• Aucun résultat trouvé

Multi-scale Modelling of the Effect a Viscoelastic Matrix has on the Strength of a Carbon Fibre Composite

N/A
N/A
Protected

Academic year: 2021

Partager "Multi-scale Modelling of the Effect a Viscoelastic Matrix has on the Strength of a Carbon Fibre Composite"

Copied!
29
0
0

Texte intégral

(1)

HAL Id: hal-00605958

https://hal.archives-ouvertes.fr/hal-00605958

Submitted on 5 Jul 2011

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Frank Ralph Jones, Joel P Foreman, David Porter, Shabnam Behzadi

To cite this version:

Frank Ralph Jones, Joel P Foreman, David Porter, Shabnam Behzadi. Multi-scale Modelling of the Effect a Viscoelastic Matrix has on the Strength of a Carbon Fibre Composite. Philosophical Magazine, Taylor & Francis, 2010, pp.1. �10.1080/14786435.2010.495361�. �hal-00605958�

(2)

For Peer Review Only

Multi-scale Modelling of the Effect a Viscoelastic Matrix has on the Strength of a Carbon Fibre Composite

Journal: Philosophical Magazine & Philosophical Magazine Letters Manuscript ID: TPHM-09-Oct-0451.R1

Journal Selection: Philosophical Magazine Date Submitted by the

Author: 07-Apr-2010

Complete List of Authors: Jones, Frank; The University of Sheffield, Engineering Materials Foreman, Joel; University of Sheffield, Engineering Materials Porter, David; University of Oxford, Zoology; University of Sheffield, Engineering Materials

Behzadi, Shabnam; University of Sheffield, Engineering Materials Keywords: composite materials, modelling, strength

Keywords (user supplied): Epoxy Resins

(3)

For Peer Review Only

Multi-scale Modelling of the Effect a Viscoelastic Matrix has on the Strength of a Carbon Fibre Composite

Joel P. Foreman1, Shabnam Behzadi1, David Porter2, Frank R. Jones1

1 Department of Engineering Materials, University of Sheffield, Mappin Street, Sheffield, S1 3JD. UK

2 Department of Zoology, University of Oxford, South Parks Road, Oxford, OX1 3PS, UK

Abstract

A recently developed multi-scale model has proven successful in predicting the tensile strength of a unidirectional fibre composite from fundamental molecular inputs. This technique is now extended to subtle changes in the properties of the matrix. The chemistry of the resin matrix is varied on the functional group level resulting in a series of stress-strain profiles predicted using Group Interaction Modelling. The transfer of strain as the result of a fibre break in the composite is modelled using finite element methods. The resulting characteristic ineffective lengths and strain concentration factors are incorporated into a statistical simulation of the propagation of fibre failure events in a typical composite. This allows the prediction of ultimate tensile strength for a composite containing a yielding and non-yielding matrix phase.

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(4)

For Peer Review Only

1. Introduction

The use of thermosetting epoxy resins as the matrix phase in high performance composite systems is commonplace. Such resins offer a number of benefits over alternative systems, such as increased strength, toughness, environmental resistance, good electrical and thermal properties and minimal shrinkage [1]. Commercially available resin systems used by the transport industry for mission critical applications often involve blending two or more epoxy resins together with a multifunctional amine curing agent. The composite part is then created by impregnating an array of carbon fibres with the blended resin and curing to a particular schedule. In the past the effect the resin matrix has on the composite properties has been relatively poorly understood. It is only recently that the importance of the viscoelastic nature of the matrix has been shown to directly influence the strength of composite systems [2].

The complex polymeric structure of the resin allows energy to be dissipated as heat via strain rate dependent molecular relaxation processes. This ductile behaviour can be tailored to suit the particular application and this paper shows the results of an attempt to predict this effect using a simple modelling approach.

A number of multifunctional epoxy resins are used in commercial formulations, including those based on glycidyl ethers and amines that typically have functionalities varying between 2 and 4. In particular, diglycidyl ether of bisphenol-A (DGEBA) has two epoxy groups attached to a central phenyl-CMe2-phenyl segment via two ether linkages. Tetraglycidyl 4,4’ diaminodiphenylmethane (TGDDM) has four epoxy groups attached via amine linkages and triglycidyl p-aminophenol (TGAP) has three epoxy groups attached via one ether and one amine. A typical curing agent for these epoxies is the multifunctional amine diaminodiphenylsulphone (DDS). The chemical structures of the DGEBA, TGDDM, TGAP and DDS monomers are shown in Figure 1.

