• Aucun résultat trouvé

Towards a general kinetic model for the thermal oxidation of epoxy-diamine networks. Effect of the molecular mobility around the glass transition temperature

N/A
N/A
Protected

Academic year: 2021

Partager "Towards a general kinetic model for the thermal oxidation of epoxy-diamine networks. Effect of the molecular mobility around the glass transition temperature"

Copied!
18
0
0

Texte intégral

(1)

Science Arts & Métiers (SAM)

is an open access repository that collects the work of Arts et Métiers Institute of

Technology researchers and makes it freely available over the web where possible.

This is an author-deposited version published in:

https://sam.ensam.eu

Handle ID: .

http://hdl.handle.net/10985/19198

To cite this version :

Xavier COLIN, Fatima ESSATBI, Justine DELOZANNE, Gurvan MOREAU - Towards a general

kinetic model for the thermal oxidation of epoxy-diamine networks. Effect of the molecular mobility

around the glass transition temperature Polymer Degradation and Stability Vol. 181, p.117

-2020

Any correspondence concerning this service should be sent to the repository

(2)

Towards

a

general

kinetic

model

for

the

thermal

oxidation

of

epoxy-diamine

networks.

Effect

of

the

molecular

mobility

around

the

glass

transition

temperature

Xavier

Colin

a,∗

,

Fatima

Essatbi

a

,

Justine

Delozanne

b

,

Gurvan

Moreau

b

a PIMM, Arts et Métiers Institute of Technology, CNRS, CNAM, HESAM University, 151 boulevard de l’Hôpital, 75013 Paris, France b Safran Composites, A platform of Safran Tech, 33 avenue de la Gare, 91760 Ittevile, France

Keywords: Epoxy-diamine network Thermal oxidation Kinetic model Oxygen consumption Carbonyl build-up Molecular mobility

a

b

s

t

r

a

c

t

The kineticmodel previouslyestablished fordescribing the thermal oxidationof polymethylenic sub-strateshasbeensuccessfullygeneralizedtoaseriesofsixepoxy-diaminenetworks(EPO-DA) character-izedbyverydifferentglasstransition temperatures.Thismodelis derivedfromthe so-called “closed-loop” mechanisticschemewhichconsists inaradical chainreactioninitiatedbythe decompositionof hydroperoxidesand propagatingviatheC-Hbondslocatedinα ofheteroatoms(NandO).The numer-ousmodelparametersweredeterminedbyapplyinga“stepbystep” procedurecombiningexperiment andsimulation.Ontheone hand,oxygentransportproperties(i.e.coefficients ofoxygen diffusionand solubility)wereestimatedfromacompilation ofliteraturedata.Ontheotherhand,rate constantsand formation yieldsweredetermined byinversesolving methodfrom themeasurements ofoxygen con-sumption andcarbonylbuild-up performedonsixdifferent EPO-DAnetworksbetween 25 and200°C andbetween0.16and 20barsofoxygenpartial pressureinourlaboratory orinthe literature.It was foundthatthemolecularmobilitymainlyaffectstherateconstantsoftheelementaryreactions involv-ingthereactivespeciesinthelowestconcentration,i.e.peroxyradicals.Infact,therateconstantk6of theapparentterminationofperoxyradicalsisreducedbyaboutfiveordersofmagnitudewhenpassing fromrubberytoglassystateduetothefreezingoflargeamplitudecooperativemolecularmovements. Incontrast,therateconstantk3 ofthepropagationofoxidation,involvingperoxyradicalsbutalsothe polymersubstrate,isonlychangedbyoneorderofmagnitudearoundtheglasstransitiontemperature. TheintroductionoftheeffectofmolecularmobilityintotheArrheniuslawsofk6 andk3allowsbuilding mastercurvesandfinally,proposingasinglekineticmodelforthewholefamilyofEPO-DAnetworks.

1. Introduction

Thevast majorityofcompositematerial structuresusedinthe civil aeronautical sector are made of diamine cross-linked epoxy matrix(EPO-DA)reinforcedwithcarbonfibers.Thedevelopmentof a numericaltool forpredictingthedegradationstate ofthis poly-merfamilyunderthecombinedeffectsoftemperature,oxygenand mechanicalstressremainsanopenissueinthissector.Indeed,this toolcouldhelpaeronauticalmanufacturerstotakeintoaccountthe possiblealterationofthethermomechanicalpropertiesof compos-itestructuresfromthemomentoftheirdesignandsizing,butalso toconsidertheuseofthesematerials inmoreaggressive thermo-chemicalenvironments.

Corresponding author.

E-mail address: xavier.colin@ensam.eu (X. Colin).

Thereisalargeamountofliteratureworksdevotedtothe ther-maldegradationmechanisms[1-9]andkinetics[10-17]ofEPO-DA networks,showingthatoxidationisclearlythepredominantaging process. However, in thesepublications, each network is consid-eredasacompletelydifferentmaterialfromothernetworksofthe EPO-DA family, because it is assumedto be characterized by its own chemistry, its own oxidation mechanisms, its own degrada-tionkineticsanditsspecifickineticmodel.Therefore,thevaluesof therateconstantsdeterminedbyusingthesekineticmodelsasan inversesolvingmethodarefoundtobequitedifferent,butwithout givinganyrealexplanationexcept,perhaps,the possibleeffectof molecularmobility[15].

However, contrary to popular belief, there are not so great chemicaldifferencesbetweenallthenetworksoftheEPO-DA fam-ily,especially fromachemicalkineticspointofview,asshownin Table1.Indeed, letusrecall that, accordingtoKorceketal.[18],

(3)

Table 1

Main oxidation sites in EPO-DA networks.

Starting epoxy monomer Oxidation sites

DGEBA, DGEBF, DGEBD, TGPAP, TGTPM, TGMDA, etc…

TGPAP and TGMDA only

Abbreviations : Diglycidyl ether of bisphenol A (DGEBA), Diglycidyl ether of bisphenol F (DGEBF), Diglycidyl ether of bisphenol D (DGEBD), Triglycidyl ether of para-amino phenol (TGPAP), Triglycidyl ether of triphenyl methane (TGTPM), Tetraglycidyl ether of methylene dianiline (TGMDA).

oxidation propagates via the most labile hydrogens of the con-stitutive repeating unit. Thus, in EPO-DAnetworks, the main ox-idation sites are oxy-methylene (-O-CH2 -) and amino-methylene (>N-CH2 -)groups,whicharealreadypresentinthestartingepoxy monomers, but also amino-methylene (-CH2 -N<) and methanol groups(>CH-OH)whichareformedduringthepolymerization re-action between the epoxy monomer and diamine hardener. All theseoxidation sitesare reportedin Table 1,as well asthe cor-respondingnameofthestartingepoxymonomer.

Basedontheseobservations,itisdifficulttoimaginelarge fun-damental differences between the oxidation mechanisms of the differentEPO-DAnetworks.However,astheirglasstransition tem-peraturecanvaryinawidetemperaturerange,typicallybetween 0°C and 221°C [19], it is expected that the molecular mobility plays indeed a key role in their oxidation kinetics, in particular onbothsidesoftheglasstransitiontemperature. Inother words, atasameagingtemperature, itwouldnotbe surprisingtofinda large difference betweenthe oxidation ratesof two EPO-DA net-worksthatare indifferent physicalstates(i.e.in glassy and rub-bery states). Unfortunately, until now, there are too few experi-mentaldataavailableintheliteratureonthethermaloxidationof EPO-DAnetworkstodoubtlesscheckthisassumption.Inaddition, almostallstudiesarefocusedonthermalaginginglassystatethat is,of course, the usual temperature range for industrial applica-tionsofthesematerials.

The impact of molecular mobility on the oxidation kinetics was clearly put in evidence for another thermodynamic transi-tion:themelting,inanelectron-beamirradiationcross-linked lin-earpolyethylene(XLPE)[20,21].Ajumpofabouthalfadecadewas observedfor the pre-exponential factoraround the melting tem-perature(130°C)intheArrheniusdiagramoftheoxidation induc-tiontime.However,thisdiscontinuityisrelativelysubtleandthus, canbehiddenbythewiderscatteringofexperimentaldatainPE compilations[22].

The effect of glass transition on the oxidation kinetics is ex-pected to be much more marked because all the diffusion pro-cessesarefavoredbylargeamplitudecooperativemolecular move-mentsinrubberystate, whereasthey are almosttotallyfrozen in glassystate.Infact,asforacommonpolymerizationreaction,two kineticregimesshouldbedistinguished:i)OneaboveTg ,forwhich thereactionkineticsisgovernedbytheintrinsicreactivityof

reac-tivespeciesandthus,canbesatisfyinglydescribedbytheclassical rulesofchemicalkinetics;ii)AnotherbelowTg ,forwhichthe re-action kinetics is governed by the diffusivity of reactive species. Thislatterregime wasspecificallyinvestigatedby Waite [23]and Russianresearchers[24,25].

That is the reason why some aromatic polymers which con-tainsalso few hydrocarbongroupsreputed to be veryeasily oxi-dizable(e.g.methynicandmethylenicgroups),suchaspolyimides of poly(bismaleimide) type [26] or polymerization ofmonomeric reactant (PMR) type [27], can be considered as thermostable in glassy state. This is also the case for many aromatic epoxies [10,28,29].

The present article aims at laying down the main lines of a general kinetic model for the thermal oxidation of networks of the EPO-DA family and checking its validity. After having re-called its theoretical bases and structure, its different param-eters will be determined by applying a “step by step” pro-cedure combining experience and simulation. First of all, oxy-gen transport properties(i.e. coefficients ofoxygen diffusionand solubility) will be estimated from a compilation of literature data. Then, rate constants and formation yields will be deter-minedby inversesolvingmethodfromthemeasurements of oxy-gen consumption andcarbonyl build-upperformedon six differ-ent EPO-DA networks between 25 and 200°C and between 0.16 and 20 bars of oxygen partial pressure in our laboratory or in literature.

