyration and stiffness of the

backbone.227–229 For comb-polyelectrolytes with charged moieties at side chain terminals

more complex conformational responses are expected due to the presence of both

electrostatic and excluded-volume interactions. To the best of our knowledge, there have

not been any attempts in developing a coherent picture of the conformational properties

of comb-like polyelectrolytes with hydrophobic backbone and charge-terminated side

chains.

This chapter strives to rationalize the backbone conformation and the local

rigidity of single chains of comb-polyelectrolytes as a function of changes in chemical

structure and environmental parameters. Section 4.3 introduces the coarse-grained

chain model, the type of interactions considered, and the simulation protocol. Section 4.4

provides a systematic evaluation of backbone conformational changes as a function of

the variation in pendant side chain length and density, strength of electrostatic

interactions, and counterion valence.

Page 84: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

66 66

4.3. Chain model

In a coarse-grained model, the interacting species are groups of atoms or beads

that interact via effective forces. We model a bead-spring polyelectrolyte chain

consisting of a hydrophobic backbone and pendant side chains in implicit solvent with

explicit counterions. Fig. 4.2 is a schematic illustration of the branched polyelectrolyte

chain considered in the model.

Figure 4.2. Schematic illustration of comb-polyelectrolyte architecture. Backbone and side chain hydrophobic beads are shown in blue and green, respectively. Red beads represent anionic headgroups and yellow beads represent the counterions.

The model was devised to emulate PFSA-type ionomers. Each backbone bead

resembles one (CF2-CF2) repeating unit with diameter of approximately 2.4 Å. The

simulated backbone chain consists of 104 beads with regularly spaced pendant side

chains, representing a real backbone contour length of ~ 25 nm. Anionic headgroups

carry a -1e charge and they are represented by a single bead (red) with the same size

as backbone and side chain beads. Ls denotes the number of beads in each side chain,

and Nb is the number of backbone beads per ionomer repeat unit (containing 1 side

chain), as illustrated in Fig. 4.2. We explore values Ls = 1, 3, 5, and 7 and Nb = 1, 2, 3, 5.

Since all side chains contain a charged unit, we define a charge fraction f = 1/Nb, for

which we will consider values f = 1, 1/2, 1/3, 1/5, with the lower range resembling

ionomers and the upper range corresponding to bottle-brush (f = 1) and polyelectrolyte

systems (f = 1/2, 1/3).

Page 85: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

67 67

Electroneutrality is guaranteed by adding Nc = Ns/Zc counterions (yellow beads in

Fig. 4.2) where Zc is the counterion valence and Ns is the total number of side chains. No

other electrolyte is added to the solution system. As a reference, we simulated all comb-

like architectures in their uncharged form in electrolyte-free solution, for all considered

combinations of f and LS. As a simplification, all beads are assumed to have the same

diameter σ. Adjacent monomer beads in the backbone and side chains are connected by

the FENE potential with spring constant k = 30 ε/ σ2 and maximum extension

of R0 =1.5 σ.130,140,230 The flexibility of the backbone is modeled with the bond angle

potential of Eq. 2.4. We considered values kθ = 0, 15, 30, 45 kBT for the backbone chain.

Pendant side chains are freely rotating. Van der Waals interactions between monomer

beads are modeled through a standard LJ potential of Eq. 2.2. The cut-off distance was

set to rc = 2.5 σ, for interactions between uncharged polymer beads (backbone and side

chain beads), and rc = 21/6 σ for all other pairwise interactions (neutral-anionic,

neutral-cationic, anionic-anionic, anionic-cationic), resulting in the purely repulsive part of

the LJ potential. The reference value of the Lennard-Jones interaction parameter is set

to εLJ = 1 kBT. The units of energy, diameter and mass of all polymer beads are set to

unity, i.e. kBT = 1, σ = 1, and m = 1. The θ solvent condition for the polymer backbone

corresponds to εLJ = 0.33 kBT.131,134 The reference value of εLJ for interactions between

uncharged beads was set to 1.0 kBT, representing an “intermediate” level of

hydrophobicity. For interactions between charged beads, we used Eq. 2.5. The baseline

value of the Bjerrum length is λB = 3.0 σ, which corresponds to about 0.7 nm in aqueous

electrolyte solution (σ ≈ 0.24 nm). In our parametric analysis, we explore effects of LS, f,

λB, ZC, and kθ. Table 4.1 summarizes the baseline values and parameter ranges

explored.

Page 86: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

68 68

Table 4.1. List of system parameters and their baseline values along with the explored range.

Parameter Baseline value

Parameter range explored in this work

Parameter value for Nafion

LS 3 1-7 3

f 1/2 1 - 1/5 1/5

λB 3.0 σ 0.5 - 12 σ 3.0 σ

0 kBT 0 - 45 kBT 25 kBT

ZC 1 e- 1 - 3 e- 1 e-

The particle–particle particle–mesh (PPPM)144 was used for the calculation of

electrostatic interactions. Simulations were carried out in the NVT ensemble with

periodic boundary conditions in all three directions for a cubic simulation box with length

L = 110 σ. The temperature was maintained by a Langevin thermostat.126

The following procedures were employed in simulations. A single polyelectrolyte

chain was placed in the middle of the simulation box and simulated for 7×106 time steps.

The first 5.0×106 timesteps were used for equilibration, tracked through convergence in

the total energy and the radius of gyration. For equilibrium state analysis, data

collections were performed for the last 2×106 timesteps and data were saved every

2×104 timesteps, resulting in 100 regularly time-spaced snapshot. Computer code

developed for generating the data file of ionomer chain with various structures is

provided in Appendix B.

Before presenting, simulation results, a remark on the implicit solvent treatment

is in order. The effects of solvent molecules are implemented through the friction term of

the Langevin thermostat. This specific treatment, which is a common practice in coarse-

grained MD simulations of polyelectrolyte systems, implies the absence of explicit

solvent molecules. Absence of these molecules lowers the computational costs and thus

longer simulation runs become practically possible, allowing for a proper equilibration of

the system. However, as argued in Refs. [202,231,232], the absence of many-body

Page 87: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

69 69

effects caused by solvent molecules can significantly influence the collapse and

conformational dynamics of both neutral and polyelectrolyte solutions under poor solvent

conditions. Since we aim at capturing the equilibrium conformational features of long

comb-polyelectrolyte chains and not the relaxation dynamics of our model system, the

use of implicit solvent molecules does not decisively alter the reported conformational

behavior.

We have also assumed a uniform polarizability of the implicit solvent by fixing the

values of the dielectric constant. This assumption implies that the electrostatic

interaction among any pair of charged groups depends only on their separation

distances. In reality, local charging effects result in polarization of solvent molecules and

alteration of local solvent dipole moments. This implies subtle variations in dielectric

constant and spatially dependent electrostatic interactions. Solvent polarization and local

dielectric effects can be introduced to the studied system either through incorporation of

explicit solvent molecules or by developing polarizable coarse-grained solvent

models.225,233 Even though the consideration of polarization effects yields more realistic

results for local conformational properties and needs to be the focus of future studies,

the imposed computational complexity shortens the equilibration time period. For the

sake of simulating long comb-like chains and capturing the global conformational trends,

it seems that consideration of uniform dielectric media is a legitimate assumption.

4.4. Results and discussion

4.4.1. Effects of side chain length and density

As discussed above, the conformational behavior of a branched polyelectrolyte

chain originates from both steric effects due to branching and the electrostatic

interactions of charged units. To be able to discriminate both of these effects, we

simulated grafted polymers in both their charged and neutral states (UCoul = 0). We used

the backbone gyration radius, Rg (σ), as the main conformational descriptor, which is

defined as

Page 88: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

70 70

Rg2 =

1N

ri − rCOM2

i=1

N

∑ , rCOM =1N

rii=1

N

∑ , (4.1)

where i is the backbone bead index, N is the number of backbone beads, ri is the

position vector of the ith bead, and rCOM the center-of-mass (COM) of the chain. Angular

brackets indicate averaging over chain configurations.

Fig. 4.3 shows the results for Rg obtained in the case of freely rotating chains.

Zero bending rigidity implies that all conformational changes stem from excluded-volume

and electrostatic effects. Fig. 4.3a shows that increasing LS and f increases Rg. These

effects become more prominent as f increases. No significant change in Rg with LS is

observed for f = 1/5. The increase in backbone Rg upon increasing LS and f for the

neutral polymer is due to enhanced excluded-volume interactions of side chain beads.

This trend reflects the classical behavior that was previously captured through extensive

Monte Carlo (MC)227–229,234 and MD224,235 simulations.

Page 89: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

71 71

Figure 4.3. Backbone Rg for (a) neutral comb-like polymers, (b) comb-polyelectrolyte chains, and (c) pure electrostatic contribution to the backbone Rg. All the error bars in this chapter represent the standard deviations obtained over the equilibrium trajectories.

Fig. 4.3b shows the changes in backbone Rg for comb-polyelectrolyte chains in

the presence of monovalent counterions at λB = 3 σ. For comb-polyelectrolytes, higher f

increases the backbone Rg for all LS values explored. Higher density of side chains

results in closer proximity of charged units and stronger electrostatic repulsion that

causes greater stretching of the backbone. Charged polymers maintain larger Rg values

Page 90: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

72 72

compared to the uncharged equivalent polymer, clearly demonstrating that eliminating

the charges leads to smaller and more compact conformations.

Regarding the effect of the side chain length, two different trends are observed

depending on the fraction of side chains. At f = 1, increasing LS increases Rg, while f ≤

1/2 results in the opposite trend. The change in LS induces a transition in the interplay of

excluded-volume and Coulombic interactions. Increasing LS leads to a larger distance

between charged headgroups and correspondingly weakened electrostatic repulsion,

which favors a reduction in Rg. On the other hand, higher LS increases the steric

congestion of side chains, which favors a stretching of the backbone chain, to increase

Rg. At f = 1, the steric effect of side chain congestion outweighs the electrostatic effect

and the increase in LS leads to enhanced Rg.

The milder increase in Rg at f = 1 for the charged chain as compared to an

uncharged chain reflects the opposite influence of increasing LS on the chain size. For

f ≤ 1/2 excluded-volume effects are weakened due to the greater separation of pendant

side chains and the electrostatic interactions among anionic headgroups exert the

dominant effect on Rg. Weakened Coulombic repulsion at higher LS lowers the

electrostatically induced stretching of the backbone and Rg decreases.

The decrease in backbone size upon increasing the length of side chains was

also experimentally reported for polypeptide chains bearing longer pendant side

chains.236–238 Fig. 4.3c shows the pure electrostatic contributions to the backbone Rg

once the Rg values in Fig. 4.3a, which account for excluded-volume effects, are

subtracted from Rg values presented in Fig. 4.3b. There is a noticeable difference in the

pure electrostatic contribution to Rg between short and long regimes of LS for all f values.

This contribution is more pronounced for regimes of small LS values. The trend seen in

Fig. 4.3c shows that for short side chain polymers, coulombic effects prevail. This effect

is owed to the closer proximity of anionic moieties at smaller LS values that strengthen

the electrostatic repulsion among charged groups. The electrostatic contribution

continuously diminishes as LS increases and the excluded-volume effect gains weight in

determining Rg.

Page 91: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

73 73

Fig. 4.4 shows the snapshots of the comb-polyelectrolyte chains at different LS

and f values. The conformational behavior observed in Fig. 4.3b is in line with the

conformational trends in Fig. 4.4. For f = 1/5 we can see the dominant effect of LS on the

size of the chain: while chains with LS ≥ 3 exhibit a globular conformation, chains with

LS = 1 assume an expanded coil-like conformation. For f = 1, increasing LS results in the

transformation of an expanded chain at LS = 1 to a rod-like chain at LS = 7. Increase in

the fraction of pendant side chains results in the monotonic expansion of collapsed

chains. All-atomistic200 and coarse-grained225 MD simulation studies of sodium

polystyrenesulfonate (NaPSS) in water bearing different fractions of anionic-terminated

branches demonstrated a similar conformational trend.

Figure 4.4. Equilibrated comb-polyelectrolyte chains at different values of f and LS. Color-coding is the same as Fig. 4.2.

Conformational structures obtained for f < 1/5 are indistinguishable from f = 1/5

and are not discussed in this work. The globular structure observed for f = 1/5 and LS ≥ 3

is in line with the globular structure of a single chain Nafion ionomer reported in Ref

Page 92: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

74 74

[118]. Distally located branches weaken the steric congestion of side chains and lower

the coulombic anion-anion charge repulsion. Minimization of the surface energy of the

hydrophobic polymer triggers the collapse of the ionomer chain to form a globular

structure. Fig. 4.5 shows the effect of different bending rigidities, kθ, on Rg of both

charged and uncharged comb-polymer for the explored range of f and LS. The value

f = 1/2 closely resembles the charge fraction of the coarse-grained Nafion models

employed by Malek et al239,240. Imposing bending rigidity to the backbone chain

continuously increases Rg for both the charged and uncharged polymers.

Figure 4.5. Effect of kθ on backbone Rg. Comb-polyelectrolyte: (a) LS = 3 (b) f = 1/2. Neutral counterpart: (c) LS = 3 (d) f = 1/2. All the plots share the legend of (a).

More rigid chains exhibit a smaller sensitivity of Rg to the structural changes for

both charged and uncharged polymers. For very stiff chains, Rg is nearly independent of

f and LS. Increased backbone rigidity introduces an energetic barrier for the

conformational changes induced either by topological effects (excluded-volume effect) or

by electrostatic repulsions.

Page 93: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

75 75

In what follows, the effect of changes in length and density of side chains on the

local orientational correlation or local persistence length, LP, of the backbone chain is

explored. The local persistence length is the scalar projection of the end-to end vector,

Rb, onto each bond in the backbone, rb, which is defined by224,227

Lpk =

rb!"

rb!"! •Rb!"!

,

(4.2)

where k is the backbone monomer or bead index and the angular brackets, <...>, denote

averaging over chain configurations. Fig. 4.6 shows the results for LP as a function of

monomer index, k, along the backbone chain for neutral and charged comb-polymers.

For both charged and uncharged polymers, greater f results in higher LP.

Page 94: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

76 76

Figure 4.6. Local persistence length for the comb-polyelectrolyte and the neutral branched polymer at different side chain fractions and lengths.

Backbones of a comb-polyelectrolyte chain maintain a greater local rigidity than

the uncharged polymer due to charge-induced chain stiffening. Increasing LS results in a

monotonic increase in LP at all f values for neutral polymers and at f = 1 for the comb-like

polyelectrolytes. Increasing LS increases the backbone flexibility at f ≤ 1/2 for charged

chains. No special features are observed in LP for f = 1/5 at LS > 1 in both uncharged and

charged chains. In the regime of low side chain content, LP values are close to zero,

which suggests that the entire backbone is in a compact globular conformation.

The LP values of f = 1, 1/2, 1/3 for the charged polymer, shown in Fig. 4.6, reveal

a curving of LP upon increase in LS. This curvature arises from lower rigidity at both ends

of the backbone chains. The monomers at the ends of a chain experience reduced steric

Page 95: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

77 77

repulsion. Shorter branching, i.e., smaller LS, yields more homogenous backbone rigidity

and diminished end effects as the long-range Coulomb forces gain in importance.