The mechanism for curing epoxy resins with amines is less than straightforward [3]

and is shown in Figure 2. All three reactions involve opening the relatively unstable oxirane ring in the epoxy group to form a CH(OH) group. The initial step in reaction 1 involves a primary amine reacting to form a secondary amine which can then react further to form a tertiary amine in reaction 2. The CH(OH) product of reactions 1 and 2, can also react to form an ether bridge as shown in reaction 3. The likelihood of each reaction occurring is 1>2>3 in the ratio 100:40:10 [3]. The number of epoxy groups in the monomer (the functionality) determines how crosslinked the cured resin will be. Hence, the number of crosslinks between polymer chains increases with increasing functionality (DGEBA<TGAP<TGDDM). The amine curing agent DDS also has a functionality of 4 and hence contributes to the degree of crosslinking in the system.

Fully cured, a typical epoxy resin system is a complex 3-D network which is amorphous and isotropic in nature. From a traditional modelling perspective such as molecular modelling this is quite a disadvantage as such techniques rely on knowing the Cartesian coordinates of the all atoms in the molecule. In a typical epoxy resin system, the coordinates of many thousands of atoms would be required to create a representative sample for modelling. This in turn increases the computing power necessary and hence the cost. Where a representative morphology exists, a repeat unit

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(5)

For Peer Review Only

can reduce the computation required significantly but this is generally not the case in crosslinked polymers. To solve this problem modelling techniques have been developed which use a mean field approach to predicting the properties of polymers, such as Group Interaction Modelling (GIM) [4] and similar techniques [5]. In GIM, the need for exact atomic coordinates is removed by defining the equations of state and hence predictive constitutive equations in terms of the polymers’ representative mer unit(s). In this way, the chemical structure of the polymer is defined by the constituent functional groups of the mer unit (e.g. CH2, Phenyl, Epoxy), the resin component ratio (e.g. epoxy to amine ratio) and the degree of cure. This relatively simple information allows the computational time to be drastically reduced while retaining the sound physical basis for accurate predictions [6,7,8].

One of the strengths of GIM is the relatively simple way in which the chemistry is defined (i.e. number and type of functional groups). The work presented here will show how the GIM predicted properties of cured epoxy resins change when the chemistry is altered in a systematic way. Three aspects of epoxy resin chemistry are investigated, the epoxy-amine ratio, the isomer of DDS and specific changes in the chemistry of DGEBA. While it is unlikely that a technique like GIM could ever replace experiments, it is worth noting that the calculations presented here were performed in a tiny fraction of the time (and cost) required to measure these quantities experimentally. It is envisaged that a technique such as GIM could be used in the future as a complementary technique, enabling the resin formulator to screen possible candidates for suitability at a fraction of the current cost.

As part of a larger multi-scale modelling effort, the GIM predicted properties such as modulus and stress-strain curves are used in a finite element model that determines the effect a fibre failure has on neighbouring fibres. A fibre failure event triggers an increase in strain concentration in surrounding fibres via the viscoelastic matrix phase, the properties of which are predicted by GIM. This information is then used in a statistical model to predict ultimate composite strength. A Monte Carlo type simulation assesses the propagation of fibre failure events through the composite as a function of the GIM and FEA predicted properties of the matrix and fibres [9,10].

2. Methods

2.1 Group Interaction Modelling

Group Interaction Modelling uses a mean field potential function approach that predicts the bulk properties of polymers from molecular input. The technique uses a modified Lennard Jones potential to define the total interaction energy between neighbouring polymer chains, Etotal. The potential is defined in terms of thermodynamic energy terms to describe the attractive and repulsive contributions to the total energy function. The equation of state for the system linking energy and volume, V, is the thermodynamic potential function shown in equation [1].

M C T coh coh

total E H H H

V V V

E V

E =− + + +





 

 

− 



 

= 

3 0 6

0 2 [1]

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(6)

For Peer Review Only

The minimum of the potential energy well, Ecoh, represents the cohesive energy of the system which is acted against by the temperature dependent volumetric term. The polymer chains are arranged in a hexagonal array with an interchain separation distance, r. The polymer chains are assumed to be many orders of magnitude longer than they are wide and therefore, for the purposes of the model, invariant in length.

This means that V ∝ r2 and hence the unusual powers in the Lennard Jones potential in equation [1]. The system functions as a strong 1-D oscillator in a weak 3-D field.