2. Theory

2.1. Oxidationsites

Asmentionedinintroduction,themainoxidationsitesin EPO-DAnetworksareoxy-methylene(-O-CH2 -),amino-methylene(> N-CH2 -) andmethanolgroups(>CH-OH)(seeTable 1).Indeed,C−H bonds are characterized by a much lower dissociation energy (ED ≈376kJ.mol−1 )whenthey arelocatedin

α

ofaheteroatom(O orN)ratherthaninpolymethylenicsequences(ED ≈ 393kJ.mol−1 ) [30].Table2recalls theArrheniusparameters andorders of mag-nitudeat30°Coftherateconstantk3 forthepropagationof oxi-dationdetermined intheprevious studiesconductedinour labo-ratory[22,31-37].

(4)

Table 2

Arrhenius parameters and orders of magnitude at 30 °C of the rate constant k 3 for the propagation of oxidation on

the main types of methylenic C −H bonds.

C-H bond k 30 (L.mol −1 .s −1 ) E 3 (kJ.mol −1 ) k 3 at 30 °C (L.mol −1 .s −1 ) Polymer References

−CH 2 −CH 2 − 1.5 × 10 10 73 4.0 × 10 −3 PE [22 , 31 , 32]

9.9 × 10 10 78 3.6 × 10 −3 EPDM [33]

−CH = CH −CH 2 − 4.0 × 10 9 68 7.7 × 10 −3 BR [34]

−O −CH 2 − 1.8 × 10 9 63 2.5 × 10 −2 PET, EPO-AN [35 , 36]

> N −CH 2 − 1.8 × 10 9 63 2.5 × 10 −2 PA 6–6 [37]

Abbreviations: Polyethylene (PE), Ethylene-Propylene terpolymer (EPDM), Polybutadiene (BR), Poly(ethylene tereph- thalate) (PET), Anhydride cross-linked epoxy (EPO-AN), Polyamide 6–6 (PA 6–6).

Thus, as linear polyesters [35], epoxy-anhydride networks [36] andaliphaticpolyamides[37],EPO-DAnetworkswillbe sub-jectedtoamuchmoresevereoxidationthanPE [22,31,32].

2.2. Mechanisticscheme

Theso-called“closed-loop” oxidationmechanism,developedfor polymethylenicsubstratesinpreviousstudies[22,31-37],was nat-urallychosenforEPO-DAnetworksinthepresentstudy.Its gener-alizationtoEPO-DAnetworksrequiredare-adjustmentinthe con-sumptionyieldsofPHsubstrateinthedifferentbalancereactions (i.e.involvingseveralelementarychemicalevents).Itsfinalversion, alreadydetailedandexplainedinreference[37],isgivenbelow:

Initiation:

1u) POOH 2P+(1

γ

CO )P-OH+

γ

CO P=O+

γ

1 S {−(2+

γ

1 )PH}(k1u ) 1b) 2POOH → P+PO2 +(1−

γ

CO )P-OH+

γ

CO P=O+

γ

1 S {−(1+

γ

1 )PH}(k1b ) Propagation: 2) P+O2 →PO2 (k2 ) 3) PO2+PH→POOH+P(k3)

(Terminatingandnon-terminating)bimolecularcombinations: 4) P+P

γ

4P-P+(1−

γ

4)PH+(1−

γ

4)F+

γ

4B{+(1−

γ

4)PH}

(k4 )

5) P+PO2

γ

5 POOP+(1−

γ

5 )POOH+(1−

γ

5 )F+

γ

5 B(k5 )

6a)PO2 +PO2 →[PO••OP]cage +O2 (k6a ) 6b)[PO••OP]cage →POOP+B(k6b )

6c)[PO••OP]cage →P-OH+P=O(k6c )

6d) [PO••OP]cage → 2P+2(1−

γ

CO )P-OH+2

γ

CO P=O+2

γ

1 S

{−2(1+

γ

1 )PH}(k6d )

where PH;POOH; P, PO2 andPO;POOP, P-OH andP=O;F; S and B designate an oxidation site; an hydroperoxide; alkyl, per-oxyandalkoxylradicals;peroxide,hydroxyl andcarbonylgroups; adoublebond;achainscissionandacrosslinknode,respectively. kiare thereactionrateconstantsand

γ

iare theformationyields

ofthedifferentinactiveproductsinthecorrespondingbalance re-actions.

Verybriefly,letusrecallthemaincharacteristicsofthis oxida-tionmechanism:

i) Thepropagationofoxidation,byabstractingthemostlabile hydrogenatoms(3),leadstotheformationofaveryinstable species:hydroperoxidegroupPOOH.Indeed,thedissociation energyoftheO−Obondisextremelylow:ED ≈ 150kJ.mol−1 , againstmorethan280kJ.mol−1 forallotherchemicalbonds (i.e.C−O,C−N,C−CandC−H)composingtheEPO-DA net-works [30]. Therefore, this main propagation product will be, either from the starting of exposure, or very quickly, themainsourceofradicals.This“closedloop” character al-lowsaccountingforthesuddenandsharpauto-acceleration ofoxidation attheendofan incubationperiod,commonly

called the “induction period” [38]. It is noteworthy that, in the kinetic model, the initial hydroperoxide concentra-tion [POOH]0 is kinetically equivalent to all structural de-fects(e.g.oxidationproductsformedduringthe polymeriza-tionreaction ofthemacromolecular network)andintrinsic species(e.g.catalytic residues,PH-O2 complexes,other im-purities,etc.)responsiblefortheveryfirstoxidationevents. ii) Thethermaldecomposition ofPOOHoccursintwo modes: unimolecular (1u) and bimolecular (1b). When the poly-merisnot toopre-oxidized(i.e.when[POOH]0 < [POOH]C ), thePOOHdecompositionstartsinunimolecularmode.Then, when [POOH] exceeds a critical value [POOH]C , which is a decreasing function of temperature obeying an Arrhe-nius law, the bimolecular decomposition becomes in turn thepredominantmode.ForEPO-DAnetworks, itwasfound that: [POOH]C=6.8 × 103 Exp

(

−40,000/RT

)

. On the

con-trary,when[POOH]0 >[POOH]C ,oxidationisinitiatedbythe bimolecularmodethroughouttheexposureduration.Thisis generallythecaseforallpolymersatlowtemperatures (typ-icallyforT<100°C).

iii) The writingofthe two initiations (1uand1b) can be sur-prisingduetotheapparentnon-compliancewith stoichiom-etry. In fact, these initiations are balance reactions involv-ing several elementary chemical events. Their writing will not be detailed here, but the reader can consult reference [37]for further details.Letusjustpointoutthatthiswriting isentirelyjustifiedbecausethedissociationofO−Obondis thelimitingstep.Inaddition,thesameformationyieldscan beused forthe differentinactiveproducts (P-OH,P=Oand S) becausetheyresultfromthesame reactivespecies(PO andPH)involved inthesetwo initiations. However,due to theignoranceoftheexactnatureofcarbonylgroupP=Oand thus,thevalueofitsmolarextinctioncoefficient,itwas nec-essarytointroduceanapparent formationyield

γ

CO andto distinguishitfromtheactualyieldofchainscissions

γ

1 .The sameconsiderationsapply to thenon-terminating bimolec-ularcombination(6d).

iv) Terminationsbetweentwo alkyl radicalsP (4)orbetween onealkyl P andone peroxy radicalPO2 (5)are also bal-ance reactions.Indeed, thesetwo reactions involvetwo el-ementarychemicalevents:couplinganddisproportionation. Thus,

γ

4and

γ

5aretheformationyieldsincrosslinknodes.

v) Infact,thecombinationbetweentwo peroxyradicals PO2 (6) involves four competitive elementary chemical events (6ato6d).Thiswritingallowedhighlightingtheexistenceof anon-terminatingcombination,butalsobetter distinguish-ing crosslinking B from chain scissions S. Indeed, in poly-methylenic substrates,the termination of PO2 is not very efficient[39,40].Asan example,itwasfound thatabout70 and85% of alkoxyl radicals escape from the cage to initi-atenewradicaloxidation chainsat160°C inPE andPA6– 6,respectively[37].However, forasakeofsimplicity,these four chemical events can be grouped into a single

(5)

appar-ent terminating combination(PO2 +PO2 →inactive prod-ucts+O2 )whosethecorrespondingrateconstantk6 canbe expressed asa functionofall elementaryrateconstants, as follows[22]:

k6 = 2

(

k6b +k6c

)

2 k6a

(

k6b +k6c +k6d

)

(

2k6b +k6c +k6d

)

(1)

vi) Finally, it is noteworthythat inactive molecular (P-OH and P=O)andmacromolecularproducts(SandB)areexclusively formed in initiations (1u and1b) and in the (terminating and non-terminating) bimolecular combinations of peroxy radicals(6cand6d).

