02.06.2013 Views

Analysis and Design of Sandwich Panels with Brittle Matrix ...

Analysis and Design of Sandwich Panels with Brittle Matrix ...

Analysis and Design of Sandwich Panels with Brittle Matrix ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Vrije Universiteit Brussel<br />

Faculteit Toegepaste Wetenschappen<br />

Mechanica van Materialen en Constructies<br />

Pleinlaan 2, 1050 Brussel<br />

<strong>Analysis</strong> <strong>and</strong> <strong>Design</strong> <strong>of</strong> S<strong>and</strong>wich <strong>Panels</strong><br />

<strong>with</strong> <strong>Brittle</strong> <strong>Matrix</strong> Composite Faces<br />

for Building Applications<br />

ir. Heidi Cuypers<br />

Proefschrift ingediend voor het behalen van de graad van Doctor in<br />

de Toegepaste Wetenschappen<br />

Promotor: Pr<strong>of</strong>. dr. ir. Jan Wastiels november 2001


Dankwoord<br />

Het is in de eerste plaats aan mijn promotor Pr<strong>of</strong>. Jan Wastiels te danken dat dit<br />

werk onstaan is. De wekelijkse samenkomsten zijn dikwijls een goede motivatie<br />

geweest om verder te gaan en nieuw opgekomen vragen te bestuderen. Hij staat<br />

bekend, <strong>of</strong> beter gezegd berucht, als een kritische geest bij de studenten en heeft<br />

die reputatie alle eer aangedaan gedurende de vier jaar dat ik met hem heb<br />

gewerkt. Ondanks vele <strong>and</strong>ere verplichtingen heeft hij toch de tijd weten te vinden<br />

voor een goede begeleiding.<br />

Myriam Bourlau en Agnes Vanaeken zorgden altijd voor snelle en goede<br />

administratieve ondersteuning en een schouderklopje op zijn tijd. Rene Heremans,<br />

Daniël Debondt en Gabriël Van den Nest zijn een grote hulp geweest in het<br />

praktische aspect van dit werk. Adri Vrijdag heeft voor enkele geslaagde foto’s<br />

gezorgd. Om de inbreng van Frans Boulpaep te beschrijven, komen woorden te<br />

kort. Zelden ben ik iem<strong>and</strong> met groter enthousiasme voor zijn werk en<br />

hulpvaardigheid ten opzichte van zijn collega’s tegengekomen.<br />

Gu Jun en Kurt De Pr<strong>of</strong>t zijn altijd zeer t<strong>of</strong>fe bureaugenoten geweest. Jun heeft me<br />

op tijd en stond de humor van mijn werk laten inzien, ook al leek die soms ver<br />

weg. Net als Myriam en Agnes is ze dikwijls mijn grote zus geweest. Met Kurt<br />

heb ik fijne gesprekken gehad, zowel over het werk als over meer wereldse<br />

onderwerpen. Hem wens ik verder het allerbeste met de afwerking van zijn<br />

doctoraat.<br />

Jun, Joëlle, Jan, Joeri, Ronald, Kurt en Michaël hebben zich met veel moed door<br />

de soms nogal droge st<strong>of</strong> gewerkt om fouten op te sporen. Ik heb hier alle<br />

appreciatie voor. Gu Jun en Mazen Alshaer hebben in de laatste fase van dit werk<br />

een extra duwtje in de rug gegeven door hun medewerking in het uitvoeren van<br />

enkele experimenten.<br />

Verder hebben collega-assisten Kris Hoes, Jurgen Maes, Pascal Bouquet en Kim<br />

Croes gedurende drie <strong>of</strong> vier jaar voor een t<strong>of</strong>fe werkomgeving gezorgd. Ik wens


hun ook het beste met de verderzetting van hun doctoraat <strong>of</strong> succes in hun nieuwe<br />

job. Ook de nieuwkomers Gunther, Georgios, Jan, Danniëlla en Shi wens ik succes<br />

met hun toekomstig onderzoek.<br />

Ik heb gedurende al die jaren veel steun gehad van mijn vader en Rie-Jeanne en<br />

mijn moeder en Karel. Mijn broer Stefan is dikwijls een rustpunt voor me geweest<br />

in de drukke schrijfperiode, ook al besefte hij het waarschijnlijk niet altijd. Ik<br />

wens hem en Isabelle het beste toe. Sommige personen van de oude garde en het<br />

nieuwe bloed binnen BEST-Brussel zijn me ook dierbaar gebleven <strong>of</strong> geworden.<br />

Enkele mensen die ik op Aero-Kiewit heb leren kennen, hebben de laatste twee<br />

jaren beduidend aangenamer gemaakt. Vrienden die ik al ettelijke jaartjes langer<br />

ken, hebben mijn vorderingen met interesse gevolgd: Frank, Ann, Geert, Aldo en<br />

Imogen, Bert, Wil, Maja, Joeri, Nico en Karen en Ronald. Vooral Ronald heeft me<br />

enorm gesteund. Een betere vriend kan je je jezelf niet toewensen.<br />

Heidi Cuypers,<br />

november 2001


Symbols <strong>and</strong> abbreviations<br />

Symbols concerning representative length scales<br />

. along <br />

.<br />

.<br />

.<br />

along < cs><br />

. along ( su)<br />

.<br />

along ( su)<br />

N<br />

. alongδ<br />

0<br />

alongδ N<br />

along <br />

−2δ<br />

0<br />

f<br />

average <strong>of</strong> variable along <br />

average <strong>of</strong> variable along f<br />

average <strong>of</strong> variable along (su)<br />

average <strong>of</strong> variable along (su)N<br />

average <strong>of</strong> variable along δ0<br />

average <strong>of</strong> variable along δN<br />

average <strong>of</strong> variable along -2δ0<br />

Symbols concerning strain<br />

∆ε el elastic strain increment<br />

∆ε pl plastic strain increment<br />

∆εc<br />

composite strain variation<br />

∆εc elastic<br />

elastic term <strong>of</strong> composite strain variation<br />

∆εc slip slip term <strong>of</strong> composite strain variation<br />

∆εc zoneII extra composite strain term, due to multiple cracking<br />

composite strain variation, cycle N<br />

∆εc,N<br />

∆εc,N elastic elastic term <strong>of</strong> composite strain variation, cycle N<br />

∆εc,N repeat<br />

extra strain variation term, introduced by repeated loading<br />

∆εc,N slip slip term <strong>of</strong> composite strain variation, cycle N<br />

∆εf<br />

fibre strain variation<br />

∆εf elastic elastic term <strong>of</strong> fibre strain variation


Symbols <strong>and</strong> abbreviations<br />

∆εf slip slip term <strong>of</strong> fibre strain variation<br />

∆εf,N fibre strain variation, cycle N<br />

∆εf,N slip<br />

slip term <strong>of</strong> fibre strain variation, cycle N<br />

∆εm<br />

matrix strain variation<br />

εc<br />

composite strain<br />

εc max<br />

composite strain at maximum composite cycle stress<br />

εc min<br />

composite strain at minimum composite cycle stress<br />

εc thermal composite strain due to thermal load<br />

zone III<br />

εc<br />

εc zoneII εc,N<br />

composite strain after multiple cracking<br />

max εc,N<br />

maximum cycle strain, cycle N<br />

min minimum cycle strain, cycle N<br />

εc,x<br />

composite strain in zone III, ACK theory<br />

composite strain, measured in x-direction<br />

composite strain, measured in z-direction<br />

εc,z<br />

εexp(σj) experimental strain, at stress point σj<br />

εf<br />

εf,N max<br />

εm<br />

εm,x<br />

εm,z<br />

εmc<br />

fibre strain<br />

fibre strain at maximum composite stress, cycle N<br />

matrix strain<br />

matrix strain, measured in x-direction<br />

matrix strain, measured in z-direction<br />

composite multiple cracking strain<br />

ultimate matrix strain<br />

εmu<br />

εtheo(σj) theoretical strain, at stress point σj<br />

Symbols concerning stress<br />

∆σc<br />

composite stress variation<br />

∆σf<br />

fibre stress variation<br />

∆σf slip slip term <strong>of</strong> fibre stress variation<br />

matrix stress variation, cycle N<br />

∆σf,N<br />

∆σf,N slip slip term <strong>of</strong> fibre stress variation, cycle N<br />

∆σm matrix stress variation<br />

∆σm elastic elastic term <strong>of</strong> matrix stress variation<br />

∆σm slip slip term <strong>of</strong> matrix stress variation<br />

∆σm,N elastic elastic term <strong>of</strong> matrix stress variation, cycle N<br />

∆σm,N slip slip term <strong>of</strong> matrix stress variation, cycle N<br />

σ normal stress<br />

σ0<br />

initial failure stress (parameter in Weibull model)<br />

initial bending failure stress (parameter in Weibull model)<br />

σ0b


σc<br />

σc max<br />

σc min<br />

σcu<br />

σf<br />

σf,max<br />

σf,min<br />

Symbols <strong>and</strong> abbreviations<br />

composite stress<br />

maximum composite cycle stress<br />

minimum composite cycle stress<br />

composite failure stress<br />

fibre stress<br />

maximum stress in fibres<br />

minimum stress in fibres<br />

σf,N max fibre stress at maximum composite stress, cycle N<br />

σf,N min fibre stress at minimum composite stress, cycle N<br />

σfu<br />

σi<br />

σ y i-<br />

σ y i+<br />

σij<br />

σ y ijσ<br />

y ij+<br />

σj<br />

failure stress <strong>of</strong> the fibres<br />

normal stress in i-direction (i = x, y or z)<br />

normal yield stress in i-direction, compression (i = x, y or z)<br />

normal yield stress in i-direction, tension (i = x, y or z)<br />

shear stress in ij-plane (ij = xy, yz or xz)<br />

shear yield stress in ij-plane, negative sign (ij = xy, xz or yz)<br />

shear yield stress in ij-plane, positive sign (ij = xy, xz or yz)<br />

experimental stress point j<br />

matrix stress<br />

far field matrix stress (far from crack)<br />

σm<br />

σm ff<br />

σm ff,max<br />

far field matrix stress at maximum composite stress<br />

σm max matrix stress at maximum composite cycle stress<br />

σm thermal matrix stress due to thermal load<br />

σmax<br />

σmc<br />

σm,max<br />

σm,max max<br />

σm,min<br />

maximum stress<br />

composite multiple cracking stress<br />

maximum stress in matrix under static load<br />

maximum matrix stress at maximum composite stress<br />

minimum stress in matrix under static load<br />

σm,N ff,max far field matrix stress at maximum composite stress, cycle N<br />

σm,N max<br />

σm,N<br />

matrix stress at maximum composite stress, cycle N<br />

min<br />

σmu<br />

matrix stress at minimum composite stress, cycle N<br />

ultimate failure stress <strong>of</strong> the matrix<br />

σm,x matrix stress, measured in x-direction<br />

σnom nominal stress<br />

σR<br />

reference failure stress (parameter in Weibull model)<br />

σRb<br />

reference bending failure stress (parameter in Weibull model)<br />

σwr<br />

wrinkling stress<br />

τ shear stress<br />

frictional matrix-fibre interface shear stress, first loading<br />

τ0<br />

τ0 x shear stress component along x-axis<br />

τ0sf<br />

single fibre frictional matrix-fibre shear stress


τa<br />

τau<br />

τd0<br />

τi0<br />

τN<br />

τu<br />

Other symbols<br />

Symbols <strong>and</strong> abbreviations<br />

adhesion matrix-fibre shear bond<br />

adhesion matrix-fibre shear bond strength<br />

dynamic frictional shear stress (matrix-fibre interface)<br />

frictional shear stress (matrix-fibre interface)<br />

frictional matrix-fibre interface shear stress, cycle N<br />

ultimate shear stress<br />

α matrix / fibre stiffness ratio <strong>of</strong> the composite<br />

αf<br />

thermal expansion coefficient <strong>of</strong> the fibres<br />

αm<br />

thermal expansion coefficient <strong>of</strong> the matrix<br />

βl<br />

variable used to calculate fibre length efficiency factor<br />

δt time increment<br />

δ0<br />

<strong>Matrix</strong>-fibre debonding length at first loading<br />

δf<br />

length along which normal stresses are transferred to fibres<br />

δN<br />

debonding length, cycle N<br />

∆T temperature difference<br />

γG<br />

safety coefficient on permanent load<br />

γQ<br />

safety coefficient on variable load<br />

ηθ<br />

orientation efficiency factor<br />

ηl<br />

length efficiency factor<br />

ηlev<br />

parameter in Largest extreme value probability distribution<br />

ϕcot<br />

creep coefficient <strong>of</strong> the core<br />

µi<br />

snow shape coefficient<br />

µlev<br />

parameter in Largest extreme value probability distribution<br />

µlog<br />

parameter in Lognormal probability distribution<br />

νm<br />

Poisson’s ratio <strong>of</strong> the matrix<br />

νfxz<br />

Poisson’s ratio <strong>of</strong> the fibres<br />

νxz<br />

Poisson’s ratio <strong>of</strong> the composite<br />

θ local cylindrical coordinate<br />

θx<br />

angle between fibre <strong>and</strong> x-axis<br />

ρ local cylindrical coordinate<br />

ρf<br />

density <strong>of</strong> glass fibres<br />

σlog<br />

parameter in Lognormal probability distribution model<br />

ω matrix-fibre interface degradation parameter<br />

ψ0,i<br />

combination coefficient<br />

ψ1,i<br />

combination coefficient<br />

combination coefficient<br />

ψ2,i


Symbols <strong>and</strong> abbreviations<br />

A constant, concerning contribution <strong>of</strong> term(s) to bending stiffness<br />

Ac<br />

composite section, transverse to the loading direction<br />

Aexposed exposed bundle surface area<br />

Af<br />

Af<br />

fibre section (Af = VfAc)<br />

* equivalent fibre section<br />

Am<br />

matrix section (Am=VmAc)<br />

Atotal total bundle surface area<br />

b width<br />

average crack spacing<br />

f average crack spacing at the end <strong>of</strong> matrix multiple cracking<br />

C1 constant in formulation <strong>of</strong> interface degradation parameter<br />

C2 constant in formulation <strong>of</strong> interface degradation parameter<br />

cd<br />

dynamic coefficient (wind load)<br />

ce<br />

exposure coefficient (snow load)<br />

cpe<br />

external pressure coefficient (wind load)<br />

cpi<br />

internal pressure coefficient (wind load)<br />

ct<br />

thermal coefficient (snow load)<br />

D bending stiffness <strong>of</strong> s<strong>and</strong>wich panel<br />

Df<br />

sum <strong>of</strong> the separate flexural rigidities <strong>of</strong> the faces<br />

DQ<br />

shear stiffness <strong>of</strong> the core<br />

e distance between the centroid <strong>of</strong> the core <strong>and</strong> the neutral axis<br />

Ec<br />

composite stiffness<br />

Ec1<br />

composite stiffness as determined by the rule <strong>of</strong> mixtures<br />

Ec3<br />

experimental obtained composite stiffness in post-cracking zone<br />

Eco<br />

E-modulus <strong>of</strong> core<br />

Ecycle linearised stiffness <strong>of</strong> one loading-unloading stress-strain cycle<br />

Ecycle,N linearised stiffness, cycle N<br />

E el<br />

elastic stiffness<br />

Ef<br />

E-modulus <strong>of</strong> fibre<br />

Efa<br />

E-modulus <strong>of</strong> face<br />

Ef b stiffness <strong>of</strong> lower face (bottom)<br />

Ef t stiffness <strong>of</strong> upper face (top)<br />

Ei,+ pl<br />

plastic stiffness in i direction, tension (i = x, y or z)<br />

Ei,+ pl plastic stiffness in i direction, compression (i = x, y or z)<br />

Ei, el elastic stiffness in i direction (i = x, y or z)<br />

Em<br />

E-modulus <strong>of</strong> matrix<br />

E pl<br />

plastic stiffness<br />

E T<br />

tangent stiffness<br />

F cumulative distribution function<br />

f probability distribution function


Symbols <strong>and</strong> abbreviations<br />

F x external external applied force parallel <strong>with</strong> the x-axis<br />

F x fibre total force in the x-direction taken by all fibres<br />

F x matrix total force in the x-direction, taken by the matrix<br />

F y external external applied force parallel <strong>with</strong> the y-axis<br />

F y fibre total force in the y-direction taken by all fibres<br />

F y matrix total force in the y-direction, taken by the matrix<br />

Gco<br />

shear modulus <strong>of</strong> the core<br />

Gij el<br />

elastic shear modulus ij plane (ij = xy, xz or yz)<br />

Gij pl<br />

plastic shear modulus ij plane (ij = xy, xz or yz)<br />

Gk<br />

permanent load<br />

Gm<br />

shear modulus <strong>of</strong> the matrix<br />

h height <strong>of</strong> specimen<br />

K efficiency factor in post-cracking zone<br />

Ku<br />

unloading efficiency factor<br />

l length <strong>of</strong> specimen<br />

L span <strong>of</strong> s<strong>and</strong>wich panel<br />

lf<br />

length <strong>of</strong> fibre<br />

m Weibull modulus (parameter in Weibull model)<br />

Mb<br />

moment in lower face (bottom)<br />

Mt<br />

moment in upper face (top)<br />

N number <strong>of</strong> applied load cycles<br />

Nb<br />

normal force in lower face<br />

Nexp number <strong>of</strong> experimental points in considered stress interval<br />

Nfb<br />

number <strong>of</strong> fibres in a bundle<br />

Nfibres number <strong>of</strong> fibres in a composite section<br />

nfl<br />

number <strong>of</strong> fibre layers<br />

Nt<br />

normal force in upper face<br />

P probability<br />

p pressure load on s<strong>and</strong>wich panel<br />

Pfibre force taken by one fibre<br />

P x fibre force in the x-direction taken by one fibre<br />

P y fibre force in the y-direction taken by one fibre<br />

qk<br />

characteristic wind load pressure<br />

Qk<br />

variable load<br />

Qk1<br />

variable load <strong>with</strong> largest effect<br />

Qk2<br />

other variable loads<br />

r fibre radius<br />

R radius <strong>of</strong> matrix around the fibre<br />

Rav<br />

ratio <strong>of</strong> average composite strains in two transverse directions<br />

rotz<br />

rotation around the z-axis


Symbols <strong>and</strong> abbreviations<br />

s snow load pressure on panel<br />

(su) unloading slip length<br />

(su)N unloading slip length, cycle N<br />

Sd<br />

equivalent static load, used for design<br />

sk<br />

characteristic snow pressure on the ground<br />

t time<br />

T0 x<br />

frictional shear flux along the x-axis<br />

t1<br />

initial time<br />

tc<br />

thickness <strong>of</strong> laminate or specimen<br />

tc<br />

thickness <strong>of</strong> core<br />

tf b lower face thickness (bottom)<br />

tf t upper face thickness (top)<br />

Tinside temperature inside building<br />

Toutside temperature outside building<br />

ub<br />

displacement <strong>of</strong> lower face (x-direction)<br />

uc<br />

displacement <strong>of</strong> core (x-direction)<br />

ut<br />

displacement <strong>of</strong> upper face (x-direction)<br />

ux<br />

displacement in the x-direction<br />

uy<br />

displacement in the y-direction<br />

V volume<br />

Vb<br />

shear force in lower face<br />

Vf<br />

fibre volume fraction<br />

Vf *<br />

equivalent fibre volume fraction<br />

Vfb<br />

volume fraction <strong>of</strong> fibres in a bundle<br />

Vf crit<br />

critical fibre volume fraction<br />

Vm<br />

matrix volume fraction<br />

VR<br />

reference volume<br />

Vt<br />

shear force in upper face<br />

w width <strong>of</strong> specimen<br />

w wind load pressure on panel (Chapter 7)<br />

wb<br />

displacement <strong>of</strong> lower face (y-direction)<br />

wfibrelayer weight <strong>of</strong> fibre layer (mat or weave) per m²<br />

wt<br />

displacement <strong>of</strong> upper face (y-direction)<br />

X general variable<br />

local cylindrical coordinate<br />

xl


Abbreviations<br />

Symbols <strong>and</strong> abbreviations<br />

ACK Aveston-Cooper-Kelly model<br />

CDF cumulative distribution function<br />

CLC characteristic load combination<br />

TMA thermomechanical analyser<br />

EST elementary s<strong>and</strong>wich theory<br />

FEM finite element method<br />

FLC frequent load combination<br />

IPC inorganic phosphate cement<br />

LEV largest extreme value model<br />

LSC least square coefficient<br />

LVDT linear variable differential transducer<br />

ncore number <strong>of</strong> element divisions across height <strong>of</strong> a core<br />

ndiv number <strong>of</strong> element divisions along length <strong>of</strong> a s<strong>and</strong>wich panel<br />

nface number <strong>of</strong> element divisions across height <strong>of</strong> a face<br />

PDF probability distribution function<br />

QLC quasi-permanent load combination<br />

SEV smallest extreme value model<br />

SLS serviceability limit state<br />

UD unidirectional reinforcement<br />

ULS ultimate limit state


Dankwoord<br />

Symbols <strong>and</strong> abbreviations<br />

Table <strong>of</strong> contents<br />

Table <strong>of</strong> contents<br />

Chapter 1: Introduction 1<br />

1.1 Research context 1<br />

1.2 Objectives <strong>of</strong> the thesis 2<br />

1.3 Basic assumptions in this thesis 3<br />

1.4 Outline <strong>of</strong> the thesis 4<br />

1.5 References 6<br />

Chapter 2: IPC composite specimens under monotonic loading 9<br />

2.1 Introduction 9<br />

2.2 Materials 10<br />

2.2.1 E-glass fibres as reinforcement 10<br />

2.2.2 matrix material: Inorganic Phosphate Cement (IPC) 10<br />

2.2.3 E-glass fibre reinforced IPC laminates 11<br />

2.3 IPC laminates under compressive loading 12<br />

2.3.1 introduction 12<br />

2.3.2 test set-up 12<br />

2.3.3 materials <strong>and</strong> results 13<br />

2.3.4 discussion <strong>and</strong> conclusions 15<br />

2.4 Tensile behaviour <strong>of</strong> continuous UD-reinforced IPC 17<br />

2.4.1 introduction 17<br />

2.4.2 basic assumptions 17<br />

2.4.3 formulation <strong>of</strong> the ACK theory 19<br />

2.4.3.1 zone I 19<br />

2.4.3.2 zone II 20<br />

2.4.3.3 zone III 25<br />

2.4.4 points <strong>of</strong> discussion in the ACK theory 26<br />

2.4.5 the statistical nature <strong>of</strong> the IPC matrix tensile strength 28<br />

2.4.6 the Weibull probability distribution 32


Table <strong>of</strong> contents<br />

2.4.7 formulation <strong>of</strong> the stochastic cracking theory 34<br />

2.4.7.1 formulation when > 2δ0<br />

36<br />

2.4.7.2 formulation when < 2δ0<br />

37<br />

2.4.7.3 σc at which > 2δ0 becomes < 2δ0<br />

38<br />

2.4.8 experiments on UD-reinforced IPC specimens 38<br />

2.4.8.1 specimens 38<br />

2.4.8.2 ACK model versus experimental results 39<br />

2.4.8.3 stochastic cracking model versus experimental results 40<br />

2.4.8.4 ACK model versus stochastic cracking model 45<br />

2.5 Tensile behaviour <strong>of</strong> 2D-r<strong>and</strong>omly reinforced IPC 46<br />

2.5.1 introduction 46<br />

2.5.2 derivation <strong>of</strong> the ACK based model 47<br />

2.5.2.1 linear elastic zone or pre-cracking zone (zone I) 47<br />

2.5.2.2 post-cracking zone (zone III) 48<br />

2.5.2.3 multiple cracking zone (zone II) 50<br />

2.5.3 derivation <strong>of</strong> the stochastic cracking based model 53<br />

2.5.3.1 formulation when > 2δ0<br />

53<br />

2.5.3.2 formulation when < 2δ0<br />

53<br />

2.5.4 experiments on 2D-r<strong>and</strong>omly reinforced IPC specimens 54<br />

2.5.4.1 specimens 54<br />

2.5.4.2 ACK based model versus experimental results 54<br />

2.5.4.3 stochastic cracking based model versus experimental results 55<br />

2.5.4.4 ACK based model versus stochastic cracking based model 56<br />

2.6 Tensile behaviour <strong>of</strong> IPC composite specimens: discussion <strong>and</strong> conclusions 57<br />

2.7 Evolution <strong>of</strong> matrix stresses transverse to the loading direction 58<br />

2.7.1 theoretical derivation <strong>of</strong> the Poisson’s ratio based on the ACK theory 59<br />

2.7.2 experimental verification 61<br />

2.8 Experimental determination <strong>of</strong> the shear modulus <strong>of</strong> IPC composite specimens 63<br />

2.8.1 specimens <strong>and</strong> test set-up 63<br />

2.8.2 results 64<br />

2.9 Conclusions 65<br />

2.10 References 67<br />

Chapter 3: Unloading <strong>of</strong> IPC composite specimens 71<br />

3.1 Introduction 71<br />

3.2 Definitions <strong>and</strong> assumptions 71<br />

3.3 ACK (based) theory for unloading 73<br />

3.3.1 introduction 73<br />

3.3.2 derivation <strong>of</strong> the linearised E-modulus <strong>of</strong> a cycle: Ecycle<br />

3.3.2.1 partial matrix-fibre slip during unloading 73<br />

3.3.2.2 from partial to total matrix-fibre slip 76<br />

3.3.2.3 total matrix-fibre slip during unloading 76<br />

3.4 stochastic cracking (based) theory for unloading 77<br />

73


Table <strong>of</strong> contents<br />

3.4.1 introduction 77<br />

3.4.2 derivation <strong>of</strong> linearised E-modulus <strong>of</strong> a cycle: Ecycle<br />

78<br />

3.4.2.1 partial matrix-fibre slip during unloading 78<br />

3.4.2.2 from partial to full matrix-fibre unloading slip, > 2δ0 80<br />

3.4.2.3 from partial to full unloading matrix-fibre slip, < 2δ0 83<br />

3.4.2.4 total matrix-fibre slip during unloading 83<br />

3.5 IPCstress-strain.exe: a program to predict unloading 85<br />

3.6 Experiments 86<br />

3.6.1 test set-up <strong>and</strong> materials 86<br />

3.6.2 test results 87<br />

3.6.3 unloading behaviour as predicted by the ACK based model 88<br />

3.6.4 unloading behaviour as predicted by the stochastic cracking based 93<br />

3.6.5 ACK based model versus stochastic cracking based model 94<br />

3.7 Conclusions 94<br />

3.8 References 95<br />

Chapter 4: IPC Composite specimens under repeated tensile loading 99<br />

4.1 Introduction 99<br />

4.2 Theoretical derivation: general remarks 100<br />

4.2.1 introduction 100<br />

4.2.2 assumptions <strong>and</strong> definitions 101<br />

4.3 Theoretical derivation <strong>of</strong> a fatigue theory from the ACK (based) model 103<br />

4.3.1 evolution <strong>of</strong> εc,N max <strong>with</strong> N 103<br />

4.3.2 evolution <strong>of</strong> Ecycle,N <strong>with</strong> N 106<br />

4.3.2.1 partial matrix-fibre unloading slip 106<br />

4.3.2.2 when partial becomes total matrix-fibre slip 108<br />

4.3.2.3 total matrix-fibre unloading slip 108<br />

4.4 Theoretical derivation <strong>of</strong> a fatigue theory from the stochastic cracking (based) 109<br />

4.4.1 evolution <strong>of</strong> εc,N max <strong>with</strong> N 110<br />

4.4.1.1 crack spacing exceeds twice the debonding length 110<br />

4.4.1.2 transition from >2δN to < 2δN<br />

111<br />

4.4.1.3 crack spacing smaller than twice the debonding length 112<br />

4.4.2 evolution <strong>of</strong> Ecycle,N <strong>with</strong> N 112<br />

4.4.2.1 partial matrix-fibre unloading slip 112<br />

4.4.2.2 partial to total matrix-fibre unloading slip, > 2δN 112<br />

4.4.2.2 partial to total matrix-fibre unloading slip, < 2δN 113<br />

4.4.2.4 total matrix-fibre unloading slip 113<br />

4.5 Determination <strong>of</strong> interface wear evolution from experimental observations 113<br />

4.5.1 introduction 113<br />

4.5.2 ACK (based) theory 114<br />

4.5.3 stochastic cracking (based) theory 114<br />

4.5.4 ACK (based) theory versus stochastic cracking (based) theory 114<br />

4.6 Testing program 115


Table <strong>of</strong> contents<br />

4.7 UD-reinforced IPC composite specimens under repeated loading 116<br />

4.7.1 introduction 116<br />

4.7.2 material properties <strong>and</strong> test schedule 116<br />

4.7.3 determination <strong>of</strong> material properties 116<br />

4.7.4 results under repeated loading 117<br />

4.7.5 determination <strong>of</strong> interface degradation parameter from test results 118<br />

4.7.6 discussion <strong>and</strong> conclusions 120<br />

4.8 UD-reinforced IPC composite specimen under limited repeated loading 122<br />

4.8.1 introduction 122<br />

4.8.2 material properties <strong>and</strong> test schedule 122<br />

4.8.3 determination <strong>of</strong> material properties 123<br />

4.8.4 results under repeated loading 123<br />

4.8.5 determination <strong>of</strong> interface degradation parameter from test results 124<br />

4.8.6.1 extrapolation from limited cycling to high number <strong>of</strong> cycles 124<br />

4.8.6.2 repeated loading at lower maximum cycle stress 128<br />

4.9 UD-reinforced IPC composites under repeated loading: discussion 130<br />

4.10 2D-r<strong>and</strong>omly reinforced IPC composite specimens under repeated loading 131<br />

4.10.1 introduction 131<br />

4.10.2 material properties <strong>and</strong> test schedule 131<br />

4.10.3 determination <strong>of</strong> material properties 131<br />

4.10.4 results under repeated loading 132<br />

4.10.5 theoretical versus experimental degradation behaviour 135<br />

4.10.5.1 results 135<br />

4.10.5.2 specimens cycled up to 22 or 27.5MPa: discussion 135<br />

4.10.5.3 specimens cycled up to 11 or 16.5MPa: discussion 136<br />

4.10.5.4 specimens cycled up to 6MPa: discussion 137<br />

4.10.6 discussion <strong>and</strong> conclusions 138<br />

4.11 Interpretation <strong>of</strong> hysteresis 139<br />

4.12 Conclusions 140<br />

4.12.1 conclusions from tests on UD-reinforced specimens 140<br />

4.12.2 conclusions from tests on 2D-r<strong>and</strong>omly -reinforced specimens 141<br />

4.13 References 141<br />

Chapter 5. S<strong>and</strong>wich modelling 143<br />

5.1 Introduction 143<br />

5.2 S<strong>and</strong>wich modelling: overview 143<br />

5.2.1 single layer theories 144<br />

5.2.2 layer-wise theories 144<br />

5.2.3 predictor-corrector <strong>and</strong> hierarchical modelling theories 145<br />

5.3 A 2D-continuum approach: plane stress versus plane strain 145<br />

5.4 S<strong>and</strong>wich panels under monotonic loading: the use <strong>of</strong> ANSYS 146<br />

5.4.1 introduction 146<br />

5.4.2 implementation <strong>of</strong> material behaviour 146


Table <strong>of</strong> contents<br />

5.4.2.1 implementation <strong>of</strong> material behaviour: ‘multilinear elastic’ 146<br />

5.4.2.2 implementation <strong>of</strong> material behaviour: ‘aniso’ 147<br />

5.4.3 type <strong>of</strong> model: discussion 150<br />

5.4.3.1 elementary s<strong>and</strong>wich theory 150<br />

5.4.3.2 advanced s<strong>and</strong>wich theory 153<br />

5.4.4 parameters <strong>of</strong> study 155<br />

5.4.5 a reference s<strong>and</strong>wich beam 156<br />

5.4.6 comparison <strong>of</strong> models: results <strong>and</strong> discussion 158<br />

5.4.7 influence <strong>of</strong> span 160<br />

5.5 The behaviour <strong>of</strong> s<strong>and</strong>wich panels during unloading <strong>and</strong> repeated loading 163<br />

5.7 Conclusions 164<br />

5.8 References 165<br />

Chapter 6: Experimental work on s<strong>and</strong>wich panels 169<br />

6.1 Introduction 169<br />

6.2 Strength <strong>of</strong> the core-face interface 169<br />

6.2.1 failure in tension 169<br />

6.2.2 shear failure 170<br />

6.3 Large-span s<strong>and</strong>wich panels: materials <strong>and</strong> geometry 172<br />

6.3.1 face materials 172<br />

6.3.2 core material 172<br />

6.3.3 geometry 172<br />

6.4 <strong>Panels</strong> <strong>with</strong> 2m span: test set-up 173<br />

6.5 <strong>Panels</strong> <strong>with</strong> 2m span: results 174<br />

6.6 2m span panels: theoretical predictions versus experimental results 176<br />

6.6.1 face properties <strong>of</strong> the tested panel 176<br />

6.6.2 theoretical prediction <strong>of</strong> the load-displacement behaviour 178<br />

6.6.3 theoretical versus experiments: initial loading up to 2/3 rd <strong>of</strong> maximum 179<br />

6.6.3.1 load-displacement curves 179<br />

6.6.3.2 load-strain curves 180<br />

6.6.4 theory versus experiments: initial unloading from 2/3 rd <strong>of</strong> maximum 182<br />

6.6.5 theory versus experiments: repeated loading up to 2/3 rd <strong>of</strong> maximum 183<br />

6.6.6 theory versus experiments: loading to maximum load 185<br />

6.7 Short-span panels: geometry <strong>and</strong> test set-up 186<br />

6.7.1 geometry <strong>and</strong> test set-up 186<br />

6.7.2 reliability <strong>of</strong> strains, measured by strain gauge laminated into a face 188<br />

6.8 Short-span panels: results 189<br />

6.8.1 loading up to 1/3 rd <strong>of</strong> maximum load: results 190<br />

6.8.2 loading up to 2/3 rd <strong>of</strong> maximum load: results 191<br />

6.9 Short-span panels: theoretical FEM predictions versus experimental results 192<br />

6.9.1 properties <strong>of</strong> the faces <strong>of</strong> the tested panels 192<br />

6.9.2 theory versus experiments: initial loading up to 1/3 rd <strong>of</strong> maximum 193<br />

6.9.3 theory versus experiments: initial loading up to 2/3 rd <strong>of</strong> maximum 193


Table <strong>of</strong> contents<br />

6.10 Conclusions 194<br />

6.9 References 195<br />

Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces 197<br />

7.1 Introduction 197<br />

7.2 <strong>Design</strong> methodology 198<br />

7.2.1 wind load 198<br />

7.2.2 snow load 199<br />

7.2.3 temperature load 199<br />

7.2.4 own weight 199<br />

7.2.5 combination <strong>of</strong> loads 199<br />

7.5.2.1 ultimate limit state (ULS) 200<br />

7.5.2.2 serviceability limit state (SLS) 202<br />

7.5.2.3 SLS: frequent load combination (FLC) 202<br />

7.5.2.4 SLS: characteristic load combination (FLC) 204<br />

7.2.6 design <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC faces 205<br />

7.3 Materials <strong>and</strong> modelling 206<br />

7.3.1 materials properties 206<br />

7.3.2 thermal expansion <strong>of</strong> UD-reinforced IPC 206<br />

7.4 Case studies 209<br />

7.4.1 case study 1: flat wall panel 209<br />

7.4.2 case study 2: flat ro<strong>of</strong> panel 209<br />

7.4.3 case study 3: flat ro<strong>of</strong> panel <strong>with</strong> intermediate support 209<br />

7.4.4 modelling <strong>and</strong> design 209<br />

7.5 <strong>Design</strong> under frequent <strong>and</strong> characteristic load combination (SLS) 210<br />

7.5.1 case study 1: flat wall panel 210<br />

7.5.2 case study 2: flat ro<strong>of</strong> panel 211<br />

7.5.3 case study 3: flat ro<strong>of</strong> panel <strong>with</strong> intermediate support 212<br />

7.5.4 conclusions 214<br />

7.6 <strong>Design</strong> in ultimate limit state 210<br />

7.6.1 case study 1: flat wall panel 210<br />

7.6.2 case study 2: flat ro<strong>of</strong> panel 211<br />

7.6.3 case study 3: flat ro<strong>of</strong> panel <strong>with</strong> intermediate support 212<br />

7.6.4 conclusions 216<br />

7.7 <strong>Design</strong> in under FLC after unloading from CLC 216<br />

7.7.1 introduction 216<br />

7.7.2 case study 1: flat wall panel 217<br />

7.7.3 case study 2: flat ro<strong>of</strong> panel 218<br />

7.7.4 case study 3: flat ro<strong>of</strong> panel <strong>with</strong> intermediate support 219<br />

7.7.5 conclusions 220<br />

7.8 <strong>Design</strong> under repeated loading<br />

7.8.1 introduction 220<br />

7.8.2 case study 1: flat wall panel 220


Table <strong>of</strong> contents<br />

7.8.3 case study 2: flat ro<strong>of</strong> panel 222<br />

7.8.4 case study 3: flat ro<strong>of</strong> panel <strong>with</strong> intermediate support 223<br />

7.8.5 conclusions 224<br />

7.9 Comparison <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite <strong>and</strong> steel faces 225<br />

7.9.1 introduction 225<br />

7.9.2 comparison <strong>of</strong> solutions 225<br />

7.6.3 conclusions 226<br />

7.10 Conclusions 226<br />

7.11 References 227<br />

Chapter 8: Conclusions 229<br />

8.1 Conclusions on the behaviour <strong>of</strong> IPC composite specimens 229<br />

8.2 Conclusions on s<strong>and</strong>wich modelling 231<br />

8.3 Conclusions on design criteria for s<strong>and</strong>wich panels <strong>with</strong> IPC faces for building 233<br />

8.4 Future research topics 234<br />

Appendix 1: Probability distribution functions<br />

Appendix 2: Probability plotting<br />

Appendix 3: Hysteresis<br />

Appendix 4: Macro for implementation <strong>of</strong> unloading<br />

Appendix 5: Macro for implementation <strong>of</strong> repeated loading


Chapter 1<br />

1.1 Research context<br />

Introduction<br />

S<strong>and</strong>wich panels were first used in aircraft industry. The combination <strong>of</strong> metal<br />

membrane skins <strong>and</strong> honeycomb cores leads to rather stiff lightweight structures.<br />

Only more recently s<strong>and</strong>wich panels have been introduced as wall cladding panels<br />

<strong>and</strong> as ro<strong>of</strong> panels. The materials, which are used most for this type <strong>of</strong> application,<br />

are two steel faces <strong>and</strong> a cellular core, which also functions as thermal insulation.<br />

Other material combinations are possible, but diversity in materials in s<strong>and</strong>wich<br />

panels for housing is still not common practice.<br />

Apart from s<strong>and</strong>wich elements, the use <strong>of</strong> composite materials in buildings has<br />

been the subject <strong>of</strong> numerous studies. Ferrocements are a good example <strong>of</strong> the<br />

potential <strong>of</strong> composite materials in construction <strong>of</strong> buildings. The advantages <strong>of</strong><br />

ferrocements are: relatively low cost, availability <strong>of</strong> the basic components, fireretarding<br />

properties <strong>and</strong> easy production. The use <strong>of</strong> ferrocements also ensures<br />

large flexibility in design. The concept <strong>of</strong> “housing” can be interpreted as the<br />

future house owner desires, as illustrated in figure 1.1.<br />

Figure 1.1: house structures using ferrocement shells (Naaman, 2000)<br />

1


Chapter 1: Introduction<br />

The use <strong>of</strong> ferrocement composites <strong>and</strong> <strong>of</strong> s<strong>and</strong>wich panels reduces construction<br />

time <strong>and</strong> energy. Lightweight construction elements are less likely to endanger the<br />

life <strong>of</strong> people in case <strong>of</strong> collapse, for example due to an earthquake. Moreover,<br />

lightweight constructions are more resistant to earthquakes than solid<br />

constructions. “A house, constructed mainly <strong>with</strong> lightweight panels may entrap<br />

people when it breaks down. However, the probability <strong>of</strong> suffering serious injuries<br />

or death decreases, due to the reduced weight <strong>of</strong> the collapsed elements.“<br />

(Goldsworthy, 2000)<br />

A new material, Inorganic Phosphate Cement (IPC), has been developed at the<br />

‘Vrije Universiteit Brussel’. After hardening, the IPC material properties are<br />

similar to those <strong>of</strong> cement-based materials. IPC is a two-component system,<br />

consisting <strong>of</strong> a calcium silicate powder <strong>and</strong> a phosphate acid based solution <strong>of</strong><br />

metaloxides. One <strong>of</strong> the major benefits <strong>of</strong> IPC, when compared to other existing<br />

cementitious materials, is the non-alkaline environment after hardening. Due to<br />

this, ordinary E-glass fibres are not attacked by the matrix <strong>and</strong> can be used as<br />

reinforcement. Therefore, E-glass fibre reinforced IPC is a cementitious<br />

composite, allowing production <strong>of</strong> very thin <strong>and</strong> cost-effective laminates. A<br />

combination <strong>of</strong> the application <strong>of</strong> this type <strong>of</strong> cementitious composite <strong>and</strong><br />

s<strong>and</strong>wich panels is discussed in this work.<br />

1.2 Objectives <strong>of</strong> the thesis<br />

Nowadays, numerous projects, which make use <strong>of</strong> brittle matrix composites, are<br />

under investigation. At present, the production <strong>of</strong> cost-effective cementitious<br />

composites <strong>with</strong> alkali resistant glass fibres is not achievable. Since IPC is<br />

compatible <strong>with</strong> cheaper E-glass fibres, industrial production <strong>of</strong> a new type <strong>of</strong><br />

construction elements becomes feasible. Indeed, one <strong>of</strong> the possible applications<br />

<strong>of</strong> these thin cementitious laminates is their implementation as faces in a s<strong>and</strong>wich<br />

panel. To the author’s knowledge, the use <strong>of</strong> very thin cementitious faces in<br />

s<strong>and</strong>wich panels for building applications has not been subject <strong>of</strong> pr<strong>of</strong>ound study<br />

up till now.<br />

In the first part <strong>of</strong> this thesis, focus is put on the behaviour <strong>of</strong> E-glass fibre<br />

reinforced IPC laminates. Loading, unloading <strong>and</strong> repeated loading is discussed.<br />

The behaviour <strong>of</strong> unidirectionally reinforced IPC under monotonic tensile loading<br />

<strong>and</strong> limited repeated loading has been the subject <strong>of</strong> study before (Bauweraerts,<br />

1998). The behaviour <strong>of</strong> 2D-r<strong>and</strong>omly reinforced IPC specimens under monotonic<br />

tensile loading has been studied by Gu et al. (1998). Refinements on the models,<br />

discussed by Bauweraerts (1998) <strong>and</strong> Gu et al. (1998), are introduced in this<br />

thesis. The behaviour <strong>of</strong> unidirectionally or 2D-r<strong>and</strong>omly reinforced IPC under<br />

monotonic loading, unloading <strong>and</strong> repeated loading is also tested <strong>and</strong> discussed in<br />

2


Chapter 1: Introduction<br />

this thesis. Publications on the behaviour <strong>of</strong> ceramic matrix composites (Curtin et<br />

al., 1998; Rouby <strong>and</strong> Reynaud, 1992; Evans et al., 1992) are used as references for<br />

the formulation <strong>of</strong> a theory, describing the behaviour <strong>of</strong> IPC composite specimens.<br />

The second part <strong>of</strong> this thesis provides an extensive report on the analysis <strong>and</strong><br />

design <strong>of</strong> lightweight panels <strong>with</strong> thin cementitious composite faces as ro<strong>of</strong> <strong>and</strong><br />

wall cladding panels. The most interesting features <strong>of</strong> wall <strong>and</strong> ro<strong>of</strong> panels are<br />

thermal <strong>and</strong> acoustical insulation, mechanical resistance, fire resistance <strong>and</strong><br />

resistance against environmental actions such as humidity changes, UV radiation<br />

<strong>and</strong> freezing-thawing cycles. Especially, the analysis <strong>of</strong> s<strong>and</strong>wich panels under<br />

mechanical loading is discussed in this work. The feasibility <strong>of</strong> load bearing<br />

s<strong>and</strong>wich ro<strong>of</strong> or wall panels <strong>with</strong> cementitious faces depends on the initial<br />

stiffness <strong>and</strong> strength properties <strong>and</strong> their evolution as a function <strong>of</strong> time,<br />

environment, etc. The design philosophy is based on the Preliminary<br />

Recommendations for S<strong>and</strong>wich <strong>Panels</strong> (ECCS, 1991) <strong>and</strong> the Updated<br />

Recommendations for S<strong>and</strong>wich <strong>Panels</strong> (Berner et al., 2000). However, these<br />

recommendations have been formulated <strong>with</strong> the specific behaviour <strong>of</strong> s<strong>and</strong>wich<br />

panels <strong>with</strong> steel faces in mind. In this work, the applicability <strong>of</strong> these<br />

recommendations is studied for s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces.<br />

Several s<strong>and</strong>wich design modifications, which are necessary, are introduced.<br />

Finally, the load bearing capacity <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC faces is compared<br />

to the capacity <strong>of</strong> more classical s<strong>and</strong>wich panels <strong>with</strong> steel faces.<br />

1.3 Basic assumptions in this thesis<br />

In this work, s<strong>and</strong>wich panels are assumed to work as wide beams rather than as<br />

plates. Ro<strong>of</strong> <strong>and</strong> wall s<strong>and</strong>wich panels are usually mounted to the main structure at<br />

two opposite ends. Since practically all applied loads are either pressure loads or<br />

line loads along the width <strong>of</strong> the s<strong>and</strong>wich panel, a wide beam approach is<br />

legitimised. Point loads are not considered in this work.<br />

The s<strong>and</strong>wich element should be able to sustain a combination <strong>of</strong> several loads.<br />

The loads considered in this work are:<br />

1. own weight, which is a permanent dead load<br />

2. snow load, which is a variable free action<br />

3. wind load, which is a variable free action<br />

4. temperature, which is a variable free action<br />

Before any analysis is made, the desired lifetime <strong>of</strong> the structure should be<br />

determined. Table 1.1 presents a general classification <strong>of</strong> structures <strong>and</strong> the<br />

lifetime they should survive <strong>with</strong>out any major repair (Eurocode 1).<br />

3


Chapter 1: Introduction<br />

The Preliminary Recommendations for S<strong>and</strong>wich <strong>Panels</strong> advise a lifetime <strong>of</strong> 50<br />

years. S<strong>and</strong>wich panels used as ro<strong>of</strong> or wall elements are thus structures <strong>of</strong> type 3,<br />

according to table 1.1. This work assumes a desired lifetime <strong>of</strong> 50 years.<br />

Table 1.1: classification <strong>of</strong> structures as function <strong>of</strong> their lifetime (Eurocode 1)<br />

classification lifetime (years) example<br />

1 1 - 5 temporary structures<br />

2 25 removable structural elements<br />

3 50 buildings<br />

4 100 civil engineering<br />

Both the ultimate limit state <strong>and</strong> the serviceability limit state are considered in the<br />

design <strong>of</strong> s<strong>and</strong>wich panels for building purposes. An important serviceability limit<br />

state is based on the limitation <strong>of</strong> the maximum deflection. In the Preliminary<br />

Recommendations for S<strong>and</strong>wich <strong>Panels</strong> <strong>and</strong> the Updated Recommendations for<br />

S<strong>and</strong>wich <strong>Panels</strong>, several propositions are made on this limitation. The proposed<br />

limits are 1/100 th , 1/200 th or 1/300 th <strong>of</strong> the span. In this work the maximum<br />

allowed deflection is 1/200 th <strong>of</strong> the span.<br />

1.4 Outline <strong>of</strong> the thesis<br />

Prior to any analysis <strong>of</strong> a construction element, the mechanical properties <strong>of</strong> all<br />

materials must be obtained.<br />

Polymer foams are suitable core materials <strong>and</strong> will be used in this study. They are<br />

less expensive than honeycomb cores. Polymer foam cores provide good thermal<br />

insulation <strong>and</strong> satisfactory mechanical stiffness <strong>and</strong> strength, all <strong>of</strong> which are a<br />

function <strong>of</strong> their density. The stress-strain behaviour <strong>of</strong> polymer foams has been<br />

discussed extensively in literature <strong>and</strong> is well known.<br />

The modelling <strong>of</strong> the behaviour <strong>of</strong> E-glass fibre reinforced IPC under monotonic<br />

loading is discussed in Chapter 2. Since the faces in a s<strong>and</strong>wich panel are mainly<br />

loaded in tension or in compression, the behaviour <strong>of</strong> E-glass fibre reinforced<br />

specimens under these types <strong>of</strong> loads is discussed. The assumption <strong>of</strong> linear elastic<br />

behaviour <strong>of</strong> E-glass fibre reinforced IPC in compression is checked<br />

experimentally. However, the stress-strain behaviour <strong>of</strong> E-glass fibre reinforced<br />

IPC laminates under tensile loading is non-linear. Constitutive modelling <strong>of</strong> Eglass<br />

fibre reinforced brittle matrix composites under monotonic tensile loading is<br />

also presented in Chapter 2. The presented theoretical constitutive models are<br />

formulated from meso-mechanical phenomena in the composite. The theoretical<br />

stress-strain behaviour is compared <strong>with</strong> experimental results on unidirectionally<br />

<strong>and</strong> 2D-r<strong>and</strong>omly reinforced specimens.<br />

4


Chapter 1: Introduction<br />

The meso-mechanical modelling philosophy applied for loading is used as basic<br />

theory to predict the unloading behaviour <strong>of</strong> E-glass fibre reinforced IPC<br />

laminates. The theoretical prediction <strong>of</strong> the unloading stress-strain behaviour <strong>of</strong><br />

IPC composite specimens is discussed in Chapter 3. Experimental data obtained<br />

on unidirectionally reinforced IPC specimens <strong>and</strong> on 2D-r<strong>and</strong>omly reinforced<br />

specimens are compared <strong>with</strong> these theoretical predictions.<br />

The phenomena leading to fatigue failure in composites are completely different<br />

from those <strong>of</strong> conventional materials. Whereas fatigue failure occurs from the<br />

initiation <strong>and</strong> propagation <strong>of</strong> one single crack for most conventional materials,<br />

several damage mechanisms can interact <strong>and</strong> lead to final failure under repeated<br />

loading <strong>of</strong> E-glass fibre reinforced IPC. Apart from fatigue failure, accumulation<br />

<strong>of</strong> deformations under repeated loading might be rather large for the studied type<br />

<strong>of</strong> composites. Underst<strong>and</strong>ing the meso-mechanics in E-glass fibre reinforced IPC<br />

laminates under cyclic tensile loading is thus essential. These phenomena are<br />

discussed in Chapter 4. Damage types to be considered are: matrix cracking, fibre<br />

failure, fibre pull-out <strong>and</strong> matrix-fibre interface degradation. The relative<br />

importance <strong>of</strong> a degradation mechanism is function <strong>of</strong> fibre volume fraction, type<br />

<strong>of</strong> fibre reinforcement, average stress <strong>and</strong> stress amplitude. The importance <strong>of</strong><br />

damage mechanisms in IPC composite specimens is discussed in Chapter 4 <strong>and</strong><br />

verified experimentally. A theoretical meso-mechanical based model is discussed.<br />

Various s<strong>and</strong>wich models are presented in Chapter 5. Their efficiency in analysis<br />

<strong>and</strong> design <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> cementitious composite faces is discussed.<br />

The use <strong>of</strong> a material <strong>with</strong> non-linear stress-strain behaviour makes it extra<br />

challenging to determine whether a simplified model gives appropriate predictions<br />

<strong>of</strong> deflections <strong>and</strong> internal stresses, once the tensile face is subjected to multiple<br />

cracking. Guidelines on an efficient practical use <strong>of</strong> the presented s<strong>and</strong>wich<br />

models are discussed <strong>and</strong> illustrated. The introduction <strong>of</strong> the material behaviour <strong>of</strong><br />

E-glass reinforced IPC in the finite element models is also discussed in Chapter 5.<br />

Small <strong>and</strong> larger scale s<strong>and</strong>wich panel testing is presented <strong>and</strong> discussed in<br />

Chapter 6. Since the controlled introduction <strong>of</strong> a pressure load on panels in the<br />

laboratory is difficult, the panels are subjected to a four point bending test. This<br />

type <strong>of</strong> loading is not experienced in situ, but it provides a useful benchmark. It is<br />

discussed whether the presented s<strong>and</strong>wich models indeed provide a useful <strong>and</strong><br />

accurate prediction tool for stresses, strains <strong>and</strong> displacements <strong>of</strong> s<strong>and</strong>wich panels<br />

<strong>with</strong> IPC composite faces. The panels are subjected to monotonic loading,<br />

unloading <strong>and</strong> repeated loading. Correspondence <strong>of</strong> experimental <strong>and</strong> theoretical<br />

load-deflection curves <strong>and</strong> load-strain curves <strong>of</strong> several s<strong>and</strong>wich panels is<br />

discussed.<br />

5


Chapter 1: Introduction<br />

Modifications might be necessary when the design rules for classical s<strong>and</strong>wich<br />

panels <strong>with</strong> steel faces are used for s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces.<br />

These modifications are presented <strong>and</strong> discussed in Chapter 7. Several case<br />

studies illustrate the design philosophy <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite<br />

faces. It is illustrated in Chapter 7 that the modelling <strong>of</strong> the face behaviour under<br />

various loading conditions is a conditio sine qua non for the analysis <strong>and</strong> design <strong>of</strong><br />

s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces for construction purposes. Finally, the<br />

load bearing capacity <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces is compared<br />

<strong>with</strong> the capacity <strong>of</strong> classical panels <strong>with</strong> steel faces.<br />

The main conclusions <strong>of</strong> this thesis are summarised in Chapter 8.<br />

1.5 References<br />

P. Bauweraerts, Aspect <strong>of</strong> the Micromechanical Characterisation <strong>of</strong> Fibre<br />

Reinforced <strong>Brittle</strong> <strong>Matrix</strong> Composites, Phd. thesis, VUB, 1998<br />

K. Berner, J.M. Davies, P. Hassinen, L. Heselius, Updated European<br />

recommendations for s<strong>and</strong>wich panels, Proceedings 5 th International Conference<br />

on S<strong>and</strong>wich Construction, EMAS Publishing, Sept 5-7, 2000, pp.389-400<br />

W.A. Curtin, B.K. Ahn, N. Takeda, Modeling <strong>Brittle</strong> <strong>and</strong> Tough Stressstrain<br />

Behaviour in Unidirectional Ceramic Composites, Acta mater., No. 10,<br />

1998, pp.3409-3420<br />

ECCS publication, Preliminary European Recommendations for S<strong>and</strong>wich<br />

<strong>Panels</strong>, part I, <strong>Design</strong>, Technical Committee 7, Working Group 7.4, 1991<br />

Eurocode 1, Grondslag voor ontwerp en belastingen op draagsystemen<br />

A.G. Evans, F.W. Zok, R.M. McMeeking, Fatigue <strong>of</strong> ceramic matrix<br />

composites, Acta metallurgica et materialia, 43, 1995, pp.859-875<br />

J. Gu, X. Wu, H. Cuypers <strong>and</strong> J. Wastiels, Modeling <strong>of</strong> the tensile<br />

behaviour <strong>of</strong> an E-glass fibre reinforced phosphate cement, Computer Methods in<br />

Composite Materials VI, proceedings CADCOMP 98, 1998, pp.589-598<br />

A.E. Naaman, Ferrocement & laminated cementitious composites, Techno<br />

Press 3000, 2000<br />

6


Chapter 1: Introduction<br />

D. Rouby <strong>and</strong> P. Reynaud, Fatigue behaviour related to interface<br />

modification during load cycling in ceramic-matrix fibre composites, Composite<br />

Science <strong>and</strong> Technology, 48, 1993, pp.109-118<br />

W. B. Goldsworthy, Composite Housing - why isn’t it here today?,<br />

Composites Technology, March/April 2000, pp.13<br />

7


Chapter 2<br />

2.1 Introduction<br />

IPC composite specimens<br />

under monotonic loading<br />

The material properties <strong>of</strong> pure IPC matrix <strong>and</strong> E-glass fibres are presented in this<br />

chapter. These properties are used in the formulation <strong>of</strong> the theory describing the<br />

behaviour <strong>of</strong> E-glass fibre reinforced specimens under monotonic compressive,<br />

tensile <strong>and</strong> shear loading. External tensile <strong>and</strong> compressive loads on the specimens<br />

are applied along one loading-axis: the x-axis.<br />

It is verified experimentally that E-glass fibre reinforced IPC specimens under<br />

compression show a linear elastic stress-strain relationship until failure.<br />

The stress-strain curves obtained from tensile testing are more complex due to<br />

non-linear irreversible micro- or meso-mechanical events. Unloaded specimens,<br />

which were subjected to tension, inevitably show residual strains <strong>and</strong> internal<br />

residual stresses. Two models, providing a prediction method <strong>of</strong> the stress-strain<br />

behaviour <strong>of</strong> IPC composite specimens in tension, are presented <strong>and</strong> discussed in<br />

this chapter. The goal to be achieved in this chapter is the formulation <strong>of</strong> a model<br />

based on physical micro- or meso-mechanical phenomena. However, the model<br />

should still be simple enough for analytical prediction <strong>of</strong> the macro-mechanical<br />

behaviour <strong>of</strong> a IPC composite. The theoretical stress-strain curve predictions are<br />

compared <strong>with</strong> experimental curves on unidirectionally reinforced <strong>and</strong> 2Dr<strong>and</strong>omly<br />

reinforced specimens.<br />

The compressive <strong>and</strong> tensile stress-strain formulations, obtained in this chapter,<br />

are implemented later as constitutive equations in finite element calculations <strong>of</strong><br />

s<strong>and</strong>wich panels <strong>with</strong> E-glass fibre reinforced IPC faces.<br />

9


Chapter 2: IPC composite specimens under monotonic loading<br />

Even though the theoretical stress-strain relationship <strong>of</strong> IPC laminates under shear<br />

is not discussed fundamentally, experimental results obtained from torsion testing<br />

are presented <strong>and</strong> discussed in this chapter.<br />

2.2 Materials<br />

Knowledge <strong>of</strong> the material properties <strong>of</strong> the different components is a minimum<br />

requirement for a proper description <strong>of</strong> the behaviour <strong>of</strong> E-glass fibre reinforced<br />

IPC composites. All IPC composite specimens in this chapter are fabricated as<br />

described here, unless mentioned otherwise explicitly.<br />

2.2.1 E-glass fibres as reinforcement<br />

Two types <strong>of</strong> E-glass fibre reinforcements are used: chopped glass fibre mats<br />

(“2D-r<strong>and</strong>om”) <strong>with</strong> a fibre density <strong>of</strong> 300g/m² (Owens Corning MK12) <strong>and</strong><br />

quasi-unidirectional woven roving (“UD”) <strong>with</strong> a fibre density <strong>of</strong> 158g/m².<br />

141g/m² <strong>of</strong> fibres are present as reinforcement in the main direction <strong>and</strong> 17g/m² <strong>of</strong><br />

the reinforcement is aligned along the transverse direction (Syncoglas Roviglas<br />

R17/141). The fibre diameter is 14µm for all fibres. The unidirectional fibres are<br />

continuous fibres. The length <strong>of</strong> the chopped fibres is 50mm. The E-glass fibre<br />

reinforcement consists <strong>of</strong> fibre bundles rather than <strong>of</strong> single fibres. The number <strong>of</strong><br />

fibres per bundle is ±1500 for the UD-reinforcement <strong>and</strong> ±540 for the 2D-r<strong>and</strong>om<br />

reinforcement. Properties <strong>of</strong> the E-glass fibres are listed in table 2.1.<br />

Table 2.1: properties E-glass fibres, found in literature <strong>and</strong> used in this work<br />

density<br />

(kg/m ³ tensile Young’s thermal<br />

strength modulus expansion<br />

) (MPa) (GPa) (10e -6 )<br />

Chawla (1993) 2550 1750 70 4.7<br />

Bentur <strong>and</strong> Mindess (1990) 2540 3500 72.5 -<br />

Matthews <strong>and</strong> Rawlings (1994) 2540 2200 70 -<br />

used in this work 2540 1700 72 4.7<br />

2.2.2 matrix material: Inorganic Phosphate Cement (IPC)<br />

Inorganic phosphate cement (IPC) has been developed at the ‘Vrije Universiteit<br />

Brussel’. IPC is a two-component system, consisting <strong>of</strong> a calcium silicate powder<br />

<strong>and</strong> a phosphate acid based solution <strong>of</strong> metaloxides. After hardening, the IPC<br />

material properties are similar to those <strong>of</strong> cement-based materials. Although the<br />

strength in compression is rather high, the tensile strength is low. The tensile<br />

strength <strong>of</strong> a IPC specimen is about one tenth <strong>of</strong> the compressive strength. Like<br />

ceramic or cementitious materials, IPC shows a high temperature resistance <strong>and</strong><br />

good fire retarding properties. Since all basic components are inorganic, no toxic<br />

gasses are released in case <strong>of</strong> fire hazard.<br />

10


Chapter 2: IPC composite specimens under monotonic loading<br />

One st<strong>and</strong>ard IPC mixture is chosen in this work, <strong>with</strong>out use <strong>of</strong> any fillers or<br />

retarding or accelerating components. The powder component is called N2 <strong>and</strong> the<br />

liquid mixture is noted by B23. More details cannot be given for reason <strong>of</strong><br />

confidentiality (PCT patent application: WO 97/19033). The weight ratio <strong>of</strong><br />

powder to liquid, used for all IPC mixtures in this work, is 1/1.25.<br />

IPC laminates or blocks are cured in ambient conditions for 24 hours. Post-curing<br />

is performed at 60°C for 24 hours. Like most cementitious mixtures, the strength<br />

<strong>of</strong> IPC increases <strong>with</strong> time. This effect can however be accelerated by the postcuring<br />

at 60°C. Typical values <strong>of</strong> stiffness, strength <strong>and</strong> own weight can be found<br />

in table 2.2. The presented stiffness <strong>and</strong> strength properties are average values.<br />

Variations on these values are mentioned if necessary. After cutting, the<br />

specimens are dried in ambient conditions, unless mentioned otherwise.<br />

Table 2.2: properties <strong>of</strong> the IPC matrix; after Gu et al. (1998), Bauweraerts (1998b) <strong>and</strong> Cuypers<br />

et al. (2000)<br />

tensile strength compressive strength E-modulus own weight<br />

(MPa)<br />

(MPa)<br />

(GPa) (kg/m³)<br />

IPC 6-14 80-120 18 2000<br />

2.2.3 E-glass fibre reinforced IPC laminates<br />

All laminates in this work are fabricated by h<strong>and</strong> lay-up technique. The amount <strong>of</strong><br />

matrix used per layer is 1600g/m², unless mentioned otherwise.<br />

After the laminate is fabricated, it is kept in ambient conditions for 24 hours.<br />

Afterwards, the laminate is post-cured for another 24 hours at 60°C, to accelerate<br />

the hardening process. During the curing <strong>and</strong> post-curing process, both sides <strong>of</strong> the<br />

laminate are covered <strong>with</strong> plastic to prevent early evaporation <strong>of</strong> water.<br />

The reinforcement is “UD” (unidirectional) or “2D-r<strong>and</strong>om”. The fibre volume<br />

fraction Vf is calculated from knowledge <strong>of</strong> the thickness <strong>of</strong> the composite<br />

specimen tc, the number <strong>of</strong> glass fibre layers nfl, the weight wfibrelayer <strong>of</strong> a fibre<br />

layer per m² <strong>and</strong> the density <strong>of</strong> glass fibres ρf.<br />

n flw<br />

fibrelayer<br />

V f = (2.1)<br />

ρ t<br />

The average value <strong>of</strong> the fibre volume fraction <strong>of</strong> a IPC composite specimen,<br />

made by h<strong>and</strong> lay-up technique, is about 15% for UD-reinforcement <strong>and</strong> 10% for<br />

2D-r<strong>and</strong>om reinforcement.<br />

f<br />

11<br />

c


Chapter 2: IPC composite specimens under monotonic loading<br />

2.3 IPC laminates under compressive loading<br />

2.3.1 introduction<br />

In general, when s<strong>and</strong>wich panels are subjected to bending, one face experiences<br />

compressive stresses <strong>and</strong> the other face experiences tensile stresses. Experimental<br />

results <strong>of</strong> compression testing are presented here to discuss whether the<br />

compressive properties <strong>of</strong> IPC laminates (E-modulus, compressive strength) are<br />

affected in a considerable way by the presence <strong>of</strong> E-glass fibre reinforcement.<br />

The presence <strong>of</strong> fibres might change the properties <strong>of</strong> IPC laminates in<br />

compression. A formulation <strong>of</strong> the influence <strong>of</strong> the presence <strong>of</strong> fibres on the<br />

stiffness is not presented here. Experimentally obtained results on 2D-r<strong>and</strong>omly<br />

reinforced IPC composite specimens are compared qualitatively <strong>with</strong> results on<br />

pure IPC matrix blocks. The hypothesis <strong>of</strong> linear elasticity in compression is<br />

checked experimentally on 2D-r<strong>and</strong>omly reinforced laminates, before it is used as<br />

material behaviour input in finite element calculations.<br />

2.3.2 test set-up<br />

Testing <strong>of</strong> laminates in compression is a delicate task, since most laminates are<br />

rather thin compared to their length. Small imperfections in the specimen or<br />

misalignment in the test set-up can cause instability at rather low compressive<br />

loads. One can overcome this problem by decreasing the length <strong>of</strong> the tested<br />

specimen. However, in that case the influence <strong>of</strong> the boundary conditions cannot<br />

be neglected in the middle <strong>of</strong> the specimen, where the strains are usually<br />

measured. The assumption <strong>of</strong> uniform compressive stresses is thus not acceptable.<br />

Another method, which helps to avoid buckling, is realised by supporting the<br />

specimen along both sides. Several methods to realise this support can be found in<br />

ASTM st<strong>and</strong>ards (1995) <strong>and</strong> in a document written by Mitropoulos (1996).<br />

The support should prevent buckling, <strong>with</strong>out introducing an extra stiffening<br />

effect. In the work <strong>of</strong> Mitropoulos (1996) several supporting methods were used to<br />

check whether they are practically achievable. Advantages <strong>and</strong> disadvantages <strong>of</strong><br />

each method are presented in the document <strong>of</strong> Mitropoulos (1996). One <strong>of</strong> the test<br />

set-ups, based on conclusions from Mitropoulos (1996), is illustrated in figure 2.1.<br />

This set-up is used here to obtain information on the behaviour <strong>of</strong> E-glass fibre<br />

reinforced IPC specimens in compression.<br />

Small specimens are cut from 2D-r<strong>and</strong>omly reinforced IPC plates. In ASTM<br />

st<strong>and</strong>ard D3410-76 (1995), the recommended minimum thickness <strong>of</strong> specimens is<br />

4mm, if the expected E-modulus is lower than 70GPa. This specimen thickness is<br />

used in this work. The length <strong>of</strong> the specimens is chosen to equal the length <strong>of</strong> the<br />

steel supporting jig (see figure 2.1) plus 2mm.<br />

12


Chapter 2: IPC composite specimens under monotonic loading<br />

steel<br />

supporting<br />

jig<br />

specimen<br />

compressive force on specimen<br />

teflon<br />

Figure 2.1: test set-up for E-glass fibre reinforced IPC specimens in compression<br />

Each specimen is prevented from buckling under compression by two steel<br />

supporting jigs. These jigs are bolted together. The tightening moment, which is<br />

applied to the bolts, is only large enough to prevent the supporting jigs from<br />

sliding from the specimen due to their own weight. Teflon sheets are placed<br />

between the specimen <strong>and</strong> the steel supporting jigs to minimise friction between<br />

the supporting plates <strong>and</strong> the specimen. This way, the steel supporting jigs have<br />

minimal stiffening effect on the specimens.<br />

The specimen is placed between the plates <strong>of</strong> a INSTRON 4505 testing bench. The<br />

compressive force is applied directly on the specimen by the horizontal steel plates<br />

<strong>of</strong> the testing bench.<br />

The upper steel plate <strong>of</strong> the INSTRON testing bench is fixed <strong>and</strong> the lower plate is<br />

displaced vertically at a rate <strong>of</strong> 0.2 mm/min. A 10kN load cell measures the forces,<br />

while the strains are measured by strain gauges. Strain gauges are attached at both<br />

sides <strong>of</strong> the specimens to ascertain that bending <strong>of</strong> the specimens is monitored, in<br />

case it occurs. Buckling can be detected too: the strain <strong>of</strong> one face would then<br />

decrease rapidly while the strain <strong>of</strong> the opposite face suddenly increases.<br />

2.3.3 materials <strong>and</strong> results<br />

Three bulk IPC blocks <strong>of</strong> 160x40x40mm³ are prepared. These blocks are tested in<br />

three-point bending <strong>and</strong> in compression, according to the Belgian National<br />

St<strong>and</strong>ards (NBN-B12-208 <strong>and</strong> NBN-B14-209). The stiffness <strong>and</strong> strength values,<br />

obtained from these specimens, are used as reference values for the matrix<br />

properties <strong>of</strong> the composite <strong>and</strong> are listed in table 2.3.<br />

13


Chapter 2: IPC composite specimens under monotonic loading<br />

Table 2.3: properties <strong>of</strong> solid IPC, obtained from three-point bending <strong>and</strong> compression tests<br />

specimen three-point bending compression<br />

name<br />

E-modulus<br />

(GPa)<br />

tensile strength<br />

(MPa)<br />

compressive<br />

strength<br />

(MPa)<br />

IPCI/CS/1 20.7 7.93 95.9<br />

IPCI/CS/2 20.4 9.56 98.9<br />

IPC/CS/3 19.9 8.65 93.3<br />

average 20.3 8.74 96.1<br />

A 2-D r<strong>and</strong>omly reinforced four-layer IPC laminate is prepared by h<strong>and</strong> lay-up.<br />

Specimens <strong>with</strong> dimensions <strong>of</strong> 12x79mm² are cut from this laminate.<br />

Figure 2.2 shows a typical stress-strain curve, obtained <strong>with</strong> the test set-up <strong>of</strong><br />

figure 2.1. One can notice there is only limited bending due to non-perfect<br />

positioning <strong>of</strong> the specimen or non-symmetrical composite lay-up. Buckling is<br />

prevented, failure <strong>of</strong> the specimen occurs under compression. Fibre volume<br />

fractions, stiffness properties <strong>and</strong> failure stresses are listed in table 2.4.<br />

stress(MPa)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

E1=18.5GPa<br />

E2=17.4GPa<br />

side 1<br />

side 2<br />

0 0.1 0.2 0.3 0.4 0.5 0.6<br />

strain(%)<br />

Figure 2.2: stress-strain curve in compression <strong>of</strong> IPC composite specimen<br />

Table 2.4: stiffness <strong>and</strong> strength properties, obtained from compressive testing on 2D-r<strong>and</strong>omly<br />

reinforced IPC specimens<br />

specimen<br />

name<br />

Vf<br />

(%)<br />

E side 1<br />

(GPa)<br />

E side 2<br />

(GPa)<br />

E average<br />

(GPa)<br />

compressive<br />

strength<br />

(MPa)<br />

IPC/C/4 9.56 21.1 16.4 18.8 88.7<br />

IPC/C/5 9.51 17.4 16.5 17.0 88.9<br />

IPC/C/6 9.56 18.5 17.4 18.0 88.3<br />

average 9.55 19.0 16.8 17.9 88.6<br />

14


Chapter 2: IPC composite specimens under monotonic loading<br />

2.3.4 discussion <strong>and</strong> conclusions<br />

The average composite stiffness <strong>of</strong> 17.9GPa <strong>and</strong> compressive strength <strong>of</strong> 88.6MPa<br />

are slightly lower than the pure solid IPC stiffness <strong>of</strong> 20.3GPa <strong>and</strong> strength <strong>of</strong><br />

96.1MPa.<br />

The law <strong>of</strong> mixtures predicts the E-modulus <strong>of</strong> a composite specimen Ec1,<br />

provided the matrix <strong>and</strong> fibre E-modulus (Em <strong>and</strong> Ef) <strong>and</strong> the matrix <strong>and</strong> fibre<br />

volume fraction (Vm <strong>and</strong> Vf) are known <strong>and</strong> the composite behaves linear<br />

elastically. Equation (2.2) represents the law <strong>of</strong> mixtures:<br />

∗<br />

E c1<br />

= E fV<br />

f + EmVm<br />

(2.2)<br />

In equation (2.2) the equivalent volume fraction Vf * includes effects <strong>of</strong> the<br />

alignment <strong>of</strong> the reinforcement <strong>and</strong> fibre length. If continuous unidirectional fibres<br />

are used as reinforcement, Vf * will simply equal Vf. If not all fibres are oriented<br />

along the loading axis or if they are not continuous, the relationship between Vf<br />

<strong>and</strong> Vf * is expressed as:<br />

(2.3)<br />

Vf * = ηθη1Vf<br />

ηθ is the fibre orientation efficiency factor <strong>and</strong> ηl expresses the influence <strong>of</strong> the<br />

length <strong>of</strong> the fibres.<br />

Krenchel (1964) <strong>and</strong> Cox (1952) performed a calculation <strong>of</strong> ηθ. Krenchel (1964)<br />

uses a constrained assumption, which means deformations are only allowed in the<br />

direction <strong>of</strong> the applied load. Cox (1952) considers an unconstrained assumption,<br />

which allows deformations in directions different from the loading axis. The<br />

values <strong>of</strong> ηθ are listed in table 2.5.<br />

Table 2.5: orientation efficiency factors derived for constrained <strong>and</strong> unconstrained composites<br />

orientation efficiency factor<br />

aligned UD 2D-r<strong>and</strong>om 3D-r<strong>and</strong>om<br />

Krenchel (1964) 1 3/8 1/5<br />

Cox (1952) 1 1/3 1/6<br />

In this work the unconstrained assumption, used by Cox (1952), is applied. If 2Dr<strong>and</strong>omly<br />

reinforced specimens are considered, Vf * is thus defined by:<br />

(2.4)<br />

Vf * =1/3Vf<br />

The fibre length effect ηl is expressed amongst others by Cox (1952), Laws (1971)<br />

<strong>and</strong> Bentur <strong>and</strong> Mindess (1990). Formulation <strong>of</strong> ηl, as presented by these authors,<br />

is expressed in equations (2.5) to (2.8).<br />

β1l<br />

f<br />

1−<br />

tanh( )<br />

n<br />

2<br />

l = (2.5)<br />

β l<br />

1 f<br />

2<br />

15


<strong>with</strong>:<br />

Chapter 2: IPC composite specimens under monotonic loading<br />

⎛ 2 ⎞<br />

⎜<br />

Gm<br />

β ⎟<br />

1 =<br />

⎜ 2<br />

ln(<br />

/ ) ⎟<br />

⎝ E f r R r ⎠<br />

(2.6)<br />

where: lf = length <strong>of</strong> the fibre<br />

Gm = shear modulus <strong>of</strong> the matrix<br />

r = radius <strong>of</strong> the fibre<br />

Ef = Young’s modulus <strong>of</strong> the fibre<br />

R = radius <strong>of</strong> the matrix around the fibre<br />

The value <strong>of</strong> ratio R/r depends on the fibre packing <strong>and</strong> fibre volume fraction.<br />

Piggott (1980) derived expressions for R/r for fibres <strong>with</strong> circular cross-section:<br />

square fibre packing:<br />

hexagonal fibre packing:<br />

ln<br />

1 / 2<br />

1 ⎛ ⎞<br />

( ) ⎜<br />

π<br />

ln R / r = ln ⎟<br />

⎜ ⎟<br />

(2.7)<br />

2 ⎝V<br />

f ⎠<br />

( R / r)<br />

1 ⎛<br />

= ln⎜<br />

2 ⎜<br />

⎝<br />

2π<br />

⎞<br />

⎟<br />

3V<br />

⎟<br />

f ⎠<br />

(2.8)<br />

For the studied 2D-r<strong>and</strong>omly reinforced IPC composites, Vf is 9.55% (table 2.4),<br />

Ef is 72GPa, r is 7µm, lf is 50mm <strong>and</strong> Gm is about 6GPa. When square fibre<br />

packing is adopted, the value <strong>of</strong> ηl is 0.9992. The value <strong>of</strong> ηl becomes 0.9993 if<br />

hexagonal fibre packing is considered. In this work, ηl is considered to equal 1<br />

from now on.<br />

The average fibre volume fraction <strong>of</strong> the tested composite specimens is 9.55% (see<br />

table 2.4). The average matrix stiffness is retrieved from table 2.3 <strong>and</strong> is 20.3GPa.<br />

According to equation (2.2) <strong>and</strong> (2.4), Ec1 is 20.9GPa. Although the theoretical<br />

prediction <strong>of</strong> the composite stiffness is higher than the stiffness <strong>of</strong> the pure matrix,<br />

the experimental obtained values <strong>of</strong> the composite stiffness are lower. Similarly,<br />

one would expect the failure stress <strong>of</strong> the IPC composite is higher than the failure<br />

stress <strong>of</strong> the pure IPC matrix. The opposite phenomenon is observed<br />

experimentally. It seems the presence <strong>of</strong> fibres introduces some extra effects,<br />

which lower the stiffness <strong>and</strong> the strength <strong>of</strong> the composites in compression. Two<br />

possible explanations are:<br />

-Together <strong>with</strong> the fibre reinforcement, extra air voids might be entrapped<br />

into the fibre bundle. From SEM (scanning electron microscope) pictures it is seen<br />

that the matrix does not adequately impregnate the fibre bundles. This effect is<br />

16


Chapter 2: IPC composite specimens under monotonic loading<br />

illustrated in figure 2.17 <strong>and</strong> discussed later in paragraph 2.4.8.3. The increased<br />

porosity <strong>of</strong> the composite leads to a decreased stiffness <strong>and</strong> failure stress.<br />

-During the curing process, the matrix will<br />

shrink. If free shrinkage is prevented by the<br />

presence <strong>of</strong> fibres, internal tensile matrix stresses<br />

are present. These internal stresses may cause<br />

micro-cracking in the matrix, even when no<br />

external loading is applied. Figure 2.3 (Naaman<br />

<strong>and</strong> Reinhardt, 1995) illustrates how the<br />

introduction <strong>of</strong> micro-cracks affects the<br />

compressive properties <strong>of</strong> a material. Under<br />

compressive loading (σ in figure 2.3), the edges <strong>of</strong><br />

the micro-cracks slide (illustrated by τ in figure<br />

2.3), introducing a local tensile stress field at their<br />

tips. Wing cracks are introduced as shown in<br />

figure 2.3. These wing cracks are initially<br />

unstable, but may become stable as the crack<br />

length increases. However, the presence <strong>of</strong> other<br />

micro-cracks <strong>and</strong> their interaction can lead to final<br />

failure.<br />

τ<br />

wing crack<br />

σ<br />

σ<br />

micro-crack<br />

Figure 2.3: crack propagation,<br />

Naaman <strong>and</strong> Reinhardt (1995)<br />

Conclusively, the law <strong>of</strong> mixtures predicts a higher stiffness <strong>and</strong> failure strength <strong>of</strong><br />

IPC composites, compared to pure IPC matrix material. Experimental results<br />

indicate the opposite occurs. Possible reasons are found in the fact that, due to<br />

introduction <strong>of</strong> E-glass fibres, entrapment <strong>of</strong> extra air voids in the fibre bundle <strong>and</strong><br />

introduction <strong>of</strong> extra matrix micro-cracks lower the strength <strong>and</strong> stiffness<br />

properties <strong>of</strong> the composite in compression.<br />

The value <strong>of</strong> the IPC composite face stiffness <strong>and</strong> strength, which will be used in<br />

this work from now on, are 18GPa <strong>and</strong> 90MPa respectively.<br />

2.4 Tensile behaviour <strong>of</strong> continuous UD-reinforced IPC<br />

2.4.1 introduction<br />

Unlike the stress-strain behaviour in compression, the stress-strain curve <strong>of</strong> Eglass<br />

fibre reinforced IPC in tension is highly non-linear. Several stages <strong>of</strong><br />

complexity are used here in the development <strong>of</strong> models. The presented models can<br />

be used as analytical constitutive equations in finite element calculations. In this<br />

chapter, two models are presented. The basic assumptions <strong>of</strong> these models are<br />

listed <strong>and</strong> derivation <strong>of</strong> the stress-strain equations is presented. Both models<br />

provide a stress-strain relationship, which will be verified on UD-reinforced IPC<br />

17


Chapter 2: IPC composite specimens under monotonic loading<br />

composite specimens. Advantages <strong>and</strong> disadvantages <strong>of</strong> both models are<br />

discussed.<br />

2.4.2 basic assumptions<br />

The basic assumptions, which are used in both models, are listed below:<br />

- The specimens are loaded along their longitudinal axis only, no combined 2D<br />

or 3D loads are considered.<br />

- The external loading axis is parallel <strong>with</strong> the alignment <strong>of</strong> the UDreinforcement.<br />

- The fibres are only capable <strong>of</strong> carrying load along their longitudinal axis.<br />

They provide no bending stiffness.<br />

- The matrix-fibre bond is weak. The adhesion shear bond strength τau is low.<br />

Propagation <strong>of</strong> matrix cracks leads to matrix-fibre debonding along a certain<br />

length <strong>of</strong> the fibre.<br />

- Once the matrix <strong>and</strong> the fibre are debonded, a pure frictional shear bond τ0<br />

replaces the previously existing adhesion shear bond τa totally.<br />

- Once matrix-fibre interface debonding occurred, the frictional interface<br />

sliding resistance τ0 is constant along the debonded interface.<br />

- At the debonded matrix-fibre interface, the sliding shear stress τ0 is constant<br />

<strong>with</strong> slip (figure 2.4a). In reality there is an initial static frictional sliding stress τi0<br />

at the onset <strong>of</strong> sliding, higher than the dynamic frictional slip stress τd0, which<br />

occurs once slip is activated under the application <strong>of</strong> external tensile loading<br />

(figure 2.4b). The assumption illustrated in figure 2.4a is used in this work.<br />

interfacial shear stress<br />

0<br />

τa<br />

τau<br />

τ0 = τd0<br />

0 displacement<br />

Figure 2.4a: hypothetical matrix-fibre<br />

interface shear stress τ versus slip (Bentur<br />

<strong>and</strong> Mindess, 1999)<br />

interfacial shear stress<br />

0<br />

τa<br />

τau<br />

τ0 = τi0<br />

τ0 = τd0<br />

0 displacement<br />

Figure 2.4b: hypothetical matrix-fibre<br />

interface shear stress τ versus slip<br />

(Naaman et al., 1991)<br />

- When the behaviour <strong>of</strong> the matrix-fibre interface is considered, Poisson<br />

effects in the fibre <strong>and</strong> matrix are neglected.<br />

- Global load sharing is used for the fibres. Local load sharing amongst fibres<br />

is neglected.<br />

18


Chapter 2: IPC composite specimens under monotonic loading<br />

- The matrix normal stresses are assumed to be constant in a section, transverse<br />

to the loading direction.<br />

2.4.3 formulation <strong>of</strong> the ACK theory<br />

The first model to be discussed in this chapter is based on the ACK theory<br />

(Aveston-Cooper-Kelly theory). The ACK theory has been derived for UDreinforced<br />

brittle matrix composites by Aveston et al. (1971). Verification <strong>of</strong> the<br />

ACK model on UD-reinforced IPC laminates has been studied by Bauweraerts et<br />

al. (1998a) <strong>and</strong> Bauweraerts (1998b). The theoretical background <strong>of</strong> the ACK<br />

model is presented here.<br />

According to the ACK theory, three distinct zones can be detected in the stressstrain<br />

curve <strong>of</strong> a unidirectionally reinforced composite. Figure 2.5 illustrates a<br />

theoretical stress-strain curve <strong>with</strong> three distinct zones (used in the ACK theory).<br />

stress<br />

0<br />

0<br />

zone I<br />

εmc = εmu<br />

∆εc zoneII<br />

σmc<br />

zone II<br />

εc zoneII<br />

zone III<br />

strain<br />

σmc = composite multiple cracking stress<br />

εmc = composite multiple cracking strain<br />

εmu = matrix failure strain<br />

εc zoneII = composite strain after multiple cracking<br />

∆εc zoneII = composite strain term, due to multiple cracking<br />

Figure 2.5: theoretical stress-strain curve, according to the ACK theory<br />

(Aveston et al., 1971)<br />

2.4.3.1 zone I<br />

This is the linear elastic zone. The stiffness <strong>of</strong> the composite Ec1 in this zone is<br />

function <strong>of</strong> the fibre volume fraction Vf, the volume fraction <strong>of</strong> the matrix Vm, the<br />

19


Chapter 2: IPC composite specimens under monotonic loading<br />

stiffness <strong>of</strong> the fibres Ef <strong>and</strong> the stiffness <strong>of</strong> the matrix Em. In zone I, the matrixfibre<br />

interface bond is assumed to be elastic.<br />

E = E V + E V<br />

c1<br />

f f m m<br />

(2.9)<br />

Imperfect matrix-fibre adhesion, warping or misalignment <strong>of</strong> unidirectional fibres,<br />

inclusion <strong>of</strong> air voids, etc. can lower the value <strong>of</strong> the composite stiffness.<br />

2.4.3.2 zone II<br />

This zone is <strong>of</strong>ten called “multiple cracking” zone. According to the ACK theory,<br />

the matrix has one determined tensile failure stress σmu (<strong>and</strong> failure strain εmu).<br />

Once the matrix failure stress σmu is reached, the matrix shows multiple cracking.<br />

This means matrix cracks are initiated <strong>and</strong> propagated along the whole specimen.<br />

The composite multiple cracking stress σmc, at which all matrix cracks are initiated<br />

<strong>and</strong> propagated, is:<br />

Ec1σ mu<br />

σ mc = (2.10)<br />

Em<br />

When a first matrix crack appears <strong>and</strong> reaches a fibre, debonding <strong>of</strong> the matrixfibre<br />

interface occurs, since it is assumed that the matrix-fibre interface is weak.<br />

Along the debonded interface, the assumption <strong>of</strong> the existence <strong>of</strong> a constant<br />

frictional interface shear stress τ0 is used. Figure 2.6a illustrates the interface<br />

debonding. The distance, along which this matrix-fibre interface debonding<br />

occurs, is called the slip length or debonding length δ0.<br />

adhesive interface<br />

matrix crack<br />

constant frictional stress (τ0)<br />

matrix<br />

fibre<br />

frictional interface<br />

debonded interface (slip length δ0)<br />

TENSION<br />

Figure 2.6a: matrix-fibre interface debonding, due to the introduction <strong>of</strong> a matrix crack<br />

crack<br />

σ<br />

0<br />

debonded interface (δ0)<br />

Figure 2.6b: normal stress evolution at matrix-fibre interface in the vicinity <strong>of</strong> a matrix crack<br />