Overall, smaller LS and larger f result in greater rigidity of the backbone chain for f ≤ 1/2.

These two main features, smaller LS and higher f, resemble the key characteristics of

short side chain (SSC) ionomers compared to longer side chain ionomers such as

Nafion. At the single-chain level, SSC ionomers are expected to maintain greater rigidity

than their longer side chain counterparts.

4.5. Localization of counterions

We explore how the length and fraction of pendant side chains as well as the

local rigidity of the backbone influences the localization of monovalent counterions.

Fig. 4.7a shows the radial distribution function (RDF) of anionic headgroups and

counterions, gh-c(r) for LS = 3 at various f. For an isotropic system, an RDF gA-B(r)

describes the probability of finding a particle A in a specific radial distance r of a

reference particle B. It is defined as124

gA−B r( ) =V

4πr2NANB

δ r − rij( )j=1

NB

∑i=1

NA

∑ , (4.3)

where NA and NB are the numbers of A and B particles in the simulated system,

respectively, and the triangular brackets around the δ-function indicate averaging over

the equilibrium system trajectory. Greater f increases the surface charge of the chain

and creates a stronger surface potential, thereby causing greater coulombic attraction of

the counterion, a trend that was found in other simulation studies as well.200,225,233

Fig. 4.7b demonstrates the effect of LS on the localization of counterions for f = 1/2.

Counterions have a stronger affinity to anionic headgroups for ionomers with shorter

side chains.

The enhanced surface charge density at lower LS is responsible for the stronger

surface potential and greater localization of counterions. The changes in the localization

of counterions in the vicinity of anionic headgroups as a function of f and LS can also be

Page 96: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

78 78

seen in the snapshots depicted in Fig. 4.4. Fig. 4.7c shows how the intrinsic rigidity of

the backbone chain influences the localization of counterions to anionic groups for

f = 1/2 and LS = 3.

Figure 4.7. Radial distribution functions between anionic headgroups and counterions for (a) f = 1-1/5 and LS = 3 (b) f = 1/2 and LS = 1-7 (c) f = 1/2, LS = 3, and kθ = 0 - 45 kBT.

Affinity of counterions to the headgroups decreases as the bending rigidity

increases. Greater kθ results in more stretched backbone chains with farther separated

side chains. This greater distance among charged moieties results in lower effective

surface charge of the chains and thus weaker electrostatic attraction to the counterions.

Counterion localization exhibits a smaller sensitivity to kθ than to f and LS. These

observations highlight the nontrivial effects of architectural variation in comb-

polyelectrolyte chains on the distribution and possibly the dynamics of mobile

counterions.

Page 97: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

79 79

4.6. Effect of electrostatic interaction strength

In this section, we explore how the change in the strength of electrostatic

interactions introduced through the variation in λB influences Rg. Fig. 4.8a shows Rg

plotted over λB for f = 1/2 and for different LS. Several trends are noticeable in this plot.

First, a nonmonotonic dependence of Rg on λB is evident for all LS. Secondly, the

sensitivity of the nonmonotonic dependence decreases as LS increases.

Figure 4.8. Rg values as a function of Bjerrum length (a) f = 1/2 and LS = 1-7 (b) f = 1/2, LS = 3, and kθ = 0 - 45 kBT.

A nonmonotonic dependence of Rg on λB was reported in Refs.[134,204,222,223]

for linear polyelectrolyte chains. It was attributed to the initial expansion of the chain due

to enhanced electrostatic repulsion upon increasing λB. Further increase in λB results in

Page 98: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

80 80

strong localization of counterions in the vicinity of the chain that screens anionic charges

and lowers their effective electrostatic repulsion. In the regime of large and small λB,

i.e., λB = 0.5 σ and λB ≥ 9 σ, side chains with larger LS assume a greater radius of

gyration. This behavior can be rationalized through the dominance of excluded-volume

effects in the case of weak or screened electrostatic interactions. In the midrange of λB =

1 to 6 σ, enhanced Coulombic repulsions along with shorter distance between the

charged units of chains with small LS results in higher Rg. Thus, we can see that

increasing the side chain length does not always result in enhanced chain size and this

behavior is sensitive to the strength of electrostatic interactions. Increasing LS results in

weaker sensitivity of Rg to λB, which can be attributed to the reduced impact of Coulomb

interactions.

Fig. 4.8b shows how comb-polyelectrolytes (f = 1/2, LS = 3) with different

backbone rigidity behave upon variation of λB. We can see the nonmonotonic

dependence for chains with different kθ values. However, the sensitivity to the change in

Bjerrum length, and the extent of chain expansion and shrinkage, diminishes for more

rigid backbones. Ou et al.223 also observed similar behavior and reported that increasing

the backbone rigidity diminishes the dependence of Rg on λB.

4.7. Effect of counterion valency

The results for the change in Rg upon change in ZC for chains of different f and LS

are shown in Fig. 4.9. Fig. 4.9a shows the change in Rg of comb-polyelectrolytes with

different fraction of side chains at fixed LS = 3. Increasing the valency of counterions,

from ZC = 1 to ZC = 2 and further to ZC = 3 leads to a decrease in Rg. Increasing the

counterion valence yields stronger correlation of counterions with charged headgroups.

Enhanced counterion localization triggers an effective charge screening that results in

decreased electrostatic repulsion. Weakened Coulomb interactions along with attractive

short-range hydrophobic effects causes the comb-polyelectrolyte chain to collapse to a

neutral-like structure. The reduction in Rg is more pronounced for chains with higher f,

owed to the prevailing electrostatic contribution to Rg at high f (Fig. 4.3c). Fig. 4.9b

shows the effect of an increase in ZC for f = 1/2 and different LS.

Page 99: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

81 81

Figure 4.9. Values of square root of backbone radius of gyration at different ZC (a) f = 1-1/5 and LS = 3 (b) f = 1/2 and LS = 1-7.

The extent of the decrease in Rg is more pronounced with ZC upon decreasing

LS. At ZC = 1, shorter branching yields greater Rg due to enhanced Coulombic repulsion.

This trend is opposite of that observed at ZC = 2, 3 where excluded-volume effects

primarily control the backbone Rg and higher LS yields larger backbone Rg.

4.8. Conclusion

The presented coarse-grained molecular dynamics study scrutinizes the effects

of different parameters on the conformational properties of hydrophobic comb-like

ionomers. Observed trends can be rationalized through the fine interplay of hydrophobic

Page 100: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

82 82

interactions of individual backbone units, excluded-volume effects of side chains, and

Coulombic interactions among charged moieties. It is observed that increasing the

fraction of pendant side chains increases the radius of gyration and rigidity of comb-

polyelectrolyte chains, a trend similar to the neutral counterpart. For high density of side

chains, increasing the length of side chains leads to backbone stretching. For lower

density of side chains, increasing side chain length yields a smaller backbone radius of

gyration, resulting in the formation of collapsed globular structures. Counterions possess

stronger affinity to anionic moieties for shorter side chains and higher grafting densities.

Chain radius of gyration is shown to demonstrate a nonmonotonic dependence on the

strength of electrostatic interaction. In the regime of weak Coulomb interaction steric

effects of side chains control the radius of gyration and the backbone conformation.

Multivalent counterions are shown to decrease the radius of gyration through the

screening of anionic charges. Qualitative agreement in the conformational trends with

both experimental and similar theoretical studies highlights the ability of the presented

coarse-grained model to capture the subtle conformational behavior of comb-

polyelectrolytes and ionomer chains. Conformational understandings at the level of

single chain are particularly essential for characterization of ionomer solutions and the

underlying process of self-assembly.

The single chain study provides us with an unprecedented knowledge of the

interrelation between the structural parameters of ionomer chains and their

conformational behavior. Future simulation studies can focus on parameterizing the

force field interactions through the first principle DFT calculations. Importantly, there is a

need for more systematic experimental studies on the single ionomer chains in order to

validate the majority of the observed trends.

Page 101: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

83 83

Chapter 5. MD Simulation Study of Ionomer Aggregation

5.1. Abstract

In this chapter we present the results of a coarse-grained MD study of ionomer

self-assembly, which has been discussed in detail in Ref [241].1 These MD simulations

elucidate the self-assembly of semiflexible ionomer molecules into cylindrical bundle-like

aggregates. Ionomer chains are comprised of hydrophobic backbones, grafted with

pendant side chains that are terminated by anionic headgroups. Bundles have a core of

backbones surrounded by a surface layer of charged anionic headgroups and a diffuse

halo of counterions. Parametric studies of bundle properties unravel the interplay of

backbone hydrophobicity, strength of electrostatic interactions between charged

moieties, side chain content, and counterion valence: expectedly, the size of bundles

increases with backbone hydrophobicity; the aggregate size depends nonmonotonically

on the Bjerrum length; increasing the grafting density of pendant side chains results in

smaller bundles; the counterion valence exerts a strong effect on bundle size and

counterion localization in the near-bundle region. Results reveal how ionomer

architecture and solvent properties influence ionomer aggregation and associated

electrostatic and mechanical bundle properties. These properties of ionomer aggregates

are vital for rationalizing water sorption behavior and transport phenomena as well as

the chemical and mechanical stability of ionomer membranes

1 This chapter reproduces in revised form material from Ref [241], with permission from American

Chemical Society.

Page 102: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

84 84

5.2. Introduction

In low-polarity solvents,91,205 association of acidic headgroups in ionomer

molecules occurs due to dipolar interactions leading to the formation of ionic multiples.

In high-polarity solvents, acidic groups dissociate completely resulting in two distinct

categories of ionomer behavior based on the strength of polymer-solvent interactions.

Under good solvent condition, for the polymer backbone, i.e. a solvent that dissolves the

backbone chains, solvated ionomer molecules exhibit polyelectrolyte-type behavior,

forming loose ionic aggregates. In sufficiently polar solvents, the ionomer backbones

undergo aggregation and phase separation. This process results in the formation of a

solid PEM at high ionomer concentration.1

Studies of the structure of Nafion and assorted materials have primarily focused

on the solid membrane state (with water volume fraction significantly below 50%). Using

a large variety of experimental techniques, a suite of structural models have been

proposed, including parallel cylinder model,79 cluster-channel model,77,242 lamellar243,244

or skin-type model,72 and rod network model.71,94 However, none of these simplified

modelistic views captures the actual membrane structure. From a fundamental

perspective, solution studies of ionomer aggregation are highly insightful in this regard.

Based on SAXS, SANS, ESR and 19F NMR experiments of ionomer solutions,81–

83,92,93,245 it was conjectured that hydrophobic backbones of Nafion form the core of self-

assembled ionomer structures while ionic units are located at the periphery of this

configuration. Jiang et al.81 employed dynamic light scattering (DLS) to study self-

assembly of Nafion chains in dilute aqueous solution. Observation of stable aggregates

supported the fringed-rod ionomer model proposed by Szajdzinska-Pietek et al.84,85

These structures are similar in nature to the cylindrical-like micelles formed in solutions

of semirigid synthetic polyelectrolytes such as sulfonated poly (p-phenylene).246–248

Loppinet, Gebel and others82,83,119 studied the influence of various parameters on the

self-assembling behavior of ionomer solutions. They observed that the dielectric

constant of the solvent and the grafting density of pendant side chains influence the size

of ionomer bundles. Using a combination of scattering and microscopy, Gebel and

Page 103: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

85 85

Rubatat71,94,95 provided evidence of fibrillar (rod-like) aggregates in well-hydrated PEMs.

The proton conductivity and mechanical strength of solution-cast membranes, formed

from different solvents, were correlated with the properties of polymeric bundles formed

in Nafion solutions.249 Rod-like bundles have also been reported for hydrated sulfonated

polyphenylene ionomers.250 Recent studies identified aggregates of Nafion ionomers as

key structure-forming elements in skin-type ionomer films, forming during self-assembly

in ink mixtures of PEFC catalyst layers.88,239,251

Obviously, relatively stiff bundle-like aggregates of ionomer backbones control

the properties of solid state PEMs and CLs;1 their presence has been seen to affect

water sorption behavior,70 proton conductivity,252–254 and chemical durability.147,154 Thus,

understanding the self-assembly of ionomers in dilute solution seems invaluable in

gaining a deep insight into fundamental structure and transport properties of ionomer

membranes. This insight in turn could spur the chemical design of ionomers with tailor-

made properties.

The majority of simulation studies of ionomers can be categorized as either

melt206,207,255 or hydrated membrane simulations.117,218,256–264 Formation and

morphologies of ionic aggregates in ionomer melts as a function of counterion valence

and polymer architecture have been studied with coarse-grained MD

simulations.206,207,255 Hydrated membrane studies have primarily focused on microphase-

separated morphologies in ionomer–water mixtures of Nafion. Atomistic molecular

dynamics (MD) simulations27–32,34,37 have served as the focal point of membrane

simulation studies. Other approaches such dissipative particle dynamics (DPD)

simulations,240,261,264 and mean-field approaches61 have helped to expand the time and

length scale of atomistic simulations.

Self-assembly of charged polymers is a phenomenon of general

interest.131,132,208,239,255,261,265–277 Aggregation is often observed in biological systems265,278

and highly-charged polyelectrolytes.273,279 Examples include fibrillar proteins275,276

(F-actin) and DNA strands.280,281 In these systems, self-charge attraction due to

condensation of multivalent counterions is considered to be the driving force of

Page 104: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

86 86

aggregation. Atomistic and coarse-grained MD simulations have been used to study self-

assembly in solutions of peptide amphiphiles,282 polyelectrolyte brushes,272,283 protein-

like polymers,269 as well as preformed bundles of “Hairy-rod” polyelectrolytes such as

poly(para-phenylene) (PPP).132,270,284

Theoretical and simulation studies of self-assembly in dilute ionomer solutions

are comparatively rare in the literature. A recent theory focused on the interplay of

electrostatic and hydrophobic interaction strengths in determining size and stability of

cylindrical bundles of uniformly charged and closely packed rigid rods through a mean-

field approach.90

Here we explore self-assembly of ionomers in dilute solution through coarse-

grained molecular dynamics simulations. Section 5.3 introduces the coarse-grained

ionomer model, the type of interactions considered, and the simulation protocol. Section

5.4 provides a systematic evaluation of ionomer bundle properties as a function of

strength of hydrophobic interactions (solvent quality), strength of electrostatic interaction,

ionomer side chain content, and counterion valence. Key findings are discussed in

Section 5.5 within the context of experimental studies of ionomer solutions and polymer

electrolyte membranes.

5.3. Ionomer model

The task at hand is to devise a coarse-grained model that retains essential

chemical and physical characteristics of the system. Fig. 5.1 illustrates the coarse-

grained ionomer architecture used in the present study.

Page 105: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

87 87

Figure 5.1. In the chemical structure of Nafion ionomer in (a), n corresponds to 6 CF2-CF2 repeating units between sidechains. Color coding is the same as Fig 4.2.

Unlike the model chain employed in the Chapter 4 that had 104 repeating units,

the backbone chain utilized in this chapter contains 48 backbone beads.