2.3.2.3Kineticmodel

Asystemoftennon-lineardifferentialequations(SED)was de-rived from the previous mechanistic scheme using the classical rulesofthechemicalkinetics.Thissystemgivesaccesstothelocal concentration changes of all the reactive speciesinvolved in the mechanistic scheme (i.e.[POOH], [P], [PO2 ], [PO••OP]cage , [PH] and[O2 ]=f(z,t)),butalsotothelocalconcentrationchangesofthe mostimportantinactiveproductsfromapracticalpointofview,as theycan beeasily measuredexperimentally.Inthisstudy,we fo-cusedon oxygen consumption ([O2abs]), carbonyl groups([P=O])

and macromolecular modifications (i.e. B and S). In the case of one-dimensionaloxygen diffusion (e.g.through thesample thick-ness),theresultingSEDcanbewrittenasfollows:

d[POOH] dt =−k1u[POOH]− 2k1b [POOH] 2 +k 3[PO•2 ][PH] +

(

1−

γ

5

)

k5 [P•][PO•2 ] (2) d[P•] dt =2k1u [POOH]+k1b [POOH] 2 − k2 [P•][O2 ] +k3 [PO•2 ][PH]− 2k4 [P•]2 −k5 [P•][PO•2 ]+2k6d [PO•OP]cage (3) d[PO•2 ] dt =k1b [POOH] 2 +k2 [P•][O2 ]− k3 [PO2 •][PH]− k5 [P•][PO•2 ] −2k6a [PO•2 ]2 (4) d[PO••OP]cage dt =k6a [PO • 2 ]2 −

(

k6b +k6c +k6d

)

[PO••OP]cage (5) d[PH] dt =−

(

2+

γ

1

)

k1u [POOH]−

(

1+

γ

1

)

k1b [POOH] 2 −k 3 [PO•2 ][PH] −2

(

1+

γ

1

)

k6d [PO••OP]cage (6) d[O2 ] dt =DO2

2 [O 2 ]

z2 − k2 [P •][O 2 ]+k6a [PO•2 ]2 (7) d[O2 abs ] dt =RO2 =k2 [P•][O2 ]− k6a [PO • 2 ]2 (8) d[P=O] dt =

γ

CO k1u [POOH]+

γ

CO k1b [POOH] 2 +

(

k6c +2

γ

CO k6d

)

[PO••OP]cage (9) dS dt =

γ

1 k1u [POOH]+

γ

1 k1b [POOH] 2 +2

γ

1 k6d [PO••OP]cage (10) dB dt =

γ

4 k4 [P •]2 +

γ

5 k5 [P•][PO•2 ]+k6b [PO••OP]cage (11)

whereDO2 is thecoefficient ofoxygen diffusion intothe EPO-DA networkandzisthespatialco-ordinate(i.e.thedepthbeneaththe samplesurface).RO2 istherateofoxygen chemicalconsumption, commonlycalledthe“oxidationrate”.

This SED was solved simultaneously in space (z) and time (t) withnumerical algorithms specifically recommendedfor“stiff problems” of chemical kinetics [41], asalreadyreported inmany previouspapers,forinstanceinreference[42].

Initialandboundaryconditionswere:

i) Whent=0,atanydepthzinthesamplethickness:

[POOH]

(

z, 0

)

=[POOH]0 =10−4 mol.L−1

[P•]

(

z, 0

)

=[PO•2 ]

(

z, 0

)

=[PO••OP]cage

(

z, 0

)

=0 [PH]

(

z, 0

)

=[PH]0

and[O2 ]

(

z, 0

)

=[O2 ]0

ii) Whent≥ 0,onbothsamplesurfaces(z=± L):

[O2 ]

(

−L, t

)

=[O2 ]

(

+L, 0

)

=[O2 ]0

OxygenconcentrationobeystheclassicalHenri’slaw:

[O2 ]0 =SO2 .PO2 (12)

whereSO2 is thecoefficient ofoxygen solubilityinto theEPO-DA network and PO2 is the oxygen partial pressure in the exposure environment.Thesamplethicknessis2L.

3. Materialsandexperimentalprocedure

3.1. EPO-DAnetworks

The studyfocused on sixperfectEPO-DA networks (i.e. with-out dangling chains) characterized by very differentglass transi-tion temperatures, typically ranged between 95 and 263°C (see Table3).

Fourofthemresultfromthereactionofcommonbi-functional or tri-functional epoxy monomers with an aromatic diamine hardener: 9,9-bis(3-chloro-4-aminophenyl)fluorene (CAF) or 4,4 -diaminodiphenylsulfone(DDS).Overthepasttwentyyears,these four networkswere all considered, inturn, aspotential matrices ofcarbonfiberreinforcedcompositematerialsforstructural appli-cations atlow tomoderatetemperatures (typicallybetween70°C and150°C)in thecivilaeronauticalsector. Themarketing of net-work no. 5was progressivelyabandoned in the early 2000s,but itsthermaldegradationbetween150°Cand210°Cwasthesubject ofseveralpapersafewyearsearlier,forinstance[10-14].Network no.6wasconsideredintheearly2010sforapplicationsat temper-aturesabove200°C,forwhichthereisclearlyalackofsolutionsin organicmaterials inEurope.Its thermaldegradationwasalsothe subject offew papers, forinstance [15,16]. In contrast,networks no.3and4arecompletely newmaterials whosethethermal ag-ingbehaviorshavenotstillbeenpublishedtodate.

FilmsofthesefourEPO-DAnetworks,withthicknessestypically ranging between25 and100μm,were produced by compression moldingthen post-curedunder primary vacuum(i.e.10−3 bar) in accordancewiththe recommendedindustrial curecycle, inorder toreachthemaximumcrosslinkingdensitywhileavoidingany un-desiredpre-oxidationbeforeexposuretothermalaging.Thatisthe reasonwhyaverylow concentrationinstructuraldefectswill be chosenformodelingthethermalagingkineticsofthesenetworks inthepresentstudy,typically:[POOH]0 =10−4 mol.l−1 .

After processing, thefilms were characterized by conventional laboratorytechniques.Inparticular,theirglasstransition tempera-ture(Tg )wasdeterminedbydifferentialscanningcalorimetry(DSC, witha20°C.min−1 heatingrate,undernitrogen)andconfirmedby mechanicalspectrometry(DMA, intensilemode,witha 1Hz fre-quencyand a2°C.min−1 heatingrate, undernitrogen). The mea-surementswerefoundveryclosetothetheoreticalvaluesofTg

(6)

Table 3

Molar mass of the repetitive monomer unit (m UCR ), density ( ρ) and concentration in oxidation sites (PH)

for perfect EPO-DA networks.

Number Network m UCR (g.mol −1 ) ρ [PH] (mol.L −1 ) T g ( °C)

1 Epon 828-Jeffamine D230 910 1.16 24.2 95 2 Epon 828-Ancamine 2049 918 1.2 32.7 150–155 3 DGEBF-CAF 1041 1.25 14.4 158 4 DGEBA-CAF 1097 1.25 13.7 182 5 (a)DGEBF/TGPAP-DDS 1362 1.31 15.4 211 6 (b)Tactix 123/Tactix 742-DDS 1674 1.11 10.6 263

Abbreviations : Diglycidyl ether of bisphenol A (DGEBA, Epon 828 or Tactix 123), Diglycidyl ether of bisphenol F (DGEBF), Triglycidyl ether of para-amino phenol (TGPAP), Triglycidyl ether of triphenyl methane (TGTPM or Tactix 742), poly(oxypropylene)diamine (Jeffamine D230), diamino dimethyl dicyclo- hexyl methane (Ancamine 2049), 9,9-bis(3-chloro-4-aminophenyl)fluorene (CAF), 4,4  -Diamino diphenyl sulfone (DDS).

(a) Matrix commercialized under the reference 977–2 by Cytec Fiberite. (b) Matrix commercialized under the name Tactix by Hunstman.

(i.e.without danglingchains).As anexample,themethodforthe Tg calculationofnetworkno.6isdetailedinreference[29].

In addition, EPO-DA networks no. 1 and 2, resulting from the reaction of diglycidyl ether of bisphenol A (Epon 828) with an aliphatic diamine hardener: poly(oxypropylene) diamine (Jef-famine D230) and diamino dimethyl dicyclohexyl methane (An-camine 2049) respectively, were considered in the present study fortwomainreasons:

i) Their Tg is lower than those of all previous networks (see Table3),thusallowingtoanalyzethepossibleeffectof molecu-larmobilityontheoxidationkineticsoverawidertemperature range;

ii) Theiroxygenconsumptionkineticswasmeticulouslymeasured inrealtimeatmoderatetoverylowtemperaturescloseto am-bient(typicallybetween140°Cand25°C)[44].

The concentration inoxidation sites [PH] was estimated from the knowledge ofthe repetitive monomer unit ofthe sixperfect EPO-DA networks. In Table 3, itcan be seen that [PH] is signifi-cantlyhigherforepoxiescross-linked withanaliphatic(typically, between 24and 33mol.L−1 ) than an aromaticdiamine (between 10and15mol.L−1 ),whichmightalreadyexplain asignificant dif-ference inoxidizability betweenthesetwo subfamiliesofEPO-DA networks.Itshouldbepointedoutthat,whenDDSisthe crosslink-ingagent,thesulfonylgroupprotectstheamino-methylenegroups (-CH2 -N<),formedduringthepolymerizationreaction,againstthe radicalattack fromoxidation.Indeed,thesulfonylgroupisa high electron-attracting group whose the inductive effect through the aromatic rings of DDS leads to an increase in the strength of the C−H bond. This stabilizing effect of sulfonyl group was first evidenced by comparing the polymerization kinetics of different epoxy-diaminemixtures [45].This isthereasonwhy[PH]isonly about10mol.L−1 fornetworkno.6.

3.2. Thermalagingandexperimentalprocedures

The oxidation kinetics of networks no. 3, 4 and 6 was stud-ied at 120, 150, 180 and 200°C under an oxygen partial pres-surerangedbetween0.21(ambientair)and20barsinautoclaves. Given the high impact of oxygen partial pressure on the oxida-tion kinetics, the aging experiments at the highest temperature (200°C) were limited to a maximum pressure of 1bar so asnot to observehighconversionratios ofoxidationfromthe veryfirst samplings. At the opposite, at the lowest temperatures (120 and 150°C), theagingexperiments werecarried outabove 0.21barin order to detect significant structural modifications after a mini-mum duration of 3 months. All the films were periodically re-moved fromtheagingchambersandcooledtoroom temperature

Fig. 1. Changes in the FTIR spectrum of network no. 3 during its thermal aging at 150 °C under 0.21 bar of oxygen (i.e. ambient air). The final spectrum (in sky blue) was recorded after 156 hours of thermal exposure.

inadesiccatorcontainingsilicagelforpreventinganymoisture re-covery prior to be characterized by FTIR spectrophotometry. The FTIRspectrawere recorded inatransmission mode between400 and4000cm−1 withaPerkinElmerFrontierapparatus,after hav-ing averaged the 16 scans obtained with a minimum resolution of4cm−1 .Asreportedmanytimesinliterature [1-10,46-48],the mainstructuralmodificationswereobservedinthecarbonylregion wheretwonewwideabsorptionbandsappearedandgrewrapidly with exposure time: one centered around 1680–1690 cm−1 , and the other around 1720–1730 cm−1 . As an example, Fig. 1 shows thechanges over time inthe FTIRspectrum of network no.3 at 150°Cunder0.21barofoxygen(i.e.ambientair).Thesetwobands wereassignedtoamidesandothertypesofcarbonylproducts, re-spectively.Unfortunately,thegreatvarietyoftheselatterproducts (aldehydes, carboxylic acids, phenyl formates, etc.), often result-ingfromoxidationinducedchainscissions inthehydroxylpropyl ethersegment[47],didnotallowidentifyingthemprecisely.Their averageconcentrationthroughoutthefilmthickness[P=O]was de-terminedbyapplyingtheclassicalBeer-Lambert’slaw:

[P=O]=epOD

ε

(13)

whereODistheopticaldensityoftheIRabsorptionbandcentered at1720–1730 cm−1 (dimensionless),

ε

isthecorresponding molar extinction coefficient (expressed in L.mol−1 .cm−1 ), and ep is the filmthickness(incm).