20<br />

σf<br />

σm<br />

x


Chapter 2: IPC composite specimens under monotonic loading<br />

In the vicinity <strong>of</strong> a matrix crack, the transfer <strong>of</strong><br />

normal stresses from fibres (σf) to the matrix<br />

(σm) is accomplished by the constant frictional<br />

interface shear stress (τ0). <strong>Matrix</strong> stress σm<br />

increases linearly <strong>and</strong> fibre stress σf decreases<br />

linearly <strong>with</strong> increasing distance x from the<br />

crack. This normal stress evolution is<br />

illustrated in figure 2.6b.<br />

From figure 2.6b, it can be seen that the normal<br />

matrix stress at the crack tip equals zero. In the<br />

vicinity <strong>of</strong> a crack, the matrix stress is<br />

considerably lower than far away from this<br />

crack. At both sides <strong>of</strong> this first crack, a new<br />

crack will thus not be likely to appear <strong>with</strong>in<br />

the debonded zone <strong>of</strong> length δ0.<br />

dx<br />

σf<br />

τ0<br />

σf+dσf<br />

Figure 2.7: stresses acting on fibre in<br />

debonded area<br />

The debonding length δ0 can be calculated from the equilibrium <strong>of</strong> forces along<br />

the loading axis as shown in figure 2.7 <strong>and</strong> is presented in equation 2.11. The xaxis<br />

is oriented parallel <strong>with</strong> the loading axis, r is the fibre radius.<br />

2<br />

2<br />

( σ f + dσ f ) r Π = σ f r π + τ 02πrdx<br />

(2.11)<br />

By expressing the fact that the total force on the composite is the sum <strong>of</strong> the force<br />

taken by the fibres <strong>and</strong> the force taken by the matrix in all transverse composite<br />

sections, equation (2.12) is found:<br />

σ c = σ fV<br />

f + σ mVm<br />

(2.12)<br />

From x = 0 to x = δ0, the matrix normal stresses vary from σm = 0 to σm = σmu.<br />

After integration <strong>and</strong> rearranging, combination <strong>of</strong> equation (2.11) <strong>and</strong> (2.12) gives:<br />

σ mur<br />

Vm<br />

δ 0 =<br />

2τ V<br />

(2.13)<br />

0<br />

f<br />

Since new cracks are introduced at a distance δ0 from an existing crack or further,<br />

the values <strong>of</strong> the crack spacing are situated between δ0 <strong>and</strong> 2δ0. The value <strong>of</strong> the<br />

average final crack spacing f has been determined <strong>and</strong> published first by<br />

Widom (1966) <strong>and</strong> is:<br />

cs = 1. 337δ<br />

f<br />

0<br />

(2.14)<br />

The determination <strong>of</strong> the value <strong>of</strong> the average final crack spacing, as derived by<br />

Widom (1966), is identical to the “car parking” problem, where cars <strong>with</strong> a length<br />

δ0 are parked r<strong>and</strong>omly along a road until there is no parking space left.<br />

In order to explain how Widom derived this particular value for the crack spacing,<br />

Curtin (1999) published figures similar to figure 2.8a to 2.8d here. These figures<br />

21


Chapter 2: IPC composite specimens under monotonic loading<br />

show how cracks are introduced sequentially in time (when σc = σmc) <strong>and</strong> how the<br />

average final crack spacing can be determined from this “car parking” problem.<br />

matrix<br />

stress<br />

0<br />

0 virgin composite<br />

matrix<br />

stress<br />

3<br />

0<br />

3<br />

0<br />

matrix<br />

stress<br />

3<br />

x<br />

0 15<br />

Figure 2.8a: σ in matrix at t = t1<br />

0<br />

0<br />

0<br />

first crack<br />

0 x<br />

0 15<br />

Figure 2.8b: σ in matrix at t = t1+δt<br />

0 x<br />

0 15<br />

Figure 2.8c: σ in matrix at t = t1+2δt<br />

matrix<br />

stress<br />

matrix<br />

stress<br />

matrix<br />

stress<br />

3<br />

0<br />

3<br />

0<br />

3<br />

0<br />

0<br />

0<br />

0<br />

0 x<br />

Figure 2.8d: σ in matrix at t = t1+3δt<br />

0 15<br />

0 x<br />

0<br />

Figure 2.8e: σ in matrix at t = t1+7δt<br />

0 x<br />

0 15<br />

Figure 2.8f: full multiple cracking<br />

-At time t = t1, the stresses in the matrix approach the ultimate matrix stress,<br />

σmu. The composite specimen is in virgin state, showing no cracks. The normal<br />

stresses in the matrix are constant along the whole length <strong>of</strong> the composite, as can<br />

be seen in figure 2.8a.<br />

-A first crack appears at t = t1 + δt at a r<strong>and</strong>omly chosen place in the<br />

composite (figure 2.8b). The matrix-fibre interface becomes debonded at both<br />

sides <strong>of</strong> this first crack <strong>and</strong> no new cracks can be initiated along these debonded<br />

lengths. Therefore, the region <strong>with</strong>in a distance δ0 to the left <strong>and</strong> a distance δ0 to<br />

the right <strong>of</strong> a matrix crack is sometimes called the “exclusion zone”.<br />

-At time t = t1 + 2δt, a second crack appears in the composite, <strong>with</strong> new<br />

debonded zones, excluding further matrix cracking (figure 2.8c).<br />

-At t = t1 + 3δt, a new crack occurs very close to the first crack (figure<br />

2.8d). The debonded zones (or exclusion zones) <strong>of</strong> crack 1 <strong>and</strong> crack 3 overlap. In<br />

this region the tensile stresses in the matrix are considerably lower than the matrix<br />

tensile strength. Extra cracking between cracks 1 <strong>and</strong> 3 is thus impossible from<br />

now on. However, in the zone between crack 2 <strong>and</strong> 3 new cracks can still appear.<br />

Also to the left <strong>of</strong> crack 1 <strong>and</strong> to the right <strong>of</strong> crack 2, new cracks can be initiated<br />

<strong>and</strong> propagated.<br />

22<br />

15


Chapter 2: IPC composite specimens under monotonic loading<br />

-Figure 2.8e shows an intermediate state t = t1 + 7δt <strong>with</strong> several exclusion<br />

areas, but there are still some zones where further matrix cracking is possible.<br />

-In figure 2.8f, the end <strong>of</strong> multiple cracking is reached. The matrix peak<br />

stresses vary between σmu <strong>and</strong> σmu/2. If the matrix stress would be σmu, a new<br />

crack would appear in that region. If the matrix peak stress would be lower than<br />

σmu/2, a crack would have appeared in the exclusion zone <strong>of</strong> an existing crack,<br />

which was stated to be impossible.<br />

When matrix cracks are introduced, elastic relaxation <strong>of</strong> matrix stresses occurs in<br />

the vicinity <strong>of</strong> these cracks. Average stresses in the matrix are lower than they<br />

were just before the matrix strength was reached. The matrix thus carries less<br />

external load after cracking than just before cracking. The fibres provide the load<br />

carrying capacity that is lost by the matrix. This is only possible if there are<br />

enough embedded fibres to carry this extra load. A minimum critical fibre volume<br />

fraction Vf crit is a requirement, if total failure <strong>of</strong> the composite is to be prevented at<br />

the multiple cracking stage.<br />

The critical fibre volume fraction Vf crit can be found by expressing the fact that the<br />

load taken by fibres <strong>and</strong> matrix, just before multiple cracking occurred equals the<br />

load taken by the fibres, only after multiple cracking occurred. If the critical fibre<br />

volume fraction is embedded, the fibre stress equals the fibre failure stress σfu after<br />

full matrix cracking occurred. The composite multiple cracking stress σmc can<br />

therefore be expressed as:<br />

σ mc<br />

crit<br />

crit<br />

= σ fV<br />

f + σ muVm<br />

= σ fuV<br />

f<br />

(2.15)<br />

Just before multiple cracking occurs, the fibre <strong>and</strong> matrix strains are still equal:<br />

E fσ<br />

mu<br />

σ f =<br />

Em<br />

Combination <strong>of</strong> equation (2.15) <strong>and</strong> (2.16) gives:<br />

(2.16)<br />

crit σ mu<br />

V f =<br />

E f<br />

σ fu − σ mu ( 1−<br />

)<br />

E<br />

(2.17)<br />

For E-glass fibre reinforced IPC laminates, the critical fibre volume fraction can<br />

be found by inserting the material properties listed in tables 2.1 <strong>and</strong> 2.2 in equation<br />

(2.17). The critical fibre volume fraction is about 0.5%.<br />

The fibres bear larger tensile stresses <strong>and</strong> consequently undergo larger strains after<br />

multiple cracking occurred. Since the composite strain equals the fibre strain, an<br />

extra composite strain term ∆εc zoneII is therefore introduced, due to multiple<br />

cracking. The composite strain εc zoneII after multiple cracking (see figure 2.5) can<br />

be expressed as the sum <strong>of</strong> the composite strain before multiple cracking εmc (<strong>with</strong><br />

εmc = εmu) <strong>and</strong> the extra strain term ∆εc zoneII . The normal stress evolution between<br />

23<br />

m


Chapter 2: IPC composite specimens under monotonic loading<br />

two cracks, as illustrated in figure 2.9, is used to derive the composite strain<br />

εc zoneII . In figure 2.9, a black linear curve represents the normal stress in the matrix.<br />

A grey linear curve represents the normal stress in the fibres. σm,min <strong>and</strong> σm,max are<br />

the minimum <strong>and</strong> maximum value <strong>of</strong> the matrix stress. σf,min <strong>and</strong> σf,max are the<br />

minimum <strong>and</strong> maximum value <strong>of</strong> the fibre stress. x is the distance from the crack<br />

face.<br />

The average final crack spacing f has been derived by Widom (1966) <strong>and</strong><br />

equals 1.337δ0 (equation 2.14).<br />

Figure 2.9 shows that slip length δ0 can be redefined as the distance from a matrix<br />

crack, along which the matrix stress increases linearly to the far field value σm ff .<br />

This matrix far field value σm ff is the stress in the matrix at a theoretically infinite<br />

distance from one single crack, provided the existence <strong>of</strong> other cracks is not taken<br />

into account. If the ACK theory is applied, σm ff equals the value <strong>of</strong> the failure<br />

stress <strong>of</strong> the matrix σmu.<br />

fibres<br />

TENSION<br />

crack<br />

f<br />

crack<br />

x<br />

σm<br />

σm,min<br />

σm,max<br />

σm ff = σmu<br />

σm,min = minimum normal stress in matrix<br />

δ0<br />

σf,min<br />

σm,max = maximum normal stress in matrix<br />

σf,min = minimum normal stress in fibres<br />

σf,max = maximum normal stress in fibres<br />

σm ff = far field normal stress in matrix<br />

σf<br />

σf,max<br />

Figure 2.9: normal stresses in matrix <strong>and</strong> fibres after occurrence <strong>of</strong> full multiple cracking<br />

24<br />

σ


Chapter 2: IPC composite specimens under monotonic loading<br />

Since the average final crack spacing f is smaller than 2δ0, the maximum<br />

matrix stress σm,max is lower than σmu. Equation (2.18) expresses that the ratio <strong>of</strong><br />

σm,max on σmu equals the ratio <strong>of</strong> f/2 on δ0. This relationship is also illustrated<br />

in figure 2.9 by the black dashed line.<br />

σ m,<br />

max σ m(<br />

x =< cs ><br />

=<br />

σ<br />

σ<br />

f<br />

/ 2)<br />

1.<br />

337δ<br />

0 =<br />

2δ<br />

(2.18)<br />

mu<br />

mu<br />

0<br />

The minimum matrix stress σm,min is found at the crack face (x = 0). The maximum<br />

matrix stress σm,max is formulated by equation (2.18) <strong>and</strong> is found at x = f/2.<br />

The value <strong>of</strong> the maximum fibre stress is found where the minimum matrix stress<br />

occurs <strong>and</strong> vice versa. The minimum fibre stress is found from combination <strong>of</strong><br />

equation (2.12) <strong>and</strong> (2.18):<br />

σ<br />

( x<br />

cs<br />

/ 2)<br />

σ<br />

− 0.<br />

668σ<br />

mc<br />

mu m<br />

f , min = σ f =< > f =<br />

V<br />

(2.19)<br />

f<br />

At the crack face, the fibres take the full external loading. The maximum fibre<br />

stress is thus:<br />

σ mc<br />

σ f , max = σ f ( x = 0)<br />

=<br />

V<br />

(2.20)<br />

f<br />

Since the fibre stress varies linearly from σf,min to σf,max along a distance f/2,<br />

the average fibre stress along f is:<br />

σ f<br />

along < cs><br />

f<br />

σ mc − 0.<br />

334σ<br />

muV<br />

=<br />

V<br />

f<br />

m<br />

V<br />

(2.21)<br />

The average fibre strain along f is found by dividing the average fibre stress<br />

along f by Ef. The composite strain εc equals the average fibre strain along the<br />

composite. Therefore εc zoneII is:<br />

ε<br />

zoneII<br />

c<br />

=<br />

ε<br />

f<br />

along < cs><br />

f<br />

=<br />

σ<br />

f<br />

along < cs><br />

E<br />

f<br />

f<br />

σ mc − 0.<br />

334σ<br />

muV<br />

=<br />

E V<br />

f<br />

f<br />

m<br />

(2.22)<br />

Or after implementation <strong>of</strong> equation (2.10) <strong>and</strong> (2.9):<br />

( E fV<br />

zoneII<br />

f<br />

εc<br />

=<br />

+ EmVm<br />

) ε mu − 0.<br />

334EmVmε<br />

mu<br />

E V<br />

(2.23)<br />

Or:<br />

zoneII<br />

σ mu<br />

σ<br />

ε c = ( 1+<br />

0.<br />

666α<br />

) εmu<br />

= ( 1+<br />

0.<br />

666α<br />

) = ( 1+<br />

0.<br />

666α<br />

)<br />

Em<br />

E<br />

EmVm<br />

<strong>with</strong>:<br />

α =<br />

E V<br />

f<br />

f<br />

f<br />

f<br />

mc<br />

c1<br />

(2.24)<br />

2.4.3.3 zone III<br />

This is the post-cracking zone. Once full multiple cracking has occurred, the IPC<br />

matrix stresses stay constant <strong>with</strong> increasing external load. Only the fibres<br />

25


Chapter 2: IPC composite specimens under monotonic loading<br />

dedicate to the stiffness, if the composite is further loaded into zone III. The<br />

stiffness <strong>of</strong> the composite Ec in this zone is thus:<br />

E = E V<br />

(2.25)<br />

c f f<br />

In zone III the composite strain is the sum <strong>of</strong> εc zoneII (equation (2.24)) <strong>and</strong> the extra<br />

strain term due to further stretching <strong>of</strong> the fibres:<br />

zoneIII<br />

σ mc<br />

( )<br />

( σ c − σ mc )<br />

εc<br />

( σ c)<br />

= 1+<br />

0.<br />

666α<br />

+<br />

E E V<br />

(2.26)<br />

c1<br />

The composite fails when the fibre failure stress is reached. The composite failure<br />

stress σcu is:<br />

σ cu = σ fuV<br />

f<br />

(2.27)<br />

2.4.4 points <strong>of</strong> discussion in the ACK theory<br />

The stress-strain curve <strong>of</strong> a brittle matrix composite is formulated in a rather<br />

simplified way by the ACK theory. However, the ACK theory hides many details<br />

<strong>of</strong> imperfections <strong>and</strong> complex stress fields <strong>and</strong> non-deterministic material<br />

properties <strong>with</strong>in a composite. There are two main shortcomings in this theory,<br />

concerning the implementation <strong>of</strong> the stress-strain behaviour <strong>of</strong> UD-reinforced<br />

IPC composite faces in s<strong>and</strong>wich panels.<br />

1) The ACK theory as presented above is verified experimentally by<br />

Bauweraerts et al. (1998a), Bauweraerts (1998b), Gu et al. (1998) <strong>and</strong> Cuypers<br />

(1999). Experimentally obtained stress-strain curves showed that the measured<br />

stiffness in zone III is generally lower than the theoretically predicted value.<br />

Several factors like imperfect fibre alignment or warping <strong>of</strong> unidirectional fibres,<br />

fibre bundle effect, inhomogeneous matrix impregnation, imperfections at the<br />

debonded fibre-matrix interface, existence <strong>of</strong> local stresses, etc. all have an<br />

influence on the composite behaviour in zone III. Since these effects are difficult<br />

to distinguish from each other, Gu (1998) proposed to bundle them in one<br />

efficiency factor, K. Equation (2.26) now becomes:<br />

zoneIII<br />

σ mc<br />

( )<br />

( σ c − σ mc )<br />

εc<br />

( σ c)<br />

= 1+<br />

0.<br />

666α<br />

+<br />

E KE V<br />

(2.28)<br />

c1<br />

A typical value for the efficiency factor K is 0.95 to 1 for unidirectionally<br />

reinforced IPC composites, according to Cuypers (1999).<br />

2) One <strong>of</strong> the basic assumptions in the ACK theory formulates that the<br />

matrix strength is a deterministic variable. This eliminates the true heterogeneous<br />

strength distribution in the matrix, which is a cementitious material. Multiple<br />

cracking can be initiated well below the value σmc <strong>of</strong> the ACK theory <strong>and</strong> full<br />

multiple cracking is only reached well beyond this stress.<br />

26<br />

f<br />

f<br />

f<br />

f


Chapter 2: IPC composite specimens under monotonic loading<br />

Two laminates are made to illustrate the large stress range in which multiple<br />

cracking really occurs in IPC composite specimens. Figure 2.10a illustrates the<br />

increasing number <strong>of</strong> cracks versus external load for a UD-reinforced specimen<br />

(Vf = 15%). The pictures have been taken under a stereomicroscope after the<br />

specimen was coated <strong>with</strong> an ink solution to visualise the cracks. Figure 2.10b<br />

shows the same phenomenon for a 2D-r<strong>and</strong>om reinforced IPC specimen (Vf =<br />

10%).<br />

It can be noticed from figure 2.10a <strong>and</strong> 2.10b that multiple cracking is not<br />

introduced at one certain stress level, but that the average crack spacing is<br />

function <strong>of</strong> the composite stress. The first matrix cracks appear at relatively low<br />

composite stress, typically between 0 to 5MPa. The final average crack spacing<br />

f is reached above 10MPa.<br />

1 mm<br />

TENSION<br />

Figure 2.10a: gradual multiple cracking <strong>of</strong><br />

UD-reinforced IPC specimen<br />

0MPa<br />

5MPa<br />

10MPa<br />

failure<br />

1 mm<br />

Figure 2.10b: gradual multiple cracking <strong>of</strong><br />

2D-r<strong>and</strong>omly reinforced IPC specimen<br />

From figure 2.10a <strong>and</strong> 2.10b, it can be seen that matrix cracks develop <strong>with</strong>in a<br />

composite stress region rather than at one fixed multiple cracking stress value. The<br />

first theory, which has been presented here, is the ACK theory. The second theory<br />

27


Chapter 2: IPC composite specimens under monotonic loading<br />

that is presented <strong>and</strong> discussed further in this work, is based on the same<br />

assumptions as the ACK theory. The only difference between this second theory<br />

<strong>and</strong> the ACK theory is that the assumption <strong>of</strong> a unique matrix strength is<br />

ab<strong>and</strong>oned. It would be more appropriate to assume that the distribution <strong>of</strong> the<br />

matrix strength obeys a certain probability distribution function (PDF). This<br />

assumption is used in the second theory, which is therefore called “the stochastic<br />

cracking theory”. Figure 2.11 shows schematically a stress-strain diagram <strong>of</strong> a<br />

composite, when matrix multiple cracking occurs at several stress levels (figure<br />

2.11 can be compared <strong>with</strong> figure 2.5).<br />

Figure 2.11: partial multiple cracking at several stress levels, after Bentur <strong>and</strong> Mindess (1990)<br />

<strong>and</strong> Allen (1971)<br />

2.4.5 the statistical nature <strong>of</strong> the IPC matrix tensile strength<br />

If the assumption on the deterministic matrix strength is ab<strong>and</strong>oned, the first<br />

question to be answered is how the statistical nature <strong>of</strong> the matrix strength can be<br />

modelled. Various types <strong>of</strong> models can be found in literature.<br />

The most common used models are:<br />

1. Normal model<br />

2. Cauchy model<br />

3. Lognormal model<br />

4. Weibull model (derived from this model: Exponential model)<br />

5. Largest extreme value model (LEV)<br />

6. Smallest extreme value model (SEV)<br />

All these models are based on a certain formulation <strong>of</strong> the shape <strong>of</strong> the probability<br />

distribution function. The probability distribution function (PDF) expresses the<br />

possibility that a certain variable X takes the value x:<br />

f ( x)<br />

= P(<br />

X = x)<br />

(2.29)<br />

28


Chapter 2: IPC composite specimens under monotonic loading<br />

There are several ways to select an appropriate model (Van Vinckenroy, 1994)<br />

1. The far most desirable method to choose a model is based on<br />

underst<strong>and</strong>ing <strong>of</strong> the basic physical phenomena, which lead to the statistical<br />

distribution <strong>of</strong> the measured variable. A model is then selected, which nature is<br />

based on the same phenomena.<br />

2. If the underlying phenomena leading to a statistical distribution <strong>of</strong> a<br />

variable are not understood well, a model may be used that has been applied<br />

successfully in the past for related problems. The chosen model should then be<br />

verified if data are available.<br />

3. If no a priori knowledge is available about the phenomena leading to the<br />

statistical nature <strong>of</strong> the measured variable, all possible models can be applied one<br />

by one to fit available data. Therefore, sufficient data should be available. The<br />

model that seems to fit the data best is <strong>with</strong>held, regardless <strong>of</strong> the underlying<br />

philosophy <strong>of</strong> the model.<br />

In this work a combination <strong>of</strong> the first <strong>and</strong> second selection method is used to<br />

choose an appropriate statistical model to describe the probability distribution<br />

function <strong>of</strong> the IPC matrix strength.<br />

Many books <strong>and</strong> papers deal <strong>with</strong> the statistical nature <strong>of</strong> the strength <strong>of</strong> brittle<br />

materials. Most authors recommend the use <strong>of</strong> a Weibull probability distribution<br />

function to describe the stochastic nature <strong>of</strong> the strength <strong>of</strong> brittle materials, e.g.<br />

Creyke et al. (1982), Davidge (1979), Chawla (1993), Matthews <strong>and</strong> Rawlings<br />

(1994), Curtin et al. (1993) <strong>and</strong> Zok <strong>and</strong> Spearing (1992).<br />

The Weibull model is based on the weakest-link principle (Weibull, 1952).<br />

Materials contain inherent flaws. According to the Weibull model, the propagation<br />

<strong>of</strong> the largest flaw leads to fracture. However, cementitious materials do not<br />

automatically behave as ideal brittle solids. The Weibull model neglects<br />

redistribution <strong>of</strong> stresses in the vicinity <strong>of</strong> damaged zones. For IPC, the interaction<br />

<strong>of</strong> defects might lead to final fracture, rather than the propagation <strong>of</strong> one critical<br />

flaw. Up to now, there is no knowledge about the micro-mechanics in pure IPC<br />

matrix in tension. For this reason, several distribution functions are checked here<br />

on their appropriateness to describe the statistical nature <strong>of</strong> the IPC matrix tensile<br />

strength.<br />

In Appendix 1, details on the mentioned probability distribution functions are<br />

given. First, all distribution functions allowing the existence <strong>of</strong> negative values <strong>of</strong><br />

the studied parameter are considered to be non-appropriate here. These are the<br />

Normal model, the Cauchy model <strong>and</strong> the Smallest extreme value model.<br />

29


Chapter 2: IPC composite specimens under monotonic loading<br />

The Gamma model allows probability <strong>of</strong> failure at zero stress <strong>and</strong> is therefore also<br />

removed from the possible models for IPC. The three remaining models are the<br />

Lognormal model, the Weibull model <strong>and</strong> the Largest extreme value model.<br />

Which <strong>of</strong> these models describes best tensile failure <strong>of</strong> IPC, is determined by<br />

comparison <strong>of</strong> experimental data <strong>with</strong> the theoretical probability functions.<br />

Four solid IPC plates are prepared. After demoulding, all plates are post-cured for<br />

24 hours at 60°C. In total 34 specimens <strong>of</strong> dimensions 30x15x160mm³ are cut<br />

from these IPC plates, using a water-cooled diamond saw. The specimens are<br />

tested in three-point bending <strong>with</strong> the mechanical INSTRON 1195 testing bench,<br />

according to National Belgian St<strong>and</strong>ards: NBN-B12-208 <strong>and</strong> NBN-B14-209. A<br />

load cell is used to record the failure load. The failure bending stress is calculated<br />

from knowledge <strong>of</strong> the failure load <strong>and</strong> the geometry <strong>of</strong> the specimens. The failure<br />

stress values are ranked from the lowest to the highest value. Figure 2.12 shows<br />

the number <strong>of</strong> times a bending strength is measured for the IPC specimens. The<br />

average value <strong>of</strong> the bending failure stress is 10.5MPa, which is situated in the<br />

interval defined in table 2.2.<br />

The parameters <strong>of</strong> the Lognormal model, the Weibull model <strong>and</strong> the Largest<br />

extreme value model are to be determined in such a way that the best fit is found<br />

for the theoretical model <strong>with</strong> the experimental data. Various model parameter<br />

estimation techniques are available (Van Vinckenroy, 1994). The technique, which<br />

is applied here, is probability plotting. Details on this technique are found in<br />

Appendix 2. The Lognormal model <strong>and</strong> Largest extreme value model are<br />

determined by two parameters. The number <strong>of</strong> parameters for the Weibull model<br />

is two or three, depending on the choice <strong>of</strong> the user. Both the two- <strong>and</strong> threeparameter<br />

Weibull model are discussed here.<br />

frequency<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

7.5 8.5 9.5 10.5 11.5 12.5 13.5<br />

IPC bending strength (MPa)<br />

Figure 2.12: number <strong>of</strong> times a certain bending strength is measured, pure IPC specimens<br />

Table 2.6 lists the values <strong>of</strong> the model parameters, obtained by probability<br />

plotting. The goodness-<strong>of</strong>-fit parameter is also listed in this table. The goodness<strong>of</strong>-fit<br />

parameter is a measure for the appropriateness <strong>of</strong> a model to fit experimental<br />

30


Chapter 2: IPC composite specimens under monotonic loading<br />

data. If the model completely fits the experimental probability distribution, this<br />

parameter equals one. In figure 2.13, the four discussed probability distribution<br />

functions are plotted together <strong>with</strong> the experimental data.<br />

Table 2.6: parameters <strong>of</strong> probability distribution models <strong>and</strong> goodness-<strong>of</strong>-fit parameter,<br />

obtained by probability plotting<br />

Lognormal two-parameter three-parameter Largest extreme<br />

Weibull Weibull value<br />

parameter 1 µlog = 2.35 σRb = 11.2 σRb = 8.43 µlev = 10.0<br />

parameter 2 σlog = 0.141 m = 9.32 m = 6.76 ηlev = 1.16<br />

parameter 3 - - σ0b = 2.75 -<br />

goodness-<strong>of</strong>-fit 0.950 0.974 0.976 0.947<br />

Frame 001 ⏐ 16 Aug 2001 ⏐ | | | | |<br />

f(x)<br />

0.35<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

experimental<br />

lognormal<br />

two parameter Weibull<br />

three parameter Weibull<br />

LEV<br />

5 10 15 20<br />

bending strength (MPa)<br />

Figure 2.13: theoretical versus experimental probability functions <strong>of</strong> the bending strength <strong>of</strong> pure<br />

IPC matrix<br />

It can be seen from table 2.6 that the three-parameter Weibull model fits the<br />

experimental results best.<br />

From figure 2.13 it can be noticed that the two-parameter <strong>and</strong> three-parameter<br />

Weibull model practically coincide. Also from table 2.6 it can be seen that the<br />

goodness-<strong>of</strong>-fit parameter is only slightly lower in case the two-parameter Weibull<br />

model is used instead <strong>of</strong> the three-parameter model.<br />

31


Chapter 2: IPC composite specimens under monotonic loading<br />

As a conclusion, the Weibull probability distribution seems to be the most<br />

appropriate choice to model the stochastic nature <strong>of</strong> the IPC bending strength. The<br />

use <strong>of</strong> a three-parameter Weibull model is only slightly more appropriate than the<br />

use <strong>of</strong> a two-parameter Weibull model, at the cost <strong>of</strong> the effort <strong>of</strong> determination <strong>of</strong><br />

the third parameter. A two-parameter Weibull model is therefore used in this work<br />

to describe the stochastic nature <strong>of</strong> the IPC matrix strength in tension.<br />

2.4.6 the Weibull probability distribution<br />

In previous paragraph, the choice <strong>of</strong> a Weibull model to describe the stochastic<br />

nature <strong>of</strong> the IPC matrix strength has been justified. Therefore, some information<br />

on this probability distribution function is presented here.<br />

The cumulative distribution function (CDF) is defined as the integral <strong>of</strong> the<br />

probability distribution function (PDF):<br />

x<br />

∫ ∞<br />

−<br />

F ( x)<br />

= P(<br />

X ≤ x)<br />

= f ( u)<br />

du<br />

(2.30)<br />

Weibull (1952) assumed a certain shape <strong>of</strong> the cumulative distribution function for<br />

the description <strong>of</strong> failure <strong>of</strong> a specimen in uniform tension. The shape <strong>of</strong> this CDF<br />

was obtained empirically:<br />

P<br />

m<br />

⎡ ⎛σ −σ<br />

⎞ ⎤<br />

0<br />

= 1 − exp⎢−<br />

⎜<br />

⎟ ⎥<br />

⎢<br />

⎣ ⎝ σ R ⎠ ⎥<br />

⎦<br />

(2.31)<br />

where: σ = uniform tensile stress in the material<br />

σ0 = maximum stress at which the probability <strong>of</strong> failure is still zero<br />

σR = reference failure stress<br />

m = Weibull modulus<br />

Equation (2.31) is the three-parameter Weibull CDF. If the two-parameter Weibull<br />

CDF is used, equation (2.31) becomes:<br />

m ⎡ ⎛ σ ⎞ ⎤<br />

P = 1 − exp⎢−<br />

⎜<br />

⎟ ⎥<br />

(2.32)<br />

⎢<br />

⎣ ⎝σ<br />

R ⎠ ⎥<br />

⎦<br />

The two-parameter model is used further in this work, for reasons explained in<br />

paragraph 2.4.5.<br />

Equation (2.32) can be used to obtain the CDF <strong>of</strong> a series <strong>of</strong> specimens <strong>with</strong> equal<br />

geometry, loaded in pure tension. If a new series <strong>of</strong> larger specimens is tested, the<br />

probability <strong>of</strong> failure increases for a same stress level, since the number <strong>of</strong><br />

inherent flaws is larger in larger specimens. The CDF <strong>of</strong> this new series, <strong>with</strong><br />

specimens <strong>of</strong> volume V, can be linked <strong>with</strong> the CDF <strong>of</strong> the previous series, <strong>with</strong><br />

volume VR, as follows:<br />

32


Chapter 2: IPC composite specimens under monotonic loading<br />

P<br />

m<br />

⎡ ⎛ V ⎞⎛<br />

σ ⎞ ⎤<br />

= 1 − exp⎢−<br />

⎜<br />

⎟<br />

⎜<br />

⎟ ⎥<br />

⎢<br />

⎣ ⎝VR<br />

⎠⎝<br />

σ R ⎠ ⎥<br />

⎦<br />

(2.33)<br />

<strong>with</strong> m <strong>and</strong> σR: the model parameters determined for the first series <strong>of</strong> specimens.<br />

If a non-uniform tensile stress state is to be expected in the specimen, equation<br />

(2.33) has to be modified to include these stress variations in the specimen:<br />

P<br />

m<br />

⎡ ⎛ V ⎞⎛<br />

σ ⎞ ( ) ⎤<br />

nom σ V dV<br />

= 1 − exp⎢−<br />

⎜<br />

⎟<br />

⎜<br />

⎟ ∫ ⎥<br />

⎢<br />

⎣ ⎝VR<br />

⎠⎝<br />

σ R ⎠ V σ nom V ⎥<br />

⎦<br />

(2.34)<br />

where: σnom = nominal stress in the specimen<br />

When a three-point bending test is performed, the normal stresses in the specimen<br />

are non-uniform. Equation (2.34) becomes:<br />

P<br />

m<br />

⎡<br />

l/ 2 b/ 2 m<br />

⎛ w ⎞⎛<br />

σ ⎞ ⎛ ⎞<br />

⎤<br />

max 4xy<br />

= 1−<br />

exp⎢−<br />

2 ⎜<br />

⎟<br />

⎜<br />

⎟ ∫∫ ⎜ ⎟ dxdy⎥<br />

⎢⎣<br />

⎝V<br />

R ⎠⎝<br />

σ R ⎠ 0 0 ⎝ lh ⎠ ⎥⎦<br />

(2.35)<br />

where: w = width <strong>of</strong> the specimens<br />

l = length <strong>of</strong> the specimens<br />

h = height <strong>of</strong> the specimens<br />

After integration <strong>and</strong> rearranging <strong>of</strong> equation (2.35), equation (2.36) is found:<br />

m<br />

⎡ ⎛<br />

⎞⎛<br />

⎞ ⎤<br />

= − ⎢−<br />

⎜<br />

⎟<br />

⎜<br />

⎟ ⎥<br />

⎢<br />

⎣ ⎝ R + ⎠⎝<br />

R ⎠ ⎥<br />

⎦<br />

m V<br />

P<br />

V 1 σ max<br />

1 exp<br />

2<br />

2 ( 1)<br />

σ<br />

(2.36)<br />

Equation (2.36) can be rewritten:<br />

P<br />

m<br />

⎡ ⎛<br />

⎞ ⎤<br />

⎢ ⎜<br />

⎟ ⎥<br />

m<br />

⎢ ⎜ σ ⎟ ⎥ ⎡ ⎛ ⎞ ⎤<br />

max σ max<br />

= 1 − exp⎢−<br />

⎜<br />

⎟ ⎥ = 1−<br />

exp⎢−<br />

⎜<br />

⎟ ⎥<br />

2<br />

1 / m<br />

⎢ ⎜ ⎛ 2V<br />

+ ⎞ ⎟ ⎥ ⎢<br />

⎣ ⎝ ⎠ ⎥<br />

R(<br />

m 1)<br />

σ Rb ⎦<br />

⎢ ⎜ ⎜<br />

⎟ σ R ⎟ ⎥<br />

⎢⎣<br />

⎝ ⎝ V ⎠ ⎠ ⎥⎦<br />

(2.37a)<br />

<strong>with</strong>:<br />

σ<br />

( m 1)<br />

2<br />

1 / m<br />

⎛ 2VR<br />

+ ⎞<br />

⎜ ⎟<br />

Rb = σ R<br />

(2.37b)<br />

⎜<br />

⎝<br />

V<br />

⎟<br />

⎠<br />

Equation (2.37a) <strong>and</strong> (2.37b) provide the link between the reference failure stress<br />

<strong>of</strong> brittle specimens loaded in uniform tension (σR) <strong>with</strong> volume VR <strong>and</strong> the<br />

reference failure stress <strong>of</strong> specimens loaded in three-point bending (σRb) <strong>with</strong><br />

volume V. The values <strong>of</strong> m <strong>and</strong> σRb have been determined for IPC specimens in<br />

three-point bending in previous paragraph. These values are to be determined for<br />

33


Chapter 2: IPC composite specimens under monotonic loading<br />

IPC specimens in uniform tension. The Weibull modulus m, determined under<br />

three-point bending, equals the Weibull modulus under uniform tension. The<br />

Weibull parameter σR, which represents the reference cracking stress under<br />

uniform tension, is calculated from σRb <strong>and</strong> equation (2.37b).<br />

2.4.7 formulation <strong>of</strong> the stochastic cracking theory<br />

The second model in this chapter differs from the ACK theory in one assumption.<br />

According to the ACK theory, it is assumed that the matrix failure strength is a<br />

deterministic parameter. In the “stochastic cracking theory” the nature <strong>of</strong> the<br />

matrix failure strength is introduced via a Weibull probability distribution.<br />

Price <strong>and</strong> Smith (1993) <strong>and</strong> He et al. (1994) demonstrated that the stress-strain<br />

curve <strong>of</strong> a brittle matrix composite can be calculated, provided the average crack<br />

spacing is measured in function <strong>of</strong> the applied composite stress σc.<br />

Curtin (1993) <strong>and</strong> Zok <strong>and</strong> Spearing (1992) considered the statistical evolution <strong>of</strong><br />

multiple cracking as function <strong>of</strong> the micro-mechanical parameters. They provided<br />

a formulation <strong>of</strong> as function <strong>of</strong> the applied composite stress σc.<br />

Finally Curtin et al. (1998, 1999) presented a model, which combines prediction<br />

<strong>of</strong> the stress-strain curve from knowledge <strong>of</strong> <strong>and</strong> the formulation <strong>of</strong> as a<br />

function <strong>of</strong> the material properties <strong>and</strong> the composite stress, σc. This way, Curtin et<br />

al. (1998, 1999) presented a statistical treatment <strong>of</strong> matrix crack evolution in UDreinforced<br />

ceramic micro-composites. The studied micro-composites were made<br />

<strong>of</strong> Nicalon fibres <strong>with</strong> a thin Carbon coating, which where infiltrated <strong>with</strong> a SiC<br />

matrix. Curtin et al. (1998, 1999) used a Weibull distribution function to describe<br />

the stochastic distribution <strong>of</strong> the matrix strength.<br />

The fundamental difference between the stochastic cracking approach <strong>and</strong> the<br />

deterministic ACK theory is that cracks can now occur gradually at various stress<br />

levels. The crack spacing is thus not only <strong>of</strong> stochastic nature in its spatial<br />

distribution, but also in its development as function <strong>of</strong> the external composite load.<br />

When the first matrix crack is initiated in the studied cementitious matrix<br />

composite, the fibres are assumed to bridge this matrix crack (this assumption is<br />

also used in the ACK theory). The external load can therefore be increased,<br />

leading to the initiation <strong>and</strong> propagation <strong>of</strong> new matrix cracks at higher matrix<br />

stress levels. This way, equation (2.32) provides information about the percentage<br />

<strong>of</strong> inherent matrix flaws, which are able to propagate at a certain stress level. If<br />

full multiple cracking occurred, the average crack spacing equals f,<br />

being the final crack spacing. If only partial multiple cracking occurred, is<br />

larger than the final crack spacing. The average crack spacing at a certain<br />

composite stress is found by dividing the final crack spacing f by the<br />

percentage <strong>of</strong> matrix cracks that already propagated:<br />

34


Chapter 2: IPC composite specimens under monotonic loading<br />

m<br />

⎛ ⎡ ⎞<br />

⎜ ⎛ σ ⎞ ⎤<br />

c<br />

cs = cs 1−<br />

exp⎢−<br />

⎥⎟<br />

⎜ ⎜<br />

⎟<br />

(2.38)<br />

f<br />

⎟<br />

⎝<br />

⎢<br />

⎣ ⎝ σ R ⎠ ⎥<br />

⎦⎠<br />

The stress terms in equation (2.38) are composite stresses instead <strong>of</strong> matrix<br />

stresses. Composite stresses <strong>and</strong> matrix stresses are related to each other. Since the<br />

behaviour <strong>of</strong> the composite is discussed here, it is more convenient to work <strong>with</strong><br />

composite stresses.<br />

Suppose for now that IPC matrix specimens (no fibres) <strong>with</strong> a length <strong>of</strong> 250mm, a<br />

thickness <strong>of</strong> 3mm <strong>and</strong> a width <strong>of</strong> 20mm are tested. Theoretically, the Weibull<br />

modulus <strong>of</strong> the matrix material in pure tension equals the Weibull modulus<br />

obtained in three-point bending. m would therefore be 9.3 (table 2.6). According<br />

to the obtained value <strong>of</strong> σRb in table 2.6 <strong>and</strong> equation (2.37b), the matrix reference<br />

cracking stress under uniform tensile stress is 8.2MPa.<br />

The matrix reference cracking stress <strong>of</strong> 8.2MPa is now modified towards a<br />

composite reference multiple cracking stress. The ratio between the composite<br />

reference cracking stress <strong>and</strong> the matrix reference cracking stress equals the ratio<br />

between their initial stiffness. Provided the fibre volume fraction is about 10%, the<br />

value <strong>of</strong> the composite reference cracking stress is thus about 10.5MPa.<br />

The preparation technique <strong>of</strong> IPC as homogeneous solid block <strong>and</strong> as matrix is<br />

different. Due to the different processing technique <strong>and</strong> the presence <strong>of</strong> fibres, the<br />

inherent flaw size distribution might change considerably. Therefore, the value <strong>of</strong><br />

the Weibull parameters obtained more early from tests on pure IPC specimens<br />

provide only an order <strong>of</strong> magnitude rather than an accurate prediction <strong>of</strong> the<br />

Weibull parameters <strong>of</strong> IPC used as matrix in composites.<br />

The average crack spacing <strong>and</strong> the debonding length δ0 are both function <strong>of</strong><br />

the applied composite stress. Formulation <strong>of</strong> the stress-strain relationship should<br />

be done for two cases. Both are illustrated in figure 2.14. Figure 2.14a shows the<br />

matrix <strong>and</strong> fibre stress in the composite if the average crack spacing is larger<br />

than twice the debonding length, 2δ0. Figure 2.14b shows these stresses if is<br />

smaller than 2δ0.<br />

The formulation <strong>of</strong> the debonding length is similar to equation (2.13):<br />

ff<br />

Vmrσ<br />

m δ 0 =<br />

V 2τ VmrEmσ<br />

c =<br />

E V 2τ<br />

(2.39)<br />

f<br />

0<br />

c1<br />

f<br />

σmu in equation (2.13) is replaced by σm ff in equation (2.39), since there is no<br />

unique matrix cracking stress in the stochastic cracking theory. The far field<br />

matrix stress σm ff is the stress in the matrix at theoretical infinite distance from one<br />

35<br />

0<br />

−1


Chapter 2: IPC composite specimens under monotonic loading<br />

crack, provided the presence <strong>of</strong> other cracks is neglected <strong>and</strong> is expressed by<br />

equation (2.40) (similarly to the definition <strong>of</strong> σmu, according to equation (2.10)).<br />

ff Emσ c σ m = (2.40)<br />

E<br />

c1<br />

It should be mentioned that equation (2.39) does not indicate that a matrix crack<br />

exist where the far field matrix stress σm ff is reached: it only states that the<br />

debonding length δ0 can be calculated if a crack exist. The formulation <strong>of</strong> the<br />

stress-strain relationship is further separately expressed for the two cases,<br />

illustrated in figures 2.14a <strong>and</strong> 2.14b.<br />

<br />

x<br />

δ0<br />

σm σf<br />

δ0<br />

σm,min σf,min σf,max<br />

σm,max = σm ff<br />

Figure 2.14a: stresses in matrix <strong>and</strong> fibre along<br />

a cracked composite, > 2δ0<br />

σ<br />

<br />

x<br />

σm,min<br />

σm<br />

σm,max<br />

δ0<br />

σm ff<br />

σf<br />

σf,min<br />

σf,max<br />

Figure 2.14b: stresses in matrix <strong>and</strong> fibre along<br />

a cracked composite, < 2δ0<br />

2.4.7.1 formulation when > 2δ0<br />

When > 2δ0, the formulation <strong>of</strong> matrix <strong>and</strong> fibre stresses <strong>and</strong> strains is based<br />

on figure 2.14a. The minimum matrix stress is found at the crack face (x = 0):<br />

σ m,<br />

min(<br />

x = 0)<br />

= 0<br />

(2.41)<br />

The maximum fibre stress is thus also found at the crack face <strong>and</strong> is:<br />

σ c<br />

σ f , max(<br />

x = 0)<br />

=<br />

V<br />

(2.42)<br />

The maximum matrix stress is found between (x = δ0) <strong>and</strong> (x = - δ0). Since<br />

the composite still behaves linear elastically in this zone:<br />

f<br />

36<br />

σ


Chapter 2: IPC composite specimens under monotonic loading<br />

E σ<br />

σ ( δ x<br />

=<br />

m,<br />

max<br />

[ < cs > −δ<br />

]<br />

ff m c<br />

0 < <<br />

0 ) = σ m<br />

(2.43)<br />

Ec1<br />

The minimum fibre stress is thus found between (x = δ0) <strong>and</strong> (x = - δ0) <strong>and</strong><br />

is:<br />

E fσ<br />

c<br />

σ f , min ( δ 0 < x < [ < cs > −δ<br />

0]<br />

) =<br />

Ec1<br />

The average fibre stress along δ0 is then:<br />

(2.44)<br />

/ 2<br />

0<br />

⎟<br />

1<br />

⎟<br />

σ f<br />

⎛ ⎞<br />

⎜<br />

σ E fσ<br />

c<br />

c<br />

= +<br />

alongδ<br />

⎜<br />

⎝V<br />

f Ec<br />

⎠<br />

(2.45)<br />

<strong>and</strong> the average fibre stress along is:<br />

σ f<br />

⎛⎛<br />

⎞<br />

⎞<br />

⎜⎜<br />

σ E fσ<br />

c<br />

E fσ<br />

c<br />

c<br />

= + ⎟ + ( < > − ) ⎟ 1<br />

δ cs<br />

along < cs><br />

⎜⎜<br />

⎟ 0 2δ<br />

0 ⎟<br />

⎝⎝<br />

V f Ec1<br />

⎠<br />

Ec1<br />

⎠<br />

< cs ><br />

(2.46)<br />

or rewritten:<br />

σ f<br />

σ cE<br />

f<br />

=<br />

along < cs><br />

E<br />

δ 0 EmVmσ<br />

c<br />

+<br />

< cs > V E<br />

(2.47)<br />

c1<br />

f c1<br />

The composite strain is found by dividing equation (2.47) by Ef:<br />

εc<br />

= ε f<br />

σ c EmVmδ<br />

0<br />

= + σ<br />

along < cs><br />

c<br />

E E E V<br />

or rewritten:<br />

c1<br />

f<br />

c<br />

f<br />

(2.48)<br />

⎛ ⎞<br />

c ⎜ 0 ⎟<br />

c = 1+ E ⎜ ⎟<br />

c1<br />

⎝ cs ⎠<br />

αδ σ<br />

ε (2.49)<br />

2.4.7.2 formulation when < 2δ0<br />

When the average crack spacing is smaller than 2δ0, the derivation <strong>of</strong> the<br />

composite strain versus stress is based on figure 2.14b. As can be seen from the<br />

black dashed line in figure 2.14b, the ratio <strong>of</strong> the maximum stress in the matrix on<br />

the far field matrix stress is:<br />

σ m,<br />

max cs<br />

= ff<br />

(2.50)<br />

σ m 2δ 0<br />

The minimum <strong>and</strong> maximum stresses in the matrix are, according to figure 2.14b:<br />

σ m,<br />

min(<br />

x = 0)<br />

= 0<br />

(2.51)<br />

cs ff cs Emσ<br />

c<br />

σ m,<br />

max(<br />

x =< cs > / 2)<br />

= σ m =<br />

(2.52)<br />

2δ<br />

0 2δ<br />

0 Ec1<br />

The maximum <strong>and</strong> minimum fibre stresses along the length <strong>of</strong> the composite are<br />

then:<br />

σ c<br />

σ f , max(<br />

x = 0)<br />

=<br />

V<br />

(2.53)<br />

f<br />

37


Chapter 2: IPC composite specimens under monotonic loading<br />

σ<br />

f , min<br />

( x =< cs > / 2)<br />

cs Emσ<br />

c<br />

σ c − V<br />

2δ<br />

0 Ec1<br />

=<br />

V<br />

f<br />

m<br />

(2.54)<br />

The average fibre stress along is thus:<br />

σ f<br />

⎛<br />

⎜<br />

1<br />

= σ<br />

along < cs ><br />

c ⎜<br />

⎝V<br />

f<br />

E ⎞<br />

mVm<br />

cs<br />

− ⎟<br />

4V<br />

⎟<br />

f δ 0Ec1<br />

⎠<br />

(2.55)<br />

Composite strain εc equals the average fibre strain along <strong>and</strong> is thus:<br />

ε c = ε f<br />

⎛<br />

⎜<br />

1<br />

= σ<br />

along < cs><br />

c⎜<br />

⎝ E fV<br />

f<br />

α cs ⎞<br />

− ⎟<br />

4δ<br />

⎟<br />

0Ec1<br />

⎠<br />

(2.56)<br />

2.4.7.3 σc at which > 2δ0 becomes < 2δ0<br />

If low composite stress is applied, the average crack spacing is large <strong>and</strong> exceeds<br />

2δ0. If composite stress σc is increased, the crack spacing decreases <strong>and</strong> δ0<br />

increases. The average crack spacing is formulated by equation (2.38) <strong>and</strong> the<br />

debonding length is expressed by equation (2.39). The composite stress at which<br />

the average crack spacing equals 2δ0, is thus found from equation (2.57):<br />

m<br />

⎛ ⎡ σ<br />

⎞ E c<br />

c1V<br />

fτ<br />

0<br />

σ<br />

⎜ ⎛ ⎞ ⎤<br />

c 1−<br />

exp⎢−<br />

⎟<br />

= cs<br />

⎜ ⎜ ⎥<br />

f<br />

σ ⎟<br />

⎟<br />

(2.57)<br />

R VmrEm<br />

⎝<br />

⎢⎣<br />

⎝ ⎠ ⎥⎦<br />

⎠<br />

2.4.8 experiments on UD-reinforced IPC specimens<br />

2.4.8.1 specimens<br />

Experimentally obtained stress-strain curves are compared <strong>with</strong> theoretical curves.<br />

Both the ACK theory <strong>and</strong> the stochastic cracking theory are applied <strong>and</strong> discussed.<br />

Two UD-reinforced laminates are made: UD1 <strong>and</strong> UD2. Six layers <strong>of</strong> E-glass<br />

fibres are used in both plates. Two specimens per plate are subjected to tensile<br />

loading up to failure.<br />

Figure 2.15a shows the stress-strain diagrams for the two plates. Figure 2.15b<br />

shows the same stress-strain curves, but <strong>with</strong>in the stress range <strong>of</strong> 0 to 20MPa,<br />

since this is the stress region in which multiple cracking occurs.<br />

The properties <strong>of</strong> the laminates are listed in table 2.7.<br />

Table 2.7: properties <strong>of</strong> UD-reinforced test plates<br />

plate reinforcement matrix amount thickness Vf<br />

(g/layer/m²) (mm) (%)<br />

UD1 unidirectional 750 3.21 10.4<br />

UD2 unidirectional 1100 4.20 7.95<br />

38


stress(Mpa)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Chapter 2: IPC composite specimens under monotonic loading<br />

UD1<br />

UD2<br />

0<br />

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8<br />

strain(%)<br />

Figure 2.15a: stress-strain curves <strong>of</strong> UDreinforced<br />

specimens<br />

stress(Mpa)<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

UD1<br />

UD2<br />

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4<br />

strain(%)<br />

Figure 2.15b: stress interval 0 to 20MPa<br />

2.4.8.2 ACK model versus experimental results<br />

The average value <strong>of</strong> the matrix stiffness Em is about 18MPa according to table<br />

2.2. The average failure stress σmu is situated between 6 to 14MPa, according to<br />

this table.<br />

Figure 2.16a shows the experimental <strong>and</strong> theoretical stress-strain curve <strong>of</strong> plate<br />

UD1. For the determination <strong>of</strong> the theoretical curves, the matrix cracking stress is<br />

varied from 6MPa to 12MPa <strong>with</strong> steps <strong>of</strong> 2MPa. Figure 2.16b shows the<br />

experimental <strong>and</strong> theoretical stress-strain curves <strong>of</strong> plate UD2.<br />

stress (MPa)<br />

20<br />

15<br />

10<br />

5<br />

0<br />

decreasing σmu<br />

experimental<br />

ACK model<br />

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35<br />

strain (%)<br />

Figure 2.16a: stress-strain curves <strong>of</strong> UDreinforced<br />

specimens, UD1<br />

σmu is 6, 8, 10 or 12MPa<br />

stress (MPa)<br />

20<br />

15<br />

10<br />

5<br />

0<br />

decreasing σmu<br />

experimental<br />

ACK model<br />

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35<br />

strain (%)<br />

Figure 2.16b: stress-strain curves <strong>of</strong> UDreinforced<br />

specimens, UD2<br />

σmu is 6, 8, 10 or 12MPa<br />

The value <strong>of</strong> the multiple cracking stress σmu can be estimated from tests on pure<br />

IPC matrix blocks, produced from the same fresh IPC mixture used for the<br />

composite laminates. However, the value <strong>of</strong> σmu may depend on the processing<br />

39


Chapter 2: IPC composite specimens under monotonic loading<br />

technique, efficiency <strong>of</strong> the curing process, presence <strong>of</strong> reinforcement, etc. σmu is<br />

therefore not always known accurately a priori. Another estimation method for<br />

σmu comes from comparison <strong>of</strong> the experimental <strong>and</strong> theoretical stress-strain<br />

curves <strong>of</strong> the specimens in tension. σmu can be varied until the difference between<br />

the theoretical <strong>and</strong> experimental curves is minimised, using a least squares<br />

method. This method is used here for laminates UD1 <strong>and</strong> UD2. The stress range,<br />

which is used for the comparison <strong>of</strong> the experimental <strong>and</strong> theoretical curves, is 0<br />

to 20, 40, 50 or 60MPa. Full multiple cracking occurred <strong>and</strong> no or few fibres are<br />

broken at these stress levels. A Least Squares Coefficient (LSC) is defined:<br />

LSC =<br />

N exp<br />

∑<br />

j = 1<br />

( ε ( σ ) − ε ( σ )<br />

exp<br />

j<br />

N<br />

exp<br />

<strong>with</strong>: Nexp = number <strong>of</strong> experimental points in considered stress interval<br />

σj = experimental stress point j<br />

εexp(σj) = experimental obtained strain at stress point σj<br />

εtheo(σj) = theoretical obtained strain at stress point σj<br />

theo<br />

j<br />

2<br />

(2.58)<br />

Table 2.8 lists the values <strong>of</strong> model parameter σmu, obtained by minimisation <strong>of</strong> the<br />

LSC. The LSC is also listed in this table.<br />

Table 2.8: parameter σmu, obtained from minimisation <strong>of</strong> the LSC for different stress intervals<br />

UD1 UD2<br />

stress interval (MPa) 0 - 20 0 - 40 0 - 50 0 - 60 0 – 20 0 - 40 0 - 50 0 - 60<br />

σmu (MPa) 7.2 7.9 7.8 7.8 7.2 7.2 8.2 8.5<br />

LSC (exp(-4)) 3.2 2.4 1.9 1.5 5.7 3.9 3.5 3.6<br />

It can be seen from table 2.8 that σmu is about 7.5MPa for UD1. For UD2 the<br />

average value <strong>of</strong> σmu is 7.8MPa, but the optimised value <strong>of</strong> σmu varies considerably<br />

<strong>with</strong> the stress interval <strong>of</strong> interest.<br />

2.4.8.3 stochastic cracking model versus experimental results<br />

It has been mentioned that the value <strong>of</strong> the ACK matrix tensile failure stress σmu is<br />

not always known a priori. If the stochastic cracking theory is applied, several<br />

material parameters might not be determined <strong>with</strong> high accuracy from tests on the<br />

separate components <strong>of</strong> the composites. These material properties are:<br />

1. the composite reference cracking stress, σR<br />

2. the Weibull modulus, m<br />

3. the frictional interface shear stress, τ0<br />

If a three-parameter Weibull probability function were used instead <strong>of</strong> a twoparameter<br />

Weibull model, an extra material parameter would be introduced: the<br />

composite stress at which the first crack appears, σ0.<br />

40


Chapter 2: IPC composite specimens under monotonic loading<br />

It has already been mentioned that the inherent flaw size distribution in a<br />

homogeneous solid IPC block might differ considerably from the distribution in<br />

IPC as matrix material. σR <strong>and</strong> m have been derived on pure IPC specimens <strong>and</strong><br />

their value was 10.5 <strong>and</strong> 9.3 (see paragraph 2.4.5 to 2.4.7). The values derived<br />

from the tests on solid IPC are used initially as IPC matrix material parameters in<br />

a laminate.<br />

The value <strong>of</strong> the frictional interface shear stress τ0 can be estimated <strong>with</strong> help <strong>of</strong><br />

the ACK model. The combination <strong>of</strong> equation (2.13) <strong>and</strong> (2.14) gives following<br />

relationship:<br />

1.<br />

337rσ<br />

muVm<br />

τ 0 =<br />

2V<br />

cs<br />

(2.59)<br />

f<br />

f<br />

The average crack spacing f is determined by visualisation <strong>of</strong> the cracks<br />

under the stereomicroscope <strong>with</strong> an ink-solution. The average final crack spacing<br />

f obtained for UD1 is 0.93mm. For UD2, the average final crack spacing is<br />

1.4mm. The value <strong>of</strong> σmu, obtained by minimisation <strong>of</strong> the LSC in a stress interval<br />

0 - 40MPa, is used (see table 2.8). Vf is listed in table 2.7. r is 7µm. The value <strong>of</strong><br />

the frictional interface shear stress τ0 is then about 0.35MPa for laminates UD1<br />

<strong>and</strong> UD2.<br />

The value <strong>of</strong> τ0 as determined here, should be considered <strong>with</strong> caution. The Eglass<br />

fibre reinforcement is made <strong>of</strong> fibre bundles, rather than <strong>of</strong> individual fibres.<br />

τ0 is thus the average frictional interface shear stress <strong>of</strong> the fibre bundle instead <strong>of</strong><br />

the frictional interface shear stress along one single fibre. Figure 2.17 shows an Eglass<br />

fibre reinforcement bundle in IPC specimens in a section, where failure<br />

occurred. In this picture, it can be noticed that the IPC matrix hardly impregnates<br />

the fibre bundle.<br />

Figure 2.17: E-glass fibre bundle in IPC matrix, Gu (2001)<br />

The stress transfer mechanism for fibre bundles varies from that <strong>of</strong> single fibres.<br />

The outer fibres in the bundle are in direct contact <strong>with</strong> the matrix. Stress transfer<br />

41


Chapter 2: IPC composite specimens under monotonic loading<br />

to the inner fibres might be accomplished by matrix-fibre transfer (if some matrix<br />

can penetrate the fibre bundle) <strong>and</strong> fibre-fibre stress transfer (see Bartos <strong>and</strong> Zhu,<br />

1997). If only matrix-fibre stress transfer occurs at the outer surface <strong>of</strong> the bundle,<br />

the ratio <strong>of</strong> the exposed bundle surface area (Aexposed) on the total surface area<br />

(Atotal) is as follows for hexagonally packed bundles (from Li et al., 1990):<br />

Aexposed π 1<br />

= (2.60)<br />

A 2 V N<br />

total<br />

fb<br />

Nfb is the number <strong>of</strong> fibres in one bundle. Vfb is the volume fraction <strong>of</strong> fibres in the<br />

bundle. If the bundle is packed hexagonally, the value <strong>of</strong> Vfb is 0.907 (Li et al.,<br />

1990). The studied UD-weave contains ±1500 fibres per bundle. The ratio<br />

Aexposed/Atotal is then 0.04.<br />

In theory, the ratio <strong>of</strong> the bundle frictional shear stress τ0 on the “single fibre”<br />

frictional shear stress τ0sf equals the ratio Aexposed/Atotal:<br />

τ<br />

τ<br />

0 =<br />

A exposed<br />

A<br />

0sf<br />

total<br />

If the bundle frictional shear stress τ0 is about 0.35MPa, the “single fibre”<br />

frictional shear stress τ0sf is about 8.5MPa.<br />

fb<br />

(2.61)<br />

Since the studied reinforcement works as a bundle <strong>and</strong> not as single fibres, the<br />

value <strong>of</strong> the determined bundle frictional shear stress τ0 is used. This bundle<br />

frictional shear stress can be used, as long as the number <strong>of</strong> filaments in a bundle<br />

is kept constant <strong>and</strong> the fibre packing in a bundle is not changed.<br />

The initial values <strong>of</strong> σR (10.5MP), τ0 (0.35MPa) <strong>and</strong> m (9.3) are first introduced in<br />

the formulation <strong>of</strong> a theoretical stochastic cracking stress-strain curve. The LSC is<br />

calculated from comparison <strong>of</strong> the experimental curve <strong>with</strong> the theoretical<br />

stochastic cracking model (see equation (2.58)). The LSC is determined here in a<br />

stress interval 0 - 20MPa. The LSC is then minimised to obtain a best coincidence<br />

<strong>of</strong> the theoretical stress-strain curve <strong>with</strong> the experimentally obtained curve.<br />

Table 2.9a illustrates the sequence, used to minimise the LSC for the stochastic<br />

cracking model on laminate UD1. First, m is varied until a lower value <strong>of</strong> the LSC<br />

is found (step 1.1 to 1.5 in table 2.9a). With this new value <strong>of</strong> m (m = 2 in table<br />

2.9a), the LSC is further minimised by changing σR (step 1.6 to 1.8 in table 2.9a).<br />

Finally, <strong>with</strong> use <strong>of</strong> the new m <strong>and</strong> σR, the value <strong>of</strong> τ0 is changed until the<br />

minimum value <strong>of</strong> the LSC is obtained (step 1.9 to 1.11 in table 2.9a). Previous<br />

steps represent a first minimisation sequence. The parameters <strong>of</strong> the solution <strong>with</strong><br />

the lowest LSC are now used as initial values in a second minimisation sequence.<br />

Again m is varied first (step 2.1 to 2.2), followed by σR (step 2.3 to step 2.4) <strong>and</strong><br />

finally τ0 (step 2.5 <strong>and</strong> 2.6). It can be seen in table 2.9a that solution 2.3 provides a<br />

42


Chapter 2: IPC composite specimens under monotonic loading<br />

lower value <strong>of</strong> the LSC than step 1.10. The minimisation sequence is thus<br />

performed a third time <strong>and</strong> further if necessary.<br />

The fourth sequence loop reveals no solutions, which lead to a lower value <strong>of</strong> the<br />

LSC than the solution <strong>of</strong> step 3.4. The solution <strong>of</strong> step 3.4 is thus <strong>with</strong>held.<br />

Table 2.9a: illustration <strong>of</strong> the minimisation sequence <strong>of</strong> LSC, stochastic cracking theory versus<br />

experimental stress-strain behaviour, UD1<br />

step m<br />

(-)<br />

σR<br />

(MPa)<br />

τ0<br />

(MPa)<br />

LSC<br />

(exp(-4))<br />

1.1 9 10.5 0.35 numerical problem<br />

1.2 5 10.5 0.35 1.4<br />

1.3 4 10.5 0.35 1.3<br />

1.4 2 10.5 0.35 0.94<br />

1.5 1 10.5 0.35 1.1<br />

1.6 2 10.0 0.35 0.96<br />

1.7 2 11.0 0.35 0.93<br />

1.8 2 11.5 0.35 0.94<br />

1.9 2 11.0 0.30 1.6<br />

1.10 2 11.0 0.40 0.67 (minimal)<br />

1.11 2 11.0 0.45 0.77<br />

1.10 2 11.0 0.40 0.67<br />

2.1 1 11.0 0.40 0.84<br />

2.2 3 11.0 0.40 0.71<br />

2.3 2 10.5 0.40 0.63(minimal)<br />

2.4 2 11.5 0.40 0.74<br />

2.5 2 10.5 0.35 0.94<br />

2.6 2 10.5 0.45 0.67<br />

2.3 2 10.5 0.40 0.63<br />

3.1 3 10.5 0.40 0.77<br />

3.2 1 10.5 0.40 0.72<br />

3.3 2 11.0 0.40 0.67<br />

3.4 2 10.0 0.40 0.58 (minimal)<br />

3.5 2 10.0 0.35 0.96<br />

3.6 2 10.0 0.45 0.59<br />

3.4 2 10.0 0.40 0.58<br />

4.1 3 10.0 0.40 0.71<br />

4.2 1 10.0 0.40 0.68<br />

4.3 2 10.5 0.40 0.72<br />

4.4 2 9.5 0.40 0.61<br />

4.5 2 10.0 0.45 0.59<br />

4.6 2 10.0 0.35 0.96<br />

In table 2.9b the solutions leading to the lowest LSC are listed for UD1 <strong>and</strong> UD2,<br />

together <strong>with</strong> the value <strong>of</strong> the LSC. It should be mentioned that table 2.9a is only<br />

printed here to illustrate the minimisation sequence. The real minimisation<br />

43


Chapter 2: IPC composite specimens under monotonic loading<br />

sequence is conducted <strong>with</strong> higher accuracy on m. The influence <strong>of</strong> the sequence<br />

<strong>of</strong> variation <strong>of</strong> the parameters is also checked for each specimen. It is found that<br />

this sequence does not influence the final results or has very limited influence.<br />

Table 2.9b: material parameters used in the stochastic cracking model, UD-reinforced specimens<br />

plate Vf<br />

(%)<br />

final<br />

(mm)<br />

σR<br />

(MPa)<br />

τ0<br />

(MPa)<br />

m<br />

(-)<br />

LSC<br />

(exp(-5))<br />

UD1 10.4 0.93 10 0.39 2.4 5.8<br />

UD2 7.95 1.4 14 0.36 2.1 6.1<br />

Figure 2.18a to 2.18c show how the shape <strong>of</strong> the theoretical stress-strain curve<br />

changes <strong>with</strong> variation <strong>of</strong> each <strong>of</strong> the studied parameters.<br />

stress (MPa)<br />

µ<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

m = 6<br />

m = 2<br />

experimental<br />

0 0.05 0.1 0.15 0.2 0.25<br />

strain (%)<br />

Figure 2.18a: influence <strong>of</strong> the Weibull<br />

modulus on the shape <strong>of</strong> the stress-strain<br />

curve <strong>of</strong> UD-reinforced IPC<br />

stress (MPa)<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

τ0 = 0.5MPa<br />

stress (MPa)<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

σR = 15MPa<br />

0 0.05 0.1 0.15 0.2 0.25<br />

strain (%)<br />

0 0.05 0.1 0.15 0.2 0.25<br />

strain (%)<br />

experimental<br />

σR = 5MPa<br />

Figure 2.18b: influence <strong>of</strong> the composite<br />

reference cracking stress on the shape <strong>of</strong> the<br />

stress-strain curve <strong>of</strong> UD-reinforced IPC<br />

experimental<br />

τ0 = 0.2MPa<br />

Figure 2.18c: influence <strong>of</strong> the frictional matrix-fibre interface shear stress on the shape <strong>of</strong> the<br />

stress-strain curve <strong>of</strong> UD-reinforced IPC<br />

44


Chapter 2: IPC composite specimens under monotonic loading<br />

Following conclusions can be formulated, according to table 2.9b:<br />

- The estimated value <strong>of</strong> 0.35MPa is indeed a fair “best-fit” value <strong>of</strong> τ0 for the<br />

stochastic cracking model.<br />

- The value <strong>of</strong> the estimated reference cracking stress σR <strong>of</strong> about 10MPa,<br />

obtained from tests on pure IPC, is also a “best fit” value for UD1. The “best fit”<br />

value for UD2 is higher.<br />

- For both panels, the value <strong>of</strong> Weibull modulus m is considerably lower than the<br />

value obtained from three-point bending. This may be due to several phenomena.<br />

Three possible explanations are:<br />

1. The processing technique <strong>of</strong> pure solid IPC <strong>and</strong> IPC used as matrix<br />

material in a composite is different. The fresh IPC mixture is casted into the mould<br />

<strong>and</strong> eventually de-aired when a solid block is prepared. The IPC is poured <strong>and</strong><br />

rolled into the reinforcement, if a laminate is made. Due to these differences, a<br />

broader inherent flaw size distribution could occur in IPC, used in a laminate.<br />

Therefore, the distribution <strong>of</strong> the IPC strength is also broader.<br />

2. The presence <strong>of</strong> fibres might cause non-uniform stresses in the matrix.<br />

This effect is largest when the matrix-fibre bond is strongest. Due to the presence<br />

<strong>of</strong> the fibres, matrix stress concentrations influence the apparent “matrix strength”<br />

distribution. Once a crack appears, a frictional matrix-fibre stress transfer<br />

mechanism replaces the elastic matrix-fibre stress transfer. Since the latter is much<br />

weaker, the matrix stress concentrations disappear or at least decrease <strong>with</strong><br />

advancing multiple cracking.<br />

3. During the curing process, the IPC matrix material shrinks. Free<br />

shrinkage <strong>of</strong> the matrix material is restrained by the presence <strong>of</strong> fibre<br />

reinforcement. This prevented free shrinkage introduces tensile stresses in the<br />

matrix <strong>and</strong> compressive stresses in the fibres. In the early curing process, this<br />

prevented free shrinkage may introduce extra matrix flaws <strong>with</strong> a widespread<br />

variation <strong>of</strong> inherent flaw length in the IPC matrix.<br />

2.4.8.4 ACK model versus stochastic cracking model<br />

The experimental stress-strain curve <strong>of</strong> UD1 is printed in figure 2.19a together<br />

<strong>with</strong> the curves representing the theoretical “best-fit” stress-strain curves <strong>of</strong> the<br />

ACK model <strong>and</strong> the stochastic cracking model. The same curves are printed in<br />

figure 2.19b for specimen UD2.<br />

The least squares coefficient (LSC) has been derived for the ACK model <strong>and</strong> the<br />

stochastic cracking model in the stress interval 0-20MPa. If the ACK theory is<br />

used, the value <strong>of</strong> the LSC is 3.2exp(-4) for UD1. This LSC value becomes<br />

5.8exp(-5), if the stochastic cracking theory is used. LSC is 5.7exp(-4) for the<br />

ACK model <strong>and</strong> 6.1exp(-5) for the stochastic cracking model for laminate UD2.<br />

45


Chapter 2: IPC composite specimens under monotonic loading<br />

The LSC is thus minimised <strong>with</strong> a factor 10 in case the stochastic cracking theory<br />

is used instead <strong>of</strong> the ACK model.<br />

stress (MPa)<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35<br />

strain (%)<br />

Figure 2.19a: experimental <strong>and</strong> theoretical<br />

stress-strain curves, UD1<br />

stress (MPa)<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35<br />

experiment<br />

ACK theory<br />

stochastic cracking theory<br />

strain (%)<br />

Figure 2.19b: experimental <strong>and</strong> theoretical<br />

stress-strain curves, UD2<br />

If one is interested in the stress-strain relationship <strong>of</strong> a UD-reinforced IPC<br />

composite beyond the multiple cracking region, using the ACK theory is more<br />

advantageous, since it is simpler to use than the stochastic cracking theory. The<br />

stochastic cracking theory dem<strong>and</strong>s the determination <strong>of</strong> at least three parameters<br />

not known a priori instead <strong>of</strong> one. The greatest advantage <strong>of</strong> the stochastic<br />

cracking theory - compared to the ACK theory - is found in theoretical “zone II”<br />

<strong>of</strong> the ACK model: the stress region in which multiple cracking occurs. The<br />

stochastic cracking theory provides more accurate results in the multiple cracking<br />

region <strong>of</strong> the stress-strain curve <strong>of</strong> UD-reinforced IPC composites. This is due to<br />

the introduction <strong>of</strong> the stochastic nature <strong>of</strong> the matrix tensile strength in the model.<br />

2.5 Tensile behaviour <strong>of</strong> 2D-r<strong>and</strong>omly reinforced IPC<br />

2.5.1 introduction<br />

The behaviour <strong>of</strong> UD-reinforced specimens under monotonic tensile loading has<br />

been discussed in paragraph 2.4. In this paragraph two models are derived <strong>and</strong><br />

discussed for a more complex type <strong>of</strong> reinforcement. These models are based on<br />

the previous two models for UD-reinforced IPC. These models will be referred to<br />

as “ACK based model” <strong>and</strong> “stochastic cracking based model”. A “ACK based<br />

model” has been derived by Aveston <strong>and</strong> Kelly (1973). Gu et al. (1998), Cuypers<br />

46


Chapter 2: IPC composite specimens under monotonic loading<br />

(1999) <strong>and</strong> Cuypers et al. (2000a) tested <strong>and</strong> discussed a slightly modified ACK<br />

based model on 2D-r<strong>and</strong>omly reinforced IPC specimens. The stochastic cracking<br />

model has been tested on UD-reinforced ceramic matrix micro-composites by<br />

Curtin et al. (1998, 1999). The stochastic cracking based model has been<br />

discussed by Cuypers (2001a) <strong>and</strong> used by Cuypers et al. (2001b) for 2Dr<strong>and</strong>omly<br />

reinforced IPC.<br />

2.5.2 derivation <strong>of</strong> the ACK based model<br />

The behaviour <strong>of</strong> 2D-r<strong>and</strong>omly reinforced IPC specimens is divided in three<br />

zones, similar to the three zones in the ACK theory, used for UD-reinforced<br />

specimens.<br />

2.5.2.1 linear elastic zone or pre-cracking zone (zone I)<br />

The law <strong>of</strong> mixtures is used in this zone to determine the composite stiffness. The<br />

material is assumed to behave linear elastically.<br />

∗<br />

E = E V + E V<br />

c1<br />

f f m m<br />

(2.62)<br />

The definition <strong>of</strong> Vf * equals the one presented in paragraph 2.3, where linear<br />

elastic stress-strain behaviour <strong>of</strong> IPC composites in compression has been<br />

discussed. For the studied type <strong>of</strong> reinforcement (2D-r<strong>and</strong>omly oriented glass fibre<br />

mats, <strong>with</strong> a fibre length <strong>of</strong> 50mm), Vf * is thus:<br />

∗ 1<br />

V f = V f<br />

(2.63)<br />

3<br />

The law <strong>of</strong> mixtures provides a formulation between the matrix normal stress<br />

along the x-axis σm <strong>and</strong> the composite matrix stress σc.<br />

Emσ c σ m = (2.64)<br />

E<br />

where: σc = F x external/Ac<br />

F x external = external applied force, parallel <strong>with</strong> the x-axis<br />

Ac = composite section, transverse to the loading direction<br />

c1<br />

σm = F x matrix/Am<br />

F x matrix = total force parallel <strong>with</strong> the x-axis taken by the matrix<br />

Am = matrix section (Am = VmAc)<br />

It should be mentioned here that σm is an average value <strong>of</strong> the matrix normal stress<br />

in a composite section, which is transverse to the loading direction. The presence<br />

<strong>of</strong> fibre reinforcement introduces matrix stresses in directions, non-parallel to the<br />

loading direction. The total sum or integral <strong>of</strong> the matrix <strong>and</strong> the fibre stress<br />

components transverse to the loading direction is zero, since for each fibre <strong>with</strong><br />

orientation θx from the x-axis, there is a fibre <strong>with</strong> orientation -θx in a 2Dr<strong>and</strong>omly<br />

reinforced composite. The influence <strong>of</strong> local transverse matrix stresses<br />

is neglected in the theoretical models. The end <strong>of</strong> zone I (elastic zone) is reached<br />

47


Chapter 2: IPC composite specimens under monotonic loading<br />

when the normal stress in the matrix along the x-axis σm reaches the matrix tensile<br />

failure stress, σmu.<br />

2.5.2.2 post-cracking zone (zone III)<br />

Once full multiple cracking occurred, only the fibres provide further stiffness. The<br />

composite stiffness in zone III is:<br />

Ec = ηθη l E f V f<br />

(2.65)<br />

where: ηl = fibre length efficiency factor<br />

ηθ = fibre orientation efficiency factor<br />

Assuming matrix-fibre stress transfer is pure frictional, the evolution <strong>of</strong> the normal<br />

stress in a fibre along fibre length lf is illustrated in figure 2.20. The determination<br />

<strong>of</strong> the length efficiency factor is based on this figure. Stresses are transferred from<br />

the matrix to the fibre along a certain length δf.<br />

σf<br />

lf<br />

fibre<br />

σf,cont<br />

fibre stress continuous fibres<br />

fibre stress short fibres<br />

x<br />

Figure 2.20: efficiency <strong>of</strong> fibre reinforcement <strong>with</strong> length lf, normal stress evolution in fibres<br />

From figure 2.20 it can be seen that:<br />

σ f , cont ( l f − 2δ<br />

f ) + 2δ<br />

fσ<br />

f , cont / 2 l f − δ f<br />

σ f = = σ<br />

along l<br />

f , cont<br />

f<br />

l<br />

l<br />

(2.66)<br />

f<br />

f<br />

where: lf = length <strong>of</strong> the fibres<br />

δf = length along which the normal stresses are transferred to the<br />

fibres<br />

σf,cont = stress in fibres, provided they would be continuous<br />

σ<br />

f<br />

alongl f<br />

= average stress in the fibres <strong>with</strong> length lf<br />

The reason why σ f<br />

alongl is a dashed line instead <strong>of</strong> a full line is to indicate that the<br />

f<br />

fibre stress along the fibre is not necessarily constant, but can have an evolution<br />

<strong>with</strong> the x-coordinate. Ratio σ f<br />

alongl /σf,cont represents the ratio <strong>of</strong> the average fibre<br />

f<br />

stress in a short fibre <strong>with</strong> length lf on the average fibre stress in continuous fibres<br />

<strong>and</strong> provides the length efficiency factor <strong>of</strong> a fibre: ηl. The formulation <strong>of</strong> δf is<br />

similar to the formulation <strong>of</strong> δ0, since both definitions formulate the length along<br />

which frictional matrix-fibre stress transfer occurs.<br />

Equation (2.66) can thus be rewritten (<strong>with</strong> help <strong>of</strong> equation 2.13):<br />

48


Chapter 2: IPC composite specimens under monotonic loading<br />

mu m<br />

σ<br />

f<br />

f<br />

along l l f − δ f 2τ 0V<br />

f<br />

ηl<br />

= = =<br />

(2.67)<br />

σ f , max l f l f<br />

If a IPC specimen is considered <strong>with</strong> a fibre volume fraction Vf <strong>of</strong> 10%, fibre<br />

radius r <strong>of</strong> 7µm, τ0 <strong>of</strong> 0.35MPa, a matrix multiple cracking stress σmu <strong>of</strong> 10MPa<br />

<strong>and</strong> fibre length lf <strong>of</strong> 50mm, the value <strong>of</strong> ηθ is 0.98. In this work, the value <strong>of</strong> ηl is<br />

considered to equal 1.<br />

Various authors studied the effect <strong>of</strong> the orientation <strong>of</strong> the reinforcements.<br />

Depending on the assumptions used by the author(s), the value <strong>of</strong> the orientation<br />

efficiency factor ηθ varies from 1/3 to 2/π for 2D-r<strong>and</strong>om reinforcement (see table<br />

2.10).<br />

l<br />

σ<br />

−<br />

Table 2.10: orientation efficiency factor in the post-cracking zone<br />

orientation efficiency factor<br />

2D-r<strong>and</strong>om 3D-r<strong>and</strong>om<br />

Laws (1971) 1/3 1/6<br />

Aveston et al.(1974) 2/π 1/2<br />

Allen (1972) 1/2 -<br />

In this work, the approach used by Laws (1971) is used for the implementation <strong>of</strong><br />

the orientation efficiency factor: ηθ equals 1/3. ηθ = 1/3 means that the fibres<br />

provide 1/3 rd <strong>of</strong> the load bearing capacity in the x-direction they would provide if<br />

they were aligned <strong>with</strong> the external force. Therefore, in the post-cracking zone:<br />

E fV<br />

f<br />

∆σ<br />

c = Ec∆ε c = ∆εc<br />

(2.68)<br />

3<br />

where: ∆σc = composite normal stress variation<br />

∆εc = composite normal strain variation<br />

In the post-cracking zone, the extra force taken by the fibres in the x-direction is<br />

equal to the extra-applied external loading:<br />

x<br />

∆F<br />

fibre<br />

E f V<br />

x<br />

f<br />

= ∆Fexternal<br />

= ∆σ<br />

c Ac<br />

= ∆ε<br />

c Ac<br />

3<br />

E f Af<br />

= ∆ε<br />

f<br />

3<br />

∗<br />

= E f A f ∆ε<br />

f (2.69)<br />

where: F x fibre = sum <strong>of</strong> the x-components <strong>of</strong> the forces, taken by the fibres<br />

Af = fibre section (<strong>with</strong> Af = Vf.Ac)<br />

Af * = equivalent fibre section (Af * = Af/3 if 2D-r<strong>and</strong>omly reinforced)<br />

∆εf = fibre normal strain variation (x-direction)<br />

A “normal fibre stress” σf is defined: σf is the x-component <strong>of</strong> the total fore taken<br />

by the fibres divided by the fibre section:<br />

49<br />

rV


Chapter 2: IPC composite specimens under monotonic loading<br />

∑<br />

P<br />

σ f<br />

x<br />

Ffibre<br />

fibres<br />

= =<br />

Af<br />

Af<br />

x<br />

f<br />

(2.70)<br />

where: P x f is the force in the x-direction taken by one fibre (see figure 2.21)<br />

Or after combination <strong>of</strong> (2.70) <strong>and</strong> (2.69)<br />

∆σ<br />

f<br />

x<br />

∆Ffibre<br />

1<br />

= = E fV<br />

f ∆ε<br />

f<br />

A 3<br />

∗<br />

= E fV<br />

f ∆ε<br />

f<br />

∗<br />

= E fV<br />

f ∆εc<br />

where:<br />

f<br />

V ∗ f<br />

V f =<br />

3<br />

(2.71)<br />

Gu (1998) <strong>and</strong> Cuypers (1999) used the ACK based theory on 2D-r<strong>and</strong>omly<br />

reinforced specimens. Experimentally obtained stress-strain curves showed that<br />

the measured stiffness in the post-cracking zone is generally lower than the<br />

theoretically predicted value. Several factors like imperfect alignment <strong>of</strong> fibres,<br />

warping <strong>of</strong> fibres, bad matrix impregnation, matrix spalling, quality <strong>of</strong> the<br />

debonded matrix-fibre interface, existence <strong>of</strong> local stresses <strong>and</strong> occurrence <strong>of</strong> fibre<br />

pull-out all have their influence on the behaviour in this zone. These effects are<br />

bundled in one efficiency factor K, like as been done for the UD-reinforced<br />

specimens. Equation (2.65) now becomes:<br />

V<br />

∗<br />

f<br />

E = KE V = KE<br />

(2.72)<br />

c<br />

f<br />

f<br />

f<br />

K is simply the ratio between the experimental <strong>and</strong> theoretical stiffness in zone III.<br />

For 2D-r<strong>and</strong>omly reinforced composites, a typical average value for the efficiency<br />

factor K is 0.85, according to Gu (1998) <strong>and</strong> Cuypers (1999). The stress-strain<br />

relationship in the post-cracking zone can thus be expressed as:<br />

zoneIII<br />

zoneII ( σ c −σ<br />

mc )<br />

ε c ( σ c ) = ε c +<br />

∗<br />

(2.73)<br />

KE V<br />

σmc is the composite multiple cracking stress for a 2D-r<strong>and</strong>omly reinforced<br />

composite. σmc can be calculated <strong>with</strong> equation (2.64). However, the strain at the<br />

end <strong>of</strong> the matrix multiple cracking, εc zoneII , is not determined yet. The expression<br />

<strong>of</strong> this strain is discussed below.<br />

2.5.2.3 multiple cracking zone (zone II)<br />

Once σmu is reached <strong>with</strong>in the matrix material, full multiple cracking occurs at a<br />

fixed composite stress level: σmc. Figure 2.21 represents the fibre <strong>and</strong> matrix<br />

normal stress evolution parallel <strong>with</strong> the loading axis in a 2D-r<strong>and</strong>omly reinforced<br />

IPC composite, provided full multiple cracking occurred.<br />

In figure 2.21, the 2D-r<strong>and</strong>om reinforcement is presented schematically by two<br />

fibres <strong>with</strong> angle θx <strong>and</strong> -θx from the x-axis. In each section, transverse to the<br />

50<br />

3<br />

f<br />

f


Chapter 2: IPC composite specimens under monotonic loading<br />

loading direction, the sum <strong>of</strong> the forces taken by the fibres F x fibres <strong>and</strong> the matrix<br />

F x matrix equals the external load F x external on the composite.<br />

Vertical (x-axis) <strong>and</strong> horizontal (y-axis) equilibrium <strong>of</strong> forces is formulated:<br />

x<br />

F<br />

x<br />

= F<br />

x<br />

( x)<br />

+ F ( x)<br />

(2.74)<br />

external<br />

fibres<br />

matrix<br />

y<br />

y<br />

y<br />

Fexternal<br />

= F fibres ( x)<br />

+ Fmatrix<br />

( x)<br />

(2.75)<br />

No external load is applied along the y-axis: F y external = 0. For each fibre <strong>with</strong><br />

orientation θx, there is another fibre <strong>with</strong> orientation -θx, since the reinforcement is<br />

r<strong>and</strong>omly oriented in the xy-plane. Therefore F y fibre(x) = 0 <strong>and</strong> F y matrix(x) = 0.<br />

fibre,<br />

angle -θx<br />

Pf y<br />

Pf<br />

Pf x<br />

fibre,<br />

angle θx<br />

Pf x Pf<br />

f<br />

Pf y<br />

x<br />

δ0<br />

σm,min<br />

σm<br />

σm,max<br />

σmu<br />

σf,min<br />

σf<br />

Pf = force taken by one fibre<br />

σf,max<br />

Figure 2.21: stresses in matrix <strong>and</strong> fibre along a fully cracked composite <strong>with</strong> final average crack<br />

spacing f = 1.337δ0<br />

At this loading stage (zone II), it is assumed that the fibres are stretched, but not<br />

broken. At a crack face, the fibres take all external load:<br />

x<br />

x<br />

F fibres ( x = 0)<br />

= Fexternal<br />

(2.76)<br />

At a distance x from the crack face, forces are transferred from the fibres to the<br />

matrix. The load share taken by the fibres is written in a similar way as for the<br />

post-cracking zone (zone III). Equivalent to equation (2.69) (see Krenchel (1964)):<br />

x<br />

∗<br />

∗ 1<br />

F fibres(<br />

x)<br />

= ηθ εc<br />

( x)<br />

E f Af<br />

= ε f ( x)<br />

E f Af<br />

= εc<br />

( x)<br />

E f Af<br />

= εc<br />

( x)<br />

E f Af<br />

(2.77)<br />

3<br />

51<br />

σ


Chapter 2: IPC composite specimens under monotonic loading<br />

The difference between equation (2.69) <strong>and</strong> (2.77) is found in the fact that the<br />

fibre force <strong>and</strong> strain terms are now function <strong>of</strong> the x-coordinate. The effect <strong>of</strong><br />

orientation <strong>and</strong> fibre length are equal in the expression <strong>of</strong> Af * (or Vf * , since Vf * =<br />