The other thermodynamic terms in equation [1] are all positive, bringing the system out of the potential well minimum. The thermal energy, HT, is defined using a simplification of the 1-D Debye function for skeletal mode vibrations. The cohesive and thermal energy terms are the two main contributions to the total interaction energy, with subtle modifications from the following terms. HC is a configurational energy term which allows for the presence of morphology such that an amorphous polymer has a higher HC than a semi-crystalline polymer. For amorphous systems, HC

takes a value of 0.11Ecoh. HM is a mechanical energy term which is zero when the system is in equilibrium and positive when the system is perturbed.

A series of linked constitutive equations for the prediction of properties is then derived from the fundamental equation of state given above. Each of the equations defines a particular polymer property in terms of a series of fundamental parameters and previously derived equations. The parameters required are defined on a functional group or mer unit basis and generally obtained using group contributions tables [4,11], but other techniques such as molecular modelling are useful too. The parameters are the degrees of freedom, N, the cohesive energy, Ecoh, the van der Waal’s volume, Vw, the molecular weight, M, the length, L, and the 1-D Debye temperature, θ1. Below the glass transition temperature, N is given a value of 2 for a typical small functional group (such as –O- or –CH2-) to account for the degrees of freedom normal to the polymer chain. Correspondingly, above the glass transition temperature N is 3 to account for the extra degree of freedom in the direction of the polymer chain that gets activated. The value of Ecoh is effectively determined using group contribution tables and represents the minimum energy of the system against which all the other thermodynamic terms act. In practice, the GIM equations balance the attractive Ecoh with the repulsive thermal contributions from N. Crosslinking is incorporated into the model by decreasing the value of N by 3 for each crosslink site on the mer unit. Similarly, hydrogen bonding is accounted for by increasing the value of Ecoh by 10,000 J/mol for each hydrogen bond site. The value of θ1 is obtained from the datasets of Wunderlich [12] and takes a value of 550 K for polymers containing phenyl rings in the backbone.

The epoxy resin systems investigated in this work are all viscoelastic in nature, and this is incorporated into GIM by treating the elastic and viscous contributions separately using a simple spring-dashpot type model. The viscous contribution is assumed to be due to the two significant transition events which occur in the dynamic mechanical profile. These events can be associated with strain rate dependent molecular relaxation processes which are activated at specific temperatures where energy is dissipated as heat. The glass transition is associated with a large scale increase in the intermolecular motion between polymer chains as the system moves from the glassy to rubbery phase. In this class of epoxy resins, the beta transition is associated with an intramolecular cooperative rotation of the backbone phenyl rings

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(7)

For Peer Review Only

[13]. It is likened to a ‘crankshaft’ style motion of two semi-adjacent rings where the handles of the crankshaft are the C-C bonds in the rings parallel to the backbone.

In GIM, these loss events are modelled using a Gaussian distribution function to simulate the temperature distribution of the event. Also, strain rate dependence in GIM is incorporated through the two loss events. The equations control how the speed of a perturbation event affects the temperature at which the event occurs. On a time- temperature equivalent scale, the faster the perturbation occurs, the less chance the relaxation event has to happen. This manifests itself as a relationship between strain rate and transition temperature which is modelled in GIM.

A simple molecular modelling routine is used to estimate the activation energy for each peak then an Arrhenius equation is used to determine the temperature of each transition. The loss associated with both transitions is determined using the loss tangent relationship between lost and stored energy. The cumulative (tan∆) and local (tanδ) loss tangents are determined and expressed as a function of temperature. A distribution term (i.e. the shape of the peak) is determined and the variables are applied to a Gaussian distribution function for each loss tangent profile.

At the heart of GIM is a series of linked constitutive equations designed to predict the terms in equation [1]. The heat capacity at constant pressure, Cp, is expressed in terms of the temperature, T, and N in equation [2]. It is derived from an empirical approximation to the 1-D Debye model [12].

2

1 2

1

7 1 6

7 6



 

 +



 

=

θ T . θ

T . R N

Cp [2]

The heat capacity is the ability to store energy as a function of temperature, so integrating over temperature gives the thermal energy, HT. The zero-point cohesive energy, Ecoh, is the binding energy at the minimum of the potential energy well at 0K.

A temperature dependent value of the cohesive energy is approximated by assuming a 50% loss through the glass transition, applied using the Gaussian distribution function described earlier. The sum of HT and Ecoh, including the approximations for HC and HM, give a value of Etotal as a function of temperature.