Forcarbonyl products, typicalvalues of

ε

areranged between 150L.mol−1 .cm−1 (for ketones) and 850L.mol−1 .cm−1 (for car-boxylic acids) [49-53].In a first approximation, an average value of500L.mol−1 .cm−1 waschosenfor

ε

inthepresentstudy.

(7)

Fig. 2. Arrhenius graph between 10 and 110 °C for the coefficient of oxygen solubility of EPO-DA networks.

The oxidation kineticsof networkno. 5wasmeasured in real time at180 and 200°C underan oxygen partial pressure ranged between0.05and 1barwith a homemade thermogravimeter.All detailsontheapparatusandexperimentalprocedureareavailable inreference[10].

The oxidation kineticsofnetworksno.1 and2wasmeasured inrealtimebetween25and140°Cunder0.166barofoxygenwith anultrasensitivemanometricmethod.Alldetailsontheapparatus andexperimentalprocedureareavailableinreference[44].

4. Kineticmodeling

Thefirststepinkineticstudyconsistedintheseparate determi-nationofthenumerousparametersofthekineticmodel.Theseare thesevenrateconstants(k1u ,k1b ,k2 ,k3 , k4 ,k5 andk6 ), thefour yieldsofinactiveproductsformation(

γ

1 ,

γ

CO ,

γ

4 and

γ

5 )andthe twooxygentransportproperties(SO2 andDO2 ).

4.1.Oxygentransportproperties

Figs. 2and3respectively showthe compilationsof the litera-turevaluesofSO2 andDO2 determinedbyoxygenpermeability be-tween10and100°C forEPO-DAnetworks[10,44,54-59].The val-uesmeasuredwithSystech8001andMoconOx-Tran2–21devices at room temperature for network no. 6 in our laboratory were addedtothiscompilation.

It clearlyappears that SO2 erratically varieswithtemperature. Given the relatively large scattering of the experimental data, it wasconsidered,infirstapproach,thatSO2 isindependentof tem-perature.Incontrast,DO2significantlyincreaseswithtemperature

andseemstoobeyanArrhenius’law.Itshouldbepointedoutthat, forbothoxygen transportproperties,no discontinuityisobserved whenapproachingtheglasstransitiontemperature.

Two sub-families of EPO-DA networks are clearly put in evi-dence:

a) For epoxies cross-linked with an aliphatic diamine hardener (JeffamineD230,JeffamineD400,Ancamine2049orIPDA):

SO2 ≈ 5.1 10−8 mol.L−1 .Pa−1 (14)

DO2 =3.010−5 Exp−44RT600 m2 .s−1 (15)

b) And for epoxiescross-linked withan aromatic diamine hard-ener(DDMorDDS,butalsoCAFbygeneralization):

SO2 ≈ 1.4510−7 mol.L−1 .Pa−1 (16)

DO2 =4.710−5 Exp−20RT000 m2 .s−1 (17)

Ofcourse,networkno.6followsthislatterbehavior.Therefore, it wouldseem that the presence of aromaticstructures, i.e. very rigidstructures,inthemacromolecularnetworkpromotesthe dis-solutionofoxygen,butalsoconsiderablyslowsdownitsdiffusion. Atthisstageoftheinvestigations,morein-depthstudiesona se-riesofmodelnetworksarenecessarytoelucidatetherelationships betweenthemolecularandmacromolecularstructures ofEPO-DA networksandtheiroxygentransportproperties.

4.2. Rateconstantsandformationyields

Rate constants and formation yields were determined by in-versesolvingmethod,inparticularbysimulatingascloselyas pos-siblethemeasurementsofcarbonylbuild-upandoxygen consump-tionofthesixEPO-DAnetworksunderstudy.Giventheverylarge numberofparameterstobedetermined,itwasdecidedtoapplya “stepbystep” procedure.

Inafirststep,asthemainoxidationsitesareC-Hbondslocated in

α

ofheteroatoms(NandO),thevaluesdeterminedforaliphatic polyamides in a previous study [37] were attributed to the rate constantsofEPO-DAnetworks.However,itshouldberecalledthat allthesevaluesweredeterminedinrubberystate,whilethe EPO-DAnetworksunderstudyaremainly inglassy state.Therefore,in

(8)

Fig. 3. Arrhenius graph between 10 and 110 °C for the coefficient of oxygen diffusion of EPO-DA networks.

asecondstep,itwasnecessarytograduallymodifysomerate con-stants, in particular those for which the corresponding reactions involvereactivespeciesinverylowconcentration, suchasperoxy radicals PO2 . Indeed, it was expected that these rate constants wouldbedirectlyimpactedbythefreezingoflargeamplitude co-operativemolecularmovements,whichcharacterizethemain me-chanicalrelaxation

α

associatedtoglasstransition.Inorderof pri-ority,thesearetherateconstantsk6i ofthe(terminatingand non-terminating) bimolecularcombinations,then the rateconstantk3

of the propagationof oxidation,because this latteronly involves onePO2 radical.

Figs. 4and 5show two examples ofnumericalsimulations in glassy state attemperatures well belowthe glasstransition tem-perature (typically when T< Tg – 60°C). Carbonyl build-up was calculatedwithEq.(9)at150°Cbetween0.21and20barsof oxy-gen(Fig.4),andbetween120and200°Cunder0.21barofoxygen for network no.6 (Fig. 5). It can be seen that the kinetic model satisfyinglyaccountsforboththeeffectsofoxygenpartialpressure andtemperatureontheoxidationkineticsofthisEPO-DAnetwork. Therateconstantsandformationyieldsusedforthese simula-tionsarereportedinTable4.Inaddition,thevaluesfoundat150°C foraliphaticpolyamidesinapreviousstudy[37]aregivenfor in-formation.Indeed,theselatterwillbe veryhelpfulforthe follow-ing discussion,becausethey were obtainedinrubberystate, well abovetheglasstransitiontemperatureofthesematerials(typically whenT≥ Tg+100°C).

Allthesevaluescallforseveralcomments:

i) For network no. 6, the initiation rate constants (k1u and k1b) are about 3–4 times higher compared to aliphatic

polyamides throughoutthe temperaturerangeunder inves-tigation.Thisresultsuggestsahigherinstabilityof hydroper-oxidesinEPO-DAnetworkspresumablyduetothepresence of heteroatoms (N or O) in a

β

position (see Table 1). As it will be seen in the next Tables 5 and6, this is general behavior for EPO-DA networks. The Arrhenius laws found for k1u and k1b forthe whole family of EPO-DA networks

throughoutthe25–200°Cintervalare:

k1u =1.31013 Exp−130RT000s−1 (16)

and

k1b =1.9109 Exp−90000

RT L.mol

−1 .s−1 (17)

ii) For network no.6,the rateconstant k6 of apparent termi-nationofperoxyradicals isfiveorders ofmagnitudelower comparedto aliphaticpolyamidesthroughoutthe tempera-turerangeunderinvestigation.Thislargegapwasmainly at-tributedtotheeffectofmolecularmobilityonthe bimolec-ular combinations of peroxy radicals on both sides of the glasstransitiontemperature.Incontrastandasexpected,the impactof molecular mobilityon the propagationof oxida-tionismuchmoremoderate.Indeed,fornetworkno.6,the rateconstantk3 isonlytwoordersofmagnitudelower com-paredtoaliphaticpolyamides.

iii) The formation yield incarbonyls

γ

CO isquite different

be-tween network no. 6 and aliphatic polyamides throughout the temperature range under investigation. This difference couldbeexplainedbytheformationofamuchgreater vari-etyofcarbonylproductsinEPO-DAnetworks.Inaddition,as therelativepredominancebetweenthesedifferentcarbonyl products depends on temperature,

γ

CO varies throughout the temperatureinterval under investigation. The fact that

γ

CO decreases withtemperaturejustconfirmsthat

aldehy-desare muchlessoxidizable atlowtemperature.As a con-sequence,theaveragemolarextinctioncoefficient

ε

,usedto estimatetheglobalconcentration ofcarbonylproductsfrom FTIRmeasurements(seeEq.13),shouldbeprogressively de-creasedwithtemperatureinordertotakeintoaccountthis changesin thechemical composition. As it willbe seen in thenext Table5, thisalso seems to be a general behavior forEPO-DAnetworks.

iv) Allotherrateconstants(k2 ,k4 andk5 )andformationyields (

γ

1 ,

γ

4 and

γ

5 ) were kept identical, if not very close, to

(9)

Fig. 4. Carbonyl build-up in 100 μm thick films of network no. 6 (Tactix 123/Tactix 742-DDS) at 150 °C between 0.21 and 20 bars of oxygen. Comparison between simulations with Eq. (9) (solid lines) and experimental data (points).