Af * /Ac) for the post-cracking zone <strong>and</strong> the multiple cracking zone.<br />

In each transverse section, transfer <strong>of</strong> forces between fibres <strong>and</strong> matrix along the<br />

x-axis is accomplished by a “frictional shear flux”, T0 x .<br />

2π<br />

⎡<br />

⎤<br />

x<br />

T0 = ∑ ⎢∫τ<br />

0sf<br />

( θ , r)<br />

dθ<br />

cosθ<br />

x ⎥<br />

fibres ⎣ 0<br />

⎦<br />

θ, ρ <strong>and</strong> xl are local cylindrical coordinates,<br />

defined for each fibre (see figure 2.22). The<br />

integral <strong>of</strong> the frictional shear stress around the<br />

fibre perimeter is represented in equation<br />

(2.78), where ρ = r <strong>and</strong> θ is varied between 0<br />

<strong>and</strong> 2π. T0 x is constant <strong>with</strong> x.<br />

z<br />

x<br />

y<br />

fibre<br />

xl<br />

θx<br />

ρ<br />

Figure 2.22: local fibre<br />

cylindrical coordinate axes<br />

(2.78)<br />

If UD-reinforced aligned fibres were used, T0 x would be:<br />

x<br />

T0 = T0<br />

= N fibresτ<br />

02πr<br />

(2.79)<br />

where: Nfibres = number <strong>of</strong> fibres in composite section<br />

If non-unidirectionally aligned fibres are used, an extra orientation efficiency<br />

factor ηθ should be introduced. The value <strong>of</strong> the orientation factor, which should<br />

be added in equation (2.79) for T0 x , equals the one <strong>of</strong> F x fibres in equation (2.77),<br />

since Pf <strong>and</strong> τ0sf are parallel <strong>and</strong> F x fibre <strong>and</strong> T0 x are parallel. Thus, since the<br />

approach <strong>of</strong> Laws is used here:<br />

x 1<br />

T0 = N fibresτ<br />

02πr<br />

(2.80)<br />

3<br />

An “average shear stress component”, τ0 x can now be defined: the total shear flux<br />

T0 x divided by the total fibre perimeter in a composite section equals τ0 x . Thus:<br />

x<br />

x T0<br />

1<br />

τ 0 = = τ 0<br />

(2.81)<br />

2πrN<br />

fibres 3<br />

Equilibrium <strong>of</strong> forces along the x-axis is similar to equation (2.11), except for the<br />

fact that the equilibrium is written in terms <strong>of</strong> forces instead <strong>of</strong> normal stresses:<br />

x x<br />

dF = T dx<br />

(2.82)<br />

or rewritten:<br />

thus:<br />

dF<br />

fibres 0<br />

x x<br />

x<br />

x<br />

= dF − dF = −dF<br />

= T dx<br />

(2.83)<br />

x<br />

fibres total matrix matrix 0<br />

52<br />

θ


Chapter 2: IPC composite specimens under monotonic loading<br />

τ 0<br />

Amdσ m = 2πrN<br />

fibresdx<br />

(2.84)<br />

3<br />

<strong>and</strong> since Af = r²πNfibres:<br />

∗<br />

2τ<br />

2τ<br />

V<br />

0<br />

0 f<br />

dσ<br />

m = Af<br />

dx = dx<br />

(2.85)<br />

3rAm<br />

rVm<br />

When x = 0, σm(x) = 0 <strong>and</strong> for x = δ0, σm(x) = σmu (see figure 2.21). Thus after<br />

integration <strong>and</strong> rearranging <strong>of</strong> equation (2.85):<br />

rVm δ 0 = σ ∗ mu<br />

2τ<br />

V<br />

(2.86)<br />

0<br />

f<br />

Equation (2.86) is equivalent to equation (2.13), except for the fibre volume<br />

fraction Vf, which is replaced <strong>with</strong> an equivalent fibre volume fraction Vf * .<br />

Further derivation <strong>of</strong> the stress-strain relationship is equivalent to equation (2.18)<br />

to (2.24) for UD-reinforced specimens, except that Vf is replaced <strong>with</strong> Vf * .<br />

Finally, the composite strain at the end <strong>of</strong> multiple cracking is for 2D-r<strong>and</strong>omly<br />

reinforced specimens:<br />

zoneII<br />

σ mu<br />

σ mc<br />

ε c = ( 1+<br />

0.<br />

666α<br />

) εmu<br />

= ( 1+<br />

0.<br />

666α<br />

) = ( 1+<br />

0.<br />

666α<br />

)<br />

Em<br />

Ec1<br />

EmV<br />

(2.87)<br />

m<br />

<strong>with</strong> α = ∗<br />

E V<br />

f<br />

f<br />

2.5.3 derivation <strong>of</strong> the stochastic cracking based model<br />

The derivation <strong>of</strong> the stochastic cracking based theory for 2D-r<strong>and</strong>omly reinforced<br />

laminates is formulated from combination <strong>of</strong> the ACK based theory for 2Dr<strong>and</strong>omly<br />

reinforced laminates <strong>and</strong> the stochastic cracking theory on UDreinforced<br />

composite specimens. Finally, the stress-strain relationship found for<br />

UD-reinforced laminates is used on 2D-r<strong>and</strong>omly reinforced laminates, provided<br />

Vf is replaced <strong>with</strong> an equivalent fibre volume fraction Vf * . Details on the<br />

mathematics are given by Cuypers (2001a).<br />

2.5.3.1 formulation when > 2δ0<br />

⎛ ⎞<br />

c ⎜ 0 ⎟<br />

c = 1+ E ⎜ ⎟<br />

c1<br />

⎝ cs ⎠<br />

αδ σ<br />

EmVm<br />

ε <strong>with</strong> α = ∗<br />

(2.88)<br />

E fV<br />

f<br />

2.5.3.2 formulation when < 2δ0<br />

⎛<br />

⎞<br />

⎜<br />

1 α cs<br />

⎟<br />

EmVm<br />

ε c = σ c −<br />

<strong>with</strong> α =<br />

⎜ ∗ ⎟<br />

∗<br />

(2.89)<br />

⎝ E fV<br />

f 4δ<br />

0Ec1<br />

⎠<br />

E fV<br />

f<br />

53


Chapter 2: IPC composite specimens under monotonic loading<br />

2.5.4 experiments on 2D-r<strong>and</strong>omly reinforced IPC specimens<br />

2.5.4.1 specimens<br />

One 2D-r<strong>and</strong>omly reinforced plate is made (R1), according to the directions <strong>of</strong><br />

paragraph 2.2. Six layers <strong>of</strong> E-glass fibre reinforcement are used. Specimens <strong>with</strong><br />

a length <strong>of</strong> 250mm <strong>and</strong> a width <strong>of</strong> 25mm are cut from the plate. The properties <strong>of</strong><br />

the laminate are found in table 2.11. Three specimens <strong>of</strong> this plate are subjected to<br />

tensile testing.<br />

Table 2.11: properties <strong>of</strong> 2D-r<strong>and</strong>omly reinforced test plates<br />

specimen reinforcement matrix amount thickness Vf<br />

name<br />

(g/m²/layer) (mm)<br />

(%)<br />

R1-1 2D-r<strong>and</strong>om 1800 3.85 12.3<br />

R1-2 2D-r<strong>and</strong>om 1800 4.20 11.3<br />

R1-3 2D-r<strong>and</strong>om 1800 3.95 12.0<br />

Figure 2.23a shows the stress-strain diagrams, obtained from tensile testing <strong>of</strong> the<br />

three specimens. Figure 2.23b shows the same stress-strain curves <strong>with</strong>in a stress<br />

interval <strong>of</strong> 0-20MPa. The experimental stress-strain curves obtained on the 2Dr<strong>and</strong>omly<br />

reinforced specimens are compared <strong>with</strong> theoretical models: the ACK<br />

based <strong>and</strong> the stochastic cracking based model. The determination <strong>of</strong> the material<br />

parameters, not necessarily known a priori, is performed in a similar way as has<br />

been done for UD-reinforced specimens (see paragraph 2.4.8.2 <strong>and</strong> 2.4.8.3).<br />

stress (MPa)<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

R1-1<br />

R1-3<br />

R1-2<br />

0<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0<br />

strain (%)<br />

Figure 2.23a: stress-strain curves <strong>of</strong> 2Dr<strong>and</strong>omly<br />

reinforced specimens<br />

stress (MPa)<br />

20<br />

15<br />

10<br />

5<br />

0<br />

R1-1<br />

0.0 0.2 0.4 0.6 0.8<br />

strain (%)<br />

R1-3<br />

R1-2<br />

Figure 2.23b: stress-strain curves <strong>of</strong> 2Dr<strong>and</strong>omly<br />

reinforced specimens, 0 to 20MPa<br />

2.5.4.2 ACK based model versus experimental results<br />

The theoretical <strong>and</strong> experimental composite stiffness are first determined in the<br />

post-cracking zone (zone III) for all specimens. Experimental stiffness Ec3 is the<br />

slope <strong>of</strong> the curve in the stress-strain diagram between 15 <strong>and</strong> 30MPa. The<br />

theoretical stiffness in zone III, Ec, is defined by equation (2.72). The value <strong>of</strong> K<br />

from equation (2.72) is thus determined as:<br />

54


Chapter 2: IPC composite specimens under monotonic loading<br />

K<br />

E<br />

E V<br />

c3<br />

c3<br />

= = ∗ (2.90)<br />

f<br />

f<br />

3E<br />

E V<br />

f<br />

The value <strong>of</strong> K is determined by equation (2.90) for all specimens <strong>and</strong> is listed in<br />

table 2.12. The value <strong>of</strong> matrix stiffness Em is about 18MPa according to table 2.2.<br />

The determination <strong>of</strong> the matrix failure stress σmu is done in a similar way as for<br />

UD-reinforced specimens earlier. σmu is varied until the difference between the<br />

theoretical <strong>and</strong> experimental curves is minimised, using the least squares method.<br />

The definition <strong>of</strong> the least squares coefficient (LSC) is expressed in equation<br />

(2.58). The stress interval, in which the “best-fit” matrix failure stress σmu is<br />

determined, is 0-20MPa.<br />

Table 2.12 lists the values <strong>of</strong> model parameter σmu, obtained from minimisation <strong>of</strong><br />

the LSC. The LSC is also listed in this table.<br />

Table 2.12: parameter σmu, obtained from least squares method<br />

R1-1 R1-2 R1-3<br />

EfVf * (GPa) 2.21 2.03 2.17<br />

Ec3 (GPa) 1.93 1.69 1.89<br />

K (-) 0.87 0.83 0.87<br />

σmu (MPa) 11.1 11.2 11.5<br />

LSC (exp(-3)) 9.3 9.5 11<br />

2.5.4.3 stochastic cracking based model versus experimental results<br />

In the stochastic cracking based theory, three material parameters are varied to<br />

obtain a “best fit” <strong>of</strong> the experimental <strong>and</strong> theoretical stress-strain curves. These<br />

three material parameters are not always known accurately a priori: the reference<br />

cracking stress σR, the Weibull modulus m <strong>and</strong> the frictional matrix-fibre interface<br />

shear stress τ0. The frictional matrix-fibre interface shear stress τ0 should be<br />

considered as a bundle shear stress, rather than a single fibre shear stress. The<br />

number <strong>of</strong> fibres per bundle is ±1500 for the UD weave <strong>and</strong> ±540 for the 2Dr<strong>and</strong>om<br />

chopped mat. Equation (2.60) <strong>and</strong> (2.61) provide the relation between τ0<br />

<strong>and</strong> τ0sf for a bundle <strong>with</strong> hexagonal packing. The experimentally obtained value<br />

<strong>of</strong> τ0 is about 0.40MPa (table 2.9) for the UD-reinforced specimens <strong>with</strong><br />

±1500fibres/bundle. The related value <strong>of</strong> τ0sf is then about 10.5MPa. If<br />

±540fibres/bundle are packed hexagonally, the predicted value <strong>of</strong> τ0 <strong>of</strong> this bundle<br />

is about 0.80MPa.<br />

The average final crack spacing f is determined under the stereomicroscope,<br />

after the cracks are visualised <strong>with</strong> an ink solution, <strong>and</strong> is about 0.9mm.<br />

f<br />

55


Chapter 2: IPC composite specimens under monotonic loading<br />

The initial values <strong>of</strong> the material parameters used in the first sequence <strong>of</strong> “best<br />

fitting” are thus m = 2, σR = 12MPa <strong>and</strong> τ0 = 0.80MPa, in accordance <strong>with</strong> the<br />

results obtained on the UD-reinforced specimens (table 2.9). The determination <strong>of</strong><br />

the value <strong>of</strong> these material parameters, such that a minimum LSC is found, is<br />

illustrated in table 2.13 for specimen R1-1.<br />

It should be mentioned that table 2.13 is only listed here to illustrate the<br />

minimisation sequence. The real minimisation sequence is performed <strong>with</strong> higher<br />

accuracy on the parameters. The influence <strong>of</strong> the sequence <strong>of</strong> variation <strong>of</strong> the<br />

parameters is also checked for each specimen. It is found that this sequence does<br />

not influence the final results or has very limited influence.<br />

Table 2.13: illustration <strong>of</strong> the minimisation sequence <strong>of</strong> LSC, stochastic cracking based theory<br />

versus experimental stress-strain behaviour, specimen R1-1<br />

step m<br />

(-)<br />

σR<br />

(MPa)<br />

τ0<br />

(MPa)<br />

LSC<br />

(exp(-4))<br />

1.1 2 12.0 0.80 5.2<br />

1.2 3 12.0 0.80 1.7<br />

1.3 4 12.0 0.80 0.65 (minimal)<br />

1.4 5 12.0 0.80 1.2<br />

1.5 4 12.5 0.80 1.7<br />

1.6 2 11.5 0.80 1.3<br />

1.7 2 12.0 0.75 0.72<br />

1.8 2 12.0 0.85 1.2<br />

In table 2.14 the solutions leading to the lowest LSC are listed for specimens R1-1<br />

to R1-3, together <strong>with</strong> the value <strong>of</strong> the LSC.<br />

Table 2.14: material parameters in stochastic cracking model, 2D-r<strong>and</strong>omly reinforced specimens<br />

plate Vf<br />

(%)<br />

f<br />

(mm)<br />

σR<br />

(MPa)<br />

τ0<br />

(MPa)<br />

m<br />

(-)<br />

LSC<br />

(exp(-5))<br />

R1-1 12.3 0.9 12.0 0.81 4.4 6.5<br />

R1-2 11.3 0.9 12.0 0.80 5.7 14.4<br />

R1-3 12.0 0.9 12.5 0.78 5.3 8.4<br />

2.5.4.4 ACK based model versus stochastic cracking based model<br />

The LSC has been derived for the ACK based model <strong>and</strong> the stochastic cracking<br />

based model in the stress interval 0-20MPa on 2D-r<strong>and</strong>omly reinforced IPC<br />

specimens. If the ACK theory is used, the average value <strong>of</strong> the LSC is about<br />

10exp(-3). If the stochastic cracking theory is used, the average value <strong>of</strong> the LSC<br />

is about 10exp(-5). The LSC is thus minimised <strong>with</strong> a factor 100 if the stochastic<br />

cracking theory is used instead <strong>of</strong> the ACK model.<br />

56


Chapter 2: IPC composite specimens under monotonic loading<br />

2.6 Tensile behaviour <strong>of</strong> IPC composite specimens:<br />

discussion <strong>and</strong> conclusions<br />

Some conclusions can be formulated on the stress-strain behaviour <strong>of</strong> UDreinforced<br />

<strong>and</strong> 2D-r<strong>and</strong>omly reinforced IPC composite specimens in tension:<br />

- The ACK (Aveston-Cooper-Kelly) model provides a rather good model for<br />

predicting the tensile stress-strain behaviour <strong>of</strong> both unidirectionally <strong>and</strong> 2Dr<strong>and</strong>omly<br />

E-glass fibre reinforced IPC composites in zone I (linear elastic zone)<br />

<strong>and</strong> zone III (post-cracking zone).<br />

- The major shortcoming <strong>of</strong> the ACK model - when applied to E-glass fibre<br />

reinforced IPC - is that the tensile matrix strength is assumed to be unique in the<br />

whole composite. The true underlying existence <strong>of</strong> matrix flaws <strong>with</strong> various<br />

lengths is thus hidden in this model. The information about true gradual multiple<br />

cracking in a “multiple cracking zone” is compressed into one “theoretical<br />

multiple cracking stress”.<br />

- Several probability distribution models on the IPC matrix strength are compared<br />

<strong>with</strong> results from three-point bending on pure IPC specimens. From comparison <strong>of</strong><br />

theoretical distribution functions <strong>with</strong> these experimental results, it is concluded<br />

that a Weibull distribution model describes the nature <strong>of</strong> the probability<br />

distribution <strong>of</strong> the tensile matrix failure stress appropriately.<br />

- The stochastic cracking model combines the basic ideas <strong>of</strong> the ACK theory <strong>with</strong><br />

a Weibull distribution <strong>of</strong> the tensile matrix strength. Comparison <strong>of</strong> experimental<br />

results <strong>with</strong> theory shows that this stochastic cracking model has the capability <strong>of</strong><br />

providing very accurate prediction <strong>of</strong> the stress-strain composite behaviour.<br />

Whereas the ACK fails in predicting the composite strains in zone II (multiple<br />

cracking zone) accurately, the stochastic cracking model provides a useful tool in<br />

refining the theoretical stress-strain curve in the multiple cracking zone. This is<br />

verified on UD-reinforced <strong>and</strong> 2D-r<strong>and</strong>omly reinforced IPC composite specimens.<br />

- The stochastic cracking model needs input <strong>of</strong> several parameters, which are not<br />

easily known a priori. The most delicate parameters are the Weibull modulus, the<br />

Weibull reference cracking stress <strong>and</strong> the matrix-fibre interface shear stress.<br />

- From three-point bending tests, performed on pure IPC matrix blocks, a<br />

reference composite cracking stress <strong>of</strong> about 11MPa is determined. It is verified<br />

that this value is a fair estimation <strong>of</strong> the reference cracking stress for UDreinforced<br />

<strong>and</strong> 2D-r<strong>and</strong>omly reinforced IPC specimens.<br />

57


Chapter 2: IPC composite specimens under monotonic loading<br />

- The value <strong>of</strong> the matrix-fibre interface shear stress can be determined from visual<br />

determination <strong>of</strong> the final crack spacing (e.g. under a microscope). It has to be<br />

mentioned that the measured frictional interface shear stress is an average shear<br />

stress for a fibre bundle, rather than a single fibre shear stress. The average bundle<br />

shear stress τ0 can be estimated from knowledge <strong>of</strong> the single fibre shear stress<br />

τ0sf, the number <strong>of</strong> fibres per bundle <strong>and</strong> the fibre packing. The value <strong>of</strong> the single<br />

fibre shear stress τ0sf obtained here is about 8.5MPa. This value is obtained from<br />

experiments on UD-reinforced <strong>and</strong> on 2D-r<strong>and</strong>omly reinforced specimens.<br />

- From three-point bending tests, performed on pure IPC matrix blocks, a Weibull<br />

modulus <strong>of</strong> 9 is determined. For all IPC composite panels, the value <strong>of</strong> the Weibull<br />

modulus m is considerably lower than the value obtained from three-point bending<br />

on pure IPC. The value <strong>of</strong> the Weibull modulus, leading to a “best fit” <strong>of</strong> the<br />

stochastic cracking model, is about 5 for the 2D-r<strong>and</strong>omly reinforced specimens<br />

<strong>and</strong> 2 for the UD-reinforced specimens. Three possible explanations are:<br />

1. The processing technique <strong>of</strong> pure solid IPC <strong>and</strong> IPC used as matrix in a<br />

composite is different. Due to these differences, a broader inherent flaw size<br />

distribution may occur in IPC used in a laminate.<br />

2. The presence <strong>of</strong> fibres cause non-uniform stresses in the matrix. This<br />

effect is largest when the matrix-fibre bond is strongest. Due to the presence <strong>of</strong> the<br />

fibres, matrix stress concentrations influence the apparent “matrix strength”<br />

distribution.<br />

3. Prevented shrinkage <strong>of</strong> the IPC matrix during the curing process<br />

introduces tensile stresses in the matrix <strong>and</strong> compressive stresses in the fibres. In<br />

the early curing process, this prevented free shrinkage might introduce extra<br />

matrix flaws, <strong>with</strong> a widespread variation <strong>of</strong> inherent flaw length, in the IPC<br />

matrix.<br />

2.7 Evolution <strong>of</strong> matrix stresses transverse to the loading<br />

direction<br />

The behaviour <strong>of</strong> a IPC composite specimen in a direction transversal to the<br />

loading axis is discussed in this paragraph. A theoretical evolution <strong>of</strong> the<br />

composite Poisson’s ratio as a function <strong>of</strong> the composite stress is derived. This<br />

theoretical evolution is compared <strong>with</strong> experimental results. The results <strong>and</strong><br />

conclusions derived here are used in Chapter 5. In paragraph 5.2, it is explained<br />

why only the ACK based theoretical derivation <strong>of</strong> the evolution <strong>of</strong> the composite<br />

Poisson’s ratio is derived here.<br />

58


Chapter 2: IPC composite specimens under monotonic loading<br />

2.7.1 theoretical derivation <strong>of</strong> the Poisson’s ratio based on the ACK<br />

theory<br />

The average ratio <strong>of</strong> the transversal strain on the longitudinal strain along the<br />

whole length <strong>of</strong> composite Rav is formulated by equation (2.91). A formulation <strong>of</strong><br />

Rav is obtained here, using the ACK theory for UD-reinforced composite<br />

specimens.<br />

Rav<br />

ε<br />

c,<br />

z<br />

= −<br />

(2.91)<br />

ε<br />

c,<br />

x<br />

Figure 2.24 shows the evolution <strong>of</strong> the transverse matrix strain εm,z(x) along the<br />

length <strong>of</strong> a UD-reinforced IPC composite specimen, according to the ACK theory.<br />

In the ACK theory, multiple cracking (zone II) occurs at a fixed stress level. From<br />

then on, matrix parts <strong>with</strong> average length f slide along the fibre reinforcement<br />

if extra loading is applied. Theoretically, normal stress σm,x(x) in the matrix does<br />

not change <strong>with</strong> increasing load in zone III (post-cracking zone). Consequently,<br />

matrix strain εm,x(x) does not vary any more. If the Poisson’s ratio <strong>of</strong> the IPC<br />

matrix is invariable, the transverse strain εm,z(x) remains constant <strong>with</strong> increasing<br />

load in zone III.<br />

x<br />

y<br />

f<br />

f<br />

f<br />

f<br />

no load introduction <strong>of</strong> cracking<br />

(zone II)<br />

x<br />

εm,z<br />

f<br />

f<br />

x<br />

εm,z<br />

beyond multiple cracking stress<br />

(zone III)<br />

Figure 2.24: evolution <strong>of</strong> transverse strains in the matrix for UD-reinforced specimens<br />

Following determination <strong>of</strong> Rav is appropriate for a unidirectionally reinforced<br />

specimen.<br />

59


Chapter 2: IPC composite specimens under monotonic loading<br />

Before multiple cracking occurred (zone I), the law <strong>of</strong> mixtures can be used to<br />

express the Poisson’s ratio <strong>of</strong> a transversely isotropic composite (see e.g. Chawla,<br />

1987)<br />

Rav = ν xz = ν fxzV<br />

f + ν mVm<br />

(2.92)<br />

where: νxz = Poisson’s ratio <strong>of</strong> an elastic composite<br />

νfxz = Poisson’s ratio <strong>of</strong> the fibres<br />

νm = Poisson’s ratio <strong>of</strong> the matrix<br />

In zone I, Rav is thus independent from the composite stress.<br />

Once multiple cracking occurred (zone II <strong>and</strong> zone III), f is the average length<br />

between two matrix cracks. In zone II <strong>and</strong> III, the longitudinal composite strain is<br />

formulated by equation (2.93), equal to equation (2.26):<br />

( )<br />

( σ c −σ<br />

mc )<br />

εc,<br />

x = εmu<br />

1+<br />

0.<br />

666α<br />

+<br />

E V<br />

(2.93)<br />

The average transversal composite strain as measured by a strain gauge is mainly<br />

determined by the value <strong>of</strong> the matrix transversal strain. Once full multiple<br />

cracking occurred, the transversal matrix strain remains constant, since the<br />

longitudinal matrix strain remains constant. The formulation <strong>of</strong> the transversal<br />

composite strain in zone III is thus:<br />

εc , z ≈ ε m,<br />

z = −ν<br />

m ε m,<br />

x<br />

(2.94)<br />

along cs<br />

f<br />

f<br />

f<br />

along cs<br />

f<br />

<strong>with</strong>:<br />

ε m,<br />

x along cs<br />

f<br />

1 1.<br />

337<br />

= ε mu<br />

2 2<br />

(2.95)<br />

The composite transversal strain is thus<br />

1 1.<br />

337<br />

εc,<br />

z ≈ − ν m ε mu<br />

2 2<br />

The formulation <strong>of</strong> Rav is:<br />

(2.96)<br />

1<br />

ε<br />

ν m1.<br />

337ε<br />

mu<br />

c,<br />

z<br />

R = − ≈ 4<br />

av<br />

εc,<br />

x<br />

⎛ ⎞<br />

( ) ⎜<br />

σ c − σ mc<br />

ε + + ⎟<br />

mu 1 0.<br />

666α<br />

⎜ ⎟<br />

⎝ E fV<br />

f ⎠<br />

(2.97)<br />

Equation (2.97) represents the evolution <strong>of</strong> the Poisson’s ratio <strong>of</strong> UD-reinforced<br />

IPC in zone III.<br />

The evolution <strong>of</strong> Rav can be measured experimentally on UD-reinforced composite<br />

specimens under monotonic tensile loading. Next paragraph discusses the<br />

comparison <strong>of</strong> experimental <strong>and</strong> theoretical curves <strong>of</strong> the evolution <strong>of</strong> the<br />

Poisson’s ratio as a function <strong>of</strong> the composite stress.<br />

60


Chapter 2: IPC composite specimens under monotonic loading<br />

2.7.2 experimental verification<br />

One four-layer UD-reinforced <strong>and</strong> one four-layer 2D-r<strong>and</strong>omly reinforced IPC<br />

laminate are made. The fibre volume fraction is 12% for the UD-reinforced<br />

laminate <strong>and</strong> 11% for the 2D-r<strong>and</strong>omly reinforced laminate. The dimensions <strong>of</strong> the<br />

UD-reinforced <strong>and</strong> 2D-r<strong>and</strong>omly reinforced specimens are: 25x250mm². All<br />

specimens are loaded by a INSTRON 4505 until failure occurs. The INSTRON<br />

load cell measures the loads. Strain gauges are attached to both sides <strong>of</strong> each<br />

specimen. At each side, one strain gauge is aligned in the longitudinal direction<br />

<strong>and</strong> one is aligned in the transversal direction (as shown in figure 2.25). The length<br />

<strong>of</strong> the strain gauges is 15mm <strong>and</strong> the width is 5mm. The strain gauges are applied<br />

on both sides <strong>of</strong> the specimens to monitor<br />

possible unsymmetrical behaviour <strong>of</strong> the<br />

specimens (e.g. the IPC matrix not spread<br />

homogeneously, misalignment <strong>of</strong> specimen).<br />

According to the ACK theory, two events<br />

should be distinguished in the experimental<br />

evolution <strong>of</strong> Rav:<br />

1. Once the multiple cracking stress is<br />

reached, Rav suddenly drops from a Poisson’s<br />

ratio <strong>of</strong> the virgin composite, predicted by the<br />

rule <strong>of</strong> mixtures, to a much lower value, which<br />

can be predicted by equation (2.97), <strong>with</strong> σc =<br />

σmc.<br />

2. From then on, extra loading <strong>of</strong> the<br />

specimen will lead to a further decreasing value<br />

<strong>of</strong> Rav. Finally, the ratio will reach zero.<br />

cracked<br />

IPC<br />

specimen<br />

transversal<br />

strain gauge<br />

longitudinal<br />

strain gauge<br />

F<br />

F<br />

Figure 2.25: test set-up<br />

Figures 2.26 <strong>and</strong> 2.27 show the evolution <strong>of</strong> the<br />

ratio <strong>of</strong> the transverse composite strain on the longitudinal composite strain as a<br />

function <strong>of</strong> the applied tensile stress for a UD-reinforced <strong>and</strong> a 2D-r<strong>and</strong>omly<br />

reinforced specimen respectively. The Poisson’s ratio <strong>of</strong> the matrix νm is<br />

determined experimentally from determination <strong>of</strong> the natural frequencies <strong>of</strong> small<br />

beam <strong>and</strong> plate specimens <strong>and</strong> is about 0.35. νfxz is found in literature <strong>and</strong> is about<br />

0.35. σmu is determined by minimisation <strong>of</strong> the LSC (see paragraph 2.4.8.2 <strong>and</strong><br />

2.5.4.2).<br />

Figure 2.26 shows discrepancy between the theoretical <strong>and</strong> experimental evolution<br />

<strong>of</strong> the Poisson’s ratio. As was already mentioned earlier, multiple cracking does<br />

not occur at one stress level, according to the stochastic cracking theory, but in a<br />

stress interval. The experimental drop <strong>of</strong> Poisson’s ratio is thus not as dramatically<br />

as predicted by the ACK theory, since multiple cracking is spread over a stress<br />

interval.<br />

61


Chapter 2: IPC composite specimens under monotonic loading<br />

R av (-)<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

theory<br />

experiment<br />

0 20 40 60 80 100 120<br />

σ c (MPa)<br />

Figure 2.26: evolution <strong>of</strong> the ratio <strong>of</strong> transversal strain on longitudinal strain, UD-reinforced IPC<br />

R av (-)<br />

0.45<br />

0.4<br />

0.35<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

theory<br />

experiment<br />

0 10 20 30<br />

σ c (MPa)<br />

Figure 2.27: evolution <strong>of</strong> the ratio <strong>of</strong> transversal strain on longitudinal strain, 2D-r<strong>and</strong>omly<br />

reinforced IPC<br />

For the theoretical prediction <strong>of</strong> the evolution <strong>of</strong> the Poisson’s ratio for 2Dr<strong>and</strong>omly<br />

reinforced specimens, equation (2.97) is slightly modified. Vf has been<br />

replaced by Vf * , as has been done earlier. Figure 2.27 shows that the transverse<br />

behaviour <strong>of</strong> a 2D-r<strong>and</strong>omly reinforced specimen is more complex than<br />

formulated in expressions (2.93) to (2.97). The Poisson’s ratio decreases notably<br />

<strong>with</strong> increasing stress, but remain positive, thus different from zero. For 2Dr<strong>and</strong>omly<br />

reinforced specimens, the measured transversal strains are not<br />

determined by the Poisson’s effect in the matrix only, but also by transfer <strong>of</strong><br />

stresses from the fibres to the matrix in directions, non-parallel to the external<br />

loading axis. However, a decreasing trend <strong>of</strong> the measured ratio <strong>of</strong> the transverse<br />

strains on the longitudinal strains due to multiple cracking is still experienced.<br />

Conclusively, the Poisson’s ratio decreases considerably <strong>with</strong> increasing stress.<br />

The value <strong>of</strong> the Poisson’s ratio reaches zero for UD-reinforced specimens <strong>and</strong> a<br />

relatively low value for 2D-r<strong>and</strong>omly reinforced specimens.<br />

62


Chapter 2: IPC composite specimens under monotonic loading<br />

Another important conclusion, which can be formulated from theoretical <strong>and</strong><br />

experimental observations here <strong>and</strong> which will be used in Chapter 5, is that the<br />

transverse matrix stresses (perpendicular to loading axis) never become very high<br />

in a IPC composite specimen, which is loaded in the x-direction. This is the case<br />

for s<strong>and</strong>wich beam specimens, but also for s<strong>and</strong>wich wide-beam specimen (plate<br />

loaded in one direction).<br />

2.8 Experimental determination <strong>of</strong> the shear modulus <strong>of</strong> IPC<br />

composite specimens<br />

The aim <strong>of</strong> this paragraph is to obtain an order <strong>of</strong> magnitude <strong>of</strong> the shear modulus<br />

<strong>of</strong> E-glass fibre reinforced IPC. No model is presented for the description <strong>of</strong> the<br />

stress-strain behaviour in shear. The shear modulus is determined from tests on a<br />

torsion bench. The value <strong>of</strong> the shear modulus, to be implemented into finite<br />

element calculations later, is determined here <strong>with</strong>out further micro-mechanical or<br />

meso-mechanical based theory.<br />

2.8.1 specimens <strong>and</strong> test set-up<br />

One 2D-r<strong>and</strong>omly (IPCS1) <strong>and</strong> one UD-reinforced (IPCS2) plate are made. Four<br />

layers <strong>of</strong> glass fibre are used as reinforcement in both plates. The average fibre<br />

volume fraction is 11% for the UD-reinforced plate <strong>and</strong> 10% for the 2D-r<strong>and</strong>omly<br />

reinforced plate. Specimens <strong>of</strong> 20x200mm² are cut from these plates. The thin<br />

laminates are clamped in a torsion bench, which has one fixed clamp <strong>and</strong> one<br />

clamp that rotates <strong>with</strong> a user-defined angular rotation velocity. A load cell<br />

measures the torsion moment. The torsion angle is also monitored. From the<br />

torsion moment-torsion angle curve, the shear modulus is determined.<br />

Six specimens <strong>of</strong> the 2D-r<strong>and</strong>omly reinforced laminate are tested. Three<br />

specimens are clamped <strong>and</strong> tested in the torsion bench after preparation <strong>and</strong><br />

cutting: IPCS1-1 to IPCS1-3. Three specimens are first subjected to a tensile load:<br />

IPCS1-4 to IPCS1-6. Specimens IPCS1-4 to IPCS1-6 are loaded up to 20MPa in<br />

tension <strong>and</strong> are then again unloaded. Whereas specimens IPCS1-1 to IPCS1-3 are<br />

virgin specimens (pre-cracked state) at the start <strong>of</strong> the torsion test, specimens<br />

IPCS1-4 to IPCS1-6 did already undergo matrix multiple cracking, prior to torsion<br />

testing (post-cracked state).<br />

Three specimens are cut from the UD-reinforced laminate, <strong>with</strong> their length axis<br />

parallel <strong>with</strong> the main fibre reinforcement: IPCS2-1 to IPCS2-3. Three specimens<br />

are cut from the UD-reinforced plate, <strong>with</strong> their length axis transverse to the main<br />

fibre reinforcement: IPCS2-4 to IPCS2-6. Specimens IPCS2-1 to IPCS2-6 are<br />

subjected to torsion testing in virgin state (pre-cracked state). Three specimens are<br />

63


Chapter 2: IPC composite specimens under monotonic loading<br />

cut from the UD-reinforced laminate, length axis parallel to the main<br />

reinforcement, <strong>and</strong> are loaded in tension up to 20MPa prior to torsion testing:<br />

IPCS2-7 to IPCS2-9. Specimens IPCS2-7 to IPCS2-9 are thus in post-cracked<br />

state at the beginning <strong>of</strong> the torsion test.<br />

2.8.2 results<br />

In tables 2.15 to 2.19, the experimentally obtained values <strong>of</strong> the shear modulus are<br />

listed.<br />

Table 2.15: evolution <strong>of</strong> the shear modulus, 2D-r<strong>and</strong>omly reinforced virgin specimens<br />

shear modulus (GPa)<br />

range angle 1-2° 1-3° 1-4° 1-8°<br />

IPCS1-1 4.4 4.0 3.7 2.9<br />

IPCS1-2 4.1 3.7 3.4 2.6<br />

IPCS1-3 3.7 3.5 3.2 2.5<br />

average 4.1 3.7 3.4 2.7<br />

Table 2.16: evolution <strong>of</strong> the shear modulus, 2D-r<strong>and</strong>omly reinforced multiple cracked specimens<br />

shear modulus (GPa)<br />

range angle 1-2° 1-3° 1-4° 1-8°<br />

IPCS1-4 2.9 2.9 2.7 2.4<br />

IPCS1-5 2.7 2.6 2.5 2.3<br />

IPCS1-6 2.8 2.8 2.6 2.4<br />

average 2.8 2.8 2.6 2.4<br />

Table 2.17: evolution <strong>of</strong> the shear modulus, UD-reinforced virgin specimens<br />

shear modulus (GPa)<br />

range angle 1-2° 1-3° 1-4° 1-8°<br />

IPCS1-1 6.4 6.2 6.2 5.8<br />

IPCS1-2 5.5 5.4 5.4 5.1<br />

IPCS1-3 6.9 6.5 6.4 5.9<br />

average 6.3 6.0 6.0 5.6<br />

Table 2.18: evolution <strong>of</strong> the shear modulus, UD-reinforced virgin specimens, loaded transverse to<br />

reinforcement direction<br />

shear modulus (GPa)<br />

range angle 1-2° 1-3° 1-4° 1-8°<br />

IPCS1-4 6.0 6.1 5.8 5.2<br />

IPCS1-5 6.4 6.2 6.0 5.2<br />

IPCS1-6 6.2 6.1 6.1 5.4<br />

average 6.2 6.1 6.0 5.3<br />

64


Chapter 2: IPC composite specimens under monotonic loading<br />

Table 2.19: evolution <strong>of</strong> the shear modulus, UD-reinforced multiple cracked specimens<br />

shear modulus (GPa)<br />

range angle 1-2° 1-3° 1-4° 1-8°<br />

IPCS1-7 5.0 5.1 4.9 4.4<br />

IPCS1-8 5.1 5.2 4.6 4.6<br />

IPCS1-9 5.3 5.2 4.6 4.6<br />

average 5.1 5.2 4.7 4.5<br />

From tables 2.15 to 2.19, it can be seen that the shear modulus decreases <strong>with</strong><br />

increasing torsion moment (or torsion angle) for all specimens. The initial value <strong>of</strong><br />

the experimentally obtained shear modulus is about 6GPa for the UD-reinforced<br />

specimens <strong>and</strong> 4GPa for the 2D-r<strong>and</strong>omly reinforced specimens. If the specimens<br />

are subjected to tensile load up to the post-cracking zone, prior to torsion testing,<br />

the initial value <strong>of</strong> the shear modulus is lower. The shear modulus <strong>of</strong> the postcracked<br />

specimens is about 5GPa for UD-reinforced specimens <strong>and</strong> 3GPa for 2Dr<strong>and</strong>omly<br />

reinforced specimens. These values will be used further as material<br />

properties in finite element calculations.<br />

2.9 Conclusions<br />

The behaviour <strong>of</strong> unidirectionally reinforced IPC specimens <strong>and</strong> 2D-r<strong>and</strong>omly<br />

reinforced specimens under monotonic loading is tested <strong>and</strong> discussed in this<br />

chapter.<br />

- The behaviour <strong>of</strong> IPC composites in compression is tested <strong>and</strong> discussed. The<br />

introduction <strong>of</strong> fibres changes the material properties <strong>of</strong> IPC only slightly. The<br />

stress-strain behaviour <strong>of</strong> IPC composite specimens in compression is linear<br />

elastic up to failure. This has been verified experimentally. An order <strong>of</strong> magnitude<br />

<strong>of</strong> IPC composite material properties in compression is about 18GPa for the<br />

Young’s modulus <strong>and</strong> 90MPa for the compressive strength. These values are to be<br />

used as compressive material properties in finite element calculations further in<br />

this work.<br />

- The stress-strain behaviour <strong>of</strong> IPC composite specimens in tension is highly nonlinear.<br />

Two models, based on meso-mechanical phenomena, are presented here for<br />

UD-reinforced specimens.<br />

1. The first model (referred to as the ACK model) uses the assumption <strong>of</strong><br />

unique tensile strength <strong>of</strong> the matrix along the whole composite. The<br />

phenomenon, at which matrix cracks appear along the whole composite, is<br />

called “multiple cracking”. Before multiple cracking occurs, the composite<br />

obeys the law <strong>of</strong> mixtures. After full matrix multiple cracking occurred,<br />

only the fibres provide further stiffness. This model has been derived by<br />

65


Chapter 2: IPC composite specimens under monotonic loading<br />

Aveston et al. (1971) <strong>and</strong> has been tested on UD-reinforced IPC specimens<br />

by Bauweraerts et al. (1998a) <strong>and</strong> Bauweraerts (1998b).<br />

2. The second model (referred to as the stochastic cracking model)<br />

implements the stochastic nature <strong>of</strong> the matrix tensile strength in the<br />

composite. It is verified that a two-parameter Weibull distribution function<br />

is an appropriate statistical model for the IPC matrix tensile strength. The<br />

stochastic cracking model has been derived <strong>and</strong> tested on ceramic<br />

composites by Curtin (1998, 1999). This model is used <strong>and</strong> discussed in<br />

this work on UD-reinforced IPC specimens.<br />

- The stress-strain behaviour <strong>of</strong> 2D-r<strong>and</strong>omly reinforced IPC specimens is also<br />

tested <strong>and</strong> discussed in this chapter.<br />

1. Aveston <strong>and</strong> Kelly (1973) presented an extended theoretical model for<br />

2D-r<strong>and</strong>omly reinforced brittle matrix composites. Gu et al. (1998) use this<br />

model on 2D-r<strong>and</strong>omly reinforced IPC specimens.<br />

2. An extended stochastic cracking model is discussed on 2D-r<strong>and</strong>omly<br />

reinforced IPC specimens in this work.<br />

- The ACK (UD-reinforced) <strong>and</strong> ACK based (2D-r<strong>and</strong>omly reinforced) models<br />

provide a predictive tool for the stress-strain behaviour <strong>of</strong> IPC composite<br />

specimens in tension. If the stress-strain behaviour is to be predicted in the precracking<br />

or post-cracking stress-strain zone, the ACK model includes only one (or<br />

two) material parameter(s), not always accurately known a priori: the matrix<br />

failure stress σmu (<strong>and</strong> efficiency factor K). The stochastic cracking theory needs<br />

input <strong>of</strong> three material parameters, not necessarily known a priori. The advantage<br />

<strong>of</strong> the stochastic cracking theory (UD-reinforced) or stochastic cracking based<br />

theory (2D-r<strong>and</strong>omly reinforced) is found in the fact that it provides a more<br />

accurate <strong>and</strong> physically more correct prediction <strong>of</strong> the stress-strain curve in the<br />

matrix multiple cracking zone.<br />

- The evolution <strong>of</strong> the transverse matrix <strong>and</strong> composite strains <strong>with</strong> increasing<br />

tensile stress is formulated theoretically according to the ACK theory. An<br />

important conclusion, which is obtained theoretically <strong>and</strong> experimentally, is that<br />

the transverse matrix stresses remain very low. This is the case for beam<br />

specimens, but also for wide-beam specimens (plate loaded in one direction). The<br />

Poisson’s ratio <strong>of</strong> the composite specimens decreases considerably <strong>with</strong> increasing<br />

composite loading.<br />

- An order <strong>of</strong> magnitude <strong>of</strong> the value <strong>of</strong> the shear modulus is obtained<br />

experimentally here. The shear modulus <strong>of</strong> UD-reinforced specimens is about<br />

6GPa in the pre-cracked state <strong>and</strong> decreases to 5GPa in the post-cracked state<br />

(multiple cracking in tension). The shear modulus <strong>of</strong> 2D-r<strong>and</strong>omly reinforced<br />

specimens is about 4GPa in pre-cracked state <strong>and</strong> 3GPa in post-cracked state.<br />

66


Chapter 2: IPC composite specimens under monotonic loading<br />

2.10 References<br />

H.G. Allen, The purpose <strong>and</strong> methods <strong>of</strong> fibre reinforcement, In Prospects<br />

<strong>of</strong> Fibre Reinforced Construction Materials, Proc. Int. Building Exhibition<br />

Conference, Building Research Station, UK, 1971, pp.3-14<br />

H.G. Allen, The strength <strong>of</strong> thin composites <strong>of</strong> finite width, <strong>with</strong> brittle<br />

matrices <strong>and</strong> r<strong>and</strong>om discontinuous reinforcing fibres, J. Phys. D. Appl. Phys.,<br />

Vol. 5, 1972, pp.331-343<br />

ASTM D3410-75, Compressive properties <strong>of</strong> unidirectional or cross-ply<br />

fibre-resin composites, reapproved 1985<br />

J. Aveston, G.A. Cooper <strong>and</strong> A Kelly, Single <strong>and</strong> multiple fracture, The<br />

Properties <strong>of</strong> Fibre Composites, Proc. Conf. National Physical Laboratories, IPC<br />

Science & Technology Press Ltd. London, 1971, pp.15-24<br />

J. Aveston <strong>and</strong> A. Kelly, Theory <strong>of</strong> multiple fracture <strong>of</strong> fibrous composites,<br />

J. Mat. Sci., Vol. 8, 1973, pp.411-461<br />

J. Aveston, R.A. Mercer, J.M. Sillwood, Fibre reinforced cements –<br />

scientific foundations for specifications, In Composites – St<strong>and</strong>ards, Testing <strong>and</strong><br />

<strong>Design</strong>, Proc. National Physical Laboratories Conference, UK, 1974, pp.93-103<br />

P.J.M. Bartos <strong>and</strong> W. Zhu, Assessment <strong>of</strong> interfacial microstructure <strong>and</strong><br />

bond properties in aged GRC using a novel microindentation method, Cement &<br />

Concrete Research, 27, 1997, pp.1701-1712<br />

P. Bauweraerts, J. Wastiels, X. Wu, H. Cuypers <strong>and</strong> J. Gu, Evaluation <strong>of</strong><br />

damage accumulation <strong>of</strong> glass fibre reinforced brittle matrix composite after<br />

cyclic loading, Durability <strong>of</strong> Composites for Construction, Quebec, august 5-7,<br />

1998a<br />

P. Bauweraerts, Aspects <strong>of</strong> the Micromechanical Characterisation <strong>of</strong> Fibre<br />

Reinforced <strong>Brittle</strong> <strong>Matrix</strong> Composites, Phd. thesis, VUB 1998b<br />

A. Bentur <strong>and</strong> S. Mindess, Fibre Reinforced Cementitious Composites,<br />

Elsevier Applied Science, 1990<br />

K.K. Chawla, Composite materials, Springer-Verlag New York, 1987<br />

K.K. Chawla, Ceramic matrix composites, Chapman & Hall, 1993<br />

67


Chapter 2: IPC composite specimens under monotonic loading<br />

H.L. Cox, The elasticity <strong>and</strong> strength <strong>of</strong> paper <strong>and</strong> other fibrous materials,<br />

British Journal <strong>of</strong> Applied Physics, Vol. 3, 1952, pp.72-79<br />

W.E.C. Creyke, I.E.J. Sainsbury, R.Morrell, <strong>Design</strong> <strong>with</strong> non-ductile<br />

materials, Applied Science Publishers, 1982<br />

W.A. Curtin, Multiple matrix cracking in brittle matrix composites, Acta<br />

metall. mater., No. 41, 1993, pp.1369<br />

W.A. Curtin, B.K. Ahn; N. Takeda, Modeling <strong>Brittle</strong> <strong>and</strong> Tough Stressstrain<br />

Behaviour in Unidirectional Ceramic Composites, Acta mater., No. 10,<br />

1998, pp.3409-3420<br />

W.A. Curtin, Stochastic Damage Evolution <strong>and</strong> Failure in Fibre-<br />

Reinforced Composites, Advances in Applied Mechanics; Vol. 36, 1999, pp.163-<br />

253<br />

H. Cuypers, The behaviour <strong>of</strong> E-glass fibre reinforced brittle matrix<br />

composite subjected to increasing cyclic loading, internal report, department<br />

MEMC, Vrije Universiteit Brussel, December 1999<br />

H. Cuypers, J. Gu, K. Croes, S. Dumortier, J. Wastiels, Evaluation <strong>of</strong><br />

fatigue <strong>and</strong> durability properties <strong>of</strong> E-glass fibre reinforced phosphate<br />

cementitious composites, Proc. Int. Symp. <strong>Brittle</strong> <strong>Matrix</strong> Composites 6, A.M.<br />

Br<strong>and</strong>t, V.C. Li, I.H. Marshall, Warsaw, October 9-11, 2000<br />

H. Cuypers, Use <strong>of</strong> a stochastic cracking based model on the behaviour <strong>of</strong><br />

2D-r<strong>and</strong>om reinforced IPC composite specimens, internal report, department<br />

MEMC, Vrije Universiteit Brussel, 2001a<br />

H. Cuypers, The behaviour <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> E-glass fibre<br />

reinforced cementitious faces under repeated loading, Proc. Composites in<br />

Construction International Conference, October 5-7, 2001b<br />

R.W. Davidge, Mechanical Behaviour <strong>of</strong> Ceramics, Cambridge University<br />

Press, 1979<br />

J. Gu, internal report, Characterisation <strong>and</strong> evaluation <strong>of</strong> IPC laminates<br />

<strong>and</strong> other cementitious laminates, VUB, 2001<br />

J. Gu, X. Wu, H. Cuypers <strong>and</strong> J. Wastiels, Modeling <strong>of</strong> the tensile<br />

behaviour <strong>of</strong> an E-glass fibre reinforced phosphate cement, Computer Methods in<br />

Composite Materials VI, proceedings CADCOMP 98, 1998, pp.589-598<br />

68


Chapter 2: IPC composite specimens under monotonic loading<br />

M.Y. He, B.-X. Wu, A.G. Evans, J.W. Hutchinson, Inelastic strains due to<br />

matrix cracking in unidirectional fiber-reinforced composites, Mechan. Mater.,<br />

Vol. 18, 1994, pp.213<br />

H. Krenchel, Fibre Reinforcement, Akademisk Forlag, Copenhagen, 1964<br />

V. Laws, The efficiency <strong>of</strong> fibrous reinforcement <strong>of</strong> brittle matrices, J.<br />

Phys. D. Appl. Phys., Vol. 4, 1971, pp.1737-1746<br />

V.C. Li, Y. Wang <strong>and</strong> S. Backer, Effect <strong>of</strong> inclining angle, bundling <strong>and</strong><br />

surface treatment on synthetic fibre pull-out from a cement matrix, Composites,<br />

Vol. 21, No. 2, 1990, pp.132-140<br />

F.L. Matthews <strong>and</strong> R.D. Rawlings, Composite Materials: Engineering <strong>and</strong><br />

Science, Chapman <strong>and</strong> Hall, 1994<br />

A.I. Mitropoulos, A study <strong>of</strong> the time-dependent mechanical behaviour <strong>of</strong> a<br />

graphite/epoxy composite material system under uniaxial compression loading,<br />

Master thesis VUB, 1996<br />

A.E. Naaman, G.G. Namur, J.M. Alwan <strong>and</strong> H.S. Najm, Fiber pullout <strong>and</strong><br />

bond slip.I: Analytical study, Journal <strong>of</strong> Structural Engineering, Vol. 117, No. 9,<br />

1991, pp.2769-2790<br />

National Belgian St<strong>and</strong>ards, NBN-B12-208<br />

National Belgian St<strong>and</strong>ards, NBN-B14-209<br />

PCT Patent Application: WO 97/19033. Patent Assignee: Vrije Universiteit<br />

Brussel, Dept. <strong>of</strong> R&D, Pleinlaan 2, 1050 Brussel, Belgium, Inventor: Wu X. &<br />

Gu J. (MEMC-TW-VUB)<br />

1980<br />

M.R. Piggott, Load Bearing Fibre Composites, Permagon Press, Oxford,<br />

A.W. Pryce, P.A. Smith, <strong>Matrix</strong> cracking in uniderictional ceramic<br />

composites under quasi-static <strong>and</strong> cyclic loading, Acta metall. mater., No. 41,<br />

1993, pp.1269-1281<br />

H. Van Herck, Numerieke en experimentele studie van s<strong>and</strong>wichbalken met<br />

IPC glasvezel verstevigde huiden, Master thesis VUB, 1999-2000<br />

69


Chapter 2: IPC composite specimens under monotonic loading<br />

G. Van Vinckenroy, Monte Carlo-based Stochastic Finite Element Method:<br />

an Application to Composite Structures, Phd. thesis VUB, 1994<br />

W. Weibull, A statistical distribution function <strong>of</strong> wide applicability, ASME<br />

J., 1952, pp.293-297<br />

B. Widom, R<strong>and</strong>om sequential addition <strong>of</strong> hard spheres to a volume, J.<br />

Chem. Phys., 44, 1966, pp.3888-3894<br />

F.W. Zok, S.M. Spearings, <strong>Matrix</strong> crack spacing in brittle matrix<br />

composites, Acta metall. mater., Vol. 40, 1992,pp.2033<br />

70


Chapter 3<br />

Unloading <strong>of</strong> IPC composite specimens<br />

3.1 Introduction<br />

In previous chapter, the tensile stress-strain behaviour <strong>of</strong> E-glass fibre reinforced<br />

IPC under loading has been discussed. The two models, presented in previous<br />

chapter, are used as basic models to describe unloading <strong>of</strong> IPC composites here:<br />

the ACK (based) theory <strong>and</strong> the stochastic cracking (based) theory. The derivation<br />

<strong>of</strong> a model for the unloading stress-train behaviour <strong>of</strong> IPC composite laminates is<br />

thus presented in this chapter. Theoretical predictions <strong>of</strong> the stress-strain<br />

unloading behaviour <strong>of</strong> IPC composites are compared <strong>with</strong> experimental<br />

observations.<br />

3.2<br />

Definitions <strong>and</strong> assumptions<br />

The<br />

hypotheses, which are used for the formulation <strong>of</strong> a theoretical unloading<br />

behaviour in this work, equal the ones used for the prediction <strong>of</strong> the stress-strain<br />

loading behaviour <strong>of</strong> IPC composite specimens (see Chapter 2). All assumptions<br />

formulated in paragraph 2.4.2 are thus also adopted here.<br />

Figure<br />

3.1 illustrates the stress-strain behaviour <strong>of</strong> E-glass fibre reinforced IPC<br />

during loading <strong>and</strong> unloading. This figure is based on an experimentally obtained<br />

stress-strain unloading-reloading curve (grey curve in figure 3.1) <strong>of</strong> a 2Dr<strong>and</strong>omly<br />

reinforced specimen. Newly defined variables are illustrated in figure<br />

3.1:<br />

<strong>and</strong> the<br />

the com<br />

- The IPC composite is first loaded up to maximum composite stress σc max<br />

refore undergoes a composite strain εc<br />

ss level, σc . The strain εc is<br />

ive sign<br />

max .<br />

min min<br />

- The composite is unloaded to a lower stre<br />

posite strain, after the composite is unloaded from σc max to σc min .<br />

- The sign if the maximum composite stress σc max has posit<br />

min<br />

(composite is loaded in tension). The minimum composite stress σc can be<br />

positive (tension), zero or negative (compression).<br />

71


Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

- (σc max - σc min ) is defined as the unloading composite stress interval ∆σc <strong>and</strong><br />

(εc max - εc min ) is defined as the unloading composite strain interval ∆εc.<br />

- The slope <strong>of</strong> the straight line, connecting points (σc min , εc min ) <strong>and</strong> (σc max ,<br />

εc max ), is defined as the linearised E-modulus <strong>of</strong> the cycle: Ecycle.<br />

For the implementation <strong>of</strong> the theoretical unloading stress-strain behaviour <strong>of</strong> IPC<br />

composites in finite element calculations, the macro-mechanical<br />

composite<br />

behaviour<br />

is to be described here. As a result, focus is put on expressing Ecycle,<br />

which is a macro-mechanical variable. The formulation <strong>of</strong> Ecycle is based on<br />

physical meso-mechanical phenomena. This formulation <strong>of</strong> Ecycle is derived here as<br />

function <strong>of</strong> the material properties, σc max <strong>and</strong> ∆σc.<br />

stress(MPa)<br />

25 (σc max , εc max )<br />

20<br />

15<br />

10<br />

5<br />

0<br />

(σc min , εc min )<br />

Ecycle<br />

0 0.2 0.4 0.6 0.8<br />

strain (% )<br />

Figure 3.1: loading-unloading stress-strain behaviour <strong>of</strong> a IPC composite specimen<br />

Homogenisation<br />

<strong>of</strong> the internal stresses in the matrix <strong>and</strong> fibres in a composite<br />

section is used as an assumption, as has been done in Chapter 2 for monotonic<br />

tensile<br />

loading. No local stress concentrations are considered <strong>with</strong>in a composite<br />

section, transverse to the loading direction. Fibre stresses, matrix stresses <strong>and</strong><br />

matrix-fibre interface stresses are defined <strong>and</strong> interpreted as in Chapter 2.<br />

During unloading, elastic unloading <strong>of</strong> fibres <strong>and</strong> matrix occurs together <strong>with</strong><br />

frictional matrix-fibre interface slip. Since the assumptions adopted for the<br />

formulation<br />

<strong>of</strong> unloading equal the ones for the formulation <strong>of</strong> loading, the<br />

magnitude <strong>of</strong> the frictional matrix-fibre interface shear stress is equal for loading<br />

<strong>and</strong> unloading. However, it has to be mentioned that the matrix-fibre frictional<br />

interface shear stress is <strong>of</strong> opposite sign for unloading (compared to loading).<br />

3.3 ACK (based) theory for unloading<br />

3.3.1 introduction<br />

72


Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

Aveston et al. (1971) <strong>and</strong> Keer (1981 <strong>and</strong> 1985) used the ACK model to predict<br />

stress-strain<br />

curves for loading <strong>and</strong> unloading <strong>of</strong> brittle matrix composites. The<br />

ACK stress-strain loading<br />

<strong>and</strong> unloading model has been verified by Bauweraerts<br />

(1998) for UD-reinforced IPC specimens. Cuypers (1999) used a ACK based<br />

unloading stress-strain formulation to describe the behaviour <strong>of</strong> 2D-r<strong>and</strong>omly<br />

reinforced IPC specimens. This derivation has been performed analogously to the<br />

derivation <strong>of</strong> the stress-strain behaviour <strong>of</strong> 2D-r<strong>and</strong>omly reinforced IPC composite<br />

specimens under monotonic tensile loading. The derivation <strong>of</strong> Ecycle as presented in<br />

paragraph 3.3.2 can be used for UD-reinforced <strong>and</strong> 2D-r<strong>and</strong>omly reinforced IPC<br />

composites.<br />

3.3.2 derivation <strong>of</strong> the linearised E-modulus <strong>of</strong> a cycle: Ecycle<br />

3.3.2.1 partial<br />

matrix-fibre slip during unloading<br />

Figure<br />

3.2 shows schematically the evolution <strong>of</strong> normal stresses in fibres <strong>and</strong><br />

matrix at composite stress σc max (loading) <strong>and</strong> σc min (unloading).<br />

fibres<br />

f<br />

y cracked composite from maximum<br />

cycle stress σc max Figure 3.2: evolution <strong>of</strong> stresses in matrix <strong>and</strong> fibres in full<br />

min<br />

to minimum cycle<br />

stress σc , partial matrix-fibre unloading slip<br />

The presented formulation <strong>of</strong> the desription <strong>of</strong> the unloading stress-strain<br />

behaviour<br />

is based on figure 3.2.<br />

max<br />

- σ m,<br />

max is the maximum matrix stress in the composite, when maximum<br />

lied.<br />

composite cycle stress σc max is app<br />

x<br />

∆σm elastic<br />

σ<br />

max<br />

m,<br />

max<br />

matrix stress<br />

fibre stress<br />

unloading slip length (su)<br />

loading<br />

elastic unloading<br />

elastic unloading & m atrix-fibre interface slip<br />

73<br />

σ


Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

- In figure 3.2, the grey dashed lines represent the fibre <strong>and</strong> matrix stresses<br />

in the composite at maximum composite load, σc max .<br />

- The full grey lines represent the matrix <strong>and</strong> fibre stresses after unloading<br />

occurred, provided unloading from σc d occur linear elastically.<br />

In<br />

reality<br />

ce shear stress during loading, but occurs in opposite direction. The<br />

full bl<br />

max to σc min woul<br />

- As can be seen from the full grey line in figure 3.2, the matrix at the crack<br />

face would undergo compression, if unloading were to occur linear elastically.<br />

the matrix crack face is free <strong>of</strong> stresses. As a result, the compressive matrix<br />

stresses in the vicinity <strong>of</strong> the crack face are released by slip <strong>of</strong> the matrix along the<br />

fibre: the fibres slip back into the matrix during unloading in the vicinity <strong>of</strong> the<br />

crack face.<br />

- The matrix-fibre unloading slip stress is equal to the frictional matrixfibre<br />

interfa<br />

ack lines in figure 3.2 show the stresses in the matrix <strong>and</strong> fibres after<br />

unloading, as a combination <strong>of</strong> elastic unloading <strong>of</strong> the matrix <strong>and</strong> fibres <strong>and</strong><br />

matrix-fibre unloading slip. These stresses are found in matrix <strong>and</strong> fibres after<br />

unloading is performed, provided full crack closure does not occur. The distance,<br />

along which matrix-fibre slip occurs during unloading, is called the unloading slip<br />

length: (su).<br />

ory predicts linear elastic stress-strain composite behaviour as long<br />

s the maximum composite stress σc max The ACK the<br />

a<br />

does not exceed the multiple cracking<br />

stress. Thus in zone I (linear elastic zone):<br />

∆σ<br />

c ∆ εc<br />

=<br />

(3.1)<br />

E<br />

c1<br />

Once σmc is exceeded, irreversible phenomena occur. If σc max > σmc, the unloading<br />

strain variation ∆ε c is the sum <strong>of</strong> the strain variation due to composite elastic<br />

unloading <strong>and</strong> strain variation due to matrix-fibre unloading slip:<br />

elastic slip<br />

∆ εc = ∆εc<br />

+ ∆εc<br />

(3.2)<br />

The unloading composite strain variation ∆εc equals the average unloading fibre<br />

strain variation ∆ ε along the composite:<br />

∆ε<br />

∆<br />

f<br />

along cs<br />

f<br />

elastic slip<br />

c = ε f = ∆ε<br />

alo<br />

f + ∆ε<br />

ng cs<br />

f<br />

along cs<br />

(3.3)<br />

f<br />

f<br />

The elastic component <strong>of</strong> the fibre strain interval ∆εf is:<br />

elastic<br />

elastic elastic ∆σ<br />

c<br />

∆ ε f = ∆εc<br />

=<br />

(3.4)<br />

Ec1<br />

The formulation <strong>of</strong> the second term <strong>of</strong> equation ( 3.3), which is the fibre slip strain<br />

variation term, is obtained from knowledge <strong>of</strong> the matrix <strong>and</strong> fibre stresses <strong>with</strong>in<br />

the composite (see figure 3.2).<br />

If composite unloading were to occur linear elastically, the matrix stress variation<br />

due to unloading would be:<br />

74


Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

E<br />

∆σ<br />

elastic m c<br />

∆ σ m =<br />

(3.5)<br />

Ec1<br />

Elastic unloading would thus introduce compressive matrix stresses at the crack<br />

face. In reality, the matrix is stress free in the vicinity <strong>of</strong> the crack face. As a<br />

result, the slip term <strong>of</strong> matrix stress variation at the crack face equals the elastic<br />

matrix stress variation (compare the black full line <strong>with</strong> the grey full line in figure<br />

(3.2)):<br />

slip<br />

elastic Em∆σ<br />

c<br />

∆ σ m ( x = 0)<br />

= ∆σ<br />

m =<br />

(3.6)<br />

Ec1<br />

Since matrix-fibre slip is an internal redistribution <strong>of</strong> stresses (<strong>with</strong>out external<br />

force):<br />

∗ slip<br />

slip<br />

V ∆σ<br />

( x = 0)<br />

= V ∆σ<br />

( x = 0)<br />

(3.7)<br />

f<br />

f<br />

m<br />

(su) is defined as the unl ce (su) from a crack face<br />

σ slip (x = (su)) = 0, thus ∆σ slip oading slip length. At a distan<br />

∆ m f (x = (su)) = 0 as well. The average fibre slip<br />

stress along (su) is thus found after equations (3.7) <strong>and</strong> (3.6) are combined:<br />

slip Vm Em∆σ<br />

c<br />

∆σ f =<br />

along su)<br />

∗<br />

V E<br />

(3.8)<br />

m<br />

(<br />

f 2 c1<br />

The average fibre slip stress along f is found after is multiplied <strong>with</strong><br />

2(su)/f:<br />

slip 2(<br />

su)<br />

slip ( su)<br />

Vm<br />

Em<br />

∆σ<br />

f = ∆σ<br />

along cs<br />

f =<br />

∆σ<br />

< ><br />

along su<br />

∗<br />

c<br />

f cs<br />

( ) cs V E<br />

(3.9)<br />

f<br />

From figure (3.2), it can be seen that ratio 2(su)/f can be expressed:<br />

elastic<br />

2(<br />

su) ∆σ<br />

m = max<br />

cs σ<br />

f<br />

m,<br />

max<br />

The combination <strong>of</strong> (3.9) <strong>and</strong> (3.10) gives:<br />

slip<br />

∆σ f<br />

along cs<br />

f<br />

elastic<br />

Vm Em<br />

∆σ<br />

m ∆σ<br />

c<br />

= ∗ max<br />

V E σ 2<br />

f<br />

c1<br />

m,<br />

max<br />

f<br />

f<br />

c1<br />

(3.10)<br />

(3.11)<br />

After equation (3.5) is inserted in equation (3.11) <strong>and</strong> taking into account m,<br />

max =<br />

1.337σ mu (see equation (2.18)), the slip term <strong>of</strong> the average fibre strain variation<br />

is:<br />

slip<br />

∆ε f<br />

along cs<br />

f<br />

=<br />

slip<br />

∆σ<br />

f<br />

along cs<br />

E<br />

f EmVm = ∗<br />

E V<br />

2<br />

Em(<br />

∆σ<br />

c )<br />

2<br />

E 2.<br />

674σ<br />

(3.12)<br />

f<br />

f<br />

f<br />

( c ) mu<br />

Combination <strong>of</strong> equation (3.3), (3.4) <strong>and</strong> (3.12) gives:<br />

( ) 2<br />

2<br />

∆σ<br />

c ( ∆σ<br />

c)<br />

Emα ∆ ε c = +<br />

Ec1 2. 647σ<br />

mu Ec1<br />

-εc min , which is the ratio <strong>of</strong> (σc ) is:<br />

max -σc min ) /(εc max<br />

Thus Ecycle,<br />

75<br />

1<br />

σ max<br />

(3.13)


Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

Ec1<br />

Ecycle<br />

=<br />

max<br />

αEm(<br />

σ c −σ<br />

1+<br />

2.<br />

674Ec1σ<br />

mu<br />

E<br />

<strong>with</strong>: α = mVm<br />

∗<br />

E fV f<br />

min<br />

c<br />

)<br />

(3.14)<br />

3.3.2.2<br />

from partial to total matrix-fibre slip<br />

From the combination <strong>of</strong> equation (3.10) <strong>with</strong><br />

(3.5), it can be noticed that<br />

unloading slip length (su) increases <strong>with</strong> increasing (σc max - σc min ). The fibres slip<br />

along their whole length in the matrix during unloading, when 2(su) equals f<br />

in equation (3.10). The unloading stress variation at which total matrix-fibre<br />

unloading slip is to be expected, is found from the combination <strong>of</strong> equation (3.5)<br />

<strong>and</strong> (3.10), taking into account that 2(su) equals f:<br />

max min 1.<br />

337σ<br />

muEc1<br />

( σ c − σ c ) =<br />

= 1.<br />

337σ<br />

mc<br />

(3.15)<br />

E<br />

m<br />

3.3.2.3 total matrix-fibre slip during unloading<br />

Once the fibres slip in the matrix along their whole<br />

length during unloading,<br />

further unloading occurs due to elastic unloading <strong>of</strong> the fibres only. Figure 3.3<br />

shows the stresses <strong>and</strong> strains in the matrix <strong>and</strong> fibres, once total matrix-fibre slip<br />

occurs during unloading. Again a formulation <strong>of</strong> Ecycle can be found by expressing<br />

the strain variation (εc max - εc min ) as a function <strong>of</strong> (σc max - σc min ). The unloading<br />

matrix stress variation is shown in figure 3.3. From this figure, it can be noticed<br />

that:<br />

m x<br />

f<br />

m σ mu<br />

σ<br />

σ 337 . 1 2 )<br />

∆ ( =<br />

cs<br />

max<br />

= , max<br />

2<br />

=<br />

(3.16)<br />

<strong>and</strong>:<br />

∆σ m(<br />

x = 0)<br />

= 0<br />

The unloading stress variation in the fibres is:<br />

(3.17)<br />

∆σ<br />

c<br />

∆σ<br />

f ( x = 0)<br />

= ∗<br />

V<br />

(3.18)<br />

cs ∆σ c −1.<br />

337V<br />

f<br />

m∆σ<br />

m(<br />

x = cs 2)<br />

f ∆σ<br />

c −1.<br />

337Vmσ<br />

∆σ<br />

f ( x = ) =<br />

=<br />

2<br />

V<br />

V<br />

The value <strong>of</strong> the average strain variation in the composite is thus:<br />

∆εc<br />

= ∆ε<br />

f<br />

along cs<br />

f<br />

∆σ<br />

c − 0.<br />

668Vmσ<br />

mu<br />

=<br />

∗<br />

E V<br />

The linearised E-modulus <strong>of</strong> an unloading-reloading cycle is thus:<br />

∗<br />

f<br />

76<br />

f<br />

f<br />

f<br />

∗<br />

f<br />

mu<br />

(3.19)<br />

(3.20)


fibres<br />

Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

E<br />

x<br />

matrix<br />

stress<br />

E V<br />

∗<br />

f f<br />

cycle =<br />

0.<br />

668σ<br />

muVm<br />

1−<br />

2σm,max max<br />

∆σ<br />

c<br />

loading<br />

unloading<br />

fibre stress<br />

Figure 3.3: evolution <strong>of</strong> stresses in the matrix <strong>and</strong> fibres in fully cracked composite from<br />

maximum cycle stress to minimum cycle stress, total matrix-fibre unloading slip<br />

(3.21)<br />

Effects <strong>of</strong> fibre orientation <strong>and</strong> length (ηθ <strong>and</strong> ηl) are included for unloading in a<br />

way that is similar to the loading <strong>of</strong> the IPC specimens (Laws, 1971; Aveston et<br />

al.,. 1974; Allen, 1972; Gu et al.,1998; Cuypers, 1999 <strong>and</strong> Cuypers et al,.2000).<br />

3.4 Stochastic cracking (based) theory for unloading<br />

3.4.1 introduction<br />

Similar to the derivation <strong>of</strong> Ecycle based on the ACK (based) model, Ecycle is now<br />

formulated <strong>with</strong>in the stochastic cracking (based) model. Rouby <strong>and</strong> Reynaud<br />

(1992) discuss the unloading behaviour <strong>of</strong> a brittle matrix composite <strong>with</strong> small<br />

( < 2δ0) <strong>and</strong> large ( > 2δ0) crack spacing. In the publication <strong>of</strong> Rouby <strong>and</strong><br />

Reynaud (1992), the relation between the average crack spacing <strong>and</strong> maximum<br />

cycle stress is not introduced yet. The crack spacing is determined visually by<br />

these authors for each value <strong>of</strong> σc max . Price <strong>and</strong> Smith (1993) worked on the<br />

relationship between the value <strong>of</strong> the frictional matrix-fibre interface shear stress<br />

<strong>and</strong> on hysteresis in an unloading-reloading cycle. Ahn <strong>and</strong> Curtin (1997) discuss<br />

a stress-strain loading-unloading-reloading evolution <strong>of</strong> UD-reinforced ceramic<br />

composites, based on the stochastic cracking theory, as presented in Chapter 2 in<br />

77<br />

σ


Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

this work. Cuypers (2001) discusses the stochastic cracking based unloading<br />

model for UD-reinforced <strong>and</strong> 2D-r<strong>and</strong>omly reinforced IPC composites.<br />

3.4.2 derivation <strong>of</strong> linearised E-modulus <strong>of</strong> a cycle: Ecycle<br />

3.4.2.1 partial matrix-fibre slip during unloading<br />

Figure 3.4a shows the evolution <strong>of</strong> the matrix <strong>and</strong> fibre stresses during unloading,<br />

provided > 2δ0 <strong>and</strong> only partial matrix-fibre slip occurs during unloading.<br />

Figure 3.4b shows the matrix <strong>and</strong> fibre stresses, provided < 2δ0 <strong>and</strong> partial<br />

matrix-fibre slip occurs during unloading.<br />

<br />

x<br />

∆σm elastic<br />

σm,max<br />

matrix<br />

stress<br />

(su)<br />

fibre<br />

stress<br />

Figure 3.4a: stresses in matrix <strong>and</strong> fibres<br />

during unloading <strong>of</strong> composite, partial matrixfibre<br />

unloading slip, > 2δ0<br />

σ<br />

<br />

∆σm elastic<br />

x<br />

loading<br />

elastic unloading<br />

real unloading<br />

matrix<br />

stress<br />

σm,max<br />

(su)<br />

fibre<br />

stress<br />

Figure 3.4b: stresses in matrix <strong>and</strong> fibres<br />

during unloading <strong>of</strong> composite, partial matrixfibre<br />

unloading slip, < 2δ0<br />

The derivation <strong>of</strong> Ecycle based on the stochastic (based) cracking theory is similar<br />

to the derivation based on the ACK (based) theory. The derivation <strong>of</strong> equations<br />

(3.22) to (3.33) shows some similarities <strong>with</strong> the derivation presented in equations<br />

(3.1) to (3.14). The composite strain interval can be written:<br />

elastic slip<br />

∆ εc = ∆εc<br />

+ ∆εc<br />

(3.22)<br />

<strong>with</strong>:<br />

elastic ∆σ<br />

c<br />

∆ εc<br />

=<br />

Ec1<br />

<strong>and</strong> similar to equation (3.9):<br />

(3.23)<br />

slip<br />

∆εc =<br />

slip<br />

∆ε<br />

f =<br />

along < cs><br />

slip<br />

∆σ<br />

f<br />

along < cs><br />

( su)<br />

Vm<br />

Em∆σ<br />

c<br />

=<br />

∗<br />

cs V E E<br />

(3.24)<br />

E f<br />

f c1<br />

78<br />

f<br />

σ


Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

Since matrix cracking is modelled as a stochastic process now, is function <strong>of</strong><br />

the maximum composite stress σc max .<br />

cs = cs<br />

max −1<br />

( 1−<br />

exp( −F<br />

( σ c ))<br />

f<br />

(3.25)<br />

<strong>with</strong>:<br />

F<br />

max ( σ )<br />

c<br />

⎛σ<br />

max<br />

c =<br />

⎜<br />

σ R<br />

⎝<br />

⎞<br />

⎟<br />

⎠<br />

m<br />

(3.26)<br />

If linear elastic unloading were to occur, the matrix stress at the crack face would<br />

be (after unloading):<br />

elastic Em<br />

∆σ<br />

m = ∆σ<br />

c<br />

(3.27)<br />

E<br />

c1<br />

In reality, the matrix is stress free at the crack face: matrix-fibre slip occurs at the<br />

crack face during unloading. The frictional matrix-fibre interface shear stress<br />

during unloading is equal to the frictional matrix-fibre interface shear stress during<br />

loading, but in opposite direction. (su) in equation (3.24) can be rewritten as a<br />

function <strong>of</strong> the debonding length δ0. The ratio <strong>of</strong> the unloading fibre slip length<br />

(su) versus the debonding length δ0 is illustrated in figure 3.5a <strong>and</strong> 3.5b.<br />

x<br />

x<br />

δ0<br />

∆σm elastic<br />

matrix stress<br />

(su)<br />

σm ff,max<br />

fibre stress<br />

Figure 3.5a: illustration <strong>of</strong> (su), δ0, σm ff,max <strong>and</strong><br />

∆σm elastic , partial matrix-fibre unloading slip,<br />

> 2δ0<br />

σ<br />

δ0<br />

∆σm elastic<br />

loading<br />

elastic unloading<br />

real unloading<br />

From figure 3.5a <strong>and</strong> 3.5b, it can be seen that:<br />

79<br />

matrix stress<br />

(su)<br />

σm ff,max<br />

fibre stress<br />

Figure 3.5b: illustration <strong>of</strong> (su), δ0, σm ff,max <strong>and</strong><br />

∆σm elastic , partial matrix-fibre unloading slip,<br />

< 2δ0<br />

σ


Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

∆σ<br />

m<br />

( su)<br />

= 2<br />

δ σ<br />

elastic<br />

(3.28)<br />

ff , max<br />

0 m<br />

σm ff,max is the far field matrix stress at maximum composite stress σc max <strong>and</strong> is<br />

formulated by equation (2.40). σm ff,max can thus be expressed as:<br />

E σ<br />

ff , max m max<br />

σ m = c<br />

(3.29)<br />

Ec1<br />

A formulation <strong>of</strong> (su) is found after equation (3.29) is inserted in equation (3.28):<br />

elastic<br />

∆σ<br />

m Ec1<br />

∆σ<br />

c<br />

( su)<br />

= δ 0 = δ<br />

max 0 max<br />

2Emσ<br />

c 2σ<br />

c<br />

(3.30)<br />

δ0 can be written, in accordance to equation (2.39):<br />

max<br />

rVmEmσ<br />

c<br />

δ 0 =<br />

∗<br />

2τ E V<br />

(3.31)<br />

0<br />

c1<br />

After some rearrangements, combination <strong>of</strong> equations (3.22), (3.23), (3.24) <strong>and</strong><br />

(3.30) gives:<br />

( ) ⎟ ⎛ ⎞<br />

max min ∆σ<br />

c ⎜<br />

αδ 0∆σ<br />

c<br />

ε c − εc<br />

= 1+<br />

⎜<br />

max<br />

(3.32)<br />

Ec1 ⎝ 2 cs σ x ⎠<br />

Finally, Ecycle is:<br />

Ec1<br />

Ecycle<br />

=<br />

αδ 0∆σ<br />

c 1+<br />

(3.33)<br />

max<br />

2 cs σ<br />

Equation (3.33) is the expression <strong>of</strong> Ecycle in case partial matrix-fibre slip occurs<br />

during unloading. Another formulation <strong>of</strong> Ecycle should be derived, once full<br />

matrix-fibre slip occurs during unloading. Moreover, the conditions under which<br />

partial matrix-fibre unloading slip becomes total matrix-fibre unloading slip<br />

should be expressed. The formulation, which describes when partial becomes full<br />

matrix-fibre slip, is different for > 2δ0 (based on figures 3.4a <strong>and</strong> 3.5a) <strong>and</strong><br />

< 2δ0 (based on figures 3.4b <strong>and</strong> 3.5b).<br />

3.4.2.2 from partial to full matrix-fibre unloading slip, > 2δ0<br />

When > 2δ0, partial matrix-fibre unloading slip becomes full matrix-fibre<br />

unloading slip once (su) reaches the debonding length δ0 (see figure 3.5a). This<br />

means the ratio in equation (3.28) equals one, when partial matrix-fibre slip is<br />

replaced by total matrix-fibre slip. Thus:<br />

, max<br />

2 ff<br />

elastic<br />

∆ σ m = σ m<br />

(3.34)<br />

or after combination <strong>of</strong> equation (3.34) <strong>with</strong> equations (3.27) <strong>and</strong> (3.29):<br />

max<br />

∆ σ c = 2σ c<br />

(3.35)<br />

f<br />

80<br />

c


Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

According to equation (3.35) <strong>and</strong> if > 2δ0, partial matrix-fibre unloading slip<br />

is not replaced <strong>with</strong> full matrix-fibre unloading slip during unloading, as long as<br />

the minimum cycle stress σc min is a tensile stress or equals zero.<br />

If > 2δ0, partial matrix-fibre unloading slip cannot be replaced by total<br />

matrix-fibre unloading slip. It will be verified here that matrix cracks will always<br />

close before the total matrix-fibre unloading slip phenomenon occurs, still<br />

provided > 2δ0. Once the matrix crack faces come in contact <strong>with</strong> each other,<br />

the matrix crack face is not free <strong>of</strong> stresses any more. From then on, compression<br />

stresses can be transferred again through the matrix at the crack face.<br />

Suppose a IPC composite is unloaded from a tensile stress σc max into compression.<br />

(negative σc min ). The composite stress value, at which the cracks are then closed, is<br />

determined here.<br />

fibres<br />

matrix<br />

crack<br />

matrix<br />

matrix<br />

crack<br />

tensile loading<br />

σc max<br />

matrix crack<br />

closed<br />

matrix crack<br />

closed<br />

composite unloading<br />

σc min<br />

Figure 3.6: matrix crack closure, unloading from σc max to σc min<br />

The value <strong>of</strong> the crack opening is determined by the discrepancy between the<br />

average strain <strong>of</strong> the fibres along the composite <strong>and</strong> the average strain <strong>of</strong> the<br />

matrix along the composite. During loading in tension, the average fibre strain<br />

along is higher than the average matrix strain along, once matrix<br />

cracks are introduced. If a composite specimen is unloaded from tension, the<br />

average unloading strain variation experienced by the fibres (along) is<br />

higher than the average unloading strain variation experienced by the matrix<br />

(along).<br />

When a IPC composite specimen is first loaded in tension <strong>and</strong> then unloaded into<br />

compression, crack closure occurs when:<br />

81


Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

ε f<br />

along < cs ><br />

− ε m along < cs > = ∆ε<br />

f<br />

along < cs ><br />

− ∆ε<br />

m along < cs ><br />

(3.36)<br />

Equation (3.36) should now be rewritten in terms <strong>of</strong> composite stresses. In<br />

Chapter 2, equation (2.48) expresses along as follows:<br />

ε f<br />

σ c<br />

=<br />

along < cs > Ec1<br />

αδ 0 max<br />

+ σ c<br />

Ec1<br />

(3.37)<br />

The average matrix strain along along the composite is derived here,<br />

according to figure (2.14a) <strong>and</strong> equation (2.41) to (2.49). It can be found from this<br />

figure <strong>and</strong> equations that:<br />

εm max<br />

σ c<br />

alongδ<br />

0 2Ec1<br />

= (3.38)<br />

<strong>and</strong> from the same figure <strong>and</strong> equations it can be noticed that:<br />

εm max<br />

σ c<br />

along cs 2δ 0 Ec1<br />

=<br />

< > −<br />

(3.39)<br />

thus:<br />

εm<br />

along < cs><br />

max<br />

σ c =<br />

Ec1 cs − δ 0<br />

cs<br />

(3.40)<br />

The left-h<strong>and</strong> term <strong>of</strong> equation (3.36) is thus:<br />

ε f<br />

max<br />

σ c δ 0<br />

− ε = ( + )<br />

along < cs><br />

m<br />

1 α<br />

along < cs><br />

< cs ><br />

(3.41)<br />

Ec1 The average fibre unloading strain term along is expressed by equation<br />

(3.32) <strong>and</strong> is:<br />

∆σ<br />

c αδ 0∆σ<br />

c<br />

∆ε<br />

f = ( 1+<br />

)<br />

along < cs><br />

max<br />

(3.42)<br />

E 2 cs σ<br />

c1<br />

c<br />

The average matrix unloading strain along along the composite can be<br />

determined from figure (3.4a) <strong>and</strong> (3.5a) <strong>and</strong> equation (3.27) <strong>and</strong> is:<br />

∆σ<br />

c ∆σ<br />

c ( su)<br />

∆ ε m = −<br />

along < cs><br />

(3.43)<br />

E cs<br />

Ec1 c1<br />

<strong>and</strong> after combination <strong>of</strong> (3.43) <strong>with</strong> (3.28):<br />

∆σ<br />

c δ 0∆σ<br />

c<br />

∆ ε m = ( 1−<br />

)<br />

along < cs><br />

max<br />

Ec1 2 cs σ c<br />

The right-h<strong>and</strong> term <strong>of</strong> equation (3.36) is thus:<br />

(3.44)<br />

∆ ε f<br />

2 ( ∆σ<br />

c ) δ 0<br />

− ∆ε<br />

= ( + )<br />

along < cs><br />

m<br />

1 α<br />

along < cs><br />

max<br />

σ<br />

2 < cs ><br />

(3.45)<br />

c<br />

Finally, combination <strong>of</strong> equations (3.36), (3.41) <strong>and</strong> (3.45) leads to the expression<br />

<strong>of</strong> matrix crack closure:<br />

max<br />

∆ σ c = 2σ c<br />

(3.46)<br />

82<br />

Ec1


Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

Consequently, if a composite specimen is loaded in tension such that > 2δ0<br />

<strong>and</strong> later unloaded, only partial matrix-fibre unloading slip can occur. If the<br />

specimen is loaded in compression, after being loaded in tension, the matrix cracks<br />

close (see equation (3.46)) before the partial matrix-fibre unloading slip<br />

mechanism is replaced by a total matrix-fibre unloading slip mechanism (see<br />

equation (3.35)).<br />

3.4.2.3 from partial to full matrix-fibre unloading slip, < 2δ0<br />

When < 2δ0, partial matrix-fibre unloading slip becomes full matrix-fibre<br />

unloading slip once (su) becomes half the crack spacing (see figure 3.4b). (su) is<br />

expressed by equation (3.30). The condition at which partial unloading matrixfibre<br />

slip is replaced <strong>with</strong> total matrix-fibre slip is thus:<br />

elastic<br />

σ m<br />

cs δ 0 ff , max<br />

σ m<br />

∆<br />

= (3.47)<br />

<strong>and</strong> after equation (3.27) <strong>and</strong> (3.29) are inserted in equation (3.47):<br />

cs max<br />

∆ σ c = σ c<br />

(3.48)<br />

δ<br />

0<br />

In case < 2δ0 <strong>and</strong> σc min = 0, from equation (3.48) can be seen that:<br />

1. if > δ0: the partial matrix-fibre unloading slip mechanism is<br />

replaced by a total matrix-fibre unloading slip mechanism only when the<br />

composite is loaded in compression after it has been subjected to tensile load<br />

2. if < δ0: the partial matrix-fibre unloading slip mechanism is<br />

replaced by a total matrix-fibre unloading slip mechanism before the tensile load<br />