Since aggregation studies involve an ensemble of ionomer chains, lowering the chain

size was essential in reaching proper equilibration. Regular spacing of side chains is

assumed in the coarse-grained ionomer model, representing a reasonable assumption

for Nafion-type ionomers, as discussed in Refs. [29] and [242] and references therein.

A backbone segment with one grafted side chain consists of Nb apolar beads. As a

reference case, we used Nb = 7 to resemble the chemical structure of typical Nafion

ionomer with 7 CF2-CF2 units per backbone segment corresponding to one side chain

including the junction site. The backbone strand for the reference case has 6 pendant

side chains affixed to it (Ns = 6). The length of a Nafion trunk (excluding the anionic

headgroup) is around 0.8 nm.285 For the baseline case, we therefore considered side

chains consisting of 3 apolar beads (green beads in Fig. 5.1), giving a length of the

hydrophobic side chain segment of about 0.8 to 1.0 nm. We assumed complete

dissociation of acid headgroups and assign a charge of -1e to the terminal bead.

Electroneutrality of the system was achieved by adding Nc = NsM/Zc counterions, shown

as yellow beads in Fig. 5.1, to the system, where Zc is the counterion valence and M is

the total number of ionomer chains. Initially, we considered monovalent counterions with

Zc = +1e, e.g., corresponding to hydronium ions, H3O+, for a protonated ionomer system.

The FENE bond potential (Eq 2.3) with the same spring constants of Chapter 4.3

was used to model the bonding of ionomer beads. The flexibility of ionomer backbone

and side chains was modeled using the bond angle potential of Eq. 2.4. In Eq. 2.4, is kθ

Page 106: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

88 88

set to 180º for linear segments along the backbone and side chains and to 90º at the

branching points. For non-bonded bead-bead interactions, we employed the LJ potential

of Eq. 2.2. The implicit solvent approach used in this study implies the absence of

explicit interactions between polymer and solvent molecules. In this approach,

hydrophobicity of ionomer groups in the backbone and the side chain units is embodied

in the Lennard-Jones interaction parameter between apolar beads, εLJ, which must be

sufficiently attractive to adequately represent so-called “poor-solvent”

conditions.131,132,134,201,226

The cut-off distance was set to rc = 2.5 σ for interactions between apolar polymer

beads in backbone and side chains. The value of εLJ was varied over the range specified

in Table 5.1. For all other pairs of beads (apolar-anionic, apolar-cationic, anionic-anionic,

anionic-cationic), we used rc = 21/6 σ and εLJ = 1 kBT. The value of εLJ for polymer-polymer

interactions was varied between 0.5 kBT and 1.5 kBT. The baseline value εLJ = 1 kBT

represents “normal hydrophobic” behavior, while εLJ = 0.5 kBT and εLJ = 1.5 kBT

correspond to “weak” and “strong” hydrophobicity.134 The θ solvent conditions

correspond to εLJ = 0.33 kBT.131,134 Charged beads interact via the direct Coulomb

potential of Eq. 2.7. The baseline value of the Bjerrum length is λB = 3.0 σ. Table 5.1

summarizes the baseline values and the range of variations of the model parameters εLJ,

λB, Ns, and the counterion valence ZC.

Table 5.1. List of system parameters and their baseline values along with the explored range.

Parameter Baseline value Parameter range explored in this work

εLJ 1 kBT 0.5-1.5 kBT

λB 3.0 σ 1-12 σ

NS 6 3 - 12

ZC 1 e- 1- 2 - 3 e-

Page 107: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

89 89

5.3.1. Computational details

The particle–particle particle–mesh (PPPM)144 with accuracy of 10-5 was used

for the calculation of electrostatic interactions. Simulations were carried out in the NVT

ensemble with periodic boundary conditions in all three directions. A cubic box of side

length L = 110 σ was chosen that is large enough to avoid finite size effects.

Temperature was maintained through coupling of the system to a Langevin

thermostat.126

The simulation protocol was as follows. A single stiff ionomer chain was placed in

the middle of the simulation box and simulated for 2×106 time steps. The resulting chain

was then replicated and distributed on a regular 3D grid in the simulation box. The

system was run for 3 million timesteps under a fully repulsive short-range potential. This

procedure resulted in random and homogeneous distribution of chains in the simulation

box. Our ensemble of M = 25 chains corresponds to a number density of beads of ρ =

MN/L3 = 1.3×10-3 σ-3 where N is the total number of beads in each chain. This density

should be compared to the overlap monomer density of ρ* = N/(4/3π Rg3) where Rg is the

radius of gyration for a single ionomer chain. Reference conditions of εLJ = 1 kBT and

λB = 3 σ result in ρ* = 1.2×10-2 s-3. These values indicate sufficiently dilute conditions.

All simulations were run for 3.5×107 time steps, corresponding to 0.55 ms of total

simulation time. The first 3.0×107 time steps were used for equilibration, as tracked by

monitoring the total energy of the system as well as the gyration radius of the ionomer

chains. For equilibrium state analysis, data collection was performed for the last 5×106

time steps and data were saved every 1.0×105 time steps, resulting in 50 regularly

spaced snapshots.

Page 108: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

90 90

5.4. Results

5.4.1. Calibrating the backbone flexibility

As a baseline, we aimed at simulating ionomer chains with the backbone

flexibility of Nafion-type ionomers. To this end, we determined kθ by comparing the

calculated persistence length, lp, of the bare ionomer backbone (no side chains

attached), simulated in the coarse-grained model, with reported values for a

polytetrafluoroethylene (PTFE) chain.79,286,287 We calculated lp using288,289

lp=12b

b!

N /2( ) •b!

N /2( )−i +b!

N /2( ) •b!

N /2( )+ii=0

N /2( )−1

, (5.1)

where is the jth is the jth bond vector and b is the average bond length, b ≈ 0.97 σ.

Fig. 5.2 displays values of lp , obtained in backbone simulations, as a function of kθ.

Reported values for the persistence length of Teflon are in the range of 2-5 nm.79,286,287

We used the upper value of this range, lp = 5 nm, which equals 20 σ in the

coarse-grained ionomer model, to obtain an estimate of the corresponding kθ value.

Fig. 5.2 shows that kθ = 25 kBT produces the bending rigidity of a semi-flexible chain with

lp = 20 σ. This value was employed for kθ in all subsequent simulations.

b!j = r!j+1 − r!j

Page 109: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

91 91

Figure 5.2. Calculated persistence length of single linear chain mimicking the ionomer backbone. A value kθ = 25 kBT reproduces the persistence length of a PTFE chain. The snapshots show typical chain conformations in different regimes of chain rigidity. All the error bars in this chapter represent the standard deviations obtained over the equilibrium trajectories.

5.4.2. Effect of ionomer hydrophobicity

Fig. 5.3 shows the impact of εLJ on ionomer aggregation. Starting from the

random initial dispersion of chains, shown in Fig. 5.3a, cylindrical aggregates are formed

at sufficiently large values of εLJ while a more dispersed state prevails at low values of

εLJ. In aggregated structures, neutral backbone beads form the hydrophobic core region.

Ionomer side chains and their anionic headgroups protrude out of the core region into

the surrounding phase. Counterions form a diffuse halo around ionomer bundles. These

bundle-like structures with cylindrical shape are in general agreement with cylindrical-like

structures found in scattering studies of Nafion solutions.71,81–83,119,251

Page 110: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

92 92

Figure 5.3. (a) Initial configuration of chains in the simulation box; (b) to (d) show snapshots of the self-assembled structure at different values of εLJ = 0.5, 1.0, and 1.5 kBT. A close-up of a typical bundle is shown below the εLJ = 1.0 kBT box.

The aggregation number k for each bundle, defined as the number of ionomer chains in

each bundle, is calculated as the number of hydrophobic backbone beads in a single

bundle divided over the number of backbone beads in each chain, N. A bundle was

identified as the group of monomers for which the smallest pairwise distance is less than

the critical value, i.e., rij < 1.5 σ. Variation of this criterion between rij < 1.0 σ and rij < 2.5

σ did not alter the value of k. An average aggregation number <k> was calculated

according to

k = kiP ki( )i=1

M

∑ (5.2)

where M is the total number of ionomer chains in the box. In Eq. 5.2, P (ki) is the bundle

size distribution function defined as

P k

i( ) =N k

i( )N k

i( )i=1

M

∑ (5.3)

where N (ki) is the number of ki - bundles and <...> denotes the time averaging over

equilibrium trajectories. Fig. 5.4a shows the change in the average aggregation number,

<k>, as a function of εLJ. The dependence of <k> on εLJ is monotonic. The growth in

Page 111: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

93 93

bundle size observed upon increasing εLJ (level of ionomer hydrophobicity) is consistent

with a theory of ionomer aggregation presented in Ref. [90].

Figure 5.4. (a) Change in average aggregate size, <k>, as a function of the variation in the interaction parameter εLJ between apolar beads. (b) Change in the minimum (radial) and maximum (longitudinal) bundle size vs. εLJ. The inset shows the aspect ratio of ionomer bundles.

At low values εLJ ≤ 0.75 kBT, representing weak hydrophobicity of backbone

chains, the increase in free energy upon bundle formation due to the loss in chain

entropy outweighs the decrease in potential energy due to the backbone aggregation;

therefore, in this regime, ionomer chains remain in a dispersed state with bundle sizes of

<k> ≈ 1 to 2. In the regime of large εLJ > 1 kBT, the energy loss upon aggregation of

hydrophobic backbones becomes the dominant contribution to the free energy,

outweighing the impact of entropy loss and of the increase in electrostatic interaction

energy between anionic beads at the bundle surface; therefore, in this regime,

equilibrium bundle sizes shift to larger values.

Page 112: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

94 94

The shape and dimension of ionomer bundles were obtained from the radius of

gyration tensor (S) of each aggregate. For every aggregate a gyration tensor is

constructed that is then diagonalized to calculate the three eigenvalues, λ1, λ2, and λ3.

The squared gyration radius of bundles, Rg2, would be the first invariant of S,

TrS = λ1 + λ2 + λ3 while the second invariant gives the relative shape anisotropy defined

as

κ 2 =3S −TrS.E( )2 TrS( )

2

(5.4)

with the E being the unit tensor;290–293 κ2 ranges from zero for a sphere (λ1 = λ2 = λ3) to

1/4 for a planar symmetric object (λ1 = λ2, λ3 = 0) and to 1 for a rod (λ2 = λ3 = 0). These

eigenvalues are averaged over the ensemble of ionomer bundles present in the

simulation box. The imposed bending rigidity yields cylindrical (λ1 > λ2 ≈ λ3) ionomer

bundles. The largest eigenvalue, denoted as <Rg,max2>1/2, corresponds to the average

length of an ionomer bundle, while the average of the remaining two l values, denoted as

<Rg,min2>1/2, corresponds to the bundle thickness. Fig. 5.4b shows <Rg,max

2>1/2 and

<Rg,min2>1/2

as functions of εLJ; the inset depicts the bundle aspect ratio

<Rg,max2>1/2/<Rg,min

2>1/2. In the range εLJ < 1 kBT, bundles grow mainly in radial direction

upon increasing εLJ. At εLJ > 1 kBT, a pronounced growth trend in bundle length sets in.

The plot of the aspect ratio parameter shows a minimum at εLJ = 1 kBT. Isolation of the

anionic headgroups from surrounding counterions upon further radial bundle growth

would cause desolvation of anionic headgroups, causing a sudden increase in the

bundle free energy.132,270 These energetically unfavorable processes that depends on

the length of side chains prevent the further radial growth of bundles. They are

responsible for the preferential longitudinal growth at εLJ > 1 kBT.

Detailed analyses of bundle structures involve calculation of radial distribution

functions (RDFs) of interacting beads, g(r). RDF plots presented in the following are

normalized to the volume per bundle and to the number of beads of type A or B per

bundle. Fig. 5.5 shows the RDFs of apolar backbone monomers, gm-m(r), anionic

Page 113: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

95 95

headgroups, gh-h(r), and between headgroups and counterions, gh-c(r), at different values

of εLJ. Fig. 5.5a shows the increase in gm-m(r) with increasing εLJ. This trend represents

the enhanced aggregation of ionomer backbones as the result of increased segregation

strength. This plot furthermore reveals a dense hexagonal packing of ionomer chains in

the bundle core, indicated by the ratio of peak separations in gm-m(r) peaks that follows a

sequence , as indicated in the plot indicated in the plot. This hexagonal

form of packing was also found in Refs.[81–83,92,93] for self-assembled Nafion chains

in dilute solution as well as for self-assembled polyelectrolyte chains.246

Figure 5.5. Radial distribution functions of (a) monomer-monomer, gm-m(r), (b) headgroup-headgroup, gh-h(r), (c) headgroup-counterion, gh-c(r), at different values of εLJ. Peak location ratios are shown in (a) by considering the first peak as the reference peak.

1: 3 : 2 : 7 : 9

Page 114: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

96 96

The aggregation of backbone chains with pendant charged side chains results in

the formation of ion-rich regions in solution, as can be observed in the RDFs gh-h(r) and

gh-c(r), shown in Fig. 5.5b and 5.5c. There is no specific correlation among headgroups

in the regime of weak hydrophobicity, εLJ = 0.5 kBT. However, gh-h(r) grows as

εLJ increases and two correlation peaks emerge. Increasing εLJ enhances the intensity of

gh-h(r) due to the presence of a greater number of backbone chains in the bundle, which

results in a greater total number of anionic headgroups at the bundle surfaces. The

locations of the first and second peaks are not altered for εLJ = 0.75 kBT, 1 kBT, and

1.25 kBT. However, these peaks are shifted to smaller values, indicating a densification

of surface groups, in the case of strong hydrophobicity i.e. at εLJ = 1.5 kBT.

The RDF gh-c(r), depicted in Fig. 5.5c for various values of εLJ, reveals that the

counterion localization in the vicinity of headgroups increases with εLJ. Higher charge

density and greater surface potential in bundles with stronger hydrophobic interactions

induces stronger electrostatic attraction on counterions. This stronger interaction results

in a greater localization of counterions around bundles. Further inspection of Fig. 5.3

affirms the formation of a counterion cloud at the bundle surface in the regime of strong

hydrophobicity.

5.4.3. Effect of electrostatic interaction strength

In this section we explore the changes in aggregation behavior upon variation of

λB, while fixing εLJ = 1 kBT. The change of λB in experiment can be achieved through

changing the dielectric constant of the solvent. In all simulations performed with varying

values of λB the formation of cylindrical bundles was observed.

The nonmonotonic behavior in the plot of average aggregation number, <k>, vs.

λB in Fig. 5.6 can be explained as follows: the decreasing trend for λB < 6 σ is caused by

the electrostatic repulsion of anionic charges on the bundle surface that disfavors

aggregation; however, at large values of λB, for λB > 6 σ, the trend is reversed due to the

effect of counterion localization shown in Fig. 5.6. The fraction of localized counterions,

Page 115: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

97 97

Xc, is defined as the number of counterions in radial distance < 2.5 σ of the nearest

headgroup.272,283 Xc is an increasing function of λB. Counterion localization results in

effective screening of the anionic charges and a correspondingly reduced effective

Coulomb repulsions between headgroups.