Fig. 5. Carbonyl build-up in 100 μm thick films of network no. 6 (Tactix 123/Tactix 742-DDS) between 120 and 200 °C under 0.21 bar of oxygen (ambient air). Comparison between simulations with Eq. (9) (solid lines) and experimental data (points).

(10)

Table 4

Values of the rate constants and formation yields used for modeling the thermal oxidation kinetics of network no. 6 between 120 and 200 °C. Comparison with the values previously determined at 150 °C for aliphatic polyamides [37] .

Polymer Tactix 123/Tactix 742-DDS [T g = 263 °C] PA [T g ≈ 40–55 °C]

T ( °C) 120 150 180 200 150 k 1u (s −1 ) 8.0 × 10 −5 1.3 × 10 −3 1.5 × 10 −2 6.5 × 10 −2 4.4 × 10 −4 k 1b (L.mol −1 .s −1 ) 2.1 × 10 −3 1.5 × 10 −2 8.1 × 10 −2 2.2 × 10 −1 3.5 × 10 −3 k 2 (L.mol −1 .s −1 ) 10 8 10 8 10 8 10 8 10 8 k 3 (L.mol −1 .s −1 ) 7.0 × 10 −2 2.9 × 10 −1 9.9 × 10 −1 2.0 3.0 × 10 1 k 4 (L.mol −1 .s −1 ) 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 k 5 (L.mol −1 .s −1 ) 3.0 × 10 11 3.0 × 10 11 3.0 × 10 11 3.0 × 10 11 5.0 × 10 11 k 6 (L.mol −1 .s −1 ) 3.0 × 10 4 6.8 × 10 4 8.2 × 10 4 8.7 × 10 4 7.4 × 10 9 γ1 (%) 60 60 60 60 100 γCO (%) 8.5 11.2 21.5 33.5 100 γ4 (%) 50 50 50 50 55 γ5 (%) 50 50 50 50 55 Table 5

Values of the rate constants and formation yields used for modeling the thermal oxidation kinetics of networks no. 3 and 4 at 120 and 150 °C. Comparison with the values previously determined at 150 °C for aliphatic polyamides [37] .

Polymer DGEBF-CAF [T g = 158 °C] DGEBA-CAF [T g = 182 °C] PA [T g ≈ 40–55 °C]

T ( °C) 120 150 120 150 150 k 1u (s −1 ) 8.0 × 10 −5 1.3 × 10 −3 8.0 × 10 −5 1.3 × 10 −3 4.4 × 10 −4 k 1b (L.mol −1 .s −1 ) 2.1 × 10 −3 1.5 × 10 −2 2.1 × 10 −3 1.5 × 10 −2 3.5 × 10 −3 k 2 (L.mol −1 .s −1 ) 10 8 10 8 10 8 10 8 10 8 k 3 (L.mol −1 .s −1 ) 2.5 × 10 −1 3.0 3.1 × 10 −1 2.0 3.0 × 10 1 k 4 (L.mol −1 .s −1 ) 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 k 5 (L.mol −1 .s −1 ) 3.0 × 10 11 3.0 × 10 11 3.0 × 10 11 3.0 × 10 11 5.0 × 10 11 k 6 (L.mol −1 .s −1 ) 7.7 × 10 4 4.3 × 10 6 5.5 × 10 4 3.0 × 10 5 7.4 × 10 9 γ1 (%) 60 60 60 60 100 γCO (%) 25 30 30 35 100 γ4 (%) 50 50 50 50 55 γ5 (%) 50 50 50 50 55 Table 6

Values of the rate constants and formation yields used for modeling the thermal oxidation kinetics of networks no. 1 and 2 at 125 and 140 °C, and network no. 5 at 180 and 200 °C.

Network Epon 828-Jeffamine D230 [T g = 95 °C] Epon 828-Ancamine 2049 [T g = 150–155 °C] DGEBF/TGPAP- DDS [T g = 211 °C] T ( °C) 125 140 125 140 180 200 k 1u (s −1 ) 1.3 × 10 −4 5.4 × 10 −4 1.3 × 10 −4 5.4 × 10 −4 1.5 × 10 −2 6.4 × 10 −2 k 1b (L.mol −1 .s −1 ) 3.0 × 10 −3 8.0 × 10 −3 3.3 × 10 −3 8.0 × 10 −3 8.1 × 10 −2 2.2 × 10 −1 k 2 (L.mol −1 .s −1 ) 10 8 10 8 10 8 10 8 10 8 10 8 k 3 (L.mol −1 .s −1 ) 2.1 4.6 3.0 × 10 −1 7.0 × 10 −1 8.0 × 10 −1 3.0 k 4 (L.mol −1 .s −1 ) 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 k 5 (L.mol −1 .s −1 ) 3.0 × 10 11 3.0 × 10 11 3.0 × 10 11 3.0 × 10 11 3.0 × 10 11 3.0 × 10 11 k 6 (L.mol −1 .s −1 ) 2.9 × 10 9 6.9 × 10 9 6.5 × 10 5 9.9 × 10 6 1.7 × 10 6 2.7 × 10 7 γ1 (%) 60 60 60 60 100 100 γ4 (%) 50 50 50 50 55 55 γ5 (%) 50 50 50 50 55 55

the valuesof aliphaticpolyamides [37].As it can be noted inTables4,5and6,alltheseparametersareindependentof temperature.

Figs.6and7showtwoexamplesofnumericalsimulationsstill inglassystate,butnow,uptotemperaturesneartheglass transi-tion temperature(typicallyforTg – 60°C <T≤ Tg – 10°C). Car-bonylbuild-upswerecalculatedwithEq.(9)at120and150°C be-tween 0.21and10bars ofoxygen fornetworksno.3and4.Here also, it can be seen that the kinetic model satisfyingly accounts forboththeeffectsoftheoxygenpartialpressureandtemperature ontheoxidationkineticsofthesetwoEPO-DAnetworks,despitea slightlylargerscatteringofexperimentalresults.

Therateconstantsandformationyieldsusedforallthese sim-ulations are reported inTable 5.In addition, thevalues found at

150°Cforaliphaticpolyamidesinapreviousstudy[37]arerecalled forinformation.

Thesenewresults call forthe followingadditional comments. Asexpected,theimpactofthemolecularmobilityonthe bimolec-ularcombinations ofperoxy radicals andthe propagation of oxi-dationissignificantlyreducedwhenincreasingtheexposure tem-peratureandapproachingtheglasstransitiontemperatureofboth EPO-DAnetworks.Asanexample,atT=Tg – 30°Cfornetworkno. 3,k6 isonlythreeordersofmagnitudelowercomparedtoaliphatic polyamides(againstfiveordersofmagnitudewhenT<Tg – 60°C). Similarly,atT=Tg – 30°Cfornetworkno.3,k3 isonlyoneorderof magnitudelower comparedtoaliphaticpolyamides (againstthree whenT<Tg – 60°C).

Finally,Fig.8showsone exampleofnumericalsimulations on bothsidesoftheglasstransitiontemperature(typicallywhenTg –

(11)

Fig. 6. Carbonyl build-up in 25 μm thick films of network no. 4 (DGEBA-CAF) at 120 °C between 0.21 (ambient air) and 10 bars of oxygen. Comparison between simulations with Eq. (9) (solid lines) and experimental data (points).

Fig. 7. Carbonyl build-up in 25 μm thick films of network no. 3 (DGEBF-CAF) at 150 °C between 0.21 (ambient air) and 10 bars of oxygen. Comparison between some simulations with Eq. (9) (solid lines) and experimental data (points).

(12)

Fig. 8. Arrhenius graphs between 25 and 200 °C of the maximal oxidation rate under 0.166 bar of oxygen for 100 μm thick films of networks no. 1 and 2 (Epon 828-Jeffamine D230 and Epon 828-Ancamine 2049, respectively), and 50 μm thick films of network no. 5 (DGEBF/TGPAP-DDS). Comparison between the simulations with Eq. (8) (solid lines) and literature data (points) [10 , 44] .

130°C≤ T ≤ Tg +45°C). Maximaloxidation rateswerecalculated withEq.(8)between25and140°Cunder0,166bar ofoxygen for networks no. 1 and2, and at180 and 200°C under 0,166bar of oxygen fornetworkno.5. Hereagain, itcan be seenthat the ki-netic modelsatisfyinglypredicts theeffectoftemperatureonthe oxidationkineticsofthesethreeEPO-DAnetworksthroughoutthis widetemperaturerange.

For a sake ofbrevity and clarity, only the rate constants and formationyieldsusedforthesimulationsatthetwohighest tem-peratures, for each EPO-DAunder study, are reportedin Table6. Theparametervaluesobtainedatallothertemperaturesare avail-ableinAppendixAfornetworksno.1and2.

These newresultscall for thefollowing additionalcomments. As expected, the gap betweenthe oxidationkinetics of the EPO-DAnetworksandaliphaticpolyamidescontinues tonarrowwhen increasingthetemperatureofexposure,andendsupdisappearing almost totallywhen equalizingandexceedingtheglass transition temperatureofEPO-DAnetworks.Asanexample,atT=Tg +30°C fornetworkno.1,k6 andk3 takealmostthesamevaluesthanin aliphaticpolyamides(atT≈ Tg +100°C).

5. Discussion

Thevaluesoftherateconstantk6 determined forthesix EPO-DA networksunderstudyhave beenplottedasa function ofthe reciprocaltemperature1/TinFig.9.Thevaluesfoundforaliphatic polyamidesinapreviousstudy[37]havebeenaddedfor compari-son.Theverticalsolidstraight-linesindicatethepositionoftheTg ofthedifferentmaterials.

Atfirstglance, itwouldseemthat theeffectofmolecular mo-bilitycausesonlyashiftofthedifferentcurvesalongtheabscissa axis.Thus,ifassumingthatk6 obeysanArrheniuslaw,the molecu-larmobilitywouldonlyaffectthepre-exponentialfactorofk6 (i.e.

thefrequencyofchemicaleventsbetweenperoxyradicals)andnot itsactivationenergy.