The linear thermal expansion coefficient, αl, is defined in terms of the previously predicted temperature dependent heat capacity and cohesive energy and is shown in equation [3].

coh p

l R E

C 3

38 .

= 1

α [3]

The thermal expansion coefficient is a change in dimension as a function of temperature, so integrating over temperature, using the van der Waal’s volume, Vw, as a scaling factor, gives the volume, V. The temperature dependent energy and volume

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(8)

For Peer Review Only

terms of equation [1] have now been determined and in principle any property of interest can now be determined. The elastic bulk modulus, Be, can be defined as an energy density term and this relationship is shown for amorphous polymers in equation [4].

V

Be =18 Etotal [4]

The term defined in equation [4] refers to the elastic contribution to the full bulk modulus (i.e. the spring in the spring-dashpot model). The full viscoelastic term is obtained by applying the loss tangent profiles for the glass and beta transitions. An expression for the tensile modulus, E, is obtained using the known relationship between the loss tangent and the temperature gradient of the previously defined elastic bulk modulus [14]. This leads to an expression for E as shown in equation [5].

( )

M θ

L A x

where B

A

B

E

e β g

e

1

10 5

5 . tan 1

exp tan

 =



− +

= [5]

The proportionality constant, A, is expressed in terms of standard GIM parameters defined earlier. The temperature dependent tensile modulus predicted by equation [5]

can be compared to experimental values. However, the resin samples undergo brittle failure in tension so compressive moduli are used instead. The comparison is valid as tensile and compressive moduli are very similar at the low strain values used in this work. Once the bulk and tensile moduli are known, Poisson’s ratio, ν, can be determined along with the shear modulus.

Finally, the stress-strain curves are obtained by predicting the stress, σ, and strain, ε, separately as a function of a dummy variable. The dummy variable set is chosen as the temperature range from the observation temperature, T0, (often room temperature) to a value T1 which is arbitrarily chosen to be well above Tg. The elastic contribution to the strain, εe, is determined by integrating the thermal expansion coefficient over temperature (which is analogous to the prediction of volume). This is shown in equation [6] along with the expression for the full strain which also includes the viscous part of the strain.





 +

=

=

∫ ∫

1

0 1

0

tan 1

&

T

T e T

T l

e α dT ε ε δ dT

ε [6]

Now that the tensile modulus and strain as a function of temperature are known, the tensile stress can be determined. The compressive stress, σc, is then obtained by correcting by a factor of twice Poisson’s ratio to account for expansion in the two axes normal to the compression as shown in equation [7].

ν α

σ 2

1

0

=

T

T l c

dT E

[7]

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(9)

For Peer Review Only

2.2. Experimental

For validation, the GIM predicted properties are compared to equivalent experimental values where possible. A series of experimental data has been gathered on the compressive modulus and compressive stress-strain curves for TGDDM/DDS, TGAP/DDS and a 50-50 blend of the two at three different strain rates [15]. The epoxy and amine monomers are mixed in the appropriate ratio and heated to 120 oC then de-gassed for at least 30 minutes. The samples are cured in cylindrical moulds according to the following schedule. Ramp to 130 oC at 3 K/min, dwell for 1 hour, ramp to 180 oC at 2 K/min, cure for 2 hours and leave to cool overnight. Uniaxial compression testing was performed using a Hounsfield Universal instrument (H100KS/05) on enough samples to provide an acceptable statistical average.

3. Results and Discussion

3.1 Group Interaction Modelling of the Mechanical Properties of an Epoxy Blend

The GIM method as described above has been applied to TGDDM and TGAP cured with DDS and a 50-50 blend of the two epoxies. Table 1 shows the important GIM input parameters for the constituent functional groups present in either the epoxy or amine monomers. The GIM parameters for the epoxy and amine mer units are summed from their constituent functional groups by reference to the monomer structures in Figure 1. Finally, the mer units for cured TGDDM/DDS, TGAP/DDS and a 50-50 blend of the two are reported, taking into account epoxy-amine ratio, degree of cure and the presence of crosslinking and hydrogen bonding. These values are used in the equations presented in the previous section to predict properties.