Figure 5.6. (a) Dependence of average bundle size on the value of λB. Imbedded plots show the distribution function of the bundle number vs. k for λB = 1 σ , 6 σ , and 12 σ; (b) shows the fraction of counterions within 2.5 σ of the anionic headgroups.

For λB < 6 σ, the majority of counterions remain dispersed (non-localized) in the

electrolyte. In this regime of weak counterion localization, the free energy change upon

increasing λB is mainly determined by the increasing electrostatic repulsion between

anionic headgroups. This dependency disfavors aggregation, which is responsible for

the decrease in <k> with λB and the evolution of a more dispersed ionomer solution, as

shown by changes in the size distribution functions shifting from λB values of 1 σ to 6 σ.

Page 116: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

98 98

At λB > 6 σ, the effect of counterion localization overcompensates the increase in

direct Coulomb interaction between headgroups, leading to a net decrease in effective

electrostatic energy with increasing λB and growing <k>, and formation of a phase-

separated ionomer solution. A nonmonotonic dependence of bundle sizes on Bjerrum

length was also observed in Refs.[272,283] for the variation in the thickness of

polyelectrolyte brushes.

Fig. 5.7 shows the backbone monomer-monomer radial distribution function for

the range of explored λB values. gm-m(r) demonstrates a nonmonotonic dependence on

λB. The spatial ordering of the ionomer chains in the bundle core is retained for the

explored values of Bjerrum length. Ionomer bundles with smallest λB (λB = 1 σ) exhibit

the largest magnitudes of peaks in gm-m(r). The RDF gm-m(r) decreases as λB increases to

6 σ. Further increase in λB above 6 σ results in an upsurge in gm-m(r). These trends are

consistent with the trends in the bundle size reported in Fig. 5.6a.

Figure 5.7. Monomer-monomer radial distribution functions for different values of the Bjerrum length.

Page 117: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

99 99

5.4.4. Effect of side chain content

The effect of the side chain density on the aggregation behavior of ionomer

chains is studied by increasing the number of grafting sites, Ns, at fixed length of the

ionomer backbone. Fig. 5.8 shows the solution of ionomer aggregates for Ns = 3, 6, 9, 15

with fixed εLJ = 1 kBT and λB = 3 σ. Increasing NS is accompanied by evolution of

larger number of aggregates with smaller bundle sizes.

Figure 5.8. Snapshots showing the ionomer bundles at different values of Ns (a) Ns = 3, (b) Ns = 6, (c) Ns = 9, (d) Ns = 15.

Fig. 5.9a shows the dependence of the aggregation number on Ns. Larger Ns

results in shorter distance between the anionic headgroups. The decrease in <k> with

increasing Ns, seen in Fig. 5.9a, can be attributed to the enhanced electrostatic repulsion

between anionic headgroups that increases the bundle free energy and disfavors

aggregation. The size distribution functions also clearly show the shift of the ionomer

ensemble from a phase-separated system of large aggregates to a system consisting of

small but abundant ionomer bundles upon increasing NS.

Fig. 5.9b shows the increase of the Coulomb interaction energy of headgroups

(per headgroup) as a function of NS. This interaction energy is obtained by isolating the

electrostatic and short-range portions of the mutual headgroups interactions.

Page 118: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

100 100

Figure 5.9. The effect of Ns on (a) equilibrium aggregate number, <k> (imbedded plots show the distribution function of the bundle number vs. k), (b) Coulombic energy of anionic headgroups, and (c) maximum and minimum bundle gyration radius.

Fig. 5.9c shows the change in the longitudinal and radial bundle dimension as a function

of Ns. The decrease in the <Rg,min2>1/2 follows the same trend as <k>. An increase in

<Rg,max2>1/2, approximately 2 σ, occurs upon increasing NS from 3 to 15. This range of

variation in the bundle length mostly originates from the Coulombic-induced chain

stretching upon increasing the number of pendant ionic side chains.

Page 119: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

101 101

Fig. 5.10 displays the RDFs among headgroup units for the final aggregated

structures at different Ns. Bundles formed from chains with smaller Ns are more

populated of headgroups than the bundles formed from chain with greater Ns values.

The observed trend in gh-h(r) is attributed to the larger bundles that form from ionomer

chains with fewer side chains. Increase in bundle size upon decrease in the number of

pendant chains is in agreement with the recent water sorption model of Ref [294] where

thermally-aged membranes in which majority of acidic headgroups were lost during the

aging process demonstrated higher pore wall module. Considering the ionomer bundles

as the main structural motif in a hydrated membrane, increase in pore wall modulus can

be translated to increased bundle sizes upon loss of acidic headgroups during the aging

process. The mean-field approach adopted in Ref [90] also captures increased bundle

sizes upon decrease in the surface charge density of ionomer rods, that can be

translated to smaller number of anionic headgroups per each ionomer chain i.e. smaller

NS.

Figure 5.10. Radial distribution function among anionic headgroups, gh-h(r), for the explored values of Ns.

The side chain content in commercial PFSAs such as 3M, Hyflon, and Nafion is

primarily expressed in terms of the ion exchange capacity (IEC) of the polymer

membrane, defined as the number of moles of acid groups over the mass of dry

membrane. Major dissimilarity in PFSA ionomers with different side chain chemistries is

manifested in terms of their different side chain contents. The side chain lengths of

Page 120: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

102 102

well-known ionomer materials such as 3M, Hyflon, and Nafion vary in the range from

5.27 Å to 8.44 Å.285 We simulated systems with ls varying from 1 σ to 6 σ. We did not

find any clearly discernible trend in bundle sizes as a function of ls. The length of side

chains has however a strong impact on the dynamics of individual ionomer chains;

therefore, over the simulation time of 3.5 × 107 timesteps, no equilibration was observed

for ionomer systems with ls = 6 σ.

5.4.5. Effect of counterion valence

Effect of the counterion valance on the aggregation behavior of the ionomer

chains is investigated by comparing results for Zc = +1, +2, and +3. Fig. 5.11 depicts the

final snapshots for the systems with different counterion valences. As Zc increases the

solution contains fewer but larger ionomer bundles. In the case of trivalent counterions,

we can see the formation of a weakly bent and elongated worm-like bundle. The

localization of counterions in the vicinity of ionomer bundles increases dramatically with

Zc because of the stronger electrostatic attraction of multivalent counterions to the

anionic headgroups. This trend is clearly observed in Fig. 5.11 where almost all the

counterions are strongly localized with the anionic headgroups for Zc = +3.

Figure 5.11. Morphologies of ionomer solutions for solutions containing (a) monovalent, Zc = +1, (b) divalent, Zc = +2, and (c) trivalent, Zc = +3, counterions.

Page 121: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

103 103

Enhanced localization results in screening the repulsion of anionic headgroups,

thereby enabling the growth of larger clusters. Strong localization of multivalent

counterions also induces an attractive interaction196 among charged units known as the

bridging295 effect. Fig. 5.12a shows that the average aggregation number, <k>, increases

significantly with Zc. Fig. 5.12b shows the fraction of strongly localized counterions

(within 2.5 σ) at different values of ZC. XC increases monotonously with ZC. Fig. 5.12c

reveals a two-fold increase in average bundle length with about 0.8 σ increase in bundle

thickness over ZC. This observation points out the preferential longitudinal assembly and

growth of ionomer bundles upon increase in ZC.

Figure 5.12. Effect of counterion valence ZC on (a) average bundle sizes, (b) fraction of localized counterions and (c) effective length and diameter of bundles.

Page 122: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

104 104

It is interesting to analyze the influence of counterion valence on the dynamics of

backbone monomers and counterions by calculating the mean square displacement

(MSD) of these species. The MSD is calculated according to

, (5.4)

Fig. 5.13 shows the MSD of counterions and backbone monomer beads during the

simulation for different Zc. Fig. 5.13a reveals that the strong localization of counterions

for Zc = +3 markedly diminishes their mobility. The backbone dynamics slows down with

increased Zc, because of the larger number of chains in the bundle core and the

increased number of monomer-monomer contacts. Yan et al.274 also observed the

decrease in the counterion dynamics upon increasing its valency due to stronger

bonding to the polyelectrolyte.

Figure 5.13. (a) MSD of counterions with different valences and (b) changes in MSD of backbone monomer beads for bundles formed in the presence of counterion with different valence ZC.

MSD( t )= ri t( )− ri 0( )2

Page 123: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

105 105

5.5. Discussion

We have systematically explored the impact of primary parameters of the

ionomer-solvent system on the aggregation behavior of ionomers. Here, we will re-

iterate the main trends and evaluate them in comparison to experimental findings

reported in the literature.

Simulation results suggest that aggregation numbers of PFSA-type ionomers can

be expected to lie in the range of <k> = 6 to 10 chains, as seen in Figs. 5.4, 5.6a, 5.9a,

and 5.12a for our reference set of parameters. This range is consistent with stable

aggregate sizes of <k> = 9 that the theory of ionomer aggregation presented in Ref [90]

predicted for the baseline parameters corresponding to PFSA ionomers in aqueous

solution.

Increased hydrophobicity results in the formation of larger bundles,

corresponding to greater <k>. A change in the strength of the hydrophobic interactions

between apolar polymer beads, represented in the model study, by the parameter εLJ,

can be induced through alteration of the ionomer chemistry, for instance by replacing

fluorocarbon with hydrocarbon chains. Less hydrophobic hydrocarbon ionomers exhibit

smaller values of εLJ that diminish the tendency for bundle formation resulting in more

dispersed solution characteristics. Comparative experimental investigation of the

solution features of hydrocarbon and PFSA ionomers would be highly insightful in this

regard.

The second major parametric effect refers to the grafting density of side chains.

In the model study this parameter is evaluated by varying the number Ns of side chains

per ionomer backbone chain (with fixed length). We observed in Fig. 5.9a that an

increase in Ns results in a monotonic decrease in bundle size, leading to a dispersed

ionomer state in the high Ns limit. The ionic content of ionomers is controlled during the

synthesis process. Experimental studies of Loppinet and Gebel82 confirm the decrease

in bundle size in solutions of PFSA-type ionomers with higher side chain content. A

similar decreasing trend in aggregation behaviour was seen in SAXS studies by

Page 124: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

106 106

Rulkens et al.246 performed on solutions of semirigid poly(p-phenylene)s (PPP) chains

upon increasing the sulfonate ion content. In the limit of very high ionic content such

systems assumed the configuration of non-aggregating and dispersed polyelectrolyte

chains.

A way to isolate the impact of direct Coulomb interactions on ionomer

aggregation is to add salt ions to the solution that screen the electric charges and reduce

the value of λB. Jiang et al.81 studied the size of Nafion bundles in solution at different

salt contents. Increasing the salt concentration resulted in an increase in the size of

Nafion bundles, consistent with the trend seen in Fig. 5.6a at low λB. Experimental

studies exploring regimes of strong electrostatic interactions are needed to capture the

predicted nonmonotonic dependence of bundle size on the strength of electrostatic

interactions.

Employing solvents with different dielectric properties exerts a more complex

impact on the aggregation behavior. In simulations, changes in solvent properties affect

the interplay between solvent-mediated interactions of apolar ionomer groups, embodied

in the simulation parameter εLJ, and direct Coulomb interactions, embodied in the

Bjerrum length, λB. A less polar solvent will decrease the value of εLJ for the hydrophobic

ionomer, thus disfavoring the aggregation of backbones, and increase the value of λB,

representing an increase in the strength of direct Coulomb interactions. Looking at

Fig. 5.6a, we should distinguish two regimes. In the regime of small λB, i.e., λB < 6 σ,

Coulomb interactions of anionic headgroups are weakly screened by diffuse counterions.

In this regime of weak counterion localization, the less polar solvent causes consistent

trends in aggregation behavior exerted by εLJ and λB: the net effect is a decrease in <k>

for the less polar solvent. In the regime of high λB, i.e., λB > 6 σ, the screening effect due

to strong counterion localization in the vicinity of bundles comes into play, reverting the

impact of solvent polarity on effective electrostatic interactions. In this regime of strong

counterion localization, the impacts of variations in εLJ and λB on ionomer aggregation

oppose each other: the less polar solvent tends to weaken aggregation via εLJ and

enhance aggregation via λB; a general trend of the net effect cannot be predicted.

Page 125: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

107 107

Loppinet et al.82,83 observed, using SAXS and SANS, that the radius of rod-like

Nafion aggregates increased when more polar solvents, that is solvent with higher

dielectric constant and greater εLJ, were employed. They observed a shift in the ionomer

peak to lower q values for aqueous solvents compared to alcoholic solvents (with

λb ≈ 9 σ by considering ε ≈ 25) indicating larger radius of ionomer aggregates in solvents

with higher polarity. Consistently, a decrease in aggregate size of Nafion ionomer was

seen by Welch et al.86 when water was replaced with glycerol, a less polar solvent.

These trends are in accord with the trends predicted by simulations in the regime of

weak counterion localization.

Experimentally, there is no direct study regarding the influence of counterion

valency on the colloidal aggregation of hydrophobic ionomer chains or PFSA chains in

dilute solutions. On the membrane level, counterion valence was seen to have a strong

impact on the mechanical properties of the ionomer membrane.217,218 These changes

can possibly originate from the fundamental structural reorganizations at the bundle-

level. However, trying to relate these membrane studies to our solution-based

simulations would be highly speculative at this point.

5.6. Conclusions

In this study, we have applied coarse-grained molecular dynamics simulations to

study the aggregation of ionomer chains in dilute solution with no added electrolyte.

Simulations account for the short-range interactions of apolar hydrophobic polymer

groups in ionomer backbone and side chains as well as explicit Coulomb indications

between charged moieties, viz. anionic headgroups and counterions in solution.

The interplay of these interactions controls the formation of ionomer aggregates or

bundles of finite size. Stable bundles attain a cylindrical-like shape with dense hexagonal

packing arrangement of ionomer backbones. The size of stable bundles, the density of

anionic headgroups at their surface and the distribution of counterions in the surrounding

solution depend on the primary ionomer chemistry and architecture, viz. the chemical

nature of polymer groups, the grafting density of side chains, as well as the properties of

Page 126: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

108 108

the solvent. We found that a stronger backbone hydrophobicity of polymer groups results

in larger bundle sizes. Increasing the grafting density of acid-terminated side chains

leads to a monotonic decrease in bundle size. A variation in the strength of Coulomb

interactions gives rise to a nonmonotonic trend in bundle size: in a regime of weak

counterion localization, corresponding to a small Coulomb interaction parameter, bundle

size decreases with increasing interaction strength; in a regime of strong counterion

localization, attained for large Coulomb interaction parameter, the bundle size increases

with increasing interaction strength. Ionomers with higher number of pendant groups

form smaller bundles due to stronger electrostatic repulsion among their anionic

moieties. Increasing the valence of counterions from one to three results in a

pronounced increase in bundle size and more strikingly it leads to an almost complete

depletion of the solution from counterions owed to pronounced counterion condensation

at the bundle surface. This effect is expected to have a strong impact on ion transport

phenomena in the membrane. Generally, trends predicted based on our computational

study are consistent with experimental observations. The ability to correlate primary

ionomer parameters and solvent properties with the structure and electrostatic properties

of ionomer bundles is crucial for a rational design of membrane-forming materials.