Tocheck thisassumption, the valuesof k6 were plotted as a functionofthedifference(1/T– 1/Tg ),whichamountsto position-ing1/Tg ofallmaterials ontheordinate axispassingthrough the origin(i.e.at1/T=0).Asexpected,theresultingshiftalongthe ab-sciseaxisofthedifferentcurvesallows obtaining asinglemaster curve fork6 (see Fig. 10), whosethe generalshape can be

satis-fyingly described by the statistical model of Mahieux and Reifs-nider[60,61].Althoughthismodelwasfirstlydevelopedfor simu-latingthetemperaturedependenceofthestoragemodulusinlarge temperaturerange,duetothepresenceofseveralthermodynamic transitionssuchas:secondaryrelaxations,glasstransition,melting, etc.,itseemstobeperfectlysuitedforthisstudy.Aroundtheglass transitiontemperature,thismodelcanberewrittenasfollows:

k6 =k6r +

(

k6v +k6r

)

× Exp





11/T− 1/Tv /Tr − 1/T



m



(18)

whereTv andTr are thetemperaturesdelimiting theactionzone oftheglasstransitionfromthefullyglassyandrubberydomains, respectively; m is a statisticalparameter relatedto the width of theactionzone; andk6v andk6r are thevaluesofk6 inthefully glassyandrubberydomains,respectively.

Itsnumericalapplicationto themastercurve ofFig.10allows the determinationof the statisticalparameter (m≈ 1.3) and the temperaturedependenceofk6 onbothsidesoftheglasstransition temperature.Itwasfoundthatk6 tendstobealmostindependent oftemperatureintherubberydomain:

k6r =1.41010 L.mol−1 .s−1 (19)

Incontrast,k6 obeysanArrheniuslawintheglassydomain:

k6v =1.4105 Exp



−16000 R



1 T− 1 Tg



L.mol−1 .s−1 (20)

(13)

Fig. 9. Arrhenius graphs between 25 and 250 °C of the rate constant k 6 for the six EPO-DA networks under study and aliphatic polyamides. The vertical solid straight-lines

indicate the position of the T g of the different materials.

Fig. 10. Master curve around the glass transition temperature for the rate constant k 6 of EPO-DA networks and aliphatic polyamides. Modeling of the impact of the thermo-

(14)

Fig. 11. Arrhenius graphs between 25 and 250 °C of the rate constant k 3 for the six EPO-DA networks under study and aliphatic polyamides. The vertical solid straight-lines

indicate the position of the T g of the different materials.

It hasbeentried to apply thesame methodologyto rate con-stant k3 .However, in this case, the molecular mobilitydoes not causeasimpleshiftofthedifferentcurvesalongtheabscissaaxis. Indeed, in Fig. 11, it can be noted that some curves are already very close to each other, while the corresponding EPO-DA net-workshaveverydifferentvaluesofTg .Thisthecasefornetworks no.2,3and4ononehand,andnetworksno.5and6ontheother hand.

AccordingtoKorceketal.[18],thepre-exponentialfactork30 of therateconstantk3 dependsaboveall onthedissociationenergy of theC-H, butalsosecondarily onthe reactivityof theinvolved peroxyradicalPO2 .However,intheprevioussection,ithasbeen shown that their reactivity (throughk6) is significantly impacted

by the molecular mobility. As a consequence, k3 is expected to show alittle morecomplicatedvariationwithTg thank6 inboth therubberyandglassydomains(seeEqs.19and21).

Atfirst,the valuesofk3 were plottedasafunction ofthe dif-ference(1/T– 1/Tg ),aspreviouslydonefork6 .Asexpected,the re-sultingshiftalongtheabsciseaxisofthedifferentcurvesdoesnot allowobtainingasinglemastercurvefork3 (seeFig.12).However, iflookingatthevariation ofk30 withthe Tg ofthedifferent ma-terials (mostofthe dataare inglassy state),one foundsthat k30 increasesexponentiallywithTg

k30 ∝ Exp



Tg 30



(21)

Then,thecurveofaliphaticpolyamideswaschosenasthe refer-encecurve.Eq.21wasusedtoshiftalongtheordinateaxisallthe curvesof EPO-DAnetworks, thus obtaining asinglemaster curve fork3 (seeFig.13).Asfork6 ,thegeneralshape ofk3 canbe sat-isfyinglydescribedby thestatisticalmodelofMahieuxand

Reifs-nider[60,61]: k3 =k3r +

(

k3v +k3r

)

× Exp





1/T− 1/Tv 1/Tr − 1/T



m



(22)

wherek3v andk3r arethevaluesofk3 inthefullyglassyand rub-berydomains,respectively.

Itsnumericalapplicationto themastercurve ofFig.13allows thedetermination of thestatistical parameter(m ≈ 0.9)and the temperaturedependenceofk3 onbothsidesoftheglasstransition temperature.Itwasfoundthatk3 obeysanArrheniuslawwiththe sameactivationenergyon bothsidesof theglasstransition tem-perature.Intherubberydomain:

k3r =1.010−1 Exp



Tg− 50 30



Exp



−64000 R



1 T− 1 Tg



L.mol−1 .s−1 (23)

Intheglassydomain:

k3v =1.910−2 Exp



Tg − 50 30



Exp



−64R000



1 T− 1 Tg



L.mol−1 .s−1 (24)

Thescatteringofdatainthe glassydomainappears tobe suf-ficientlynarrow(overonlyonedecade) toallowtheuseofa sin-glekineticmodelforthewholefamilyofEPO-DAnetworks.If ne-glectinganyexperimentalerror,thisscatteringcanbeduetothree sources:

i) Errors in the calculation of the concentration of oxidation sites.

ii) Differencesinthe reactivityofoxidationsites (i.e.between oxy-methylene,amino-methyleneandmethanolgroups)and

(15)

Fig. 12. Result of the shift along the abscise axis of the curves of Fig. 11 . It can be noted an increase k 30 with the T g of the different materials.

Fig. 13. Master curve around the glass transition temperature for the rate constant k 3 of EPO-DA networks and aliphatic polyamides. Modeling of the impact of the thermo-

(16)

therefore also,differences inthe reactivity ofthe resulting peroxyradicals.

iii) A possible loss of aromaticity leading to the formation of highlyoxidizablealiphaticstructuresduringthermalaging. Theimpactoftwofirstassumptionsonthevaluesofk3 should be very limited(of a factor lower than 2) and thus, not allow a full explanation of the width of data cloud. In contrast, the last assumption should be carefully considered, in particular epoxies cross-linkedwithanaromatichardener,becauseitcould theoreti-callyimpactbyseveralordersofmagnitudethevaluesofk3 .Such a radical recombination has already been evidenced in a highly aromaticpolymer,forinstanceinpoly(etheretherketone)[62].

6. Conclusion

Thekineticmodeldevelopedforthethermaloxidationof poly-methylenic substrates in previous studies [22,31-37]was success-fullygeneralizedtoEPO-DAnetworksinthepresentstudy.Its nu-merous parametersweredeterminedby applyinga“stepbystep” procedurecombiningexperimentandsimulation.

It wasfoundthat theoxygen transport propertiesdepend(i.e. oxygendiffusionandsolubility)onthearomaticityoftheEPO-DA network, whereas the rateconstantsof theapparent termination of peroxyradicals (k6 ) andthe propagationof oxidation (k3 ) are highlyimpactedby themolecularmobilityaroundthe glass tran-sition temperature(Tg ). The introductionoftheeffect of molecu-larmobilityintotheArrheniuslawsofk6 andk3 allowed building mastercurvesandfinally,proposingasinglekinetic modelforthe wholefamilyofEPO-DAnetworks.Inaddition,thesharpvariation ofbothrateconstantswithtemperaturewithintheglasstransition zone(i.e.whenpassingfromglassytorubberystate)was success-fully describedusing the statisticalmodel ofMahieux and

Reifs-Table A.1

Values of the rate constants and formation yields used for modeling the thermal oxidation kinetics of networks no. 1 between 25 and 140 °C. Network Epon 828-Jeffamine D230 [T g = 95 °C]

T ( °C) 25 37 50 65 80 95 110 125 140 k 1u (s −1 ) 2.6 × 10 −10 2.0 × 10 −9 1.5 × 10 −8 1.3 × 10 −7 8.9 × 10 −7 5.4 × 10 −6 2.8 × 10 −5 1.3 × 10 −4 5.4 × 10 −4 k 1b (L.mol −1 .s −1 ) 3.3 × 10 −7 1.3 × 10 −6 5.5 × 10 −6 2.4 × 10 −5 9.4 × 10 −5 3.3 × 10 −4 1.0 × 10 −3 3.0 × 10 −3 8.0 × 10 −3 k 2 (L.mol −1 .s −1 ) 10 8 10 8 10 8 10 8 10 8 10 8 10 8 10 8 10 8 k 3 (L.mol −1 .s −1 ) 5.0 × 10 −4 1.1 × 10 −3 2.7 × 10 −3 7.0 × 10 −3 2.8 × 10 −2 2.6 × 10 −1 1.0 2.1 4.6 k 4 (L.mol −1 .s −1 ) 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 k 5 (L.mol −1 .s −1 ) 3.0 × 10 11 3.0 × 10 11 3.0 × 10 11 3.0 × 10 11 3.0 × 10 11 3.0 × 10 11 3.0 × 10 11 3.0 × 10 11 3.0 × 10 11 k 6 (L.mol −1 .s −1 ) 4.9 × 10 4 5.9 × 10 4 7.8 × 10 4 1.1 × 10 5 1.6 × 10 6 1.8 × 10 8 1.2 × 10 9 2.9 × 10 9 6.9 × 10 9 γ1 (%) 60 60 60 60 60 60 60 60 60 γ4 (%) 50 50 50 50 50 50 50 50 50 γ5 (%) 50 50 50 50 50 50 50 50 50 Table A.2

Values of the rate constants and formation yields used for modeling the thermal oxidation kinetics of networks no. 2 between 25 and 140 °C. Network Epon 828-Ancamine 2049 [T g = 150–155 °C]