σc max is released completely<br />

3.4.2.4 total matrix-fibre slip during unloading<br />

If < 2δ0, partial matrix-fibre unloading slip can be replaced by total matrixfibre<br />

unloading slip. A new formulation <strong>of</strong> Ecycle is then to be derived. The stresses<br />

in matrix <strong>and</strong> fibres after loading <strong>and</strong> unloading are illustrated in figure 3.7. Once<br />

total matrix-fibre slip occurred during unoading, the stresses in the matrix do not<br />

vary any more. If the composite is unloaded any further, only the fibres provide<br />

stiffness.<br />

From figure 3.7, it can be seen that the maximum <strong>and</strong> minimum matrix stress<br />

variations are expressed as:<br />

∆σ m(<br />

x = 0)<br />

= 0<br />

(3.49)<br />

∆ σ m(<br />

x =<br />

cs<br />

max<br />

) = 2σ<br />

m ( x =<br />

2<br />

cs<br />

max<br />

) = 2σ<br />

m,<br />

max<br />

2<br />

(3.50)<br />

The fibre unloading stress variation is thus:<br />

83


<strong>and</strong>:<br />

thus:<br />

matrix<br />

stress<br />

Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

∆σ<br />

∆σ<br />

f ( x = 0)<br />

=<br />

V<br />

(3.51)<br />

c<br />

∗<br />

f<br />

cs ∆σ<br />

c − 2σ<br />

Vm<br />

∆σ<br />

f ( x = ) =<br />

(3.52)<br />

2 V<br />

0<br />

max<br />

m,<br />

max<br />

∗<br />

f<br />

cs / 2 ff , max cs Em<br />

max<br />

σ m,<br />

max = σ m = σ c<br />

(3.53)<br />

δ<br />

2δ<br />

E<br />

∆σ<br />

( x =<br />

f<br />

x<br />

cs σ<br />

∆σ<br />

c −<br />

cs<br />

δ E<br />

) =<br />

2<br />

V<br />

2σm,max<br />

0<br />

0<br />

∗<br />

f<br />

c1<br />

max<br />

c<br />

loading<br />

unloading<br />

c1<br />

E<br />

m<br />

V<br />

m<br />

fibre stress<br />

σ<br />

(3.54)<br />

Figure 3.7 stresses in matrix <strong>and</strong> fibres after loading <strong>and</strong> unloading <strong>of</strong> composite, total matrixfibre<br />

unloading slip, < 2δ0<br />

The average value <strong>of</strong> the unloading strain variation along the composite is then:<br />

∆εc = ∆ε<br />

f<br />

1<br />

= ( ∆σ<br />

along ( su)<br />

c −<br />

∗<br />

E V<br />

max<br />

cs EmVmσ<br />

c<br />

)<br />

2δ<br />

E<br />

(3.55)<br />

Finally, Ecycle is:<br />

E<br />

cycle<br />

f<br />

f<br />

∗<br />

E fV<br />

f<br />

=<br />

cs Em<br />

σ<br />

1−<br />

Vm<br />

2δ<br />

E ∆σ<br />

0<br />

c1<br />

84<br />

max<br />

c<br />

c<br />

0<br />

c1<br />

(3.56)


Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

3.5 IPCstress-strain.exe: a program to predict unloading<br />

A program has been written in Visual Basic for this thesis: IPCstress-strain.exe.<br />

The aim <strong>of</strong> this program is the prediction <strong>of</strong> theoretical stress-strain loading <strong>and</strong><br />

unloading curves <strong>of</strong> IPC composite laminates. The algorithm on which IPCstressstrain.exe<br />

is based, is illustrated in figure 3.8 <strong>and</strong> explained here.<br />

EXPERIMENTAL INPUT<br />

stress-strain curve from simple<br />

tensile testing<br />

σ<br />

USER INPUT<br />

σc min to which each<br />

unloading occurs,<br />

limits <strong>of</strong> σc max<br />

ε<br />

Ecycle<br />

USER INPUT<br />

Em, Ef, Vf, type <strong>of</strong> reinforcement, f<br />

PROGRAM IPCstress-strain.exe<br />

module : loading<br />

optimising σmu, K, τ0, σR, m for best fit <strong>of</strong><br />

theoretical curves <strong>with</strong> experimental curves<br />

OUTPUT<br />

optimised values <strong>of</strong> σmu, K, τ0, σr, m<br />

INPUT<br />

PROGRAM IPCstress-strain.exe<br />

module : unloading<br />

calculation <strong>of</strong> Ecycle as function <strong>of</strong><br />

σc max , from which unloading occurs, <strong>with</strong><br />

ACK theory <strong>and</strong> stochastic cracking theory<br />

OUTPUT<br />

theoretical curves Ecycle versus σc max<br />

σc max<br />

ACK theory<br />

stochastic cracking theory<br />

Figure 3.8: algorithm <strong>of</strong> program IPCstress-strain.exe, prediction <strong>of</strong> unloading <strong>with</strong> ACK (based)<br />

model <strong>and</strong> stochastic cracking (based) model<br />

85


Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

- The user can insert material parameters: Em, Ef, Vf, type <strong>of</strong> reinforcement,<br />

r, m, τ0, f, σR, K <strong>and</strong> σmu,. A theoretical stress-strain curve is then determined<br />

under monotonic tensile loading.<br />

- If experimental stress-strain curves are obtained under monotonic tensile<br />

loading (loading up to failure in tension), the experimental <strong>and</strong> theoretical stressstrain<br />

curves under loading can be compared. As was already mentioned in<br />

Chapter 2, accurate a priori knowledge <strong>of</strong> τ0, σR <strong>and</strong> m is not always available in<br />

case the stochastic cracking (based) theory is used. In case the ACK (based) theory<br />

is used, σmu, <strong>and</strong> K are not always determined in an accurate way a priori. A “best<br />

fit” <strong>of</strong> a theoretical <strong>with</strong> an experimental stress-strain curve can be obtained by<br />

varying these material parameters until the LSC (least squares coefficient) is<br />

minimised (see equation (2.58)).<br />

- The value <strong>of</strong> the material parameters obtained from “best fit” <strong>of</strong> theory<br />

<strong>with</strong> experiments under loading are now used to predict the behaviour <strong>of</strong> the<br />

composite during unloading.<br />

- The user inserts a value <strong>of</strong> σc min , the program returns the evolution <strong>of</strong><br />

Ecycle as a function <strong>of</strong> σc max . For the prediction <strong>of</strong> this theoretical evolution, the<br />

values <strong>of</strong> τ0, σR, m σmu, <strong>and</strong> K are used as determined from comparison <strong>of</strong> the<br />

theoretical <strong>and</strong> experimental stress-strain loading curves.<br />

3.6 Experiments<br />

3.6.1 test set-up <strong>and</strong> materials<br />

Four different 2D-r<strong>and</strong>omly reinforced IPC laminates are made for testing. Their<br />

properties are listed in table 3.1.<br />

Table 3.1: characteristics <strong>of</strong> tested laminates<br />

plate nr. # fibre layers matrix weight fibre mat Vf<br />

(- ) (g/m²/layer) (g/m²/layer) (%)<br />

RU1 2 1800 300 9.42<br />

RU2 4 1800 300 11.6<br />

RU3 2 2000 300 7.53<br />

RU4 2 1800 450 10.6<br />

For each laminate made, some <strong>of</strong> the fresh matrix mixture was taken apart <strong>and</strong><br />

poured into three blocks <strong>of</strong> 160x40x40mm³. Plates RU1 <strong>and</strong> RU2 <strong>and</strong> the matrix<br />

blocks from mixtures RU1 <strong>and</strong> RU2 are tested one month after casting. Plates<br />

RU3 <strong>and</strong> RU4 are tested one week after casting. Therefore, the matrix from the<br />

latter plates did probably not obtain full strength yet.<br />

86


Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

From each laminate, six specimens are cut (250x18mm²). Per plate, three <strong>of</strong> these<br />

specimens are tested on an INSTRON 4505 bench up to failure under monotonic<br />

tensile loading, similar to paragraph (2.5.4). A load cell measures the load <strong>and</strong> the<br />

strains are monitored by an extensometer. Three other specimens are measured<br />

under increasing cyclic loading, as illustrated in figure 3.9. The maximum stress <strong>of</strong><br />

each cycle σc max is 5.5MPa higher than the previous maximum cycle stress. The<br />

minimum stress σc min is 1.0MPa for each cycle, to prevent accidentally<br />

compression <strong>of</strong> the specimens.<br />

stress(MPa)<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 0.2 0.4 0.6 0.8 1 1.2<br />

strain (% )<br />

Figure 3.9: experimental stress-strain curve under increasing cyclic loading, 2D-r<strong>and</strong>omly<br />

reinforced specimen<br />

3.6.2 test results<br />

From the experimentally obtained stress-strain curves under limited cyclic loading,<br />

an experimental curve <strong>of</strong> the evolution <strong>of</strong> Ecycle as function <strong>of</strong> σc max is determined.<br />

An average evolution <strong>of</strong> Ecycle as function <strong>of</strong> σc max is determined for each plate<br />

from results on three specimens. The evolution <strong>of</strong> Ecycle is shown for the four<br />

tested plates in figure 3.10.<br />

E modulus (GPa)<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

plate RU1<br />

plate RU2<br />

plate RU3<br />

plate RU4<br />

Figure 3.10: Ecycle versus maximum cycle stress σc max 0<br />

0 5 10 15 20 25 30 35<br />

stress (MPa)<br />

<strong>of</strong> cycled specimens<br />

87


Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

Some preliminary conclusion can be formulated from figure 3.10:<br />

- Plate RU1 <strong>and</strong> RU2 are tested one month after preparation. Plate RU3 <strong>and</strong><br />

RU4 are tested one week after preparation. Like most cementitious materials, the<br />

matrix strength still increases <strong>with</strong> increasing age. This effect can be seen in figure<br />

3.10. For the first load cycles (5.5MPa <strong>and</strong> 11MPa), Ecycle is higher for plates RU1<br />

<strong>and</strong> RU2 (tested after one month) than for plates RU3 <strong>and</strong> RU4 (tested after one<br />

week). <strong>Matrix</strong> cracking occurs at higher stresses if the matrix failure stress is<br />

higher.<br />

- Plate RU2 <strong>and</strong> RU4 have higher fibre volume fraction than plate RU1 <strong>and</strong><br />

RU3. The value <strong>of</strong> Ecycle is higher for plate RU2 <strong>and</strong> plate RU4 for higher values <strong>of</strong><br />

the unloading stress σc max . If the composite specimens are loaded up to relatively<br />

high tensile stress levels, the stiffness provided by the fibre becomes a dominant<br />

term in the composite stiffness (both for loading <strong>and</strong> unloading).<br />

3.6.3 unloading behaviour as predicted by the ACK based model<br />

From each mixture, three matrix blocks are casted <strong>and</strong> are subjected to a threepoint<br />

bending test. From this test, pure matrix properties are determined: the<br />

ultimate stress σmu <strong>and</strong> strain εmu <strong>of</strong> the matrix (according to st<strong>and</strong>ards NBN-B12-<br />

208 <strong>and</strong> NBN-B14-209). The average value <strong>of</strong> the failure stress <strong>and</strong> strain <strong>of</strong> the<br />

three tested specimens is listed in table 3.2.<br />

Table 3.2: properties <strong>of</strong> pure IPC matrix<br />

nr σmu<br />

εmu<br />

(MPa) (%)<br />

RU1 11.5 0.0672<br />

RU2 10.6 0.0641<br />

RU3 8.65 0.0543<br />

RU4 8.57 0.0526<br />

The value <strong>of</strong> σmu in table 3.2 is determined from three-point bending on pure<br />

matrix blocks. As has been mentioned several times in Chapter 2, the value <strong>of</strong> σmu<br />

as determined on pure IPC might differ from σmu <strong>of</strong> IPC matrix in a composite.<br />

Similar to paragraph 2.5.4, K <strong>and</strong> σmu are also obtained by determination <strong>of</strong> a<br />

“best fit” <strong>of</strong> experimental <strong>and</strong> theoretical stress-strain loading curves.<br />

Before the LSC (least squares coefficient) is to be minimised, efficiency factor K<br />

is determined, according to equation (2.90). The resulting values <strong>of</strong> σmu <strong>and</strong> K are<br />

listed in table 3.3. From knowledge <strong>of</strong> σmu, the value <strong>of</strong> σmc is determined<br />

(equation (2.10)).<br />

88


Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

From table 3.3, it can be noticed there is only small discrepancy between the value<br />

<strong>of</strong> the multiple cracking stress, as determined from pure IPC blocks <strong>and</strong> from “best<br />

fitting” <strong>of</strong> theoretical <strong>and</strong> experimental composite stress-strain curves. The results<br />

from three-point bending are good approximations <strong>of</strong> the multiple cracking<br />

stresses.<br />

Table 3.3: material parameters for ACK based model, plates RU1 to RU4<br />

plate nr RU1 RU2 RU3 RU4<br />

Vf (%) 9.42 11.6 7.53 10.6<br />

K (-) 0.83 0.86 0.93 0.85<br />

σmu 3-point bending (MPa) 11.5 10.6 8.65 8.57<br />

σmu stress-strain fitting (MPa) 12.2 11.9 8.49 9.10<br />

σmc 3-point bending (MPa) 14.8 14.3 10.6 11.3<br />

σmc stress-strain fitting (MPa) 15.7 16.1 10.4 12.0<br />

From figure 3.10 <strong>and</strong> table 3.3, it can be noticed that, in contrast <strong>with</strong> the ACK<br />

based theoretical prediction in the pre-cracking zone (zone I), there is already<br />

serious matrix cracking at 5.5MPa for plates RU3 <strong>and</strong> RU4. Ecycle seriously<br />

decreases when σc max = 5.5MPa (first cycle), whereas multiple cracking stress σmc<br />

is about 10MPa. For the cycle <strong>with</strong> σc max = 11MPa, the value <strong>of</strong> Ecycle decreases<br />

considerably for all plates. For plates RU1 <strong>and</strong> RU2, this means serious<br />

degradation <strong>of</strong> the laminates occurred below the theoretical ACK multiple<br />

cracking stress <strong>of</strong> about 15MPa, listed in table 3.3, has been reached.<br />

The ACK based theory is now used for prediction <strong>of</strong> the evolution <strong>of</strong> Ecycle as<br />

function <strong>of</strong> σc max for each plate. Equation (3.14) or (3.21) should be used,<br />

depending on the fact whether partial or full matrix-fibre unloading slip occurs<br />

during unloading. According to equation (3.15) partial matrix-fibre unloading slip<br />

is replaced by total matrix-fibre unloading slip, once (σc max - σc min ) > 1.337σmc.<br />

The value <strong>of</strong> the maximum cycle stress σc max at which this occurs, is listed in table<br />

3.4 (<strong>with</strong> σc min = 1.0MPa).<br />

Table 3.4: σc max at which partial matrix-fibre slip is replaced by total matrix-fibre unloading slip<br />

plate nr RU1 RU2 RU3 RU4<br />

Vf (%) 9.42 11.6 7.53 10.6<br />

σmc (stress-strain fitting) (MPa) 15.7 16.1 10.4 12.0<br />

σc max partial to total matrix-fibre slip (MPa) 21.5 22.0 14.4 16.5<br />

Figures 3.11a to 3.11d show the experimental evolution <strong>of</strong> Ecycle versus σc max for<br />

all laminates, together <strong>with</strong> the theoretical evolution predicted <strong>with</strong> the ACK based<br />

theory.<br />

89


Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

In all figures (3.11a to 3.11d) there is a large discrepancy between the theoretical<br />

<strong>and</strong> experimental curves below the theoretical multiple cracking stress. In figure<br />

3.11a <strong>and</strong> figure 3.11b the theoretical ACK curve <strong>of</strong> plates RU1 <strong>and</strong> RU2 <strong>and</strong> the<br />

experimentally obtained curve coincide at 16.5MPa <strong>and</strong> higher. For plates RU3<br />

<strong>and</strong> RU4, good coincidence is obtained between the theoretical <strong>and</strong> experimental<br />

evolution <strong>of</strong> Ecycle at <strong>and</strong> above 11MPa <strong>and</strong> 16.5MPa respectively.<br />

cycle (GPa)<br />

E<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 5 10 15 20 25 30<br />

max<br />

σc (MPa)<br />

experimental<br />

ACK theory<br />

Figure 3.11a: Ecycle versus σc max , plate RU1, ACK<br />

theory<br />

cycle (GPa)<br />

E<br />

20<br />

15<br />

10<br />

5<br />

0<br />

experimental<br />

ACK theory<br />

0 5 10 15<br />

max<br />

σc (MPa)<br />

20 25 30<br />

Figure 3.11c: Ecycle versus σc max , plate RU3,<br />

ACK theory<br />

Ecycle (GPa)<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 5 10 15 20 25 30 35<br />

max<br />

σc (MPa)<br />

experimental<br />

ACK theory<br />

Figure 3.11b: Ecycle versus σc max , plate RU2, ACK<br />

theory<br />

Ecycle (GPa)<br />

20<br />

15<br />

10<br />

5<br />

0<br />

experimental<br />

ACK theory<br />

0 5 10 15 20 25 30<br />

max<br />

σx (MPa)<br />

Figure 3.11d: Ecycle versus σc max , plate RU4,<br />

ACK theory<br />

From figures 3.11a to 3.11d, it can be noticed that, once full multiple cracking<br />

occurred (zone III), a good coincidence <strong>of</strong> the predicted <strong>and</strong> measured unloading<br />

behaviour is obtained for all tested laminates. When σc max is situated below this<br />

stress zone, (in zone I <strong>and</strong> zone II) the ACK based model fails to predict the value<br />

<strong>of</strong> Ecycle. For example, a theoretical overestimation <strong>of</strong> Ecycle <strong>of</strong> almost 100% is<br />

found in cycle number 2 <strong>of</strong> plate RU1 (figure 3.11a), when σc max is about 10MPa.<br />

The experimental value <strong>of</strong> Ecycle is measured to be slightly lower than 10GPa,<br />

whilst the theoretical obtained value is still about 18 GPa.<br />

An efficiency factor K has been introduced in paragraph 2.4.4 for UD-reinforced<br />

specimens <strong>and</strong> in paragraph 2.5.2 for 2D-r<strong>and</strong>omly reinforced specimens. This<br />

efficiency factor K represents the ratio <strong>of</strong> the experimental on the theoretical<br />

stiffness in zone III under monotonic tensile loading. The values <strong>of</strong> K have been<br />

90


Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

determined for laminates RU1 to RU4 <strong>and</strong> are listed in table 3.3. A similar<br />

efficiency factor is now defined for unloading: Ku. Ku is the ratio <strong>of</strong> the<br />

experimentally obtained stiffness Ecycle on the value predicted by the ACK based<br />

theory.<br />

Ecycle(<br />

experimental)<br />

Ku<br />

E ( theoretical<br />

ACK)<br />

= (3.57)<br />

cycle<br />

The evolution <strong>of</strong> Ku <strong>with</strong> σc max , is listed in table 3.5 for plate RU1. The other<br />

laminates give similar results.<br />

Table 3.5: unloading efficiency factor, 2D-r<strong>and</strong>omly reinforced IPC, plate RU1<br />

2D-r<strong>and</strong>omly reinforced IPC<br />

σc max<br />

(MPa)<br />

Ku<br />

(-)<br />

16.5 0.99<br />

22 0.95<br />

27.5 0.92<br />

One can see that the value <strong>of</strong> Ku decreases slightly as σc max increases. The value <strong>of</strong><br />

Ku is considerably higher than K (table 3.3). One would expect that Ku would be<br />

considerably lower than 1.0, since phenomena like fibre pull-out, loss <strong>of</strong> matrixfibre<br />

interface properties from repeated loading (the specimens are already cycled<br />

five times before stress level 27.5MPa is reached) would lead to lower<br />

experimentally obtained stiffness Ecycle.<br />

As a comparative test case, the evolution <strong>of</strong> Ku <strong>of</strong> a UD-reinforced laminate is also<br />

tested <strong>and</strong> discussed. A UD-reinforced laminate is made (UDU1), <strong>with</strong> a fibre<br />

volume fraction Vf <strong>of</strong> 15.7%. Three specimens <strong>of</strong> 18x250mm² are cut <strong>and</strong><br />

subjected to a simple tensile test up to failure. From these tests an average value<br />

for K <strong>of</strong> 0.93 <strong>and</strong> for σmc <strong>of</strong> 11MPa are obtained from minimisation <strong>of</strong> the LSC<br />

(cfr. paragraph 2.4.8.2) between 0 <strong>and</strong> 40MPa. Since continuous unidirectional<br />

fibres are used, effects <strong>of</strong> fibre pull-out, fibre bending, etc. are less likely to occur.<br />

One specimen (UDU1-4) is now subjected to limited cycling. σc max <strong>of</strong> each load<br />

cycle is 6.0MPa higher than the previous one. Unloading is performed to a<br />

minimum stress σc min = 1.0MPa. The stress-strain curve obtained this way is<br />

similar to the one presented in figure 3.9. The experimental <strong>and</strong> theoretical<br />

evolution <strong>of</strong> Ecycle are both determined as has been done for the 2D-r<strong>and</strong>omly<br />

reinforced specimens earlier. In table 3.6 the evolution <strong>of</strong> Ku <strong>of</strong> specimen UDU-4<br />

versus σc max is listed.<br />

As can be seen from table 3.6, the unloading factor Ku is higher than one for all<br />

unloading stresses. This means there are some phenomena, leading to stiffening <strong>of</strong><br />

the composite during unloading. One small specimen <strong>of</strong> 9x50mm² is cut from<br />

laminate UDU1: UDU1-5. This specimen is cycled on a micro-testing bench in a<br />

91


Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

SEM (Scanning Electron Microscope) microscopic unit. This way, the surface <strong>of</strong><br />

the specimen can be visualised during loading <strong>and</strong> unloading. Tests on the microtesting<br />

bench in the electron microscope revealed that matrix particles are<br />

sometimes ripped <strong>of</strong>f from the crack face. This way, small matrix particles are left<br />

in <strong>and</strong> around the crack, sometimes preventing it to close again. This effect is<br />

illustrated in figure 3.12.<br />

Table 3.6: unloading efficiency factor Ku for unidirectionally reinforced IPC<br />

Unidirectionally reinforced IPC<br />

σc max<br />

(MPa)<br />

Ku<br />

(-)<br />

18 1.35<br />

24 1.26<br />

30 1.23<br />

36 1.20<br />

42 1.20<br />

48 1.13<br />

60 1.13<br />

90 1.06<br />

Figure 3.12: SEM picture <strong>of</strong> crumbling matrix <strong>of</strong> UD-reinforced IPC: UDU1-5<br />

In the oval shape in figure 3.12, a small particle <strong>of</strong> crumbled matrix can be seen in<br />

the vicinity <strong>of</strong> the crack. In the black circle a new matrix particle, situated at the<br />

crack face, is about to be ripped <strong>of</strong>f from the composite. This type <strong>of</strong> particles can<br />

prevent the crack from closing at unloading.<br />

It is the competition between the stiffening (prevented crack closure) <strong>and</strong> the<br />

degradation (matrix-fibre interface degradation, fibre pullout, etc.) effects that<br />

determine the value <strong>of</strong> Ku. Since the experimental value <strong>of</strong> Ku is smaller for 2Dr<strong>and</strong>omly<br />

reinforced specimens than for UD-reinforced laminates, degradation is<br />

more pronounced or crack closure is less prevented for 2D-r<strong>and</strong>omly reinforced<br />

92


Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

specimens. The degradation effects are further discussed in Chapter 4. It has also<br />

been illustrated that Ku may be considerably higher than one for UD-reinforced<br />

specimens.<br />

3.6.4 unloading behaviour as predicted by the stochastic cracking<br />

based model<br />

In this paragraph, the stochastic cracking based theory is used to predict the<br />

evolution <strong>of</strong> Ecycle versus σc max . First, material parameters τ0, σR <strong>and</strong> m are to be<br />

determined. The values <strong>of</strong> these parameters are obtained for each plate from<br />

comparison <strong>of</strong> experimental stress-strain curves under monotonic tensile loading<br />

<strong>with</strong> theoretical curves.<br />

τ0, σR <strong>and</strong> m are varied as described in paragraph 2.5.4 to find a “best fit” <strong>of</strong> the<br />

theoretical <strong>and</strong> experimental stress-strain loading curves. The resulting parameters<br />

are listed in table 3.7. f is determined from visual crack counting under the<br />

stereomicroscope.<br />

Table 3.7: parameters obtained from fitting <strong>of</strong> the stochastic cracking based theory <strong>with</strong><br />

experimental stress-strain curves under monotonic tensile loading<br />

plate 1 plate2 plate3 plate4<br />

Vf (%) 9.4 11.6 7.5 10.6<br />

f (mm) 1.1 0.95 1.2 1.0<br />

m (-) 4.8 4.1 4.7 4.3<br />

σR (MPa) 15 15 9.5 10<br />

τ0 (MPa) 1.3 1.0 1.3 1.1<br />

The material parameters, listed in table 3.7, are used in equation (3.33) or (3.56) to<br />

predict the evolution <strong>of</strong> Ecycle <strong>with</strong> σc max . The program IPCstress-strain.exe is used<br />

for this purpose. The experimental evolutions <strong>of</strong> Ecycle as a function <strong>of</strong> σc max are<br />

printed together <strong>with</strong> the theoretical evolutions, as predicted by the stochastic<br />

cracking based theory, in figures 3.13a to 3.13d.<br />

cycle (GPa)<br />

E<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

experiment<br />

stochastic theory<br />

0 5 10 15 20 25 30<br />

max<br />

σc (MPa)<br />

Figure 3.13a: Ecycle versus σc max , plate RU1,<br />

stochastic cracking based theory<br />

E cycle (GPa)<br />

93<br />

20<br />

15<br />

10<br />

5<br />

0<br />

experimental<br />

stochastic theory<br />

0 10 20 30<br />

max<br />

σc (MPa)<br />

Figure 3.13b: Ecycle versus σc max , plate RU2,<br />

stochastic cracking based theory<br />

40


)<br />

cycle (GPa<br />

E<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

experimental<br />

stochastic theory<br />

0 5 10 15<br />

max<br />

σc (MPa)<br />

20 25 30<br />

Figure 3.13c: Ecycle versus σc max , plate RU3,<br />

stochastic cracking based theory<br />

E cycle (GPa)<br />

20<br />

15<br />

10<br />

5<br />

0<br />

experimental<br />

stochastic theory<br />

0 5 10 15 20 25 30<br />

max<br />

σ c (MPa)<br />

Figure 3.13d: Ecycle versus σc max , plate RU4,<br />

stochastic cracking based theory<br />

From figures 3.13a to 3.13d, it can be seen for all tested laminates that the<br />

linearised E-modulus <strong>of</strong> each cycle (Ecycle) is predicted rather well <strong>with</strong> the<br />

stochastic cracking based model.<br />

3.6.5 ACK based model versus stochastic cracking based model<br />

Four 2D-r<strong>and</strong>omly reinforced laminates have been tested. The ACK based theory<br />

<strong>and</strong> the stochastic cracking based theory have been compared <strong>with</strong> the<br />

experimental evolution <strong>of</strong> Ecycle. The ACK based theory for unloading is an<br />

appropriate model to predict Ecycle as a function <strong>of</strong> σc max , when σc max is situated in<br />

zone III. Where the ACK theory fails to predict the unloading behaviour at lower<br />

stress levels (pre-cracking <strong>and</strong> multiple cracking zones), the stochastic cracking<br />

based model seems to overcome this problem. The stochastic cracking based<br />

model is verified on four 2D-r<strong>and</strong>omly reinforced laminates, containing different<br />

fibre volume fractions <strong>and</strong> matrix quality.<br />

3.7 Conclusions<br />

- The two models presented in Chapter 2, to describe the stress-strain behaviour <strong>of</strong><br />

UD-reinforced <strong>and</strong> 2D-r<strong>and</strong>omly reinforced IPC composite specimens under<br />

monotonic tensile loading, are extended in this chapter to predict unloading <strong>of</strong> IPC<br />

composite specimens. The evolution <strong>of</strong> Ecycle, which is the linearised E-modulus <strong>of</strong><br />

one cycle, is formulated as function <strong>of</strong> the maximum composite cycle stress σc max .<br />

This variable will be inserted later into finite element calculations. The evolution<br />

<strong>of</strong> Ecycle, according to both models, is compared <strong>with</strong> observations on specimens <strong>of</strong><br />

four different laminates.<br />

- It has been mentioned in Chapter 2 that not all material parameters are known<br />

accurately a priori. In case the ACK (based) theory is used, the matrix failure<br />

94


Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

stress σmu <strong>and</strong> efficiency factor K are not always known a priori. In case the<br />

stochastic cracking (based) theory is used, the reference cracking stress (σR), the<br />

Weibull modulus (m) <strong>and</strong> the frictional matrix-fibre interface shear stress (τ0) are<br />

difficult to obtain a priori. In Chapter 2 it has been mentioned that these<br />

parameters can be obtained from “best-fitting” <strong>of</strong> the theoretical stress-strain curve<br />

<strong>with</strong> an experimental curve under monotonic tensile loading.<br />

- One goal <strong>of</strong> this chapter is to check whether the value <strong>of</strong> σmu, obtained by “best<br />

fit” <strong>of</strong> the experimental <strong>and</strong> theoretical stress-strain curve under loading (see<br />

Chapter 2), can be inserted into a ACK (based) model for prediction <strong>of</strong> the<br />

unloading behaviour. In a similar way σR, m <strong>and</strong> τ0 are determined from loading<br />

experiments. These values are then inserted into a stochastic cracking based<br />

unloading model.<br />

- The ACK (based) theory can be used to predict the macro-mechanical behaviour<br />

<strong>of</strong> a E-glass fibre reinforced IPC laminate during unloading, when the maximum<br />

composite stress is higher than the theoretical ACK multiple cracking stress.<br />

- In contrast <strong>with</strong> what is assumed by the ACK (based) theory, the multiple<br />

cracking zone is not limited to one stress value in the stress-strain curve. This is<br />

noticed in the experimentally obtained evolution <strong>of</strong> Ecycle as a function <strong>of</strong> the<br />

maximum cycle stress. Multiple cracking starts notably below the theoretical ACK<br />

multiple cracking stress. The ACK based model for unloading seriously (up<br />

overestimates Ecycle (up to 100%), provided the maximum cycle stress is situated at<br />

or below the theoretical ACK multiple cracking stress.<br />

- It has been mentioned that the ACK (based) model is not suited to predict the<br />

unloading stiffness <strong>of</strong> a IPC laminate in the pre-cracking or multiple cracking<br />

zones. The stochastic cracking (based) model can be used in these zones. If the<br />

material <strong>and</strong> interface properties (m, σR <strong>and</strong> τ0) are obtained under monotonic<br />

tensile loading, the knowledge <strong>of</strong> these material parameters can be used to predict<br />

the unloading behaviour in an accurate way for all values <strong>of</strong> the maximum<br />

composite cycle stress. This has been verified on 2D-r<strong>and</strong>omly reinforced<br />

specimens.<br />

3.8 References<br />

B.K. Ahn <strong>and</strong> W.A Curtin, Strain <strong>and</strong> hysteresis by stochastic matrix<br />

cracking in ceramic matrix composites, J. Mech. Phys. Solids, Vol. 45, No. 2,<br />

1997 pp.177-209<br />

95


Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

H.G. Allen, The purpose <strong>and</strong> methods <strong>of</strong> fibre reinforcement, In Prospects<br />

<strong>of</strong> Fibre Reinforced Construction Materials, Proc. Int. Building Exhibition<br />

Conference, Building Research Station, UK, 1971, pp.3-14<br />

J. Aveston, G.A. Cooper <strong>and</strong> A Kelly, Single <strong>and</strong> multiple fracture, The<br />

Properties <strong>of</strong> Fibre Composites, Proc. Conf. National Physical Laboratories, IPC<br />

Science & Technology Press Ltd. London, 1971, pp.15-24<br />

J. Aveston, R.A. Mercer, J.M. Sillwood, Fibre reinforced cements –<br />

scientific foundations for specifications, In Composites – St<strong>and</strong>ards, Testing <strong>and</strong><br />

<strong>Design</strong>, Proc. National Physical Laboratories Conference, UK, 1974, pp.93-103<br />

P. Bauweraerts, Aspects <strong>of</strong> the Micromechanical Characterisation <strong>of</strong> Fibre<br />

Reinforced <strong>Brittle</strong> <strong>Matrix</strong> Composites, Phd. thesis, VUB 1998<br />

H. Cuypers, The behaviour <strong>of</strong> E-glass fibre reinforced brittle matrix<br />

composite subjected to increasing cyclic loading, internal report, department<br />

MEMC, Vrije Universiteit Brussel, December 1999<br />

H. Cuypers, J. Gu, K. Croes, S. Dumortier, J. Wastiels, Evaluation <strong>of</strong><br />

fatigue <strong>and</strong> durability properties <strong>of</strong> E-glass fibre reinforced phosphate<br />

cementitious composites, Proc. Int. Symp. <strong>Brittle</strong> <strong>Matrix</strong> Composites 6, A.M.<br />

Br<strong>and</strong>t, V.C. Li, I.H. Marshall, Warsaw, October 9-11, 2000<br />

H. Cuypers, Use <strong>of</strong> a stochastic cracking based model on the behaviour <strong>of</strong><br />

2D-r<strong>and</strong>om reinforced IPC composite specimens, internal report, department<br />

MEMC, Vrije Universiteit Brussel, 2001<br />

J. Gu, X. Wu, H. Cuypers <strong>and</strong> J. Wastiels, Modeling <strong>of</strong> the tensile<br />

behaviour <strong>of</strong> an E-glass fibre reinforced phosphate cement, Computer Methods in<br />

Composite Materials VI, proceedings CADCOMP 98, 1998, pp.589-598<br />

J.G. Keer, Behaviour <strong>of</strong> cracked fibre composites under limited cyclic<br />

loading, Int. J. Cem. Comp. & Ltwt. Concr., Vol. 3, 1981, pp.179-186<br />

J.G. Keer, Some observations on hysteresis effects in fibre cement<br />

composites, J. Mat. Sci. Letters, Vol. 4, 1985, pp.363-366<br />

V. Laws, The efficiency <strong>of</strong> fibrous reinforcement <strong>of</strong> brittle matrices, J. Phys.<br />

D. Appl. Phys., Vol. 4, 1971, pp.1737-1746<br />

96


Chapter 3: Unloading <strong>of</strong> IPC composite specimens<br />

A.W. Pryce <strong>and</strong> P.A. Smith, <strong>Matrix</strong> cracking in unidirectional ceramic<br />

composites under quasi-static <strong>and</strong> cyclic loading, Acta metall. mater., No. 41, 1993,<br />

pp.1269-1281<br />

D. Rouby <strong>and</strong> P. Reynaud, Fatigue behaviour related to interface<br />

modification during load cycling in ceramic-matrix fibre composites, Composite<br />

Science <strong>and</strong> Technology, Vol. 48, 1993, pp.109-118<br />

97


Chapter 4<br />

4.1 Introduction<br />

IPC composite specimens under<br />

repeated tensile loading<br />

S<strong>and</strong>wich<br />

panels, which<br />

are used as ro<strong>of</strong> or wall cladding, are loaded repeatedly<br />

under wind, snow <strong>and</strong> temperature loads. Since the concept <strong>of</strong> s<strong>and</strong>wich<br />

construction leads to the use <strong>of</strong> lightweight panels, the value <strong>of</strong> the variable load<br />

may be quite larger than the dead load. These construction elements should<br />

therefore be analysed for fatigue, if severe accumulation <strong>of</strong> damage is to be<br />

expected. Main factors contributing to accumulated damage <strong>and</strong> fatigue failure<br />

are:<br />

- the number <strong>of</strong> applied load cycles<br />

- the mean stress, experienced in each load cycle<br />

- the load cycle amplitude<br />

- the presence <strong>of</strong> local stress concentrations<br />

Phenomena, leading to<br />

fatigue failure in composites, are completely different from<br />

those <strong>of</strong> conventional materials. Fatigue failure occurs due to initiation <strong>and</strong><br />

propagation <strong>of</strong> one single crack for most conventional materials, whereas several<br />

damage mechanisms can interact <strong>and</strong> lead to final failure under repeated loading<br />

for composites. In Chapter 2, it has been verified that it can be assumed that the<br />

stress-strain behaviour <strong>of</strong> IPC composites in compression is linear elastic.<br />

However, IPC composites under tensile loading experience damage introduction at<br />

low stress levels. Therefore, further accumulation <strong>of</strong> damage under repeated<br />

tensile loading <strong>of</strong> IPC composite specimens is discussed here. Initiation <strong>and</strong><br />

propagation <strong>of</strong> damage in a brittle-matrix composite are mainly determined by<br />

following phenomena:<br />

- matrix cracking<br />

- fibre failure<br />

- fibre pull-out<br />

- degradation <strong>of</strong> the matrix-fibre interface<br />

99


Chapter 4: IPC composite specimens under repeated tensile loading<br />

The main parameters, which might influence the relative importance <strong>of</strong> the<br />

damage mechanisms, are fibre volume fraction, matrix strength <strong>and</strong> stiffness,<br />

matrix-fibre interface properties, mean stress level, load cycle amplitude <strong>and</strong> load<br />

history.<br />

A modified ACK model <strong>and</strong> modified stochastic cracking model are used in this<br />

chapter to formulate the evolution <strong>of</strong> the deformations <strong>and</strong> the stiffness <strong>of</strong> UDreinforced<br />

<strong>and</strong> 2D-r<strong>and</strong>omly reinforced IPC composites as a function <strong>of</strong> the<br />

number <strong>of</strong> elapsed load cycles in tension. Theoretical formulations, describing<br />

accumulation <strong>of</strong> deformations due to cyclic loading, are based on the formulation<br />

<strong>of</strong> a single load cycle, as discussed in Chapter 3. These theoretical formulations<br />

are compared <strong>with</strong> the observations on UD-reinforced <strong>and</strong> 2D-r<strong>and</strong>omly<br />

reinforced IPC specimens under repeated tensile loading.<br />

4.2 Theoretical derivation: general remarks<br />

4.2.1 introduction<br />

Formulations for the description <strong>of</strong> the stress-strain behaviour <strong>of</strong> ceramic matrix<br />

composites under repeated loading are derived <strong>and</strong>/or discussed by several<br />

authors.<br />

Rouby <strong>and</strong> Reynaud (1992) discuss the behaviour <strong>of</strong> UD-reinforced ceramic<br />

composites under repeated loading: the behaviour <strong>of</strong> SiC fibre (Nicalon) bundles<br />

embedded in a SiC matrix is studied. According to Rouby <strong>and</strong> Reynaud (1992),<br />

full matrix multiple cracking occurs <strong>and</strong> some fibres break at first loading. The<br />

ACK theory is used to describe the distance between the matrix cracks. It is<br />

assumed by Rouby <strong>and</strong> Reynaud (1992) that further degradation <strong>of</strong> the composite<br />

after first loading (cycle number one) is due to the decreasing frictional shear<br />

stress at the matrix-fibre interface, as a result <strong>of</strong> “interfacial degradation”. Rouby<br />

<strong>and</strong> Reynaud (1992) conclude that decreasing matrix-fibre shear stress leads to<br />

increasing failure probability <strong>of</strong> the fibres <strong>and</strong> a widening <strong>of</strong> the hysteresis loop.<br />

Evans et al. (1994) adopt most <strong>of</strong> the basic assumptions, formulated by Rouby <strong>and</strong><br />

Reynaud (1992). <strong>Matrix</strong> multiple cracking occurs during first loading. After first<br />

loading, degradation <strong>of</strong> the frictional matrix-fibre interface is the major damage<br />

mechanism under further repeated loading. No additional matrix cracks are<br />

introduced under repeated loading, after first loading. While Rouby <strong>and</strong> Reynaud<br />

(1992) determined the average crack spacing <strong>with</strong> help <strong>of</strong> the ACK theory, Evans<br />

et al. (1994) propose the average crack spacing should be determined by visual<br />

crack counting. The advantage <strong>of</strong> the approach <strong>of</strong> Evans et al. (1994) is found in<br />

the fact that the behaviour <strong>of</strong> composite specimens under partial multiple cracking<br />

can be discussed. This was impossible in the approach <strong>of</strong> Rouby <strong>and</strong> Reynaud<br />

100


Chapter 4: IPC composite specimens under repeated tensile loading<br />

(1992). A fatigue methodology is presented by Evans et al. (1994), in which the<br />

evolution <strong>of</strong> the frictional matrix-fibre interface shear stress is to be determined<br />

from low-cycle fatigue experiments. The obtained evolution <strong>of</strong> the interface shear<br />

stress can then be introduced to predict S-N curves <strong>of</strong> laminates under different<br />

conditions (fibre volume fraction, stress amplitude, fibre diameter, etc.) Limited<br />

experimental verification showed that some extra matrix cracks might be<br />

introduced under repeated loading, after first loading. It is reported that this<br />

phenomenon only occurs during the first ten load cycles. Evans et al. (1994) also<br />

mention that the matrix fibre “interface degradation” occurs by reduction in the<br />

height asperities along the fibre coating in some MMCs (metal matrix<br />

composites). Direct observations <strong>of</strong> related effects have not been performed on<br />

CMCs (ceramic matrix composites) yet.<br />

Curtin (1999) follows the same approach as Evans et al. (1994), but takes into<br />

account that the average crack spacing can be formulated in function <strong>of</strong> the<br />

maximum composite cycle stress analytically. The matrix failure stress is assumed<br />

to obey a Weibull probability distribution function.<br />

The mentioned publications are used in this chapter as basic documents to model<br />

<strong>and</strong> discuss the behaviour <strong>of</strong> IPC composite specimens.<br />

4.2.2 assumptions <strong>and</strong> definitions<br />

Definitions <strong>of</strong> stresses, strains <strong>and</strong> stiffness, used to describe the behaviour <strong>of</strong> Eglass<br />

fibre reinforced IPC specimens under repeated load, are plotted in figure 4.1.<br />

Cycling is performed between a fixed maximum (σc max ) <strong>and</strong> minimum stress<br />

(σc min ). The amplitude <strong>of</strong> the stress cycles is ∆σc = (σc max - σc min ). The slope <strong>of</strong> the<br />

straight line which connects the points (σc min , εc,N min ) <strong>and</strong> (σc max , εc,N max ) is defined<br />

as the linearised E-modulus <strong>of</strong> the N th cycle: Ecycle,N. Based on the documents <strong>of</strong><br />

Rouby <strong>and</strong> Renaud (1991), Evans et al. <strong>and</strong> Curtin (1999), it is assumed that the<br />

frictional matrix-fibre interface shear stress decreases <strong>with</strong> increasing number <strong>of</strong><br />

elapsed load cycles. This assumption is discussed later for the studied composites<br />

(see experimental part <strong>of</strong> this chapter). After N load cycles are applied, the matrixfibre<br />

interface shear stress is τN, <strong>with</strong> τN ≤ τ0.<br />

The composite properties εc,N max , εc,N min , τN <strong>and</strong> Ecycle,N are all function <strong>of</strong> the<br />

number <strong>of</strong> elapsed cycles load cycles, N.<br />

The relationship between εc,N min , Ecycle,N <strong>and</strong> εc,N max is expressed in equation (4.1).<br />

max min<br />

min max ( σ c − σ c ) max ∆σ<br />

c<br />

εc,<br />

N εc,<br />

N −<br />

= εc,<br />

N −<br />

E cycle,<br />

N<br />

Ecycle,<br />

N<br />

= (4.1)<br />

101


Chapter 4: IPC composite specimens under repeated tensile loading<br />

The ratio <strong>of</strong> the actual (N th load cycle) versus initial frictional matrix-fibre<br />

interface shear stress is ω :<br />

τ N ω = (4.2)<br />

τ<br />

σc<br />

(σc max , εc,1 max ) (σc max , εc,N max )<br />

first<br />

cycle<br />

Ecycle,1<br />

(σc min , εc,1 min )<br />

0<br />

N th<br />

cycle<br />

Ecycle,N<br />

(σc min , εc,N min )<br />

σc max = maximum cycle stress<br />

σc min = minimum cycle stress<br />

εc,1 max = maximum cycle strain, cycle 1<br />

εc,1 min = minimum cycle strain, cycle 1<br />

εc,N max = maximum cycle strain, cycle N<br />

εc,N min = minimum cycle strain, cycle N<br />

Ecycle,1 = linearised E-modulus, cycle 1<br />

Ecycle,N = linearised E-modulus, cycle N<br />

N = number <strong>of</strong> applied load cycles<br />

Figure 4.1: stress, strain <strong>and</strong> stiffness notations used in modelling <strong>of</strong> the behaviour <strong>of</strong> E-glass<br />

fibre reinforced IPC under repeated loading<br />

All basic assumptions, mentioned in Chapter 2 <strong>and</strong> Chapter 3, are again used to<br />

model the behaviour <strong>of</strong> IPC composites under repeated loading. At present, three<br />

extra hypotheses are introduced. These will be discussed <strong>and</strong>/or verified<br />

experimentally later. Rouby <strong>and</strong> Reynaud (1992), Evans et al. (1994) <strong>and</strong> Curtin<br />

(1999) also use or discuss these assumptions in their formulation <strong>of</strong> the behaviour<br />

<strong>of</strong> CMCs under repeated loading.<br />

1. Introduction <strong>of</strong> matrix cracking occurs at first loading <strong>of</strong> the specimens, no<br />

matter if the fatigue model is based on the ACK (based) theory or on the stochastic<br />

cracking (based) theory. Further cycling after initial loading does not introduce<br />

extra matrix cracking. The average crack spacing is function <strong>of</strong> σc max , but is<br />

independent from the number <strong>of</strong> applied load cycles.<br />

102<br />

εc


Chapter 4: IPC composite specimens under repeated tensile loading<br />

2. During repeated cycling, the frictional matrix-fibre interface suffers<br />

“interface degradation”. The frictional matrix-fibre interface shear stress decreases<br />

<strong>with</strong> increasing number <strong>of</strong> cycles. This degradation effect is most pronounced at<br />

the first few cycles. Accumulation <strong>of</strong> residual deformations occurs due to this<br />

matrix-fibre interface degradation.<br />

3. At present, one extra hypothesis will be used. Fibre breakage <strong>and</strong> pull-out<br />

are not considered. The consequences <strong>of</strong> this hypothesis will be discussed later.<br />

Before the theoretical derivation <strong>of</strong> the stress-strain behaviour <strong>of</strong> IPC composites<br />

under repeated load is presented, it is stressed that the reader should be aware <strong>of</strong><br />

the fact that the “matrix-fibre interface degradation” phenomenon as used here<br />

includes many micro-mechanical phenomena. As has been mentioned in paragraph<br />

2.4.8, the matrix-fibre interface shear stress τ0 is a bundle property rather than a<br />

single fibre property. <strong>Matrix</strong>-fibre stress transfer is probably rather high for the<br />

outer fibres <strong>and</strong> lower for those fibres, situated in the middle <strong>of</strong> the bundle. Inside<br />

the fibre bundle fibre-fibre stress transfer might also exist. τ0 is thus only an<br />

averaged material property, describing stress transfer from matrix to fibres. In a<br />

similar way ω is a parameter describing the average loss <strong>of</strong> frictional matrix-fibre<br />

interface stress transfer <strong>and</strong> fibre-fibre stress transfer in <strong>and</strong> around a fibre bundle.<br />

4.3 Theoretical derivation <strong>of</strong> a fatigue theory from the ACK<br />

(based) model<br />

According to the ACK (based) theory, no matrix cracking is initiated in the<br />

specimens as long as the multiple cracking stress σmc is not reached. A fatigue<br />

theory based on the ACK theory, which considers that frictional matrix-fibre<br />

interface degradation is the sole damage mechanism, can thus only be expressed<br />

for damage accumulation in zone III (post-cracking zone). At present, formation<br />

<strong>of</strong> extra matrix cracks <strong>and</strong> fibre pull-out or breakage under repeated loading are<br />

not considered <strong>and</strong> degradation <strong>of</strong> the matrix-fibre interface is considered to be the<br />

only fatigue mechanism. This assumption is studied experimentally later.<br />

4.3.1 evolution <strong>of</strong> εc,N max <strong>with</strong> N<br />

The formulation <strong>of</strong> εc,N max as a function <strong>of</strong> internal stresses <strong>and</strong> material properties<br />

is based on figure 4.2a <strong>and</strong> 4.2b. These figures illustrate the evolution <strong>of</strong> the<br />

matrix <strong>and</strong> fibre stresses as a function <strong>of</strong> the distance x from the crack face at a<br />

maximum composite cycle stress σc max for initial loading (N = 1) <strong>and</strong> after N load<br />

cycles are applied respectively.<br />

Figure 4.2a shows the matrix <strong>and</strong> fibre stresses in between two cracks at initial<br />

loading (N=1). As the number <strong>of</strong> elapsed load cycles increases, the frictional<br />

103


Chapter 4: IPC composite specimens under repeated tensile loading<br />

matrix-fibre interface shear stress decreases. At the end <strong>of</strong> the N th load cycle, the<br />

frictional matrix-fibre interface shear stress is τN, which is smaller than τ0. Since<br />

the frictional shear stress τN decreases as the number <strong>of</strong> elapsed cycles increases,<br />

the maximum matrix stress in the middle <strong>of</strong> two cracks σm,N max (x = f/2)<br />

decreases as well, as can be seen in figure 4.2b.<br />

x<br />

matrix<br />

stress<br />

σm,1 max<br />

σf,1 min<br />

fibre<br />

stress<br />

σf,1 max<br />

Figure 4.2a: stresses in matrix <strong>and</strong> fibres at<br />

maximum composite cycle stress, first cycle<br />

σ<br />

N cycles<br />

x<br />

matrix<br />

stress<br />

σm,1 max<br />

σf,1 min<br />

fibre<br />

stress<br />

σf,1 max<br />

Figure 4.2b: stresses in matrix <strong>and</strong> fibres at<br />

maximum composite cycle stress, N th cycle<br />

The expression <strong>of</strong> σm,N max (x = f/2) is derived here. It is assumed that the<br />

average distance between two cracks does not vary <strong>with</strong> the number <strong>of</strong> applied<br />

load cycles, thus equivalent to equation (2.14):<br />

cs f = 1. 337δ<br />

0<br />

(4.3)<br />

<strong>with</strong> (similar to equation (2.13)):<br />

σ mur<br />

Vm<br />

δ 0 =<br />

∗<br />

2τ V<br />

(4.4)<br />

0<br />

f<br />

Debonding length δN increases <strong>with</strong> increasing number <strong>of</strong> elapsed load cycles:<br />

σ mur<br />

Vm<br />

δ N =<br />

∗<br />

(4.5)<br />

2τ<br />

V<br />

N<br />

f<br />

The far field matrix stress is the stress in the matrix at distance δN <strong>of</strong> a crack,<br />

provided the existence <strong>of</strong> other cracks is neglected (see equation (2.39)). If the<br />

ACK theory is used, the far field matrix stress equals σmu. Thus:<br />

∗<br />

2δ<br />

NV<br />

ff<br />

f τ<br />

, max<br />

N<br />

σ m,<br />

N = = σ mu<br />

(4.6)<br />

Vmr<br />

The ratio <strong>of</strong> the maximum matrix stress on the far field matrix stress equals the<br />

ratio <strong>of</strong> the distances from the crack face, at which these stresses are found:<br />

104<br />

σ


Thus:<br />

Chapter 4: IPC composite specimens under repeated tensile loading<br />

σ<br />

σ<br />

max<br />

m,<br />

N<br />

( x =<br />

σ<br />

cs<br />

ff , max<br />

m,<br />

N<br />

1.<br />

337δ<br />

f<br />

/ 2)<br />

cs<br />

=<br />

2δ<br />

N<br />

τ<br />

(4.7)<br />

max<br />

0<br />

N<br />

m,<br />

N = σ mu = 0.<br />

668 σ mu<br />

(4.8)<br />

2δ<br />

N<br />

τ 0<br />

The ratio <strong>of</strong> the actual versus initial frictional matrix-fibre interface shear stress<br />

has been defined as the “degradation parameter” ω (equation (4.2)):<br />

max<br />

σ 0.<br />

668ωσ<br />

m,<br />

N<br />

mu<br />

= (4.9)<br />

The fibre stress at the crack face is:<br />

max σ<br />

σ f , N ( x = 0)<br />

=<br />

V<br />

(4.10)<br />

max<br />

c<br />

∗<br />

f<br />

<strong>and</strong> in the middle <strong>of</strong> two cracks:<br />

σ<br />

σ<br />

668<br />

∗ ∗<br />

− = =<br />

max<br />

f , N ( x cs / 2)<br />

max<br />

c<br />

V f<br />

0.<br />

m ωσ mu<br />

V f<br />

(4.11)<br />

The average fibre stress along the total composite length is then:<br />

∗ ∗<br />

− =<br />

max<br />

σ f N<br />

along cs<br />

f<br />

σ c<br />

Vm<br />

0.<br />

334ωσ<br />

mu<br />

V<br />

V<br />

V<br />

, (4.12)<br />

f<br />

f<br />

The composite strain, after multiple cracking at first loading occurs <strong>with</strong><br />

subsequent degradation <strong>of</strong> the matrix-fibre frictional interface, due to repeated<br />

loading is thus:<br />

∗ ∗<br />

− =<br />

=<br />

=<br />

max<br />

σ f , N<br />

max<br />

along cs<br />

max max<br />

f σ c<br />

Vm<br />

ε c N ε f N<br />

0.<br />

334ωσ<br />

(4.13)<br />

, ,<br />

along cs<br />

mu<br />

f E E V<br />

E V<br />

f<br />

f f<br />

When σc max = σmc, this strain can thus be rewritten:<br />

max<br />

ε c , N = ε mu ( 1+<br />

( 1−<br />

0.<br />

334ω<br />

) α )<br />

EmVm<br />

<strong>with</strong>: α = ∗<br />

E V<br />

f<br />

f<br />

f<br />

f<br />

(4.14)<br />

Equation (4.14) represents the composite strain at maximum composite stress,<br />

provided the maximum composite stress equals the multiple cracking stress (end<br />

<strong>of</strong> zone II). If cycling occurs into the post-cracking zone (zone III), an extra strain<br />

term should be introduced. In zone III, only the fibres provide further stiffness.<br />

Therefore, the strain term in the post-cracking zone is:<br />

max<br />

max<br />

( σ c,<br />

N − σ mc )<br />

εc,<br />

N = ε mu(<br />

1 + ( 1−<br />

0.<br />

334ω)<br />

α ) +<br />

∗<br />

(4.15)<br />

KE V<br />

105<br />

f<br />

f


Chapter 4: IPC composite specimens under repeated tensile loading<br />

When ω = 1, equation (4.14) <strong>and</strong> (4.15) become equivalent to equation (2.87) <strong>and</strong><br />

(2.73) respectively.<br />

4.3.2 evolution <strong>of</strong> Ecycle,N <strong>with</strong> N<br />

Ecycle,N is formulated as a function <strong>of</strong> ω. Different formulations are to be derived<br />

for the cases <strong>of</strong> partial <strong>and</strong> total matrix-fibre slip during unloading. The derivation<br />

<strong>of</strong> Ecycle,N is similar to the derivation <strong>of</strong> Ecycle in paragraph 3.3.2.<br />

It is assumed in this work that the value <strong>of</strong> the frictional matrix-fibre interface<br />

shear stress τN remains constant <strong>with</strong>in one loading cycle <strong>and</strong> decreases stepwise<br />

between the loading cycles. This assumption is not totally correct, but becomes<br />

acceptable after 10 to 100 load cycles, as will be shown later from experimentally<br />

obtained curves (see paragraph 4.11).<br />

4.3.2.1 partial matrix-fibre unloading slip<br />

Figure 4.3a shows how the matrix stresses evolve when the composite is unloaded<br />

from σc max to σc min (see figure 4.1) for the first cycle. Figure 4.3b shows the same<br />

evolution for cycle number N. The total unloading strain variation (∆εc,N) is<br />

expressed as the sum <strong>of</strong> a strain variation under elastic unloading (∆εc,N elastic ) <strong>and</strong> a<br />

term representing fibre slip variation in the matrix in the vicinity <strong>of</strong> the crack<br />

surface (∆εc,N slip ).<br />

x<br />

(su)1<br />

σm<br />

Figure 4.3a: evolution <strong>of</strong> stresses in the matrix<br />

in fully cracked composite from σc max to σc min ,<br />

partial slip, first cycle<br />

N cycles<br />

x<br />

(su)N<br />

σm<br />

Figure 4.3b: evolution <strong>of</strong> stresses in the matrix<br />

in fully cracked composite from σc max to σc min ,<br />

partial slip, N th cycle<br />

In figures 4.3a <strong>and</strong> 4.3b, the dashed lines represent the stresses in the matrix at<br />

maximum composite stress, σc max . The grey full lines show the stresses in the<br />

matrix, if unloading would occur linear elastically. The grey full line shows that<br />

the matrix would then undergo compression at the crack faces. However, since in<br />

reality the matrix is stress free at the crack face, a correction is needed. During<br />

unloading, the matrix slips along the fibres. This matrix-fibre interface unloading<br />

106


Chapter 4: IPC composite specimens under repeated tensile loading<br />

slip shear stress has the same magnitude for loading <strong>and</strong> unloading, but in the<br />

reverse direction. The black curve shows the matrix stress distribution, after<br />

combined elastic unloading <strong>and</strong> partial matrix-fibre slip occurred.<br />

The composite strain interval, due to elastic unloading is:<br />

max min<br />

elastic ( σ c −σ<br />

c ) ∆σ<br />

c<br />

∆ ε c,<br />

N =<br />

=<br />

Ec1<br />

Ec1<br />

(4.16)<br />

If only linear elastic unloading would occur, the matrix will undergo compression<br />

at the crack face. This stress is released by matrix-fibre slip:<br />

slip<br />

elastic Em<br />

∆ σ m,<br />

N ( x = 0)<br />

= ∆σ<br />

m,<br />

N = ∆σ<br />

c<br />

Ec1<br />

<strong>and</strong> at a distance (su)N from the crack:<br />

(4.17a)<br />

slip<br />

∆σ m,<br />

N ( x = ( su)<br />

N ) = 0<br />

(4.17b)<br />

The extra fibre stress variation term, due to matrix-fibre unloading slip at the crack<br />

face is:<br />

slip<br />

elastic Vm<br />

∆σ<br />

f , N ( x = 0)<br />

= ∆σ<br />

m,<br />

N ( x = 0)<br />

= ∆σ<br />

∗ c<br />

V<br />

EmVm<br />

∗<br />

E V<br />

(4.18a)<br />

<strong>and</strong> at a distance (su)N from the crack face:<br />

slip<br />

σ ( x = ( su)<br />

) = 0<br />

(4.18b)<br />

∆ f , N<br />

N<br />

The average unloading slip stress in the fibres along the average crack space f<br />

is:<br />

slip ∆σ<br />

c EmVm<br />

2(<br />

su)<br />

N<br />

∆σ<br />

f , N =<br />

along cs<br />

∗<br />

(4.19)<br />

f E V cs<br />

The composite slip strain is then:<br />

∆<br />

=<br />

∆<br />

2 c1<br />

∆σ<br />

=<br />

f<br />

f<br />

E<br />

V<br />

f<br />

c1<br />

( su)<br />

slip slip<br />

c m m<br />

N<br />

ε c,<br />

N ε f , N<br />

along cs<br />

∗<br />

(4.20)<br />

f 2 E f Ec1V<br />

f cs f<br />

The total composite unloading fibre strain variation ∆εc,N is thus:<br />

∆σ<br />

c ∆σ<br />

c EmVm<br />

( su)<br />

N ∆σ<br />

c ∆σ<br />

c ( su)<br />

N<br />

∆εc, N = +<br />

= + α<br />

∗ (4.21)<br />

E E E V cs E E cs<br />

c1<br />

c1<br />

f<br />

f<br />

f<br />

The unloading slip length (su)N is still not expressed in equation (4.21). From<br />

figure 4.3a <strong>and</strong> 4.3b, it can seen that the ratio <strong>of</strong> ∆σm,N elastic /2 on σm,N max (x =<br />

f/2) equals the ratio <strong>of</strong> the slip length (su)N on half the length <strong>of</strong> the crack<br />

spacing (f/2).<br />

( su)<br />

N<br />

=<br />

cs / 2 σ<br />

f<br />

max<br />

m,<br />

N<br />

∆σ<br />

elastic<br />

m,<br />

N<br />

c1<br />

/ 2<br />

( x =< cs ><br />

Combination <strong>of</strong> equation (4.22) <strong>with</strong> equations (4.17a) <strong>and</strong> (4.9) gives:<br />

107<br />

f<br />

/ 2)<br />

2<br />

c1<br />

f<br />

f<br />

(4.22)


Chapter 4: IPC composite specimens under repeated tensile loading<br />

After inserting (4.23) into (4.21):<br />

( su) N Em∆σ<br />

c<br />

=<br />

cs 4E<br />

σ 0.<br />

668ω<br />

(4.23)<br />

f<br />

c1<br />

mu<br />

( ∆σ<br />

)<br />

2<br />

∆σ<br />

c ∆σ<br />

c EmVm<br />

( su)<br />

N ∆σ<br />

c<br />

c Em<br />

∆εc, N = +<br />

= + α<br />

∗ 2<br />

(4.24)<br />

Ec1 Ec1<br />

E fV<br />

f cs Ec<br />

4(<br />

Ec<br />

) 0.<br />

668ωσ<br />

f 1<br />

1<br />

mu<br />

The linearised E-modulus, Ecycle,N is by definition the ratio <strong>of</strong> the unloading stress<br />

on the unloading strain:<br />

∆σ<br />

c Ec1<br />

Ecycle,<br />

N = =<br />

∆ε<br />

α∆σ<br />

,<br />

cE<br />

c N<br />

m<br />

1+<br />

(4.25)<br />

4Ec1σ<br />

muω0.<br />

668<br />

When ω equals 1, equation (4.25) equals equation (3.14).<br />

4.3.2.2 when partial becomes total matrix-fibre unloading slip<br />

Partial matrix-fibre slip during unloading occurs along a distance (su)N. The<br />

subscript N indicates that the unloading matrix-fibre slip distance is function <strong>of</strong> the<br />

number <strong>of</strong> elapsed load cycles. As illustrated in figures 4.3a <strong>and</strong> 4.3b, the<br />

unloading slip distance (su)N enlarges when a composite is subjected to repeated<br />

loading. The occurrence <strong>of</strong> total matrix-fibre slip is no longer just a function <strong>of</strong> the<br />

stresses, but also <strong>of</strong> the number <strong>of</strong> elapsed load cycles. Specimens might undergo<br />

total matrix-fibre slip during unloading if they experience enough load cycles. A<br />

new formulation should therefore be formulated to express the number <strong>of</strong> load<br />

cycles that the specimen has been subjected to, before partial matrix-fibre<br />

unloading slip changes to total matrix-fibre unloading slip.<br />

Total unloading matrix-fibre slip can be formulated by expressing that the slip<br />

length (su),N equals half the crack spacing, f/2. This occurs when formula<br />

(4.22) equals 1, thus:<br />

elastic<br />

su ∆σ<br />

/ 2<br />

( )<br />

cs<br />

f<br />

N<br />

=<br />

/ 2 σ<br />

max<br />

m,<br />

N<br />

m,<br />

N<br />

( x =< cs ><br />

f<br />

/ 2)<br />

= 1<br />

(4.26)<br />

Inserting equations (4.9) <strong>and</strong> (4.17a) into equation (4.26) gives after rearranging:<br />

∆σ<br />

cEm<br />

ω =<br />

1.<br />

337E<br />

σ<br />

(4.27)<br />

c1<br />

mu<br />

4.3.2.3 total matrix-fibre unloading slip<br />

Figure 4.4a <strong>and</strong> 4.4b illustrate the evolution <strong>of</strong> matrix stresses during unloading<br />

for the first <strong>and</strong> the N th cycle respectively <strong>and</strong> are similar to figure (3.3). The<br />

formulation <strong>of</strong> Ecycle,N under the condition <strong>of</strong> full matrix-fibre unloading slip is<br />

determined from figures 4.4a <strong>and</strong> 4.4b.<br />

108


Chapter 4: IPC composite specimens under repeated tensile loading<br />

x<br />

2σm,1 max<br />

σm<br />

Figure 4.4a: evolution <strong>of</strong> stresses in the matrix<br />

in fully cracked composite at σc max <strong>and</strong> σc min ,<br />

total matrix-fibre slip, first cycle<br />

N cycles<br />

loading<br />

unloading<br />

x<br />

2σm,N max<br />

σm<br />

Figure 4.4b: evolution <strong>of</strong> stresses in the matrix<br />

in fully cracked composite at σc max <strong>and</strong> σc min ,<br />

total matrix-slip, N th cycle<br />

From figure 4.4b, it can be noticed that the matrix unloading stress variation at the<br />

crack face ∆σm(x = 0) is zero. ∆σm(/2) equals 2σm,N max . The maximum fibre<br />

stress variation at x = f/2 <strong>and</strong> <strong>of</strong> minimum fibre stress variation at x = 0 after<br />

unloading <strong>with</strong> full matrix-fibre slip are thus:<br />

∆σ<br />

c<br />

∆σ<br />

f , N ( x = 0)<br />

= ∗<br />

(4.28)<br />

V<br />

∆σ<br />

cs ∆σ<br />

− 2V<br />

σ ( x =< cs > / 2)<br />

ω<br />

f<br />

max<br />

c m m,<br />

N<br />

∆σ<br />

c −1.<br />

337σ<br />

mu Vm<br />

f , N ( ) =<br />

=<br />

∗<br />

∗<br />

(4.29)<br />

2<br />

V f<br />

V f<br />

The average composite strain is:<br />

∆ε<br />

c,<br />

N = ∆ε<br />

f , N<br />

along<<br />

cs><br />

f<br />

∆σ<br />

=<br />

E V<br />

1.<br />

337σ<br />

−<br />

2E<br />

ωVm<br />

(4.30)<br />

<strong>and</strong> the linearised E-modulus Ecycle,N is:<br />

∗<br />

E fV<br />

f<br />

Ecycle,<br />

N =<br />

0.<br />

668ωVmσ<br />

1−<br />

∆σ<br />

f<br />

c<br />

∗<br />

f<br />

c<br />

mu<br />

mu<br />

∗<br />

f V f<br />

4.4 Theoretical derivation <strong>of</strong> a fatigue theory from the<br />

stochastic cracking (based) model<br />

(4.31)<br />

In the stochastic cracking (based) model, is function <strong>of</strong> the applied maximum<br />

composite load, σc max . In paragraph 4.2.2 it has been mentioned that one<br />

assumption is made on : no extra matrix cracking occurs after first loading.<br />

109


Chapter 4: IPC composite specimens under repeated tensile loading<br />

This implies that is not a function <strong>of</strong> the number <strong>of</strong> cycles <strong>and</strong> will thus not<br />

change under repeated loading. This assumption will be verified experimentally<br />

later. Degradation <strong>of</strong> the frictional matrix-fibre interface is the only damage<br />

mechanism to be considered in the theoretical derivations here. The expression in<br />

equation (4.32) <strong>of</strong> equals equation (2.38):<br />

cs<br />

=<br />

cs<br />

f<br />

⎛ ⎡<br />

⎜ ⎛ σ<br />

1−<br />

exp⎢−<br />

⎜ ⎜<br />

⎝<br />

⎢<br />

⎣ ⎝ σ R<br />

max<br />

c<br />

⎞<br />

⎟<br />

⎠<br />

m<br />

⎤⎞<br />

⎥<br />

⎟<br />

⎥<br />

⎟<br />

⎦⎠<br />

−1<br />

(4.32)<br />

4.4.1 evolution <strong>of</strong> εc,N max <strong>with</strong> N<br />

4.4.1.1 crack spacing exceeds twice the debonding length ( > 2δN)<br />

Figures 4.5a <strong>and</strong> 4.5b show how the internal stresses in the matrix <strong>and</strong> fibres<br />

change at σc max <strong>with</strong> increasing number <strong>of</strong> elapsed loading cycles.<br />

<br />

δ0<br />

δ0<br />

x<br />

σm,1 max<br />

σf,1 min<br />

σ<br />

σf,1 max<br />

Figure 4.5a: stresses in matrix <strong>and</strong> fibres at<br />

σc max , first cycle<br />

N cycles<br />

<br />

δN<br />

δN<br />

x<br />

σm,N max<br />

σf,N min<br />

σ<br />

σf,N max<br />

Figure 4.5b: stresses in matrix <strong>and</strong> fibres at<br />

σc max , N th cycle<br />

Debonding length δN is function <strong>of</strong> the frictional matrix-fibre interface shear stress<br />

τN <strong>and</strong> is therefore also function <strong>of</strong> ω. With increasing number <strong>of</strong> load cycles, τN<br />

decreases <strong>and</strong> the debonding length increases. The expression <strong>of</strong> δN in the<br />

stochastic cracking (based) theory is similar to the expression used in the ACK<br />

(based) theory (see equation (4.5)) <strong>and</strong> based on figure 4.5b:<br />

ff , max<br />

max<br />

max<br />

max<br />

σ m,<br />

N r V σ m m,<br />

N r V σ m m,<br />

N r Vm<br />

σ c r EmVm<br />

δ N =<br />

= = =<br />

∗<br />

∗<br />

∗<br />

∗<br />

(4.33)<br />

2τ V 2τ<br />

V 2ωτ<br />

V 2ωτ<br />

E V<br />

N<br />

f<br />

N<br />

f<br />

When figure 4.2 <strong>and</strong> figure 4.5 are compared, one essential difference can be<br />

noticed, concerning possible introduction <strong>of</strong> extra matrix cracks after the first<br />

loading cycle.<br />

0<br />

110<br />

f<br />

0<br />

c1<br />

f


Chapter 4: IPC composite specimens under repeated tensile loading<br />

- Figure 4.2a <strong>and</strong> 4.2b: As can be seen from these figures, according to the<br />

ACK theory, matrix multiple cracking occurs at one fixed stress level. From first<br />

loading on, the average matrix stress at maximum composite stress is lower than<br />

σmu. With increasing number <strong>of</strong> elapsed load cycles, σm,N max decreases (see<br />

equation (4.8). Once multiple cracking occurs, the formation <strong>of</strong> extra matrix<br />

cracks becomes very unlikely, since matrix stresses are too low <strong>and</strong> are further<br />

decreasing upon cycling.<br />

- Figure 4.5a <strong>and</strong> 4.5b: From these figures, it can be noticed that the matrix<br />

stress in the middle <strong>of</strong> two cracks equals the value <strong>of</strong> the matrix stress as<br />

calculated from the law <strong>of</strong> mixtures, as long as 2δN < . This means σm,N max<br />

stays constant <strong>with</strong> increasing number <strong>of</strong> load cycles, as long as 2δN < . It is<br />

therefore more likely that extra matrix cracks occur due to repeated cycling <strong>of</strong> the<br />

specimen. However, interface degradation <strong>and</strong> matrix cracking are two competing<br />

degradation mechanisms under repeated loading. Equation (4.33) <strong>and</strong> figure 4.5a<br />

<strong>and</strong> 4.5b show that the debonding length δN increases <strong>with</strong> increasing number <strong>of</strong><br />

load cycles. If matrix-fibre interface degradation occurs very rapidly, δN becomes<br />

larger than /2 after only few load cycles are applied. Once 2δN > , the<br />

maximum matrix stress σm,N max starts to decrease <strong>with</strong> the number <strong>of</strong> elapsed load<br />

cycles (see matrix stress situation in figure 4.2), decreasing the probability that<br />

extra matrix cracks occur due to further repeated loading.<br />

If > 2δN the maximum average composite strain εc,N max is expressed as in<br />

paragraph 2.5.3.1, thus equivalent to equation (2.88). However, δN is now function<br />

<strong>of</strong> the number <strong>of</strong> elapsed load cycles <strong>and</strong> defined by equation (4.33). is<br />

defined by equation (4.32):<br />

max<br />

max σ c αδ N<br />

ε c,<br />

N = ( 1+<br />

)<br />

(4.34)<br />

E cs<br />

c1<br />

4.4.1.2 transition from > 2δN to < 2δN<br />

Since δN increases <strong>with</strong> increasing number <strong>of</strong> elapsed load cycles <strong>and</strong> is<br />

considered to be independent <strong>of</strong> the number <strong>of</strong> elapsed load cycles, 2δN might<br />

exceed after several load cycles, even when 2δN < at first loading. The<br />

transition from > 2δN to < 2δN can be expressed by inserting equation<br />

(4.33) <strong>and</strong> equation (4.32) in the condition = 2δN. After some<br />

rearrangements, an expression for ω is found:<br />

max<br />

τ N σ c EmrVm<br />

ω = =<br />

∗<br />

(4.35)<br />

τ E τ V < cs ><br />

0<br />

c1<br />

0<br />

f<br />

111


Chapter 4: IPC composite specimens under repeated tensile loading<br />

4.4.1.3 crack spacing smaller than twice the debonding length ( < 2δN)<br />

When < 2δN, the maximum strain can be formulated, based on equation<br />

(2.89), taking into account δN is expressed by equation (4.33). is formulated<br />

by equation (4.32):<br />

⎛<br />

⎞<br />

max max ⎜<br />

1 α cs<br />

ε = − ⎟<br />

c,<br />

N σ c ⎜ ∗ ⎟<br />

(4.36)<br />

⎝ E fV<br />

f 4δ<br />

N Ec1<br />

⎠<br />

4.4.2 evolution <strong>of</strong> Ecycle,N <strong>with</strong> N<br />

4.4.2.1 partial matrix-fibre unloading slip<br />

In case only partial matrix-fibre slip occurs during unloading, the derivation <strong>of</strong><br />

Ecycle,N can be written similarly to paragraph 3.4.2.1. Ecycle,N can be expressed<br />

similar to equation (3.33):<br />

Ec1<br />

Ecycle,<br />

N =<br />

αδ N ∆σ<br />

c<br />

1+<br />

(4.37)<br />

max<br />

2 cs σ<br />

4.4.2.2 when partial becomes total matrix-fibre unloading slip, > 2δN<br />

When the number <strong>of</strong> elapsed load cycles increases, a cycling mechanism can<br />

change from partial to full matrix-fibre slip. The change from partial to total<br />

matrix-fibre unloading slip under repeated loading is based on the expression,<br />

derived for a single cycle in paragraph 3.4.2.2 <strong>and</strong> 3.4.2.3.<br />

If the crack space exceeds two times the debonding length ( > 2δN), the<br />

switch from partial to full matrix-fibre slip during unloading occurs when (su)N<br />

equals δN . Partial matrix-fibre slip becomes full matrix-fibre slip once:<br />

elastic<br />

( su)<br />

N ∆σ<br />

m Em∆σ<br />

c ∆σ<br />

c<br />

=<br />

=<br />

= = 1<br />

max<br />

ff , max max<br />

(4.38)<br />

δ 2σ<br />

( x = δ ) 2E<br />

σ 2σ<br />

N<br />

m,<br />

N<br />

N<br />

c1<br />

From equation (4.38) it can be seen that, as long as > 2δN, full matrix-fibre<br />

slip only occurs when the specimen is loaded in compression after releasing the<br />

tensile force. Equation (4.38) equals equation (3.35) for a single load cycle (first<br />

cycle). However, in paragraph 3.4.2.2 it has been verified that crack closure<br />

always occurs before partial matrix-fibre slip becomes total matrix-fibre unloading<br />

slip.<br />

As a conclusion, total matrix-fibre slip does not occur under repeated loading, as<br />

long as > 2δN.<br />

c<br />

m<br />

112<br />

c


Chapter 4: IPC composite specimens under repeated tensile loading<br />

4.4.2.3 when partial becomes total matrix-fibre unloading slip, < 2δN<br />

If the debonding length exceeds half the crack spacing (2δN > ), the existence<br />

<strong>of</strong> full matrix-fibre slip is formulated by expressing that (su)N equals half the crack<br />

spacing, as shown in equation (4.39).<br />

elastic<br />

( su)<br />

∆σ<br />

N<br />

m,<br />

N<br />

=<br />

= 1<br />

max<br />

cs / 2 2σ<br />

m,<br />

N ( x =< cs > / 2)<br />

(4.39)<br />

<strong>and</strong> since:<br />

( su)<br />

N<br />

δ N<br />

elastic<br />

∆σ<br />

m Em∆σ<br />

c<br />

= =<br />

max<br />

ff , max<br />

2σ<br />

m,<br />

N ( x = δ N ) 2Ec1σ<br />

m,<br />

N<br />

∆σ<br />

c = max<br />

2σ<br />

c<br />

(4.40)<br />

equation (3.39) can be rewritten:<br />

max<br />

< cs > σ c<br />

∆ σ c =<br />

δ N<br />

(4.41)<br />

After some rearrangements, the combination <strong>of</strong> equation (4.33) <strong>and</strong> (4.41) gives:<br />

∆σ<br />

cαE<br />

f r<br />

ω =<br />

2E < cs > τ<br />

(4.42)<br />

c<br />

4.4.2.4 total matrix-fibre unloading slip<br />

Once ω reaches the value expressed by equation (4.42), total matrix-fibre slip<br />

occurs during unloading (<strong>and</strong> reloading), provided 2δN > . The formulation <strong>of</strong><br />

Ecycle,N is similar to the one derived in paragraph 3.4.2.4.<br />

∗<br />

E fV<br />

f<br />

Ecycle,<br />

N =<br />

max<br />

cs Em<br />

σ<br />

(4.43)<br />

c<br />

1−<br />

Vm<br />

2δ<br />

E ∆σ<br />

N<br />

4.5 Determination <strong>of</strong> the interface degradation evolution<br />

from experimental observations<br />

0<br />

c1<br />

4.5.1 introduction<br />

The main goal <strong>of</strong> this chapter is to discuss the evolution <strong>of</strong> damage in<br />

unidirectionally reinforced <strong>and</strong> 2D-r<strong>and</strong>omly reinforced IPC laminates under<br />

repeated loading. Until now, it has been assumed that degradation in IPC<br />

composite specimens under repeated loading is only caused by loss <strong>of</strong> the matrixfibre<br />

stress transfer.<br />

The evolution <strong>of</strong> the matrix-fibre interface shear stress is included in an interface<br />

degradation parameter ω, <strong>with</strong> ω = τN/τ0. The evolution <strong>of</strong> ω cannot be measured<br />

directly, but can be determined from measurements <strong>of</strong> Ecycle,N <strong>and</strong> εc,N max as a<br />

function <strong>of</strong> N. Given an experimentally obtained evolution <strong>of</strong> Ecycle,N <strong>and</strong> εc,N max , ω<br />

113<br />

c


Chapter 4: IPC composite specimens under repeated tensile loading<br />

can be calculated according to the ACK (based) or stochastic cracking (based)<br />

model. This methodology is explained, tested <strong>and</strong> discussed further.<br />

4.5.2 ACK (based) theory<br />

If the ACK theory is used, no damage is introduced, when σc max is situated in zone<br />

I (elastic zone). If σc max is located into zone III (post-cracking zone), the<br />

expression <strong>of</strong> ω versus the maximum cycle strain is derived from equation (4.15)<br />

<strong>and</strong> is:<br />

⎡ max ⎛<br />

max ε ( ) ⎞ ⎤<br />

⎢ ⎜ c,<br />

N σ c −σ<br />

mc<br />

= − −<br />

− ⎟<br />

1<br />

ω 2.<br />

994 1<br />

1 ⎥<br />

⎢<br />

⎜<br />

∗ ⎟<br />

(4.44)<br />

⎣ ⎝ ε mu KE fV<br />

f ε mu ⎠α<br />

⎥⎦<br />

If only partial matrix-fibre slip occurs (when the composite is unloaded <strong>and</strong><br />

reloaded), equation (4.25) can be rearranged a follows:<br />

∆σ<br />

⎛<br />

⎞<br />

cEmα<br />

Ecycle,<br />

N<br />

ω =<br />

⎜<br />

⎟<br />

⎜<br />

⎟<br />

(4.45)<br />

2. 667Ecσ<br />

mu ⎝ Ec1<br />

− Ecycle,<br />

N ⎠<br />

<strong>and</strong> when total matrix-fibre slip occurs during unloading (after rearrangement <strong>of</strong><br />

equation (4.31)):<br />

∗<br />

∆σ<br />

⎛ E ⎞ fV<br />

c<br />

f<br />

ω =<br />

⎜1−<br />

⎟<br />

⎜ ⎟<br />

(4.46)<br />

0. 668V<br />

mσ<br />

mu ⎝ Ecycle,<br />

N ⎠<br />

4.5.3 stochastic cracking (based) theory<br />

If > 2δN, combination <strong>of</strong> equation (4.33) <strong>and</strong> (4.34) gives:<br />

2 max max<br />

E f rα<br />

σ c ⎛ E ⎞<br />

cεc<br />

, N<br />

ω = ⎜ 1⎟<br />

max<br />

2 ⎜<br />

−<br />

τ<br />

⎟<br />

0 cs Ec1<br />

⎝ σ c ⎠<br />

(4.47)<br />

Once < 2δN, equation (4.48) is obtained after rearrangement <strong>of</strong> the<br />

combination <strong>of</strong> equations (4.33) <strong>and</strong> (4.36):<br />

⎛ max 2rE<br />

⎞<br />

f ⎜<br />

σ c max<br />

ω = − ε ⎟<br />

⎜ ∗ c,<br />

N<br />

cs τ<br />

⎟<br />

0 ⎝ E fV<br />

f ⎠<br />

(4.48)<br />

Equation (4.37), which expresses the evolution <strong>of</strong> Ecycle,N, can be rewritten. Thus in<br />

case only partial matrix-fibre slip occurs along the debonded interface:<br />

−1<br />

−1<br />

2<br />

rE ⎛ ⎞<br />

fα<br />

∆σ<br />

c ⎜<br />

Ec1<br />

ω =<br />

−1⎟<br />

(4.49)<br />

4Ec1<br />

cs τ ⎜<br />

0 ⎝ Ecycle,<br />

N<br />

⎟<br />

⎠<br />

If total matrix-fibre slip occurs during unloading, equation (4.43) becomes:<br />

ω =<br />

∗<br />

r∆σ<br />

⎛ E ⎞ fV<br />

c<br />

f ⎜1−<br />

⎟<br />

∗<br />

cs τ V ⎜ ⎟<br />

0 f ⎝ Ecycle,<br />

N ⎠<br />

(4.50)<br />

114


Chapter 4: IPC composite specimens under repeated tensile loading<br />

4.6 Testing program<br />

Specimens <strong>of</strong> three IPC composite plates are tested <strong>and</strong> discussed in this chapter.<br />

The first two laminates are UD-reinforced laminates. The first laminate is RUD1<br />

(Repeated load, UD-reinforced, plate 1) <strong>and</strong> the second is RUD2. The third<br />

laminate contains 2D-r<strong>and</strong>om reinforcement <strong>and</strong> is noted RR3.<br />

For all three laminates, six steps are followed in the testing schedule <strong>and</strong> the<br />

discussion:<br />

1. objective <strong>of</strong> the testing program, applied on the laminate<br />

2. determination <strong>of</strong> a priori known material properties (Vf, r, Em,<br />

etc.) <strong>and</strong> test schedule<br />

3. determination <strong>of</strong> material properties, not known a priori:<br />

determination <strong>of</strong> stochastic cracking model parameters σR, m <strong>and</strong> τ0<br />

from monotonic tensile loading (see Chapter 2)<br />

4. results under repeated loading: Ecycle,N <strong>and</strong> εc,N max as function <strong>of</strong> N<br />

5. determination <strong>of</strong> evolution <strong>of</strong> ω from monitored evolution <strong>of</strong><br />

Ecycle,N <strong>and</strong> εc,N max <strong>with</strong> help <strong>of</strong> equations (4.47) to (4.50)<br />

6. discussion <strong>and</strong> conclusions<br />

Only the behaviour predicted <strong>with</strong> the stochastic cracking based theory under<br />

repeated loading is compared to the experimental results. The ACK based theory<br />

is not used. This choice is justified later in paragraph 4.9. Discussion <strong>of</strong> the results<br />

<strong>of</strong> the three tested plates will lead to conclusions on several topics. These topics<br />

are:<br />

- RUD1: preliminary determination <strong>of</strong> the evolution <strong>of</strong> the matrix-fibre<br />

interface degradation parameter ω from repeated loading tests <strong>with</strong> various mean<br />

stresses <strong>and</strong> amplitudes. All maximum cycle stress levels are situated in the postcracking<br />

zone (zone III <strong>of</strong> ACK theory). ω is determined from measured<br />

evolutions <strong>of</strong> Ecycle,N <strong>and</strong> εc,N max , obtained under relatively high number <strong>of</strong> applied<br />

load cycles (about one million).<br />

- RUD2: two questions are discussed:<br />

1. Is it possible to obtain an evolution <strong>of</strong> ω from a limited number<br />

<strong>of</strong> load cycles (typically one thous<strong>and</strong>), which can be<br />

extrapolated (typically up to one million load cycles)?<br />

2. Are additional matrix cracks introduced under further repeated<br />

loading, after first loading occured?<br />

115


Chapter 4: IPC composite specimens under repeated tensile loading<br />

- RR3: tests on specimens <strong>of</strong> this plate are used to verify if the same<br />

assumptions used for UD-reinforced continuous specimens are acceptable for 2Dr<strong>and</strong>omly<br />

reinforced specimens.<br />

4.7 UD-reinforced IPC composite specimens under repeated<br />

loading<br />

4.7.1 introduction<br />

In paragraph 4.7, a preliminary evolution <strong>of</strong> degradation parameter ω will be<br />

obtained. ω is determined from measured evolutions <strong>of</strong> Ecycle,N <strong>and</strong> εc,N max ,<br />

obtained during repeated loading up to a relatively high number <strong>of</strong> applied load<br />

cycles (typically one million). Three specimens are cycled up to a different<br />

maximum stress to study the influence <strong>of</strong> this maximum cycle stress in the<br />

evolution <strong>of</strong> ω. The chosen maximum stress levels are 30MPa, 40MPa <strong>and</strong><br />

60MPa. At present, all chosen stress levels are situated in the post-cracking zone<br />

(zone III <strong>of</strong> the ACK theory).<br />

4.7.2 material properties <strong>and</strong> test schedule<br />

The fibre volume fraction <strong>and</strong> minimum <strong>and</strong> maximum composite cycle stress are<br />

listed in table 4.1. Two specimens, cut from these plates (RUD1-1 <strong>and</strong> RUD1-2)<br />

are tested under monotonic tensile loading on the mechanical INSTRON 4505<br />

testing-bench. The width <strong>of</strong> these specimens is 18mm <strong>and</strong> the length is 300mm.<br />

Damage evolution under a high number <strong>of</strong> load cycles (RUD1-9 to RUD1-11) is<br />

studied from the test results, obtained by cycling <strong>of</strong> the specimens <strong>with</strong> a hydraulic<br />

testing-bench (typically up to one or two million cycles, regarding the availability<br />

<strong>of</strong> the test set-up).<br />

Table 4.1: load schedule <strong>and</strong> fibre volume fraction UD-reinforced IPC specimens, plate RUD1<br />

specimen calculated Vf minimum stress maximum stress<br />

name<br />

%<br />

(MPa)<br />

(MPa)<br />

RUD1-1 14.8 monotonic tensile loading<br />

RUD1-2 14.8 monotonic tensile loading<br />

RUD1-9 14.0 0 40<br />

RUD1-10 14.1 10 30<br />

RUD1-11 14.0 20 60<br />

4.7.3 determination <strong>of</strong> material properties<br />

The stress-strain curves under monotonic tensile loading (specimens RUD1-1 <strong>and</strong><br />