There are different venues for improving the present study. Instances include exploring

the effect of backbone rigidity on the size and shape of aggregated bundles,

parameterizing the force field interactions for specific ionomer material, and comparing

the results to all-atomistic simulations incorporating explicit solvent molecules.

Increasing the number of ionomer chains i.e. shifting from dilute to concentrated

solutions is expected to reveal the possible interconnection and network formation of

ionomer bundles, which in turn sheds light on the morphological characteristics of

hydrated membranes. In related work, we study how the mechanical and electrostatic

properties of microscopic bundles determine water sorption and swelling, transport

phenomena, chemical degradation and mechanical failure of ionomer-based

membranes. The present work is a vital step towards a basic understanding on how to

modify ionomer and solvent properties in order to obtain polymer electrolyte membranes

with improved performance and stability.

Page 127: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

109 109

Chapter 6. Summary and future work

In Chapter 3, we introduced a novel kinetic model of side chain-initiated radical

attack and chemical degradation in fuel cell ionomers. This zero dimensional model

addresses the ex situ degradation conditions, in which radical species are produced

from a solution that contains iron ions and hydrogen peroxide. In the first part of this

work, we developed analytical relationships capturing the �OH content as a function of

Fe2+ and H2O2 concentrations. Considering the experimental challenges in direct

measurement of �OH content, the results can provide a robust estimate for predicting

radical contents under various experimental conditions. Iron ions were shown to have

more significant impact on the �OH content, which in turn suggests that certain

strategies must be adopted in order to minimize the level of iron contamination in the

PEM. The model of radical attack to the course-grained ionomers takes into account, for

the first time, the mechanism of radical attack on vulnerable sites in the ionomer side

chain. The coarse-grained model that is specifically designed for chemically stabilized

PFSA ionomers demonstrates reasonable agreement with experimental data for the

fluorine loss and IEC change in ex situ degradation studies. Application of the model to

degradation data of various PFSA membranes highlights the common mechanism of

headgroup loss in these materials. The zero dimensional model assumes an uniform

distribution of radical species as well as ionomer groups. This simplistic assumption

mostly originates from the lack of a generally accepted morphological model of hydrated

ionomer membranes. Coupling the degradation model to a structural model can be

considered as a viable next step for improving the degradation model. As part of

validating the parametric studies, we highlight the need of systematic experimental

studies on capturing the effect of pivotal ionomer properties such as IEC and side chain

length on the chemical degradation of PFSA materials.

Page 128: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

110 110

Numerous challenges regarding the hitherto unresolved PFSA morphology stem

from a lack of a bottom-up approach in understanding the basic structural properties of

ionomer materials. Basic knowledge of the conformation of single chain comb-like

ionomers seems essential for a deeper insight into the more complex PEM morphology.

In Chapter 4, we captured the fine conformational details of comb-like ionomer chains

upon systematic changes in the length and density of pendant side chains. The interplay

between long-range electrostatic interactions and short-range excluded volume effects

defines the conformational behaviour of ionomer chains. We discussed the decisive role

of a series of structural parameters on the conformational behaviour of single ionomer

chains and the corresponding affinity of neutralizing counterions. Our simulation studies

captured the intrinsic configuration of Nafion chains as globular-like structures in

accordance with AFM studies of adsorbed Nafion chains of Ref [116]. One possible

approach to further the single chain study would be parameterizing the interaction

parameters, through first principle DFT calculations, to target specific PFSA of

hydrocarbon ionomers. Comparing the results of coarse-grained MD study to all-

atomistic simulations, in which the solvent molecules are explicitly taken into account,

can reveal the effect of solvent molecules and their interactions with the ionomer chains

on the conformational behaviour of single ionomer molecules.

Spontaneous aggregation of coarse-grained ionomer chains was the focus of

Chapter 5. In this chapter, we explored the self-assembly of ionomer chains upon

variation in strength of hydrophobicity, Bjerrum length, density of pendant side chains,

and valency of counterions. Minimization of hydrophobic surface energy of ionomer

backbones is the driving force for the aggregation while the self-charge repulsion among

anionic headgroups acts as the opposing factor. Increase in the ionomer hydrophobicity

resulted in the evolution of larger ionomer bundles while at lower hydrophobicity

dispersed solution behaviour was observed. The nonmonotonic dependence of bundle

sizes upon a change in the strength of electrostatic interactions highlights headgroup-

headgroup repulsions and the condensation of counterions on the anionic moieties.

Ionomers with higher density of pendant side chains yield smaller bundles. This

phenomenon is due to enhanced electrostatic repulsion among the anionic headgroups

that is increased upon increasing the ionomer grafting density. This study confirms the

aggregated state of hydrophobic PFSA ionomer and rationalizes a series of

Page 129: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

111 111

experimental observations on ionomer solutions. Within the context of poroelectroelastic

theory of water sorption32, mechanical properties of ionomer bundles that form the pore

walls control the sorption behaviour of the PEM and they determine the mechanical

properties of the hydrated membrane. Mechanical properties of ionomer bundles are

shown to have significant impact on the fracture formation and propagation in networks

of ionomer bundles.296 In the future, simulation studies focusing on stretching the

ionomer bundles with the aim of producing the stress-strain curve of individual bundles

can provide us with an unprecedented knowledge of the elastic behaviour of self-

assembled ionomer bundles. Experimental studies measuring the elastic properties of

self-assembled biological macromolecules such as F-actin and DNA have been

performed using optical tweezers.297, 298 Thus, we suggest similar micro-scale elastic

studies on PFSA ionomer bundles.

6.1. Future work

Possible future works lie within the context of theoretical simulation of ionomer

networks or membranes employing similar coarse-grained chains. Different experimental

studies highlight the evolution of a gel-like structure upon increasing the ionomer

concentration in solution.27,71,94 Interconnection of ionomer bundles is believed to give

rise to the formation of a network structure that can possibly serve as the skeleton of

ionomer membranes used as the PEM of PEFCs. Simulating the gel-like network can

give us the ability to test different proposed structural models of PFSA ionomers and

asses the relation between the changes in bundle sizes (discussed in Chapter 5) and the

network morphology. Establishing the relation between ionomer properties such as the

density of pendant chains and the morphological properties of the gel-like network can

provide us with an unparalleled insight toward the architecture-morphology in PFSA

materials.

Another relevant outlook can be the simulation of the self-assembling behaviour

and network formation in hydrocarbon ionomers. The ease of synthesis of hydrocarbon

ionomers, which are similar to PFSAs that incorporate ionic and non-ionic segments, has

resulted in a wide range of possible molecular arrangements such as diblock and triblock

topologies.27 Phase separation of ionic and non-ionic segments results in a plethora of

Page 130: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

112 112

nanoscale structures such as spherical, lamellar, and cylindrical morphologies.64, 68, 210

Molecular modeling, at the regime of dilute solution, can elucidate the basic structural

forming processes that result in evolution of larger-scale morphologies in hydrocarbon

ionomers. Studying the network formation behaviour and the membrane morphology, on

the other hand, can rationalize the relation between the chain topology, membrane

morphology, and the ionic conductivity in these materials and can ultimately guide

designing hydrocarbon copolymers with predefined structures and properties.

However, studying concentrated systems faces different challenges. Proper

equilibration of dense polymer systems with charged moieties is a huge challenge.

Kinetic trapping and freezing effects become progressively more difficult to resolve in

concentrated self-associating systems. A possible network formation study for both

PFSA and hydrocarbon polymers would require sufficient computational power in order

to perform the simulations parallelized over numerous CPUs for sufficiently long

trajectories in order to attain properly equilibrated systems.

Page 131: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

113 113

References

(1) Eikerling, M.; Kulikovsky, A. Polymer Electrolyte Fuel Cells: Physical Principles of Materials and Operation; CRC Press: Boca Raton, FL, 2014.

(2) Asia/Worl Energy Outlook; Tokyo, 2014.

(3) Scherson, D. A.; Palencsar, A. Elctrochemical Soc. Interface 2008, 15, 17–22.

(4) Gasik, M. Materials for Fuel Cells; CRC Press: Boca Raton, FL.

(5) Bose, S.; Kuila, T.; Nguyen, T. X. H.; Kim, N. H.; Lau, K. -t.; Lee, J. H. Prog. Polym. Sci. 2011, 36, 813–843.

(6) Diat, O.; Gebel, G. Nat. Mater. 2008, 7, 13–14.

(7) Parsons, E. G. S. No Title; 2000; Vol. 5.

(8) Appleby, J.; Foulkes, R. L. In Fuel Cell Handbook; NY, 1989.

(9) Devanathan, R. Energy Environ. Sci. 2008, 1, 101–119.

(10) Wang, C.-Y. Chem. Rev. 2004, 104, 4727–4766.

(11) Smitha, B.; Sridhar, S.; Khan, a. a. J. Memb. Sci. 2005, 259, 10–26.

(12) Zhang, L.; Brostowitz, N. R.; Cavicchi, K. A.; Weiss, R. A. Macromol. React. Eng. 2014, 8, 81–99.

(13) Steele, B. C. H.; Heinzel, A. Nature, 2001, 414, 345–352.

(14) Reitz, W. Mater. Manuf. Process. 2007, 22, 789–789.

Page 132: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

114 114

(15) Wang, Y.; Chen, K. S.; Mishler, J.; Cho, S. C.; Adroher, X. C. Appl. Energy 2011, 88, 981.

(16) Hamrock, S. J.; Yandrasits, M. A. J. Macromol. Sci. Part C Polym. Rev. 2006, 46, 219–244.

(17) Vishnyakov, V. M. Vacuum, 2006, 80, 1053–1065.

(18) Zawodzinski, T. A.; Davey, J.; Valerio, J.; Gottesfeld, S. Electrochim. Acta 1995, 40, 297.

(19) Appleby, A. J.; Yeager, E. B. Energy 1986, 11, 137.

(20) Jahnke, T.; Futter, G.; Latz, A.; Malkow, T.; Papakonstantinou, G.; Tsotridis, G.; Schott, P.; Gérard, M.; Quinaud, M.; Quiroga, M.; Franco, A. A.; Malek, K.; Calle-Vallejo, F.; Ferreira de Morais, R.; Kerber, T.; Sautet, P.; Loffreda, D.; Strahl, S.; Serra, M.; Polverino, P.; Pianese, C.; Mayur, M.; Bessler, W. G.; Kompis, C. J. Power Sources 2016, 304, 207–233.

(21) Gubler, L.; Gursel, S. A.; Scherer, G. G. Fuel Cells 2005, 5, 317–335.

(22) Eisenberg, A.; Kim, J.-S. Introduction to Ionomers; John Wiley & Sons: New York, 1998.

(23) Huang, X.; Solasi, R.; Zou, Y.; Feshler, M.; Reifsnider, K.; Condit, D.; Burlatsky, S.; Madden, T. J. Polym. Sci. Part B Polym. Physics, 2006, 44, 2346–2357.

(24) Pu, H. Polymers for PEM Fuel Cells; John Wiley & Sons: Hoboken, NJ, 2014.

(25) Merle, G.; Wessling, M.; Nijmeijer, K. J. Memb. Sci. 2011, 377, 1–35.

(26) Elabd, Y. A.; Hickner, M. A. Macromolecules 2011, 44, 1–11.

(27) Kreuer, K. D. Chem. Mater. 2014, 26, 361–380.

(28) Mauritz, K. a; Moore, R. B. Chem. Rev. 2004, 104, 4535–4585.

(29) Grot, W. G. Macromol. Symp. 1994, 82, 161–172.

Page 133: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

115 115

(30) Eikerling, M.; Kornyshev, A. A. J. Electroanal. Chem. 2001, 502, 1–14.

(31) Schlick, S. Ionomers: Characterization, Theory, and Applications; CRC Press: Boca Raton, FL, 1996.

(32) Eikerling, M. H.; Berg, P. Soft Matter 2011, 7, 5976–5990.

(33) Zawodzinski, T. A.; Derouin, C.; Radzinski, S.; Sherman, R. J.; Smith, V. T.; Springer, T. E.; Gottesfeld, S. Elctrochemical Soc. 1993, 140, 1041–1047.

(34) Dixon, D. A.; Smart, B. E.; Krusic, P. J.; Matsuzawa, N. J. Fluor. Chem. 1995, 72, 209–214.

(35) Houchins, C.; Kleen, G. J.; Spendelow, J. S.; Kopasz, J.; Peterson, D.; Garland, N. L.; Ho, D. L.; Marcinkoski, J.; Martin, K. E.; Tyler, R.; Papageorgopoulos, D. C. Membranes (Basel). 2012, 2, 855–878.

(36) Andrew, H. W.; Thomas, C. K.; John, C. S. Copolym. Polyblends, Compos. Am. Chem. Soc. 1975, 142, 300–308.

(37) Peron, J.; Mani, A.; Zhao, X.; Edwards, D.; Adachi, M.; Soboleva, T.; Shi, Z.; Xie, Z.; Navessin, T.; Holdcroft, S. J. Memb. Sci. 2010, 356, 44–51.

(38) Edmondson, C. A.; Stallworth, P. E.; Chapman, M. E.; Fontanella, J. J.; Wintersgill, M. C.; Chung, S. H.; Greenbaum, S. G. Solid State Ionics 2000, 135, 419–423.

(39) Ghielmi, A.; Vaccarono, P.; Troglia, C.; Arcella, V. J. Power Sources 2005, 145, 108–115.

(40) Merlo, L.; Ghielmi, A.; Cirillo, L.; Gebert, M.; Arcella, V. J. Power Sources 2007, 171, 140–147.

(41) Radice, S.; Oldani, C.; Merlo, L.; Rocchia, M. Polym. Degrad. Stab. 2013, 98, 1138–1143.

(42) Merlo, L.; Arcella, V. In 224th ECS Meeting; The Electrochemical Society: San Francisco, 2013.

(43) Hamrock, S. In Fuel Cell Seminar and Energy Exposition; Orlando, 2011.

Page 134: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

116 116

(44) Hamrock, S. HTMWG Meet. 2005.

(45) Wei, J.; Stone, C.; Steck, A. E. No Title. 5 422 411, 1995.

(46) Steck, A. E. No Title; Montreal, 1995.

(47) Gebel, G. Macromolecules 2013, 46, 6057–6066.

(48) Borup, R.; Meyers, J.; Pivovar, B.; Kim, Y. S.; Mukundan, R.; Garland, N.; Myers, D.; Wilson, M.; Garzon, F.; Wood, D.; Zelenay, P.; More, K.; Stroh, K.; Zawodzinski, T.; Boncella, J.; McGrath, J. E.; Inaba, M.; Miyatake, K.; Hori, M.; Ota, K.; Ogumi, Z.; Miyata, S.; Nishikata, A.; Siroma, Z.; Uchimoto, Y.; Yasuda, K.; Kimijima, K.-I.; Iwashita, N. Chem. Rev. 2007, 107, 3904–3951.