The epoxy amine ratio used for TGDDM/DDS was 1:1 due to both monomers having a functionality of 4. The epoxy amine ratio for TGAP/DDS is 4:3 as TGAP has one less epoxy group. However, this is changed slightly due to the increased likelihood of the epoxy-hydroxyl (E-OH) reaction (reaction 3 in Figure 2) in TGAP/DDS. The exact reason for this is not clear but it seems likely due to increased mobility in TGAP over TGDDM. An illustration of this is to consider the molecular structure as the E- OH reaction becomes a possibility. At this stage, the network is forming and reducing the mobility of reactants but the impact this has on the two epoxies is different. In TGDDM the mobility is lower because 3 of 4 epoxy groups are reacted and it is unable to rotate to meet an incoming reactant. In TGAP the mobility is higher because only 2 of 3 epoxy groups are reacted and the molecule is therefore more able to rotate and hence is more able to participate in the E-OH reaction.

In experimental work, the amine is normally present in excess to ensure as high a degree of cure as possible. The experimental samples tested in this work used the standard recipe of 74 and 64 % by weight DDS for TGDDM and TGAP respectively.

However, the epoxy amine ratios used in GIM are near to the theoretical values because it is assumed that the parts of the network structure that impart properties are fully cured. A number of potential diluents can be present in the cured resin (such as unreacted hardener, impurities and water) and are known to affect properties [16].

However, the GIM predictions do not currently require this input in order to be

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(10)

For Peer Review Only

accurate. Work is progressing on developing GIM to predict the effect water (and other diluents) has on the properties of cured epoxy resins [17].

The tensile modulus, calculated from equation [5] is presented in Figure 3 as a function of temperature and blend type at a strain rate of 0.00167 /s. The step-wise drops in modulus as the temperature increases are a direct result of the beta and glass transitions. The broad beta transition is centred around ~ -50 oC while the glass transition occurs around 275 oC and has a more abrupt and catastrophic effect on the modulus. At low temperatures (< -100 oC) TGAP has a higher modulus than TGDDM, presumably a manifestation of the former being smaller and able to pack more efficiently at lower temperatures. As the temperature increases the modulus of TGDDM rises above that of TGAP as observed experimentally.

Compressive stress-strain curves, calculated from equations [6] and [7] are presented in Figure 4 as a function of blend type at a strain rate of 0.00167 /s. The curves are typical for this class of polymers with an initial elastic section at extremely low strains followed by increasing viscous contributions. Eventually the polymer yields as large scale translation between adjacent polymers causes ductile failure. The yield point is defined here as the point at which the gradient of the curve reaches zero. The TGDDM curve is consistently above that of TGAP which is a reflection of the fact that the simulations were performed at room temperature.

A comparison between the GIM predicted and experimental stress-strain curve for the 50-50 blend at a strain rate of 0.00167 /s is presented in Figure 5. The curves agree very well and in particular the yield condition is predicted with good accuracy. There is some post yield strain softening and hardening present in the experimental curve which is not replicated in the GIM predicted curve. This is not currently part of the model but will be in future versions as predicting such effects requires a simple change in the value of N beyond the yield condition to account for the activating of extra degrees of freedom.

One of the key aspects of GIM is its ability to predict properties as a function of strain rate. Figures 6, 7, and 8 compare GIM predicted and experimental moduli and yield stress values as a function of strain rate and blend type. Each pair of lines represents a GIM to experiment comparison and in general the agreement between the two is excellent. In particular the yield stresses are predicted very accurately across all blends and strain rates. The predicted modulus values at the lower rates are slightly inaccurate but this may be in part due to creep effects. The trends observed in both modulus and yield stress is for the values to increase with increasing strain rate. This is consistent with the viscoelastic model presented earlier where loss events are caused by molecular relaxation processes. As the strain rate increases, the molecular processes have less chance to occur and hence less loss is observed which in turn leads to a higher modulus and yield stress.

3.2 Modelling Structural Changes of the Epoxy Resin

A significant advantage of GIM over similar techniques is the relatively simple method used to define the polymer system where the representative mer unit is described using its constituent functional groups. This means that the chemistry can be easily altered to investigate any changes to the predicted properties. Figure 9 shows

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(11)

For Peer Review Only

the modulus and yield stress of TGDDM/DDS as a function of the epoxy-amine ratio for a given strain rate and temperature. The values at 50% epoxy (i.e. 50% epoxy mer units, 50% amine mer units) are the same as those reported earlier while increasing or decreasing the amount of epoxy has a consistent effect on the properties. As expected, adding more hardener increases the modulus and yield stress while more epoxy has the opposite effect. It is worth noting that practical issues are an issue here as there would be little benefit in attempting to make a resin where 80% of the monomer molecules were either epoxy or amine. However, an appropriate amount of epoxy could be between 40 and 60% (mer units) and within this range the properties don’t vary as much as might be expected. The plateau value of both properties at high epoxy values indicates that there is a limiting low value for each property where ductile effects dominate. As the amount of hardener is increased the brittle nature of the resin increases until a balance between ductile and brittle effects is obtained defining the working range of the resin.