RUD1-2) are printed in figure 4.6. Three material parameters, not known a priori,<br />

are determined by finding <strong>of</strong> the “best fit” <strong>of</strong> experimental <strong>and</strong> theoretical stressstrain<br />

curves, as explained in Chapter 2. r is 7µm, Em is 18 GPa (see table 2.2) <strong>and</strong><br />

Ef is 72 GPa (see table 2.1). The final crack spacing f is counted on the<br />

116


Chapter 4: IPC composite specimens under repeated tensile loading<br />

specimens after testing. Vf is listed in table 4.1. The values <strong>of</strong> σR, m <strong>and</strong> τ0 are<br />

listed in table 4.2.<br />

stress (MPa)<br />

200<br />

150<br />

100<br />

50<br />

0<br />

RUD1-1<br />

RUD1-2<br />

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6<br />

strain (%)<br />

Figure 4.6: stress-strain curve under monotonic tensile loading; unidirectionally reinforced plate<br />

RUD1<br />

Table 4.2: determination <strong>of</strong> material parameters from specimens RUD1-1 <strong>and</strong> RUD1-2<br />

m τ0 σR<br />

RUD1-1 3.3 0.74 8.9<br />

RUD1-2 3.5 0.66 11<br />

average 3.4 0.70 10<br />

The values <strong>of</strong> σR, m <strong>and</strong> τ0 listed in table 4.2, are used further as material<br />

properties in equations (4.47) to (4.50).<br />

4.7.4 results under repeated loading<br />

Cycling <strong>of</strong> the specimens is performed by a hydraulic INSTRON testing system:<br />

sinusoidal load <strong>with</strong> a frequency <strong>of</strong> 5Hz is applied. A load cell measures the force<br />

on the specimens <strong>and</strong> the strain is monitored by an extensometer. Both the force<br />

<strong>and</strong> strain are measured at a sampling rate <strong>of</strong> 400Hz. Three specimens are<br />

subjected to repeated loading up to one million cycles. The force <strong>and</strong> strain<br />

evolution are measured for cycle number 100, 500, 1000, 5000, 10000, etc.<br />

The maximum cycle stress <strong>of</strong> all three specimens tested under repeated loading is<br />

situated in the post-cracking zone, such that full matrix cracking occurs at first<br />

loading (load cycle one). However, stresses are probably still too low for serious<br />

fibre breakage to occur, since the highest applied cycle stress is 60MPa <strong>and</strong> the<br />

failure stress <strong>of</strong> the composite is about 140MPa (see figure 4.6).<br />

Figures 4.7a <strong>and</strong> 4.7b show the evolution <strong>of</strong> Ecycle,N <strong>and</strong> εc,N max respectively as a<br />

function <strong>of</strong> the number <strong>of</strong> elapsed load cycles for the three specimens under<br />

repeated loading.<br />

117


cycle, N (GPa)<br />

E<br />

22<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

Chapter 4: IPC composite specimens under repeated tensile loading<br />

1.E+00 1.E+02 1.E+04 1.E+06 1.E+08<br />

# cycles<br />

Figure 4.7a: evolution <strong>of</strong> Ecycle,N under<br />

repeated loading, laminate RUD1<br />

Figure 4.7b: evolution <strong>of</strong> εc,N max under repeated<br />

loading, laminate RUD1<br />

One preliminary conclusion can be formulated from figures 4.7a <strong>and</strong> 4.7b. If the<br />

matrix-fibre interface would be degraded totally after repeated loading occurred,<br />

eventually only the continuous fibres would provide stiffness <strong>and</strong> Ecycle,N would<br />

reach EfVf. Theoretically, EfVf is about 10GPa (since Vf is 14%, see table 4.1).<br />

The experimentally obtained values <strong>of</strong> Ecycle,N are higher than 10GPa for all tested<br />

specimens in figure 4.7a. Since the matrix still contributes to Ecycle,N (compare<br />

Ecycle,N in figures 4.7a <strong>and</strong> 4.7b <strong>with</strong> EfVf <strong>of</strong> 10GPa), the matrix <strong>and</strong> fibres still<br />

interact <strong>and</strong> the matrix-fibre interface is not totally degraded under repeated<br />

loading.<br />

4.7.5 determination <strong>of</strong> interface degradation parameter from test<br />

results<br />

Figures 4.8 to 4.10 show the evolution <strong>of</strong> ω as determined from measurement <strong>of</strong><br />

εc,N max (figure 4.7b) <strong>and</strong> <strong>of</strong> Ecycle,N (figure 4.7a),. The white dots/squares/triangles<br />

represent the evolution <strong>of</strong> ω as calculated from the measured evolution <strong>of</strong> Ecycle,N.<br />

The black dots/squares/triangles show the evolution <strong>of</strong> ω, as calculated from the<br />

measured evolution <strong>of</strong> εc,N max .<br />

ω<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0-40MPa<br />

10-30MPa<br />

20-60MPa<br />

Figure 4.8: evolution <strong>of</strong> ω, retrieved from experiments, 0-40MPa (RUD1-9)<br />

εc,N max (%)<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0-40MPa<br />

10-30MPa<br />

20-60MPa<br />

1.E+00 1.E+02 1.E+04 1.E+06 1.E+08<br />

# cycles<br />

E cycle<br />

maximum strain<br />

0.4<br />

1.E+00 1.E+02 1.E+04<br />

# cycles<br />

1.E+06 1.E+08<br />

118


Chapter 4: IPC composite specimens under repeated tensile loading<br />

ω<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

E cycle<br />

maximum strain<br />

1.E+00 1.E+02 1.E+04 1.E+06 1.E+08<br />

# cycles<br />

Figure 4.9: evolution <strong>of</strong> ω retrieved from experiments, 10-30MPa (RUD1-10)<br />

ω<br />

0.8<br />

0.75<br />

0.7<br />

0.65<br />

E cycle<br />

maximum strain<br />

0.6<br />

1.E+00 1.E+02 1.E+04<br />

# cycles<br />

1.E+06 1.E+08<br />

Figure 4.10: evolution <strong>of</strong> ω, retrieved from experiments, 20-60MPa (RUD1-11)<br />

Following conclusions can be retrieved from figures 4.8 to 4.10:<br />

1. The evolution <strong>of</strong> the matrix-fibre interface degradation parameter ω is<br />

obtained from the measurement <strong>of</strong> Ecycle,N <strong>and</strong> εc,N max . Ecycle,N <strong>and</strong> εc,N max are<br />

determined independently from each other from the experimental stress-strain<br />

curve. The evolution <strong>of</strong> ω is thus determined in two independent ways. From<br />

figure 4.98 to 4.10, it can be noticed that the evolutions <strong>of</strong> ω, determined from<br />

Ecycle,N <strong>and</strong> from εc,N max , show rather good coincidence. This is an indication that<br />

indeed it is the evolution <strong>of</strong> the frictional matrix-fibre interface shear stress that<br />

determines the evolution <strong>of</strong> the remaining stiffness <strong>of</strong> the composite under<br />

repeated loading. (Ecycle,N <strong>and</strong> εc,N max are both concerned to the stiffness).<br />

2. In figure 4.8, it can be seen that the value <strong>of</strong> ω is systematically lower if<br />

it is determined from the evolution <strong>of</strong> εc,N max than when it is determined from the<br />

evolution <strong>of</strong> Ecycle,N (see Evans et al. (1995)). Specimen RUD1-9 (figure 4.8), is<br />

the only specimen which is fully unloaded. The minimum cycle stress <strong>of</strong> the other<br />

two specimens is considerably higher than zero (10MPa for RUD1-10 <strong>and</strong> 20MPa<br />

119


Chapter 4: IPC composite specimens under repeated tensile loading<br />

for RUD1-11). In paragraph 3.6.3 it has been mentioned that full crack closure<br />

may be prevented when a specimen is totally unloaded. Figure 4.8 shows that the<br />

experimental value <strong>of</strong> Ecycle,N will be higher due to prevented crack closure, but<br />

εc,N max is not influenced by this effect. Therefore, the calculated value <strong>of</strong> ω is<br />

overestimated when it is determined from equation (4.49) or (4.50). Consequently,<br />

in figure 4.8 the values <strong>of</strong>, ω determined from the evolution Ecycle,N, are<br />

overestimated values.<br />

σ<br />

Ecycle,N <strong>with</strong>out prevented crack closure<br />

ε<br />

Ecycle,N <strong>with</strong> prevented crack closure<br />

Figure 4.11: Ecycle,N calculated from a cycle <strong>with</strong> or <strong>with</strong>out prevented crack closure,<br />

according to Evans et al. (1995)<br />

4.7.6 discussion <strong>and</strong> conclusions<br />

A mathematical expression <strong>of</strong> the evolution <strong>of</strong> the frictional matrix-fibre interface<br />

shear stress as a function <strong>of</strong> the number <strong>of</strong> cycles should be derived. This<br />

expression can later be used to predict the evolution <strong>of</strong> the stiffness properties <strong>of</strong><br />

IPC composite specimens under repeated loading. According to figure 4.8 to 4.10,<br />

the mathematical expression <strong>of</strong> ω seems to meet three requirements:<br />

1. ω is different from one, starting from the first load cycle<br />

2. ω decreases rapidly during the first load cycles<br />

3. when a large number <strong>of</strong> load cycles are applied, ω keeps on decreasing<br />

slightly<br />

Several mathematical expressions meet these requirements: an exponential<br />

function, a logarithmic function, a power law, etc. Almost every author, discussing<br />

evolution <strong>of</strong> the matrix-fibre interface degradation, works <strong>with</strong> another<br />

mathematical expression for ω (Rouby <strong>and</strong> Reynaud, 1992, Evans et al., 1994,<br />

Curtin, 1999)<br />

120


Chapter 4: IPC composite specimens under repeated tensile loading<br />

The mathematical formulation <strong>of</strong> ω, which will be used in this work, is formulated<br />

by equation (4.51):<br />

ω = C1 − C2<br />

ln( N)<br />

(4.51)<br />

C1 is thus an estimation <strong>of</strong> the frictional matrix-fibre interface degradation in the<br />

first loading-unloading-reloading cycle. The matrix-fibre interface degradation<br />

speed after the first load cycle is implemented in C2.<br />

The constants C1 <strong>and</strong> C2 are obtained by finding a best fit <strong>of</strong> formula (4.51) <strong>with</strong><br />

the experimentally obtained evolution <strong>of</strong> ω. For each specimen the average<br />

evolution <strong>of</strong> ω, determined from the experimental evolution <strong>of</strong> Ecycle,N <strong>and</strong> <strong>of</strong> εc max<br />

is used for a best fit <strong>of</strong> equation (4.51). The values <strong>of</strong> C1 <strong>and</strong> C2 are listed in table<br />

4.3.<br />

Table 4.3: constants in formulation <strong>of</strong> matrix-fibre interface degradation; laminate RUD1-1<br />

specimen stress interval C1 C2<br />

RUD1-9 0-40MPa 0.91 0.019<br />

RUD1-10 10-30MPa 0.83 0.010<br />

RUD1-11 20-60MPa 0.78 0.0057<br />

From table 4.3 <strong>and</strong> figures 4.8 to 4.10, several preliminary conclusions can be<br />

formulated:<br />

- As has been mentioned, all degradation, which occurs in the first cycle, is<br />

implemented in C1. If C1 equals one, no damage occurs in the first loadingunloading-reloading<br />

cycle. In table 4.3 it can be seen that C1 is lower for specimen<br />

RUD1-11, which is cycled up to 60MPa. Most damage is introduced in the first<br />

load cycle if the maximum cycle stress is 60MPa (compared to maximum cycle<br />

stresses <strong>of</strong> 30MPa or 40MPa).<br />

- C2 is a measure for the speed <strong>of</strong> further matrix-fibre interface degradation after<br />

the first load cycle. The higher C2, the faster matrix-fibre stress transfer is lost. As<br />

can be seen in table 4.2 the loss <strong>of</strong> stress transfer, after the first cycle, is higher for<br />

those specimens, which are cycled up to lower σc max . This - at first sight strange -<br />

conclusion is explained by the fact that the matrix-fibre interface already<br />

experienced major degradation during the first loading-unloading-reloading cycle<br />

for specimen RUD1-11 (cycled up to 60MPa). If the maximum cycle stress is high<br />

(here typically 60MPa), practically all degradation occurs in the first cycle(s).<br />

Further cycling only introduces minor extra matrix-fibre interface degradation,<br />

since the matrix-fibre stress transfer is already quite low.<br />

121


Chapter 4: IPC composite specimens under repeated tensile loading<br />

- From figures 4.8 to 4.10 it can be seen that, regardless <strong>of</strong> the maximum cycle<br />

stress or cycle amplitude, the value <strong>of</strong> ω is about 0.6 to 0.7 after one million load<br />

cycles are applied.<br />

- If the σc max is lower (here for example 30 or 40MPa), full matrix-fibre<br />

degradation is postponed. The degradation is still worst for the first few cycles.<br />

The degradation speed then slowly diminishes <strong>with</strong> increasing number <strong>of</strong> elapsed<br />

load cycles.<br />

4.8 UD-reinforced IPC composite specimen under limited<br />

repeated loading<br />

4.8.1 introduction<br />

As has already been mentioned, two questions are discussed in this paragraph:<br />

1. Is it possible to obtain an evolution <strong>of</strong> ω from limited repeated loading<br />

(typically one thous<strong>and</strong> load cycles), which can be extrapolated (typically up to<br />

one million loading cycles)?<br />

2. Is degradation <strong>of</strong> the matrix-fibre interface being the sole damage<br />

mechanism still correct, when σc max is relatively low (zone I <strong>and</strong> II <strong>of</strong> ACK<br />

theory)?<br />

The preliminary conclusions, obtained in paragraph 4.7 are used to answer these<br />

questions, together <strong>with</strong> extra experimental observations.<br />

4.8.2 material properties <strong>and</strong> test schedule<br />

A unidirectionally reinforced IPC plate RUD2, reinforced <strong>with</strong> 6 layers <strong>of</strong> glass<br />

fibre is laminated. Specimens, which are cut from this plate (18x250mm²), are<br />

tested on the mechanical INSTRON 4505. The fibre volume fraction, test schedule<br />

<strong>and</strong> maximum <strong>and</strong> minimum cycle stress levels <strong>of</strong> each specimen are listed in<br />

tables 4.4 to 4.6. In table 4.4 the specimens, which are subjected to monotonic<br />

tensile loading, are listed. Specimens, which are tested to study the question<br />

concerning extrapolation <strong>of</strong> the evolution <strong>of</strong> ω, can be found in table 4.5. In table<br />

4.6 specimens are listed, which are tested repeatedly at lower stress levels,<br />

typically in zone I (linear elastic zone) <strong>and</strong> zone II (multiple cracking zone) <strong>of</strong> the<br />

ACK theory.<br />

Table 4.4: specimens <strong>of</strong> plate RUD2, subjected to monotonic tensile loading<br />

specimen calculated Vf minimum stress maximum stress<br />

name<br />

%<br />

(MPa)<br />

(MPa)<br />

RUD2-1 12.8 monotonic tensile loading<br />

RUD2-2 13.7 monotonic tensile loading<br />

RUD2-3 12.5 monotonic tensile loading<br />

122


Chapter 4: IPC composite specimens under repeated tensile loading<br />

Table 4.5: specimens <strong>of</strong> plate RUD2, subjected to limited repeated loading (one thous<strong>and</strong> cycles)<br />

specimen calculated Vf minimum stress maximum stress<br />

name<br />

%<br />

(MPa)<br />

(MPa)<br />

RUD2-5 14.0 0 20<br />

RUD2-6 12.8 0 40<br />

RUD2-7 13.8 0 60<br />

Table 4.6: specimens <strong>of</strong> plate RUD2, subjected to repeated loading up to lower stress levels (zone<br />

I <strong>and</strong> II in ACK theory)<br />

specimen calculated Vf minimum stress maximum stress<br />

name<br />

%<br />

(MPa)<br />

(MPa)<br />

RUD2-8 13.0 1.0 5<br />

RUD2-9 12.5 1.0 10<br />

4.8.3 determination <strong>of</strong> material properties<br />

Figure 4.11a shows the stress-strain curve <strong>of</strong> three specimens under monotonic<br />

tensile loading. In figure 4.11b these stress-strain curves are printed between 0-<br />

20MPa (multiple cracking region).<br />

stress (MPa)<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8<br />

strain (%)<br />

Figure 4.11a: stress-strain curve under<br />

monotonic tensile loading, laminate HCUDF<br />

stress (MPa)<br />

20<br />

15<br />

10<br />

5<br />

0<br />

RUD1-3<br />

RUD1-2<br />

RUD1-1<br />

0 0.025 0.05 0.075 0.1 0.125 0.15<br />

strain (%)<br />

Figure 4.11b: stress-strain curve under<br />

monotonic tensile loading, 0-20MPa<br />

From the stress-strain curves illustrated in figure 4.11b, parameters τ0, m <strong>and</strong> σR<br />

are determined by finding a ‘best fit’ <strong>of</strong> experimental <strong>and</strong> theoretical curves (see<br />

Chapter 2 <strong>and</strong> paragraph 4.7.3). The average values are: m = 3.7, τ0 = 0.5 <strong>and</strong> σR<br />

= 9.2. The material parameters are used as input in equations (4.47) to (4.50).<br />

4.8.4 results under repeated loading<br />

Specimens RUD2-5 to RUD2-9 are subjected to repeated loading up to 1000 load<br />

cycles. The force <strong>and</strong> strain are monitored at cycle number 1, 2, 50, 100, 200, 300,<br />

400, 500 … up to 1000 cycles. Figure 4.12a shows the evolution <strong>of</strong> Ecycle,N <strong>of</strong> the<br />

specimens, which are loaded repeatedly. Figure 4.12b shows the evolution <strong>of</strong> the<br />

maximum strain, εc,N max .<br />

123


Chapter 4: IPC composite specimens under repeated tensile loading<br />

E cycle (GPa)<br />

24<br />

22<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

0-5 Mpa 0-10MPa 0-20MPa<br />

0-40MPa 0-60MPa<br />

0 100 200 300 400 500 600 700 800 900 1000<br />

# cycles<br />

Figure 4.12a: evolution <strong>of</strong> Ecycle,N versus the number <strong>of</strong> cycles under repeated loading<br />

εc max (GPa)<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0-5 Mpa 0-10MPa 0-20MPa<br />

0-40MPa 0-60MPa<br />

0 100 200 300 400 500 600 700 800 900 1000<br />

# cycles<br />

Figure 4.12b: evolution <strong>of</strong> εc,N max versus the number <strong>of</strong> cycles under repeated loading<br />

4.8.5 determination <strong>of</strong> interface degradation parameter from test<br />

results<br />

4.8.5.1 extrapolation from limited cycling to high number <strong>of</strong> cycles<br />

Specimens RUD2-5 to RUD2-7 are loaded well into the post-cracking zone, like<br />

all specimens <strong>of</strong> plate RUD1 (paragraph 4.7). The evolution <strong>of</strong> ω is obtained from<br />

the evolution <strong>of</strong> Ecycle,N <strong>and</strong> εc,N max for specimens RUD2-5 (0-20MPa), RUD2-6 (0-<br />

40MPa) <strong>and</strong> RUD2-7 (0-60MPa) <strong>with</strong> equations (4.47) to (4.50). Figure 4.13<br />

shows the evolution <strong>of</strong> ω, retrieved for specimen RUD2-5 (0-20MPa). From figure<br />

4.13 it is noticed that the effect <strong>of</strong> the interface degradation might be<br />

underestimated if it is derived from the evolution <strong>of</strong> Ecycle,N.: prevented crack<br />

closure may lead to overestimation <strong>of</strong> Ecycle,N <strong>and</strong> thus to overestimation <strong>of</strong> ω, as<br />

was already mentioned before (see paragraph 4.7, specimen RUD1-9).<br />

124


Chapter 4: IPC composite specimens under repeated tensile loading<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

E cycle<br />

maximum strain<br />

0 200 400 600 800 1000<br />

# cycles<br />

Figure 4.13: evolution <strong>of</strong> matrix-fibre interface degradation, retrieved from experiments 0-20MPa<br />

(RUD2-5)<br />

The logarithmic evolution <strong>of</strong> ω, as represented in paragraph 4.7, is also used here.<br />

The values <strong>of</strong> C1 <strong>and</strong> C2 are determined by finding the “best fit” <strong>of</strong> equation<br />

(4.51) <strong>with</strong> the experimentally obtained evolution. Results are listed in table 4.7.<br />

Table 4.7: constants in formulation <strong>of</strong> interface degradation evolution, laminate RUD2<br />

specimen stress interval C1 C2<br />

RUD2-5 0.5-20MPa 1.02 0.0135<br />

RUD2-6 0.5-40MPa 0.84 0.0239<br />

RUD2-7 0.5-60MPa 0.67 0.0136<br />

From table 4.7, following conclusions can be formulated:<br />

- From the value <strong>of</strong> C1 it can be seen that serious matrix-fibre degradation<br />

is introduced in the first cycle at high maximum cycle stress (0-60MPa, specimen<br />

RUD2-7). The degradation <strong>of</strong> the matrix-fibre interface in the first loadingunloading-reloading<br />

cycle is considerably lower for specimens, which are applied<br />

to lower σc max (RUD2-5, σc max =20MPa <strong>and</strong> RUD2-6, σc max = 40MPa).<br />

- The value <strong>of</strong> C2, which is a measure <strong>of</strong> matrix-fibre interface degradation<br />

speed when further cycling is applied, is larger for specimen RUD2-6 (0-40MPa)<br />

than for specimen RUD2-7 (0-60MPa). The reason can be found in the fact that<br />

major major-fibre interface degradation already occurs in the first cycle for<br />

specimen RUD2-7 (0-60MPa). This situation is similar to the one that was noticed<br />

for specimens RUD1-10 (10-30MPa) <strong>and</strong> RUD1-11 (20-60MPa), discussed in<br />

previous paragraph (paragraph 4.7).<br />

- If the value <strong>of</strong> C2 <strong>of</strong> specimen RUD2-5 (0-20MPa) is compared to C2 <strong>of</strong><br />

RUD2-6 (0-40MPa), apparently the speed <strong>of</strong> introduction <strong>of</strong> the matrix-fibre<br />

interface degradation is higher for specimen RUD2-6, which has been cycled up to<br />

a higher maximum composite stress.<br />

125


Chapter 4: IPC composite specimens under repeated tensile loading<br />

The question that is still to be answered is: is extrapolation <strong>of</strong> the stiffness<br />

evolution under repeated loading from limited cyclic testing (few applied loading<br />

cycles) possible? Specimen RUD1-9 has been cycled one million times between 0<br />

<strong>and</strong> 40MPa. Specimen RUD2-6 has been cycled one thous<strong>and</strong> times between<br />

0MPa <strong>and</strong> 40MPa. The constants C1 <strong>and</strong> C2 have been determined in both cases.<br />

Their values are listed in table 4.8.<br />

Table 4.8: constants in formulation <strong>of</strong> matrix-fibre interface degradation<br />

specimen stress interval # elapsed<br />

cycles<br />

C1<br />

C2<br />

RUD1-9 0-40MPa one million 0.91 0.019<br />

RUD2-6 0-40MPa one thous<strong>and</strong> 0.84 0.024<br />

discrepancy - - ±10% ±25%<br />

The values <strong>of</strong> the discrepancy on C1 <strong>and</strong> C2 are used further in the discussion <strong>of</strong><br />

the sensitivity <strong>of</strong> theoretical predictions <strong>of</strong> Ecycle,N <strong>and</strong> εc,N max to changing values <strong>of</strong><br />

the material parameters. The material parameters, which are not always known<br />

accurately a priori, are τ0, σR, m, C1 <strong>and</strong> C2. Representative values <strong>of</strong> the<br />

discrepancy on these material parameters are listed in table 4.9.<br />

Table 4.9: representative values <strong>of</strong> discrepancy on material parameters<br />

parameter τ0<br />

(MPa)<br />

σR<br />

(MPa)<br />

m<br />

(-)<br />

C1<br />

(-)<br />

C2<br />

(-)<br />

discrepancy ±25% ±25% ±75% ±10% ±25%<br />

Starting from a reference solution, parameters τ0, σR, m, C1 <strong>and</strong> C2 are varied one<br />

by one. The reference solution, which is used here, is the evolution <strong>of</strong> specimen<br />

RUD1-9. The reference parameters are thus Vf = 14.0 (table 4.1), τ0 = 0.70MPa,<br />

σR = 10MPa, m = 3.4 (table 4.2), C1 = 0.91 <strong>and</strong> C2 = 0.019 (table 4.8). Figure<br />

4.14 shows the theoretical evolution <strong>of</strong> Ecycle,N versus N, when the material<br />

parameters (listed in table 4.9) are varied one by one. The evolution <strong>of</strong> εc,N max is<br />

illustrated in figure 4.15.<br />

Following remarks can be formulated from figures 4.14 <strong>and</strong> 4.15:<br />

- Variations <strong>of</strong> m <strong>and</strong> σR have minor or no influence on the prediction <strong>of</strong><br />

Ecycle,N <strong>and</strong> εc,N max as function <strong>of</strong> N. However, it should be mentioned that the<br />

composites studied here are loaded into the post-cracking zone. Since matrix<br />

multiple cracking is fully developed at initial loading, the statistical development<br />

<strong>of</strong> this matrix cracking is not <strong>of</strong> importance any more. This situation can be<br />

compared <strong>with</strong> the behaviour <strong>of</strong> composite specimens in Chapter 2: once full<br />

matrix multiple cracking occurred, the ACK theory predicts the stress-strain<br />

126


Chapter 4: IPC composite specimens under repeated tensile loading<br />

behaviour as satisfactory as the stochastic cracking model. The information on the<br />

statistical nature <strong>of</strong> matrix cracking is not necessary, when the specimens are<br />

loaded into zone III once.<br />

- If there is a variation on τ0 <strong>of</strong> 25%, the discrepancy between the predicted<br />

<strong>and</strong> measured value <strong>of</strong> Ecycle,N <strong>and</strong> <strong>of</strong> εc,N max is higher than when a variation <strong>of</strong> 10%<br />

is applied on C1.<br />

- When a variation <strong>of</strong> C2 <strong>of</strong> 25% is applied, the error on the prediction <strong>of</strong><br />

Ecycle,N <strong>and</strong> εc,N max is relatively low, compared to variations <strong>of</strong> τ0 <strong>and</strong> C1 in figure<br />

4.14 <strong>and</strong> 4.15. However, when C2 is varied, the discrepancy increases <strong>with</strong><br />

increasing number <strong>of</strong> elapsed load cycle.<br />

Ecycle,N (GPa)<br />

16.5<br />

16<br />

15.5<br />

15<br />

14.5<br />

14<br />

13.5<br />

13<br />

12.5<br />

12<br />

∆C1<br />

∆C2<br />

∆τ0<br />

reference = ∆m = ∆σR<br />

1 100 10000 1000000<br />

# cycles<br />

Figure 4.14 theoretical evolution <strong>of</strong> Ecycle,N as a function <strong>of</strong> the number <strong>of</strong> elapsed cycles<br />

εc,N max (MPa)<br />

0.36<br />

0.35<br />

0.34<br />

0.33<br />

0.32<br />

0.31<br />

∆C1<br />

∆C2<br />

reference = ∆m = ∆σR<br />

∆τ0<br />

1 100 10000 1000000<br />

# cycles<br />

Figure 4.15: theoretical evolution <strong>of</strong> εc,N max as a function <strong>of</strong> the number <strong>of</strong> elapsed cycles<br />

127


Chapter 4: IPC composite specimens under repeated tensile loading<br />

4.8.6.2 repeated loading at lower maximum cycle stress<br />

Two specimens <strong>of</strong> plate RUD2 are loaded in the pre-cracking zone (zone I) or in<br />

the multiple cracking zone (zone II) <strong>of</strong> the ACK theory. The theoretical evolutions<br />

<strong>of</strong> Ecycle,N <strong>and</strong> εc,N max are calculated, according to the stochastic cracking model for<br />

repeated tensile loading, as presented in paragraph 4.4. These theoretical<br />

evolutions are compared <strong>with</strong> the experimental observations.<br />

An estimation <strong>of</strong> matrix-fibre interface degradation evolution parameters C1 <strong>and</strong><br />

C2 (see equation (4.51)) is found in table 4.7. The values <strong>of</strong> C1 <strong>and</strong> C2, obtained<br />

from interpretation <strong>of</strong> the behaviour <strong>of</strong> specimen RUD2-5 (cycled between 0 <strong>and</strong><br />

20MPa) under repeated loading will be used. C1 is thus 1.02 <strong>and</strong> C2 is 0.014 (see<br />

table 4.7). Material parameters Vf, τ0, σR <strong>and</strong> m have been determined in<br />

paragraph 4.8.3.<br />

Figure 4.17a <strong>and</strong> 4.17b show the theoretical evolution <strong>of</strong> Ecycle,N <strong>and</strong> εmax,N versus<br />

the experimental degradation <strong>of</strong> specimen RUD2-9, which is cycled between<br />

1.0MPa <strong>and</strong> 10MPa.<br />

E cycle,N (GPa)<br />

18.6<br />

18.4<br />

18.2<br />

18.0<br />

17.8<br />

17.6<br />

17.4<br />

17.2<br />

experiment<br />

theory<br />

0 200 400 600 800 1000<br />

# cycles<br />

Figure 4.17a: theoretical <strong>and</strong> experimental evolution <strong>of</strong> Ecycle ,N, specimen RUD2-9, 1-10MPa<br />

εc max (%)<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0 200 400 600 800 1000<br />

# cycles<br />

experiment<br />

theory<br />

Figure 4.17b: theoretical <strong>and</strong> experimental evolution <strong>of</strong> εc,N max , specimen RUD2-9, 1-10MPa<br />

128


Chapter 4: IPC composite specimens under repeated tensile loading<br />

It can be seen from figure 4.17a that the evolution <strong>of</strong> Ecycle,N is predicted rather<br />

well. The experimentally measured maximum strains in figure 4.17b are slightly<br />

higher than the theoretical values. From figure 4.17b, one can see that the<br />

experimental maximum cycle strain increases rapidly during the first cycles. This<br />

rapid increment is not predicted theoretically. There are two possible explanations<br />

for this discrepancy.<br />

1. The evolution <strong>of</strong> ω occurs more rapidly for the first few cycles<br />

than predicted.<br />

2. Some extra matrix cracking occurs due to repeated loading in the<br />

first loading cycles. Since full matrix cracking might have not<br />

occurred yet at initial loading, the matrix stresses are still<br />

relatively high at certain parts <strong>of</strong> the composite. Repeated cycling<br />

may introduced extra cracks.<br />

Figure 4.18a <strong>and</strong> 4.18b show the predicted versus experimentally obtained<br />

evolution <strong>of</strong> Ecycle,N <strong>and</strong> εc,N max for specimen RUD2-8, cycled between 1 <strong>and</strong> 5<br />

MPa.<br />

E cycle (GPa)<br />

22.1<br />

22.0<br />

21.9<br />

21.8<br />

21.7<br />

21.6<br />

21.5<br />

21.4<br />

0 200 400 600 800 1000<br />

# cycles<br />

experiment<br />

theory<br />

Figure 4.18a: theoretical <strong>and</strong> experimental evolution <strong>of</strong> Ecycle ,N, 1-5MPa (RUD2-8)<br />

maximum strain (%)<br />

0.035<br />

0.03<br />

0.025<br />

0.02<br />

0.015<br />

0.01<br />

0 200 400 600 800 1000<br />

# cycles<br />

experiment<br />

theory<br />

Figure 4.18b: theoretical <strong>and</strong> experimental evolution <strong>of</strong> εc,N max , 1-5MPa (RUD2-8)<br />

129


Chapter 4: IPC composite specimens under repeated tensile loading<br />

From figures 4.17 <strong>and</strong> 4.18 it can be seen that the comments, formulated on the behaviour<br />

<strong>of</strong> a specimen under repeated loading between 1.0-10MPa (RUD2-9), are also useful for<br />

the specimen loaded between 1.0 <strong>and</strong> 5MPa (RUD2-8).<br />

4.9 UD-reinforced IPC composites under repeated loading:<br />

discussion<br />

From the performed experiments on UD-reinforced specimens following<br />

conclusions can be formulated:<br />

- A damage model, based on the stochastic cracking model (presented in Chapter<br />

2 <strong>and</strong> Chapter 3) is presented in this chapter. The matrix-fibre interface<br />

degradation is adopted as the principal degradation mechanism. A logarithmic law<br />

is proposed for the matrix-fibre interface degradation evolution.<br />

- The evolution <strong>of</strong> Ecycle,N <strong>and</strong> <strong>of</strong> εc,N max is obtained from experimental stress-strain<br />

curves. Testing indicates that the measured evolution <strong>of</strong> Ecycle,N <strong>and</strong> <strong>of</strong> εc,N max <strong>of</strong><br />

specimens, which are subjected to a limited number <strong>of</strong> cycles (about one<br />

thous<strong>and</strong>), can be used to predict the behaviour <strong>of</strong> a specimen under a high<br />

number <strong>of</strong> applied load cycles (about one million) <strong>with</strong> fair accuracy.<br />

- It has been mentioned in paragraph 4.6 that the ACK theory is not used in this<br />

chapter for the interpretation <strong>of</strong> experimental observations. Two reasons can now<br />

be formulated to support this statement:<br />

1. If the maximum cycle stress is situated below the theoretical multiple<br />

cracking stress <strong>and</strong> frictional matrix-fibre interface degradation is assumed to be<br />

the only damage mechanism, the ACK theory does not predict loss <strong>of</strong> stiffness<br />

under repeated loading. The stochastic cracking theory predicts further loss <strong>of</strong><br />

stiffness, even at low cycle stress levels. It has been observed experimentally that<br />

loss <strong>of</strong> stiffness <strong>of</strong> the composite occurs indeed under repeated loading.<br />

2. If on the contrary the maximum cycle stress is high (post-cracking zone),<br />

the ACK theory does not lead to a less complex formulation <strong>of</strong> Ecycle,N <strong>and</strong> εc,N max<br />

than the stochastic cracking theory, as was the case for the behaviour <strong>of</strong> specimens<br />

under monotonic tensile loading.<br />

- The implementation <strong>of</strong> the evolution <strong>of</strong> ω, obtained from cycling into the<br />

post-cracking zone (e.g. cycling between 0 <strong>and</strong> 20MPa) leads to rather good<br />

theoretical predictions <strong>of</strong> the evolution <strong>of</strong> the stiffness properties <strong>of</strong> specimens<br />

cycled in zone I (pre-cracking zone in ACK theory, e.g. cycling between 1 <strong>and</strong><br />

5MPa <strong>and</strong> cycling between 1 <strong>and</strong> 10MPa). It has been mentioned that some<br />

limited extra matrix cracking might occur during the first few loading cycles,<br />

130


Chapter 4: IPC composite specimens under repeated tensile loading<br />

when the maximum cycle stress is situated <strong>with</strong>in the “pre-cracking” <strong>and</strong><br />

“multiple cracking” zones.<br />

4.10 2D-r<strong>and</strong>omly reinforced IPC composite specimens<br />

under repeated loading<br />

4.10.1 introduction<br />

The behaviour <strong>of</strong> 2D-r<strong>and</strong>omly reinforced IPC specimens under repeated loading<br />

is discussed here. The evolution <strong>of</strong> matrix-fibre interface degradation ω, as<br />

obtained on UD-reinforced specimens is used here in the theoretical prediction <strong>of</strong><br />

Ecycle,N <strong>and</strong> εc,N max as a function <strong>of</strong> N. This prediction is calculated, according to<br />

equation (4.37) or (4.43) for Ecycle,N <strong>and</strong> equation (4.34) or (4.36) for εc,N max . The<br />

theoretical predictions are compared <strong>with</strong> experimental observations.<br />

4.10.2 material properties <strong>and</strong> test schedule<br />

A four layer 2D-r<strong>and</strong>omly reinforced laminate is prepared. All specimens are cut<br />

from this plate. Three specimens are subjected to monotonic tensile loading to<br />

obtain material parameters τ0, σR <strong>and</strong> m. Several other specimens are tested under<br />

repeated loading up to four different values <strong>of</strong> σc max : 11MPa, 16.5MPa, 22MPa<br />

<strong>and</strong> 27.5MPa. The minimum cycle stress is 0.5MPa. For each value <strong>of</strong> σc max , two<br />

or three specimens are tested. The measured fibre volume fractions are all situated<br />

between 12.2 <strong>and</strong> 12.4%. The average fibre volume fraction Vf is 12.3%<br />

4.10.3 determination <strong>of</strong> material properties<br />

Figure 4.19 shows the stress-strain curves <strong>of</strong> the three specimens, tested under<br />

monotonic tensile loading up to failure. With help <strong>of</strong> these stress-strain curves, the<br />

average values <strong>of</strong> m, τ0 <strong>and</strong> σR are obtained, as explained in paragraph 2.5.4. The<br />

resulting values are: m = 3.4, τ0 = 1.3MPa <strong>and</strong> σR = 12MPa. The average<br />

composite failure stress σcu is 34MPa.<br />

stress (MPa)<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 0.5 1 1.5 2<br />

strain (%)<br />

Figure 4.19: stress-strain curve under monotonic tensile loading, RR1<br />

131


Chapter 4: IPC composite specimens under repeated tensile loading<br />

4.10.4 results under repeated loading<br />

The specimens loaded up to 11MPa are loaded into their theoretical ACK multiple<br />

cracking zone. The three higher maximum stresses (16.5MPa, 22MPa, 27.5MPa)<br />

are situated in zone III <strong>of</strong> the tensile curve <strong>of</strong> the laminate. For each stress level,<br />

two or three specimens are tested. The width <strong>of</strong> the specimens is 30mm, the<br />

average thickness about 4mm <strong>and</strong> the length is 250mm. Cycling <strong>of</strong> the specimens<br />

is performed by a MTS hydraulic testing system, <strong>with</strong> a sinusoidal load at a<br />

frequency <strong>of</strong> 5Hz. Both the force <strong>and</strong> strain are measured at a sampling rate <strong>of</strong><br />

400Hz during the test. The load <strong>and</strong> strain are measured for 10 successive cycles<br />

at cycle numbers 10, 100, 500, 1000, 5000, 10000, ... until failure.<br />

Figure 4.20 shows the number <strong>of</strong> loading cycles, which have been applied at<br />

failure for each stress level. The black squares indicate that the specimens failed<br />

during testing. The grey squares indicate that cycling was stopped after about one<br />

million cycles (regarding availability <strong>of</strong> the test set-up), before failure occurred.<br />

max<br />

σc (MPa)<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

1.E+00 1.E+02 1.E+04 1.E+06 1.E+08<br />

# cycles<br />

Figure 4.20: number <strong>of</strong> cycles at failure for different stress levels<br />

The experimentally obtained evolutions <strong>of</strong> Ecycle,N are shown in figures 4.21 to<br />

4.24. If the last point in the curve is a cross inside a grey box, it means the<br />

specimen failed. As can be seen from these figures, the reproducibility <strong>of</strong> the<br />

fatigue test is rather good, <strong>with</strong> the exception <strong>of</strong> figure 4.21. In this figure, the<br />

triangle shaped curve is however obtained from a specimen which has been cycled<br />

up to 16.5MPa for the first 10 cycles <strong>and</strong> then subsequently repeatedly up to<br />

11MPa. The difference between this specimen <strong>and</strong> the specimens, which are<br />

cycled initially up to 11MPa is distinct for the first thous<strong>and</strong>s <strong>of</strong> cycles. After<br />

more than 10000 loading cycles, this discrepancy is vanishing.<br />

132


Chapter 4: IPC composite specimens under repeated tensile loading<br />

Ecycle,N (GPa)<br />

11<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

1.E+00 1.E+02 1.E+04 1.E+06<br />

# cycles<br />

Figure 4.21: evolution <strong>of</strong> Ecycle,N for specimens tested up to 11MPa<br />

Ecycle,N (GPa)<br />

7.0<br />

6.5<br />

6.0<br />

5.5<br />

5.0<br />

4.5<br />

4.0<br />

3.5<br />

3.0<br />

1.E+00 1.E+02 1.E+04 1.E+06<br />

# cycles<br />

Figure 4.22: Evolution <strong>of</strong> Ecycle,N for specimens tested up to 16.5MPa<br />

4.4<br />

4.2<br />

4.0<br />

3.8<br />

3.6<br />

3.4<br />

3.2<br />

3.0<br />

2.8<br />

1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05<br />

# cycles<br />

E cycle,N (GPa)<br />

Figure 4.23: Evolution <strong>of</strong> Ecycle,N for specimens tested up to 22MPa<br />

4.2<br />

4.0<br />

3.8<br />

3.6<br />

3.4<br />

3.2<br />

3.0<br />

2.8<br />

2.6<br />

Ecycle,N (GPa)<br />

1.E+00 1.E+02 1.E+04 1.E+06<br />

# cycles<br />

Figure 4.24: Evolution <strong>of</strong> Ecycle,N for specimens tested up to 27.5MPa<br />

133


Chapter 4: IPC composite specimens under repeated tensile loading<br />

All specimens cycled up to 22MPa or 27.5MPa<br />

failed. At the fracture surface, fibres stick out <strong>of</strong><br />

the matrix for all broken specimens. This is an<br />

indication that failure <strong>of</strong> the composite occurs due<br />

to fibre pull-out rather than fibre fracture. Figure<br />

4.25 shows such a fracture surface <strong>of</strong> a composite,<br />

which has been tested under repeated loading.<br />

Figure 4.26 <strong>and</strong> 4.27 show the overview <strong>of</strong> the<br />

average evolution <strong>of</strong> Ecycle,N <strong>and</strong> εc,N min <strong>with</strong> the<br />

number <strong>of</strong> elapsed loading cycles for all tested<br />

stress-levels respectively. A black cross in a grey<br />

box indicates failure <strong>of</strong> the specimen.<br />

E cycle,N (GPa)<br />

11<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

10mm<br />

Figure 4.25: fracture surface<br />

11MPa<br />

17.5MPa<br />

22MPa<br />

27.5MPa<br />

2<br />

1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05 1.E+06<br />

# cycles<br />

Figure 4.26: overview <strong>of</strong> the average evolution <strong>of</strong> Ecycle,N for all tested stress levels<br />

εc min (%)<br />

0.35<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05 1.E+06<br />

# cycles<br />

11MPa<br />

16.5MPa<br />

22MPa<br />

27.5MPa<br />

Figure 4.27: overview <strong>of</strong> the average evolution <strong>of</strong> the residual strains for all tested stress levels<br />

134


Chapter 4: IPC composite specimens under repeated tensile loading<br />

4.10.5 theoretical versus experimental degradation behaviour <strong>of</strong> 2Dr<strong>and</strong>omly<br />

reinforced specimens<br />

4.10.5.1 results<br />

In this paragraph the experimental evolutions <strong>of</strong> Ecycle,N obtained on the 2Dr<strong>and</strong>omly<br />

reinforced specimens are compared <strong>with</strong> the theoretical predictions. The<br />

stochastic cracking (based) model - <strong>with</strong> assumption <strong>of</strong> matrix-fibre interface<br />

degradation being the only fatigue mechanism - is used. The material parameters<br />

Vf, m, τ0, σR are determined in paragraph 4.10.2 <strong>and</strong> 4.10.3. The theoretical<br />

evolution <strong>of</strong> Ecycle,N is calculated <strong>with</strong> equation (4.37) or (4.43). C1 <strong>and</strong> C2<br />

determine the evolution <strong>of</strong> ω. These parameters are set equal to the values<br />

obtained for UD-reinforced specimens. The value <strong>of</strong> these constants are set equal<br />

to the values, obtained on specimen RUD2-5 (cycled between 0 <strong>and</strong> 20MPa)<br />

earlier. C1 is thus 1.0 <strong>and</strong> C2 is 0.014 (see table 4.7).<br />

4.10.5.2 specimens cycled up to 22 or 27.5MPa: discussion<br />

The experimental observations on 2D-r<strong>and</strong>omly reinforced specimens, which have<br />

been loaded repeatedly up to 22MPa, are similar to the observations on specimens,<br />

which have been cycled up to 27.5MPa. Therefore, these specimens are discussed<br />

together.<br />

Figures 4.28a <strong>and</strong> 4.28b represent the theoretical prediction <strong>of</strong> the evolution <strong>of</strong><br />

Ecycle,N for specimens cycled up to 22MPa <strong>and</strong> 27.5MPa versus the experimentally<br />

obtained curve respectively.<br />

Ecycle,N (Gpa)<br />

5<br />

4.5<br />

4<br />

3.5<br />

3<br />

2.5<br />

2<br />

experimental<br />

1.5<br />

1<br />

0.5<br />

0<br />

theoretical<br />

1.E+00 1.E+02 1.E+04 1.E+06<br />

# cycles<br />

Figure 4.28a: theoretical versus experimental<br />

evolution Ecycle,N for 2D-r<strong>and</strong>omly reinforced<br />

specimens tested up to 22MPa<br />

Ecycle,N (Gpa)<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

experimental<br />

theoretical<br />

1.E+00 1.E+02 1.E+04 1.E+06<br />

# cycles<br />

Figure 4.28b: theoretical versus experimental<br />

evolution Ecycle for 2D-r<strong>and</strong>omly reinforced<br />

specimens tested up to 27.5MPa<br />

If the maximum cycle stress is 22MPa or 27.5MPa, the descending trend <strong>of</strong> the<br />

experimental evolution <strong>of</strong> Ecycle,N versus N is much steeper than that <strong>of</strong> the<br />

theoretical curve: degradation <strong>of</strong> the specimens occurs much faster than predicted.<br />

There might be two reasons for this behaviour:<br />

1. the degradation <strong>of</strong> the matrix-fibre interface occurs faster<br />

than expected<br />

135


Chapter 4: IPC composite specimens under repeated tensile loading<br />

2. another damage mechanism interferes <strong>with</strong> the matrixfibre<br />

interface degradation.<br />

Let us initially assume matrix-fibre interface degradation is still the sole damage<br />

mechanism. The evolution <strong>of</strong> the specimens cycled up to 22MPa is considered.<br />

From figure 4.28a, it can be seen that the value <strong>of</strong> ω after 1000 load cycles is<br />

seriously overestimated, since the predicted Ecycle,N is considerably higher than the<br />

experimental value, after 1000 load cycles have been applied. In paragraph 4.5.3,<br />

the equations have been derived for the determination <strong>of</strong> ω from the experimental<br />

value <strong>of</strong> Ecycle,N. The experimental value <strong>of</strong> Ecycle,N at load cycle 1000 is 2.99GPa.<br />

When this value is now inserted in equation (4.50) <strong>of</strong> paragraph 4.5.3, the obtained<br />

value <strong>of</strong> ω is about –0.1. Therefore, when the value <strong>of</strong> ω is calculated from the<br />

experimentally obtained value <strong>of</strong> Ecycle,N at failure, a negative value <strong>of</strong> ω is found,<br />

which is physically not possible. The assumption on matrix-fibre interface<br />

degradation being the sole fatigue mechanism should therefore be ab<strong>and</strong>oned here.<br />

In paragraph 4.10.4, it has been mentioned that composite specimen failure occurs<br />

due to fibre pull-out rather than fibre failure. The most plausible explanation for<br />

the rapid decrease <strong>of</strong> Ecycle,N is thus that matrix-fibre interface degradation <strong>and</strong><br />

fibre pull-out co-exist as damage mechanisms for the studied 2D-r<strong>and</strong>omly<br />

reinforced specimens cycled up to 22 or 27.5MPa.<br />

In design calculations, where repeated loading is to be expected, the tensile stress<br />

in 2D-r<strong>and</strong>omly oriented short-fibre (50mm fibre length) reinforced IPC is not to<br />

equal or exceed 22MPa, since this would lead to early failure upon cycling.<br />

4.10.5.3 specimens cycled up to 11 or 16.5MPa: discussion<br />

The experimental observations on the 2D-r<strong>and</strong>omly reinforced specimens, which<br />

have been loaded repeatedly up to 11MPa, are similar to the observations on the<br />

specimens, which have been cycled up to 16.5MPa. Therefore, these specimens<br />

are discussed together.<br />

Figures 4.29a <strong>and</strong> 4.29b represent the theoretical prediction <strong>of</strong> Ecycle,N for<br />

specimens cycled up to 11MPa <strong>and</strong> 16.5MPa versus the experimentally obtained<br />

curve respectively.<br />

From figures 4.29a <strong>and</strong> 4.29b, it can be noticed that the slope <strong>of</strong> the experimental<br />

<strong>and</strong> the theoretical evolution <strong>of</strong> Ecycle,N versus N are practically equal, once one<br />

thous<strong>and</strong> load cycles are applied, if the maximum cycle stress is 11 or 16.5MPa. If<br />

the specimens are cycled up to 11 (figure 4.28a) or 16.5MPa (figure 4.28b), the<br />

experimental value <strong>of</strong> Ecycle,N is initially higher than the theoretical obtained value.<br />

This effect might be the result <strong>of</strong> prevented crack closure, as has already been<br />

discussed in Chapter 3. However, both figures show the experimentally obtained<br />

136


Chapter 4: IPC composite specimens under repeated tensile loading<br />

curve <strong>of</strong> Ecycle,N decreases rapidly during the first one hundred to one thous<strong>and</strong><br />

load cycles. This indicates another damage mechanism might occur at these stress<br />

levels, during the first one hundred to one thous<strong>and</strong> load cycles. This last<br />

statement will be discussed more in detail in paragraph 4.10.5.3.<br />

(Gpa)<br />

E cycle,N<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

1.E+00 1.E+02 1.E+04 1.E+06<br />

# cycles<br />

experimental<br />

theoretical<br />

Figure 4.29a: theoretical versus experimental<br />

evolution Ecycle,N for 2D-r<strong>and</strong>omly reinforced<br />

specimens tested up to 22MPa<br />

E cycle,N (GPa)<br />

7<br />

experimental<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

theoretical<br />

1.E+00 1.E+02<br />

# cycles<br />

1.E+04 1.E+06<br />

Figure 4.29a: theoretical versus experimental<br />

evolution Ecycle for 2D-r<strong>and</strong>omly reinforced<br />

specimens tested up to 27.5MPa<br />

4.10.5.4 specimen cycled up 6MPa: discussion<br />

One specimen is cycled between 2 <strong>and</strong> 6MPa up to 1 million cycles. Figure 3.30<br />

shows the measured <strong>and</strong> predicted evolution <strong>of</strong> Ecycle,N. As can be seen from this<br />

figure, the experimental decrease <strong>of</strong> Ecycle occurs considerably faster than<br />

theoretically predicted for the first one hundred to one load thous<strong>and</strong> cycles. After<br />

about one hundred to one thous<strong>and</strong> load cycles are applied, the slope <strong>of</strong> the<br />

experimental curve <strong>of</strong> the evolution <strong>of</strong> Ecycle,N becomes less steep <strong>and</strong> becomes<br />

more equal to the slope <strong>of</strong> the theoretical curve. This is again an indication that<br />

another damage mechanism might interact <strong>with</strong> matrix-fibre interface degradation<br />

during the first loading cycles. To obtain a better idea <strong>of</strong> the meso-mechanics,<br />

occurring upon cycling, pictures are taken <strong>of</strong> the surface <strong>of</strong> the specimen during<br />

cycling, after the specimen has been coated <strong>with</strong> an ink solution each time. Figure<br />

3.31 shows pictures taken at virgin state, after first loading <strong>and</strong> at load cycle 2500.<br />

E cycle<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

1.E+00 1.E+02 1.E+04 1.E+06 1.E+08<br />

# cycles<br />

Figure 4.30: theoretical versus experimental evolution Ecycle,N for 2D-r<strong>and</strong>omly reinforced<br />

specimens tested up to 6MPa<br />

137


Chapter 4: IPC composite specimens under repeated tensile loading<br />

(a)<br />

0MPa, virgin specimen<br />

(b)<br />

6MPa, first loading cycle<br />

(c)<br />

6MPa, loading cycle 2500<br />

Figure 4.31: introduction <strong>of</strong> matrix cracking due to repeated cycling between 2 <strong>and</strong> 6MPa<br />

From figures 4.30 <strong>and</strong> 4.31 it can be seen that extra matrix cracks <strong>and</strong> crack<br />

extensions are introduced during cycling, when a 2D-r<strong>and</strong>omly reinforced<br />

specimen is loaded up to 6MPa (pre-cracking zone <strong>of</strong> the ACK theory). In figure<br />

4.32 a new proposal is illustrated for description <strong>of</strong> the evolution <strong>of</strong> Ecycle,N for 2Dr<strong>and</strong>omly<br />

reinforced specimens, when σc max is situated in the theoretical ACK precracking<br />

zone.<br />

Ecycle,N<br />

matrix cracking &<br />

matrix-fibre interface<br />

degradation regime<br />

matrix-fibre interface<br />

degradation regime<br />

log(cycles)<br />

Figure 4.32: proposal theoretical evolution Ecycle,N for 2D-r<strong>and</strong>omly reinforced specimens,<br />

σc max situated in the theoretical ACK pre-cracking zone<br />

4.10.6 discussion <strong>and</strong> conclusions<br />

The behaviour <strong>of</strong> IPC composite specimens <strong>with</strong> 2D-r<strong>and</strong>omly oriented short fibre<br />

reinforcement (50mm length) is studied in this paragraph. The values <strong>of</strong> the<br />

matrix-fibre interface damage evolution parameters (C1 <strong>and</strong> C2), obtained from<br />

experiments on UD-reinforced specimens are inserted into the theoretical<br />

formulation <strong>of</strong> the evolution <strong>of</strong> Ecycle,N <strong>and</strong> εc,N max for 2D-r<strong>and</strong>omly reinforced<br />

specimens. The main observations from comparison <strong>of</strong> experimental <strong>and</strong><br />

theoretical stress-strain curves are:<br />

- When the maximum cycle stress is situated well into the post-cracking zone <strong>of</strong><br />

the ACK theory (in this case 22MPa <strong>and</strong> 27.5MPa), the hypothesis <strong>of</strong> matrix-fibre<br />

138


Chapter 4: IPC composite specimens under repeated tensile loading<br />

interface degradation being the only damage mechanism is not appropriate. Fibre<br />

pull-out occurs from the very first load cycle. The combination <strong>of</strong> matrix-fibre<br />

interface degradation <strong>and</strong> fibre pull-out lead to final failure <strong>of</strong> the composite at one<br />

thous<strong>and</strong> (σc max = 27.5MPa) to ten thous<strong>and</strong> (σc max = 22MPa) loading cycles. The<br />

average tensile strength under monotonic tensile loading is about 35MPa.<br />

- When the maximum cycle stress is situated below or around the theoretical ACK<br />

multiple cracking stress (6MPa, 11MPa or 16.5MPa), the hypothesis <strong>of</strong> matrixfibre<br />

degradation being the sole damage mechanism can still be justified, once one<br />

hundred to one thous<strong>and</strong> load cycles have been applied, according to the<br />

comparison <strong>of</strong> the experimental <strong>and</strong> theoretical stress-strain curves. However,<br />

extra matrix cracks are introduced during the first one hundred to one thous<strong>and</strong><br />

cycles. This phenomenon is monitored from the evolution <strong>of</strong> the stress-strain curve<br />

<strong>and</strong> visually under the stereomicroscope.<br />

4.11 Interpretation <strong>of</strong> hysteresis<br />

Until now, Ecycle,N <strong>and</strong> εc,N max have been used to describe the evolution <strong>of</strong> the<br />

stiffness <strong>of</strong> the composite. However, it should be stressed that the stochastic<br />

cracking (based) theory can be used to predict the whole shape <strong>of</strong> the stress-strain<br />

unloading-reloading hysteresis loop. It has been mentioned in paragraph 4.2.1 that<br />

Rouby <strong>and</strong> Reynaud (1992) concluded that degradation <strong>of</strong> the matrix-fibre<br />

interface leads to increasing failure probability <strong>of</strong> the fibres <strong>and</strong> a widening <strong>of</strong> the<br />

hysteresis loop. Figure 4.33 illustrates this is not the case for all tested IPC<br />

specimens in this work, since this measured hysteresis loop tightens <strong>with</strong><br />

increasing number <strong>of</strong> load cycles.<br />

stress (MPa)<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

cycle 1<br />

cycle 100<br />

0 0.1 0.2 0.3 0.4 0.5<br />

strain (%)<br />

Figure 4.33: tightening <strong>of</strong> stress-strain hysteresis loop <strong>of</strong> IPC composite specimen <strong>with</strong> increasing<br />

number <strong>of</strong> load cycles, 2D-r<strong>and</strong>omly reinforced IPC specimen<br />

Cuypers (2001) verified that stress-strain loop widening is an indication towards<br />

partial matrix-fibre unloading slip, while loop tightening gives an indication that<br />

full matrix-fibre slip occurs during unloading. The verification <strong>of</strong> this statement is<br />

139


Chapter 4: IPC composite specimens under repeated tensile loading<br />

enclosed in Appendix 3 <strong>and</strong> is based on documents <strong>of</strong> Keer (1981 <strong>and</strong> 1985),<br />

Rouby <strong>and</strong> Reynaud (1993), Pryce <strong>and</strong> Smith (1993) <strong>and</strong> Ahn <strong>and</strong> Curtin (1997),<br />

discussing hysteresis effects in fibre cement composites.<br />

From figure 4.37, another conclusion can be formulated. The stress-strain loop <strong>of</strong><br />

the first loading cycle is not nicely closed, as can be seen at the maximum cycle<br />

stress. This indicates considerable damage occurs <strong>with</strong>in this first loading cycle.<br />

However, stress-strain cycle 100 is nicely closed, which means no or minor<br />

damage occurs <strong>with</strong>in this load cycle, as has been mentioned in paragraph 4.3.2.<br />

4.12 Conclusions<br />

The evolution <strong>of</strong> two IPC composite stiffness parameters is studied upon cycling:<br />

the linearised cycle stiffness (Ecycle,N) <strong>and</strong> the strain at maximum composite stress<br />

(εc max ). A damage model is presented in this chapter, based on the stochastic<br />

cracking model under monotonic tensile loading <strong>and</strong> unloading as presented in<br />

Chapter 2 <strong>and</strong> Chapter 3. The degradation <strong>of</strong> the matrix-fibre interface is<br />

formulated as being a function <strong>of</strong> the logarithm <strong>of</strong> the number <strong>of</strong> elapsed cycles.<br />

Due to this logarithmic function, two parameters (C1 <strong>and</strong> C2) determine the<br />

damage introduced in the first load cycle <strong>and</strong> the speed <strong>of</strong> further degradation due<br />

to further repeated loading. This formulation has been verified experimentally on<br />

UD-reinforced <strong>and</strong> 2D-r<strong>and</strong>omly reinforced specimens. Some conclusions can be<br />

formulated from comparison <strong>of</strong> theory <strong>and</strong> experiments. All conclusions are<br />

preliminary conclusions <strong>and</strong> most <strong>of</strong> them require further extended verification.<br />

4.12.1 Conclusions from tests on UD-reinforced specimens<br />

- One argument has been found, which defends the assumption <strong>of</strong> the frictional<br />

matrix-fibre interface degradation being the sole damage mechanism under<br />

repeated loading for UD-reinforced IPC laminates, after initial loading occurred.<br />

The evolution <strong>of</strong> the matrix-fibre interface degradation parameter ω can be<br />

obtained from the measurement <strong>of</strong> Ecycle,N <strong>and</strong> εc,N max <strong>of</strong> a specimen. Ecycle,N <strong>and</strong><br />

εc,N max are determined independently from each other from an experimental stressstrain<br />

curve <strong>of</strong> a specimen. The evolution <strong>of</strong> ω is thus determined in two<br />

independent ways. It has been verified that the evolutions <strong>of</strong> ω, determined from<br />

Ecycle,N <strong>and</strong> from εc,N max , show rather good coincidence. This is an indication that<br />

indeed it is the evolution <strong>of</strong> the frictional matrix-fibre interface shear stress that<br />

determines the evolution <strong>of</strong> the remaining stiffness <strong>of</strong> the composite under<br />

repeated loading. (Ecycle,N <strong>and</strong> εc,N max are both concerned to the stiffness).<br />

- When the matrix-fibre interface degradation evolution parameters (C1 <strong>and</strong> C2)<br />

are determined from experimental results under limited cycling (typically one<br />

140


Chapter 4: IPC composite specimens under repeated tensile loading<br />

thous<strong>and</strong>), these values have been inserted in the theoretical formulations <strong>of</strong><br />

Ecycle,N <strong>and</strong> εc,N max for prediction <strong>of</strong> Ecycle,N <strong>and</strong> εc,N max after one million cycles <strong>with</strong><br />

good result.<br />

4.12.2 conclusions from tests on composite specimens <strong>with</strong> shortfibre<br />

2D-r<strong>and</strong>omly oriented reinforcement<br />

- When relatively high maximum cycle stress (theoretical ACK post-cracking<br />

zone) is applied, combined matrix-fibre interface degradation <strong>and</strong> fibre failure or<br />

pull-out occur, which might lead to final composite failure after a several<br />

thous<strong>and</strong>s <strong>of</strong> load cycles have been applied.<br />

- When low or intermediate maximum cycle stress (theoretical ACK pre-cracking<br />

<strong>and</strong> multiple cracking zones) are applied, combined matrix-fibre interface<br />

degradation <strong>and</strong> matrix cracking occurs <strong>with</strong>in the first one hundred to one<br />

thous<strong>and</strong> cycles. If further repeated loading occurs, matrix-fibre interface<br />

degradation is indeed again the sole degradation mechanism.<br />

4.13 References<br />

B.K. Ahn <strong>and</strong> W.A Curtin, Strain <strong>and</strong> hysteresis by stochastic matrix<br />

cracking in ceramic matrix composites, J. Mech. Phys. Solids, Vol. 45, No. 2,<br />

1997 pp.177-209<br />

J. Aveston, G.A. Cooper <strong>and</strong> A Kelly, Single <strong>and</strong> multiple fracture, The<br />

Properties <strong>of</strong> Fibre Composites, Proc. Conf. National Physical Laboratories, IPC<br />

Science & Technology Press Ltd. London, 1971, pp.15-24<br />

J. Aveston <strong>and</strong> A. Kelly, Theory <strong>of</strong> multiple fracture <strong>of</strong> fibrous composites,<br />

J. Mat. Sci., Vol. 8, 1973, pp. 411-461<br />

J. Aveston, R.A. Mercer, J.M. Sillwood, Fibre reinforced cements –<br />

scientific foundations for specifications, In Composites – St<strong>and</strong>ards, Testing<br />

<strong>and</strong> <strong>Design</strong>, Proc. National Physical Laboratories Conference, UK, 1974,<br />

pp.93-103<br />

P. Bauweraerts, J. Wastiels, X. Wu, H. Cuypers <strong>and</strong> J. Gu, Evaluation <strong>of</strong><br />

damage accumulation <strong>of</strong> glass fibre reinforced brittle matrix composite after<br />

cyclic loading, Durability <strong>of</strong> Composites for Construction, Quebec, august 5-7,<br />

1998a<br />

P. Bauweraerts, Aspects <strong>of</strong> the Micromechanical Characterisation <strong>of</strong> Fibre<br />

Reinforced <strong>Brittle</strong> <strong>Matrix</strong> Composites, Phd. thesis, VUB 1998b<br />

141


Chapter 4: IPC composite specimens under repeated tensile loading<br />

H. Cuypers, Use <strong>of</strong> a stochastic cracking based model on the behaviour<br />

<strong>of</strong> 2D-r<strong>and</strong>om reinforced IPC composite specimens, internal report,<br />

department MEMC, Vrije Universiteit Brussel, 2001<br />

W.A. Curtin, Stochastic Damage Evolution <strong>and</strong> Failure in Fibre-<br />

Reinforced Composites, Advances in Applied Mechanics; Volume 36; 1999;<br />

pp.163-253<br />

A.G. Evans, F.W. Zok, R.M. McMeeking, Fatigue <strong>of</strong> ceramic matrix<br />

composites, Acta metallurgica et materialia, 43, 1995, pp.859-875<br />

J. G. Keer, Some observations on hysteresis effects in fibre cement<br />

composites, Journal <strong>of</strong> Material Science letters, Vol. 4, 1985, pp. 363-366<br />

J. G. Keer, behaviour <strong>of</strong> cracked fibre composites under limited cyclic<br />

loading, International Journal <strong>of</strong> Cement Composites, Vol. 3, 1981, pp. 197-186<br />

A.W. Pryce <strong>and</strong> P.A. Smith, <strong>Matrix</strong> cracking in uniderictional ceramic<br />

composites under quasi-static <strong>and</strong> cyclic loading, Acta metall. mater., No. 41,<br />

1993, pp.1269-1281<br />

D. Rouby <strong>and</strong> P. Reynaud, Fatigue Behaviour related to interface<br />

modification during load cycling in ceramic-matrix fibre composites, Composite<br />

Science <strong>and</strong> Technology, 48; 1993, pp.109-118<br />

142


Chapter 5<br />

5.1 Introduction<br />

S<strong>and</strong>wich modelling<br />

S<strong>and</strong>wich panels are construction elements, <strong>of</strong>fering high stiffness to weight ratio.<br />

Common s<strong>and</strong>wich panels for building purposes consist <strong>of</strong> a cellular core <strong>and</strong> two<br />

faces. Diversity in face materials in construction <strong>of</strong> buildings is rather limited at<br />

present. S<strong>and</strong>wich panels <strong>with</strong> steel faces are frequently used in industrial<br />

complexes. Hardboard faces are regularly used in housing applications for<br />

esthetical <strong>and</strong> economical reasons.<br />

This chapter discusses some computational methods (analytical <strong>and</strong> numerical),<br />

useful for optimum design <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> a new type <strong>of</strong> face material: Eglass<br />

fibre reinforced cementitious composites.<br />

The implementation <strong>of</strong> several finite element models <strong>and</strong> <strong>of</strong> the face material<br />

behaviour in ANSYS is discussed. This widespread finite element package is used<br />

for analysis <strong>and</strong> design <strong>of</strong> s<strong>and</strong>wich panels in this work. The constitutive<br />

equations, derived for IPC composites in previous chapters, are introduced into<br />

finite element s<strong>and</strong>wich calculations in this chapter.<br />

5.2 S<strong>and</strong>wich modelling: overview<br />

Many research groups working on s<strong>and</strong>wich structures developed a particular<br />

s<strong>and</strong>wich model that fits the specific needs <strong>and</strong> interests <strong>of</strong> the studied application.<br />

Burton <strong>and</strong> Noor (1995), Noor et al. (1996) <strong>and</strong> Ha (1990) wrote extensive state<strong>of</strong>-the-art<br />

documents on existing models <strong>and</strong> their applications.<br />

Most models can be roughly classified in three classes <strong>of</strong> approximation:<br />

simplified models, continuum models <strong>and</strong> plate <strong>and</strong> beam models<br />

143


Chapter 5: S<strong>and</strong>wich modelling<br />

Simplified models are used only when one is interested in a specific global<br />

response value <strong>of</strong> the s<strong>and</strong>wich panel, like global buckling load, resonance<br />

frequency or face wrinkling load. These models are frequently used to ensure that<br />

the studied s<strong>and</strong>wich panel does not fail due to an instability phenomenon, like<br />

buckling or wrinkling.<br />

If a continuum approach is used, no a priori assumptions are made on the<br />

displacements, stresses or strains across the thickness <strong>of</strong> the s<strong>and</strong>wich. If enough<br />

finite elements are used along the length <strong>and</strong> across the thickness <strong>and</strong> width <strong>of</strong> a<br />

s<strong>and</strong>wich panel, this method provides accurate stresses, strains <strong>and</strong> displacements<br />

at any point <strong>of</strong> the s<strong>and</strong>wich, but at the cost <strong>of</strong> higher analysis-effort. However,<br />

this type <strong>of</strong> model can be very useful as a benchmark for comparison <strong>of</strong> accuracy<br />

<strong>of</strong> the more simplified models.<br />

Different plate <strong>and</strong> beam models are based on different assumptions <strong>of</strong> the<br />

through-thickness displacements <strong>and</strong>/or stresses. They are <strong>of</strong>ten categorised into<br />

single layer theories, layer-wise theories, <strong>and</strong> predictor-corrector methods.<br />

Several types <strong>of</strong> s<strong>and</strong>wich modelling <strong>with</strong> different level <strong>of</strong> complexity are<br />

presented further in this chapter. Their feasibility to predict the behaviour <strong>of</strong><br />

s<strong>and</strong>wich panels <strong>with</strong> cementitious composite faces is discussed.<br />

5.2.1 single layer theories<br />

If a single layer theory is used, an equivalent single layer plate replaces the<br />

s<strong>and</strong>wich or shell. Global through-the-thickness approximations are introduced for<br />

the displacements, strains <strong>and</strong>/or stresses. Within the single layer theories, higher<br />

order displacement functions are sometimes used to allow for some warping <strong>of</strong> the<br />

cross section. Since in the first-order shear deformation theories, the transverse<br />

shear strains or stresses are assumed to be constant across the whole thickness <strong>of</strong><br />

the s<strong>and</strong>wich, correction factors are used to adjust the transverse shear stiffness.<br />

The appropriateness <strong>of</strong> these shear deformation theories generally depends highly<br />

on the factors used to correct the shear stiffness. References using single layer<br />

theories can be found in Noor et al. (1996).<br />

5.2.2 layer-wise theories<br />

Both discrete layer theories <strong>and</strong> zig-zag theories are categorised in the layer-wise<br />

theories. They are both based on the assumption <strong>of</strong> a certain displacement field for<br />

each separate layer instead <strong>of</strong> the whole s<strong>and</strong>wich panel thickness, as is the case<br />

for single layer theories. The main difference between these two theories is found<br />

in the fact that in the discrete layer theories each extra layer introduces extra<br />

degrees <strong>of</strong> freedom in the plate or beam element. In the zig-zag theory, the number<br />

<strong>of</strong> degrees <strong>of</strong> freedom remains constant by analytically satisfying continuity<br />

conditions at the interface for each new layer <strong>of</strong> the laminate. The displacement<br />

field across the thickness <strong>of</strong> one layer can be a first or higher order function. In<br />

144


Chapter 5: S<strong>and</strong>wich modelling<br />

general, both the discrete layer theory <strong>and</strong> the zig-zag theory lead to a good<br />

prediction <strong>of</strong> the distribution <strong>of</strong> the stresses in the different layers. Many<br />

references on layer-wise theories can be found in Noor et al. (1996)<br />

5.2.3 predictor-corrector theories <strong>and</strong> hierarchical modelling theories<br />

Also quite recently, some algorithms are developed for the automatic optimisation<br />

<strong>of</strong> shear correction factors or <strong>of</strong> the thickness distribution <strong>of</strong> the displacements or<br />

their derivatives. This method is a predictor-corrector method. When no a priori<br />

assumption is made on the order <strong>of</strong> the displacement field, the predictor-corrector<br />

method is used to establish the order <strong>of</strong> this field. This method is <strong>of</strong>ten referred to<br />

as hierarchical modelling. The sensitivity <strong>of</strong> the response <strong>of</strong> the structure to<br />

variations <strong>of</strong> each parameter <strong>of</strong> the model is checked, when its complexity is<br />

increased. This promising approach differs totally from the classical way <strong>of</strong> using<br />

finite elements <strong>with</strong> a predefined shape function <strong>and</strong> requires further investigation.<br />

Up to now these predictor-corrector methods are developed for linear elastic<br />

materials, so they are not discussed further in this thesis. Information on the<br />

implementation <strong>of</strong> this method in s<strong>and</strong>wich calculations can be found in papers<br />

written by Noor et al. (1994), Actis et al. (1999), Schwab (1999) <strong>and</strong> Sze et al.<br />

(2000).<br />

5.3 A 2D-continuum approach: plane stress versus plane<br />

strain<br />

The s<strong>and</strong>wich panels, studied in this work, are used to span one direction. The<br />

most accurate approach to model a s<strong>and</strong>wich panel is found in using 3D<br />

continuum finite elements <strong>and</strong> separately discretise the different s<strong>and</strong>wich layers.<br />

In most cases this method provides results <strong>with</strong> good accuracy, but it will also lead<br />

to an extensive FE (finite element) mesh. As a result, this method is only<br />

applicable for small <strong>and</strong> simple s<strong>and</strong>wich panels. As soon as plasticity or damage<br />

mechanics get involved <strong>and</strong>/or optimisation loops are made, this might lead to<br />

extensive computing time. However, when a s<strong>and</strong>wich panel is used as a wide<br />

s<strong>and</strong>wich beam, 2D structural solid elements can be used along the length <strong>and</strong><br />

across the thickness <strong>of</strong> the panel. Provided the load on the s<strong>and</strong>wich is a constant<br />

pressure load across the width <strong>of</strong> the s<strong>and</strong>wich (z-axis in figure 5.1), a 2D plane<br />

stress or plane strain assumption can be used to represent the width.<br />

Z<br />

Y<br />

X<br />

?<br />

Figure 5.1: z-axis: plane stress versus plane strain<br />

145


Chapter 5: S<strong>and</strong>wich modelling<br />

The transverse behaviour <strong>of</strong> E-glass fibre reinforced IPC under monotonic tensile<br />

loading has been discussed in Chapter 2. In paragraph 2.7, it has been verified that<br />

transverse strains <strong>and</strong> stresses never become high due to the introduction <strong>of</strong> matrix<br />

multiple cracking. Poisson’s ratio reaches zero for UD-reinforced specimens <strong>and</strong><br />

is very low for 2D-r<strong>and</strong>omly specimens, after introduction <strong>of</strong> matrix cracking<br />

occurred. Once multiple cracking occurs, plane stress <strong>and</strong> plane strain behaviour<br />

almost coincide.<br />

As a conclusion, the plane strain <strong>and</strong> plane stress assumption in the z-direction<br />

should lead to rather similar results for the overall s<strong>and</strong>wich behaviour <strong>and</strong> for the<br />

determination <strong>of</strong> maximum stresses in the tensile face. However, plane strain is<br />

still a more logical option for the compressive face, when the behaviour <strong>of</strong> widebeams<br />

under loading is discussed.<br />

5.4 S<strong>and</strong>wich panels under monotonic loading: the use <strong>of</strong><br />

ANSYS to model s<strong>and</strong>wich panels<br />

5.4.1 introduction<br />

The finite element package, used in this work, is ANSYS, which is a widely used<br />

finite element package. The introduction <strong>of</strong> a constitutive equation, representing<br />

the stress-strain behaviour <strong>of</strong> the IPC composite face in tension under monotonic<br />

loading, is presented in this paragraph. Also the feasibility <strong>of</strong> several finite<br />

element models <strong>and</strong> their implicit hypotheses, accuracy <strong>and</strong> computing effort are<br />

discussed. At present, only the behaviour <strong>of</strong> s<strong>and</strong>wich panels under monotonic<br />

loading is considered.<br />

5.4.2 implementation <strong>of</strong> material behaviour<br />

The implementation <strong>of</strong> the IPC composite face behaviour in ANSYS is discussed<br />

in this paragraph. The introduction <strong>of</strong> the stress-strain behaviour <strong>of</strong> IPC under<br />

monotonic loading in ANSYS can be performed in several ways.<br />

5.4.2.1 implementation <strong>of</strong> material behaviour: ‘multilinear elastic’ option<br />

If it is clear a priori which parts <strong>of</strong> the faces are loaded in compression <strong>and</strong> which<br />

are loaded in tension, they can be defined as different materials. The regions in the<br />

faces, which are loaded in compression, are characterised as linear elastic<br />

materials. For those parts <strong>of</strong> the faces loaded in tension, material parameters like<br />

the fibre <strong>and</strong> matrix volume fraction <strong>and</strong> E-modulus, the Weibull properties <strong>of</strong> the<br />

matrix <strong>and</strong> the frictional matrix-fibre interface shear stress are introduced into the<br />

program stress-strainIPC.exe. This program has been presented in paragraph 3.5.<br />

The output <strong>of</strong> this program is a stress-strain curve, but also a table <strong>with</strong> a number<br />

<strong>of</strong> stress-strain points from the calculated stress-strain curve. The user can define<br />

146


Chapter 5: S<strong>and</strong>wich modelling<br />

the number <strong>of</strong> desired stress-strain points. This table can be read by or copied into<br />

the finite element program. If those regions <strong>of</strong> the faces, which are loaded in<br />

tension, are defined to be ‘multilinear elastic’ materials, a stress-strain table up to<br />

100 points can be inserted in ANSYS.<br />

Figure 5.2 shows a typical stress-strain curve, as introduced in the ‘multilinear<br />

elastic’ option. If this material option is used, the unloading stress-strain path<br />

equals the loading stress-strain path. ‘multilinear elastic’ can thus be used to<br />

calculate the deflections, stresses <strong>and</strong> strains <strong>of</strong> a s<strong>and</strong>wich panel <strong>with</strong> IPC<br />

composite faces under monotonic loading, but is useless to predict residual<br />

deflections if the panel is unloaded.<br />

σ<br />

Figure 5.2: stress-strain behaviour <strong>with</strong> option <strong>of</strong> ‘multilinear’ behaviour<br />

5.4.2.2 implementation <strong>of</strong> material behaviour: ‘aniso’ option<br />

If one is not sure a priori about the occurrence <strong>of</strong> compression or tension in a face<br />

at a certain location, the material behaviour ‘aniso’ can be used in ANSYS. The<br />

material option ‘aniso’ is used if the compression <strong>and</strong> tension behaviour can be<br />

different <strong>and</strong>/or if the material can have different properties in three perpendicular<br />

directions. The yield surface is a modified Von Mises criterion, <strong>with</strong> introduction<br />

<strong>of</strong> different material behaviour in three perpendicular directions <strong>and</strong> for tension<br />

<strong>and</strong> compression. A “best fit” bilinear curve in tension represents the stress-strain<br />

behaviour <strong>of</strong> IPC in tension.<br />

Figure 5.3 shows an example <strong>of</strong> bilinear stress-strain curves in the x- <strong>and</strong> ydirection,<br />

when the ‘aniso’ material behaviour is used. In the example <strong>of</strong> figure<br />

5.3, the material yield stresses, which are to be inserted, are σ y x-, σ y x+, σ y y-, σ y y+<br />

<strong>and</strong> σ y xy, where:<br />

σ y x- = compressive yield stress in x-direction<br />

σ y x+ = tensile yield stress in x-direction<br />

σ y y- = compressive yield stress in y-direction<br />

σ y y+ = tensile yield stress in y-direction<br />

σ y xy = shear yield stress in xy-plane<br />

147<br />

ε


Chapter 5: S<strong>and</strong>wich modelling<br />

σ<br />

y-direction<br />

ε<br />

x-direction<br />

Figure 5.3: example <strong>of</strong> bilinear stress-strain curves for ‘aniso’ material option<br />

These material yielding stresses, which are introduced in ANSYS by the user to<br />

define the face material behaviour, should satisfy certain requirements leading to a<br />

proper description <strong>of</strong> the behaviour <strong>of</strong> E-glass fibre reinforced IPC:<br />

1. Plasticity should be triggered by the normal stress in the xdirection<br />

<strong>and</strong> is preferably insensitive to shear stresses <strong>and</strong> normal stresses<br />

in any direction transverse to the x-direction. If a high value <strong>of</strong> the shear<br />

yield stresses is inserted, the yield criterion becomes insensitive to shear.<br />

2. If high values <strong>of</strong> yield stresses in the directions perpendicular to<br />

the x-direction are chosen, the yield criterion becomes also insensitive to<br />

yield in those directions.<br />

3. If the face is loaded in compression along the x-axis, the yield<br />

stress σx- should be reached only at failure. A typical value <strong>of</strong> σx- is thus<br />

90MPa (see Chapter 2).<br />

Once “yield” occurs due to multiple cracking, the stiffness in the x-direction<br />

should decrease considerably, whilst the stiffness properties are only slightly<br />

affected in the other directions (see paragraph 2.8 for shear). Within ANSYS a<br />

Young’s modulus E el is to be defined as the modulus before ‘yielding’ occurs. E T<br />

is the modulus, which should be defined by the user as the remaining stiffness<br />

after yielding occurs.<br />

The definition <strong>of</strong> the elastic E-modulus (E el ), the tangent E-modulus (E T ) <strong>and</strong> the<br />

plastic E-modulus (E pl ), as used in ANSYS, are illustrated in figure 5.4:<br />

148


∆σ<br />

σ<br />

Chapter 5: S<strong>and</strong>wich modelling<br />

∆ε el<br />

E el<br />

∆ε pl<br />

Figure 5.4: definition <strong>of</strong> elastic, plastic <strong>and</strong> tangent stiffness, bi-axial ‘aniso’ behaviour<br />

∆ε el = elastic strain increment<br />

∆ε pl = plastic strain increment<br />

From figure 5.4 it can be seen that:<br />

pl pl<br />

∆ σ = ∆ε<br />

E<br />

(5.1)<br />

el el<br />

∆ σ = ∆ε<br />

E<br />

( )<br />

(5.2)<br />

T el<br />

pl<br />

∆ σ = ∆ε<br />

+ ∆ε<br />

E<br />

Thus, E<br />

(5.3)<br />

pl can be rewritten in terms <strong>of</strong> E el <strong>and</strong> E T :<br />

T el<br />

pl E E<br />

E = el T<br />

E − E<br />

(5.4)<br />

It is the value <strong>of</strong> E T that can be inserted in ANSYS by the user, representing the<br />

stiffness after yielding occurs. Since yielding in the x-direction should have minor<br />

influence on the IPC composite behaviour in the y-direction <strong>and</strong> z-direction<br />

direction, the tangent moduli in the y-direction <strong>and</strong> z-directions (E T in y-direction<br />

<strong>and</strong> in z-direction) are chosen slightly lower than the initial elastic stiffness<br />

modulus E el . The tangent modulus in the x-direction is chosen considerably lower<br />

than E el . E pl is very high in the y-direction <strong>and</strong> in the z-direction (but not infinite),<br />

so that the plastic strains in the y-direction <strong>and</strong> z-direction are very low according<br />

to equation (5.1). This way, after the onset <strong>of</strong> yielding is triggered by the normal<br />

stresses in the x-direction, the total strains in the y-direction <strong>and</strong> z-direction are<br />

mainly defined by the elastic strain component, but there is a small contribution <strong>of</strong><br />

plastic strains in any direction different from the length axis.<br />

Similarly to the behaviour <strong>of</strong> the faces in the y-direction <strong>and</strong> z-direction, the shear<br />

stiffness <strong>of</strong> the faces is reduced slightly, once the yielding criterion is fulfilled (see<br />

experimental verification in paragraph 2.8).<br />

As a conclusion, the st<strong>and</strong>ard material behaviour options in ANSYS, which can be<br />

considered as useful to describe the behaviour <strong>of</strong> the IPC composite faces, are<br />

‘multilinear elastic’ <strong>and</strong> ‘aniso’ behaviour. If the ‘multilinear elastic’ behaviour is<br />

149<br />

E pl<br />

E T<br />

ε


Chapter 5: S<strong>and</strong>wich modelling<br />

chosen, the stress-strain curve <strong>of</strong> the IPC composite faces can be modelled as<br />

presented in Chapter 2 <strong>with</strong> rather high accuracy. The main drawback <strong>of</strong> this<br />

material option is that it has to be clear a priori, which regions <strong>of</strong> the faces are<br />

loaded in tension <strong>and</strong> which in compression. The material option ‘aniso’ can thus<br />

be used if there is no certainty about the occurrence <strong>of</strong> compression or tension in a<br />

s<strong>and</strong>wich face. The drawback <strong>of</strong> the ‘aniso’ material behaviour is that the real<br />

stress-strain behaviour <strong>of</strong> the IPC composite face in tension is replaced by a<br />

bilinear curve, which is slightly less accurate in the calculation <strong>of</strong> face strains <strong>and</strong><br />

stresses.<br />

5.4.3 type <strong>of</strong> model: discussion<br />

5.4.3.1 elementary s<strong>and</strong>wich theory (EST)<br />

Several s<strong>and</strong>wich theories use the assumption <strong>of</strong> a transverse incompressible core.<br />

This is a basic assumption in elementary s<strong>and</strong>wich theory (EST). Figure 5.5 shows<br />

the evolution <strong>of</strong> the normal <strong>and</strong> shear stress across the thickness <strong>of</strong> the s<strong>and</strong>wich,<br />

when an increasing number <strong>of</strong> assumptions is used.<br />

Figure 5.5: normal stress <strong>and</strong> shear stress evolution across the thickness <strong>of</strong> a s<strong>and</strong>wich <strong>with</strong><br />

increasing number <strong>of</strong> assumptions (from Allen, 1969 <strong>and</strong> Zenkert, 1995)<br />

If the elementary s<strong>and</strong>wich theory (EST) is used, several hypotheses are made:<br />

- The bending stiffness <strong>of</strong> the core is neglected. Normal stresses σx ≠ 0 in<br />

the faces only. σx = 0 in the core.<br />

- Since there are no normal stresses in the core, the shear stresses are<br />

constant across the thickness <strong>of</strong> the core.<br />

- The faces act as membranes. The constant shear stresses in the core<br />

contribute to a shear deformation. Since the faces act as membranes, they do not<br />

provide extra bending resistance to the shear deformation.<br />

- If the thickness <strong>of</strong> a face is small compared to the distance between the<br />

middle <strong>of</strong> this face <strong>and</strong> the neutral axis, the normal stresses in the faces are<br />

constant across the thickness.<br />

150


Chapter 5: S<strong>and</strong>wich modelling<br />

The bending stiffness <strong>of</strong> a s<strong>and</strong>wich beam is the sum <strong>of</strong> the stiffness contributions<br />

<strong>of</strong> all layers <strong>and</strong> is thus:<br />

b 3<br />

t 3<br />

( t f ) b(<br />

t f ) b(<br />

t )<br />

3<br />

b b<br />

t<br />

c<br />

2<br />

D = E f + E f + Eco<br />

+ Ecobtc()<br />

e +<br />

12 12 12<br />

2<br />

2<br />

(5.5)<br />

t<br />

b<br />

⎛ t + t ⎞ ⎛ t + t ⎞<br />

t t f c<br />

E bt ⎜ − e⎟<br />

b b f c<br />

+ E bt ⎜ + e⎟<br />

f f ⎜ 2 ⎟ f f ⎜ 2 ⎟<br />

⎝ ⎠ ⎝ ⎠<br />

tf b <strong>and</strong> tf t are the lower <strong>and</strong> upper face thickness respectively. b is the width <strong>of</strong> the<br />

s<strong>and</strong>wich panel. tc is the thickness <strong>of</strong> the core <strong>and</strong> e is the distance between the<br />

centroid <strong>of</strong> the core <strong>and</strong> the neutral axis. Ef b , Ef t <strong>and</strong> Eco are the lower <strong>and</strong> upper<br />

face E-moduli <strong>and</strong> the E-modulus <strong>of</strong> the core.<br />