(49) Wang, Y.; Chen, K. S.; Mishler, J.; Cho, S. C.; Adroher, X. C. Appl. Energy 2011, 88, 981–1007.

(50) Nasef, M. M.; Hegazy, E. S. A. Prog. Polym. Sci. 2004, 29, 499–561.

(51) Dargaville, T. R.; George, G. A.; Hill, D. J. T.; Whittaker, A. K. Prog. Polym. Sci. 2003, 28, 1355–1376.

(52) Hickner, M. A.; Ghassemi, H.; Kim, Y. S.; Einsla, B. R.; McGrath, J. E. Chem. Rev. 2004, 104, 4587–4611.

(53) Yu, J.; Yi, B.; Xing, D.; Liu, F.; Shao, Z.; Fu, Y.; Zhang, H. Phys. Chem. Chem. Phys. 2003, 5, 611–615.

(54) Roziere, J.; Jones, D. J. Annu. Rev. Mater. Res. 2003, 33, 503–555.

(55) Kim, D. J.; Jo, M. J.; Nam, S. Y. J. Ind. Eng. Chem. 2015, 21, 36–52.

(56) Sambandam, S.; Ramani, V. J. Power Sources 2007, 170, 259–267.

(57) Silva, V. S.; Ruffmann, B.; Silva, H.; Gallego, Y. A.; Mendes, A.; Madeira, L. M.; Nunes, S. P. J. Power Sources 2005, 140, 34–40.

(58) Jiang, Z.; Zhao, X.; Fu, Y.; Manthiram, A. J. Mater. Chem. 2012, 22, 24862.

Page 135: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

117 117

(59) Devrim, Y.; Erkan, S.; Baç, N.; Eroǧlu, I. Int. J. Hydrogen Energy 2009, 34, 3467–3475.

(60) Wu, X.; Wang, X.; He, G.; Benziger, J. J. Polym. Sci. Part B Polym. Phys. 2011, 49, 1437–1445.

(61) Komarov, P. V; Veselov, I. N.; Chu, P. P.; Khalatur, P. G. Soft Matter 2010, 6, 3939–3956.

(62) Kerres, J. A. J. Memb. Sci. 2001, 185, 3–27.

(63) Kreuer, K. D. J. Memb. Sci. 2001, 185, 29–39.

(64) Titvinidze, G.; Kreuer, K.-D.; Schuster, M.; de Araujo, C. C.; Melchior, J. P.; Meyer, W. H. Adv. Funct. Mater. 2012, 22, 4456–4470.

(65) Wohlfarth, A.; Smiatek, J.; Kreuer, K. D.; Takamuku, S.; Jannasch, P.; Maier, J. Macromolecules 2015, 48, 1134–1143.

(66) Saito, T.; Moore, H. D.; Hickner., M. A. Macromolecules 2009, 43, 599–601.

(67) Yoon, K.; Seok, J.; Lee, Y.; Kim, T.-H.; Yu, D. M.; Hong, S.-K.; Hong, Y. T. Eur. Polym. J. 2015, 66, 1–11.

(68) Tsang, E. M. W.; Zhang, Z.; Shi, Z.; Soboleva, T.; Holdcroft, S. J. Am. Chem. Soc. 2007, 129, 15106–15107.

(69) Eikerling, M.; Kornyshev, A. a.; Kucernak, A. R. Phys. Today 2006, 59, 38–44.

(70) Eikerling, M. H.; Berg, P. Soft Matter 2011, 7, 5976–5990.

(71) Rubatat, L.; Rollet, A. L.; Gebel, G.; Diat, O. Macromolecules 2002, 35, 4050–4055.

(72) Berrod, Q.; Lyonnard, S.; Guillermo, A.; Ollivier, J.; Frick, B.; Manseri, A.; Améduri, B.; Gébel, G. Macromolecules 2015, 48, 6166–6176.

(73) Gebel, G. Polymer (Guildf). 2000, 41, 5829–5838.

Page 136: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

118 118

(74) Bass, M.; Berman, A.; Singh, A.; Konovalov, O.; Freger, V. J. Phys. Chem. B 2010, 114, 3784–3790.

(75) Kreuer, K. D.; Portale, G. Adv. Funct. Mater. 2013, 23, 5390–5397.

(76) Gebel, G. Polymer (Guildf). 2000, 41, 5829–5838.

(77) Gierke, T. D. .; Munn, G. E. .; Wilson, F. C. J. Polym. Sci. Polym. Phys. Ed 1981, 19, 1687–1704.

(78) Yeager, H. L.; Steck., A. J. Electrochem. Soc. 1981, 128, 1880–1884.

(79) Schmidt-Rohr, K.; Chen, Q. Nat. Mater. 2008, 7, 75–83.

(80) Melchior, J. P.; Bräuniger, T.; Wohlfarth, A.; Portale, G.; Kreuer, K. D. Macromolecules 2015, 48, 8534–8545.

(81) Jiang, S.; Xia, K.-Q.; Xu, G. Macromolecules 2001, 34, 7783–7788.

(82) Loppinet, B.; Gebel, G. Langmuir 1998, 14, 1977–1983.

(83) Loppinet, B.; Gebel, G.; Williams, C. E. J. Phys. Chem. B 1997, 5647, 1884–1892.

(84) Szajdzinska-Pietek, Ewa Schlick, S.; Plonka, A. Langmuir 1994, 10, 1101–1109.

(85) Szajdzinska-Pietek, Ewa Schlick, S.; Plonka, A. Langmuir 1994, 10, 2188–2196.

(86) Welch, C.; Labouriau, A.; Hjelm, R.; Orler, B.; Johnston, C.; Kim, Y. S. ACS Macro Lett. 2012, 1, 1403–1407.

(87) Kim, Y. S.; Welch, C. F.; Hjelm, R. P.; Mack, N. H.; Labouriau, A.; Orler, E. B. Macromolecules 2015, 48, 2161−2172.

(88) Yamaguchi, M.; Matsunaga, T.; Amemiya, K.; Ohira, A.; Hasegawa, N.; Shinohara, K.; Ando, M.; Yoshida, T. J. Phys. Chem. B 2014, 118, 14922–14928.

(89) Aldebert, P. B. D. J. Phys. Fr. 1988, 49, 2101–2109.

Page 137: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

119 119

(90) MeIch, P A.; Eikerling, M Phys. Rev. E 2013, 39, 137603.

(91) Nomula, S.; Cooper, S. L. Macromolecules 2001, 34, 2653–2659.

(92) Aldebert, P.; Dreyfus, B.; Gebel, G.; Nakamura, N.; Pineri, M.; Volino, F. J. Phys. Fr. 1988, 49, 2101–2109.

(93) Aldebert, P.; Dreyfus, B.; Pineri, M. Macromolecules 1986, 19, 2651–2653.

(94) Rubatat, L.; Gebel, G.; Diat, O. Macromolecules 2004, 37, 7772–7783.

(95) Rubatat, L.; Diat, O. Macromolecules 2007, 40, 9455–9462.

(96) Jiang, S.; Xia, K.-Q.; Xu, G. Macromolecules 2001, 34, 7783–7788.

(97) Loselevich, A. S.; Kornyshev, A. A.; Steinke, J. H. G. J. Phys. Chem. B 2004, 108, 11953–11963.

(98) Eikerling, M.; Kulikovsky, A. Polymer Electrolyte Fuel Cells: Physical Principles of Materials and Operation; CRC Press: Boca Raton, FL, 2014.

(99) Akiyama, Y.; Sodaye, H.; Shibahara, Y.; Honda, Y.; Tagawa, S.; Nishijima, S. J. Power Sources 2010, 195, 5915–5921.

(100) Carlsson, A.; Jörissen, L. Fundam. Dev. Fuel Cells Conf. 2008 2008.

(101) Cipollini, N. E. ECS Trans. 2007, 11, 1071–1082.

(102) Gummalla, M.; Atrazhev, V. V.; Condit, D.; Cipollini, N.; Madden, T.; Kuzminyh, N. Y.; Weiss, D.; Burlatsky, S. F. J. Electrochem. Soc. 2010, 157, B1542–B1548.

(103) Fernandes, A. C.; Ticianelli, E. A. J. Power Sources 2009, 193, 547–554.

(104) F. D. Coms, H. Xu, T. McCallum, C. M. ECS Trans. 2012, 50, 907–918.

(105) Coms, F. D. ECS Trans. 2008, 16, 235–255.

(106) Collier, A.; Wang, H.; Zi Yuan, X.; Zhang, J.; Wilkinson, D. P. Int. J. Hydrogen

Page 138: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

120 120

Energy 2006, 31, 1838–1854.

(107) Ghassemzadeh, L.; Holdcroft, S. J. Am. Chem. Soc. 2013, 135, 8181–8184.

(108) Ghassemzadeh, L.; Kreuer, K.-D.; Maier, J.; Müller, K. J. Phys. Chem. C 2010, 114, 14635–14645.

(109) Danilczuk, M.; Perkowski, A. J.; Schlick, S. Macromolecules 2010, 43, 3352–3358.

(110) Curtin, D. E.; Lousenberg, R. D.; Henry, T. J.; Tangeman, P. C.; Tisack, M. E. J. Power Sources 2004, 131, 41–48.

(111) Ghassemzadeh, L.; Peckham, T. J.; Weissbach, T.; Luo, X.; Holdcroft, S. J. Am. Chem. Soc. 2013, 135, 15923–15932.

(112) Ghassemzadeh, L.; Kreuer, K. D.; Maier, J.; Müller, K. J. Power Sources 2011, 196, 2490–2497.

(113) Xie, T.; Hayden, C. A. Polymer (Guildf). 2007, 48, 5497–5506.

(114) Gubler, L.; Dockheer, S. M.; Koppenol, W. H. J. Electrochem. Soc. 2011, 158, B755–B769.

(115) Shah, A. A.; Ralph, T. R.; Walsh, F. C. J. Electrochem. Soc. 2009, 156, B465–B484.

(116) Vishnyakov, A.; Neimark, A. V. J. Phys. Chem. B 2000, 104, 4471–4478.

(117) Vishnyakov, A.; Neimark, A. V. J. Phys. Chem. B 2001, 105, 9586–9594.

(118) Koestner, R.; Roiter, Y.; Kozhinova, I.; Minko, S. Langmuir 2011, 27, 10157–10166.

(119) Gebel, G.; Loppinet, B.; Hara, H.; Hirasawa, E. J. Phys. Chem. B 1997, 101, 3980–3987.

(120) Arnold, A.; Holm, C. In Advanced Computer Simulation Approaches for Soft Matter Sciences II,; Springer,: Berlin, 2002; pp. 59–109.

Page 139: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

121 121

(121) Ballenegger, V.; Cerdà, J. J.; Holm, C. Comput. Phys. Commun. 2011, 182, 1919–1923.

(122) Frenkel, D.; Smit, B. Understanding Molecular Simulation; 2nd ed.; Academic Press: San Diego, 2002.

(123) Jones, R. A. L. Soft Condensed Matter; Oxford University Press: Oxford, 2002.

(124) Allen, M. P.; Tildesley, D. J. Computer Simulation of Liquids; Clarendon Press: New York, NY, USA, 1989.

(125) Van Gunsteren, W.; Berendsen, H. Mol. Phys. 1982, 45, 637–647.

(126) Grest, G. S.; Kremer, K. Phys. Rev. A 1986, 33, 3628–3631.

(127) Gennes, P. De. Scaling concepts in polymer physics; Cornell university press, 1979.

(128) Doi, M.; Edwards, S. F. The Theory of Polymer Dynamics; Oxford University Press: Oxford, NY, 1986.

(129) Li, Y.; Abberton, B. C.; Kröger, M.; Liu, W. K. Challenges in multiscale modeling of polymer dynamics; 2013; Vol. 5.

(130) Kremer, K.; Grest, G. S. J. Chem. Phys. 1990, 92, 5057–5086.

(131) Grest, G. S.; Murat, M. Macromolecules 1993, 26, 3108–3117.

(132) Limbach, H. J.; Holm, C.; Kremer, K. Macromol. Chem. Phys. 2005, 206, 77–82.

(133) Košovan, P.; Limpouchová, Z.; Procházka, K. Macromolecules 2006, 39, 3458–3465.

(134) Micka, U.; Holm, C.; Kremer, K. Langmuir 1999, 15, 4033–4044.

(135) Müller-Plathe, F. ChemPhysChem 2002, 3, 754–769.

(136) Pathria, R. No Title; 2nd ed.; Elsevier Science, 1996.

Page 140: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

122 122

(137) Allen, M. P.; Tildesley, D. J. No Title; 1st ed.; Clarendon Press: Oxford, 1987.

(138) Weeks, J. D.; Chandler, D.; Andersen, H. C. J. Chem. Phys. 1971, 54, 5237.

(139) Warner Jr, H. R. Ind. Eng. Chem. Fundam. 1972, 11, 379–387.

(140) Binder, K. Monte Carlo and Molecular Dynamics Simulations in Polymer Science; Oxford University Press: New York, 1995.

(141) Soddemann, T.; Dünweg, B.; Kremer, K. Eur. Phys. J. E Soft Matter 2001, 6, 409.

(142) Ewald, P. P. Ann. Phys. (Paris). 1921, 369, 253–287.

(143) Swope, W. C.; Andersen, H. C.; Berens, P. H.; Wilson, K. R. J. Chem. Phys. 1982, 76, 637–649.

(144) Darden, T.; York, D.; Pedersen, L. J. Chem. Phys. 1993, 98, 10089–10092.

(145) Plimpton, S. J. Comput. Phys. 1995, 117, 1–19. http://lammps.sandia.gov.

(146) Humphrey, W.; Dalke, A.; Schulten, K. J. Molec. Graph. 1996, 14, 33–38. http://www.ks.uiuc.edu/Research/vmd/.

(147) Ghelichi, M.; Melchy, P.-É. A.; Eikerling, M. H. J. Phys. Chem. B 2014, 118, 11375–11386.

(148) Pianca, M.; Barchiesi, E.; Esposto, G.; Radice, S. J. Fluor. Chem. 1999, 95, 71–84.

(149) Mittal, V. O.; Russell Kunz, H.; Fenton, J. M. Electrochem. Solid-State Lett. 2006, 9, A299–A302.

(150) Liu, W.; Zuckerbrod, D. J. Electrochem. Soc. 2005, 152, A1165–A1170.

(151) Pozio, A.; Silva, R. F.; De Francesco, M.; Giorgi, L. Electrochim. Acta 2003, 48, 1543–1549.

(152) Gubler, L.; Koppenol, W. H. J. Electrochem. Soc. 2012, 159, B211–B218.

Page 141: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

123 123

(153) Rodgers, M. P.; Brooker, R. P.; Mohajeri, N.; Bonville, L. J.; Kunz, H. R.; Slattery, D. K.; Fenton, J. M. J. Electrochem. Soc. 2012, 159, F338–F352.