There are two common isomers of the amine hardener used in this work, 4,4-DDS and 3,3-DDS, differing only in relative positions of the amine and sulphonyl groups on the phenyl rings. In 4,4-DDS (as shown in Figure 1) the groups are arranged in the para position which allows the phenyl ring to rotate more easily than in 3,3-DDS where the groups are in the meta position. In GIM parameter terms, the phenyl ring in 3,3-DDS has a higher number of degrees of freedom as it’s rotation is not symmetrical. Figure 10 shows the modulus and yield stress of TGDDM/DDS as a function of the percentage of 4,4-DDS for a given strain rate and temperature. The results show that a higher percentage of 4,4-DDS leads to higher modulus and yield stress, a consequence of the fact that the 4,4 isomer dissipates less energy through molecular relaxation than the 3,3 isomer.

Another way to alter the chemistry of the resin is to change the functional groups present. A series of epoxies exist based on whichever variant of bisphenol was used in their manufacture. The most common is the diglycidyl ether of bisphenol-A (or DGEBA, see Figure 1) where two epoxy-ether segments surround a central phenyl- CMe2-phenyl group. The central CMe2 group can be changed to vary the chemistry of the epoxy and is denoted X for convenience. The GIM parameters for six different X groups are given in Table 2. The various functional groups contribute the same number of degrees of freedom to the mer unit as they are all just a single link in the polymer chain. The differences are seen in the cohesive energy and the volume and this leads to different GIM parameters for the five mer units as shown in Table 3. The modulus and yield stress are plotted as a function of the central X group in Figure 11 for a given strain rate and temperature when cured with DDS. The decrease in both modulus and yield stress when moving from DGEBF to DGEBB is a consequence of having more alkyl groups present. These are able to dissipate more energy and the properties are correspondingly lower. The value of DGEBS is high as the SO2 group is present which has a ratio of cohesive energy to thermal energy which results in high modulus and yield stress. Indeed, the SO2 group is one of the main reasons DDS is an effective, popular hardener. DGEBC is used as a fire retardant in polycarbonate and its properties are included here to show that GIM is also capable of predicting the properties of more exotic structures. In the case of DGEBA/DDS, the comparison between GIM predicted modulus (2.35 GPa) and experiment (~ 2.3 GPa [18]) is very good.

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(12)

For Peer Review Only

3.3 Modelling of Composite Strength

The work presented here exists as part of a larger multi-scale approach to the prediction of composite properties. GIM has proven effective at predicting the non- linear properties of an epoxy resin matrix as a function of temperature, strain rate and chemistry. These properties are then used to parameterise a finite element model which predicts the effect a fibre failure has on the surrounding fibres. The impact a fibre failure has on its neighbours is quite different depending on whether the matrix is elastic or viscoelastic [2]. A 1/12th slice of a 100 µm long unit cell of UD composite is created with hexagonal symmetry in Ansys 11. The unit cell is constructed using the SOLID45 element so that the fibre fraction is 50% as shown in Figure 12.

Orthotropic properties were applied to the fibre elements (acquired from experimental measurements on HTA3151 carbon fibres [19]). Isotropic properties were applied to the matrix elements (obtained from the previously described GIM calculations on TGAP/DDS resin). In order to simulate the fibre break, the top layer of elements is removed from the central fibre only. A 1% axial strain is then applied to the system so that the front face displaces outwards and the central fibre remains in place. The strain concentrations associated with the broken fibre are measured in the surrounding fibres. The ineffective length of the broken fibre is defined as twice the length over which 90% strain recovery occurs.