In general, a term in bending stiffness equation (5.5) is neglected if its contribution<br />

is less than 1%. The core does not contribute to the total bending stiffness if the<br />

sum <strong>of</strong> the third <strong>and</strong> fourth term in equation (5.5) contributes less than 1% to the<br />

total stiffness. The bending stiffness is then:<br />

b 3<br />

t<br />

( t ) b(<br />

t )<br />

3<br />

12 12<br />

2<br />

2 ⎟ b b f<br />

D = E f<br />

t<br />

+ E f<br />

f<br />

t ⎛ t + t ⎞<br />

t t f c<br />

+ E ⎜ − e⎟<br />

f bt f ⎜ ⎟<br />

⎝ ⎠<br />

b ⎛ t + t ⎞<br />

b b f c<br />

+ E bt ⎜<br />

f f + e<br />

⎜<br />

⎝ ⎠<br />

(5.6)<br />

The weak core assumption is thus valid if:<br />

( t )<br />

2<br />

b 3<br />

t<br />

( t ) b(<br />

t )<br />

3<br />

b<br />

2 b<br />

c<br />

b f t f<br />

Eco<br />

+ Ecobtc()<br />

e < A(<br />

E f + E f +<br />

12<br />

12 12<br />

2<br />

2<br />

(5.7)<br />

t<br />

b<br />

⎛ t + t ⎞ ⎛ t + t ⎞<br />

t t f c<br />

E bt ⎜ − e⎟<br />

b b f c<br />

+ E bt ⎜ + e⎟<br />

f f<br />

)<br />

⎜ 2 ⎟ f f ⎜ 2 ⎟<br />

⎝ ⎠ ⎝ ⎠<br />

A is a constant, concerning the contribution <strong>of</strong> a term to the total bending stiffness.<br />

As mentioned before, A is usually chosen 0.01 (1%).<br />

In case the faces are thin, the bending stiffness <strong>of</strong> the faces along their own axis is<br />

neglected, compared to the contribution <strong>of</strong> the other terms. Thus:<br />

b 3 t<br />

( t ) b(<br />

t )<br />

b b f<br />

D = E f<br />

12<br />

t<br />

+ E f<br />

f<br />

12<br />

(5.8)<br />

The approximation <strong>of</strong> thin faces can be used in case:<br />

b 3<br />

t<br />

( t ) b(<br />

t )<br />

3<br />

t<br />

b<br />

b<br />

⎛ t + t ⎞ ⎛ t + t ⎞<br />

b f t f<br />

t t f c<br />

( ⎜ ⎟ b b f c<br />

E + E < A E bt − e + E bt ⎜ + e⎟<br />

f f<br />

f f<br />

)<br />

12 12<br />

⎜ 2 ⎟ f f<br />

(5.9)<br />

⎜ 2 ⎟<br />

⎝ ⎠ ⎝ ⎠<br />

In case both faces have the same stiffness <strong>and</strong> thickness, e equals zero <strong>and</strong><br />

equation (5.5) becomes:<br />

( ) ( ) ( ) 2<br />

t 3<br />

3<br />

t<br />

t b t<br />

t + t<br />

2<br />

3<br />

b t f<br />

c t t f c<br />

D = E f + Eco<br />

+ E f bt<br />

(5.10)<br />

f<br />

6 12<br />

2<br />

151<br />

3<br />

2<br />

2


Chapter 5: S<strong>and</strong>wich modelling<br />

The weak core assumption is then valid if:<br />

3<br />

t 3<br />

t 2<br />

( t ) b(<br />

t f ) ( t f + tc<br />

) c<br />

t<br />

t t<br />

(<br />

)<br />

b<br />

Eco<br />

12<br />

< A E f<br />

6<br />

+ E f bt f<br />

2<br />

(5.11)<br />

In case (5.11) is fulfilled, (5.10) becomes:<br />

( ) ( ) 2<br />

3 t<br />

t<br />

t<br />

t + t<br />

b t<br />

D = E f<br />

f<br />

6<br />

t t<br />

+ E f bt f<br />

f<br />

2<br />

c<br />

(5.12)<br />

The assumption <strong>of</strong> thin faces is valid when condition (5.13) is fulfilled.<br />

b t f<br />

E f<br />

6<br />

The bending stiffness is then:<br />

( ) ( ) 2<br />

3 t<br />

t<br />

t<br />

t + t<br />

t t f c<br />

< AE f bt<br />

(5.13)<br />

f<br />

2<br />

( ) 2 t<br />

t + t<br />

t t f c<br />

D = E f bt<br />

(5.14)<br />

f<br />

2<br />

If s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces are calculated, the non-linear<br />

material behaviour <strong>of</strong> the faces might make it more delicate to decide whether the<br />

thin face <strong>and</strong> weak core approximations are valid. If a s<strong>and</strong>wich panel is made<br />

<strong>with</strong> initially equal bottom <strong>and</strong> top face, the weak core <strong>and</strong> thin face assumption<br />

can be checked <strong>with</strong> equation (5.11) <strong>and</strong> (5.13). When the external load on the<br />

s<strong>and</strong>wich panel increases, the face in tension will develop matrix cracking. The<br />

stiffness <strong>of</strong> this face decreases <strong>and</strong> consequently equation (5.7) <strong>and</strong> (5.9) should be<br />

used to check the thin face <strong>and</strong> weak core assumption, instead <strong>of</strong> equations (5.11)<br />

<strong>and</strong> (5.13).<br />

If the faces are dissimilar, two extra inequalities can be checked. If equation (5.9)<br />

is not fulfilled, it may be due to high contribution <strong>of</strong> local bending <strong>of</strong> the lower<br />

face or the upper face only. Each face can be checked separately to fulfil the thin<br />

face approximation. The upper face is checked by condition (5.15):<br />

t ( t )<br />

3<br />

b t f<br />

E f<br />

12<br />

t<br />

A ⎛ t + t ⎞<br />

t t f c<br />

< ( E bt ⎜ − e⎟<br />

f f<br />

)<br />

2 ⎜ 2 ⎟<br />

⎝ ⎠<br />

(5.15)<br />

<strong>and</strong> the lower face by condition (5.16):<br />

b ( t )<br />

3<br />

(<br />

12 2 2 ⎟ b<br />

b<br />

⎛ t + t ⎞<br />

b f A b b f c<br />

E < E bt ⎜<br />

f f f + e ) (5.16)<br />

⎜<br />

⎝ ⎠<br />

Once multiple cracking occurs in a IPC composite face, the neutral axis will shift<br />

towards the non-cracked face. The slope <strong>of</strong> the normal stress evolution across the<br />

face thickness increases in the non-cracked face <strong>and</strong> decreases in the face showing<br />

multiple cracking. This effect is illustrated in figure 5.6a <strong>and</strong> 5.6b<br />

152<br />

2<br />

2


Chapter 5: S<strong>and</strong>wich modelling<br />

σx<br />

middle axis<br />

neutral axis<br />

Figure 5.6a: evolution <strong>of</strong> normal stresses<br />

across thickness <strong>of</strong> s<strong>and</strong>wich panel <strong>with</strong> IPC<br />

faces, initial condition<br />

e<br />

σx<br />

Figure 5.6b: evolution <strong>of</strong> normal stresses<br />

across thickness <strong>of</strong> s<strong>and</strong>wich panel <strong>and</strong> shift<br />

<strong>of</strong> neutral axis<br />

after initiation <strong>of</strong> multiple cracking<br />

In ANSYS, SHELL 91 is a finite element, working <strong>with</strong> the assumptions <strong>of</strong> thin<br />

faces <strong>and</strong> weak core <strong>and</strong> should therefore lead to accurate prediction <strong>of</strong> the<br />

deflections, provided equations (5.7) <strong>and</strong> (5.9) are fulfilled.<br />

5.4.3.2 advanced s<strong>and</strong>wich theory<br />

There may be reasons to reject the hypothesis <strong>of</strong> the transverse incompressibility<br />

<strong>of</strong> the core, the thin face assumption <strong>and</strong>/or the weak core assumption, which are<br />

used in the EST model. Frostig (1992) presented a higher-order s<strong>and</strong>wich<br />

formulation for flat s<strong>and</strong>wich beams, still <strong>of</strong>fering the possibility <strong>of</strong> finding a<br />

solution analytically for a s<strong>and</strong>wich panel <strong>with</strong> transversely compressible core.<br />

The core is modelled as a special orthotropic elastic material. The in-plane<br />

Young’s modulus is zero. The in-plane normal stresses in the core are thus zero.<br />

The core material only possesses an out-<strong>of</strong>-plane E-modulus (thickness <strong>of</strong> the<br />

core). Normal stresses are present in the core in the transverse direction (ydirection,<br />

thickness <strong>of</strong> the core). Shear stresses are constant across the thickness <strong>of</strong><br />

the core. The faces are considered to act as elastic beams on an elastic foundation<br />

(the core). Figure 5.7 illustrates the forces <strong>and</strong> stresses, acting on the faces <strong>and</strong> the<br />

core <strong>of</strong> a flat s<strong>and</strong>wich panel <strong>and</strong> shows the displacement field. The responses <strong>of</strong><br />

the core <strong>and</strong> the faces are coupled by requiring continuity <strong>of</strong> the displacements<br />

along the core-face interface.<br />

For straight panels, an analytical solution for the differential equations (obtained<br />

from expression <strong>of</strong> equilibrium <strong>of</strong> faces <strong>and</strong> core) is derived <strong>and</strong> discussed for<br />

many load cases <strong>and</strong> specific boundary conditions by Frostig (1992) <strong>and</strong> Swanson<br />

(1999). The influence <strong>of</strong> local loading <strong>and</strong> <strong>of</strong> the supports is discussed in these<br />

publications. Frostig (1999) derived differential equations for curved s<strong>and</strong>wich<br />

panels. Frostig (1999) <strong>and</strong> Thomsen <strong>and</strong> Vinson (2000) discuss the use <strong>of</strong> this<br />

higher order theory for s<strong>and</strong>wich panels <strong>with</strong> high curvature. Although this<br />

advanced s<strong>and</strong>wich theory can reproduce effects that are not detected by the EST,<br />

153


Chapter 5: S<strong>and</strong>wich modelling<br />

it has been subject <strong>of</strong> study on s<strong>and</strong>wich panels <strong>with</strong> linear elastic materials only at<br />

present. However, the background may be useful in choosing a finite element<br />

modelling method <strong>with</strong> limited number <strong>of</strong> degrees <strong>of</strong> freedom, but high enough<br />

complexity to predict features that may be hidden in more simplified s<strong>and</strong>wich<br />

models. These features are high core-face interface shear stresses, high core-face<br />

interface transverse normal stresses <strong>and</strong> local face bending effects.<br />

Nt<br />

Nb<br />

Mt<br />

Mb<br />

Vt<br />

Vb<br />

qt<br />

qb<br />

b.τ<br />

b.σy(y=0)<br />

b.σy(y=tc)<br />

b.τ<br />

uc<br />

(wb , ut)<br />

(wt , ut)<br />

Figure 5.7: forces <strong>and</strong> stresses acting on faces <strong>and</strong> core <strong>and</strong> displacement field (after Frostig,<br />

1992)<br />

Nb = normal force in lower face Nt = normal force in upper face<br />

Mb = moment in lower face Mt = moment in upper face<br />

Vb = shear force in lower face Vt = shear force in upper face<br />

wb = displacement <strong>of</strong> lower face (y-direction) wt = displacement <strong>of</strong> upper face (y-direction)<br />

ub = displacement <strong>of</strong> lower face (x-direction) ut = displacement <strong>of</strong> upper face (x-direction)<br />

uc = displacement <strong>of</strong> core (x-direction)<br />

No st<strong>and</strong>ard element is available <strong>with</strong>in ANSYS, which includes all assumptions<br />

<strong>and</strong> advantages described above. However, D.J. O’Connor (1987) presented an<br />

approach, which is almost similar <strong>and</strong> has the possibility <strong>of</strong> easy implementation<br />

in ANSYS. The author suggests the core should be represented by one element<br />

over the thickness. This element is a 2D structural solid element <strong>with</strong> eight nodes<br />

<strong>and</strong> two degrees <strong>of</strong> freedom per node: a translation in the x-direction (ux) <strong>and</strong> in<br />

the y-direction (uy). The faces are represented by one element over the thickness.<br />

This element is a beam element <strong>with</strong> two nodes per element <strong>and</strong> three degrees <strong>of</strong><br />

freedom per node: ux, uy <strong>and</strong> a rotational degree <strong>of</strong> freedom (rotz). Figure 5.8a<br />

illustrates the element representation <strong>of</strong> a small section <strong>of</strong> a s<strong>and</strong>wich beam.<br />

Figure 5.8b shows the use <strong>of</strong> this element representation for a whole panel. Figure<br />

154


Chapter 5: S<strong>and</strong>wich modelling<br />

5.8c illustrates that element refinement is possible at certain locations, for example<br />

in the vicinity <strong>of</strong> a support. This refinement can be done along the length <strong>and</strong><br />

across the thickness <strong>of</strong> the panel.<br />

beam<br />

element<br />

beam<br />

element<br />

beam<br />

element<br />

2D structural element<br />

beam<br />

element<br />

Figure 5.8a: combination <strong>of</strong> 2D<br />

structural <strong>and</strong> beam elements for<br />

a s<strong>and</strong>wich section<br />

Symmetry conditions<br />

Figure 5.8b: combined 2D structural <strong>and</strong> beam element<br />

modelling for s<strong>and</strong>wich panels<br />

Symmetry conditions<br />

Figure 5.8c: refinement <strong>of</strong> combined model<br />

The main objection that can be made against this type <strong>of</strong> modelling is the fact that<br />

the number <strong>of</strong> degrees <strong>of</strong> freedom <strong>of</strong> the coupled elements is not equal. D. J.<br />

O’Connor (1987) announces in his work that this disadvantage does not prevent<br />

this model from reproducing quite accurate results in combination <strong>with</strong> relatively<br />

low computing effort. The reason this model provides accurate results can be<br />

found in the low bending stiffness <strong>of</strong> the core. If the E-modulus <strong>of</strong> the core is<br />

rather low, bending <strong>of</strong> the faces is hardly transferred to the core. Therefore, if rotz<br />

(rotation around the z-axis) is transferred from a beam element to the neighbour<br />

beam elements <strong>and</strong> not to the core, this method provides accurate results, provided<br />

the bending stiffness <strong>of</strong> the core is indeed very low.<br />

Within ANSYS, PLANE 82 elements <strong>with</strong> plane stress or plain strain option can<br />

be used to represent the core elements. BEAM 188 is used to represent the face<br />

elements. All types <strong>of</strong> non-linear behaviour are supported by this element. Node<br />

<strong>of</strong>fset is also an option if this type <strong>of</strong> elements is used. Moreover, in post<br />

processing the beam stresses are directly plotted in the face beam elements.<br />

5.4.4 parameters <strong>of</strong> study<br />

The aim <strong>of</strong> this paragraph is the comparison <strong>of</strong> different finite elements available<br />

in ANSYS for the computation <strong>of</strong> s<strong>and</strong>wich panels. The models, which are<br />

compared to each other, are:<br />

155


Chapter 5: S<strong>and</strong>wich modelling<br />

a) The reference model, which uses 2D quadrilateral element <strong>with</strong> 8-nodes. The<br />

number <strong>of</strong> divisions along the length is 250. The number <strong>of</strong> divisions across<br />

the thickness is 10 for each face <strong>and</strong> 20 for the core.<br />

b) One 2D quadrilateral element (PLANE 82) is used to represent the upper face<br />

across the thickness. One PLANE 82 element represents the core <strong>and</strong> one<br />

represents the lower face along the thickness. The number <strong>of</strong> divisions along<br />

the length is varied.<br />

c) A s<strong>and</strong>wich section is represented by one multi-layer element, <strong>with</strong> the<br />

s<strong>and</strong>wich option activated (SHELL 91, keyopt (9)= 1)<br />

d) The upper face <strong>and</strong> lower face are both represented by a beam element (BEAM<br />

188). The core is represented by one 2D quadrilateral element across the<br />

thickness <strong>of</strong> the s<strong>and</strong>wich (PLANE 82). Although the number <strong>of</strong> degrees <strong>of</strong><br />

freedom is not equivalent for nodes <strong>of</strong> the the BEAM 188 <strong>and</strong> PLANE 82<br />

elements, it has been mentioned by D.J. O’Connor (1987) that this type <strong>of</strong><br />

modelling leads to rather good prediction <strong>of</strong> deflections, stresses <strong>and</strong> strains.<br />

This model is referred to as “COMBI” in this work<br />

Manet (1998) performed a similar study on s<strong>and</strong>wich panels <strong>with</strong> aluminium<br />

faces, which are thinner than the studied IPC composite faces <strong>and</strong> behave linear<br />

elastically.<br />

5.4.5 a reference s<strong>and</strong>wich beam<br />

A reference beam, subjected to a uniform pressure, is analysed. The geometry <strong>of</strong><br />

this reference s<strong>and</strong>wich beam is chosen to represent a typical practical s<strong>and</strong>wich<br />

panel, as will be illustrated in Chapter 7. The total length <strong>of</strong> the studied reference<br />

beam is 2500mm, the core thickness is 60mm <strong>and</strong> the face thickness is 3.75mm.<br />

The width is 1m. The beam is simply supported <strong>and</strong> is subjected to a uniform<br />

pressure <strong>of</strong> 2.5kN/m².<br />

The reference model uses 2D quadrilateral element <strong>with</strong> 8-nodes (PLANE 82).<br />

The number <strong>of</strong> divisions along the length is 250. The number <strong>of</strong> divisions across<br />

the thickness is 10 for each face <strong>and</strong> 20 for the core.<br />

The core material is a PUR foam <strong>with</strong> a stiffness Eco = 8MPa <strong>and</strong> a density <strong>of</strong><br />

40kg/m³. The faces contain 2D-r<strong>and</strong>omly oriented reinforcement. The IPC matrix<br />

stiffness Em is 18GPa, the fibre volume fraction Vf is 10% <strong>and</strong> the fibre radius r =<br />

7µm. The final crack spacing is 1mm. The reference cracking stress σR is 14MPa.<br />

τ0 is chosen to be 1.0MPa. The Weibull modulus m is 3.0. The choice <strong>of</strong> these<br />

values <strong>of</strong> the reference cracking stress, matrix-fibre interface shear stress <strong>and</strong><br />

Weibull modulus are justified by the experiments in Chapter 2, Chapter 3 <strong>and</strong><br />

Chapter 4. The density <strong>of</strong> the face material is about 2000kg/m³. Figure 5.9 shows<br />

the theoretical stress-strain curve under monotonic tensile loading.<br />

156


stress (MPa)<br />

15<br />

10<br />

5<br />

0<br />

Chapter 5: S<strong>and</strong>wich modelling<br />

0.0 0.1 0.2 0.3 0.4 0.5<br />

strain (%)<br />

Figure 5.9: theoretical stress-strain curve 2D-r<strong>and</strong>omly reinforced IPC under monotonic loading<br />

The moment is maximal at mid-span <strong>of</strong> the panel. The normal stresses in the faces<br />

are thus maximal at this location. The maximum deflection is also determined at<br />

mid-span.<br />

The transverse force is maximal close to the supports. However, close to the<br />

support local stress transfer mechanisms are expected. If one is interested in the<br />

‘s<strong>and</strong>wich action’ it would be more interesting to determine shear stresses in a<br />

section far enough from the local support. This section should also be located far<br />

enough from mid-span, since transverse force is minimal at mid-span. Therefore,<br />

shear stresses are retrieved at 1/4 th <strong>of</strong> the total span. A point A is chosen in this<br />

section, at the core-face interface (see figure 5.10). Figure 5.10 illustrates how the<br />

reference s<strong>and</strong>wich panel is modelled <strong>and</strong> also indicates the placement <strong>of</strong> point A,<br />

where core <strong>and</strong> face stresses are retrieved.<br />

2.5kN/m²<br />

A<br />

Figure 5.10: reference s<strong>and</strong>wich model<br />

The parameters <strong>of</strong> interest, which are retrieved from the result files are thus:<br />

1. the normal stress at the bottom <strong>of</strong> the lower face at mid-span<br />

2. the normal stress at the top <strong>of</strong> the upper face at mid-span<br />

3. the normal stress at the bottom <strong>of</strong> the core at point A (see figure 5.10)<br />

4. the normal stress at the top <strong>of</strong> the lower face at point A<br />

5. the shear stress in the core at point A<br />

6. the maximum defelction (at mid-span)<br />

Since only the behaviour <strong>of</strong> the panel under monotonic loading is considered now,<br />

the ‘multilinear elastic’ material behaviour is used in ANSYS to represent the<br />

157


Chapter 5: S<strong>and</strong>wich modelling<br />

behaviour <strong>of</strong> the lower face. 25 stress-strain points are extracted from the curve in<br />

figure 5.9 <strong>and</strong> inserted in the material stress-strain material data table <strong>of</strong> ANSYS.<br />

The upper face is defined to be a linear elastic material. The law <strong>of</strong> mixtures<br />

formulates the E-modulus <strong>of</strong> the composite in the upper face.<br />

5.4.6 comparison <strong>of</strong> models: results <strong>and</strong> discussion<br />

A s<strong>and</strong>wich <strong>with</strong> 2500mm span, 60mm core thickness <strong>and</strong> 3.75mm face thickness<br />

(3layers/face) is studied here. A pressure load <strong>of</strong> 2.5kN/m² is applied to the simply<br />

supported s<strong>and</strong>wich. For all finite element models, the number <strong>of</strong> elements along<br />

the length <strong>of</strong> the s<strong>and</strong>wich (ndiv) is increased.<br />

maximumσx<br />

(MPa)<br />

face point A (MPa)<br />

x<br />

σ<br />

8.35<br />

8.25<br />

8.15<br />

8.05<br />

7.95<br />

reference<br />

plane 82<br />

shell 91<br />

combi<br />

0 100 200 300<br />

ndiv (-)<br />

Figure 5.11a: maximum normal tensile stress in<br />

the lower face at mid-span<br />

6.20<br />

6.10<br />

6.00<br />

5.90<br />

5.80<br />

5.70<br />

reference<br />

plane 82<br />

shell 91<br />

combi<br />

0 100 200 300<br />

ndiv (-)<br />

Figure 5.11c: normal tensile stress in the lower<br />

face at point A<br />

τxy core (MPa)<br />

2.65E-02<br />

2.60E-02<br />

2.55E-02<br />

2.50E-02<br />

2.45E-02<br />

2.40E-02<br />

reference<br />

plane 82<br />

shell 91<br />

combi<br />

0 100 200 300<br />

ndiv (-)<br />

|minimum σx| (MPa)<br />

9.5<br />

9.0<br />

8.5<br />

8.0<br />

7.5<br />

7.0<br />

reference<br />

plane 82<br />

shell 91<br />

combi<br />

0 50 100 150 200 250 300<br />

ndiv (-)<br />

Figure 5.11b: maximum normal compressive stress<br />

in the upper face at mid-span<br />

|Uz| (mm)<br />

σx core point A (MPa)<br />

4.E-03<br />

3.E-03<br />

2.E-03<br />

1.E-03<br />

0.E+00<br />

reference<br />

plane 82<br />

shell 91<br />

combi<br />

0 100 200 300<br />

ndiv (-)<br />

Figure 5.11d: normal tensile stress in core at<br />

point A<br />

22.8<br />

22.6<br />

22.4<br />

22.2<br />

22.0<br />

21.8<br />

reference<br />

plane 82<br />

shell 91<br />

combi<br />

0 100 200 300<br />

ndiv (-)<br />

Figure 5.11e: shear stress in the core at point A Figure 5.11f: maximum defelction<br />

158


Chapter 5: S<strong>and</strong>wich modelling<br />

If the external load is 2.5kN/m², figure 5.11a represents the evolution <strong>of</strong> the<br />

maximum normal tensile stress in the lower face <strong>with</strong> increasing (ndiv). Figure<br />

5.11b shows the maximum normal compressive stress in the upper face as function<br />

<strong>of</strong> (ndiv). These stresses are retrieved at mid-span <strong>of</strong> the s<strong>and</strong>wich panel.<br />

The normal tensile core <strong>and</strong> face stresses in point A are printed in figure 5.11c <strong>and</strong><br />

5.11d. The shear stress in the core in point A is shown in figure 5.11e <strong>and</strong> the<br />

maximum deflection in figure 5.11f.<br />

From figure 5.11a to 5.11f, following conclusions can be formulated:<br />

1. In general, the PLANE 82 model gives better predictions <strong>of</strong> stresses than<br />

the other finite element models, <strong>with</strong> exception <strong>of</strong> the normal stresses in the face at<br />

point A. Even when a rather course mesh is used across the thickness <strong>of</strong> the face -<br />

one element for each layer - the accuracy <strong>of</strong> the solution compared to the reference<br />

solution is rather high.<br />

2. If SHELL 91 elements are used, only about 10 elements are to be used<br />

along the length <strong>of</strong> half the s<strong>and</strong>wich panel to obtain convergence. Unfortunately<br />

the converged solution <strong>of</strong> this model usually gives the least accurate results <strong>of</strong> all<br />

tested models. Possible explanations are discussed later in this paragraph.<br />

3. The combined PLANE 82/BEAM 188 model gives more satisfactory<br />

results for the prediction <strong>of</strong> resulting parameters (stresses <strong>and</strong> maximum<br />

deflections) than the SHELL 91 model. The normal stress in the core at point A is<br />

predicted more accurately <strong>with</strong> this model than <strong>with</strong> the PLANE 82 model.<br />

As can be seen from figure 5.11b, the SHELL 91 finite element provides a rather<br />

inaccurate prediction <strong>of</strong> the maximum compressive stress in the s<strong>and</strong>wich face.<br />

The SHELL 91 finite element is based on the assumptions made for EST. If the<br />

SHELL 91 element is used, one assumes that stresses are constant across the face<br />

thickness. From figure 5.11a, it can be seen that the discrepancy on the prediction<br />

<strong>of</strong> the maximum tensile stress between the reference model <strong>and</strong> the SHELL 91<br />

model is about 5%. However, figure 5.11b shows that this discrepancy is 25% for<br />

the prediction <strong>of</strong> the maximum stress in the face in compression. When a s<strong>and</strong>wich<br />

panel <strong>with</strong> same geometry <strong>and</strong> external load is calculated assuming linear elastic<br />

behaviour <strong>of</strong> the faces occurs, the discrepancy between the SHELL 91 finite<br />

element model <strong>and</strong> the reference model can also be determined. When (ndiv) is<br />

250 <strong>and</strong> linear elastic face behaviour is assumed, the discrepancy between both<br />

models is about 10% for the upper <strong>and</strong> lower face.<br />

The non-linear behaviour <strong>of</strong> the IPC faces in tension makes the EST based model<br />

(<strong>with</strong> SHELL 91 finite elements) less appropriate to predict the maximum<br />

compressive stress in the faces. The reason for this high discrepancy <strong>of</strong> the<br />

maximum compressive stress has been illustrated in figure 5.6a <strong>and</strong> 5.6b. The<br />

neutral axis will shift towards the compressive face, due to introduction <strong>of</strong><br />

159


Chapter 5: S<strong>and</strong>wich modelling<br />

multiple cracking in the tensile face. The relative variation <strong>of</strong> the normal stresses<br />

across the face thickness decreases in the lower face, but increases in the upper<br />

face.<br />

A final comment can be made on the PLANE 82/BEAM 188 combination. If a<br />

s<strong>and</strong>wich panel is designed, the problem is <strong>of</strong>ten split into a global <strong>and</strong> local<br />

s<strong>and</strong>wich response. A global s<strong>and</strong>wich theory is used to predict the global<br />

s<strong>and</strong>wich deformations, while the local stresses in the vicinity <strong>of</strong> a support are<br />

calculated later, using the Wrinkler or two parametric foundation model<br />

(Thomsen, 1994; Thomsen, 1995; Zenkert, 1995). The face in contact <strong>with</strong> the<br />

support or other local loading is considered to act as a beam on an elastic<br />

foundation. If s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces are calculated, the main<br />

objection to this local-global decoupling is the fact that the superposition principle<br />

is only valid if the faces behave linear elastically, which is clearly not the case.<br />

The foundation models are useful to give an idea <strong>of</strong> the wavelength <strong>of</strong> the local<br />

loading or the support: how far is the influence <strong>of</strong> the local loading experienced<br />

before local stresses are redistributed. The PLANE 82/BEAM 188 model is one <strong>of</strong><br />

the more suitable models to predict local bending <strong>of</strong> a face in the vicinity <strong>of</strong> a<br />

support, even <strong>with</strong> introduction <strong>of</strong> multiple cracking in this face. The SHELL 91<br />

model is not capable to h<strong>and</strong>le this problem at all. The PLANE 82 model <strong>with</strong> one<br />

element for each layer can give some idea about face stresses, but is not really<br />

designed to deal <strong>with</strong> very localised bending <strong>of</strong> a face, as is the case close to a<br />

support.<br />

5.4.7 influence <strong>of</strong> span<br />

In this paragraph the influence <strong>of</strong> the span on the distribution <strong>of</strong> normal stresses in<br />

the faces is discussed. Starting from the reference solution, the span <strong>of</strong> the<br />

s<strong>and</strong>wich panel is now varied. The pressure load is varied <strong>with</strong> the span, to obtain<br />

an identical bending moment at mid-section for all s<strong>and</strong>wich panels. This<br />

condition is satisfied when:<br />

2<br />

1<br />

2 1 ⎟<br />

2<br />

⎟<br />

⎛ l ⎞<br />

= p<br />

⎜<br />

l<br />

p (5.17)<br />

⎝ ⎠<br />

p1 <strong>and</strong> l1 are the pressure load on <strong>and</strong> the length <strong>of</strong> the reference panel<br />

respectively, p2 is the pressure load that should be applied on a s<strong>and</strong>wich <strong>of</strong> length<br />

l2 to get the same value <strong>of</strong> the bending moment at mid-section. Table 5.1 shows<br />

how the value <strong>of</strong> the applied pressure load varies <strong>with</strong> the span <strong>of</strong> the discussed<br />

s<strong>and</strong>wich panels.<br />

Table 5.1: pressure applied on panels <strong>with</strong> different span<br />

length (mm) 1000 1500 2000 2500 3000 3500 4000<br />

pressure (kN/m²) 15.6 6.95 3.90 2.50 1.74 1.28 0.977<br />

160


Chapter 5: S<strong>and</strong>wich modelling<br />

The reference finite element model <strong>with</strong> (ndiv) = 250, (ncore) = 20 <strong>and</strong> (nface) =<br />

10 is calculated, but this time the span is varied. The absolute value <strong>of</strong> the<br />

maximum tensile <strong>and</strong> compressive stress is listed for all spans in table 5.2. The<br />

relative variation <strong>of</strong> the normal stress across one face (∆σacross) is defined by<br />

equation (5.18). The values <strong>of</strong> (∆σacross) calculated for the upper <strong>and</strong> lower faces<br />

are listed in table 5.3.<br />

σ ( at top <strong>of</strong> face)<br />

−σ<br />

( at bottom <strong>of</strong> face)<br />

∆σ across =<br />

* 100%<br />

(5.18)<br />

maximum stress in face<br />

Table 5.2: maximum compressive stress in upper face <strong>and</strong> maximum tensile stress in lower face<br />

as function <strong>of</strong> the span<br />

length (mm) 1000 1500 2000 2500 3000 3500 4000<br />

max compressive stress (MPa) -11 -10 -9.5 -9.3 -9.2 -9.1 -8.9<br />

max tensile stress (MPa) 8.6 8.4 8.3 8.3 8.3 8.3 8.3<br />

Table 5.3: relative variation <strong>of</strong> normal stresses across upper face <strong>and</strong> lower face (∆σacross)<br />

length (mm) 1000 1500 2000 2500 3000 3500 4000<br />

lower face variation (%) 9 5 3 3 3 2 2<br />

upper face variation (%) 65 36 27 22 20 18 18<br />

Figure 5.12 shows the normal stress field in a s<strong>and</strong>wich panel at mid-span when<br />

the span is 1m (black curve) <strong>and</strong> when the span is 4m (grey curve). It can be seen<br />

from figure 5.12 <strong>and</strong> table 5.3 that the relative variation <strong>of</strong> the normal stress in the<br />

upper face is much higher, if the length <strong>of</strong> the panel decreases.<br />

span 1m<br />

span 4m<br />

-14 -12 -10 -8 -6 -4 -2-1 -3<br />

0 2 4 6 8 10<br />

7<br />

5<br />

3<br />

1<br />

-5<br />

-7<br />

stress (MPa)<br />

Figure 5.12: stress evolution across the thickness <strong>of</strong> the s<strong>and</strong>wich: span length 1m <strong>and</strong> 4m<br />

From table 5.3, it can be seen that the relative variation <strong>of</strong> the normal stress in the<br />

upper face varies from 18% for a span <strong>of</strong> 4m to 65% for a span <strong>of</strong> 1m. It has been<br />

mentioned in paragraph 5.4.3 that the applicability <strong>of</strong> the elementary s<strong>and</strong>wich<br />

theory is function <strong>of</strong> the relative material stiffness <strong>and</strong> thickness properties <strong>of</strong><br />

faces <strong>and</strong> core. Here it is shown that the span is also <strong>of</strong> importance. The SHELL<br />

161


Chapter 5: S<strong>and</strong>wich modelling<br />

91 element will not meet the requirements <strong>of</strong> accurate stress, strain <strong>and</strong> deflections<br />

predictions if the span is relatively short.<br />

In paragraph 5.4.3.1 the appropriateness <strong>of</strong> EST has been formulated as a function<br />

<strong>of</strong> the relative stiffness <strong>and</strong> thickness <strong>of</strong> faces <strong>and</strong> core in a section <strong>of</strong> the<br />

s<strong>and</strong>wich, as presented amongst others by Allen (1969) <strong>and</strong> Zenkert (1995). This<br />

classification helps the designer to decide <strong>with</strong>out too much trouble, which<br />

modelling simplifications can be used <strong>and</strong> is therefore widely used as assistance in<br />

choosing the modelling complexity. However, previous calculations show that this<br />

classification lacks an important parameter: the span. Just like the Bernouilli or<br />

Timoshenko approach for homogeneous beams, the relative importance <strong>of</strong> the<br />

bending <strong>and</strong> shear effects in the determination <strong>of</strong> the maximum deflection is<br />

function <strong>of</strong> the length to section ratio <strong>of</strong> the tested panel. The classification,<br />

presented in paragraph 5.4.3.1 should be h<strong>and</strong>led <strong>with</strong> care for short s<strong>and</strong>wich<br />

beams. The classification <strong>of</strong> the behaviour <strong>of</strong> s<strong>and</strong>wich beams as proposed in<br />

equation (5.5) to (5.16) can be used, provided the panel is relatively long-span.<br />

Figure 5.13a shows the normal stress field in a section <strong>of</strong> a s<strong>and</strong>wich panel, close<br />

to mid-span. It is assumed for now that the faces meet the requirements <strong>of</strong><br />

equations (5.9), (5.15) <strong>and</strong> (5.16). They are therefore relatively thin. At mid-span<br />

the normal stresses are rather constant <strong>and</strong> can be predicted well <strong>with</strong> EST. Figure<br />

5.13b shows the normal stress field across the section <strong>of</strong> the s<strong>and</strong>wich, if the span<br />

is decreased, but the maximum bending at mid-span is kept constant.<br />

bending<br />

+<br />

shear total<br />

Figure 5.13a: normal stresses in s<strong>and</strong>wich panel <strong>with</strong> relatively large span L;<br />

effects from global bending >>>> effects from shear deformation<br />

bending<br />

+<br />

shear total<br />

Figure 5.13b: normal stresses in s<strong>and</strong>wich panel <strong>with</strong> medium span L;<br />

effects from global bending > effects from shear deformation<br />

162<br />

=<br />

=


Chapter 5: S<strong>and</strong>wich modelling<br />

In figure 5.13a <strong>and</strong> figure 5.13b it is illustrated that the total normal stress in the<br />

faces is the sum <strong>of</strong> the effects due to bending <strong>and</strong> due to shear. The normal<br />

stresses due to bending are relatively high in the faces <strong>and</strong> may be considered<br />

constant, provided the faces are relatively thin compared to the distance between<br />

the two faces. The core takes the shear force. When the faces resist the shear<br />

deformation <strong>of</strong> the core, a local bending moment is created in both faces (Allen,<br />

1969; Zenkert, 1995). The sign <strong>of</strong> this local bending moment is the same for both<br />

faces. The total normal stresses in the faces are the sum <strong>of</strong> the stresses from the<br />

global bending moment <strong>and</strong> <strong>of</strong> the resistance to shear. When the span <strong>of</strong> the panel<br />

is large (like in figure 5.13a), the contribution <strong>of</strong> the local face bending moment,<br />

due to the resisting <strong>of</strong> the core shear deformation by the faces, is still relatively<br />

small. In figure 5.13a the resulting normal stresses are thus approximately constant<br />

across the face thickness.<br />

If the span <strong>of</strong> the panel is decreased, like in figure 5.13b the relative importance <strong>of</strong><br />

the normal stresses due to local bending moments (from resisting the shear<br />

deformation <strong>of</strong> the core) becomes more important. The sum <strong>of</strong> the normal stresses<br />

in the faces due to bending <strong>and</strong> shear effects cannot be considered being constant<br />

any more across the face thickness.<br />

Allen <strong>and</strong> Feng (1998) proposed a master diagram, which uses the relative<br />

importance <strong>of</strong> span, thickness <strong>of</strong> faces <strong>and</strong> core <strong>and</strong> their stiffness, to make a<br />

classification <strong>of</strong> the s<strong>and</strong>wich behaviour. It is a fairly simple <strong>and</strong> good guidance<br />

for the classification <strong>of</strong> s<strong>and</strong>wich panels, but it is assumed that all materials<br />

behave linear elastically. This classification method can be used for s<strong>and</strong>wich<br />

panels <strong>with</strong> IPC faces, but only <strong>with</strong> great care if the s<strong>and</strong>wich elements show<br />

non-linear behaviour. Possibly, recalculation <strong>of</strong> the final design solution should be<br />

performed <strong>with</strong> a more complex s<strong>and</strong>wich model.<br />

5.5 The behaviour <strong>of</strong> s<strong>and</strong>wich panels during unloading <strong>and</strong><br />

repeated loading: implementation in ANSYS<br />

From Chapter 3 <strong>and</strong> Chapter 4, two important conclusions should be kept in mind<br />

for the calculation <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces. (1) Residual<br />

strains, due to tensile loading <strong>and</strong> unloading <strong>of</strong> a IPC composite specimen can be<br />

large. (2) Accumulated strains due to repeated loading can be large. Translated to<br />

the behaviour <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces, this means residual<br />

deformations <strong>and</strong>/or extra displacement terms due to repeated loading might<br />

become very large.<br />

The behaviour <strong>of</strong> IPC during unloading <strong>and</strong> repeated loading cannot be<br />

represented by any st<strong>and</strong>ard finite element code. Therefore, two macros are<br />

163


Chapter 5: S<strong>and</strong>wich modelling<br />

written to represent the unloading behaviour <strong>of</strong> a face in tension <strong>and</strong> the behaviour<br />

<strong>of</strong> a face under repeated tensile loading.<br />

- The macro unload.mac, which is used after initial loading is calculated,<br />

loops over all face elements loaded in tension. The stiffness properties <strong>and</strong> ‘yield<br />

stresses’ are recalculated such that unloading occurs <strong>with</strong> element stiffness Ecycle<br />

for the x-component <strong>of</strong> the stiffness (Ex el = Ecycle). Ecycle is calculated according to<br />

Chapter 3. A different unloading stiffness Ecycle is thus attributed to each element,<br />

loaded in tension. More information on the methodology <strong>of</strong> the macro unload.mac<br />

can be found in Appendix 4.<br />

- A user-written macro, repeat.mac, provides an estimation <strong>of</strong> the extra<br />

strain terms in the IPC face elements under repeated loading. The program loops<br />

over all face elements in tension, providing a theoretical estimation <strong>of</strong> the<br />

accumulated strains for each element as a function <strong>of</strong> the maximum stress <strong>and</strong><br />

number <strong>of</strong> load cycles experienced by this element. The calculation <strong>of</strong> the<br />

accumulation <strong>of</strong> strains is based on the accumulation <strong>of</strong> damage in IPC composite<br />

specimens as discussed in Chapter 4. These extra element strains are then inserted<br />

in the s<strong>and</strong>wich model for the calculation <strong>of</strong> additional deflections <strong>of</strong> the s<strong>and</strong>wich<br />

panel, due to repeated loading. More information on the methodology <strong>of</strong> the macro<br />

repeat.mac can be found in Appendix 5.<br />

5.6 Conclusions<br />

Several topics have been discussed in this chapter: s<strong>and</strong>wich models <strong>and</strong> their<br />

applicability, the implementation <strong>of</strong> the IPC face behaviour under monotonic<br />

loading, unloading <strong>and</strong> repeated loading.<br />

It has been illustrated in paragraph 5.4 that the choice <strong>of</strong> an appropriate finite<br />

element can be rather delicate for the studied panels. A finite element model,<br />

which provides accurate results on maximum stresses <strong>and</strong> deflections while the<br />

IPC composite face behaviour is still linear elastical, may provide unreliable<br />

results once multiple cracking occurs in a face.<br />

Two material models, implemented in the st<strong>and</strong>ard finite element code in ANSYS<br />

are discussed for their feasibility to describe the behaviour <strong>of</strong> IPC under<br />

monotonic loading. The ‘multilinear elastic’ material behaviour can be used if it is<br />

clear a priori which parts <strong>of</strong> the faces are stressed in tension. This ‘multilinear<br />

elastic’ model can approximate the real stress-strain behaviour <strong>of</strong> IPC in tension in<br />

a more accurate way than the ‘aniso’ material option. The ‘aniso’ material option<br />

has the advantage that no a priori knowledge <strong>of</strong> the sign <strong>of</strong> the normal stresses in<br />

the faces is needed.<br />

164


Chapter 5: S<strong>and</strong>wich modelling<br />

A methodology is discussed for implementation <strong>of</strong> the unloading material<br />

behaviour <strong>of</strong> the IPC composite faces. This methodology is based on the stochastic<br />

cracking (based) model for unloading in Chapter 3. A macro has been written in<br />

ANSYS for this purpose, since this material behaviour does not come near any<br />

st<strong>and</strong>ard finite element material unloading law.<br />

Finally, the implementation <strong>of</strong> the behaviour <strong>of</strong> s<strong>and</strong>wich panels under repeated<br />

loading is in calculations discussed. The presented methodology provides an<br />

estimation <strong>of</strong> the additional deflections <strong>of</strong> the s<strong>and</strong>wich under repeated loading.<br />

The advantage <strong>of</strong> the methodology is found in the fact that it is not very timeconsuming<br />

<strong>and</strong> the implementation into the finite element calculations is rather<br />

simple.<br />

5.7 References<br />

R.L. Actis, B.A. Szabo, C. Schwab, Hierarchic models for laminated plates<br />

<strong>and</strong> shells, Comput. Methods Appl. Mech. Engrg. Vol. 172, 1999, pp.79-107<br />

H. G. Allen, <strong>Analysis</strong> <strong>and</strong> <strong>Design</strong> <strong>of</strong> Structural S<strong>and</strong>wich <strong>Panels</strong>, Permagon<br />

Press, 1969<br />

H. G. Allen, Z. Feng, Classification <strong>of</strong> structural s<strong>and</strong>wich panel<br />

behaviour, Proceedings EUROMECH, 13-15 May 1997, Kluwer Academic<br />

Publishers, pp.1-12<br />

ANSYS manuals, Release 5.5, 1998<br />

K. Berner, J.M. Davies, P. Hassinen, L. Heselius, Updated European<br />

recommendations for s<strong>and</strong>wich panels, Proceedings 5 th International Conference<br />

on S<strong>and</strong>wich Construction, EMAS Publishing, Sept 5-7, 2000,pp. 389-400<br />

W.S. Burton, A.K. Noor, Assessment <strong>of</strong> computational models for s<strong>and</strong>wich<br />

panels ans shells, Comput. Methods Appl. Mech. Engrg., Vol.124, 1995, pp.125-<br />

151<br />

Y. Frostig, Bending <strong>of</strong> Curved S<strong>and</strong>wich <strong>Panels</strong> <strong>with</strong> a Transversely<br />

Flexible Core – Closed – Form High-Order Theory, Journal <strong>of</strong> S<strong>and</strong>wich<br />

Structures <strong>and</strong> Materials, Vol. 1, 1999, pp.4-41<br />

Y. Frostig, M. Baruch, O. Vilnay <strong>and</strong> I. Scheinman, High-Order Theory for<br />

S<strong>and</strong>wich-Beam Behaviour <strong>with</strong> Transversely Flexible Core, Journal <strong>of</strong><br />

Engineering Mechanics, Vol.118, No.5, 1992, pp.1026-1043<br />

165


Chapter 5: S<strong>and</strong>wich modelling<br />

K.H. Ha, Finite element analysis <strong>of</strong> s<strong>and</strong>wich plates: an overview,<br />

Computers <strong>and</strong> Structures, Vol. 37, No. 4, 1990, pp.397-403<br />

V.Manet, The use <strong>of</strong> ANSYS to calculate the behaviour <strong>of</strong> s<strong>and</strong>wich<br />

structures, Composites Science <strong>and</strong> technology, Vol. 58, 1998, pp.1899-1905<br />

A.K. Noor, W.Scott. Burton, J.M. Peters, Hierarchical adaptive modeling<br />

<strong>of</strong> structural s<strong>and</strong>wiches <strong>and</strong> multilayered composite panels, Applied Numerical<br />

Mathematics, Vol.14, 1994, pp.69-90<br />

A.K. Noor <strong>and</strong> W. Scott Burton, Ch. W. Bert, Computational models for<br />

s<strong>and</strong>wich panels <strong>and</strong> shells, Appl. Mech. Rev., Vol. 49, No. 3, 1996, pp.155-199.<br />

J. O’Connor, A Finite Element Package for the <strong>Analysis</strong> <strong>of</strong> S<strong>and</strong>wich<br />

Constructions, Composite Structures, Vol. 8, 1987, pp.143-161<br />

C. Schwab, Hierarchic Modelling in Elasticity by Generalised p- <strong>and</strong> hp-<br />

FEM, Lecture notes summer school ‘Adaptive FEM for linear <strong>and</strong> nonlinear Solid<br />

<strong>and</strong> Structural Mechanics, C.I.M.E., Italy, Oct. 4-8, 1999<br />

V. Sokloinsky, Y. Frostig, On the response <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong><br />

transversely flexible core – special behaviour, Proceedings 5 th International<br />

Conference on S<strong>and</strong>wich Construction, EMAS Publishing, Sept 5-7, 2000, pp.49-<br />

60<br />

Stephen R. Swanson, An examination <strong>of</strong> a higher order theory for s<strong>and</strong>wich<br />

beams, Composite structures, 1999, pp.169-177<br />

Stephen R. Swanson <strong>and</strong> Jongman Kim, Comparison <strong>of</strong> a higher order<br />

theory for s<strong>and</strong>wich beams <strong>with</strong> finite element <strong>and</strong> elasticity analyses, Journal <strong>of</strong><br />

S<strong>and</strong>wich Structures <strong>and</strong> Materials, Vol.2-January 2000, pp.33-49<br />

K.Y. Sze, L.W. He, Y.K. Cheung, Predictor – corrector procedure for<br />

analysis <strong>of</strong> laminated plates using st<strong>and</strong>ard mindlin finite element models,<br />

Composite Structures, Vol. 50, 2000, pp.171-182<br />

O.T. Thomsen, J.R. Vinson, <strong>Design</strong> study <strong>of</strong> non-circular pressurized<br />

s<strong>and</strong>wich fuselage section using a high-order s<strong>and</strong>wich theory formulation,<br />

Proceedings 5 th International Conference on S<strong>and</strong>wich Construction, EMAS<br />

Publishing, Sept 5-7, 2000, pp.3-14<br />

166


Chapter 5: S<strong>and</strong>wich modelling<br />

O.T. Thomsen, <strong>Analysis</strong> <strong>of</strong> local bending effects in s<strong>and</strong>wich panels<br />

subjected to concentrated loads, Proc. S<strong>and</strong>wich Constructions 2, Gainesville,<br />

Florida, USA, March 9-12, 1992<br />

O.T. Thomsen, Theoretical <strong>and</strong> experimental investigation <strong>of</strong> local bending<br />

effects in s<strong>and</strong>wich plates, Composite Structures, Vol. 30, 1995, pp.85-101<br />

D. Zenkert, An introduction to s<strong>and</strong>wich construction, Chameleon Press<br />

Ltd, 1995<br />

167


Chapter 6<br />

Experimental work on s<strong>and</strong>wich panels<br />

6.1 Introduction<br />

Results <strong>of</strong> four-point bending tests on s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

are presented in this chapter. The influence <strong>of</strong> the core <strong>and</strong> face thickness <strong>and</strong> the<br />

span is examined <strong>and</strong> discussed. The strains <strong>and</strong> deflections <strong>of</strong> these panels are<br />

monitored under monotonic loading, unloading <strong>and</strong> repeated loading. The<br />

experimental load-deflection <strong>and</strong> load-strain curves, obtained on each s<strong>and</strong>wich<br />

panel, are compared <strong>with</strong> the theoretical curves. The constitutive equations<br />

obtained from the stochastic cracking model, as presented in Chapter 2 <strong>and</strong><br />

Chapter 3, are used as face material behaviour. The behaviour <strong>of</strong> the IPC<br />

composites under repeated loading, as obtained <strong>and</strong> discussed in Chapter 4, is<br />

used to predict the behaviour <strong>of</strong> a s<strong>and</strong>wich panel <strong>with</strong> IPC composite faces under<br />

repeated loading. Prior to any testing on s<strong>and</strong>wich panels, the strength <strong>of</strong> the coreface<br />

adhesion in tension <strong>and</strong> shear is determined experimentally. These failure<br />

modes are to be foreseen <strong>and</strong> - even more desirably - avoided.<br />

6.2 Strength <strong>of</strong> the core-face interface<br />

Most s<strong>and</strong>wich panels for building construction purposes contain an adhesive to<br />

assure proper co-operation <strong>of</strong> faces <strong>and</strong> core. The core is usually foamed between<br />

the faces. The same process can be used when IPC composites are used as faces,<br />

but it might be more advantageous if no adhesive is needed. It is verified<br />

experimentally here whether it is possible to assure sufficient core-face interface<br />

adhesion when the IPC faces are laminated directly on a core material.<br />

Experimental results on the core-face interface are presented below.<br />

6.2.1 failure in tension<br />

A low core-face interface tensile strength could lead to early failure <strong>of</strong> a s<strong>and</strong>wich<br />

panel. The tensile interface strength <strong>of</strong> a s<strong>and</strong>wich panel <strong>with</strong> a PUR core <strong>of</strong><br />

169


Chapter 6: Experimental work on s<strong>and</strong>wich panels<br />

35kg/m³ <strong>and</strong> IPC faces is determined here<br />

according to the ECCS publication:<br />

Preliminary European Recommendations for<br />

S<strong>and</strong>wich <strong>Panels</strong>, part I, <strong>Design</strong>, (1991). On<br />

both sides <strong>of</strong> a PUR core <strong>of</strong> 80mm thickness, a<br />

2D-r<strong>and</strong>omly reinforced composite is<br />

laminated on this core. After the faces are<br />

post-cured, small s<strong>and</strong>wich specimens <strong>of</strong><br />

50x50mm² are cut from this panel. On both<br />

IPC faces <strong>of</strong> such a small s<strong>and</strong>wich specimen,<br />

a metal connection element is glued <strong>with</strong> a<br />

PUR adhesive. These metal connection<br />

elements are used to transfer a tensile force<br />

from a INSTRON 1905 testing bench to the<br />

s<strong>and</strong>wich element. The test set-up is illustrated<br />

in figure 6.1. It can be seen in this picture that<br />

failure<br />

Figure 6.1: test set-up core-face<br />

interface tensile strength<br />

the specimen failed in the core material <strong>and</strong> not at the core-face interface.<br />

Conclusively, the core-face interface is stronger than the core material, which is<br />

advantageous.<br />

Several specimens were tested. The average measured failure stress is 0.20MPa,<br />

which is an order <strong>of</strong> magnitude found in literature for the tensile failure stress <strong>of</strong> a<br />

PUR foam <strong>with</strong> a density <strong>of</strong> 35kg/m³. (Gibson <strong>and</strong> Ashby, 1988)<br />

6.2.2 shear failure<br />

The European Recommendations for S<strong>and</strong>wich <strong>Panels</strong> (ECCS, 1991) do not<br />

provide a st<strong>and</strong>ard test for the determination <strong>of</strong> core-face interface shear strength<br />

properties. The ASTM st<strong>and</strong>ards E229-70, D 3163 <strong>and</strong> D3164 provide test set-ups<br />

for the determination <strong>of</strong> an adhesive shear strength. The core-face shear failure test<br />

is based on ASTM st<strong>and</strong>ards D3163 <strong>and</strong> D3164, describing the strength <strong>of</strong><br />

adhesively bonded plastic lap-shear joints. The test set-up used here is illustrated<br />

in figure 6.2b <strong>and</strong> a picture <strong>of</strong> the test set-up is printed in figure 6.2c. The purpose<br />

<strong>of</strong> this experiment is to know whether the core-face interface shear strength is<br />

higher than the core material strength or vice versa.<br />

A s<strong>and</strong>wich panel <strong>with</strong> core thickness <strong>of</strong> 75mm, length <strong>of</strong> 300mm <strong>and</strong> width <strong>of</strong><br />

300mm is made. Both E-glass fibre reinforced IPC faces are laminated directly on<br />

the core. Two layers <strong>of</strong> 2D-r<strong>and</strong>omly oriented reinforcement are inserted in each<br />

face. The fibre volume fraction <strong>of</strong> the faces is approximately 10%. The s<strong>and</strong>wich<br />

panels are cured in ambient conditions for 24 hours <strong>and</strong> are post-cured at 60°C for<br />

another 24 hours. Small s<strong>and</strong>wich panels <strong>of</strong> 250mm by 25mm are cut from the<br />

large s<strong>and</strong>wich panel. The core is then partially removed from the face: core <strong>and</strong><br />

170


Chapter 6: Experimental work on s<strong>and</strong>wich panels<br />

face are still connected to each other across 20mm height. The geometry <strong>of</strong> the<br />

specimens is illustrated in figure 6.2a.<br />

20mm<br />

IPC composite<br />

core<br />

core-face interface<br />

Figure 6.2a: geometry specimens for determination <strong>of</strong> the core-face interface shear strength<br />

fixed steel element,<br />

preventing movement<br />

in horizontal direction<br />

tensile force from<br />

INSTRON clamp<br />

fixed steel element, preventing<br />

movement <strong>of</strong> core in vertical<br />

direction<br />

Figure 6.2b: core-face interface shear strength<br />

test-setup<br />

clamp<br />

fixed steel element<br />

IPC laminate<br />

fixed steel elements<br />

core<br />

all fixed steel elements fixed to<br />

INSTRON beam<br />

Figure 6.2c: core-face interface shear strength<br />

test-setup<br />

As can be seen in figures 6.2b <strong>and</strong> 6.2c, the foam is clamped to the bottom <strong>of</strong> the<br />

mechanical INSTRON 1905 testing bench by a fixed steel element. It is the upper<br />

face <strong>of</strong> the core that is thus restrained in vertical direction. The upper part <strong>of</strong> the<br />

IPC laminate, free from foam, is clamped. This side is pulled upwards during the<br />

test. The IPC is supported at the left side by another fixed steel element,<br />

preventing the IPC specimen from moving horizontally. Since tilting <strong>of</strong> the<br />

specimen is thus minimised, the core-face interface will be loaded in shear mainly.<br />

The height <strong>of</strong> the core-face interface is 20mm. The width is 25mm.<br />

171


Chapter 6: Experimental work on s<strong>and</strong>wich panels<br />

Failure occurs in the core, close to the core-face interface for all tested specimens.<br />

This means the core fails in shear, before the core-face interface fails. A value <strong>of</strong><br />

the shear failure stress is not determined from this test, since the shear stress is not<br />

uniform across the height <strong>of</strong> the specimen. The nominal value <strong>of</strong> the shear failure<br />

stress depends on the height <strong>of</strong> the core-face interface. The shear failure stress <strong>of</strong><br />

the core is determined further in this chapter.<br />

6.3 Large-span s<strong>and</strong>wich panels: materials <strong>and</strong> geometry<br />

6.3.1 face materials<br />

The faces are laminated directly on the core. The 2D-r<strong>and</strong>om reinforcement, <strong>with</strong><br />

a surface density <strong>of</strong> 300g/m² per layer, is used in these faces. The amount <strong>of</strong><br />

matrix used, is 1600 g/m²/layer. At the core-face interface 10% extra matrix is<br />

used. Some <strong>of</strong> the matrix will penetrate the core cells at surface, assuring a better<br />

mechanical core-face adhesion. The average fibre volume fraction obtained in this<br />

way is about 10%. All panels are kept in ambient conditions for 24 hours <strong>and</strong> are<br />

then post-cured for another 24 hours at 60°C.<br />

6.3.2 core material<br />

The core is a PUR layer <strong>with</strong> a density <strong>of</strong> 35kg/m³. The thickness <strong>of</strong> the core plate<br />

is 40mm, 60mm or 80mm. All panels are cut from one large block, so no large<br />

variations <strong>of</strong> the material properties along the thickness <strong>of</strong> one core plate are to be<br />

expected. Several beam specimens are cut from the panel <strong>of</strong> 40mm thickness to<br />

obtain an idea <strong>of</strong> the shear stiffness <strong>and</strong> strength <strong>of</strong> the core material. Small beams<br />

<strong>with</strong> variable length (400mm or less) <strong>and</strong> a section <strong>of</strong> 40x40mm² are used. The<br />

shear modulus <strong>and</strong> the E-modulus are determined by means <strong>of</strong> measurement <strong>of</strong> the<br />

eigenfrequencies. The shear modulus <strong>and</strong> strength are measured on a torsion<br />

bench. The results from these tests are:<br />

• from eigenfrequencies: G = 1.8MPa, E = 4.0MPa<br />

• torsion bench: G = 2MPa, τxy failure = 0.2MPa<br />

6.3.3 geometry<br />

Table 6.1: geometry s<strong>and</strong>wich panels, 2m span<br />

Name core thickness # face layers<br />

mm -<br />

HC/S7 80 2<br />

HC/S8 80 3<br />

HC/S9 60 2<br />

HC/S10 60 3<br />

HC/S11 40 2<br />

HC/S12 40 3<br />

172


Chapter 6: Experimental work on s<strong>and</strong>wich panels<br />

Several s<strong>and</strong>wich panels are made. The length <strong>of</strong> all panels is 2000mm, the width<br />

is 300mm <strong>and</strong> the thickness is variable. The geometrical properties <strong>of</strong> the tested<br />

panels are found in table 6.1. Six different combinations <strong>of</strong> core <strong>and</strong> face thickness<br />

are made.<br />

6.4 panels <strong>with</strong> 2m span: test set-up<br />

All s<strong>and</strong>wich panels are subjected to four-point bending. Figure 6.3 illustrates the<br />

geometry <strong>of</strong> the test set-up. The distance between the lower supports is 1800mm.<br />

The distance between the upper supports is 600mm. A 10kN load cell measures<br />

the force.<br />

strain gauge 3<br />

strain gauge 2<br />

600mm<br />

strain gauge 1<br />

1800mm<br />

LVDT 1 & 2<br />

LVDT 1<br />

LVDT 2<br />

Figure 6.3: test set-up four point bending test on 2m span s<strong>and</strong>wich panels <strong>and</strong> placement <strong>of</strong><br />

strain gauges<br />

Two LVDT’s measure the displacement at 100mm from mid-span <strong>of</strong> the s<strong>and</strong>wich<br />

panel. It would be more convenient to place these LVDT’s at mid-span, but this<br />

was not possible for practical reasons. However, the results <strong>of</strong> the theoretical<br />

calculation <strong>of</strong> the force-deflection curves will be given for a location at 100mm<br />

from mid-span for all panels. If two LVDT’s are used along the width <strong>of</strong> the<br />

s<strong>and</strong>wich panel, misalignment <strong>of</strong> the set-up is detected in time. From the resulting<br />

force-displacement curves, it has been noticed that the measured differences<br />

between the two LVDT’s could be neglected for all panels. Therefore, the mean<br />

displacement will be given.<br />

Three strain gauges are attached to the surface <strong>of</strong> each s<strong>and</strong>wich panel. One strain<br />

gauge is located at mid-span at the bottom <strong>of</strong> the lower face (strain gauge 1). A<br />

second strain gauge is attached to the bottom <strong>of</strong> the lower face at 1/4 th <strong>of</strong> the<br />

length (strain gauge 2). The third strain gauge is glued at the top <strong>of</strong> the upper face<br />

at 1/4 th <strong>of</strong> the length (strain gauge 3). Figures 6.4a <strong>and</strong> 6.4b show the total test setup<br />

<strong>and</strong> the placement <strong>of</strong> the upper supports in detail.<br />

173


Chapter 6: Experimental work on s<strong>and</strong>wich panels<br />

Figure 6.4a: test set-up panels <strong>with</strong> 2m span Figure 6.4b: test set-up panels <strong>with</strong> 2m span,<br />

detail upper support<br />

At the upper <strong>and</strong> lower supports, steel plates <strong>of</strong> 4mm thickness, <strong>and</strong> 60x300mm²<br />

surface are attached to the s<strong>and</strong>wich panel to ensure more distributed transfer <strong>of</strong><br />

forces from the supports to the s<strong>and</strong>wich panel. Two extra rubber pieces are placed<br />

between the upper steel plates <strong>and</strong> the s<strong>and</strong>wich for further minimisation <strong>of</strong> stress<br />

concentrations.<br />

In paragraph 6.3.2, it has been verified that the measured core material shear<br />

failure stress is about 0.2MPa. Below 0.75MPa no fracture or serious non-linear<br />

behaviour under core shear are thus expected. When 0.75MPa is adopted as the<br />

maximum core shear stress, the maximum load applied on the s<strong>and</strong>wich panels is<br />

1200N when the core thickness is 40mm. This load is 1800N for the panels <strong>with</strong><br />

60mm thick core <strong>and</strong> 2400N for the panels <strong>with</strong> 80mm core thickness. These<br />

values are further referred to as maximum loads.<br />

All panels are subjected to subsequent loading steps. A panel is first loaded up to<br />

2/3 rd <strong>of</strong> the maximum load, defined in paragraph 6.4. The panel is cycled 10 times<br />

between 2/3 rd <strong>of</strong> the maximum load <strong>and</strong> 1/3 rd <strong>of</strong> this 2/3 rd <strong>of</strong> the maximum load<br />

(thus 2/9 th <strong>of</strong> the maximum load). Immediately after the panels are cycled 10<br />

times, they are subsequently loaded up to the maximum defined load (1200N for<br />

40mm core thickness, etc.).<br />

6.5 panels <strong>with</strong> 2m span: results<br />

The measured maximum deflections are listed in table 6.2 for the subsequent load<br />

steps, defined in paragraph 6.4<br />

174


Chapter 6: Experimental work on s<strong>and</strong>wich panels<br />

Table 6.2: measured maximum deflections at several load stages<br />

HC/S7 HC/S8 HC/S9 HC/S10 HC/S11 HC/S12<br />

2/3 rd , 1 st cycle 19.1 12.0 22.5 12.2 27.0 16.3<br />

2/3 rd , 10 th maximum deflection (mm)<br />

cycle 21.5 13.2 25.3 13.3 30.5 18.4<br />

maximum load 33.4 21.5 40.7.9 24.1 54.1 30.8<br />

Figures 6.5a <strong>and</strong> 6.5b are two pictures taken at the lower face <strong>of</strong> the s<strong>and</strong>wich<br />

panel after loading. Picture 6.5a is taken at 1/4 th <strong>of</strong> the length, in the vicinity <strong>of</strong><br />

strain gauge 2. Picture 6.5b shows the lower face at mid-span, in the vicinity <strong>of</strong><br />

strain gauge 1. It can be noticed from these figures that the matrix crack density is<br />

highest at mid-span. The crack density decreases at 1/4 th <strong>of</strong> the span-length <strong>and</strong><br />

becomes nearly zero towards the lower support.<br />

Figure 6.5a: lower face at 1/4 th <strong>of</strong> the length,<br />

after loading, HC/S7 (80mm core, 2<br />

layers/face)<br />

Figure 6.5b: lower face at 1/2 nd <strong>of</strong> the<br />

length, after loading HC/S7 (80mm core, 2<br />

layers/face)<br />

To illustrate the effect <strong>of</strong> the number <strong>of</strong> face layers, the s<strong>and</strong>wich behaviour <strong>of</strong> two<br />

panels <strong>with</strong> equal geometry are illustrated more in detail. The only difference<br />

between these panels is the number <strong>of</strong> face layers. Panel HC/S9 contains a core<br />

thickness <strong>of</strong> 60mm <strong>and</strong> two layers <strong>of</strong> 2D-r<strong>and</strong>omly reinforced IPC layers per face.<br />

Panel HC/S10 contains a core <strong>with</strong> the same thickness <strong>and</strong> three layers <strong>of</strong> 2Dr<strong>and</strong>omly<br />

reinforced IPC laminae per face. The evolution <strong>of</strong> the strains, measured<br />

by the strain gauges is presented in figure 6.6a for panel HC/S9 <strong>and</strong> in figure 6.7a<br />

for panel HC/S10. The evolution <strong>of</strong> the deflections is illustrated in figure 6.6b <strong>and</strong><br />

6.7b for panels HC/S9 <strong>and</strong> HC/S10 respectively.<br />

Preliminary conclusions from figures 6.6a <strong>and</strong> 6.6b are:<br />

- Multiple cracking is obviously monitored at both strain gauges at the<br />

lower face (strain gauge 1 <strong>and</strong> strain gauge 2).<br />

175


Chapter 6: Experimental work on s<strong>and</strong>wich panels<br />

- If figure 6.6a <strong>and</strong> 6.6b are compared, it can be noticed that, once matrixmultiple<br />

cracking is clearly detected by strain gauge 1 on the lower face, the forcedeflection<br />

curve starts to deflect as well.<br />

force (N)<br />

strain gauge 3<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

strain gauge 2<br />

strain gauge 1<br />

-0.1 0 0.1 0.2 0.3<br />

strain (%)<br />

Figure 6.6a: strain-force curve<br />

HC/S9 (60mm core thickness, 2layers/face)<br />

strain gauge 3<br />

force (N)<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

strain gauge 2<br />

strain gauge 1<br />

-0.04 -0.02 0 0.02 0.04 0.06 0.08<br />

strain (%)<br />

Figure 6.7a: strain-force curve<br />

HC/S10 (60mm core thickness, 3layers/face)<br />

force (N)<br />

force (N)<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

0 2 4 6 8 10 12 14 16 18 20 22 24 26<br />

displacement (mm)<br />

Figure 6.6b: deflection-force curve<br />

HC/S9 (60mm core thickness, 2layers/face)<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

0 2 4 6 8 10 12 14<br />

displacement (mm)<br />

Figure 6.7b: deflection-force curve<br />

HC/S10 (60mm core thickness, 3layers/face)<br />

From comparison <strong>of</strong> figures 6.7a <strong>and</strong> 6.7b <strong>with</strong> 6.6a <strong>and</strong> 6.6b, it can be noticed<br />

that, when one extra lamina per face is added, the effects due to multiple cracking<br />

are less pronounced during first loading up to 1200N. However, during unloading<br />

residual strains are noticed clearly in figure 6.7a <strong>and</strong> residual deflections can be<br />

found in figure 6.7b. Upon cycling, this effect becomes even more pronounced.<br />

6.6 2m span panels: theoretical predictions versus<br />

experimental results<br />

6.6.1 face properties <strong>of</strong> the tested panels<br />

The s<strong>and</strong>wich panels are removed from the test set-up after cycling. None <strong>of</strong> the<br />

panels has been loaded up to failure. The faces are removed from the core<br />

176


Chapter 6: Experimental work on s<strong>and</strong>wich panels<br />

carefully. The thickness <strong>of</strong> the faces is measured at 5 locations along the length to<br />

obtain an idea about the variation <strong>of</strong> the thickness <strong>of</strong> the faces along each<br />

s<strong>and</strong>wich panel. Specimens are cut from each face at length locations 0mm,<br />

500mm, 1000mm, 1500mm <strong>and</strong> 2000mm. At each length location, two IPC<br />

composite specimens are cut from the face. The average value <strong>and</strong> the st<strong>and</strong>ard<br />

variation <strong>of</strong> the face thickness are listed in table 6.3.<br />

Table 6.3: thickness variation <strong>of</strong> tested s<strong>and</strong>wich panel faces<br />

HC/S7 HC/S8 HC/S9 HC/S10 HC/S11 HC/S12<br />

lower face thickness (mm)<br />

average 2.96 3.89 3.32 4.09 2.92 4.26<br />

st<strong>and</strong>ard dev 0.19 0.47 0.28 0.46 0.31 0.36<br />

upper face thickness (mm)<br />

average 2.89 4.01 2.88 4.26 3.24 4.27<br />

st<strong>and</strong>ard dev 0.28 0.46 0.41 0.36 0.46 0.53<br />

From the upper face, four extra composite specimens are cut <strong>and</strong> tested under<br />

monotonic tensile loading. These specimens are cut far from the middle <strong>of</strong> the<br />

s<strong>and</strong>wich. Specimens are thus cut from a region, where only minor compressive<br />

stresses occurred during testing. Two specimens are cut from the left side <strong>and</strong> two<br />

from the right side <strong>of</strong> the upper face <strong>of</strong> the panel. The program IPCstressstrain.exe<br />

is used to obtain the material parameters, which are needed as input into<br />

the finite element calculations (m, σR, τ0). The fibre volume fraction Vf is<br />

calculated from knowledge <strong>of</strong> the average thickness <strong>of</strong> the specimens, which were<br />

cut from the faces <strong>of</strong> the panels (see equation 2.1 <strong>and</strong> table 6.3).<br />

In table 6.4 the fibre volume fraction <strong>and</strong> parameters m, σR <strong>and</strong> τ0, as obtained by<br />

the program IPCstress-strain.exe from fitting <strong>of</strong> the theoretical stress-strain curves<br />

<strong>with</strong> the experiments, are listed. For each panel, the average value <strong>of</strong> Vf, m, σR <strong>and</strong><br />

τ0 <strong>of</strong> the four tested specimens is used as material input for s<strong>and</strong>wich calculations.<br />

It should be kept in mind that the material properties <strong>of</strong> the lower face <strong>and</strong> the<br />

upper face might be slightly different. However, since the lower face has been<br />

loaded in tension already, the initial material properties (m, σR <strong>and</strong> τ0) cannot be<br />

obtained any more from the lower face.<br />

Table 6.4: IPC composite material parameters from upper faces <strong>of</strong> tested s<strong>and</strong>wich panel faces<br />

HC/S7 HC/S8 HC/S9 HC/S10 HC/S11 HC/S12<br />

average tupper (mm) 2.96 3.89 3.32 4.09 2.92 4.26<br />

averageVf (%) 8.01 9.18 7.07 8.71 8.15 7.97<br />

Weibull modulus m(-) 3.3 4.2 3.8 3.4 4.5 4.1<br />

average σR (N/mm²) 9.2 12 12 7.0 11 12<br />

average τ0 (N/mm²) 0.44 0.92 0.89 0.54 0.45 0.61<br />

177


Chapter 6: Experimental work on s<strong>and</strong>wich panels<br />

The properties, listed in table 6.3 <strong>and</strong> 6.4 are used as input for the finite element<br />

calculations for prediction <strong>of</strong> the behaviour <strong>of</strong> the s<strong>and</strong>wich panels.<br />

One important comment on the testing program in this chapter is necessary. The<br />

material properties Vf, σR, m <strong>and</strong> τ0 are determined on each panel separately,<br />

before they are inserted into the finite element calculations. This might seem<br />

rather odd <strong>and</strong> very time-consuming. However, the aim <strong>of</strong> this chapter is to check<br />

whether the force-deflection behaviour <strong>of</strong> a panel <strong>with</strong> IPC faces can be predicted<br />

by a very limited number <strong>of</strong> material properties, characteristic for the panel in<br />

question. It is the quality <strong>of</strong> the model that is discussed <strong>and</strong> not the quality <strong>of</strong> the<br />

panels. There are at this point still too many factors that influence the matrix <strong>and</strong><br />

impregnation quality <strong>of</strong> the composite faces to assume all panels have equal face<br />

behaviour. Not only the quality <strong>of</strong> the IPC powder component varies from one<br />

batch to another, also the h<strong>and</strong> lay-up technique introduces large variations in face<br />

quality from one panel to another. Once the IPC fabrication <strong>and</strong> s<strong>and</strong>wich panel<br />

fabrication are better controlled, these values might be determined on one panel <strong>of</strong><br />

a batch.<br />

6.6.2 theoretical prediction <strong>of</strong> the load-displacement behaviour <strong>of</strong> the<br />

tested panels <strong>with</strong> ANSYS<br />

The implementation <strong>of</strong> the IPC composite stress-strain behaviour <strong>and</strong> choice <strong>of</strong><br />

finite element model(s) in ANSYS are first discussed in this paragraph.<br />

The upper IPC composite face is assumed to behave linear elastically up to failure<br />

(see Chapter 2). The Young’s modulus <strong>of</strong> the IPC composite in compression is<br />

chosen 18GPa, according to table 2.4.<br />

The core material is modelled as an isotropic linear elastic material, <strong>with</strong> the value<br />

<strong>of</strong> the material parameters as obtained in paragraph 6.3.2.<br />

The behaviour <strong>of</strong> the lower IPC face in ANSYS is introduced by the ‘multilinear<br />

elastic’ material option. The material properties in table 6.4 are thus inserted into<br />

the consitutive equations <strong>of</strong> IPC composite faces (see Chapter 2 to Chapter 4) to<br />

calculate the global force-deflection behaviour <strong>of</strong> the tested s<strong>and</strong>wich panels.<br />

The face thicknesses <strong>of</strong> the upper <strong>and</strong> lower faces, used in the finite element<br />

geometry, are the average value obtained in table 6.3. The core thickness, used in<br />

the finite element model, is 40mm, 60mm or 80mm.<br />

Only half the s<strong>and</strong>wich panel is modelled <strong>and</strong> symmetrical boundary conditions<br />

are used. The displacements in the x-direction (length <strong>of</strong> the panel) are blocked in<br />

the nodes at mid-span <strong>of</strong> the panel. The s<strong>and</strong>wich panel is modelled by 2Dcontinuum<br />

elements along the length <strong>and</strong> across the thickness. The number <strong>of</strong><br />

178


Chapter 6: Experimental work on s<strong>and</strong>wich panels<br />

elements along the length is 100, 10 elements are used across the core thickness<br />

<strong>and</strong> 5 element are used across the face thickness.<br />

6.6.3 theory versus experiments: initial loading to 2/3 rd <strong>of</strong> the<br />

maximum load<br />

6.6.3.1 load-displacement curves<br />

Figures 6.8 to 6.13 show the measured displacement <strong>and</strong> the displacements<br />

calculated <strong>with</strong> FEM versus increasing force up to 2/3 rd <strong>of</strong> the maximum load. It<br />

can be seen from these figures that the evolution <strong>of</strong> the displacement versus force<br />

is predicted rather well theoretically. The theoretical <strong>and</strong> experimental curves<br />

show good coincidence for most panels.<br />

One can notice from figure 6.9 that the experimental load deflection behaviour <strong>of</strong><br />

s<strong>and</strong>wich panel HC/S8 (80mm core, 3 layers/face) is considerably stiffer than the<br />

theoretical curve. Possible explanations are discussed later in this paragraph.<br />

force (N)<br />

force (N)<br />

2000<br />

1500<br />

1000<br />

500<br />

1500<br />

1000<br />

0<br />

FEM calculation<br />

experiment<br />

0 5 10 15 20<br />

displacement (mm)<br />

Figure 6.8: FEM versus experimental<br />

deflection-force curve, HC/S7<br />

80mm core thickness, 2 face layers<br />

500<br />

0<br />

FEM<br />

calculation<br />

0 5 10 15 20 25<br />

displacement (mm)<br />

Figure 6.10: FEM versus experimental<br />

deflection-force curve, HC/S9<br />

60mm core thickness, 2 face layers<br />

force (N)<br />

force (N)<br />

179<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

1500<br />

1000<br />

FEM calulation<br />

experiment<br />

0 5 10 15 20<br />

displacement (mm)<br />

Figure 6.9: FEM versus experimental<br />

deflection-force curve, HC/S8<br />

80mm core thickness, 3 face layers<br />

500<br />

0<br />

FEM<br />

calculation<br />

0 5 10 15<br />

displacement (mm)<br />

Figure 6.11: FEM versus experimental<br />

deflection-force curve, HC/S10<br />

60mm core thickness, 3 face layers


force (N)<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

Chapter 6: Experimental work on s<strong>and</strong>wich panels<br />

FEM calculation<br />

experiment<br />

0 5 10 15 20 25<br />

displacement (mm)<br />

Figure 6.12 FEM versus experimental<br />

deflection-force curve, HC/S11<br />

40mm core thickness, 2 face layers<br />

30<br />

force (N)<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

FEM calculation<br />

experiment<br />

0 5 10 15 20<br />

displacement (mm)<br />

Figure 6.13: FEM versus experimental<br />

deflection-force curve, HC/S12<br />

40mm core thickness, 3 face layers<br />

In figures 6.8, 6.10 <strong>and</strong> 6.12 it can be seen that the experimental load-deflection<br />

curve <strong>of</strong> the s<strong>and</strong>wich panels <strong>with</strong> 2 layers <strong>of</strong> 2D-r<strong>and</strong>omly reinforced IPC shows<br />

noticeably non-linear behaviour. The theoretical load-deflection curves <strong>of</strong> the<br />

s<strong>and</strong>wich panels <strong>with</strong> 2 layers/face, as calculated <strong>with</strong> FEM, predict this nonlinear<br />

behaviour rather well.<br />

From figures 6.9, 6.11 <strong>and</strong> 6.13 it can be seen that the load-deflection behaviour is<br />

considerably stiffer if 3 layers/face are used instead <strong>of</strong> 2 layers/face. This effect is<br />

predicted theoretically from the FEM calculations. The deflection from linear<br />

load-deflection behaviour is less pronounced experimentally as well as<br />

theoretically.<br />

6.6.3.2 load-strain curves<br />

Figures 6.14 <strong>and</strong> 6.15 show the evolution <strong>of</strong> the strain measured by the strain<br />

gauges, in function <strong>of</strong> the applied load for panels HC/S9 <strong>and</strong> HC/S10. Both the<br />

experimental <strong>and</strong> theoretical (FEM calculation) curves are presented.<br />

force (N)<br />

1500<br />

1000<br />

500<br />

0<br />

FEM calculation<br />

experiment<br />

-0.1 0 0.1 0.2 0.3<br />

strain (%)<br />

Figure 6.14: FEM versus experimental strain<br />

versus force curve, HC/S9<br />

60m m core thickness, 2 face layers<br />

180<br />

force (N)<br />

1500<br />

1000<br />

500<br />

0<br />

FEM calculation<br />

experiment<br />

-0.04 -0.02 0 0.02 0.04 0.06 0.08<br />

strain (%)<br />

Figure 6.15: FEM versus experimental strain<br />

versus force curve, HC/S10<br />

60mm core thickness, 3 face layers


Chapter 6: Experimental work on s<strong>and</strong>wich panels<br />

The strain is a local variable, whereas the displacement is a more global variable.<br />

This is an explanation why the theoretical load-strain curves show higher<br />

discrepancy <strong>with</strong> the experimental load-strain curves (see figure 6.14 <strong>and</strong> 6.15)<br />

than was the case for the load-displacement curves (see figure 6.10 <strong>and</strong> 6.11).<br />

Figure 6.16 illustrates probably why the theoretical (FEM) predicted maximum<br />

deflections (see figure 6.9) have been overestimated for panel HC/S8 (80mm core,<br />

3layers/face). The measured normal strains are considerably lower in the<br />

experimental curves than is predicted theoretically. As has been mentioned earlier,<br />

the IPC composite material properties were derived from tests on specimens cut<br />

from the upper face. In case <strong>of</strong> s<strong>and</strong>wich panel HC/S8, the upper <strong>and</strong> lower face<br />

material properties are probably considerably different.<br />

force (N)<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

FEM calculation<br />

experiment<br />

-0.04 -0.02 0 0.02 0.04 0.06 0.08 0.1<br />

strain (%)<br />

Figure 6.16: FEM versus experimental strain versus force curve, HC/S8<br />

80mm core thickness, 3 face layers<br />

In table 6.5, the theoretical (FEM) <strong>and</strong> experimental values <strong>of</strong> the displacement at<br />

2/3 rd <strong>of</strong> the maximum load are listed in this table. This displacement is calculated<br />

two times <strong>with</strong> FEM:<br />

1. The theoretical value <strong>of</strong> the maximum displacement is calculated <strong>with</strong><br />

‘multilinear elastic’ material face behaviour for the tensile face.<br />

2. All panels are also calculated a second time, but now the assumption <strong>of</strong><br />

linear elastic behaviour <strong>of</strong> both faces is used. The face material properties in<br />

tension are now obtained from the law <strong>of</strong> mixtures, considering no matrix cracking<br />

occur (same behaviour in tension <strong>and</strong> compression).<br />

Table 6.5: maximum deflection under 2/3 rd <strong>of</strong> the maximum load, initial loading, experiment,<br />

theory <strong>and</strong> theory <strong>with</strong> linear elastic behaviour.<br />

maximum deflection (mm)<br />

HC/S7 HC/S8 HC/S9 HC/S10 HC/S11 HC/S12<br />

experiment 19.1 12.0 22.5 12.2 27.0 16.3<br />

FEM, stochastic cracking 19.9 14.8 19.8 13.2 24.7 16.2<br />

error (%) +1 +23 -12 +8 -9 -1<br />

FEM, linear elastic 14.9 13.7 13.5 12.1 16.1 13.8<br />

error (%) -22 -14 -40 -1 -40 -15<br />

181


Chapter 6: Experimental work on s<strong>and</strong>wich panels<br />

Table 6.5 is discussed here, <strong>with</strong> exclusion <strong>of</strong> the results obtained on panel HC/S8.<br />

It has been mentioned before that the quality <strong>of</strong> the upper <strong>and</strong> lower face in this<br />

panel must have been considerably different.<br />

From table 6.5, following conclusions can be taken:<br />

- If both faces are supposed to behave linear elastically, the predicted<br />

deflections <strong>of</strong> the s<strong>and</strong>wich panels are systematically lower than the<br />

experimentally obtained values. The discrepancy between the theoretical<br />

calculations <strong>with</strong> linear elastic face behaviour <strong>and</strong> experimental values <strong>of</strong> the<br />

deflections varies between –1% <strong>and</strong> –40%<br />

- If multiple cracking behaviour <strong>of</strong> the tensile face is inserted into the finite<br />

element calculations, the theoretical predictions are not longer systematically<br />

higher or lower than the experimental values. The discrepancy now varies between<br />

–12 <strong>and</strong> +8, which means the absolute value <strong>of</strong> the discrepancy is now<br />

considerably lower, compared to the assumption <strong>of</strong> linear elastic face behaviour.<br />

6.6.4 theory versus experiments: initial unloading from 2/3 rd <strong>of</strong> the<br />

maximum load<br />

In this paragraph the first unloading-reloading force-deflections cycle <strong>of</strong> the tested<br />

s<strong>and</strong>wich panel is studied. Theoretical unloading <strong>of</strong> the panels is performed by the<br />

user-written macro unload.mac, as explained in Chapter 5 (paragraph 5.5).<br />

All s<strong>and</strong>wich panels have first been loaded to 2/3 rd <strong>of</strong> the maximum load <strong>and</strong> then<br />

unloaded to 2/9 th <strong>of</strong> the maximum load. The displacement variation, which will be<br />

obtained experimentally <strong>and</strong> theoretically, is illustrated in figure 6.17.<br />

force (N)<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

deflection<br />

variation<br />

0 2 4 6 8 10 12 14 16 18 20 22<br />

displacement (mm)<br />

Figure 6.17: definition <strong>of</strong> residual displacements at initial unloading<br />

The theoretical <strong>and</strong> experimental deflection variations are listed in table 6.6. The<br />

theoretical calculation <strong>of</strong> the displacement variation is performed first, <strong>with</strong> the<br />

stochastic cracking theory implemented in the lower face. The displacement<br />

variation is also calculated <strong>with</strong> the assumption <strong>of</strong> linear elastic material behaviour<br />

for the lower face.<br />

182


Chapter 6: experimental work on s<strong>and</strong>wich panels<br />

Table 6.6: residual deflection first unloading, from 2/3 rd <strong>of</strong> the maximum load to 2/9 th <strong>of</strong> the<br />

maximum load<br />

deflection (mm)<br />

HC/S7 HC/S8 HC/S9 HC/S10 HC/S11 HC/S12<br />

unloading,<br />

experimental<br />

8.9 6.7 10 8.8 12 9.4<br />

unloading,<br />

stochastic cracking<br />

9.9 9.2 9.1 8.4 11 9.2<br />

error (%) +10 +27 -14 -4.5 -11 -2.2<br />

unloading,<br />

linear<br />

9.9 9.2 9.1 8.1 11 9.2<br />

error (%) +10 +27 -14 -8.0 -11 -2.2<br />

From table 6.6, it can be noticed that for all tested panels implementation <strong>of</strong> the<br />

stochastic cracking theory in the lower face has no advantages compared to a<br />

linear elastic calculation. Although the computing time is larger when the multiple<br />

cracking unloading behaviour is implemented (see paragraph 5.5), the discrepancy<br />

between the FEM calculated <strong>and</strong> the experimentally obtained deflection variation<br />

is not lower than when linear elastic face behaviour is used. It should be<br />

mentioned that this conclusion is true for the test cases studied here. If the<br />

geometry <strong>and</strong>/or load conditions are changed this might not be the case any more.<br />

6.6.5 theory versus experiments: repeated loading between 2/9 th <strong>and</strong><br />

2/3 rd <strong>of</strong> the maximum load<br />

force (N)<br />

∆εc repeat (strain 2) ∆εc repeat (strain 1)<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