(154) Rodgers, M. P.; Bonville, L. J.; Kunz, H. R.; Slattery, D. K.; Fenton, J. M. Chem. Rev. 2012, 112, 6075–6103.

(155) Stewart, S. M.; Spernjak, D.; Borup, R.; Datye, a.; Garzon, F. ECS Electrochem. Lett. 2014, 3, F19–F22.

(156) Zhou, C.; Guerra, M. A.; Qiu, Z.-M.; Zawodzinski, Jr, T. A.; Schiraldi, D. A. Macromolecules 2007, 40, 8695–8707.

(157) Yuan, X.-Z.; Zhang, S.; Ban, S.; Huang, C.; Wang, H.; Singara, V.; Fowler, M.; Schulze, M.; Haug, A.; Andreas Friedrich, K.; Hiesgen, R. J. Power Sources 2012, 205, 324–334.

(158) Schwiebert, K. E.; Raiford, K. G.; Escobedo, G.; Nagarajan, G. ECS Trans. 2006, 1, 303–311.

(159) Healy, J.; Hayden, C.; Xie, T.; Olson, K.; Waldo, R.; Brundage, M.; Gasteiger, H.; Abbott, J. Fuel Cells 2005, 5, 302–308.

(160) Yu, T. H.; Sha, Y.; Liu, W.-G.; Merinov, B. V.; Shirvanian, P.; Goddard, W. A. J. Am. Chem. Soc. 2011, 133, 19857–19863.

(161) Yu, T. H.; Liu, W.-G.; Sha, Y.; Merinov, B. V.; Shirvanian, P.; Goddard, W. A. J. Memb. Sci. 2013, 437, 276–285.

(162) Lindsey, M. E.; Tarr, M. a. Chemosphere 2000, 41, 409–417.

(163) Peter Wardman, L. P. C. Radiat. Res. 1996, 145, 523–531.

(164) Chemizmu, K.; Fentona, R. Ecol. Chem. Eng. 2009, 16, 347–358.

(165) Rigg, T.; Taylor, W.; Weiss, J. J. Chem. Phys 1954, 22, 575–577.

(166) Walling, C.; Goosen, A. J. Am. Chem. Soc. 1973, 95, 2987–2991.

(167) Jayson, G. G.; Parsons, B. J.; Swallow, A. J. J. Chem. Soc., Faraday Trans. 1 1972, 68, 2053–2058.

Page 142: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

124 124

(168) Lundström, T.; Christensen, H.; Sehested, K. Radiat. Phys. Chem. 2004, 69, 211–216.

(169) Jayson, G. G.; Parsons, B. J.; Swallow, A. J. J. Chem. Soc., Faraday Trans. 1 1973, 69, 236–242.

(170) Rush, J. D.; Bielski, B. H. J. J. Phys. Chem. 1985, 89, 5062–5066.

(171) Lee, C.; Yoon, J. Chemosphere 2004, 56, 923–934.

(172) Giguère, P. A.; Liu, I. D. Can. J. Chem. 1957, 35, 283–293.

(173) George V. Buxton, Clive L. Greenstock, W. Phillips Helman, A. B. R. J. Phys. Chem. Ref. Data 1988, 17, 513–534.

(174) Chrlstensen, H.; Sehested, K.; Corfitzen, H. J. Phys. Chem. 1982, 86, 1588–1590.

(175) Koppenol, W. H.; Butler, J.; Leeuwen, J. W. van. Photochem. Photobiol. 1978, 28, 655–658.

(176) Dainton, F. S.; Rowbottom, J. Trans. Faraday Soc. 1953, 49, 1160–1173.

(177) Bielski, B. H. J. Photochem. Photobiol. 1978, 28, 645–649.

(178) Lundström, T.; Christensen, H.; Sehested, K. Radiat. Phys. Chem. 2001, 61, 109–113.

(179) Danilczuk, M.; Coms, F. D.; Schlick, S. Fuel Cells 2008, 8, 436–452.

(180) Gubler, L.; Dockheer, S. M.; Koppenol, W. H. ECS Trans. 2011, 41, 1431–1439.

(181) Danilczuk, M.; Coms, F. D.; Schlick, S. J. Phys. Chem. B 2009, 113, 8031–8042.

(182) Danilczuk, M.; Schlick, S.; Coms, F. D. Macromolecules 2009, 42, 8943–8949.

(183) Kumar, M. Modeling chemical degradation and proton transport in perfluorosulfonic acid ionomers, University of Tennessee, Knoxville, 2011.

Page 143: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

125 125

(184) Kumar, M.; Paddison, S. J. J. Mater. Res. 2012, 27, 1982–1991.

(185) Deranleau, D. A. Experientia 1982, 38, 661–662.

(186) Dreizler, A. M.; Roduner, E. Fuel Cells 2012, 12, 132–140.

(187) Hommura, S.; Kawahara, K.; Shimohira, T.; Teraoka, Y. J. Electrochem. Soc. 2008, 155, A29–A33.

(188) Rodgers, M. P.; Pearman, B. P.; Mohajeri, N.; Bonville, L. J.; Slattery, D. K. Electrochim. Acta 2013, 100, 180–187.

(189) Ghelichi, M.; Eikerling, M. H. J. Phys. Chem. B 2016, 120, 2859–2867.

(190) Hara, M. Polyelectrolytes: Science and Technology; Marcel Dekker: New York, 1993.

(191) Holm, C.; Joanny, J. F.; Kremer, K.; Netz, R. R.; Reineker, P.; Seidel, C.; Vilgis, T. A.; Winkler, R. G. Adv. Polym. Sci. 2004, 166, 67–111.

(192) Khokhlov, A. R.; Khalatur, P. G. Curr. Opin. Colloid Interface Sci. 2005, 10, 22–29.

(193) Dobrynin, A. V. Curr. Opin. Colloid Interface Sci. 2008, 13, 376–388.

(194) Burak, Y.; Ariel, G.; Andelman, D. Biophys. J. 2003, 85, 2100–2110.

(195) Dobrynin, A. V; Colby, R. H.; Rubinstein, M. Macromolecules 1995, 28, 1859–1871.

(196) Dobrynin, A. V. Macromolecules 2006, 39, 9519–9527.

(197) Dobrynin, A. V.; Rubinstein, M. Prog. Polym. Sci. 2005, 30, 1049–1118.

(198) Dobrynin, A. V.; Rubinstein, M. Macromolecules 1999, 32, 915–922.

(199) Limbach, H.-J.; Holm, C. J. Phys. Chem. B 2003, 107, 8041–8055.

Page 144: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

126 126

(200) Carrillo, J. M. Y.; Dobrynin, A. V. J. Phys. Chem. B 2010, 114, 9391–9399.

(201) Carrillo, J.-M. Y.; Dobrynin, a V. Macromolecules 2011, 44, 5798–5816.

(202) Chang, R.; Yethiraj, A. Macromolecules 2006, 39, 821–828.

(203) Liao, Q.; Dobrynin, A. V.; Rubinstein, M. Macromolecules 2006, 39, 1920–1938.

(204) Stevens, M. J.; Kremer, K. J. Chem. Phys. 1995, 103, 1669.

(205) Nomula, S.; Cooper, S. L. J. Phys. Chem. B 2000, 104, 6963–6972.

(206) Hall, L. M.; Stevens, M. J.; Frischknecht, A. L. Macromolecules 2012, 45, 8097–8108.

(207) Hall, L. M.; Seitz, M. E.; Winey, K. I.; Opper, K. L.; Wagener, K. B.; Stevens, M. J.; Frischknecht, A. L. J. Am. Chem. Soc. 2012, 134, 574–587.

(208) Hall, L. M.; Stevens, M. J.; Frischknecht, A. L. Phys. Rev. Lett. 2011, 106, 1–4.

(209) Ting, C. L.; Stevens, M. J.; Frischknecht, A. L. Macromolecules 2015, 48, 809–818.

(210) Skalski, T. J. G.; Britton, B.; Peckham, T. J.; Holdcroft, S. J. Am. Chem. Soc. 2015, 137, 12223–12226.

(211) Wright, A. G.; Holdcroft, S. ACS Macro Lett. 2014, 3, 444–447.

(212) Thomas, O. D.; Soo, K. J. W. Y.; Peckham, T. J.; Kulkarni, M. P.; Holdcroft, S. J. Am. Chem. Soc. 2012, 134, 10753–10756.

(213) Arges, C. G.; Ramani, V.; Pintauro, P. N. Elctrochemical Soc. Interface 2010, 31–35.

(214) Luo, X.; Holdcroft, S.; Mani, A.; Zhang, Y.; Shi, Z. PCCP 2011, 13, 18055–18062.

(215) Ghassemzadeh, L.; Marrony, M.; Barrera, R.; Kreuer, K. D.; Maier, J.; Müller, K. J. Power Sources 2009, 186, 334–338.

Page 145: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

127 127

(216) Chang, Y.; Mohanty, A. D.; Smedley, S. B.; Abu-Hakmeh, K.; Lee, Y. H.; Morgan, J. E.; Hickner, M. A.; Jang, S. S.; Ryu, C. Y.; Bae, C. Macromolecules 2015, 48, 7117–7126.

(217) Fan, Y.; Tongren, D.; Cornelius, C. J. Eur. Polym. J. 2014, 50, 271–278.

(218) Daly, K. B.; Panagiotopoulos, A. Z.; Debenedetti, P. G.; Benziger, J. B. J. Phys. Chem. B 2014, 118, 13981–13991.

(219) Eikerling, M.; Kulikovsky, A. In Polymer Electrolyte Fuel Cells: Physical Principles of Materials and Operation; 2014; pp. 59–154.

(220) Vishnyakov, A.; Neimark, A. V. J. Phys. Chem. B 2001, 105, 7830–7834.

(221) Wang, C.; Duscher, G.; Paddison, S. J. RSC Adv. 2015, 5, 2368–2373.

(222) Liu, S.; Muthukumar, M. J. Chem. Phys. 2002, 116, 9975–9982.

(223) Ou, Z.; Muthukumar, M. J. Chem. Phys. 2005, 123, 1–10.

(224) Elder, R. M.; Jayaraman, A. Macromolecules 2012, 45, 8083–8096.

(225) Mantha, S.; Yethiraj, A. J. Phys. Chem. B 2015, 119, 11010–11018.

(226) Košovan, P.; Limpouchová, Z.; Procházka, K. J. Phys. Chem. B 2007, 111, 8605–8611.

(227) Connolly, R.; Bellesia, G.; Timoshenko, E. G.; Kuznetsov, Y. a.; Elli, S.; Ganazzoli, F. Macromolecules 2005, 38, 5288–5299.

(228) Hsu, H. P.; Paul, W.; Binder, K. Macromolecules 2010, 43, 3094–3102.

(229) Yethiraj, A. J. Chem. Phys. 2006, 125, 1–10.

(230) Dunweg, B.; Kremer, K. J. Chem. Phys 1993, 99, 6983–6997.

(231) Reddy, G.; Yethiraj, a. Macromolecules 2006, 39, 8536–8542.

Page 146: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

128 128

(232) Reddy, G.; Yethiraj, A. J. Chem. Phys. 2010, 132, 074903.

(233) Park, S.; Yethiraj, A. J. Phys. Chem. B 2012, 116, 1–30.

(234) Kramarenko, E. Y.; Pevnaya, O. S.; Khokhlov, A. R. J. Chem. Phys. 2005, 122, 084902–084910.

(235) E. Theodorakis, P.; G. Fytas, N. Am. J. Condens. Matter Phys. 2012, 2, 101–108.

(236) Lu, H.; Wang, J.; Bai, Y.; Lang, J. W.; Liu, S.; Lin, Y.; Cheng, J. Nat. Commun. 2011, 2, 206.

(237) Zhang, Y.; Lu, H.; Lin, Y.; Cheng, J. Macromolecules 2011, 44, 6641–6644.

(238) Deming, T. J. Chem. Rev. 2016, 116, 786–808.

(239) Malek, K.; Eikerling, M.; Wang, Q.; Navessin, T.; Liu, Z. J. Phys. Chem. C 2007, 111, 13627–13634.

(240) Malek, K.; Eikerling, M.; Wang, Q.; Liu, Z.; Otsuka, S.; Akizuki, K.; Abe, M. J. Chem. Phys. 2008, 129, 204702–204710.

(241) Ghelichi, M.; Malek, K.; Eikerling, M. H. Macromolecules 2016, 49, 1479–1489.

(242) Mauritz, K. A.; Moore, R. B. Chem. Rev. 2004, 104, 4535–4585.

(243) Litt, M. H. Polym. Prepr 1997, 38, 80–81.

(244) Haubold, H. G.; Vad, T.; Jungbluth, H.; Hiller, P. Electrochim. Acta 2001, 46, 1559–1563.

(245) Rollet, A. L.; Gebel, G.; Simonin, J. P.; Turq, P. J. Polym. Sci. Part B Polym. Phys. 2001, 39, 548–558.

(246) Rulkens, R.; Wegner, G.; Thurn-Albrecht, T. Langmuir 1999, 15, 4022–4025.

(247) Kroeger, A.; Deimede, V.; Belack, J.; Lieberwirth, I.; Fytas, G.; Wegner, G. Macromolecules 2007, 40, 105–115.

Page 147: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

129 129

(248) Zaroslov, Y. D.; Gordeliy, V. I.; Kuklin, A. I.; Islamov, A. H.; Philippova, O. E.; Khokhlov, A. R.; Wegner, G. Macromolecules 2002, 35, 4466–4471.

(249) Lin, H. L.; Yu, T. L.; Huang, C. H.; Lin, T. L. J. Polym. Sci. Part B Polym. Phys. 2005, 43, 3044–3057.

(250) He, L.; Fujimoto, C. H.; Cornelius, C. J.; Perahia, D. Macromolecules 2009, 42, 7084–7090.

(251) Paul, D. K.; Karan, K.; Docoslis, A.; Giorgi, J. B.; Pearce, J. Macromolecules 2013, 46, 3461–3475.

(252) Eikerling, M.; Kornyshev, A. A.; Kuznetsov, A. M.; Ulstrup, J.; Walbran, S. J. Phys. Chem. B 2001, 105, 3646–3662.

(253) Eikerling, M.; Kornyshev, A. A.; Spohr, E. In Fuel Cells I; Springer Berlin Heidelberg, 2008; Vol. 215, pp. 15–54.

(254) Vartak, S.; Roudgar, A.; Golovnev, A.; Eikerling, M. J. Phys. Chem. B 2013, 117, 583–588.

(255) Goswami, M.; Kumar, S. K.; Bhattacharya, A.; Douglas, J. F. Macromolecules 2007, 40, 4113–4118.

(256) Cui, S.; Liu, J.; Selvan, M. E.; Keffer, D. J.; Edwards, B. J.; Steele, W. V. J. Phys. Chem. B 2007, 111, 2208–2218.

(257) Knox, C. K.; Voth, G. A. J. Phys. Chem. B 2010, 114, 3205–3218.

(258) Lucid, J.; Meloni, S.; Mackernan, D.; Spohr, E.; Ciccotti, G. J. Phys. Chem. C 2013, 117, 774–782.