T ( °C) 25 37 50 65 80 95 110 125 140 k 1u (s −1 ) 2.6 × 10 −10 2.0 × 10 −9 1.5 × 10 −8 1.3 × 10 −7 8.9 × 10 −7 5.4 × 10 −6 2.8 × 10 −5 1.3 × 10 −4 5.4 × 10 −4 k 1b (L.mol −1 .s −1 ) 3.3 × 10 −7 1.3 × 10 −6 5.5 × 10 −6 2.4 × 10 −5 9.4 × 10 −5 3.3 × 10 −4 1.0 × 10 −3 3.0 × 10 −3 8.0 × 10 −3 k 2 (L.mol −1 .s −1 ) 10 8 10 8 10 8 10 8 10 8 10 8 10 8 10 8 10 8 k 3 (L.mol −1 .s −1 ) 6.0 × 10 −4 1.6 × 10 −3 4.0 × 10 −3 1.2 × 10 −2 3.0 × 10 −2 6.8 × 10 −2 1.5 × 10 −1 3.0 × 10 −1 7.0 × 10 −1 k 4 (L.mol −1 .s −1 ) 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 8.0 × 10 11 k 5 (L.mol −1 .s −1 ) 3.0 × 10 11 3.0 × 10 11 3.0 × 10 11 3.0 × 10 11 3.0 × 10 11 3.0 × 10 11 3.0 × 10 11 3.0 × 10 11 3.0 × 10 11 k 6 (L.mol −1 .s −1 ) 2.0 × 10 4 2.4 × 10 4 3.3 × 10 4 4.0 × 10 4 4.8 × 10 4 6.2 × 10 4 9.8 × 10 4 6.5 × 10 5 9.9 × 10 6 γ1 (%) 60 60 60 60 60 60 60 60 60 γ4 (%) 50 50 50 50 50 50 50 50 50 γ5 (%) 50 50 50 50 50 50 50 50 50

nider [60,61],although initially developedfor the storage modu-lus. It is thus clear that the molecular mobilityplays a key role inboththemechanicalbehaviorandinthedurabilityaround the glasstransitiontemperatureofEPO-DAnetworks.

Despitethisapparentsuccess,atleasttwomainchallenges re-main to be met. The effect of aromaticity on oxygen transport propertiesisnotwell understoodandwoulddeservean in-depth andspecificstudyonaseriesofmodelEPO-DAnetworks,butalso on a series of linearpolymers having almost the same chemical composition than EPO-DA networks (e.g. polycarbonate, polysul-fone,etc.).

Similarly, the effect of aromaticity on the thermal oxidation kinetics should be carefully considered because it could explain smallbehavioraldifferencesobserved betweenthedifferent EPO-DA networks in the glassy state. In particular, a possible loss of aromaticity could certainly allow reducing the scattering of data observedonthemastercurveofk3 (Fig.13).

Initscurrentform,thekineticmodelcanbenowusedto eluci-datetheoxidationbehaviorofmanyindustrialmaterialsexhibiting Tg gradients such as: thick structures in composite material and jointsofglueinassembliesofmetalpartswithcompositematerial structures.

DeclarationofCompetingInterest

Theauthorsdeclarethattheyhavenoknowncompeting finan-cialinterestsorpersonalrelationshipsthatcouldhaveappearedto influencetheworkreportedinthispaper.

AppendixA

(17)

References

[1] T. Dyakonov , P.J. Mann , Y. Chen , W.T.K. Stevenson , Thermal analysis of some aromatic amine cured model epoxy resin systems—II: Residues of degradation, Polym. Degrad. Stab. 54 (1) (1996) 67–83 .

[2] P. Musto , G. Ragosta , P. Russo , L. Mascia , Thermal oxidative degradation of epoxy and epoxy-bismaleimide networks: Kinetics and mechanism, Macromol. Chem. Phys. 202 (18) (2001) 3445–3458 .

[3] P. Musto , Two-dimensional FTIR spectroscopy studies on the thermal oxidative degradation of epoxy and epoxy −bis(maleimide) networks, Macromol 36 (9) (2003) 3210–3221 .

[4] B. Mailhot , S. Morlat-Therias , M. Ouhioune , J.-L. Gardette , Study of the degrada- tion of an epoxy/amine resin, Macromol. Chem. Phys. 206 (5) (2005) 575–584 .

[5] F. Delor-Jestin , D. Drouin , P.Y. Cheval , J. Lacoste , Thermal and photochemical ageing of epoxy resin – Influence of curing agents, Polym. Degrad. Stab. 91 (6) (2006) 1247–1255 .

[6] B. Dao , J. Hodgkin , J. Krstina , J. Mardel , W. Tian , Accelerated aging versus real- istic aging in aerospace composite materials. I- The chemistry of thermal aging in a low-temperature-cure epoxy composite, J. Appl. Polym. Sci. 102 (5) (2006) 4291–4303 .

[7] N. Longerias , M. Sebban , P. Palmas , A. Rivaton , J.-L. Gardette , Degradation of epoxy resins under high energy electron beam irradiation: Radio-oxidation, Polym. Degrad. Stab. 92 (12) (2007) 2190–2197 .

[8] C. Galant , B. Fayolle , M. Kuntz , J. Verdu , Thermal and radio-oxidation of epoxy coatings, Prog. Org. Coat. 69 (4) (2010) 322–329 .

[9] Y.-M. Pei , K. Wang , M.-S. Zhan , W. Xu , X.-J. Dhing , Thermal-oxidative aging of DGEBA/EPN/LMPA epoxy system: Chemical structure and thermal-mechanical properties, Polym. Degrad. Stab. 96 (7) (2011) 1179–1186 .

[10] X. Colin , C. Marais , J. Verdu , A new method for predicting the thermal oxida- tion of thermosets. Application to an amine crosslinked epoxy, Polym. Test. 20 (7) (2001) 795–803 .

[11] X. Colin , C. Marais , J. Verdu , Kinetic modelling and simulation of gravimetric curves. Application to the oxidation of bismaleimide and epoxy resins, Polym. Degrad. Stab. 78 (3) (2002) 545–553 .

[12] J. Decelle , N. Huet , V. Bellenger , Oxidation induced shrinkage for thermally aged epoxy networks, Polym. Degrad. Stab. 81 (2) (2003) 239–248 .

[13] X. Colin , C. Marais , J. Verdu , Kinetic modelling of the stabilizing effect of car- bon fibre on thermal ageing thermoset matrix composites, Compos. Sci. Tech- nol. 65 (2005) 117–127 .

[14] M.-C. Lafarie-Frenot , J.-C. Grandidier , M. Gigliotti , L. Olivier , X. Colin , J. Verdu , J. Cinquin , Thermo-oxidation behaviour of composite materials at high temper- atures: A review of research activities carried out within the COMEDI program, Polym. Degrad. Stab. 95 (6) (2010) 965–974 .

[15] X. Colin , B. Fayolle , J. Cinquin , Nouvelles avancées en modélisation cinétique de la thermo-oxydation des matrices époxy-diamines, Mater. Tech.104 (2) (2016) paper no. 202 .

[16] J. Cinquin , X. Colin , B. Fayolle , M. Mille , S. Terekhina , L. Chocinski-Arnault , M. Gigliotti , J.-C. Grandidier , M.-C. Lafarie-Frenot , M. Minervino , C. Cluzel , F. Daghia , P. Ladeveze , F. Zhang , Thermo-oxidation behavior of organic matrix composite materials at high temperatures, Adv. Aircr. Spacecr. Sci. 3 (2) (2016) 171–195 .

[17] E. Ernault , J. Dirrenberger , E. Richaud , B. Fayolle , Prediction of stress induced by heterogeneous oxidation: Case of epoxy/amine networks, Polym. Degrad. Stab. 162 (2019) 112–121 .

[18] S. Korcek , J.H.B. Chenier , J.A. Howard , K.U. Ingold , Absolute rate constants for hydrocarbon autoxidation. XXI. Activation energies for propagation and the correlation of propagation arte constants with carbon-hydrogen bond strengths, Can. J. Chem. 50 (14) (1972) 2285–2297 .

[19] J.-P. Pascault , H. Sautereau , J. Verdu , R.J.J. Williams , Thermosetting Polymers, Marcel Dekker, New York, 2002 .

[20] V. Langlois , L. Audouin , J. Verdu , P. Courtois , Thermooxidative aging of crosslinked linear polyethylene: Stabilizer consumption and lifetime predic- tion, Polym. Degrad. Stab. 40 (3) (1993) 399–409 .

[21] L. Audouin , X. Colin , B. Fayolle , J. Verdu , Sur l’utilisation de la loi d’Arrhenius dans le domaine du vieillissement des polymères, Mater. Techn. 95 (2007) 167–177 .

[22] N. Khelidj , X. Colin , L. Audouin , J. Verdu , C. Monchy-Leroy , V. Prunier , Oxida- tion of polyethylene under irradiation at low temperature and low dose rate. Part II- Low temperature thermal oxidation, Polym. Degrad. Stab. 91 (7) (2006) 1598–1605 .

[23] T.R. Waite , Bimolecular reaction rates in solids and liquids, J. Chem. Phys. 32 (1) (1960) 21–23 .

[24] N.M. Emanuel , A.L. Buchachenko , Chemical Physics of Polymer Degradtion and Stabilization, VNU Science Press, Ultrech, The Netherlands, 1987 .

[25] E.T. Denisov , Polymer oxidation and antioxidant action, in: S. Halim Hamid (Ed.), Handbook of Polymer Degradation, Marcel Dekker, New-York, 20 0 0, pp. 383–419. Chap. 9 .

[26] X. Colin , C. Marais , J. Verdu , Thermal oxidation kinetics for a poly-(bismaleimide), J. Appl. Polym. Sci. 82 (14) (2001) 3418–3430 .

[27] K. Abdeljaoued , Study of matrix thermal oxidation in carbon fibers-PMR-15 composites, PhD thesis, ENSAM, Paris, 1999 .