-0.04<br />

0 0.04 0.08 0.12<br />

strain (%)<br />

force (N)<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

∆deflection<br />

0 2 4 6 8 10 12 14 16 18 20 22<br />

displacement (mm)<br />

Figure 6.18a: extra strains due to repeated loading Figure 6.18b: extra deflection due to<br />

repeated loading<br />

Figure 6.18a illustrates how the extra strain terms, due to repeated loading, are<br />

defined here. From figure 6.18a, it can be seen that, due to repeated loading,<br />

accumulation <strong>of</strong> strains is monitored in strain gauge 1 <strong>and</strong> strain gauge 2, which<br />

are both located at the lower face in tension. Almost no variation <strong>of</strong> strains is<br />

measured in strain gauge 3. The additional strain term at mid-span <strong>of</strong> the lower<br />

face is noted as ∆εc repeat (strain 1). ∆εc repeat (strain 2) is the extra strain term at 1/4 th<br />

183


Chapter 6: experimental work on s<strong>and</strong>wich panels<br />

<strong>of</strong> the length at the bottom <strong>of</strong> the lower face. These strain increments in the lower<br />

face lead to an increasing maximum deflection <strong>of</strong> the s<strong>and</strong>wich panel. The extra<br />

deflection term, due to repeated loading, is illustrated in figure 6.18b<br />

In Chapter 4 it has been mentioned that the main damage mechanism for 2Dr<strong>and</strong>omly<br />

reinforced IPC composite specimens under repeated loading is the<br />

matrix-fibre interface degradation, provided the maximum cycle stress is low<br />

enough (at or below 16.5MPa, see paragraph 4.10.6). Two constants, C1 <strong>and</strong> C2<br />

determine the speed <strong>of</strong> this degradation phenomenon. C1 is chosen to equal 1.0<br />

<strong>and</strong> C2 is 0.02, similar to the results from experiments in Chapter 4. However, it<br />

has been mentioned in paragraph 4.10.6 that another damage mechanism occurs at<br />

low maximum cycle stress: extra matrix cracking due to repeated cycling. It has<br />

been mentioned that this effect typically occurs in the first cycles, but is not<br />

implemented in the model in this work: since no quantitative evolution <strong>of</strong> the<br />

matrix cracking due to repeated cycling has been determined yet, only the matrixfibre<br />

interface degradation is considered here<br />

The implementation <strong>of</strong> the lower face degradation is performed as explained in<br />

paragraph 5.5 (macro repeat.mac).<br />

The theoretical predicted extra strain terms are listed together <strong>with</strong> the<br />

experimental extra strain terms in table 6.7 for strain gauge 1. These values are<br />

listed in table 6.8 for strain gauge 2.<br />

Table 6.7: extra strain term at lower face, 1/2 nd <strong>of</strong> the length, under 2/3 rd <strong>of</strong> the maximum load,<br />

repeated loading, experiment <strong>and</strong> theory<br />

repeated load,<br />

experimentally<br />

repeated load,<br />

predicted<br />

strain gauge 1 (%)<br />

HC/S7 HC/S8 HC/S9 HC/S10 HC/S11 HC/S12<br />

0.05 0.02 0.03 0.008 0.02 0.02<br />

0.04 0.008 0.04 0.007 0.03 0.01<br />

From table 6.7 (strain gauge 1, lower face at 1/2 nd <strong>of</strong> span), it is noticed that the<br />

predicted theoretical additional strain terms approximate the experimentally<br />

measured additional strain terms. The predicted value <strong>of</strong> the extra strain term due<br />

to repeated loading is sometimes higher <strong>and</strong> sometimes lower than the<br />

experimentally obtained value. From the finite element calculations, it can be<br />

found that the value <strong>of</strong> the tensile normal stress in the faces is about 6 to 8MPa at<br />

mid-span <strong>of</strong> the s<strong>and</strong>wich panels (where strain gauge 1 is located). These values<br />

will be compared later <strong>with</strong> the values at the location <strong>of</strong> strain gauge 2.<br />

184


Chapter 6: experimental work on s<strong>and</strong>wich panels<br />

Table 6.8: extra strain term at lower face, 1/4 th <strong>of</strong> the length, under 2/3 rd <strong>of</strong> the maximum load,<br />

repeated loading, experiment <strong>and</strong> theory<br />

strain gauge 2 (%)<br />

HC/S7 HC/S8 HC/S9 HC/S10 HC/S11 HC/S12<br />

repeated load,<br />

0.02 0.01 0.04 0.007 0.02 0.003<br />

experimentally<br />

repeated load,<br />

predicted<br />

0.008 0.004 0.01 0.004 0.008 0.003<br />

From table 6.8, it can be seen that the experimentally obtained extra strain terms,<br />

measured in strain gauge 2 (lower face, 1/4 th <strong>of</strong> length), due to repeated loading,<br />

are systematically higher than the theoretical predictions. According to the finite<br />

element calculations, the maximum normal stresses in the faces are 3 to 5MPa, at<br />

1/4 th <strong>of</strong> the span length. At 1/4 th <strong>of</strong> the span length, the maximum stress is thus still<br />

situated far below the ACK multiple cracking stress. It has been mentioned in<br />

Chapter 4 that matrix cracking due to repeated loading leads to extra loss <strong>of</strong><br />

composite stiffness, when the maximum cycle stress is situated in the theoretical<br />

ACK pre-cracking zone. This effect is not implemented in the calculation <strong>of</strong> the<br />

extra strain term.<br />

Table 6.9 shows the theoretical <strong>and</strong> experimental extra deflection terms from<br />

repeated loading (10 times) between 2/3 rd <strong>and</strong> 2/9 th <strong>of</strong> the maximum load.<br />

Table 6.9: extra deflection term under repeated loading up to 2/3 rd <strong>of</strong> the maximum load, 10<br />

loading cycles, experiment <strong>and</strong> theory<br />

deflections (mm)<br />

repeated load,<br />

experimentally<br />

repeated load,<br />

theoretically<br />

HC/S7 HC/S8 HC/S9 HC/S10 HC/S11 HC/S12<br />

2.1 1.2 2.4 0.9 3.4 1.9<br />

1.6 1.0 2.0 1.4 3.0 1.7<br />

The results in table 6.9 indicate that the methodology used here to predict extra<br />

deflections due to repeated loading provides - in all its simplicity – a proper<br />

estimation <strong>of</strong> extra deformations due to accumulated damage.<br />

6.6.6 theory versus experiments: loading to maximum load<br />

All s<strong>and</strong>wich panels are now loaded up to the maximum load, as defined in<br />

paragraph 6.4. Table 6.10 lists the experimental deflections <strong>and</strong> the theoretical<br />

predictions.<br />

185


Chapter 6: experimental work on s<strong>and</strong>wich panels<br />

The influence <strong>of</strong> the earlier applied 10 loading cycles - before the panels were<br />

loaded up to the maximum load - is introduced in the theoretical predictions here<br />

by adding the extra deflections terms, as given in table 6.9.<br />

Table 6.9: maximum deflection under maximum load, initial loading, experiment, theory <strong>and</strong><br />

theory <strong>with</strong> linear elastic behaviour.<br />

displacements (mm)<br />

HC/S7 HC/S8 HC/S9 HC/S10 HC/S11 HC/S12<br />

experiment 33.4 21.5 40.7 24.6 54.1 30.8<br />

theory, stochastic<br />

cracking<br />

37.7 28.1 41.9 28.6 49.7 33.9<br />

error +13 +31 +3 +16 -8 +10<br />

theory, linear<br />

elastic behaviour<br />

24.5 20.5 25.1 18.1 27.0 22.9<br />

error -27 -5 -39 -26 -50 -26<br />

It can be seen from table 6.9 that implementation <strong>of</strong> the stochastic cracking theory<br />

leads to fairly good prediction <strong>of</strong> the global behaviour <strong>of</strong> the s<strong>and</strong>wich panels: the<br />

deflection. Again the results from panel HC/S8 are not taken into account. If the<br />

stochastic cracking theory is used, the discrepancy between the experimental <strong>and</strong><br />

theoretical maximum deflections varies between –8% <strong>and</strong> +16%. If the<br />

assumption <strong>of</strong> linear elastic behaviour <strong>of</strong> the faces is used, this discrepancy varies<br />

from –5% to as high as -50%.<br />

Whereas the maximum deflections is systematically underestimated if FEM<br />

calculation is performed <strong>with</strong> assumption <strong>of</strong> linear elastic face behaviour, this is<br />

not the case when multiple cracking face behaviour is implemented in the FEM<br />

calculations. The maximum absolute value <strong>of</strong> the discrepancy between prediction<br />

<strong>and</strong> experiment decreases from 50% to 16%, when linear elastic face behaviour is<br />

replaced by multiple cracking behaviour.<br />

6.7 Short-span panel: geometry <strong>and</strong> test set-up<br />

6.7.1 geometry <strong>and</strong> test set-up<br />

Until now, focus has been put on core thickness <strong>and</strong> face thickness variations. In<br />

this paragraph, the length <strong>of</strong> the s<strong>and</strong>wich panel is changed from 2000mm to<br />

850mm. If the s<strong>and</strong>wich length is shortened, the assumption <strong>of</strong> constant normal<br />

stresses along the face thickness will become less appropriate, as was already<br />

discussed in Chapter 5. In this paragraph the strain variation across the thickness<br />

<strong>of</strong> the lower face at mid-span is studied for a “relatively short panel”.<br />

186


Chapter 6: experimental work on s<strong>and</strong>wich panels<br />

In order to obtain some information on the strain variation along the thickness <strong>of</strong><br />

the faces, strain gauges are now laminated inside the face layers.<br />

A s<strong>and</strong>wich panels is prepared. Strain gauges are laminated into the lower face at<br />

mid-span <strong>of</strong> the s<strong>and</strong>wich. The panel length is 850mm instead <strong>of</strong> 2000mm. The<br />

core thickness is 60mm. The number <strong>of</strong> face layers is 3. Table 6.10 summarises<br />

geometrical properties <strong>of</strong> the tested panel.<br />

Table 6.10: properties s<strong>and</strong>wich panel <strong>with</strong> modification on length<br />

length core material core thickness # face layers<br />

(mm)<br />

(mm)<br />

HC/S3 850 PUR 35kg/m³ 60 3<br />

The short-span panel is subjected to four-point bending. The distance between the<br />

lower supports is 750mm. Between the upper supports, there is a separation <strong>of</strong><br />

150mm. The total force on the s<strong>and</strong>wich is measured <strong>with</strong> a load cell. A LVDT is<br />

used to measure the maximum displacement <strong>of</strong> the panels at mid-span. Several<br />

strain gauges are located in <strong>and</strong> on the panel. Figure 6.19 illustrates the position <strong>of</strong><br />

the strain gauges<br />

strain gauge 1<br />

strain gauge 2<br />

strain gauge 3<br />

150mm<br />

850mm<br />

strain gauge 4<br />

strain gauge 5<br />

Figure 6.19: test set-up four-point bending test on small-scale s<strong>and</strong>wich panels <strong>and</strong> placement <strong>of</strong><br />

strain gauges.<br />

Three strain gauges are attached to the surface <strong>of</strong> the s<strong>and</strong>wich panel:<br />

- The first strain gauge is attached to the upper (compressive) side at 1/4 th<br />

<strong>of</strong> the length <strong>of</strong> the s<strong>and</strong>wich.<br />

- The second strain gauge is fastened at the lower (tensile) side <strong>of</strong> the panel,<br />

also at 1/4 th <strong>of</strong> the length <strong>of</strong> the panel.<br />

- Strain gauge number 3 is located at the middle <strong>of</strong> the length <strong>of</strong> the<br />

s<strong>and</strong>wich at the lower (tensile) side.<br />

Two strain gauges were put into the lower face during lamination:<br />

187


layer.<br />

Chapter 6: experimental work on s<strong>and</strong>wich panels<br />

- Strain gauge 4 is placed in the matrix, between the core <strong>and</strong> the first fibre<br />

- Strain gauge 5 is placed between the first <strong>and</strong> the second fibre layer.<br />

The test set-up is illustrated in figure 6.20a. Figure 6.20b shows in detail how the<br />

forces are transferred from the supports to the s<strong>and</strong>wich panel.<br />

Figure 6.20a: test set-up short span panels Figure 6.20b: test set-up short-span panels,<br />

detail upper support<br />

The panel is subjected to four-point bending in two steps. The panel is first loaded<br />

up to 800N maximum <strong>and</strong> then unloaded again. This loading-unloading is<br />

displacement controlled at a rate <strong>of</strong> 2mm per minute. Then, the panel is loaded up<br />

to 1600N. As was mentioned in paragraph 6.4, the core material will undergo no<br />

yield or fracture in shear at these stress levels.<br />

6.7.2 reliability <strong>of</strong> the strains, measured by strain gauges laminated<br />

into a face layer<br />

One <strong>of</strong> the main concerns, when using internal strain gauges, is the reliability <strong>of</strong><br />

the measurements. The strain gauge may slip inside the material if the adhesion<br />

between the strain gauge <strong>and</strong> the material is not satisfactory, leading to<br />

underestimation <strong>of</strong> the internal strains. On the other h<strong>and</strong>, the presence <strong>of</strong> liquid<br />

<strong>and</strong>/or charged particles may lead to secondary electric circuits <strong>and</strong> can influence<br />

the measurement <strong>of</strong> the electrical resistance <strong>of</strong> the strain gauge. Therefore, the<br />

reliability <strong>of</strong> the strain measurements from strain gauges, which are laminated into<br />

IPC matrix, is to be tested first.<br />

A laminate is made <strong>with</strong> two layers <strong>of</strong> 2D-r<strong>and</strong>omly oriented reinforcement <strong>and</strong><br />

embedded strain gauges in between the two layers. Before the strain gauges were<br />

embedded, a thin electrical insulating polyurethane coating was applied on all<br />

strain gauges <strong>and</strong> their connecting wires. Specimens are cut from this plate for<br />

simple tensile testing. During simple tensile testing, a load cell measures the load,<br />

188


Chapter 6: experimental work on s<strong>and</strong>wich panels<br />

while the strains are measured by an internal strain gauge <strong>and</strong> an extensometer on<br />

the specimen.<br />

Figure 6.21 shows a stress-strain curve <strong>of</strong> a specimen, obtained by measurement <strong>of</strong><br />

the strain <strong>with</strong> the extensometer in grey <strong>and</strong> <strong>with</strong> the internal strain gauge in black.<br />

The other specimens give similar results.<br />

stress (Mpa)<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

extensometer<br />

embedded strain gauge<br />

0 0.25 0.5 0.75 1 1.25 1.5<br />

strain (%)<br />

Figure 6.21: stress-strain curve <strong>with</strong> embedded strain gauge <strong>and</strong> extensometer<br />

One can notice that there is small discrepancy between the two curves, due to<br />

small misalignment <strong>of</strong> the strain gauge or to the fact that the gauge length <strong>of</strong> the<br />

two strain measurements is different. The strain gauge measures the strain along<br />

10mm length, the extensometer along 50mm.<br />

Use <strong>of</strong> an internal strain gauge gives accurate knowledge <strong>of</strong> the strains inside an<br />

IPC composite specimen. Other methods could have been used, such as the<br />

insertion <strong>of</strong> optical fibres, but strain gauges have the advantage that they are easy<br />

to use <strong>and</strong> quite robust during h<strong>and</strong>ling. If the matrix impregnates the next layer <strong>of</strong><br />

glass fibres by rolling <strong>and</strong> pressing, the strain gauge stays intact.<br />

6.8 Short-span panels: results<br />

Figure 6.22 shows the evolution <strong>of</strong> the deflections versus applied force during<br />

testing for s<strong>and</strong>wich HC/S3. Loading up to 1/3 rd <strong>of</strong> the maximum load is printed in<br />

black, loading up to 2/3 rd <strong>of</strong> the maximum load is printed in grey.<br />

If the shape <strong>of</strong> the deflection-force curve from this panel is compared to those<br />

curves, obtained on 2m span panels, the non-linear deflection-force relationship<br />

during loading was more obvious in the curves <strong>of</strong> the 2m span panels. The reason<br />

can be found in the fact that the moment to shear force ratio decreases when the<br />

span is decreased. For short span panels, the moment is thus still not very high,<br />

when the core fails. When the moment is low, the normal face stresses are also<br />

189


Chapter 6: experimental work on s<strong>and</strong>wich panels<br />

very low. Multiple cracking does occur in the faces <strong>of</strong> the short-span panels, but<br />

only moderate, since normal stresses are relatively small during the whole test.<br />

force (N)<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

0 2 4 6 8<br />

displacement (mm)<br />

Figure 6.22: deflection versus force evolution s<strong>and</strong>wich HC/S3<br />

6.8.1 loading up to 1/3 rd <strong>of</strong> the maximum load: results<br />

The aim <strong>of</strong> the testing program on the short-span panel is to visualise the evolution<br />

<strong>of</strong> normal strains across the face thickness <strong>of</strong> a IPC composite face. As has been<br />

mentioned in Chapter 5, it is expected that there is a large strain variation across<br />

the face thickness for short-span panels. Figure 6.23 shows the evolution <strong>of</strong> the<br />

strains, measured in the strain gauges, located in the lower face at mid-span. The<br />

location <strong>of</strong> the strain gauges is illustrated in figure 6.19.<br />

From figure 6.23, one can notice that the strain variation across the face thickness<br />

at mid-span is indeed high for panel HC/S3. The measured strain is very low at the<br />

core-face interface for panel HC/S3. In Chapter 5 (see paragraph 5.4.6) it was<br />

mentioned that the stress evolution along a face thickness is higher if the panel is<br />

short <strong>and</strong> the face thickness is high, which is the case for panel HC/S3. Therefore,<br />

the variation <strong>of</strong> the stresses is high across the face thickness <strong>of</strong> panel HC/S3.<br />

Consequently, the measured strain variations are higher in panel HC/S3.<br />

force (N)<br />

strain gauge 4<br />

layer1/layer2 lower face<br />

mid-span<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

strain gauge 5<br />

core/layer1 lower face<br />

mid-span<br />

strain gauge 3<br />

bottom lower face<br />

mid-span<br />

0 0.01 0.02 0.03 0.04<br />

strain (%)<br />

Figure 6.23 strain-force evolution <strong>of</strong> strain gauges HC/S3, at 1/3 rd <strong>of</strong> the maximum load<br />

190


Chapter 6: experimental work on s<strong>and</strong>wich panels<br />

strain gauge 1<br />

top upper face<br />

at 1/4 th <strong>of</strong> span<br />

force (N)<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

strain gauge 2<br />

bottom lower face<br />

at 1/4 th <strong>of</strong> span<br />

strain gauge 3<br />

bottom lower face<br />

mid span<br />

-0.02 -0.01 0 0.01 0.02 0.03 0.04<br />

strain (%)<br />

Figure 6.24 strain-force evolution <strong>of</strong> strain gauges, attached to the surface s<strong>and</strong>wich panel<br />

HC/S3, 1/3 rd <strong>of</strong> the maximum load<br />

Figure 6.24 shows the strain at the top <strong>of</strong> the upper face at 1/4 th <strong>of</strong> the span length<br />

in black <strong>and</strong> <strong>with</strong> negative strain values (strain gauge 1). The grey line represents<br />

the strain measured at the bottom <strong>of</strong> the lower face at 1/4 th <strong>of</strong> the span (strain<br />

gauge 2). The black line at the right represents the strain evolution at the bottom <strong>of</strong><br />

the lower face at mid-span (strain gauge 3).<br />

6.8.2 loading up to 2/3 rd <strong>of</strong> the maximum load: results<br />

Figure 6.25 <strong>and</strong> 6.26 show the strain evolutions as measured on panel HC/S3,<br />

when the panel is loaded up to 2/3 rd <strong>of</strong> the maximum load.<br />

strain gauge 4<br />

layer1/layer2 lower face<br />

mid span<br />

strain gauge 5<br />

core/layer1 lower face<br />

2000<br />

mid span<br />

force (N)<br />

1500<br />

1000<br />

500<br />

0<br />

strain gauge 3<br />

bottom lower face<br />

mid span<br />

0 0.05 0.1 0.15 0.2<br />

strain (%)<br />

Figure 6.25 strain-force evolution <strong>of</strong> strain gauges; HC/S3, at 2/3 rd <strong>of</strong> the maximum load<br />

191


Chapter 6: experimental work on s<strong>and</strong>wich panels<br />

strain gauge 1<br />

top upper face<br />

at 1/4<br />

2000<br />

th <strong>of</strong> span strain gauge 2<br />

bottom lower face<br />

at 1/4 th <strong>of</strong> span<br />

force (N)<br />

1500<br />

1000<br />

500<br />

0<br />

-0.05 0 0.05 0.1 0.15 0.2<br />

strain (%)<br />

strain gauge 3<br />

bottom lower face<br />

mid span<br />

Figure 6.26 strain-force evolution <strong>of</strong> strain gauges attached to surface s<strong>and</strong>wich<br />

HC/S3, at 2/3 rd <strong>of</strong> the maximum load<br />

Where the strain measured in strain gauge 3 (at bottom <strong>of</strong> the lower face, midspan)<br />

is relatively high <strong>and</strong> multiple cracking already occurred, the strain at the<br />

core-face interface is still low, so probably only minor matrix cracking occurs at<br />

the interface. The normal strain is not constant across the thickness in the lower<br />

face at mid-span. At the bottom <strong>of</strong> the face, multiple cracking already occurred as<br />

can be seen from the distinct bending point in this force-strain curve. Close to the<br />

core-face interface the strains are still low: multiple cracking does not only<br />

advance from the mid-span to the supports, but also from the bottom <strong>of</strong> the lower<br />

face to the top <strong>of</strong> the lower face, if the external force is increased.<br />

6.9 Short-span panels: theoretical FEM predictions versus<br />

experimental results<br />

6.9.1 properties <strong>of</strong> the faces <strong>of</strong> the tested panels<br />

After panel HC/S3 is tested, several specimens are cut from both faces, like has<br />

been done for the 2m span panels. The average thickness <strong>of</strong> the upper faces is<br />

4.2mm. The average thickness <strong>of</strong> the lower face is 3.9mm. Four specimens from<br />

the upper face are subjected to simple tensile testing to obtain parameters m, σR<br />

<strong>and</strong> τ0. The average value <strong>of</strong> m is 4.2, the average value <strong>of</strong> σR is 8.9MPa. The<br />

average value <strong>of</strong> τ0 is 0.65MPa.The average fibre volume fraction is 8% in the<br />

upper face <strong>and</strong> is 9% in the lower face. These parameters are used in the<br />

theoretical prediction <strong>of</strong> the stress-strain behaviour <strong>of</strong> the face in tension.<br />

192


Chapter 6: experimental work on s<strong>and</strong>wich panels<br />

6.9.2 theory versus experiments: initial loading to 1/3 rd <strong>of</strong> the failure<br />

load<br />

Table 6.11 lists the theoretical (FEM) <strong>and</strong> the experimental values <strong>of</strong> the strains,<br />

when the applied load is 1/3 rd <strong>of</strong> the maximum load.<br />

Table 6.11: strains measured by strain gauges, 1/3 rd <strong>of</strong> maximum load<br />

strain 1 strain 2 strain 3 strain 4 strain 5<br />

(%) (%) (%) (%) (%)<br />

experiment, HC/S3 -0.008 0.01 0.03 0.001 0.01<br />

FEM theory, HC/S3 -0.007 0.007 0.02 0.0009 0.005<br />

If panel HC/S3 is loaded up to 1/3 rd <strong>of</strong> the maximum load, the measured maximum<br />

deflection is 3.6mm <strong>and</strong> the theoretical prediction is 4.0mm. As can be seen from<br />

table 6.11, the theoretical predictions <strong>of</strong> the strains are systematically lower than<br />

the experimentally measured strains. However, stresses are still relatively low in<br />

the faces. The maximum stress in the faces is about 3MPa. Although the (FEM)<br />

calculated strains are lower than the experimental strains, the predicted<br />

displacement is higher than the experimental. Since HC/S3 is a short-span panel,<br />

the total displacement is mainly determined by the shear deformation <strong>of</strong> the core.<br />

The extra displacement due to bending is only a minor term, compared to the shear<br />

deformation. If the core material properties are not determined exactly, this will<br />

have a larger influence on an accurate prediction <strong>of</strong> the deflection than inaccuracy<br />

<strong>of</strong> the face material properties.<br />

It has been mentioned earlier that the measured strain at the core/face interface<br />

(strain gauge 4 in figure 6.23) is very low, almost zero. This effect is also<br />

predicted theoretically. The high variation <strong>of</strong> normal strains across the thickness <strong>of</strong><br />

the faces is also found in the theoretical predictions <strong>of</strong> strains.<br />

6.9.3 theory versus experiments: loading up to 2/3 rd <strong>of</strong> the maximum<br />

load<br />

When panel HC/S3 is now further loaded up to 2/3 rd <strong>of</strong> the maximum load, the<br />

measured deflection is 7.3mm, the theoretical prediction is 8.1mm.<br />

Table 6.19: strains measured at strain gauges, 2/3 rd <strong>of</strong> maximum load, HC/S3<br />

strain 1 strain 2 strain 3 strain 4 strain 5<br />

(%) (%) (%) (%) (%)<br />

experiment, -0.016 0.021 0.17 0.015 0.049<br />

theory -0.013 0.016 0.052 0.009 0.025<br />

Table 6.19 lists the theoretical <strong>and</strong> experimental strains at the strain gauge<br />

locations. The theoretical strains underestimate the experimentally obtained values<br />

systematically. The theoretical displacement slightly overestimates the<br />

193


Chapter 6: experimental work on s<strong>and</strong>wich panels<br />

experimental obtained values. The conclusions on table 6.19 are similar to the<br />

conclusions on table 6.18.<br />

6.10 Conclusions<br />

Typical spans <strong>of</strong> s<strong>and</strong>wich panels for building purposes in Europe vary from 2m to<br />

4m (Aerts, 1998). In this chapter, panels are tested <strong>with</strong> a length <strong>of</strong> 2m. The core<br />

<strong>and</strong> face thickness is varied for these panels. The s<strong>and</strong>wich models discussed in<br />

Chapter 5 <strong>and</strong> the IPC composite material behaviour, derived in Chapter 2 to<br />

Chapter 4, are inserted in FEM calculations to obtain theoretical predictions <strong>of</strong> the<br />

behaviour <strong>of</strong> the tested panels. Theoretical FEM calculations <strong>and</strong> experimental<br />

results are compared. Conclusions from four-point bending on panels <strong>with</strong> 2m<br />

length are:<br />

- When the panels are subjected to monotonic loading in four-point bending, there<br />

is discrepancy (up to 15%) between the theoretical predictions <strong>of</strong> the maximum<br />

deflections <strong>and</strong> the experimental values. However, this discrepancy does not lead<br />

to a systematic underestimation or overestimation <strong>of</strong> the maximum deflection <strong>of</strong><br />

the panels. This is an indication that the discrepancy is probably rather due to<br />

difficult accurate determination <strong>of</strong> material parameters than to inaccuracy <strong>of</strong> the<br />

model, which has been used.<br />

- When the behaviour <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces is calculated,<br />

assuming the faces behave linear elastically, this leads to a systematic<br />

underestimation <strong>of</strong> the maximum deflection. From comparison <strong>of</strong> FEM<br />

calculations <strong>and</strong> experimental results it is noticed that this discrepancy can be<br />

rather high (up to 50%). Therefore, it is a necessity that multiple cracking<br />

constitutive equations, (derived in Chapter 2) are inserted in the FEM calculations.<br />

- Concerning the unloading behaviour <strong>of</strong> the panel, fair accuracy is obtained on the<br />

prediction <strong>of</strong> the unloading deflection variation by implementation <strong>of</strong> the<br />

stochastic cracking theory as presented in Chapter 3. However, for all tested<br />

panels it has been noticed that the prediction <strong>of</strong> the unloading deflection variation<br />

by implementation <strong>of</strong> linear elastic face behaviour gives very similar results to<br />

when the stochastic cracking theory has been used. Further study should reveal if<br />

this is still true when the panel length or/<strong>and</strong> the applied loading cycle amplitude is<br />

increased.<br />

- If the behaviour <strong>of</strong> the s<strong>and</strong>wich panels under repeated loading is to be predicted,<br />

linear elastic behaviour <strong>of</strong> the faces cannot be adopted, since no accumulated<br />

deformations are predicted this way, although this effect has been observed<br />

experimentally. The stochastic cracking based model in Chapter 4 has been<br />

194


Chapter 6: experimental work on s<strong>and</strong>wich panels<br />

implemented as proposed in Chapter 5 to predict the behaviour <strong>of</strong> s<strong>and</strong>wich panels<br />

<strong>with</strong> IPC composite faces under repeated loading. When stochastic matrix<br />

cracking is used to predict the behaviour <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite<br />

faces under repeated loading, a fair (but not very accurate) estimation <strong>of</strong> additional<br />

strains <strong>and</strong> deflection terms is obtained in this chapter for all panels.<br />

The deflection <strong>of</strong> short-span panels is determined by the core properties, rather<br />

than by the face properties. However, s<strong>and</strong>wich panels, which are smaller than 2m<br />

have little practical use. One main conclusion is however found from comparison<br />

<strong>of</strong> FEM calculations <strong>with</strong> experiments: the high strain variation across the face<br />

thickness at mid-span, which is measured experimentally, is also predicted<br />

theoretically by the FEM model<br />

6.11 References<br />

E. Aerts, S<strong>and</strong>wichpanelen als dakelement in de bouw, Master thesis VUBhogeschool<br />

Antwerpen, 1998-1999<br />

ANSYS manuals, Release 5.5, 1998<br />

ASTM st<strong>and</strong>ard D3163<br />

ASTM st<strong>and</strong>ard D3164<br />

ECCS publication, Preliminary European Recommendations for S<strong>and</strong>wich<br />

<strong>Panels</strong>, part I, <strong>Design</strong>, Technical Committee 7, Working Group 7.4, 1991<br />

L. J. Gibson <strong>and</strong> M. F. Ashby, Cellular solids, Structure & Properties,<br />

Permagon Press, 1988<br />

195


Chapter 7<br />

7.1 Introduction<br />

<strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC<br />

composite faces<br />

In this chapter, several test cases illustrate the design <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> Eglass<br />

fibre reinforced faces. <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

is compared <strong>with</strong> design <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> steel faces.<br />

Three types <strong>of</strong> s<strong>and</strong>wich panels are studied in this chapter. These types are:<br />

1. simply supported flat wall panels<br />

2. simply supported flat ro<strong>of</strong> panels<br />

3. flat ro<strong>of</strong> panels, <strong>with</strong> an intermediate support<br />

The loads, which are considered to act on the s<strong>and</strong>wich panels, are own weight,<br />

snow load, wind load <strong>and</strong> temperature load. Each <strong>of</strong> the three studied s<strong>and</strong>wich<br />

panel types differs slightly from the two others.<br />

- The wall panels are subjected to wind <strong>and</strong> temperature load.<br />

- The ro<strong>of</strong> panels are also subjected to wind <strong>and</strong> temperature load like wall<br />

panels, but in addition snow load should be considered.<br />

- When an intermediate support is added, the panel becomes hyperstatic.<br />

No extra service loads, like extra ro<strong>of</strong> covering, are considered since these loads<br />

do not occur inevitably, in contrast <strong>with</strong> wind, snow or temperature loading.<br />

For all three panel types, the panel length is varied from 2m to 4m <strong>with</strong> steps <strong>of</strong><br />

0.5m. The allowed core thicknesses are 40mm, 60mm, 80mm <strong>and</strong> 100mm. These<br />

core thicknesses are chosen because they are commonly used in s<strong>and</strong>wich panels<br />

<strong>with</strong> steel faces in Belgium (see technical reference sheets, paragraph 7.11). The<br />

IPC composite face layers contain UD-reinforcement. The thickness <strong>of</strong> each extra<br />

face layer is about 0.75mm.<br />

197


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

For all case studies, the least-weight design solution is to be found, still respecting<br />

all defined limits <strong>and</strong> conditions.<br />

7.2 <strong>Design</strong> methodology<br />

The loads considered here are own weight, snow, wind <strong>and</strong> temperature effects. It<br />

has been mentioned in Chapter 1 that the lifetime <strong>of</strong> the construction elements in<br />

this work is 50 years. The magnitude <strong>of</strong> characteristic snow/wind/temperature load<br />

<strong>with</strong> a return period <strong>of</strong> 50 years is found in Eurocode 1 <strong>and</strong>/or National st<strong>and</strong>ards.<br />

A characteristic load is generally defined as the load, which will be exceeded <strong>with</strong><br />

a probability <strong>of</strong> x% during a chosen lifetime. Usually this x% is chosen to equal<br />

5%. For all case studies in this chapter, it is considered that the s<strong>and</strong>wich panels<br />

are used in buildings, situated in Belgium.<br />

7.2.1 wind load<br />

NBN B 03-002-1 is used to determine the basic wind pressure. Wind effects are<br />

most pronounced at the Belgian coast. It is thus assumed here for all case studies<br />

that the building is situated at the coast. The maximum height <strong>of</strong> the building, on<br />

which the s<strong>and</strong>wich panels are placed, is chosen to be 30m. According to NBN B<br />

03-002-1, the basic wind pressure, qb, is thus 1258 N/m².<br />

In this work, the translation <strong>of</strong> wind effects to pressure loads on the s<strong>and</strong>wich<br />

panels only considers loading perpendicular to the panels: the influence <strong>of</strong> friction<br />

on wall claddings under wind load is neglected. Wind pressure load on a surface is<br />

formulated by equation (7.1). This equation is used to define the external (outside<br />

a building) <strong>and</strong> the internal (inside a building) wind pressure.<br />

w = cpcd<br />

qk<br />

(7.1)<br />

<strong>with</strong> qk = characteristic wind pressure<br />

cd = dynamic coefficient<br />

cp = pressure coefficient (cpe = external wind pressure coefficient, cpi =<br />

internal wind pressure coefficient)<br />

When the pr<strong>of</strong>ile <strong>of</strong> the surroundings is relatively flat, the loads can act on the<br />

building during the whole lifetime <strong>and</strong> the wind may come from any direction.<br />

The characteristic value <strong>of</strong> the wind pressure qk then equals the basic value <strong>of</strong> the<br />

wind pressure qb. If the framework <strong>of</strong> the building can be classified as rigid, the<br />

value <strong>of</strong> the dynamic coefficient cd equals 1, which will be assumed here. In case<br />

the building contains windows <strong>and</strong> internal walls, the extreme values <strong>of</strong> the<br />

internal pressure coefficient cpi are –0.3 <strong>and</strong> +0.3. The external pressure coefficient<br />

cpe varies for the different studied s<strong>and</strong>wich types <strong>and</strong> will be determined later.<br />

198


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

7.2.2 snow load<br />

The snow load pressure, s, on a flat ro<strong>of</strong> is formulated as:<br />

s = µ C C s<br />

(7.2)<br />

i e t k<br />

The characteristic snow pressure load on the ground, sk, is 0.5kN/m² in Belgium.<br />

The snow shape coefficient µi is 0.8 for flat ro<strong>of</strong>s. Ce <strong>and</strong> Ct are the exposure<br />

coefficient <strong>and</strong> thermal coefficient respectively <strong>and</strong> are usually set equal to one.<br />

For a simply supported ro<strong>of</strong> panel <strong>with</strong> two supports, the snow load is considered<br />

to cover the whole length <strong>of</strong> the ro<strong>of</strong> uniformly. No snow load is considered on<br />

wall panels.<br />

7.2.3 temperature load<br />

Temperature variations have two effects on the calculation <strong>of</strong> s<strong>and</strong>wich panels<br />

<strong>with</strong> IPC faces. First <strong>of</strong> all, internal stresses are created in the IPC faces, which<br />

means the stress-strain curve <strong>of</strong> IPC composites under monotonic tensile loading<br />

has to be modified: if the matrix undergoes high tensile stresses due to thermal<br />

effects, less external mechanical load is needed to introduce matrix multiple<br />

cracking. A quantitative formulation <strong>of</strong> this effect is derived in paragraph 7.3.<br />

Secondly, if a temperature gradient exists between the inner <strong>and</strong> outer face <strong>of</strong> the<br />

s<strong>and</strong>wich panel, the panel will deform. Maps <strong>of</strong> the maximum <strong>and</strong> minimum<br />

isotherms in Belgium are found in ENV 1991-2-5:1997, p31. At the Belgian coast,<br />

the minimum <strong>and</strong> maximum outside service life temperatures are –18°C <strong>and</strong><br />

+34°C respectively. The inside temperature is kept constant at 25°C for all<br />

calculations.<br />

About thermal effects, the Preliminary European Recommendations for S<strong>and</strong>wich<br />

<strong>Panels</strong> mention that the influence <strong>of</strong> thermal effects should always be taken into<br />

account in elastic analysis. They should only be introduced in plastic analysis if<br />

the inclusion <strong>of</strong> thermal effects seems to be relevant. Thermal effects will thus be<br />

included in this chapter, since it is not verified yet that these effects are irrelevant.<br />

7.2.4 own weight<br />

Inserting the density <strong>of</strong> all materials in ANSYS <strong>and</strong> defining the direction <strong>of</strong> the<br />

gravity for all calculated panels introduce the own weight <strong>of</strong> the panel. The own<br />

weight <strong>of</strong> IPC composite faces is about 2000kg/m³ (see table 2.2). The own weight<br />

<strong>of</strong> the considered PUR core material is 40kg/m³ (commonly used in panels <strong>with</strong><br />

steel faces, see technical sheets in references, paragraph 7.11).<br />

7.2.5 combination <strong>of</strong> loads<br />

Within the chosen lifetime <strong>of</strong> a construction (element) it is very unlikely that the<br />

characteristic snow load, characteristic wind load <strong>and</strong> characteristic temperature<br />

load occur at the same moment in time. Therefore, combination coefficients are<br />

used to include this probability. The total load on the structure, used for design<br />

199


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

calculations, is the sum <strong>of</strong> all characteristic loads (own<br />

weight/wind/snow/temperature), after these loads are multiplied <strong>with</strong> the<br />

combination coefficients (<strong>and</strong> possibly <strong>with</strong> safety coefficient, see later).<br />

According to the value <strong>of</strong> these combination coefficients, several load<br />

combinations are considered. Figure 7.1 illustrates these load combinations.<br />

The design methodology used on s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces is<br />

based on the methodology in Eurocode 1, the Preliminary European<br />

Recommendations for S<strong>and</strong>wich <strong>Panels</strong> (ECCS, 1991) <strong>and</strong> the Updated European<br />

Recommendations for S<strong>and</strong>wich <strong>Panels</strong> (Berner et al., 2000). All these documents<br />

consider two types <strong>of</strong> limit states in design: the ultimate limit state <strong>and</strong> the<br />

serviceability limit state.<br />

ULS<br />

load<br />

CLC<br />

FLC<br />

ULS = ultimate limit state<br />

CLC = characteristic load combination<br />

FLC = frequent load combination<br />

Figure 7.1: variation <strong>of</strong> load on a s<strong>and</strong>wich panel <strong>with</strong> time<br />

7.2.5.1 ultimate limit state (ULS)<br />

The ultimate limit state is defined in Eurocode 1 as the (loaded) state in which the<br />

construction (element) reaches a maximum bearing capacity. One should thus<br />

verify that:<br />

S ≤ R<br />

(7.3)<br />

where:<br />

d<br />

d<br />

Sd = the design value <strong>of</strong> the action or the combination <strong>of</strong> actions working<br />

on a construction<br />

Rd = the design value <strong>of</strong> the resistance <strong>of</strong> the construction (element)<br />

The combination <strong>of</strong> actions working on a construction (element) is defined for<br />

several situations: service lifetime loading, transition loading, accidental loading<br />

200<br />

time


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

<strong>and</strong> seismic loading. In this work only the service lifetime loading will be studied<br />

in detail.<br />

When service lifetime loading is considered in the ultimate limit state, the<br />

combination <strong>of</strong> loads is defined equally in Eurocode 1 <strong>and</strong> in the Updated<br />

Recommendations for S<strong>and</strong>wich <strong>Panels</strong> from Berner et al. (2000):<br />

Sd = kGk<br />

+ γ Q1Qk1<br />

+ γ Qiψ<br />

0iQki<br />

(7.4)<br />

where:<br />

γ ∑<br />

i><br />

1<br />

Gi = own weight <strong>of</strong> the construction (element)<br />

Qk1 = characteristic value <strong>of</strong> the dominant variable load<br />

Qki = characteristic value <strong>of</strong> the other variable load(s)<br />

γi = safety factor on the own weight<br />

γQ1 = safety factor on the dominant variable load<br />

γQi = safety factor on the other variable load(s)<br />

ψ0i = combination coefficient<br />

Failure <strong>of</strong> the panels is to be avoided under the load combination <strong>of</strong> equation (7.4).<br />

If a transversely loaded s<strong>and</strong>wich panel is studied, the most common failure types,<br />

which will be discussed here, are:<br />

- face fracture in tension<br />

- face fracture in compression<br />

- core fracture in shear<br />

- face wrinkling (in compression)<br />

- core-face interface failure<br />

In Chapter 6 it has been verified that the strength <strong>of</strong> the core-face interface <strong>of</strong> the<br />

studied panels is usually higher than the strength <strong>of</strong> the core material. Therefore,<br />

core-face interface failure is not further discussed as possible failure mechanism.<br />

The values <strong>of</strong> the safety factors γi <strong>and</strong> γQi are defined in Eurocode 1 <strong>and</strong> in the<br />

Updated Recommendations for S<strong>and</strong>wich <strong>Panels</strong> from Berner et al. (2000) <strong>and</strong> are<br />

listed in table 7.1.<br />

Table 7.1: safety factors<br />

ultimate limit state serviceability limit state<br />

permanent actions 1.35 1.0<br />

variable actions 1.5 1.0<br />

creep effects 1.0 1.0<br />

Although safety factors γi <strong>and</strong> γQi are totally equal in Eurocode 1 <strong>and</strong> in the<br />

Updated Recommendations for S<strong>and</strong>wich <strong>Panels</strong> from Berner et al. (2000), this is<br />

not the case for the combination coefficients. Table 7.2 lists the combination<br />

201


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

coefficients ψ0i as defined in Eurocode 1 <strong>and</strong> in the Updated Recommendations for<br />

S<strong>and</strong>wich <strong>Panels</strong> from Berner et al. (2000).<br />

Table 7.2: combination coefficients ψ0i<br />

Eurocode 1 Recommendations<br />

wind 0.6 0.7<br />

snow 0.6 0.5<br />

temperature 0.6 0.5<br />

As can be seen from table 7.2 the values <strong>of</strong> ψ0i, which are used in equation (7.4) to<br />

calculate the ultimate limit state during service conditions, are only slightly<br />

different in Eurocode 1 <strong>and</strong> in the Updated Recommendations for S<strong>and</strong>wich<br />

<strong>Panels</strong> from Berner et al. (2000). In this work, the combination coefficients <strong>of</strong> the<br />

Updated Recommendations for S<strong>and</strong>wich <strong>Panels</strong> from Berner et al. (2000) will be<br />

used. The reason is explained in paragraph 7.2.5.2.<br />

In figure 7.1 the ULS loading combination is printed together <strong>with</strong> the “real”<br />

evolution <strong>of</strong> the load on the construction (element) <strong>with</strong> time. The ULS is defined<br />

such that there is a very limited probability that it will be reached during the<br />

service lifetime <strong>of</strong> the construction (element).<br />

7.2.5.2 serviceability limit state (SLS)<br />

The serviceability limit state is defined in Eurocode 1 as the state in which the<br />

construction (element) does not satisfy any more certain service requirements,<br />

defined for this construction (element). Examples <strong>of</strong> service requirements are<br />

limitation <strong>of</strong> deflections, limitation <strong>of</strong> crack opening, etc. Several load<br />

combinations are defined <strong>with</strong>in the serviceability limit state, according to<br />

Eurocode 1 <strong>and</strong> the Updated Recommendations for S<strong>and</strong>wich <strong>Panels</strong> from Berner<br />

et al. (2000): the frequent load combination (FLC), the characteristic load<br />

combination (CLC) <strong>and</strong> the quasi-permanent load combination (QLC).<br />

7.2.5.3 serviceability limit state: frequent load combination (FLC)<br />

Equation (7.5) formulates the frequent combination (FLC) <strong>of</strong> loads as defined in<br />

Eurocode 1:<br />

Sd ∑Gkj + ψ 11Qk1<br />

+ ∑ψ<br />

2iQki<br />

(7.5)<br />

=<br />

j≥ 1 i><br />

1<br />

In equation (7.5) the definitions <strong>of</strong> Sd, Gkj, Qk1 <strong>and</strong> Qki are equal to the ones used<br />

in the expression <strong>of</strong> the ultimate limit state (equation (7.4)). The values <strong>of</strong> the<br />

combination coefficients according to Eurocode 1 are listed in table 7.3.<br />

202


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

Table 7.3: combination coefficients from Eurocode 1<br />

snow wind temperature<br />

ψ1 0.2 0.5 0.5<br />

ψ2 0 0 0<br />

Equation (7.6) gives the FLC as defined by Berner et al. (2000) in the Updated<br />

Recommendations for S<strong>and</strong>wich <strong>Panels</strong>:<br />

S d ∑G kj + ψ 11Qk1<br />

+ ∑ψ<br />

0iψ<br />

1iQki<br />

(7.6)<br />

=<br />

j≥ 1 i><br />

1<br />

In equation (7.6) the definitions <strong>of</strong> Sd, Gkj, Qk1 <strong>and</strong> Qki are equal to the ones used<br />

in the expression <strong>of</strong> the ultimate limit state (equation (7.4)). The values <strong>of</strong> the<br />

combination coefficients according to Eurocode 1 are listed in table 7.4.<br />

Table 7.4: combination coefficients from Updated Recommendations for S<strong>and</strong>wich <strong>Panels</strong>,<br />

Berner et al. (2000)<br />

snow wind temperature<br />

ψ0 0.7 0.5 0.5<br />

ψ1 0.7 0.5 0.5<br />

The frequent load combination (FLC) is most probably experienced frequently by<br />

a construction (element) during the service lifetime as can be seen in figure 7.1. A<br />

limitation on the maximum deflection <strong>of</strong> the s<strong>and</strong>wich panel under FLC is<br />

determined in the Updated European Recommendations for S<strong>and</strong>wich panels<br />

(Berner et al. (2000)): the maximum deflection <strong>of</strong> the panel should not exceed<br />

1/200 th <strong>of</strong> the span under the frequent load combination. When equation (7.5) <strong>and</strong><br />

table 7.3 are compared to equation (7.6) <strong>and</strong> table 7.4, one can notice that Sd<br />

defined by the Updated European Recommendations for S<strong>and</strong>wich panels is<br />

usually larger than Sd defined by Eurocode 1. In Eurocode 1 no variable loads are<br />

considered in the FLC, except the dominating variable load. In the Updated<br />

European Recommendations for S<strong>and</strong>wich panels all variable loads contribute to<br />

Sd, even when these loads are not the dominating variable load. The explanation<br />

for the difference between these approaches is found in the fact that equation (7.5)<br />

<strong>and</strong> table 7.3 from Eurocode 1 are not especially formulated for lightweight<br />

structures. In most design cases, the own weight <strong>of</strong> the construction (element)<br />

contributes considerably to the total design load Sd. In the philosophy <strong>of</strong> Eurocode<br />

1, the own weight <strong>of</strong> the construction is thus generally a major contribution term<br />

in Sd, compared to the variable loads. Therefore, only the variable load <strong>with</strong> the<br />

largest effect is considered in Sd <strong>and</strong> the others are not taken into account.<br />

However, the Updated European Recommendations for S<strong>and</strong>wich panels are<br />

formulated for lightweight panels. When it is assumed that the own weight<br />

contributes only slightly to Sd, the relative importance <strong>of</strong> the variable loads (wind,<br />

203


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

snow <strong>and</strong> temperature) in Sd becomes much more important. Since the studied<br />

panels are lightweight panels indeed, the philosophy <strong>of</strong> the Updated European<br />

Recommendations for S<strong>and</strong>wich panels is followed in this work, thus equation<br />

(7.6) <strong>and</strong> table 7.4 are used. For the sake <strong>of</strong> similarity, the approach <strong>of</strong> the<br />

Updated European Recommendations for S<strong>and</strong>wich panels is also used in ultimate<br />

limit state design.<br />

7.2.5.4 serviceability limit state: characteristic load combination (CLC)<br />

During the lifetime <strong>of</strong> the s<strong>and</strong>wich panel – which is about 50 years – the frequent<br />

combination <strong>of</strong> loads is most probability exceeded <strong>of</strong>ten. Another load<br />

combination is defined <strong>with</strong>in serviceability limit state: the characteristic load<br />

combination (CLC). The characteristic load combination is most probably<br />

exceeded rarely in the service lifetime <strong>of</strong> a construction (element), as illustrated by<br />

figure 7.1. The effects studied under characteristic load combination are effects,<br />

which influence the structure by their one-time occurrence. The characteristic load<br />

combination is higher than the frequent load combination. The definition <strong>of</strong> the<br />

characteristic load combination is equal in Eurocode 1 <strong>and</strong> in the Updated<br />

European Recommendations for S<strong>and</strong>wich panels:<br />

S d ∑G kj + Qk1<br />

+ ∑ψ<br />

0iQki<br />

(7.7)<br />

=<br />

j≥ 1 i><br />

1<br />

In equation (7.7) the definitions <strong>of</strong> Sd, Gkj, Qk1 <strong>and</strong> Qki are equal to the ones used<br />

in the ultimate limit state. According to the Updated European Recommendations<br />

for S<strong>and</strong>wich panels, this load combination is applied in serviceability limit state<br />

to avoid yielding <strong>and</strong> wrinkling <strong>of</strong> the faces, occurring <strong>with</strong>out consequential<br />

failure <strong>of</strong> the panel. The characteristic combination <strong>of</strong> loads still represents a<br />

serviceability condition rather than an ultimate limit condition, since face yielding<br />

<strong>and</strong> wrinkling do not necessarily lead to immediate failure <strong>of</strong> the panel. However,<br />

these phenomena introduce irreversible damage in the s<strong>and</strong>wich panel, which<br />

might lead to failure <strong>of</strong> the s<strong>and</strong>wich panel later or to unacceptable residual<br />

deformations <strong>of</strong> the panel.<br />

The compressive face stress at which wrinkling occurs, can be found in st<strong>and</strong>ard<br />

h<strong>and</strong>books <strong>of</strong> Allen (1969) <strong>and</strong> <strong>of</strong> Zenkert (1995) <strong>and</strong> is:<br />

σ 0. 53<br />

wr = E faEcoGco<br />

(7.8)<br />

<strong>with</strong> Efa <strong>and</strong> Eco being the E-modulus <strong>of</strong> the face <strong>and</strong> the core respectively <strong>and</strong> Gco<br />

being the shear modulus <strong>of</strong> the core. The face wrinkling stress is lowest if Efa<br />

reaches its lowest value. This occurs when only the fibres provide composite<br />

stiffness, after full multiple cracking occurred in tension (<strong>and</strong> possibly serious<br />

degradation occurred under repeated loading), followed by compression in this<br />

face. The face material properties used in this study are presented in paragraph<br />

7.3.1. According to these material properties, the minimum value <strong>of</strong> the face<br />

204


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

stiffness <strong>of</strong> the UD-reinforced faces is 7.2GPa. The core material properties are<br />

listed in paragraph 7.3.1. In this chapter, the value <strong>of</strong> the compressive stress, at<br />

which face wrinkling occurs is thus in all case studies: σwr = 85MPa.<br />

7.2.5.5 serviceability limit state: quasi-static load combination (QLC)<br />

Eurocode 1 <strong>and</strong> the Preliminary Recommendations for S<strong>and</strong>wich <strong>Panels</strong> introduce<br />

a third load combination under serviceability limit state: the quasi-permanent load<br />

combination. This load combination is lower than the frequent load combination<br />

<strong>and</strong> is usually combined <strong>with</strong> creep effects. The formulation <strong>of</strong> this load<br />

combination is:<br />

S d = Gk<br />

+ ψ 2,<br />

iQk<br />

, i<br />

(7.9)<br />

∑<br />

i≥1<br />

The value <strong>of</strong> the combination coefficients ψ2,i for the quasi-permanent<br />

combination are:<br />

Table 7.5: combination coefficient ψ2,I for the quasi-permanent load combination<br />

snow wind temperature<br />

ψ2,i 0.5 0 0<br />

S<strong>and</strong>wich wall panels are subjected to wind load <strong>and</strong> temperature changes, but not<br />

to snow load. Therefore the use <strong>of</strong> the quasi-permanent load on wall panels is<br />

useless. The Preliminary Recommendations for S<strong>and</strong>wich <strong>Panels</strong> advise to analyse<br />

only ro<strong>of</strong> panels under this type <strong>of</strong> loading. However, in practical design<br />

calculations <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> steel faces, the serviceability limit state under<br />

quasi-load combination seldom or never determines the final design. For this<br />

reason, the quasi-permanent load combination has been ab<strong>and</strong>oned in the Updated<br />

Recommendations for S<strong>and</strong>wich <strong>Panels</strong> by Berner et al. (2000).<br />

7.2.6 design <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

The external loads, which are taken into consideration, are own weight, wind load,<br />

snow load <strong>and</strong> temperature. The least-weight solution, still obeying all limit states<br />

(ULS <strong>and</strong> SLS), is to be found. The maximum deflection, the maximum shear<br />

stress in the core <strong>and</strong> the maximum tensile <strong>and</strong> compressive stress in the faces<br />

should thus be determined under the different load combinations, defined more<br />

early.<br />

The steps that are discussed in the design philosophy in this chapter are:<br />

1. study <strong>of</strong> loads acting on the s<strong>and</strong>wich panel<br />

2. analysis <strong>and</strong> design <strong>of</strong> panels under frequent load combination,<br />

serviceability limit state<br />

3. check possible redesign under characteristic load combination,<br />

serviceability limit state<br />

4. check possible redesign under ultimate limit state<br />

205


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

5. panels returning from characteristic load combination to frequent load<br />

combination<br />

6. study <strong>of</strong> the panels under repeated loading<br />

7. comparison <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces <strong>and</strong> s<strong>and</strong>wich<br />

panels <strong>with</strong> steel faces<br />

7.3 Materials <strong>and</strong> modelling<br />

7.3.1 material properties<br />

The faces are made <strong>of</strong> E-glass fibre reinforced IPC, <strong>with</strong> a volume fraction <strong>of</strong> UDreinforcement<br />

<strong>of</strong> 10%. The fibre radius is 7µm, Em is 18GPa <strong>and</strong> Ef is 72GPa. The<br />

Weibull modulus is 3.0, the reference cracking stress σR is 12MPa <strong>and</strong> the initial<br />

frictional matrix-fibre interface shear stress τ0 is 0.7MPa. The final crack spacing<br />

f is 1mm. The compressive failure stress <strong>and</strong> the tensile failure stress are<br />

90MPa. These material properties can be justified from experimental results in<br />

Chapter 2.<br />

In this chapter, the E-modulus is 14MPa. The density is 45kg/m² <strong>and</strong> the<br />

maximum allowed shear stress is 0.20MPa. This type <strong>of</strong> core material is<br />

commonly used in s<strong>and</strong>wich panels <strong>with</strong> steel faces (according to technical sheets<br />

in reference list <strong>of</strong> this chapter).<br />

According to Eurocode 1, the Preliminary European Recommendations for<br />

S<strong>and</strong>wich <strong>Panels</strong> (ECCS, 1991) <strong>and</strong> the Updated European Recommendations for<br />

S<strong>and</strong>wich <strong>Panels</strong> (Berner et al., 2000) the characteristic value <strong>of</strong> material<br />

properties is to be obtained. In general, when a characteristic material strength is<br />

provided, this means there is only 5% probability that the material strength is<br />

lower than the provided value. The design value <strong>of</strong> material properties, to be used<br />

in design calculations, is the value <strong>of</strong> the characteristic material property divided<br />

by a material safety factor, which is larger than one. This way the material<br />

properties as used in design calculations should be underestimated, for reasons <strong>of</strong><br />

safety. However, no information is available at present on characteristic material<br />

properties <strong>of</strong> IPC composites. Moreover, no material safety factor is available for<br />

IPC composites. In this work, estimated average material properties are used,<br />

instead <strong>of</strong> characteristic values. No material safety factors are applied in the<br />

calculations. However, for each design case, the acquired material safety in<br />

ultimate limit state calculations will be discussed later.<br />

7.3.2 thermal expansion <strong>of</strong> UD-reinforced IPC<br />

The design examples <strong>of</strong> this chapter include external temperature variations,<br />

causing internal stresses in the composite. Consequently, the stress-strain<br />

formulation <strong>of</strong> UD-reinforced IPC presented in Chapter 2 should be extended. The<br />

206


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

implementation <strong>of</strong> thermal effects in the behaviour <strong>of</strong> the composite is discussed<br />

here, based on the work <strong>of</strong> Ahn <strong>and</strong> Curtin (1997).<br />

Initially, it is assumed that no external mechanical load is applied <strong>and</strong> no matrix<br />

cracks are formed due to thermal effects. If αf <strong>and</strong> αm are the thermal expansion<br />

coefficients <strong>of</strong> the fibre <strong>and</strong> matrix respectively <strong>and</strong> ∆T is the value <strong>of</strong> the<br />

temperature increment, the resulting matrix stress, σm thermal , is:<br />

Em<br />

E f V<br />

thermal<br />

f<br />

σ m = −(<br />

α m −α<br />

f ) ∆T<br />

(7.10)<br />

Ec1<br />

If a IPC composite section would undergo matrix cracking, due to external<br />

mechanical loading σmc <strong>with</strong>out thermal effects, the external composite cracking<br />

stress <strong>with</strong> inclusion <strong>of</strong> thermal effects would be:<br />

thermal<br />

<strong>with</strong> thermal effects σ m Ec1 σ mc = σ mc −<br />

(7.11)<br />

Em<br />

or rewritten:<br />

<strong>with</strong> thermal effects<br />

σ mc = σ mc + ( α m −α<br />

f ) ∆TE<br />

f V f = σ mc + σ th<br />

(7.12)<br />

The physical meaning <strong>of</strong> equation (7.12) is that under combined thermal <strong>and</strong><br />

mechanical loading, matrix cracking occurs as if an external mechanical load σth<br />

already exists on the composite, before any real mechanical load is applied. If<br />

thermal effects are included into the matrix cracking process, equation (2.38)<br />

should thus be rewritten:<br />

m<br />

⎛ ⎛<br />

⎞⎞<br />

⎜ ⎜ ⎛σ ⎞<br />

c − σ th<br />

cs = cs − −<br />

⎟⎟<br />

⎜ ⎜ ⎜<br />

⎟<br />

f<br />

⎟⎟<br />

⎝ ⎝ ⎝ σ R ⎠ ⎠⎠<br />

max<br />

1 exp<br />

(7.13)<br />

Equation (7.13) is not the only modification in the stress-strain behaviour <strong>of</strong> a IPC<br />

composite. If no external mechanical load is applied, the composite will deform<br />

due to thermal effects. If initially no matrix cracking is introduced, the expression<br />

<strong>of</strong> the composite strain is formulated as a function <strong>of</strong> the temperature:<br />

thermal EmVm<br />

εc<br />

= ( α m −α<br />

f ) ∆T<br />

+ α f ∆T<br />

(7.14)<br />

E<br />

c1<br />

When a composite section undergoes full multiple cracking, the thermal strain<br />

εc thermal is modified. The elastic matrix-fibre interface is released <strong>and</strong> the thermal<br />

strain <strong>of</strong> the composite finally equals the thermal strain <strong>of</strong> the fibre. Supposing full<br />

multiple cracking occurred prior to any thermal loading, the thermal strain<br />

component after multiple cracking is then:<br />

thermal<br />

ε = α ∆T<br />

(7.15)<br />

c<br />

With each crack appearing, the expression <strong>of</strong> the thermal strain <strong>of</strong> an extra section<br />

<strong>of</strong> the composite will thus be expressed by equation (7.15) instead <strong>of</strong> equation<br />

(7.14). Consequently, the stress-strain equations in Chapter 2 should be rewritten.<br />

f<br />

207


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

In case the average crack spacing exceeds twice the debonding length ( ><br />

2δ0):<br />

σ ⎛ ⎞<br />

ţ<br />

− ⎡<br />

⎤<br />

⎜<br />

αδ<br />

cs 2δ<br />

0<br />

0<br />

⎟<br />

EmVm<br />

εc<br />

= 1 + + α ∆T<br />

+ ⎢ ( − ) ∆T<br />

⎜ ⎟ f<br />

α m α f ⎥ (7.16)<br />

Ec1<br />

⎝ cs ⎠<br />

cs ⎣ Ec<br />

⎦<br />

When the average crack spacing becomes smaller than the twice the debonding<br />

length ( < 2δ0):<br />

⎛ cs ⎞<br />

⎜<br />

1 α<br />

εc<br />

= σ c − ⎟ + α f ∆T<br />

⎜ E fV<br />

f E ⎟<br />

(7.17)<br />

⎝ 4δ<br />

0 c1<br />

⎠<br />

The thermal expansion coefficient <strong>of</strong> the fibres, αf, is copied from literature <strong>and</strong> is<br />

8*10 -6 /°C. The thermal expansion coefficient <strong>of</strong> the IPC matrix is obtained from<br />

testing on the TMA (thermomechanical analyser). In figure 7.2, the experimentally<br />

obtained evolution <strong>of</strong> the thermal expansion coefficient <strong>of</strong> pure IPC is plotted<br />

versus temperature. The average value <strong>of</strong> the thermal expansion coefficient <strong>of</strong> pure<br />

IPC matrix is about 12*10 -6 /°C over the temperature range 30-70°C.<br />

α (exp-6/°C)<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

thermo1/2<br />

thermo1/4<br />

thermo1/5<br />

20 30 40 50 60 70 80<br />

temperature (°C)<br />

Figure 7.2: thermal expansion coefficient <strong>of</strong> pure IPC matrix from TMA experiments<br />

stress (MPa)<br />

20<br />

15<br />

10<br />

5<br />

0<br />

increasing T<br />

-0.05 0 0.05 0.1 0.15<br />

strain (%)<br />

Figure 7.3: stress-strain curves under different temperature conditions<br />

208


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

Figure 7.3 shows stress-strain curves for the UD-reinforced IPC material, used in<br />

this chapter. It is assumed that no internal stresses exist due to thermal effects in<br />

the composite at 25°C. The stress-strain curves, calculated at different temperature<br />

levels, are plotted. Both the effects <strong>of</strong> the shift <strong>of</strong> the matrix cracking stress<br />

(equation (7.12) <strong>and</strong> (7.13)) <strong>and</strong> the extra strain term, due to thermal effects<br />

(equation (7.16) or (7.17)) are introduced. The stress-strain curves are plotted for<br />

temperatures from 5°C to 45°C, <strong>with</strong> steps <strong>of</strong> 10°C.<br />

7.4 Case studies<br />

7.4.1 case study I: flat wall panel<br />

The loads, which are considered here, are own weight <strong>of</strong> the panels, wind <strong>and</strong><br />

temperature. The external pressure coefficient cpe is an aerodynamic coefficient<br />

<strong>and</strong> function <strong>of</strong> the geometry <strong>of</strong> the building. If the wall panel is flat, the value <strong>of</strong><br />

the external pressure coefficient on vertical walls can be found from figure 7.4,<br />

which is a plan view, according to Eurocode 1.<br />

wind<br />

+0.8<br />

-1.0<br />

-1.0<br />

-0.8<br />

-0.8<br />

-0.3<br />

Figure 7.4: external wind pressure coefficients, cpe, on vertical walls (plan view)<br />

7.4.2 case study II: flat ro<strong>of</strong> panel<br />

The loads to be considered are snow load, wind pressure, own weight <strong>and</strong><br />

temperature effects. The external pressure coefficient on a flat ro<strong>of</strong> is determined<br />

from ENV 1991-2-4:1995, p.48 <strong>and</strong> is cpe = –0.7.<br />

7.4.3 case study III: flat ro<strong>of</strong> panel <strong>with</strong> intermediate support<br />

Figure 7.5: snow load pressure configurations to be considered for ro<strong>of</strong> panels <strong>with</strong> an extra<br />

support<br />

209


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

The loads in case study III are equal to the ones <strong>of</strong> case study II. However, if an<br />

extra support is used, two snow load configurations should be checked. The first<br />

considers the ro<strong>of</strong> panel is covered <strong>with</strong> snow along the whole length. The second<br />

considers only one half <strong>of</strong> the panel is covered <strong>with</strong> snow (see figure 7.5).<br />

7.4.4 modelling <strong>and</strong> design<br />

In this chapter, finite element modelling is based on the discussion in Chapter 5.<br />

10 PLANE 82 elements are used across the thickness <strong>of</strong> the core <strong>and</strong> 5 elements<br />

across the thickness <strong>of</strong> the faces. Symmetry conditions are used where appropriate.<br />

The number <strong>of</strong> elements along the x-axis is varied, but is minimum 100. This<br />

time-consuming finite element calculation model is used, since from model<br />

comparison in Chapter 5 it has been concluded that the use <strong>of</strong> more simple<br />

s<strong>and</strong>wich element models might lead to considerably less accurate results. It was<br />

also mentioned that it is a rather delicate task to determine the performance <strong>of</strong> a<br />

simplified finite element model for the calculation <strong>of</strong> a s<strong>and</strong>wich panel <strong>with</strong> IPC<br />

composite faces. The ‘aniso’ material behaviour is used to include absence <strong>of</strong> a<br />

priori knowledge <strong>of</strong> the sign <strong>of</strong> normal stress in the faces.<br />

7.5 <strong>Design</strong> under frequent load combination (SLS) <strong>and</strong><br />

characteristic load combination (CLC)<br />

7.5.1 case study I: flat wall panel<br />

Several frequent load combinations are possible. The worst-case combination is<br />

retained for design purposes. Possible worst-case frequent load combinations are<br />

listed in table 7.6.<br />

Table 7.6: possible worst-case frequent load combinations, case study I, serviceability limit state<br />

combination cpe cpi wind pressure Tinside Toutside Q1<br />

number (-) (-) (N/m²) (°C) (°C)<br />

1 +0.8 -0.3 1384 +25 -18 wind<br />

2 -1.0 +0.3 1635 +25 +34 wind<br />

3 +0.8 -0.3 1384 +25 -18 temperature<br />

4 -1.0 +0.3 1635 +25 +34 temperature<br />

Table 7.7: least-weight solutions under frequent load combination, case study I, serviceability<br />

limit state<br />

span worst load core # face layers maximum max. allowed<br />

(m) combination thickness (-) deflection deflection<br />

(mm)<br />

(mm) (mm)<br />

2.0 2 60 1 7.9 10.0<br />

2.5 2 80 1 11.4 12.5<br />

3.0 2 80 2 11.1 15.0<br />

3.5 2 100 2 13.1 17.5<br />

4.0 2 100 3 15.1 20.0<br />

210


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

In combinations number 1 <strong>and</strong> 3, both the wind pressure load <strong>and</strong> the temperature<br />

load deform the panel to the inside <strong>of</strong> the building. In combinations number 2 <strong>and</strong><br />

4, both the wind pressure load <strong>and</strong> temperature load deform the panel to the<br />

outside.<br />

The serviceability limit state on deflections requires the maximum deflection does<br />

not exceed 1/200 th <strong>of</strong> the span under the frequent combination <strong>of</strong> loads. This<br />

requirement is found in the Updated Recommendations for S<strong>and</strong>wich <strong>Panels</strong><br />

(Berner et al. (2000)). In table 7.7, the least-weight solutions satisfying this<br />

condition are listed for all load cases <strong>of</strong> table 7.6. The value <strong>of</strong> the maximum<br />

deflection is also mentioned.<br />

The least-weight solutions <strong>of</strong> table 7.7 are recalculated under the characteristic<br />

combination <strong>of</strong> loads. The maximum compressive stresses in the faces are listed in<br />

table 7.8. It can be seen that the compressive wrinkling stress <strong>of</strong> 85MPa from<br />

paragraph 7.2.5.4 is never exceeded under the characteristic load combination. The<br />

maximum deflections are also listed (just for the reader’s information) in table 7.8.<br />

None <strong>of</strong> the least-weight solutions, presented in table 7.7 should be modified.<br />

Table 7.8: recheck solutions under characteristic load combination, case study I, serviceability<br />

limit state<br />

Span<br />

maximum<br />

max. allowed maximum<br />

(m) compressive stress compressive stress deflection (mm)<br />

(MPa)<br />

(MPa)<br />

2.0 18 85 20.7<br />

2.5 21 85 28.9<br />

3.0 15 85 28.4<br />

3.5 17 85 34.3<br />

4.0 14 85 37.8<br />

7.5.2 case study II: flat ro<strong>of</strong> panel<br />

For this case study, load combinations <strong>of</strong> snow, wind <strong>and</strong> temperature should be<br />

considered. For each case a least-weight solution under the frequent load<br />

combination is calculated. The possible worst-case frequent load combinations are<br />

listed in table 7.9.<br />

Table 7.9: possible worst-case frequent load combinations, case study II, serviceability limit state<br />

number cpe<br />

(-)<br />

cpi<br />

(-)<br />

wind<br />

pressure<br />

(N/m²)<br />

snow<br />

pressure<br />

(N/m²)<br />

Tinside<br />

(°C)<br />

Toutside<br />

(°C)<br />

1 -0.7 +0.3 1258 0 +25 +34 wind<br />

2 -0.7 +0.3 1258 0 +25 +34 temperature<br />

3 0 -0.3 377 400 +25 -18 snow<br />

4 0 -0.3 377 400 +25 -18 temperature<br />

211<br />

Q1


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

The least-weight solutions under the worst-case frequent load combinations are<br />

listed in table 7.10 for the ro<strong>of</strong> panel <strong>with</strong> two supports. The maximum deflection<br />

remains below 1/200 th <strong>of</strong> the span under FLC.<br />

The characteristic combination <strong>of</strong> loads is also checked. Table 7.11 shows the<br />

maximum compressive stress in the faces <strong>and</strong> the maximum deflection under the<br />

characteristic load combination (only for the reader’s information).<br />

Table 7.10: least-weight solutions under frequent load combination, case study II, serviceability<br />

limit state<br />

span worst load core # face layers maximum max. allowed<br />

(m) combination thickness (-) deflection deflection<br />

(mm)<br />

(mm) (mm)<br />

2.0 1 60 1 5.9 10.0<br />

2.5 1 80 1 8.1 12.5<br />

3.0 1 100 1 11.0 15.0<br />

3.5 1 80 2 15.3 17.5<br />

4.0 1 100 2 16.8 20.0<br />

Table 7.11: recheck solutions under characteristic load combination, case study II, serviceability<br />

limit state<br />

span<br />

maximum max. allowed maximum<br />

(m) compressive stress compressive stress deflection (mm)<br />

(MPa)<br />

(MPa)<br />

2.0 14 85 14.7<br />

2.5 16 85 20.8<br />

3.0 19 85 28.2<br />

3.5 16 85 39.2<br />

4.0 16 85 43.6<br />

7.5.3 case study III: flat ro<strong>of</strong> panel <strong>with</strong> intermediate support<br />

The external loads are similar to the ones in case study II. The only difference <strong>with</strong><br />

the previous calculations is that both snow load configurations <strong>of</strong> figure 7.5 are to<br />

be considered. In table 7.12 all possible worst load combinations are printed. For<br />

load combination number 3 <strong>and</strong> 4, it is assumed that the whole panel is covered<br />

<strong>with</strong> snow. Load combinations 5 <strong>and</strong> 6 are calculated, considering only half the<br />

panel is covered <strong>with</strong> snow.<br />

Figure 7.6a shows the deformation <strong>of</strong> a panel <strong>with</strong> span <strong>of</strong> 2m (total length <strong>of</strong> 4m),<br />

core thickness <strong>of</strong> 40mm <strong>and</strong> 1 layer/face under load combination 1 <strong>of</strong> table 7.12.<br />

Figure 7.6b shows this panel, when it is subjected to load combination number 5.<br />

212


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

Table 7.12: possible worst frequent combination load cases, case study III, serviceability limit<br />

state<br />

number cpe cpi wind snow Tinside Toutside Q1<br />

(-) (-) pressure pressure (°C) (°C)<br />

(N/m²)<br />

(N/m²)<br />

1 -0.7 +0.3 1258 0 +25 +34 wind<br />

2 -0.7 +0.3 1258 0 +25 +34 temperature<br />

3 0 -0.3 377 400 (full) +25 -18 snow<br />

4 0 -0.3 377 400 (full) +25 -18 temperature<br />

5 0 -0.3 377 400 (half) +25 -18 snow<br />

6 0 -0.3 377 400 (half) +25 -18 temperature<br />

Table 7.13: least-weight solutions under frequent load combination, case study III, serviceability<br />

limit state<br />

span worst load core # face layers maximum max. allowed<br />

(m) combination thickness (-) deflection deflection<br />

(mm)<br />

(mm) (mm)<br />

2.0 1 40 1 6.1 10.0<br />

2.5 1 60 1 6.7 12.5<br />

3.0 1 60 1 13.5 15.0<br />

3.5 1 80 1 14.1 17.5<br />

4.0 1 100 1 15.6 20.0<br />

Figure 7.6a: deformation <strong>of</strong> a s<strong>and</strong>wich panel under load combination 1 <strong>of</strong> table 7.12<br />

2m span, 40mm core thickness, 1 face layer<br />

Figure 7.6b: deformation <strong>of</strong> a s<strong>and</strong>wich panel under load combination 5 <strong>of</strong> table 7.12<br />

2m span, 40mm core thickness, 1 face layer<br />

213


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

The least-weight solutions <strong>of</strong> table 7.13 are recalculated under the characteristic<br />

combinations <strong>of</strong> loads. Table 7.14 shows that all panels <strong>of</strong> table 7.13 fulfil the<br />

condition under characteristic load combination. This means wrinkling <strong>of</strong> the face<br />

in compression is avoided.<br />

Table 7.14: recheck solutions under characteristic load combination, case study III, serviceability<br />

limit state<br />

span<br />

maximum<br />

max. allowed maximum<br />

(m) compressive stress compressive stress deflection (mm)<br />

(MPa)<br />

(MPa)<br />

2.0 26 85 14.1<br />

2.5 29 85 14.6<br />

3.0 40 85 33.7<br />

3.5 42 85 35.3<br />

4.0 45 85 38.6<br />

7.5.4 conclusions<br />

Three s<strong>and</strong>wich panel types were discussed under monotonic loading. For all test<br />

cases <strong>and</strong> spans, a least-weight solution is obtained.<br />

All panels are initially designed to fulfil the maximum deflection condition under<br />

frequent load combination (which is a serviceability limit state). The core<br />

thickness <strong>and</strong> number <strong>of</strong> layers are chosen in such a way, that the maximum<br />

deflection stays below 1/200 th <strong>of</strong> the span <strong>and</strong> the weight <strong>of</strong> the panel is minimal.<br />

The least-weight solutions are calculated under the characteristic load<br />

combination. The design value <strong>of</strong> the combination <strong>of</strong> loads under characteristic<br />

load combination is higher than the one under frequent load combination. For<br />

none <strong>of</strong> the previously calculated design solutions, obtained under the frequent<br />

load combination, the design had to be modified under characteristic load<br />

combinations. The face stresses stay well below the wrinkling stress for all<br />

calculated panels.<br />

In general, the weight <strong>of</strong> the least-weight s<strong>and</strong>wich wall panels is higher than the<br />

weight <strong>of</strong> the ro<strong>of</strong> panels. The reason for this can be found in the fact that in the<br />

worst-case load combination <strong>of</strong> loads, the external wind pressure is higher for wall<br />

panels than for the ro<strong>of</strong> panels.<br />

214


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

7.6 <strong>Design</strong> in ultimate limit state<br />

7.6.1 case study I: flat wall panel<br />

The load combination <strong>of</strong> equation (7.4) is used to express the ultimate limit state.<br />

As has been mentioned earlier in this chapter, the failure modes <strong>of</strong> s<strong>and</strong>wich<br />

panels in bending, which should be considered, are failure <strong>of</strong> a face in tension,<br />

failure <strong>of</strong> a face in compression <strong>and</strong> core failure in shear. The design solutions <strong>of</strong><br />

table 7.7 are checked for failure in ultimate limit state. Table 7.15 lists the<br />

maximum calculated stresses in faces <strong>and</strong> core for all considered spans.<br />

Table 7.15: recalculating the least-weight solutions, case study I, ultimate limit state<br />

span<br />

(m)<br />

maximum tensile<br />

stress<br />

(MPa)<br />

maximum<br />

compressive<br />

stress (MPa)<br />

maximum core<br />

shear stress<br />

(MPa)<br />

2.0 26 27 0.054<br />

2.5 32 30 0.054<br />

3.0 23 22 0.057<br />

3.5 25 24 0.058<br />

4.0 22 21 0.060<br />

It can be seen from table 7.15 that none <strong>of</strong> the least-weight solutions, determined<br />

under serviceability limit state <strong>of</strong> maximum deflection earlier, fails under ultimate<br />

limit state. The failure stress <strong>of</strong> UD-reinforced IPC is about 90MPa in tension <strong>and</strong><br />

in compression. The calculated stresses in the faces <strong>of</strong> the studied s<strong>and</strong>wich panels<br />

stay well below this failure value. The maximum experienced core shear stress is<br />

about 0.06MPa, which is still lower than the maximum allowed stress <strong>of</strong> 0.20MPa.<br />

The safety factor on the faces materials is about 4. The maximum allowed core<br />

shear stress is about 3 times the calculated maximum core stress.<br />

7.6.2 case study II: flat ro<strong>of</strong> panel<br />

The design solutions <strong>of</strong> table 7.10 are checked for failure in ultimate limit state.<br />

The maximum core <strong>and</strong> face stresses under ULS are listed in table 7.16. None <strong>of</strong><br />

the proposed solutions fails under ultimate limit state. The safety factors on the<br />

faces <strong>and</strong> core materials are both about 4.<br />

Table 7.16: recalculating the least-weight solutions, case study II, ultimate limit state<br />

span<br />

(m)<br />

maximum tensile<br />

stress<br />

(MPa)<br />

maximum<br />

compressive<br />

stress (MPa)<br />

maximum core<br />

shear stress<br />

(MPa)<br />

2.0 20 21 0.038<br />

2.5 23 25 0.044<br />

3.0 18 19 0.042<br />

3.5 23 25 0.048<br />

4.0 24 26 0.051<br />

215


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

7.6.3 case study III: flat ro<strong>of</strong> panel <strong>with</strong> intermediate support<br />

All least-weight solutions under frequent load combination (table 7.13) are again<br />

calculated in the ultimate limit state. The results are listed in table 7.17. Again, no<br />

panel exceeds limits in the ultimate limit state.<br />

Table 7.17: recalculating the least-weight solutions, case study III, ultimate limit state<br />

spanlength<br />

(m)<br />

maximum tensile<br />

stress<br />

(MPa)<br />

maximum<br />

compressive<br />

stress (MPa)<br />

maximum core<br />

shear stress<br />

(MPa)<br />

2.0 29 41 0.058<br />

2.5 30 44 0.060<br />

3.0 44 62 0.060<br />

3.5 44 65 0.062<br />

4.0 46 70 0.063<br />

7.6.4 conclusions<br />

The least-weight panel design solutions, obtained more early under the frequent<br />

load combination in serviceability limit state, are calculated for the possible<br />

occurrence <strong>of</strong> failure in ultimate limit state. All studied panels are still <strong>with</strong>in safe<br />

limits in ultimate limit state, <strong>with</strong>out any modifications being necessary.<br />

7.7 <strong>Design</strong> under FLC after unloading from CLC<br />

7.7.1 introduction<br />

The characteristic load combination is experienced rarely, which means there is a<br />

probability that it occurs during the lifetime. The effects studied under<br />

characteristic load combination are effects, which influence the structure by their<br />

one-time occurrence. For s<strong>and</strong>wich panels <strong>with</strong> steel faces, it is applied in<br />

serviceability limit state to avoid yielding <strong>and</strong> buckling <strong>of</strong> the face <strong>and</strong> core layers,<br />

occurring <strong>with</strong>out consequential failure <strong>of</strong> the panel under variable live <strong>and</strong><br />

environmental loads<br />

If yielding <strong>of</strong> the faces or <strong>of</strong> the core or wrinkling <strong>of</strong> a face occurs, irreversible<br />

deformations are introduced. There are two reasons why yielding or buckling <strong>of</strong><br />

steel faces should be avoided:<br />

- Even if yielding or buckling <strong>of</strong> the faces does not lead to<br />

immediate failure <strong>of</strong> the panel, it may finally lead to final failure after repeated<br />

loading.<br />

- Both face yielding <strong>and</strong> face wrinkling introduce irreversible<br />

deformations. After the characteristic combination <strong>of</strong> loads disappears, the<br />

maximum deflection under the frequent load combination will return to its original<br />

216


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

value, provided no irreversible phenomena occurred. This is the case if steel faces<br />

are used <strong>and</strong> no face wrinkling or yielding occurred.<br />

The European Recommendations for S<strong>and</strong>wich <strong>Panels</strong> are widely used for analysis<br />

<strong>and</strong> design <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> steel faces. Under the characteristic<br />

combination <strong>of</strong> loads, the focus in design <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> steel faces is<br />

put on avoiding face wrinkling <strong>and</strong> face yielding (non-linear face behaviour) under<br />

the frequent load combination. This face yielding might introduce unacceptable<br />

residual deformations.<br />

In contrast <strong>with</strong> s<strong>and</strong>wich panels <strong>with</strong> steel faces, non-linear face behaviour is<br />

allowed to occur in service conditions in the design <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC<br />

composite faces. For s<strong>and</strong>wich panels <strong>with</strong> IPC faces, the condition <strong>of</strong> avoiding<br />

non-linear face behaviour under characteristic load combination can usually not be<br />

fulfilled <strong>and</strong> will therefore be modified on limiting deflections: the maximum<br />

deflection under frequent load combination should not exceed 1/200 th <strong>of</strong> the span,<br />

even when the characteristic load combination has been experienced by the panel<br />

before. Figure 7.7 illustrates this modified interpretation.<br />

All test case panels have been designed in paragraph 7.5 such that the maximum<br />

deflection does not exceed 1/200 th <strong>of</strong> the span under the frequent load combination<br />

(noted by (1) in figure 7.7). Under the characteristic load combination, the panel is<br />

temporarily allowed to violate this serviceability limit state ((2) in figure 7.7).<br />

However, after the characteristic load disappears, a maximum deflection<br />

exceeding 1/200 th <strong>of</strong> the span under FLC should still be avoided (this condition is<br />

illustrated here as being violated in (3) <strong>of</strong> figure 7.7).<br />

external load<br />

CLC<br />

FLC<br />

(1)<br />

(3)<br />

(2)<br />

SLS<br />

maximum deflection < span/200<br />

maximum deflection<br />

Figure 7.7: load versus deflection behaviour <strong>of</strong> a s<strong>and</strong>wich panel <strong>with</strong> IPC composite faces<br />

(1) under the frequent load combination<br />

(2) under the characteristic load combination<br />

(3) under the frequent load combination, after the characteristic load combination occurred<br />

217


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

7.7.2 case study I: flat wall panel<br />

In table 7.7, the least-weight solutions are listed, still obeying the serviceability<br />

limit state under the frequent combinations <strong>of</strong> loads. These panels are now<br />

recalculated. All panels are first loaded from the frequent combination <strong>of</strong> loads to<br />

the characteristic combination <strong>of</strong> loads. Afterwards, the panels are unloaded from<br />

the characteristic to the frequent load combination.<br />

Table 7.18 lists the calculated maximum deflections <strong>of</strong> the studied panels under<br />

initial frequent load combination, characteristic load combination <strong>and</strong><br />

subsequently return to the frequent load combination. Solutions exceeding the<br />

serviceability limit state <strong>of</strong> limited deflections (deflection < 1/200 th <strong>of</strong> the span)<br />

under the frequent combination, after being loaded to the characteristic load<br />

combination are printed italic in table 7.18.<br />

Table 7.18: recalculating the least-weight solutions under frequent load combination after<br />

occurrence <strong>of</strong> the characteristic load combination, case study I: deflections<br />

span<br />

(m)<br />

FLC CLC return to FLC FLC<br />

maximum<br />

deflection<br />

(mm)<br />

maximum<br />

deflection<br />

(mm)<br />

residual maximum<br />

deflection<br />

(mm)<br />

max. allowed<br />

deflection<br />

(mm)<br />

2.0 7.9 20.7 12.0 10.0<br />

2.5 11.4 28.9 17.0 12.5<br />

3.0 11.1 28.4 15.3 15.0<br />

3.5 13.1 34.3 18.6 17.5<br />

4.0 15.1 37.8 20.0 20.0<br />

Table 7.19: recalculating the least-weight solutions under frequent load combination after<br />

occurrence <strong>of</strong> the characteristic load combination, case study I: new least-weight solutions<br />

span<br />

(m)<br />

previous solution new solution<br />

core<br />

thickness<br />

(mm)<br />

# face layers<br />

(-)<br />

core<br />

thickness<br />

(mm)<br />

# face layers<br />

(-)<br />

deflection<br />

(mm)<br />

2.0 60 1 80 1 5.9<br />

2.5 80 1 100 1 9.8<br />

3.0 80 2 100 2 9.1<br />

3.5 100 2 100 3 10.5<br />

4.0 100 3 100 3 20.0<br />

From table 7.18 it can be noticed that only the maximum deflection <strong>of</strong> the panel<br />

<strong>with</strong> 4m span does not exceed 1/200 th <strong>of</strong> the span under the frequent load<br />

combination, provided the characteristic load combination has been applied once<br />

in the load history <strong>of</strong> the panel. The other solutions should be modified: a larger<br />

core thickness should be chosen or the number <strong>of</strong> face layers should be increased.<br />