(259) Urata, S.; Irisawa, J.; Takada, A.; Shinoda, W.; Tsuzuki, S.; Mikami, M. J. Phys. Chem. B 2005, 109, 4269–4278.

(260) Jang, S. S.; Molinero, V.; Cagin, T.; Goddard III, W. A. J. Phys. Chem. B 2004, 108, 3149–3157.

(261) Vishnyakov, Aleksey Neimark, A. V. J. Phys. Chem. B 2014, 118, 11353–11364.

Page 148: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

130 130

(262) Hristov, I. H.; Paddison, S. J.; Paul, R. J. Phys. Chem. B 2008, 112, 2937–2949.

(263) Wang, C.; Paddison, S. J. Soft Matter 2014, 10, 819–830.

(264) Yamamoto, S.; Hyodo, S.-A. Polym. J. 2003, 35, 519–527.

(265) Adamcik, J.; Mezzenga, R. Macromolecules 2012, 45, 1137–1150.

(266) Karatasos, K. Macromolecules 2008, 41, 1025–1033.

(267) Srinivas, G.; Shelley, J. C.; Nielsen, S. O.; Discher, D. E.; Klein, M. L. J. Phys. Chem. B 2004, 108, 8153–8160.

(268) Wang, S.; Larson, R. G. Langmuir 2015, 31, 1262–1271.

(269) Khalatur, P. G.; Khokhlov, A. R.; Mologin, D. a.; Reineker, P. J. Chem. Phys. 2003, 119, 1232–1247.

(270) Limbach, H. J.; Sayar, M.; Holm, C. J. Phys. Condens. Matter 2004, 16, S2135–S2144.

(271) Omar, A. K.; Hanson, B.; Haws, R. T.; Hu, Z.; Bout, D. A. Vanden; Rossky, P. J.; Ganesan, V. J. Phys. Chem. B 2015, 119, 330–337.

(272) Sandberg, D. J.; Carrillo, J. M. Y.; Dobrynin, A. V. Langmuir 2007, 25, 13158–13168.

(273) Stevens, M. J. Phys. Rev. Lett. 1999, 89, 101–104.

(274) Yan, L.-T.; Guo, R. Soft Matter 2012, 8, 660.

(275) Wong, G. C.; Tang, J. X.; Lin, a; Li, Y.; Janmey, P. a; Safinya, C. R. Science 2000, 288, 2035–2039.

(276) Tang, J. X.; Janmey, P. a. J. Biol. Chem. 1996, 271, 8556–8563.

(277) Vasilevskaya, V. V.; Markov, V. A.; Ten Brinke, G.; Khokhlov, A. R. Macromolecules 2008, 41, 7722–7728.

Page 149: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

131 131

(278) Morriss-Andrews, A.; Shea, J. E. J. Phys. Chem. Lett. 2014, 5, 1899–1908.

(279) Mohammadinejad, S.; Fazli, H.; Golestanian, R. Soft Matter 2009, 5, 1522–1529.

(280) Goobes, R.; Cohen, O.; Minsky, A. Nucleic Acids Res. 2002, 30, 2154–2161.

(281) Rogers, P. H.; Michel, E.; Bauer, C. a.; Vanderet, S.; Hansen, D.; Roberts, B. K.; Calvez, A.; Crews, J. B.; Lau, K. O.; Wood, A.; Pine, D. J.; Schwartz, P. V. Langmuir 2005, 21, 5562–5569.

(282) Fu, I. W.; Markegard, C. B.; Chu, B. K.; Nguyen, H. D. Langmuir 2014, 30, 7745–7754.

(283) Carrillo, J. M. Y.; Dobrynin, A. V. Langmuir 2009, 25, 13158–13168.

(284) Hess, B.; Sayar, M.; Holm, C. Macromolecules 2007, 40, 1703–1707.

(285) Clark, J. K.; Paddison, S. J. J. Phys. Chem. A 2013, 117, 10534–10543.

(286) Rosi-Schwartz, B.; Mitchell, G. R. Polymer (Guildf). 1996, 37, 1857–1870.

(287) Chu, B.; Wu, C.; Bucks, W. Macromolecules 1989, 22, 831–837.

(288) Ullner, M.; Jönsson, B.; Peterson, C.; Sommelius, O.; Söderberg, B. J. Chem. Phys. 1997, 107, 1279–1287.

(289) Stevens, M. J.; Kremer, K. J. Chem. Phys. 1995, 103, 1669–1690.

(290) Šolc, K. J. Chem. Phys. 1971, 55, 335–344.

(291) Theodorou, D. N.; Suter, U. W. Macromolecules 1985, 18, 1206–1214.

(292) Arkin, H.; Janke, W. J. Chem. Phys. 2013, 138, 1–9.

(293) Šolc, K.; Stockmayer, W. H. J. Chem. Phys. 1971, 54, 2756–2757.

(294) Liu, H.; Paddison, S. J. J. Phys. Chem. C 2016, 5, 537–543.

Page 150: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

132 132

(295) Liu, S.; Ghosh, K.; Muthukumar, M. J. Chem. Phys. 2003, 119, 1813–1823.

(296) Melcy, P. A.; Eikerling, M. J. Phys. Condens. Matters 2014, 67, 1329123.

(297) Farré, A.; van der Horst, A.; Blab, G. A.; Downing, B. P.; Forde, N. R. J. Biophotonics. 2010, 3, 224-233.

(298) Wang, M. D.; Yin, H.; Landick, R.; Gelles, J.; Block, S. M. Biophys. J. 1997, 72, 1335-1346.

Page 151: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

133 133

Appendix A. Solution for cubic equation.

In this appendix the cubic equation derived for [�OOH] is presented and its three roots are explicated.

•OOH!" #$3+ •OOH!" #$

2 2 k1 + k2( )k9 H2O2!" #$+ 2k4k5 Fe

2+!"

#$0

2 k4 + k5( )k9

%

&

''

(

)

**−

2k1k2 H2O2!" #$

2Fe2+!"

#$0

2 k4 + k5( )k9

%

&

'''

(

)

***= 0

with: S = R+ D3 and T = R− D3

D =Q3 + R2; Q = −a2

9 and R =

− 27b+ 2a3( )54

considering: a =2 k1 + k2( )k9 H2O2

!" #$+ 2k4k5 Fe2+!

"#$0

2 k4 + k5( )k9

%

&

''

(

)

**

b =2k1k2 H2O2

!" #$2Fe2+!"

#$0

2 k4 + k5( )k9

%

&

'''

(

)

***

Three roots are:

•OOH!" #$=−a3+ S +T( ) (A1)

•OOH!" #$=−a3−

12S +T( )+ 1

2i 3 S −T( ) (A2)

•OOH!" #$=−a3−

12S +T( )− 1

2i 3 S −T( ) (A2)

From the above-mentioned roots, A1 is the only physically meaningful answer for

[�OOH] since the A2 and A3 roots contain imaginary terms.

Page 152: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

134 134

Appendix B. Python code creating the LAMMPS data file.

1. importnumpy2. #----------------filename,tacticity,polymerinformation------#3. file_name="Pymer"#filenamefordataand.xyzfile4. both_sides="yes"#"yes":branchonbothsides,"no"branchononeside

5. bond_length=0.97#bondlengthofbead

springunits,determinedfromFENEpotential6. backbone_spacing=2#numberofbeadsbetweentwobranchingpoints7. branch_length=0#lengthofthebranchesorsidechains8. repeat_num=61#numberofrepeatingunits9. #--------------------------simulationboxinformation-------------#10. xlo=0.00011. xhi=110.00012. ylo=0.00013. yhi=110.00014. zlo=0.00015. zhi=110.00016. #-----------assigningmatrixesofcoordinates,bonds,andangles----#17. bead_mat=numpy.zeros(shape=(tot_beads,7))#numberofbeads18. coord_mat=numpy.zeros(shape=(tot_beads,3))#coordinatematrix19. bond_mat=numpy.zeros(shape=(tot_bonds,4))#bondmatrix20. angle_mat=numpy.zeros(shape=(tot_angles,5))#anglesmatrix21. counter_backbone=122. counter_bond=123. counter_branch=024. counter_branch_side=0.025. #--------------------calculationsectionforcoordinates----------#26. foriinrange(0,tot_beads):27. bead_mat[i][0]=i+128. bead_mat[i][1]=129. bead_mat[i][2]=130. bead_mat[i][3]=031. bead_mat[i][6]=+5532. bead_mat[i][5]=+5533. bead_mat[0][4]=bond_length+2.534. bead_mat[i][4]=bead_mat[i-1][4]+bond_length35. counter_backbone+=136. ifboth_sides=="yes":37. ifcounter_backbone>backbone_spacing+1:38. ifcounter_branch<branch_length:39. ifcounter_branch_side%2==0:40. bead_mat[i][4]=bead_mat[i-1][4]41. bead_mat[i][5]+=(bond_length+bead_mat[i-1][5]-55)42. counter_branch+=143. bead_mat[i][2]=2#type44. ifcounter_branch>=branch_length:45. counter_branch_side+=146. counter_backbone=147. counter_branch=048. bead_mat[i][2]=349. bead_mat[i][3]=-1.050. else:

Page 153: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

135 135

51. bead_mat[i][4]=bead_mat[i-1][4]52. bead_mat[i][5]+=(-bond_length+bead_mat[i-1][5]-55)#-5553. counter_branch+=154. bead_mat[i][2]=2#type55. ifcounter_branch>=branch_length:56. counter_branch_side+=157. counter_backbone=158. counter_branch=059. bead_mat[i][2]=360. bead_mat[i][3]=-1.061. ifboth_sides=="no":62. ifcounter_backbone>backbone_spacing+1:63. ifcounter_branch<branch_length:64. bead_mat[i][4]=bead_mat[i-1][4]65. bead_mat[i][5]+=bond_length+bead_mat[i-1][5]66. counter_branch+=167. ifcounter_branch>=branch_length:68. counter_backbone=169. counter_branch=070. #--------------------calculationsectionforbonds-------------

#71. foriinrange(0,tot_bonds):72. bond_mat[i][0]=i+173. bond_mat[i][1]=174. bond_mat[i][2]=i+175. bond_mat[i][3]=i+276. ifi==counter_bond*(backbone_spacing+branch_length)-1:77. bond_mat[i][2]=i-branch_length+178. bond_mat[i][3]=i+279. counter_bond+=180. #--------------------calculationsectionforangles------------#81. ifbranch_length!=0:82. foriinrange(0,backbone_spacing+branch_length+1):83. angle_mat[i][0]=i+184. angle_mat[i][1]=185. angle_mat[i][2]=i+186. angle_mat[i][3]=i+287. angle_mat[i][4]=i+388. ifi==(backbone_spacing+branch_length)-2:89. angle_mat[i][2]=i-branch_length+390. angle_mat[i][3]=i-branch_length+291. angle_mat[i][4]=i+392. ifi==(backbone_spacing+branch_length)-1:93. angle_mat[i][2]=i-branch_length94. angle_mat[i][3]=i-branch_length+195. angle_mat[i][4]=i+296. ifi==(backbone_spacing+branch_length):97. angle_mat[i][2]=i-branch_length98. angle_mat[i][3]=i+199. angle_mat[i][4]=i+2100. foriinrange(backbone_spacing+branch_length+1,tot_angles):101. angle_mat[i][0]=i+1102. angle_mat[i][1]=1103. angle_mat[i][2]=angle_mat[i-backbone_spacing-branch_length-

1][2]+backbone_spacing+branch_length104. angle_mat[i][3]=angle_mat[i-backbone_spacing-branch_length-

1][3]+backbone_spacing+branch_length105. angle_mat[i][4]=angle_mat[i-backbone_spacing-branch_length-

1][4]+backbone_spacing+branch_length

Page 154: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

136 136

106. ifbranch_length==0:107. foriinrange(0,tot_angles):108. angle_mat[i][0]=i+1109. angle_mat[i][1]=1110. angle_mat[i][2]=i+1111. angle_mat[i][3]=i+2112. angle_mat[i][4]=i+3113. 114. counterion=[]115. counterion_copy=[]116. foriinrange(backbone_spacing+branch_length-

1,bead_mat.shape[0],backbone_spacing+branch_length):117. counterion.append(bead_mat[i,:])118. counterion_copy=numpy.copy(counterion[:])119. forminrange(0,len(counterion_copy)):120. counterion_copy[m][0]=m+1+tot_beads121. counterion_copy[m][2]=4.0#counteriontype122. counterion_copy[m][3]=1.0#counterioncharge123. ifm%2==0:124. counterion_copy[m][5]=counterion_copy[m][5]+2.0125. ifm%2!=0:126. counterion_copy[m][5]=counterion_copy[m][5]-2.0127. #----------------writtingtheoutputfiles:.dataand.xyz-------

#128. file=open("%s.data"%file_name,"w")129. file.write("LAMMPSDescription\n\n"'%d'%tot_beads+"atoms\n"'%d'%tot

_bonds+"bonds\n")130. file.write('%d'%0+"angles\n"'%d'%0+"dihedrals\n"'%d'%0+"impr

opers\n\n")131. file.write('%d'%4+"atomtypes\n"'%d'%1+"bondtypes\n"'%d'%0+"

angletypes\n")132. file.write('%d'%0+"dihedraltypes\n"'%d'%0+"impropertypes\n\n")

133. file.write(""+'%0.3f'%xlo+""+'%0.3f'%xhi+"xloxhi\n\n")134. file.write(""+'%0.3f'%ylo+""+'%0.3f'%yhi+"yloyhi\n\n")135. file.write(""+'%0.3f'%zlo+""+'%0.3f'%zhi+"zlozhi\n\n")136. file.write("Masses\n\n"+'%d'%1+""+'%d'%1+"\n"'%d'%2+""+'%d'%

1+"\n"'%d'%3+""+'%d'%1+"\n"'%d'%4+""+'%d'%1+"\n\n")137. file.write("Atoms\n\n"+'\n'.join(''.join(str(cell)forcellinrow)fo

rrowinbead_mat))138. file.write("\n\nBonds\n\n"+'\n'.join(''.join(str(cell)forcellinrow

)forrowinbond_mat.astype(int)))139. file.write("\n\nAngles\n\n"+'\n'.join(''.join(str(cell)forcellinro

w)forrowinangle_mat.astype(int)))140. file.close()141. foriinrange(0,2):142. coord_mat[:,i]=bead_mat[:,i+4]143. file=open("%s.xyz"%file_name,"w")144. file.write('%d'%tot_beads+"\n\n")145. file.write('\n'.join(''.join(str(cell)forcellinrow)forrowincoo

rd_mat))146. file.close()147. withopen("%s.xyz"%file_name,'r')asf:148. lines=f.readlines()149. withopen("%s.xyz"%file_name,'w')asf:150. fori,lineinenumerate(lines):151. ifi>1:152. f.write('n')153. f.write(line)

Page 155: Molecular Modeling of Structural Transformations in Ionomer Solutions and Membranes

137 137

154. file.close()155. 156. file=open("%s.counterion"%file_name,"w")157. file.write('\n'.join(''.join(str(cell)forcellinrow)forrowincou

nterion_copy))158. file.close()