[28] T. Devanne , A. Bry , L. Audouin , J. Verdu , Radiochemical ageing of an amine cured epoxy network. Part I: change of physical properties, Polymer (Guildf) 46 (1) (2005) 229–236 .

[29] S. Terekhina , M. Mille , B. Fayolle , X. Colin , Oxidation induced changes in vis- coelastic properties of a thermostable epoxy matrix, Polym. Sci. Ser. A 55 (10) (2013) 614–624 .

[30] X. Colin , G. Teyssèdre , M. Fois , in: A. Boudenne, L. Ibos, Y. Candau, S. Thomas (Eds.), Ageing and Degradation of Multiphase Polymer systems, Dans Hand- book of Multiphase Polymer Systems, Vol. 2/2, John Wiley & Sons Ltd, Chich- ester, 2011, pp. 797–841. Chap. 21 .

[31] X. Colin , L. Audouin , J. Verdu , M. Rozental-Evesque , B. Rabaud , F. Martin , F. Bourgine , Aging of polyethylene pipes transporting drinking water disin- fected by chlorine dioxide. I-Chemical aspects, Polym. Eng. Sci. 49 (7) (2009) 1429–1437 .

[32] A. Mikdam , X. Colin , G. Minard , N. Billon , R. Maurin , A kinetic model for pre- dicting the oxidative degradation of additive free polyethylene in bleach desin- fected water, Polym. Degrad. Stab. 146 (2017) 78–94 .

[33] X. Colin , M. Ben Hassine , M. Nait-Abdelaziz , Chemo-mechanical model for pre- dicting the lifetime of EPDM rubbers, Rubber Chem. Technol. 92 (4) (2019) 722–748 .

[34] M. Coquillat , J. Verdu , X. Colin , L. Audouin , R. Nevière , Thermal oxidation of polybutadiene. Part II-Mechanistic and kinetic schemes for additive free un- crosslinked polybutadiene, Polym. Degrad. Stab. 92 (7) (2007) 1334–1342 .

[35] L.K. Nait-Ali , X. Colin , A. Bergeret , Kinetic analysis and modeling of PET macro- molecular changes during its mechanical recycling by extrusion, Polym. De- grad. Stab. 96 (2) (2011) 236–246 .

[36] G. Minard , X. Colin , Thermal ageing of a hybrid composite rod for next gener- ation overhead power lines, J. Compos. Sci. 3 (4) (2019) article no. 103 .

[37] C. El Mazry , M. Ben Hassine , O. Correc , X. Colin , Thermal oxidation kinetics of free additive polyamide 6-6, Polym. Degrad. Stab. 98 (1) (2013) 22–36 .

[38] L. Achimsky , L. Audouin , J. Verdu , J. Rychly , L. Matisova-Rychla ,On a transition at 80 °C in polypropylene oxidation kinetics, Polym. Degrad. Stab. 58 (3) (1997) 283–289 .

[39] C. Decker , F.R. Mayo , H. Richardson , Aging and degradation of polyolefins. III– Polyethylene and ethylene–propylene copolymers, J. Polym. Sci.: Polym. Chem. Ed. 11 (1973) 2879–2898 .

[40] N. Khelidj , X. Colin , L. Audouin , J. Verdu , C. Monchy-Leroy , V. Prunier , Oxida- tion of polyethylene under irradiation at low temperature and low dose rate. Part I- The case of “pure” radiochemical initiation, Polym. Degrad. Stab. 91 (7) (2006) 1593–1597 .

[41] E. Hairer , G. Wanner , Solving Ordinary Differential equations. II-Stiff and Dif- ferential-Algebraic Problems, Springer, Berlin, 1991 .

[42] X. Colin , L. Audouin , J. Verdu , Determination of thermal oxidation rate con- stants by an inverse method. Application to polyethylene, Polym. Degrad. Stab. 86 (2004) 309–321 .

[43] E.A. Di Marzio , On the second-order transition of a rubber, J. Res. Natl. Br. Stand.: Sect. A 68 (1964) 611–617 .

[44] M. Celina , A.R. Dayile , A. Quintana , A perspective on the inherent oxidation sensitivity of epoxy materials, Polymer (Guildf) 54 (2013) 3290–3296 .

[45] E. Girard-Reydet , C.C. Riccardi , H. Sautereau , J.P. Pascault , Epoxy-aromatic amine kinetics. 1- Modeling and influence of diamine structure, Macro- molecules 28 (1995) 7599–7607 .

[46] V. Bellenger , C. Bouchard , P. Claveirolle , J. Verdu , Photo-oxidation of epoxy resins cured by non-aromatic amines, Polym. Photochem. 1 (1) (1981) 69–80 .

[47] V. Bellenger , J. Verdu , Oxidative skeleton breaking in epoxy–amine networks, J. Appl. Polym. Sci. 30 (1) (1985) 363–375 .

[48] A. Rivaton , L. Moreau , J.-L. Gardette , Photo-oxidation of phenoxy resins at long and short wavelengths. I- Identification of the photoproducts, Polym. Degrad. Stab. 58 (3) (1997) 321–332 .

[49] M.St.C. Flett , Intensities of some group characteristic infra-red bands, Spec- trochim. Acta 18 (1962) 1537–1556 .

[50] W.-D. Domke , H. Steinke , Oxidative structures in polyolefins: FT-IR method of quantitative determination, J. Polym. Sci.: Part A: Polym. Chem. 24 (10) (1986) 2701–2705 .

[51] J. Lacoste , D.J. Carlsson , Gamma-, photo-, and thermally-initiated oxidation of linear low density polyethylene: A quantitative comparison of oxidation prod- ucts, J. Polym. Sci.: Part A: Polym. Chem. 30 (1992) 493–500 .

[52] A. Barth , Infrared spectroscopy of proteins, Biochim. Biophys. Acta (2007) (1767) 1073–1101 .

[53] M. Da Cruz , L. Van Schoors , K. Benzari , X. Colin , Thermo-oxidative degradation of additive free polyethylene. Part I- Analysis of the chemical modifications at molecular and macromolecular scales, J. Appl. Polym. Sci. 133 (18) (2016) 43287 .

[54] C. Damian , E. Espuche , M. Escoubes , Influence of three ageing types (thermal oxidation, radiochemical and hydrolytic ageing) on the structure and gas trans- port properties of epoxy-amine networks, Polym. Degrad. Stab. 72 (3) (2001) 447–458 .

[55] C. Damian-Pélissier , Réseaux polyépoxydes utilisés pour l’enrobage des déchets radioactifs : analyse du vieillissement en conditions de stockage et effets sur les propriétés diffusionnelles des matériaux, PhD thesis, Université Claude Bernard, Lyon 1, France, 1999 .

[56] Y. Zahara , Dégradation des réseaux époxy −amine en ambiance nucléaire, PhD thesis, Arts et Métiers ParisTech, Campus de Paris, France, 2012 .

[57] M. Celina , A. Quintana , N. Giron , A. Dayile , An Overview of the Inherent Oxida- tion Sensitivity and DLO Behavior of Epoxy materials, 30th Meeting of PDDG, Arts et Métiers ParisTech, Paris, 1-4 Sept. 2013 .

[58] M. Celina , A. Quintana , Oxygen diffusivity and permeation trough polymers at elevated temperature, Polymer (Guildf) 150 (2018) 326–342 .

(18)

[59] J. Delozanne , Durabilité des époxys. Application au collage structural aéronau- tique, Thèse de doctorat, Arts et Métiers ParisTech, Campus de Paris, France, 2018 .

[60] C.A. Mahieux , K.L. Reifsnider , Property modeling across transition temperatures in polymers : a robust stiffness-temperature model, Polymer (Guildf) 42 (2001) 3281–3291 .

[61] C.A. Mahieux , K.L. Reifsnider , Property modeling across transition temperatures in polymers: Application to thermoplastic systems, J. Mater. Sci. 37 (2002) 911–920 .

[62] S. Giancaterina , A. Rossi , A. Rivaton , J.-L. Gardette , Photochemical evolution of poly(ether ether ketone), Polym. Degrad. Stab. 68 (20 0 0) 133–144 .

Figure

Fig. 1. Changes in  the FTIR spectrum  of network no. 3 during its  thermal aging at 150  °C under 0.21  bar of oxygen (i.e
Fig. 2. Arrhenius  graph  between  10  and  110  °C for the coefficient of oxygen solubility of EPO-DA networks.
Fig. 3. Arrhenius  graph  between 10 and 110  °C for the coefficient  of oxygen diffusion  of EPO-DA networks.
Fig.  4. Carbonyl  build-up  in  100  μm thick  films  of  network no.  6  (Tactix  123/Tactix  742-DDS) at 150  °C  between  0.21  and  20  bars of oxygen
+6

Références

Documents relatifs

the rainy season (June to October), the transition periods (October to December, April to June) and the completely dry period (January to April), where a constant value of

Domain-independent implicit and explicit transformations for Visual Data Mining using Differential Evolution and Genetic Programming aiming at preserving the similarity structure of

Pour ce faire, les objectifs poursuivis sont : 1 connaître les perceptions des infirmières par rapport à l’évolution de leur pratique au cours des années, leurs stratégies de

- Final configurations of binary alloy at low-temperature 0.08To with (a) Lennard-Jones potential, (b) Lennard-Jones potential supplemented by orientational three body

Mineralogical Society of America, Geochemical Society 32, 190-246.. Non-Newtonian rheology of igneous melts at high stresses and strain rates: Experimental results for

The predictions were confronted to the measurement of the in vitro digestibility of myofibrillar proteins coming from either slices of beef meat heated in water bath or

The thermal oxidation of polyethylene (PE) impregnated by the methyl esters of unsaturated fatty acids (UFEs) was studied using chemiluminescence, and infra-red spectrophotometry?.

The decomposition of the material occurs in three apparent steps: de- composition of M21 into a carbonaceous residue (1 st peak), thermo- oxidation of the intermediate