A statistical method is then used to predict the relationship between the strength of the composite and its size using the finite element results as input. A single layer of composite of thickness equal to the ineffective length consisting of a 20x20 array of fibres is assigned random failure strains. The applied strain is increased until the fibre with the minimum failure strain is broken and the strain concentrations are applied to its neighbours. Should any of the neighbouring fibres break as a result of this strain concentration, the process is repeated. Then the applied strain is increased to the next lowest failure strain causing another break and potentially further breaks in neighbouring fibres. The process is repeated until unstable crack growth criteria are achieved (e.g. 3 broken fibres in a row or 2% broken fibres). The applied strain is then equal to the layer failure strain. The whole process is repeated for 50 to 100 iterations until self-consistency is achieved and the average layer failure strain is recorded. The length is scaled up using standard probability tables to give a relationship between composite strength and composite length. A typical set of results is shown in Figure 13 where a plateau value of strength is observed at high lengths. The original predictions by Curtis [20] have been developed here to allow strength predictions using an elastic (no energy dissipation) or viscoelastic (dissipation via beta and glass transitions) matrix. The results indicate that the yielding matrix predictions are nearest to typical experimental values. Ultimately, the multi-scale approach presented here shows how composite strength varies with the nature of the matrix. Subtle changes to the chemistry of the matrix have a corresponding effect on the properties of the composite, and the use of GIM has allowed this to be quantified for perhaps the first time.

4. Concluding Remarks

A multi-scale modelling approach has provided insights into the importance of the viscoelastic nature of the matrix phase in composite systems. The properties of the

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(13)

For Peer Review Only

matrix phase were predicted using the mean-field based Group Interaction Modelling technique. Group Interaction Modelling allows the pragmatic calculation of a full set of non-linear volumetric, thermomechanical and engineering properties of a thermosetting multifunctional epoxy resin system. The comparison of predicted values with experimental values is very good across a range of temperatures and strain rates. The matrix phase was also altered at a fundamental molecular level to provide predicted properties as a function of chemistry. This provided predictions of properties as a function of the epoxy-amine ratio, the amine isomer and also the specific chemical structure of the epoxy. The matrix phase properties were used in a finite element model to predict the effect a broken fibre has on its neighbours. This was then used in a statistical model to predict the propagation of fibre failure events through a composite system.

Acknowledgments

This work was carried out as a part of the Weapons and Platform Effectors Domain of the UK MoD Research Program.

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(14)

For Peer Review Only

Figure Captions

Figure 1. Chemical structures of DGEBA, TGDDM, TGAP and DDS monomers.

Figure 2. General reaction scheme for amine curing an epoxy resin.

Figure 3. Plot of GIM predicted tensile modulus against temperature for TGDDM/DDS, TGAP/DDS and a 50-50 blend of the two at a strain rate of 0.00167 /s.

Figure 4. Plot of GIM predicted stress-strain curves for TGDDM/DDS, TGAP/DDS and a 50-50 blend of the two at a strain rate of 0.00167 /s.

Figure 5. Plot of GIM predicted and experimental compressive stress-strain curves for a 50-50 blend of TGDDM/DDS and TGAP/DDS at a strain rate of 0.00167 /s and room temperature.

Figure 6. Plot of GIM predicted compressive modulus and yield stress for TGDDM/DDS as a function of strain rate at room temperature. Experimental points are shown for comparison purposes.

Figure 7. Plot of GIM predicted compressive modulus and yield stress for a 50-50 blend of TGDDM/DDS and TGAP/DDS as a function of strain rate at room temperature. Experimental points are shown for comparison purposes.

Figure 8. Plot of GIM predicted compressive modulus and yield stress for TGAP/DDS as a function of strain rate at room temperature. Experimental points are shown for comparison purposes.

Figure 9. Plot of the GIM predicted compressive modulus and yield stress as a function of epoxy-amine ratio at a strain rate of 0.00167 /s and room temperature. The resin is TGDDM/DDS.

Figure 10. Plot of the GIM predicted compressive modulus and yield stress as a function of the percentage of 4,4-DDS over 3,3-DDS at a strain rate of 0.00167 /s and room temperature. The epoxy is TGDDM.

Figure 11. Plot of the GIM predicted compressive modulus and yield stress as a function of the type of Bisphenol used to create the epoxy monomer. The central X group is noted for each epoxy and the strain rate is 0.00167 /s at room temperature.

Figure 12. The hexagonal unit cell used in the finite element model. A fibre-break is simulated by removing the top layer of cells from the central fibre only and applying a 1% longitudinal strain to the rest of the system.

Figure 13. Plot of the relationship between the predicted UD carbon fibre composite failure strain and composite length. The data is presented for a viscoelastic TGAP/DDS matrix and a purely elastic matrix. The dotted line represents a typical experimental failure strain for a UD carbon fibre composite.