218


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

Table 7.19 lists minimum weight solutions for all spans, such that the residual<br />

deflections do not violate the serviceability limit state under the frequent<br />

combination <strong>of</strong> loads, even after the characteristic load combination has been<br />

applied.<br />

7.7.3 case study II: flat ro<strong>of</strong> panel<br />

Table 7.20 lists the maximum deflections <strong>of</strong> the studied panels under initial<br />

frequent load combination, characteristic load combination <strong>and</strong> subsequently<br />

return to the frequent load combination for a flat ro<strong>of</strong> panel. From table 7.20 it can<br />

be seen that the panels <strong>with</strong> span <strong>of</strong> 3.0m, 3.5m <strong>and</strong> 4.0m are to be modified. The<br />

new least-weight solution panels are listed in table 7.21.<br />

Table 7.20: recalculating the least-weight solutions under frequent load combination after<br />

occurrence <strong>of</strong> the characteristic load combination, case study II: deflections<br />

span<br />

(m)<br />

FLC CLC return to FLC FLC<br />

maximum<br />

deflection<br />

(mm)<br />

maximum<br />

deflection<br />

(mm)<br />

residual maximum<br />

deflection<br />

(mm)<br />

max. allowed<br />

deflection<br />

(mm)<br />

2.0 5.9 14.7 7.8 10.0<br />

2.5 8.1 20.8 11.3 12.5<br />

3.0 11.0 28.2 16.4 15.0<br />

3.5 15.3 39.2 21.3 17.5<br />

4.0 16.8 43.6 23.9 20.0<br />

Table 7.21: recalculating the least-weight solutions under frequent load combination after<br />

occurrence <strong>of</strong> the characteristic load combination, case study II: new least-weight solutions<br />

span<br />

(m)<br />

previous solution new solution<br />

core<br />

thickness<br />

(mm)<br />

# face layers<br />

(-)<br />

core<br />

thickness<br />

(mm)<br />

# face layers<br />

(-)<br />

deflection<br />

(mm)<br />

2.0 60 1 60 1 7.8<br />

2.5 80 1 80 1 11.3<br />

3.0 100 1 80 2 10.0<br />

3.5 80 2 100 2 12.6<br />

4.0 100 2 100 3 13.0<br />

7.7.4 case study III: flat ro<strong>of</strong> panel <strong>with</strong> intermediate support<br />

For a flat ro<strong>of</strong> panel <strong>with</strong> intermediate support, the previously obtained design<br />

solutions for a span <strong>of</strong> 3.0m, 3.5m <strong>and</strong> 4.0 m are to be modified as can be seen in<br />

table 7.22. The modified panels are printed in table 7.23.<br />

219


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

Table 7.22: recalculating the least-weight solutions under frequent load combination after<br />

occurrence <strong>of</strong> the characteristic load combination, case study III: deflections<br />

span<br />

(m)<br />

FLC CLC return to FLC FLC<br />

maximum<br />

deflection<br />

(mm)<br />

maximum<br />

deflection<br />

(mm)<br />

residual maximum<br />

deflection<br />

(mm)<br />

max. allowed<br />

deflection<br />

(mm)<br />

2.0 6.1 14.1 7.3 10.0<br />

2.5 6.7 14.6 8.2 12.5<br />

3.0 13.5 33.7 19.1 15.0<br />

3.5 14.1 35.3 20.1 17.5<br />

4.0 15.6 38.6 22.2 20.0<br />

Table 7.23: recalculating the least-weight solutions under frequent load combination after<br />

occurrence <strong>of</strong> the characteristic load combination, case study III: new least-weight solution<br />

span<br />

length<br />

(m)<br />

previous solution new solution<br />

core<br />

thickness<br />

(mm)<br />

# face layers<br />

(-)<br />

core<br />

thickness<br />

(mm)<br />

# face layers<br />

(-)<br />

maximum<br />

deflection<br />

(mm)<br />

2.0 40 1 40 1 7.3<br />

2.5 60 1 60 1 8.2<br />

3.0 60 1 80 1 9.9<br />

3.5 80 1 100 1 11.9<br />

4.0 100 1 80 2 15.8<br />

7.7.5 conclusions<br />

A slightly modified approach has been used from the one used for s<strong>and</strong>wich<br />

panels <strong>with</strong> steel faces. The different face behaviour <strong>of</strong> IPC composite faces,<br />

compared to steel faces, dem<strong>and</strong>s extra care in the design <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong><br />

IPC composite faces. If the characteristic load combination is applied to the<br />

s<strong>and</strong>wich panels <strong>with</strong> IPC faces, these panels are allowed to show deflections<br />

exceeding 1/200 th <strong>of</strong> the span. However, after this characteristic load combination<br />

occurred, they should not exceed this limit under the frequent load combination.<br />

About 50% <strong>of</strong> the presented panels, designed earlier under the frequent load<br />

combination in paragraph 7.5 exceeded this limit here. The core <strong>and</strong>/or face<br />

thickness <strong>of</strong> these panels has been increased. New least-weight solutions, slightly<br />

heavier than the panels proposed more early, are presented.<br />

7.8 <strong>Design</strong> under repeated loading<br />

7.8.1 introduction.<br />

Due to the accumulation <strong>of</strong> damage in the faces, the deformation <strong>of</strong> s<strong>and</strong>wich<br />

panels <strong>with</strong> IPC composite faces increases <strong>with</strong> elapsed number <strong>of</strong> loading cycles.<br />

220


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

Special care should be taken to avoid that the maximum deflection <strong>of</strong> the panel<br />

exceeds the serviceability limit state after several load cycles have been<br />

experienced.<br />

Figure 7.8 explains why all solutions, obtained previously, should be recalculated<br />

once more. Even if the panels are designed to limit the deflections under the FLC<br />

after occurrence <strong>of</strong> the CLC, they might still violate the serviceability limit state<br />

after being loaded repeatedly up to the frequent load combination.<br />

external load<br />

CLC<br />

FLC<br />

(1)<br />

(3)<br />

(4)<br />

(2)<br />

repeated cycling<br />

SLS<br />

maximum deflection< span/200<br />

maximum deflections<br />

Figure 7.8: load versus deflection behaviour <strong>of</strong> a s<strong>and</strong>wich panel <strong>with</strong> IPC composite faces<br />

(1) under the frequent load combination<br />

(2) under the characteristic load combination<br />

(3) under the frequent load combination, after the characteristic load combination occurred<br />

(4) under repeated loading<br />

Especially wind load can be highly variable. The description <strong>of</strong> the r<strong>and</strong>om nature<br />

<strong>of</strong> wind load has a long history <strong>and</strong> is still subject <strong>of</strong> study. The distribution <strong>of</strong><br />

mean wind speeds can be measured <strong>and</strong> modelled.<br />

Figure 7.9: measurement <strong>of</strong> wind speed <strong>and</strong> Weibull distribution model <strong>of</strong> mean wind speed,<br />

Davenport (1966)<br />

221


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

Mean speeds are generally averaged over a period <strong>of</strong> 10 minutes to 1 hour. The<br />

probability that a certain mean wind velocity is measured can be described<br />

adequately by a Weibull distribution. Figure 7.9 shows a typical mean wind speed<br />

distribution as measured <strong>and</strong> modelled <strong>with</strong> Weibull statistics by Davenport,<br />

(1966). This Weibull model can then be used to predict the number <strong>of</strong> times a<br />

certain mean wind speed will probably be exceeded during a chosen return period.<br />

This model is also adopted in the national Belgian st<strong>and</strong>ards NBN B 03-002-1.<br />

Figure 7.10 is extracted from Eurocode 1 <strong>and</strong> represents the number <strong>of</strong> times a<br />

certain value <strong>of</strong> the mean wind speed will probably be exceeded. S(w) is the<br />

maximum wind speed, <strong>with</strong> a return period <strong>of</strong> the lifetime <strong>of</strong> the structure.<br />

Figure 7.10: Number <strong>of</strong> times a certain value <strong>of</strong> mean wind speed will probably be exceeded for a<br />

return period <strong>of</strong> 1,5 or 50 years, Eurocode 1, NBN B 03-002-1, 1988<br />

The maximum normal stresses in the x-direction (length axis) <strong>of</strong> the face elements,<br />

σc max , are retrieved from the ANSYS results file under the characteristic<br />

combination <strong>of</strong> loads. These values are used, as explained in Chapter 4, for the<br />

calculation <strong>of</strong> the extra strain terms under repeated loading for all finite elements.<br />

The additional strains, ∆εc,N repeat , resulting from repeated loading are inserted in the<br />

s<strong>and</strong>wich geometry. The resulting additional deflections due to repeated loading<br />

are thus obtained. The number <strong>of</strong> experienced load cycles is set to one million for<br />

all discussed panels. The methodology <strong>and</strong> the macro repeated.mac, explained in<br />

Chapter 5 are used. Degradation <strong>of</strong> the matrix-fibre interface is considered to be<br />

the sole damage mechanism here, according to Chapter 4. The interface<br />

degradation is function <strong>of</strong> the number <strong>of</strong> elapsed cycles:<br />

ω = C1 − C2<br />

ln( N)<br />

(7.18)<br />

The values <strong>of</strong> the constants C1 <strong>and</strong> C2 are chosen 0.85 <strong>and</strong> 0.02 respectively, since<br />

these are typical experimentally obtained values in Chapter 4.<br />

7.8.2 case study I: flat wall panel<br />

The panels <strong>of</strong> table 7.19 are calculated under repeated FLC. All panels are first<br />

loaded under the characteristic combination <strong>of</strong> loads. Then, the panels are<br />

unloaded from the characteristic to the frequent load combination. Finally, they<br />

222


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

are loaded repeatedly one million times. The calculation <strong>of</strong> the extra deflections <strong>of</strong><br />

the s<strong>and</strong>wich panel under repeated loading is performed in as explained Chapter 5<br />

<strong>and</strong> performed in Chapter 6.<br />

Table 7.24 lists the maximum deflections <strong>of</strong> the panels after two loading steps:<br />

after step (3) from figure 7.8 <strong>and</strong> after step (4). Step (3) is the residual deflection<br />

under FLC after unloading from the CLC. The additional deflection terms,<br />

resulting from repeated loading under FLC (from (3) to (4) in figure 7.8) is added<br />

to this residual deflection term. The final deflections, which are due to residual<br />

deflections after occurrence <strong>of</strong> the CLC <strong>with</strong> subsequent repeated loading under<br />

FLC, are listed in table 7.24. Solutions printed in italic exceed the limitation <strong>of</strong><br />

deflections in serviceability limit state.<br />

Table 7.25 lists the new design panels, which do not exceed the limitation on the<br />

maximum deflection, even after one million load cycles in FLC are applied.<br />

Table 7.24: recalculating the least-weight solutions under repeated frequent load combination<br />

after occurrence <strong>of</strong> the characteristic load combination, case study I: deflections<br />

return to FLC repeated load FLC FLC<br />

span<br />

maximum<br />

maximum<br />

max. allowed<br />

(m) deflection (mm) deflection (mm) deflection (mm)<br />

2.0 5.9 7.3 10.0<br />

2.5 9.8 11.6 12.5<br />

3.0 9.1 11.5 15.0<br />

3.5 10.5 13.4 17.5<br />

4.0 20.0 24.7 20.0<br />

Table 7.25: recalculating the least-weight solutions under repeated frequent load combination<br />

after occurrence <strong>of</strong> the characteristic load combination, case study I: new least-weight solution<br />

span<br />

(m)<br />

previous solution new solution<br />

core<br />

thickness<br />

(mm)<br />

# face<br />

layers<br />

(-)<br />

core<br />

thickness<br />

(mm)<br />

# face<br />

layers<br />

(-)<br />

deflection<br />

(mm)<br />

2.0 80 1 80 1 7.3<br />

2.5 100 1 100 1 11.6<br />

3.0 100 2 100 2 11.5<br />

3.5 100 3 100 3 13.4<br />

4.0 100 3 100 4 16.7<br />

From table 7.25, it is clear that for a span <strong>of</strong> 4m, an extra face layer should be used<br />

at the top <strong>and</strong> bottom, to make sure the limitation <strong>of</strong> the deflection in serviceability<br />

limit state is not exceeded after repeated loading.<br />

7.8.3 case study II: flat ro<strong>of</strong> panel<br />

223


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

If the panels in table 7.21 are calculated under repeated loading, table 7.26<br />

illustrates that when one million load cycles are applied, only the 2.5m span<br />

s<strong>and</strong>wich panel should be modified.<br />

In table 7.25, the new least-weight solutions are listed, provided the number <strong>of</strong><br />

applied load cycles in FLC is one million.<br />

Table 7.26: recalculating the least-weight solutions under repeated frequent load combination<br />

after occurrence <strong>of</strong> the characteristic load combination, case study II: deflections<br />

return to FLC repeated load FLC FLC<br />

span<br />

maximum<br />

maximum<br />

max. allowed<br />

(m) deflection (mm) deflection (mm) deflection (mm)<br />

2.0 7.8 9.7 10.0<br />

2.5 11.3 13.5 12.5<br />

3.0 10.0 12.9 15.0<br />

3.5 12.6 16.0 17.5<br />

4.0 13.0 16.8 20.5<br />

Table 7.27: recalculating the least-weight solutions under repeated frequent load combination<br />

after occurrence <strong>of</strong> the characteristic load combination, case study II: new least-weight solution<br />

span<br />

(m)<br />

previous solution new solution<br />

core<br />

thickness<br />

(mm)<br />

# face<br />

layers<br />

(-)<br />

core<br />

thickness<br />

(mm)<br />

# face<br />

layers<br />

(-)<br />

deflection<br />

(mm)<br />

2.0 60 1 60 1 9.7<br />

2.5 80 1 100 1 6.7<br />

3.0 80 2 80 2 12.9<br />

3.5 100 2 100 2 16.0<br />

4.0 100 3 100 3 16.8<br />

7.8.4 case study III: flat ro<strong>of</strong> panel <strong>with</strong> intermediate support<br />

The panels <strong>of</strong> table 7.23 are calculated under repeated loading. Table 7.28 shows<br />

no panel has to be modified, if the number <strong>of</strong> applied load cycles is one million.<br />

No redesign <strong>of</strong> panels is thus required.<br />

7.28: recalculating the least-weight solutions under repeated frequent load combination after<br />

occurrence <strong>of</strong> the characteristic load combination, case study III: deflections<br />

return to FLC repeated load FLC FLC<br />

span<br />

maximum<br />

maximum<br />

max. allowed<br />

(m) deflection (mm) deflection (mm) deflection (mm)<br />

2.0 7.3 9.1 10.0<br />

2.5 8.2 10.1 12.5<br />

3.0 9.9 12.0 15.0<br />

3.5 11.9 17.1 17.5<br />

4.0 15.8 18.4 20.0<br />

224


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

7.8.5 conclusions<br />

The number <strong>of</strong> panels that is to be redesigned under repeated loading is very<br />

limited, provided the number <strong>of</strong> applied load cycles is one million. In general, only<br />

small changes to the geometry <strong>of</strong> the studied panel are required to assure no panel<br />

exceeds the limitation on the maximum deflection in serviceability limit state<br />

under repeated loading. However, redesign is needed in some cases.<br />

7.9 Comparison <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite<br />

faces <strong>and</strong> s<strong>and</strong>wich panels <strong>with</strong> steel faces<br />

7.9.1 introduction<br />

In this paragraph, the least-weight solutions, presented for the three studied<br />

s<strong>and</strong>wich panel types, are compared <strong>with</strong> the least-weight solutions <strong>of</strong> s<strong>and</strong>wich<br />

panels <strong>with</strong> steel faces. Necessary data on steel faces are extracted from technical<br />

sheets, provided by Belgian companies, producing s<strong>and</strong>wich panels <strong>with</strong> steel<br />

faces (see technical references, paragraph 7.11). The weight <strong>of</strong> the panel per m² is<br />

determined for all solutions.<br />

7.9.2 comparison <strong>of</strong> solutions<br />

It is not as easy to change the face thickness <strong>of</strong> steel faces as it is for IPC<br />

composite faces. In the technical sheets, provided by Belgian companies, the steel<br />

face thickness is usually 0.4, 0.5 or 0.7mm. These values are applied for all panels<br />

<strong>with</strong> steel faces here. The thermal expansion coefficient <strong>of</strong> steel is 12e -6 /°C. The<br />

density <strong>of</strong> steel is 7800kg/m³. The density <strong>of</strong> IPC faces is 2000kg/m³. Tables 7.29<br />

to 7.31 show the least-weight solutions for all case studies, discussed in this<br />

chapter.<br />

Table 7.29: case study I: least-weight solutions, s<strong>and</strong>wich panels <strong>with</strong> steel faces versus s<strong>and</strong>wich<br />

panels <strong>with</strong> IPC composite faces<br />

steel faces IPC composite faces<br />

span core face weight core # face weight<br />

(m) (mm) (mm) (kg/m²) (mm) layers (kg/m²)<br />

2.0 40 0.4 8.0 80 1 6.6<br />

2.5 40 0.4 8.0 100 1 7.5<br />

3.0 60 0.4 8.9 100 2 10.5<br />

3.5 60 0.4 8.9 100 3 13.5<br />

4.0 80 0.4 10.8 100 4 16.5<br />

225


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

Table 7.30: case study II: least-weight solutions, s<strong>and</strong>wich panels <strong>with</strong> steel faces versus<br />

s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

steel faces IPC composite faces<br />

span core face weight core # face weight<br />

(m) (mm) (mm) (kg/m²) (mm) layers (kg/m²)<br />

2.0 40 0.4 8.0 60 1 5.7<br />

2.5 40 0.4 8.0 100 1 7.5<br />

3.0 40 0.4 8.0 80 2 9.5<br />

3.5 60 0.4 8.9 100 2 10.5<br />

4.0 60 0.4 8.9 100 3 13.5<br />

Table 7.31: case study III: least-weight solutions, s<strong>and</strong>wich panels <strong>with</strong> steel faces versus<br />

s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

steel faces IPC composite faces<br />

span core face weight core # face weight<br />

(m) (mm) (mm) (kg/m²) (mm) layers (kg/m²)<br />

2.0 40 0.4 8.0 40 1 4.8<br />

2.5 40 0.4 8.0 60 1 5.7<br />

3.0 40 0.4 8.0 80 1 6.6<br />

3.5 40 0.4 8.0 100 1 7.5<br />

4.0 40 0.4 8.0 80 2 9.5<br />

7.9.2 conclusions<br />

From tables 7.29 to 7.31 it can be seen that the least-weight solutions for panels<br />

<strong>with</strong> steel faces <strong>and</strong> span between 2m <strong>and</strong> 3m are on the average as heavy as the<br />

panels <strong>with</strong> IPC faces. The reason can be found in the fact that for s<strong>and</strong>wich<br />

panels <strong>with</strong> steel faces, a smaller face thickness than 0.4mm is usually not<br />

produced. These panels are thus over-designed. If the span increases, s<strong>and</strong>wich<br />

panels <strong>with</strong> steel faces provide higher stiffness to weight ratio than s<strong>and</strong>wich<br />

panels <strong>with</strong> IPC faces. Still s<strong>and</strong>wich panels <strong>with</strong> IPC faces are performing<br />

relatively well. The required face thickness is definitely higher if IPC composite<br />

faces are used, but the density <strong>of</strong> steel is four times higher than the density <strong>of</strong> IPC.<br />

7.10 Conclusions<br />

In this chapter several test cases were studied to illustrate the behaviour <strong>of</strong><br />

s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces for building purposes under external<br />

loads. The introduction <strong>of</strong> thermal loads has also been discussed.<br />

It has been demonstrated that the design code <strong>of</strong> s<strong>and</strong>wich panels, as it is used in<br />

the Preliminary Recommendations for S<strong>and</strong>wich <strong>Panels</strong> <strong>and</strong> the Updated<br />

Recommendations for S<strong>and</strong>wich <strong>Panels</strong>, should be applied differently for<br />

s<strong>and</strong>wich panel <strong>with</strong> steel faces <strong>and</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces.<br />

226


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

Two slight modifications are used for the design <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC<br />

faces, compared to s<strong>and</strong>wich panels <strong>with</strong> steel faces:<br />

1. concerning unloading from CLC (characteristic load combination) to<br />

FLC (frequent load combination): The non-linear behaviour <strong>of</strong> IPC composite<br />

faces in tension is allowed in service conditions <strong>and</strong> this requires modifications in<br />

the interpretation <strong>of</strong> the characteristic load combination in the serviceability limit<br />

state. The modified requirement for panels <strong>with</strong> IPC composite faces is that<br />

residual deflections should not exceed 1/200 th <strong>of</strong> the span under frequent load<br />

combination. This requirement is to be satisfied, even when the characteristic load<br />

combination has already been applied to the panel before. This requirement may<br />

force the designer to reconsider the solution, previously obtained under the<br />

frequent combination <strong>of</strong> loads. Examples <strong>of</strong> such redesign are given in this<br />

chapter. Of all studies examples about 50% <strong>of</strong> the panels had to be redesigned<br />

under this condition.<br />

2. concerning repeated loading up to FLC (frequent load combination):<br />

Since damage introduction is allowed in the behaviour <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong><br />

IPC faces, the behaviour <strong>of</strong> these panels under repeated loading should be h<strong>and</strong>led<br />

<strong>with</strong> great care. A requirement should therefore be verified: the deflections are not<br />

to exceed 1/200 th <strong>of</strong> the span under repeated occurrence <strong>of</strong> the frequent load<br />

combination. Some examples are given in this chapter, where reconsideration <strong>of</strong><br />

the design <strong>of</strong> the s<strong>and</strong>wich panels is required under repeated loading. Of all<br />

studied examples only a few panels had to be redesigned under this condition.<br />

Compared to s<strong>and</strong>wich panels <strong>with</strong> steel faces, the stiffness to weight ratio <strong>of</strong> IPC<br />

faces is still relatively good. The required face thickness is definitely higher if IPC<br />

composite faces are used, but the density <strong>of</strong> steel is four times higher than the<br />

density <strong>of</strong> IPC.<br />

7.11 References<br />

B.K. Ahn <strong>and</strong> W.A Curtin, Strain <strong>and</strong> hysteresis by stochastic matrix<br />

cracking in ceramic matrix composites, J. Mech. Phys. Solids, Vol. 45, No. 2,<br />

1997 pp.177-209<br />

H. G. Allen, <strong>Analysis</strong> <strong>and</strong> <strong>Design</strong> <strong>of</strong> Structural S<strong>and</strong>wich <strong>Panels</strong>, Permagon<br />

Press, 1969<br />

ANSYS manuals, Release 5.5, 1998<br />

227


Chapter 7: <strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces<br />

K. Berner, J.M. Davies, P. Hassinen, L. Heselius, Updated European<br />

recommendations for s<strong>and</strong>wich panels, Proceedings 5 th International Conference<br />

on S<strong>and</strong>wich Construction, EMAS Publishing, Sept 5-7, 2000,pp. 389-400<br />

A.G. Davenport, The estimation <strong>of</strong> load repetitions on structures <strong>with</strong><br />

application to wind induced fatigue <strong>and</strong> overload, Proceedings RILEM<br />

International Symposium on the Effect <strong>of</strong> Repeated Loading <strong>of</strong> Materials <strong>and</strong><br />

Structures, Mexico City, 1966<br />

ECCS publication, Preliminary European Recommendations for S<strong>and</strong>wich<br />

<strong>Panels</strong>, part I, <strong>Design</strong>, Technical Committee 7, Working Group 7.4, 1991<br />

Eurocode 1, Grondslag voor ontwerp en belastingen op draagsystemen<br />

Eurocode 1, Windbelasting op bouwwerken, winddruk op een w<strong>and</strong> en<br />

gezamelijke windeffecten op bouwwerken. NBN B 03-002-1, 1988<br />

D. Zenkert, An introduction to s<strong>and</strong>wich construction, Chameleon Press<br />

Ltd, 1995<br />

Technical sheets, (Belgian) producers <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> steel faces:<br />

- Color Pr<strong>of</strong>il<br />

- Europan<br />

- Evopan<br />

- Fischer pr<strong>of</strong>il<br />

- Haironville<br />

- Hoesch<br />

- Isobouw<br />

- Isocab<br />

- Isometall System<br />

- Isopan<br />

- Kingspan<br />

- Opstalan PAB Nord<br />

- Perma Systems<br />

- SBC Holl<strong>and</strong><br />

- Unidek<br />

- UnilinVerpac<br />

- Wolvega<br />

228


Chapter 8<br />

Conclusions<br />

The main goal <strong>of</strong> this thesis is the proposition <strong>of</strong> a design methodology for<br />

s<strong>and</strong>wich panels <strong>with</strong> cementitious composite faces for building purposes. The<br />

studied face material in this work is a composite, which is made <strong>of</strong> inorganic<br />

phosphate cement (IPC) <strong>and</strong> reinforced <strong>with</strong> E-glass fibres. The advantage <strong>of</strong> IPC,<br />

compared to other cementitious mixtures, is the non-alkaline environment. Due to<br />

this, the matrix does not attack the E-glass fibres <strong>and</strong> the production <strong>of</strong> costeffective<br />

cementitious composites becomes a possibility. Existing publications on<br />

the design <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> more classical face materials are used as<br />

reference documents for design.<br />

Conclusions are formulated on three levels. The first level implies the formulation<br />

<strong>of</strong> the stress-strain behaviour <strong>of</strong> E-glass fibre reinforced IPC (inorganic phosphate<br />

cement) specimens under monotonic loading, unloading <strong>and</strong> repeated loading. The<br />

second level introduces the implementation <strong>of</strong> the face behaviour into s<strong>and</strong>wich<br />

finite element modelling <strong>and</strong> the comparison <strong>of</strong> this modelling <strong>with</strong> experimental<br />

observations. The formulation <strong>of</strong> s<strong>and</strong>wich design criteria for s<strong>and</strong>wich panels<br />

<strong>with</strong> IPC composite faces for building purposes is the final step, which includes<br />

the implementation <strong>of</strong> the two previous levels.<br />

8.1 Conclusions on the behaviour <strong>of</strong> IPC composite<br />

specimens<br />

The appropriateness <strong>of</strong> two models, describing the stress-strain behaviour <strong>of</strong> Eglass<br />

fibre reinforced IPC in tension has been presented <strong>and</strong> discussed. Loading,<br />

unloading <strong>and</strong> repeated loading <strong>of</strong> laminates is studied. The first model (referred<br />

to as the ACK based model or the Aveston-Cooper-Kelly model) assumes the<br />

matrix has a unique matrix failure stress. The second model (the stochastic<br />

cracking based model) implements the statistical scatter <strong>of</strong> the matrix failure<br />

stress. Both models presume the fibres provide further stiffness <strong>and</strong> strength, once<br />

intensive introduction <strong>and</strong> propagation <strong>of</strong> matrix cracks occurred. If a matrix crack<br />

229


Chapter 8: Conclusions<br />

is introduced, it is assumed in both models that a frictional matrix-fibre interface<br />

stress transfer mechanism replaces the previously elastic matrix-fibre bond in the<br />

vicinity <strong>of</strong> this matrix crack.<br />

The two types <strong>of</strong> reinforcement, which are studied in detail, are a unidirectional<br />

reinforcement <strong>and</strong> 2D-r<strong>and</strong>omly oriented short fibre reinforcement.<br />

Loading <strong>of</strong> unidirectionally reinforced IPC specimens is discussed in terms <strong>of</strong> the<br />

ACK theory <strong>and</strong> the stochastic cracking theory (see Chapter 2). It is verified that<br />

the behaviour <strong>of</strong> E-glass fibre reinforced IPC in compression is linear elastic up to<br />

failure. The tensile behaviour <strong>of</strong> IPC composites is highly non-linear. In general,<br />

three apparent zones are observed in the stress-strain curve under monotonic<br />

tensile loading: the pre-cracking zone (linear elastic behaviour <strong>of</strong> the material), the<br />

multiple cracking zone (crack initiation <strong>and</strong> propagation in the matrix) <strong>and</strong> the<br />

post-cracking zone (fibres provide further stiffness). From comparison <strong>of</strong><br />

experimental <strong>and</strong> theoretical stress-strain curves for UD-reinforced IPC<br />

composites, it is concluded that the presented models describe the ongoing mesomechanics<br />

rather well. Both models are modified theoretically for application on<br />

2D-r<strong>and</strong>omly reinforced specimens. Both the modified ACK based model <strong>and</strong> the<br />

modified stochastic cracking based model provide rather good prediction <strong>of</strong> the<br />

stress-strain curves. This indicates that both models are appropriate to describe the<br />

meso-mechanics <strong>of</strong> IPC composites in tension <strong>with</strong> more complex reinforcement<br />

than the unidirectional reinforcement, which is generally studied <strong>and</strong> discussed in<br />

literature. The main advantage <strong>of</strong> the stochastic cracking theory is found in a better<br />

prediction <strong>of</strong> the stress-strain relationship around the theoretical ACK multiple<br />

cracking stress, since it includes the existing stochastic nature <strong>of</strong> the matrix tensile<br />

strength, whereas all multiple cracking occurred at one stress level in the ACK<br />

theory.<br />

The prediction <strong>of</strong> the unloading behaviour <strong>of</strong> unidirectionally reinforced <strong>and</strong> 2Dr<strong>and</strong>omly<br />

reinforced specimens is based on the modelling <strong>of</strong> these composites<br />

under monotonic tensile loading. It is examined in Chapter 3 whether these<br />

unloading models provide good accuracy on the prediction <strong>of</strong> experiments. The<br />

ACK-based unloading model provides satisfactory results at high stress levels<br />

only, situated in the post-cracking zone (full multiple cracking occurred). The<br />

stochastic cracking based theory provides good accuracy in the prediction <strong>of</strong><br />

unloading for all stress levels.<br />

The damage model, presented in Chapter 4, is based on an extended stochastic<br />

cracking (based) theory for unidirectionally reinforced composites, when loaded<br />

repeatedly in tension. The main hypotheses, which are used in the formulation <strong>of</strong> a<br />

description <strong>of</strong> the loss <strong>of</strong> stiffness <strong>of</strong> IPC composite specimens under repeated<br />

tensile loading, are: (1) matrix cracking is the main damage mechanism, occurring<br />

230


Chapter 8: Conclusions<br />

at first loading, but extra matrix cracking, introduced during repeated loading is<br />

neglected <strong>and</strong> (2) the dominant damage mechanism during further repeated<br />

loading, after first loading occurred, is matrix-fibre interface degradation. The<br />

presented model can be used to predict extra deformations under repeated tensile<br />

loading for unidirectionally <strong>and</strong> 2D-r<strong>and</strong>omly reinforced composite specimens.<br />

It is verified in Chapter 4 that a model, which implements these hypotheses in a<br />

stochastic cracking model, as presented in previous chapters, is an appropriate<br />

model for the prediction <strong>of</strong> the behaviour <strong>of</strong> unidirectionally reinforced IPC<br />

composites under repeated tensile loading.<br />

For 2D-r<strong>and</strong>omly reinforced IPC specimens, loaded repeatedly up to a stress at or<br />

below the theoretical ACK multiple cracking stress (in the pre-cracking or<br />

multiple zone), it has been noticed that extra matrix cracking occurred upon<br />

cycling, which cannot be predicted by the presented model <strong>and</strong> is not described<br />

quantitatively in this thesis. However, this effect only occurs in the first one<br />

hundred to one thous<strong>and</strong> cycles. <strong>Matrix</strong>-fibre interface degradation is still the sole<br />

damage mechanism for 2D-r<strong>and</strong>omly short fibre reinforced IPC composites, if the<br />

maximum stress is situated below or around the multiple cracking zone <strong>and</strong><br />

already one hundred to one thous<strong>and</strong> load cycles have been applied. Due to the<br />

more complex stress fields in 2D-r<strong>and</strong>omly short fibre (50mm length) reinforced<br />

composites, fibre pull-out is a major damage mechanism at higher stress levels, far<br />

beyond the theoretical ACK multiple cracking stress. Eventually, progressing <strong>of</strong><br />

fibre pull-out under repeated loading leads to failure <strong>of</strong> the composite specimens.<br />

This is noticed from experimental observations, where failure occurred after<br />

several thous<strong>and</strong>s <strong>of</strong> load cycles, provided the maximum cycle stress is situated<br />

well beyond the multiple cracking stress.<br />

8.2 Conclusions on s<strong>and</strong>wich modelling<br />

In Chapter 5, it is verified that choosing an appropriate model can be rather<br />

delicate in the case <strong>of</strong> the studied panels. Although the conditions are fulfilled for<br />

the application <strong>of</strong> the well-known elementary s<strong>and</strong>wich theory (EST) when all<br />

materials still behave linear elastically, it might be necessary that more complex<br />

s<strong>and</strong>wich models are to be used, once multiple cracking occurs in the faces.<br />

A finite element package, ANSYS, is used in this work to calculate the behaviour<br />

<strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces. Two ANSYS material models are<br />

discussed on their feasibility to describe the behaviour <strong>of</strong> IPC composites under<br />

monotonic loading: the “multilinear elastic” <strong>and</strong> the “aniso” material behaviour.<br />

The “aniso” material behaviour is a bilinear stress-strain model, including<br />

different stress-strain behaviour in three perpendicular directions. It also takes into<br />

231


Chapter 8: Conclusions<br />

account possible different behaviour in tension <strong>and</strong> in compression. If the<br />

“multilinear elastic” model is used, up to 100 stress-strain points are used to<br />

represent a stress-strain curve. The disadvantage <strong>of</strong> this model is that the<br />

compressive <strong>and</strong> tensile behaviour are assumed to be equal. Conclusively, the<br />

“multilinear elastic” material behaviour is to be used if it is clear a priori which<br />

parts <strong>of</strong> the faces are stressed in tension. This “multilinear elastic” model can be<br />

used to represent the behaviour <strong>of</strong> regions in the faces, loaded in tension. This<br />

material option approximates the stress-strain behaviour <strong>of</strong> IPC composites in<br />

tension, as modelled by the stochastic cracking theory in Chapter 2, in a more<br />

accurate way than the “aniso” material option. The “aniso” material option has the<br />

advantage that no a priori knowledge <strong>of</strong> the sign <strong>of</strong> the normal stresses is needed.<br />

In Chapter 5 a methodology is discussed on the implementation <strong>of</strong> the unloading<br />

material behaviour <strong>of</strong> the IPC composite faces in s<strong>and</strong>wich calculations. This<br />

methodology is based on the stochastic cracking based model for unloading in<br />

Chapter 3. A macro has been written in ANSYS for this purpose, since the IPC<br />

composite unloading material behaviour does not come near any st<strong>and</strong>ard material<br />

unloading law <strong>of</strong> any commercial finite element package. Another macro is<br />

written for the implementation <strong>of</strong> the behaviour <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC<br />

composite faces under repeated loading.<br />

In Chapter 6 experimental observations on s<strong>and</strong>wich panels <strong>with</strong> IPC faces are<br />

compared to theoretical predictions. The length <strong>of</strong> the tested s<strong>and</strong>wich panels is<br />

2m. The thicknesses <strong>of</strong> core <strong>and</strong> faces are varied from one panel to another. The<br />

theoretical calculation <strong>of</strong> loading, unloading <strong>and</strong> repeated loading <strong>of</strong> the tested 2m<br />

long s<strong>and</strong>wich panels is based on the methodology explained in Chapter 5. From<br />

comparison <strong>of</strong> theoretical <strong>with</strong> experimentally obtained force-deflection curves,<br />

following conclusions can be formulated:<br />

- Loading: almost all theoretical force-deflection curves <strong>of</strong> the panels,<br />

loaded in four-point bending, show good coincidence <strong>with</strong> the experimentally<br />

obtained curves. If linear elastic face behaviour is used in the finite element<br />

calculations, the deflections are systematically <strong>and</strong> seriously underestimated. It is<br />

thus verified that implementation <strong>of</strong> the stress-strain behaviour, as represented by<br />

the stochastic cracking based theory in Chapter 2, in the tensile face is a necessity<br />

for correct analysis <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces.<br />

- Unloading: the residual deflections <strong>of</strong> almost all panels were predicted<br />

<strong>with</strong> fair accuracy.<br />

- Repeated loading: comparison <strong>of</strong> test results <strong>and</strong> theoretical predictions<br />

reveals that the presented methodology provides an estimation <strong>of</strong> the additional<br />

deflections <strong>of</strong> the s<strong>and</strong>wich under repeated loading. The advantage <strong>of</strong> the poposed<br />

methodology is that it is not very computational time-consuming <strong>and</strong> the<br />

implementation into finite element calculations is rather easy. Prediction <strong>of</strong><br />

additional strains under repeated loading is less accurate, but the comparison<br />

232


Chapter 8: Conclusions<br />

between theoretical <strong>and</strong> experimental load-deflection behaviour revealed<br />

prediction <strong>of</strong> additional deflections is fair. (see Chapter 6)<br />

8.3 Conclusions on design criteria for s<strong>and</strong>wich panels <strong>with</strong><br />

IPC composite faces for building purposes<br />

In Chapter 7, design <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces is studied <strong>and</strong><br />

discussed. This design is mainly based on three documents: Eurocode 1, the<br />

Preliminary Recommendations for S<strong>and</strong>wich <strong>Panels</strong> <strong>and</strong> the Updated<br />

Recommendations for S<strong>and</strong>wich <strong>Panels</strong>.<br />

Three case studies are discussed: a simply supported wall panel, a simply<br />

supported ro<strong>of</strong> panel <strong>and</strong> a ro<strong>of</strong> panel <strong>with</strong> an intermediate support.<br />

The external loads, which are applied to the s<strong>and</strong>wich panels in this work, are own<br />

weight, snow, wind <strong>and</strong> temperature. Three load combinations, defined in<br />

Eurocode 1 <strong>and</strong> the Recommendations for s<strong>and</strong>wich panels are applied on the<br />

studied panels.<br />

1. The “frequent load combination” is supposed to be experienced a couple<br />

<strong>of</strong> times or during a certain period in the lifetime. The condition to be fulfilled in<br />

the Recommendations is that the maximum deflection <strong>of</strong> the panel does not exceed<br />

1/200 th <strong>of</strong> the span length under this load combination (this is a condition under<br />

serviceability limit state).<br />

2. The “characteristic load combination” is rare <strong>and</strong> might possibly be<br />

experienced by the construction during the lifetime. The characteristic load<br />

combination is higher than the frequent load combination. The effects studied<br />

under characteristic load combination are effects, which influence the structure by<br />

their one-time occurrence. If a classical s<strong>and</strong>wich panel <strong>with</strong> steel faces is<br />

analysed, the s<strong>and</strong>wich layers should not yield or buckle under this load<br />

combination (serviceability limit state).<br />

3. The third load combination takes into account extra safety factors on the<br />

loads - is therefore higher than the characteristic load combination - <strong>and</strong> is used to<br />

check failure <strong>of</strong> the panel. (ultimate limit state).<br />

In Chapter 7 it is verified for all case studies how a least-weight solution can be<br />

found, still not exceeding the defined limitations on the deflections (serviceability<br />

limit state) <strong>and</strong> preventing failure <strong>of</strong> any component <strong>of</strong> the panel (ultimate limit<br />

state).<br />

<strong>Design</strong> <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC faces is compared <strong>with</strong> design <strong>of</strong> s<strong>and</strong>wich<br />

panels <strong>with</strong> steel faces in Chapter 7.<br />

233


Chapter 8: Conclusions<br />

It is verified that the design <strong>of</strong> s<strong>and</strong>wich panels, as recommended in the<br />

Preliminary Recommendations for S<strong>and</strong>wich <strong>Panels</strong> <strong>and</strong> the Updated<br />

Recommendations for S<strong>and</strong>wich <strong>Panels</strong> should be interpreted in a different way<br />

for s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces, compared to the commonly used<br />

s<strong>and</strong>wich panels <strong>with</strong> steel faces: the implementation <strong>of</strong> the non-linear behaviour<br />

<strong>of</strong> IPC composite faces requires modifications in the interpretation <strong>of</strong> the<br />

serviceability limit state.<br />

1. A modified requirement for panels <strong>with</strong> IPC composite faces is that<br />

residual deflections should not exceed 1/200 th <strong>of</strong> the span under the frequent load<br />

combination, even when the characteristic load combination already occurred in<br />

the history <strong>of</strong> the panel.<br />

2. Serious accumulation <strong>of</strong> strains due to repeated loading does hardly<br />

occur in steel faces, but is shown to be <strong>of</strong> importance in IPC composite faces.<br />

Deflections <strong>of</strong> s<strong>and</strong>wich panels <strong>with</strong> IPC composite faces are not to exceed 1/200 th<br />

<strong>of</strong> the span, even after a high number <strong>of</strong> applied load cycles occurred in the<br />

history <strong>of</strong> the panel.<br />

8.4 Future research topics<br />

Concerning the behaviour <strong>of</strong> IPC composite laminates, the presented models on<br />

loading <strong>and</strong> unloading are appropriate to be used in design calculations. The<br />

behaviour <strong>of</strong> specimens under repeated loading is still to be discussed further.<br />

Especially the behaviour <strong>of</strong> 2D-r<strong>and</strong>omly short fibre reinforced specimens under<br />

repeated loading should be studied more in detail, if this type <strong>of</strong> reinforcement is<br />

to be used.<br />

In this work a simplified methodology is used to predict the extra s<strong>and</strong>wich panel<br />

deflections due to repeated loading. The implementation <strong>of</strong> accumulated face<br />

damage in a s<strong>and</strong>wich panel under repeated loading is an interesting <strong>and</strong> complex<br />

subject. In reality, redistribution <strong>of</strong> stresses occurs continuously under repeated<br />

loading. Extensive study <strong>of</strong> algorithms, which include this redistribution <strong>of</strong><br />

stresses <strong>with</strong> the number <strong>of</strong> elapsed cycles for the studied panels, is a very<br />

interesting potential subject.<br />

On the composite level, the influence <strong>of</strong> several material parameters on the<br />

behaviour <strong>of</strong> IPC composites is still to be studied. The use <strong>of</strong> modifications on<br />

fibre seizing, matrix components, the use <strong>of</strong> fillers, additives, etc. might change<br />

the behaviour <strong>of</strong> the composite in quite an important manner. This might even<br />

have large consequences on the design philosophy, presented here for s<strong>and</strong>wich<br />

panels <strong>with</strong> IPC composite faces.<br />

234


Chapter 8: Conclusions<br />

A new design philosophy is introduced in this work for s<strong>and</strong>wich panels <strong>with</strong><br />

brittle composite faces for building purposes. Focus has been put on the<br />

introduction <strong>of</strong> mechanical loads. Study <strong>of</strong> the durability <strong>of</strong> these panels is actually<br />

as important as the resistance to mechanical loading: degradation <strong>of</strong> the faces, the<br />

core or core-face interface under freezing-thawing, UV radiation, water ingress,<br />

chemical attack <strong>and</strong> fungi can be important. Water ingress is much easier, once<br />

multiple cracking <strong>of</strong> the faces occurred. An extra condition may be defined under<br />

the serviceability limit state, limiting the opening <strong>of</strong> matrix cracks.<br />

235


Appendix 1<br />

A1.1 The Normal model<br />

Probability distribution functions<br />

The Normal model (or Gaussian model) is frequently used. The probability<br />

distribution function is:<br />

2<br />

1 ⎡<br />

( )<br />

( x − µ ) ⎤<br />

f x = exp⎢−<br />

2 ⎥⎦<br />

σ 2π<br />

⎣ 2σ<br />

(A1.1)<br />

where: -∞ < x < +∞<br />

-∞ < µ < +∞, <strong>with</strong> µ being a location parameter<br />

0 < σ, <strong>with</strong> σ being a scale parameter<br />

Figure A1.1a <strong>and</strong> figure A1.1b illustrate the influence <strong>of</strong> location parameter µ <strong>and</strong><br />

scale<br />

parameter σ respectively.<br />

0.2<br />

µ = 1, σ = 1<br />

0.35<br />

0.15<br />

µ = 2, σ = 1<br />

0.3<br />

0.25<br />

µ = 0, σ = 0.5<br />

0.2<br />

0.1<br />

µ = 3, σ = 1<br />

0.15 µ = 0, σ = 1.0<br />

0.05<br />

0.1<br />

0.05<br />

µ = 0, σ = 2.0<br />

0<br />

0<br />

-4<br />

-2 0 2<br />

x<br />

4 6 8 -8 -4 0<br />

x<br />

4 8<br />

Figure A1.1a: Normal model, influence <strong>of</strong> µ Figure A1.1b: Normal model, influence <strong>of</strong> σ<br />

f(x)<br />

A1.2 The Cauchy model<br />

The Cauchy probability distribution function is formulated by equation (A1.2):<br />

f(x)


where:<br />

1<br />

f ( x)<br />

= 2<br />

⎡ ⎛ x − µ ⎞ ⎤<br />

⎢1<br />

+ ⎜ ⎟ ⎥σπ<br />

⎢⎣<br />

⎝ σ ⎠ ⎥⎦<br />

-∞ < x < +∞<br />

-∞ < µ < +∞, <strong>with</strong> µ being a location parameter<br />

0 < σ, <strong>with</strong> σ being a scale parameter<br />

(A1.2)<br />

Figure A1.2a <strong>and</strong> figure A1.2b illustrate the influence <strong>of</strong> location parameter µ <strong>and</strong><br />

scale parameter σ respectively.<br />

-8<br />

f(x)<br />

µ = 1, σ = 1<br />

0.35<br />

0.3<br />

0.25<br />

µ = 2, σ = 1<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

µ = 3, σ = 1<br />

-4 0<br />

x<br />

4 8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

µ = 0, σ = 0.5<br />

0.3<br />

0.2<br />

µ = 0, σ = 1<br />

0.1<br />

0<br />

µ = 0, σ = 2<br />

-8 -4 0<br />

x<br />

4 8<br />

Figure A1.2a: Cauchy model, influence <strong>of</strong> µ Figure A1.2b: Cauchy model, influence <strong>of</strong> σ<br />

A1.3 The Lognormal model<br />

The lognormal probability distribution model is used on a r<strong>and</strong>om variable whose<br />

logarithm follows a normal distribution function. The formulation <strong>of</strong> the<br />

lognormal model is:<br />

f ( x)<br />

=<br />

2<br />

1 ⎡ ( ln x − µ ) ⎤<br />

exp⎢−<br />

2 ⎥⎦ 0 < x<br />

xσ<br />

2π<br />

⎣ 2σ<br />

(A1.3)<br />

where:<br />

f ( x)<br />

= 0<br />

x ≤ 0<br />

-∞ < x < +∞<br />

-∞ < µ < +∞, <strong>with</strong> µ being a scale parameter<br />

0 < σ, <strong>with</strong> σ being a shape parameter<br />

Figure A1.3a <strong>and</strong> figure A1.3b illustrate the influence <strong>of</strong> scale parameter µ <strong>and</strong><br />

shape parameter σ respectively.<br />

f(x)


f(x)<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

µ = 1, σ = 1<br />

µ = 3, σ = 1<br />

µ = 2, σ = 1<br />

0 2 4 6 8 10<br />

x<br />

f(x)<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

µ = 1, σ = 2<br />

µ = 1, σ = 1<br />

µ = 1, σ = 0.5<br />

0 2 4 6 8 10<br />

x<br />

Figure A1.3a: Lognormal model, influence <strong>of</strong> µ Figure A1.3b: Lognormal model, influence <strong>of</strong> σ<br />

A1.4 The Weibull model<br />

Failure <strong>of</strong> a component may be linked to phenomena dependent on the largest or<br />

smallest value in a sample from a particular distribution. Examples <strong>of</strong> distribution<br />

functions, which are based on this principle, are the Weibull model, the Largest<br />

extreme value model <strong>and</strong> the Smallest extreme value model.<br />

A Weibull probability distribution may include two or three parameters. The twoparameter<br />

Weibull model is formulated as follows:<br />

η −1 η<br />

η ⎛ x ⎞ ⎡ ⎛ x ⎞ ⎤<br />

f ( x)<br />

= ⎜ ⎟ exp⎢−<br />

⎜ ⎟ ⎥ 0 ≤ x<br />

σ ⎝σ<br />

⎠ ⎢⎣<br />

⎝σ<br />

⎠ ⎥⎦<br />

f ( x)<br />

= 0<br />

x < 0<br />

(A1.4)<br />

where: -∞ < x < +∞<br />

0 < σ, <strong>with</strong> σ being a scale parameter<br />

0 < η, <strong>with</strong> η being a shape parameter<br />

The three-parameter model is formulated:<br />

η −1 η<br />

η ⎛ x − µ ⎞ ⎡ ⎛ x − µ ⎞ ⎤<br />

f ( x)<br />

= ⎜ ⎟ exp⎢−<br />

⎜ ⎟ ⎥ µ ≤ x<br />

σ ⎝ σ ⎠ ⎢⎣<br />

⎝ σ ⎠ ⎥⎦<br />

f ( x)<br />

= 0<br />

x < µ<br />

where: -∞ < x < +∞<br />

-∞ < µ < +∞, <strong>with</strong> µ being a location parameter<br />

0 < σ, <strong>with</strong> σ being a scale parameter<br />

0 < η, <strong>with</strong> η being a shape parameter<br />

(A1.5)<br />

Figure A1.4a, figure A1.4b <strong>and</strong> figure A1.4c illustrate the influence <strong>of</strong> shape<br />

parameter η, scale parameter σ <strong>and</strong> location parameter µ respectively.


f(x)<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

µ = 2, σ = 5, η = 5<br />

µ = 2, σ = 5,<br />

η = 4<br />

0 2 4 6<br />

x<br />

µ = 2, σ = 5,<br />

η = 3<br />

8<br />

f(x)<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

µ = 2, σ = 5, η = 3<br />

µ = 2, σ = 6, η = 3<br />

µ = 2, σ = 7, η = 3<br />

0 2 4 6 8 10 12 14 16<br />

x<br />

Figure A1.4a: Weibull model, influence <strong>of</strong> η Figure A1.4b: Weibull model, influence <strong>of</strong> σ<br />

f(x)<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

µ = 2,<br />

σ = 5,<br />

η = 3<br />

A1.5 The Gamma model<br />

µ = 3, σ = 5, η = 3<br />

µ = 4, σ = 5, η = 3<br />

0 2 4 6 8 10 12 14 16<br />

x<br />

Figure A1.4c: Weibull model, influence <strong>of</strong> µ<br />

The Gamma distribution is frequently used to describe r<strong>and</strong>om variables, which<br />

are bounded at one end<br />

f<br />

( x)<br />

=<br />

Γ(<br />

η)<br />

( x)<br />

= 0<br />

η<br />

λ η −1<br />

−λx<br />

x<br />

e<br />

0 ≤ x<br />

f x < 0<br />

where: -∞ < x < +∞<br />

0 < λ, <strong>with</strong> λ being a scale parameter<br />

0 < η, <strong>with</strong> η being a shape parameter<br />

Γ( η)<br />

=<br />

∞<br />

∫<br />

0<br />

x<br />

e<br />

η −1<br />

− x<br />

dx<br />

(A1.6)<br />

Figure A1.5a <strong>and</strong> figure A1.5b illustrate the influence <strong>of</strong> shape parameter η <strong>and</strong><br />

scale parameter λ respectively.


f(x)<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

λ = 1, η = 1<br />

λ = 1, η = 1.5<br />

λ = 1, η = 2<br />

0 2 4 6<br />

x<br />

f(x)<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

λ = 0.5, η = 1.5<br />

λ = 1.5, η = 1.5<br />

λ = 1, η = 1.5<br />

0 2 4 6<br />

x<br />

Figure A1.5a: Gamma model, influence <strong>of</strong> η Figure A1.5b: Gamma model, influence <strong>of</strong> λ<br />

A1.6 The Largest extreme value model (LEV)<br />

Failure <strong>of</strong> a component may be linked to phenomena dependent on the largest or<br />

smallest value in a sample from a particular distribution. Examples <strong>of</strong> distribution<br />

functions, which are based on this principle, are the Weibull model, the Largest<br />

extreme value model <strong>and</strong> the Smallest extreme value model. The Largest extreme<br />

value model is defined as:<br />

1 ⎡ 1<br />

⎛ 1 ⎞⎤<br />

f ( x)<br />

= exp⎢−<br />

( x − µ ) − exp ⎜<br />

⎜−<br />

( x − µ ) ⎟<br />

⎟⎥<br />

(A1.7)<br />

η ⎣ η<br />

⎝ η ⎠⎦<br />

where: -∞ < µ < +∞<br />

0 < η<br />

Figure A1.6a <strong>and</strong> figure A1.6b illustrate the influence <strong>of</strong> µ <strong>and</strong> η respectively.<br />

-4<br />

f(x)<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

µ = 1, η = 1<br />

µ = 2, η = 1<br />

µ = 3, η = 1<br />

0<br />

-2 0 2<br />

x<br />

4 6 8 10<br />

f(x)<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

µ = 1, η = 1<br />

µ = 1, η = 2<br />

µ = 1, η = 3<br />

-4 -2 0 2 4 6 8<br />

x<br />

Figure A1.6a: LEV model, influence <strong>of</strong> µ Figure A1.6b: LEV model, influence <strong>of</strong> η


A1.7 The Smallest extreme value model (SEV)<br />

Failure <strong>of</strong> a component may be linked to phenomena dependent on the largest or<br />

smallest value in a sample from a particular distribution. Examples <strong>of</strong> distribution<br />

functions, which are based on this principle, are the Weibull model, the Largest<br />

extreme value model <strong>and</strong> the Smallest extreme value model. The Smallest extreme<br />

value model is defined as:<br />

1 ⎡1<br />

⎛ 1 ⎞⎤<br />

f ( x)<br />

= exp⎢<br />

( x − µ ) − exp ⎜ ( x − µ ) ⎟<br />

⎟⎥<br />

(A1.8)<br />

η ⎣η<br />

⎝η<br />

⎠⎦<br />

where: -∞ < µ < +∞<br />

0 < η<br />

Figure A1.7a <strong>and</strong> figure A1.7b illustrate the influence <strong>of</strong> µ <strong>and</strong> η respectively.<br />

f(x)<br />

µ = 1, η = 1<br />

0.4 µ = 2, η = 1<br />

0.3<br />

0.2<br />

0.1<br />

µ = 3, η = 1<br />

0<br />

-6 -4 -2 0<br />

x<br />

2 4 6<br />

f(x)<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

µ = 1, η = 1<br />

µ = 1, η = 2<br />

µ = 1, η = 3<br />

-5 -3<br />

0<br />

-1 1<br />

x<br />

3 5 7<br />

Figure A1.7a: LEV model, influence <strong>of</strong> µ Figure A1.7b: LEV model, influence <strong>of</strong> η


Appendix 2<br />

A2.1 Introduction<br />

Probability plotting<br />

When one (several) probability distribution function(s) is (are) chosen, the<br />

parameters <strong>of</strong> this (these) function(s) should be obtained. The model parameters<br />

achieving “best agreement” between experimental data <strong>and</strong> chosen model are<br />

called “best-fit” parameters. Several definitions <strong>of</strong> “best agreement” are possible.<br />

Therefore several methodologies finding a “best agreement” can be used. The<br />

method, which is used here, is one <strong>of</strong> the more common <strong>and</strong> simple methods:<br />

probability plotting.<br />

Probability plotting modifies the scale <strong>of</strong> the ordinate <strong>of</strong> the graph <strong>of</strong> the<br />

probability function (1-F(x)) against x in such a way that a straight line is<br />

obtained. When the experimental data are plotted in this curve, the linearity <strong>of</strong> the<br />

experimental curve is a measure for the appropriateness <strong>of</strong> the chosen model. The<br />

probability model parameters are obtained from determination <strong>of</strong> the slope <strong>and</strong><br />

intersect <strong>of</strong> the “best fit” line <strong>with</strong> the data. The probability plotting method is<br />

illustrated in this appendix, together <strong>with</strong> the obtained results on the IPC matrix<br />

bending strength for three types <strong>of</strong> models: the Lognormal model, the Weibull<br />

model <strong>and</strong> the Largest extreme value model. The bending strength (x) <strong>of</strong> 34 IPC<br />

matrix beam specimens is determined experimentally. Table A2.1 shows the<br />

results. These results are ranked in increasing order.<br />

Table A2.1: bending strength <strong>of</strong> IPC beam specimens<br />

n 1 2 3 4 5 6 7 8 9 10<br />

x (MPa) 7.25 8.65 8.93 9.02 9.34 9.51 9.75 9.8 9.92 9.93<br />

n 11 12 13 14 15 16 17 18 19 20<br />

x (MPa) 10.1 10.2 10.2 10.3 10.3 10.6 10.6 10.6 10.7 10.7<br />

n 21 22 23 24 25 26 27 28 29 30<br />

x (MPa) 10.8 11.0 11.1 11.4 11.4 11.4 11.5 11.7 11.8 12.1<br />

n 31 32 33 34<br />

x (MPa) 12.2 12.5 13.1 13.4


A2.2 The Lognormal model<br />

The experimentally obtained data are ranked. The x-axis represents ln(x). The yaxis<br />

represent a linearised function <strong>of</strong> F(x): G(x), <strong>with</strong><br />

2<br />

c0<br />

+ c1<br />

* t + c2<br />

* t<br />

G( x)<br />

= t −<br />

+ ε(<br />

p)<br />

2<br />

3<br />

(A2.1)<br />

1+<br />

d1<br />

* t + d2<br />

* t + d3<br />

* t<br />

where:<br />

c0 = 2.515517 d1 = 1.432788<br />

c1 = 0.802853 d2 = 0.189269<br />

c2 = 0.010328 d3 = 0.001308<br />

1 2<br />

⎧ 1 ⎫<br />

t = ⎨ln<br />

2 ⎬<br />

⎩ p ⎭<br />

(A2.2)<br />

<strong>and</strong><br />

ε ( p)<br />

< 4. 5e<br />

− 4<br />

(A2.3)<br />

<strong>with</strong>:<br />

p = 1− F(<br />

x)<br />

(A2.4)<br />

F(x) is determined as:<br />

n − 0.<br />

3<br />

F ( x)<br />

=<br />

N + 0.<br />

4<br />

(A2.5)<br />

where: N = total number <strong>of</strong> data points<br />

n = ranking number <strong>of</strong> the data point <strong>with</strong> value x<br />

If this linearisation is adapted on the data, obtained from three-point bending tests<br />

on the IPC specimens, the result is illustrated in figure A2.1. The grey curve<br />

shows the data points, the black curve shows the “best fit” line.<br />

G(x)<br />

3<br />

2<br />

1<br />

0<br />

-11.5<br />

2 2.5 3<br />

-2<br />

-3<br />

experimental<br />

Linear (experimental)<br />

ln(x)<br />

y = 7.1021x - 16.724<br />

R 2 = 0.9504<br />

Figure A2.1: linearised probability plot IPC data <strong>and</strong> Lognormal model<br />

From figure A2.1, the expression <strong>of</strong> G(x) can be formulated as:<br />

G ( x)<br />

= y = Ax + B<br />

(A2.6)


The value <strong>of</strong> A, B <strong>and</strong> the regression coefficient are printed on figure A2.1. Model<br />

parameters µ <strong>and</strong> σ are obtained from knowledge <strong>of</strong> A <strong>and</strong> B:<br />

1<br />

A =<br />

(A2.7)<br />

σ<br />

µ<br />

B = −<br />

(A2.8)<br />

σ<br />

The value <strong>of</strong> µ is thus 2.35 <strong>and</strong> the value <strong>of</strong> σ is 0.141.<br />

A2.3 The Weibull model<br />

For the two-parameter Weibull model, the value <strong>of</strong> the scale parameter σ <strong>and</strong><br />

shape parameter η can be obtained from the linearised probability curve. ln(x) is<br />

represented in the x-axis <strong>of</strong> the regression plot. The y-axis represents G(x), <strong>with</strong><br />

G(x) formulated by equation (A2.9):<br />

⎛ ⎛ 1 ⎞⎞<br />

G ( x)<br />

= ln⎜<br />

⎟<br />

⎜<br />

ln ⎜<br />

⎟<br />

⎟<br />

(A2.9)<br />

⎝ ⎝1<br />

− F(<br />

x)<br />

⎠⎠<br />

The experimental <strong>and</strong> the theoretical linearised two-parameter Weibull curve are<br />

plotted in figure A2.2.<br />

G(x)<br />

2<br />

-2<br />

-3<br />

-4<br />

-5<br />

experimental<br />

Chart Title<br />

1<br />

0<br />

Linear (experimental)<br />

-11.6<br />

2.1 2.6<br />

ln(x)<br />

y = 9.3158x - 22.502<br />

R 2 = 0.9736<br />

Figure A2.2: linearised probability plot IPC data <strong>and</strong> two-parameter Weibull model<br />

The value <strong>of</strong> the shape parameter η is the slope <strong>of</strong> the linearised curve, thus η is<br />

9.32. Scale parameter σ is determined from equation (A2.6) <strong>and</strong> (A2.10).<br />

B = −η<br />

lnσ<br />

(A2.10)<br />

The value <strong>of</strong> σ is thus 11.2.<br />

When the three-parameter Weibull model is used, one extra parameter is to be<br />

obtained: the location parameter µ. In the x-axis <strong>of</strong> the linearised probability curve<br />

ln(x-µ) is represented. The y-axis represents G(x), <strong>with</strong> G(x) being defined by


equation (A2.9) as for the two-parameter Weibull model. When the threeparameter<br />

Weibull model is used, initially a value <strong>of</strong> µ is chosen. The values <strong>of</strong> η,<br />

σ <strong>and</strong> the regression coefficient are determined as if a two-parameter Weibull<br />

model is used. The value <strong>of</strong> µ is then varied. For each value <strong>of</strong> µ, the value <strong>of</strong> η, σ<br />

<strong>and</strong> the regression coefficient are obtained. The “best fit” is obtained when a value<br />

<strong>of</strong> µ is found, leading to a maximum value <strong>of</strong> the regression coefficient. For this<br />

value <strong>of</strong> µ, the value <strong>of</strong> η <strong>and</strong> σ are obtained.<br />

Figure A2.3 shows the influence <strong>of</strong> µ on the three-parameter Weibull model<br />

linearisation.<br />

G(x)<br />

2<br />

1<br />

0<br />

-2<br />

-3<br />

-4<br />

-5<br />

µ = 0<br />

R 2 = 0.9736<br />

µ = 1<br />

R 2 = 0.9746<br />

1.4 1.6 1.8 2 2.2 2.4 2.6<br />

-1<br />

ln(x-µ)<br />

µ = 2<br />

R 2 = 0.9752<br />

Figure A2.3: linearised probability plot IPC data <strong>and</strong> three-parameter Weibull model<br />

When the value <strong>of</strong> µ is 2.75, the maximum regression coefficient is found, which<br />

is 0.9757. The value <strong>of</strong> σ is then 8.43<strong>and</strong> η is 6.76.<br />

A2.4 The Largest extreme value model (LEV)<br />

The x-axis <strong>of</strong> the linearised LEV model represents x. The y-axis represents G(x),<br />

<strong>with</strong> G(x) being defined by equation (A2.11):<br />

⎛ ⎛ 1 ⎞⎞<br />

G ( x)<br />

= ln⎜<br />

⎟<br />

⎜<br />

ln ⎜<br />

⎟<br />

⎟<br />

(A2.11)<br />

⎝ ⎝ F(<br />

x)<br />

⎠⎠<br />

Figure A2.4 shows the linearised probability plot, according to the LEV model for<br />

the tested IPC specimens.


G(x)<br />

3<br />

2<br />

1<br />

0<br />

-1<br />

-2<br />

-3<br />

-4<br />

6 8 10 12 14<br />

y = -0.8651x + 8.6569<br />

R 2 = 0.9471<br />

x<br />

experimental<br />

Linear (experimental)<br />

Figure A2.4: linearised probability plot IPC data <strong>and</strong> Largest extreme value method (LEV)<br />

The value <strong>of</strong> the model parameters are found from equation (A2.6) <strong>and</strong>:<br />

1<br />

A = −<br />

(A2.12)<br />

η<br />

µ<br />

B =<br />

(A2.13)<br />

η<br />

µ is 10.0 <strong>and</strong> η is 1.156.


Appendix 3<br />

A3.1 Introduction<br />

Hysteresis<br />

It has been mentioned in Chapter 4 that the shape <strong>of</strong> the hysteresis loop contains<br />

information about the ongoing meso-mechanics. It has been also mentioned that<br />

stress-strain loop widening under repeated cycling is an indication towards partial<br />

matrix-fibre slip while loop tightening gives an indication that full matrix-fibre<br />

slip occurs during unloading <strong>and</strong> reloading. The theoretical verification <strong>of</strong> this<br />

statement is enclosed in this appendix.<br />

A3.2 Theoretical determination <strong>of</strong> hysteresis<br />

Figure A1.3 shows an unloading-reloading stress-strain cycle. In this figure<br />

∆εc hysteresis is the hysteresis strain at mean cycle stress. The goal <strong>of</strong> this paragraph is<br />

to obtain a formulation <strong>of</strong> ∆εc hysteresis <strong>and</strong> to link this formulation <strong>with</strong> the ongoing<br />

meso-mechanics in the IPC composite during cyclic loading.<br />

σ<br />

σc max<br />

σc mean<br />

σc min<br />

εc min<br />

∆εc,r mean<br />

reloading<br />

εc,r mean<br />

∆εc hysteresis<br />

εc,u mean<br />

∆εc,u mean<br />

unloading<br />

εc max<br />

Figure A3.1: definition <strong>of</strong> stresses, strains <strong>and</strong> hysteresis in an unloading-reloading loop<br />

ε


A3.2.1 determination <strong>of</strong> ∆εc hysteresis : partial matrix-fibre slip<br />

A composite specimen is unloaded from a maximum stress σc max to σc min , <strong>with</strong><br />

∆σc = σc max - σc min . It is assumed that only partial matrix-fibre slip occurs during<br />

the whole process <strong>of</strong> unloading <strong>and</strong> reloading. The hysteresis strain ∆εc hysteresis will<br />

be derived at a stress level: σc mean = σc max - ∆σc/2 = σc min + ∆σc/2. Figure A3.2<br />

illustrates the stresses in the matrix at maximum cycle stress σc max (1), at mean<br />

cycle stress during unloading σc mean (2), at minimum cycle stress σc min (3), (4), at<br />

mean cycle stress during reloading σc mean (5) <strong>and</strong> at maximum cycle stress after<br />

reloading σc max (6).<br />

crack<br />

crack<br />

/2<br />

/2<br />

δN<br />

composite<br />

δN<br />

(4)<br />

(3)<br />

x<br />

(5) (6)<br />

(2) (1)<br />

(1) maximum cycle stress, σc max<br />

(2) mean cycle stress, σc mean , during unloading<br />

(3), (4) minimum cycle stress, σc min<br />

(5) mean cycle stress, σc mean , during reloading<br />

(6) maximum cycle stress, σc max , after reloading<br />

reloading<br />

unloading<br />

(su(∆σc))<br />

(su(∆σc))<br />

σm,N max σ<br />

Figure A3.2: stresses in matrix during unloading <strong>and</strong> reloading.<br />

The definitions <strong>of</strong> the composite strains used in the formulation <strong>of</strong> the hysteresis<br />

are illustrated in figure A3.1. The formulation <strong>of</strong> ∆εc hysteresis is found by expressing<br />

∆εc,u mean , ∆εc,r mean <strong>and</strong> ∆εc.<br />

Unloading from the maximum cycle stress σc max to the minimum cycle stress σc min<br />

occurs <strong>with</strong> a strain variation ∆εc, which is formulated according to equation (3.32)<br />

<strong>and</strong> is thus:


∆σ<br />

⎛ ⎞<br />

c ⎜<br />

αδ N ∆σ<br />

c<br />

ε = 1 + ⎟<br />

c ⎜<br />

⎟<br />

(A3.1)<br />

Ec1 ⎝ 2 cs σ c ⎠<br />

∆ max<br />

Unloading from the maximum cycle stress σc max to the mean cycle stress σc mean<br />

occurs <strong>with</strong> a strain variation ∆εc,u mean , according to equation (3.32), but ∆σc should<br />

be replaced by ∆σc/2. Thus:<br />

⎛ ⎞<br />

mean ∆σ<br />

c ⎜<br />

αδ N ∆σ<br />

c<br />

∆ε = + ⎟<br />

c,<br />

u 1<br />

⎜<br />

max<br />

2 ⎟<br />

(A3.2)<br />

Ec1 ⎝ 4 cs σ c ⎠<br />

Unloading from σc max to σc mean occurs <strong>with</strong> combined linear elastic unloading <strong>and</strong><br />

matrix-fibre slip. Reloading from σc min to σc mean also occurs <strong>with</strong> combined linear<br />

elastic reloading <strong>and</strong> matrix-fibre slip. Since in both cases the composite stress<br />

variation is ∆σc/2 <strong>and</strong> the same meso-mechanical phenomena occur during<br />

unloading <strong>and</strong> reloading, ∆εc,r mean equals ∆εc,u mean . Thus:<br />

⎛ ⎞<br />

mean ∆σ<br />

c ⎜<br />

αδ N ∆σ<br />

c<br />

ε = + ⎟<br />

c,<br />

r 1<br />

2 ⎜<br />

⎟<br />

(A3.3)<br />

Ec1 ⎝ 4 cs σ c ⎠<br />

∆ max<br />

From figure A3.1 it can be noticed that:<br />

hysteresis<br />

∆ ε<br />

mean mean<br />

= ∆ε<br />

− ∆ε<br />

+ ∆ε<br />

(A3.4)<br />

c c c,<br />

u c,<br />

r<br />

Combination <strong>of</strong> equations (A3.1), (A3.2), (A3.3) <strong>and</strong> (A3.4) gives:<br />

2<br />

hysteresis δ Nα<br />

( ∆σ<br />

c )<br />

∆ εc<br />

=<br />

max<br />

(A3.5)<br />

4Ec cs σ c<br />

1<br />

The debonding length is the only variable in equation (A3.5), which changes <strong>with</strong><br />

increasing number <strong>of</strong> elapsed cycles. The width <strong>of</strong> the hysteresis strain ∆εc hysteresis<br />

at the mean cycle stress increases during repeated cycling, since the debonding<br />

length δN in equation (A3.5) increases.<br />

A3.2.2 determination <strong>of</strong> ∆εc hysteresis : total matrix-fibre slip<br />

∆εc hysteresis , is found by expressing ∆εc,u mean , ∆εc,r mean , <strong>and</strong> ∆εc:<br />

hysteresis<br />

mean mean<br />

∆ ε = ∆ε<br />

− ∆ε<br />

+ ∆ε<br />

(A3.6)<br />

c c c,<br />

u c,<br />

r<br />

The formulation <strong>of</strong> ∆εc is expressed by equation (3.55) <strong>and</strong> is:<br />

cs Em<br />

max<br />

∆σ<br />

c − Vm<br />

σ c<br />

2δ<br />

N Ec1<br />

∆ε<br />

c =<br />

∗<br />

E V<br />

f<br />

f<br />

(A3.7)<br />

If total matrix-fibre slip already occurred at the intermediate stress ∆σc mean in the<br />

unloading step, it will also occur at the reloading step, since the composite slip


strain terms are equal, but <strong>of</strong> opposite sign. When unloading occurs, equation<br />

(A3.7) can thus be rewritten, <strong>with</strong> replacement <strong>of</strong> ∆σc by ∆σc/2:<br />

∆σ<br />

cs c Em<br />

max<br />

− Vm<br />

σ c<br />

mean 2 2δ<br />

N Ec1<br />

∆ε<br />

c,<br />

u =<br />

(A3.8)<br />

∗<br />

E V<br />

And for reloading:<br />

∆ε<br />

mean<br />

c,<br />

r<br />

The hysteresis strain is thus:<br />

∆σ<br />

c E<br />

− Vm<br />

2 2δ<br />

N E<br />

=<br />

E V<br />

f<br />

f<br />

cs m max<br />

σ c<br />

c1<br />

∗<br />

f f<br />

cs E<br />

cs ασ<br />

(A3.9)<br />

hysteresis<br />

∆ε c<br />

mean mean<br />

m max<br />

= ∆εc<br />

− ∆εc,<br />

u − ∆εc,<br />

r = Vm<br />

σ ∗ c<br />

2δ N E fV<br />

f Ec1<br />

=<br />

2δ<br />

N<br />

max<br />

c<br />

Ec1<br />

(A3.10)<br />

The only parameter, which changes <strong>with</strong> the number <strong>of</strong> cycles in equation<br />

(A3.10), is δN. If the number <strong>of</strong> elapsed load cycles increases, the debonding<br />

length enlarges <strong>and</strong> the hysteresis strain decreases.<br />

A3.3 Conclusion<br />

If only partial matrix-fibre slip occurs during the whole unloading-reloading loop,<br />

the width <strong>of</strong> the hysteresis loop increases linearly <strong>with</strong> δN. Therefore the width <strong>of</strong><br />

the hysteresis loop increases <strong>with</strong> increasing number <strong>of</strong> applied load cycles.<br />

If total matrix-fibre slip already occurred when the mean cycle stress ∆σc mean is<br />

reached, the width <strong>of</strong> the hysteresis loop increases linearly <strong>with</strong> increasing 1/δN.<br />

This means that the width <strong>of</strong> the hysteresis loop decreases <strong>with</strong> increasing number<br />

<strong>of</strong> applied load cycles.


Appendix 4<br />

Macro for implementation <strong>of</strong> unloading<br />

A4.1 Introduction<br />

The stress-strain behaviour <strong>of</strong> IPC composites under monotonic loading can be<br />

formulated <strong>with</strong> rather high accuracy <strong>with</strong> st<strong>and</strong>ard material laws in finite element<br />

programs. Unfortunately no st<strong>and</strong>ard finite element code approximates the<br />

unloading behaviour <strong>of</strong> IPC composite specimens. A macro has therefore been<br />

written in ANSYS for this thesis: unload.mac. The outline <strong>of</strong> this macro is<br />

presented in this appendix.<br />

A4.2 Summary <strong>of</strong> numerical implementation<br />

Figure A4.1 shows a force-displacement loading <strong>and</strong> unloading step on a s<strong>and</strong>wich<br />

panel <strong>with</strong> IPC composite faces. The loading step can usually be calculated <strong>with</strong> a<br />

st<strong>and</strong>ard material law in the finite element program. The unloading step needs<br />

some modifications. Table A4.1 illustrates step by step how loading <strong>and</strong> unloading<br />

should be performed when the macro unload.mac is used in finite element<br />

calculations to predict the unloading behaviour <strong>of</strong> a s<strong>and</strong>wich panel <strong>with</strong> IPC<br />

composite faces.<br />

load<br />

p<br />

step 1: loading (p is applied as positive<br />

load in finite element calculation)<br />

step 2: loading (p is applied as negative<br />

load in finite element calculation)<br />

displacement<br />

Figure A4.1: force-displacement loading <strong>and</strong> unloading step <strong>of</strong> s<strong>and</strong>wich panel <strong>with</strong> IPC<br />

composite faces


Table A4.1: step by step implementation <strong>of</strong> loading <strong>and</strong> unloading <strong>of</strong> a s<strong>and</strong>wich panel <strong>with</strong> IPC<br />

composite faces, calculation <strong>with</strong> finite element program: macro unload.mac in ANSYS<br />

1. Geometry <strong>and</strong> material data tables are inserted in finite element program,<br />

meshing is performed<br />

- The material behaviour ‘aniso’ is used for IPC composite faces to allow<br />

different behaviour in tension <strong>and</strong> compression <strong>and</strong> in three directions. (see figure<br />

A4.2).<br />

- The most important material parameters in the data table <strong>of</strong> IPC composite<br />

are (see figure A4.2):<br />

*Ex+ el =Ex- el = Ex el = E-modulus prior to yielding<br />

*Ex+ T = E-modulus in tension after yielding<br />

*σ y x+ = yield stress in tension<br />

*σ y x- = yield stress in compression<br />

2. Load step 1 is applied on the panel<br />

3. Loading <strong>of</strong> the s<strong>and</strong>wich panel is solved in finite element program<br />

4. The element displacements, stresses <strong>and</strong> strains are retrieved from the results<br />

file <strong>of</strong> the finite element program <strong>and</strong> saved to data file step1.dat<br />

5. Run macro unload.mac for recalculation <strong>of</strong> element stiffness<br />

unload.mac<br />

retrieve all elements attributed to IPC material definition<br />

loop over retrieved elements<br />

retrieve element stress (σelement)<br />

if element loaded in tension then<br />

if element stress ≥ yield stress (see figure A4.3)<br />

calculate crack spacing <strong>of</strong> element<br />

calculate debonding length δ0 <strong>of</strong> element<br />

Ex+ el = Ecycle (see Chapter 3 <strong>and</strong> table A4.2)<br />

σ y x+ ≅ 0<br />

else (see figure A4.4)<br />

recalculate σ y x+: σ y x+ = σ y x+ - σelement<br />

end if<br />

else (see figure A4.4)<br />

recalculate σ y x+: σ y x+ = σ y x+ - σelement<br />

end if<br />

close retrieved elements loop<br />

6. The finite element results file is cleared<br />

7. All loads are removed from the s<strong>and</strong>wich panel<br />

8. Load step 2 is applied in finite element program (unloading load)<br />

9. Load step 2 on s<strong>and</strong>wich panel is solved in finite element program<br />

10. The element displacements, stresses <strong>and</strong> strains are retrieved from the results<br />

file <strong>of</strong> the finite element program <strong>and</strong> saved to data file step2.dat<br />

11. Total displacements = displacements from step1.dat + displacements from<br />

step2.dat


σ y x,+<br />

compression yield stress<br />

∨<br />

∨<br />

tensile yield stress<br />

Ex,- pl<br />

σ<br />

Ex el<br />

Ex,+ pl<br />

σ y x,-<br />

σ y x,+= tensile yield stress<br />

σ y x,- = compressive yield stress.<br />

Ex el = initial value <strong>of</strong> the E-modulus used for compression <strong>and</strong> tension.<br />

Ex,+ pl = E-modulus in tension, once the yield stress is reached<br />

Ex,- pl = E-modulus in compression, once the yield stress is reached<br />

Figure A4.2: stress-strain behaviour ‘aniso’ for loading <strong>of</strong> IPC composite faces,<br />

length axis (x-axis)<br />

σx,+ y<br />

σ<br />

E el<br />

σ<br />

σx,+ y = 0<br />

Ex el = Ecycle<br />

Ex,+ pl = Ex,- pl<br />

compression yield stress<br />

∨<br />

∨<br />

tensile yield stress<br />

ε<br />

ε<br />

ε<br />

loading (step 1)<br />

unloading (step 2)<br />

Figure A4.3: IPC composite element material properties (yield stress <strong>and</strong> stiffness) for unloading;<br />

step 1 in grey (loading) <strong>with</strong> element stress ≥ yield stress, step 2 in black (unloading)


σx,+ Y<br />

σ<br />

σ<br />

Ex,+ pl = Ex,+ pl<br />

σx,+ y<br />

Ex el = Ex el<br />

ε<br />

compression yield stress<br />

∨<br />

∨<br />

tensile yield stress<br />

ε<br />

loading (step 1)<br />

unloading (step 2)<br />

Figure A4.4: IPC composite element material properties (yield stress <strong>and</strong> stiffness) for unloading;<br />

step 1 in grey (loading) <strong>with</strong> element stress < yield stress, step 2 in black (unloading)<br />

Table A4.1: summary <strong>of</strong> formulations for Ecycle<br />

condition Ecycle formulation<br />

> 2δ0 <strong>and</strong> partial matrix-fibre unloading slip<br />

Ec1<br />

Ecycle<br />

=<br />

αδ 0∆σ<br />

c 1+<br />

max<br />

2 cs σ<br />

> 2δ0 <strong>and</strong> total matrix-fibre unloading slip condition not possible<br />

≤ 2δ0 <strong>and</strong> partial matrix-fibre unloading slip<br />

≤ 2δ0 <strong>and</strong> total matrix-fibre unloading slip<br />

where: Ec1 = Ex el<br />

σc max = σelement<br />

E<br />

E<br />

cycle<br />

cycle<br />

Ec1<br />

=<br />

αδ 0∆σ<br />

1+<br />

2 cs σ<br />

c<br />

c<br />

max<br />

c<br />

∗<br />

E fV<br />

f<br />

=<br />

cs Em<br />

σ<br />

1−<br />

Vm<br />

2δ<br />

E ∆σ<br />

∆σc = σelement (choice for implementation <strong>of</strong> behaviour according to Chapter 3)<br />

= average crack spacing in the element <strong>and</strong> is function <strong>of</strong> σelement<br />

δ0 = debonding length in the element <strong>and</strong> is function <strong>of</strong> σelement<br />

α = ratio, function <strong>of</strong> fibre & matrix volume fraction <strong>and</strong> Young’s moduli<br />

0<br />

c1<br />

max<br />

c<br />

c


Appendix 5<br />

A5.1 Introduction<br />

Macro for implementation <strong>of</strong><br />

repeated loading<br />

The stress-strain behaviour <strong>of</strong> IPC composites under monotonic loading can be<br />

formulated <strong>with</strong> rather high accuracy by st<strong>and</strong>ard material laws in finite element<br />

programs. Unfortunately no st<strong>and</strong>ard finite element code approximates the<br />

repeated loading behaviour <strong>of</strong> IPC composite specimens. A macro has therefore<br />

been written in ANSYS for this thesis: repeat.mac. The outline <strong>of</strong> this macro is<br />

presented in this appendix.<br />

A5.2 Summary <strong>of</strong> numerical implementation<br />

Figure A5.1 shows a loading step <strong>and</strong> accumulation <strong>of</strong> deflections due to repeated<br />

loading. Table A5.1 illustrates how repeated loading can be introduced, when<br />

macro repeat.mac is used in finite element calculations. repeat.mac provides an<br />

estimation <strong>of</strong> the extra deformations due to repeated loading, but should not be<br />

considered highly accurate.<br />

load<br />

p<br />

step 1: loading<br />

N loading cycles<br />

step 2: extra deformations due to repeated<br />

loading (N cycles)<br />

displacement<br />

Figure A5.1: force-displacement loading <strong>and</strong> repeated loading step <strong>of</strong> s<strong>and</strong>wich panel <strong>with</strong> IPC<br />

composite faces


Table A5.1: step by step implementation <strong>of</strong> loading <strong>and</strong> repeated loading <strong>of</strong> a s<strong>and</strong>wich panel<br />

<strong>with</strong> IPC composite faces, calculation <strong>with</strong> finite element program: macro repeat.mac in ANSYS<br />

1. Geometry <strong>and</strong> material data tables are inserted in finite element program,<br />

meshing is performed<br />

2. The load is applied on the panel<br />

3. Loading <strong>of</strong> the s<strong>and</strong>wich panel is solved in finite element program<br />

4. The element displacements, stresses <strong>and</strong> strains are retrieved from the results<br />

file <strong>of</strong> the finite element program <strong>and</strong> saved to data file step1.dat<br />

5. Run macro repeat.mac for calculation <strong>of</strong> extra element strains<br />

repeat.mac<br />

insert number <strong>of</strong> applied loading cycles N (repeated loading)<br />

insert interface degradation constants C1 & C2 (Chapter 4)<br />

retrieve all elements attributed to IPC material definition<br />

loop over retrieved elements<br />

retrieve element stress (σelement)<br />

if element loaded in tension then<br />

calculate crack spacing <strong>of</strong> element<br />

calculate debonding length δN <strong>of</strong> element<br />

calculate extra element strain term ∆εc,N repeat due to<br />

repeated cycling (see paragraph A5.3)<br />

save extra element strain <strong>and</strong> element number in<br />

data file extrastrain.dat<br />

close retrieved elements loop<br />

6. The finite element results file is cleared<br />

7. All loads are removed from the s<strong>and</strong>wich panel<br />

8. Loop over all elements<br />

applied element strain = element strain from data file extrastrain.dat<br />

Close elements loop<br />

9. Calculate deformation s<strong>and</strong>wich panel due to applied element strains <strong>with</strong><br />

finite element program<br />

10. The displacements are retrieved from the results file <strong>of</strong> the finite element<br />

program <strong>and</strong> saved to data file step2.dat<br />

11. Total displacements = displacements from step1.dat + displacements from<br />

step2.dat<br />

It should be mentioned that the methodology presented in table A5.1 gives an<br />

estimation <strong>of</strong> extra deflections <strong>of</strong> a s<strong>and</strong>wich panel <strong>with</strong> IPC composite faces<br />

under repeated loading. No redistribution <strong>of</strong> stresses across the thickness <strong>of</strong> the<br />

s<strong>and</strong>wich composite is calculated. If more accurate results are to be found, the<br />

methodology presented in table A5.1 can be used in an iterative calculation. Small<br />

steps <strong>of</strong> elapsed loading cycles can be applied, each time <strong>with</strong> recalculation <strong>of</strong><br />

redistribution <strong>of</strong> stresses. However, more information is then needed on the<br />

influence <strong>of</strong> the actual stress on the accumulation <strong>of</strong> damage in IPC composite


specimens than provided in Chapter 4. The extra strain term in Chapter 4 is<br />

formulated as a function <strong>of</strong> the maximum applied composite cycle stress, but not<br />

<strong>of</strong> the actual composite cycle stress. The information on the influence <strong>of</strong> the actual<br />

composite cycle stress is not determined quantitatively yet.<br />

A5.3 Determination <strong>of</strong> the extra strain term due to repeated<br />

loading, ∆εc,N repeat<br />

The extra strain term, due to repeated loading is:<br />

repeat<br />

∆ ε<br />

max max<br />

= ε − ε<br />

(A5.1)<br />

c,<br />

N c,<br />

N c,<br />

0<br />

where: εc,0 max = composite strain under monotonic loading up to σc max<br />

σc max = maximum cycle stress<br />

εc,N max = the composite strain at σc max after N cycles are applied<br />

The extra strain term, due to repeated loading should be formulated for two cases.<br />

The first case assumes the average crack spacing exceeds twice the debonding<br />

length ( > 2δN). The second case assumes the opposite.<br />

If > 2δN, the composite strain at maximum stress is formulated by equation<br />

(4.34), as a function <strong>of</strong> the number <strong>of</strong> applied cycles. When equations (4.34) <strong>and</strong><br />

(4.33) are inserted in equation (A5.1), the extra strain term can be written:<br />

max ( σ )<br />

2<br />

( ) ⎟ repeat<br />

∆ε c,<br />

N =<br />

⎛<br />

⎞<br />

c rαEmVm<br />

1<br />

⎜<br />

−1<br />

2 ∗<br />

E ⎝ C1<br />

− C2<br />

ln N<br />

c1<br />

cs V f<br />

⎠<br />

(A5.2)<br />

If < 2δN, equations (4.33) <strong>and</strong> (4.36) are to be combined <strong>with</strong> equation<br />

(A5.1) to obtain equation (A5.3):<br />

∗<br />

V fτ<br />

cs<br />

repeat 0α<br />

∆ε<br />

c,<br />

N = ( 1−<br />

C1<br />

+ C2<br />

ln N )<br />

2rEmVm<br />

<strong>with</strong>: Vf<br />

(A5.3)<br />

* = equivalent fibre volume fraction<br />

Em = Young’s modulus <strong>of</strong> the matrix<br />

Ef = Young’s modulus <strong>of</strong> the fibres<br />

Ec1 = Ex el<br />

σc max = σelement<br />

= average crack spacing in the element <strong>and</strong> is function <strong>of</strong> σelement<br />

δN = debonding length after N cycles in the element, function <strong>of</strong> σelement<br />

α = ratio, function <strong>of</strong> fibre & matrix volume fraction <strong>and</strong> Young’s moduli<br />

r = fibre radius<br />

C1 & C2 = interface degradation parameter constants

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!