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(15)

For Peer Review Only

Table 1. GIM parameters for functional groups, mer units and cured mer units.

Group N Ecoh (J/mol) Vw (cm3/mol)

CH2 2 4,500 10.25

Phenyl 3 25,000 43.3

N 2 9,000 4

Epoxy 4 15,300 22

CH(OH) 2 20,800 11.5

SO2 2 45,000 20.3

O 2 6,300 5

TGAP Mer Unit 25 129,700 148.3 TGDDM Mer Unit 36 191,700 232.9 DDS Mer Unit 12 113,000 114.9 E-OH Mer Unit 22 119,700 148.3 Cured TGAP/DDS 19 121,116 134.0 Cured 50-50 Blend 21.5 136,732 153.9 Cured TGDDM/DDS 24 152,350 173.9

Table 2. GIM parameters for the central X group for various diglycidyl bisphenol epoxies.

Variant X N Ecoh (J/mol) Vw (cm3/mol)

F CH2 2 4,500 10.2

E CHMe 2 9,000 20.4

A CMe2 2 13,500 30.6

B CMeEt 2 18,000 40.8

S SO2 2 45,000 20.3

C C=CCl2 2 32,000 37.0

Table 3. GIM parameters for various diglycidyl bisphenol epoxy monomers.

X N Ecoh (J/mol) Vw (cm3/mol) DGEBF CH2 24 126,700 171.3

DGEBE CHMe 24 131,200 181.5 DGEBA CMe2 24 135,700 191.7 DGEBB CMeEt 24 140,200 201.9 DGEBS SO2 24 167,200 181.4 DGEBC C=CCl2 24 154,200 198.1

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(16)

For Peer Review Only

References

1 H. Q. Pham & M. J. Marks, 'Epoxy Resins' in Encyclopedia of Polymer Science and Technology, Wiley, 2002.

2 S. Behzadi, P. T. Curtis, F. R. Jones, Composites Science and Technology, 69 (2009) p.2421.

3 A. Gupta, M. Cizmecioglu, D. Coulter, R. H. Liang, A. Yavrouian, F. D. Tsay, J.

Moacanin, J. Appl. Pol. Sci., 28 (1983) p.1011.

4 D. Porter, Group Interaction Modelling of Polymer Properties, Dekker, New York, 1995.

5 J. Bicerano, Prediction of Polymer Properties. Dekker, New York, 2002.

6 J. P. Foreman, D. Porter, S. Behzadi, F. R. Jones, Polymer, 49 (2008) p.5588.

7 D. Porter, P. J. Gould, International Journal of Solids and Structures 46 (2009) p.1981.

8 J. P. Foreman, D. Porter, S. Behzadi, K. P. Travis, F. R. Jones, Journal of Materials Science, 41 (2006) p.6631.

9 J. P. Foreman, S. Behzadi, S. A. Tsampas, D. Porter, P. T. Curtis, F. R. Jones, Plastics, Rubber and Composites, 38 (2009) p.1.

10 J. P. Foreman, S. Behzadi, D. Porter, P. T. Curtis, F. R. Jones, Journal of Materials Science, 43 (2008) p.6642.

11 D. W. Van Krevelen, Properties of Polymers. Elsevier, Amsterdam, 1993.

12 B. Wunderlich, S. Z. D. Cheng, K. Loufakis, in Encyclopedia of Polymer Science and Engineering, Volume 16, Wiley, New York, 1989.

13 L. M .Robeson, A. G. Farnham, J. E. Mcgrath in Molecular Basis of Transitions and Relaxations, Gordon and Breach, New York, 1978.

14 A. Bondi, Physical Properties of Molecular Crystals, Liquids, and Glasses, Wiley 1969.

15 S. Behzadi, PhD Thesis, University of Sheffield, UK, 2006.

16 M. I. Caballero-Martinez, PhD thesis, University of Sheffield, UK, 2004.

17 J. P. Foreman, D. Porter, F. R. Jones, manuscript in preparation

18 S. R. White, P. T. Mather, M. J. Smith. Polymer Engineering and Science, 42 (2002) p.51.

19 M. R. Nedele, PhD Thesis, University of Bristol, UK, 1996.

20 P. T. Curtis, Composites Science and Technology, 27 (1986) p.63.

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(17)

For Peer Review Only N CH

2

N

O

O O

O

C O

O

O O

CH

3

CH

3

SO

2

NH

2

H

2

N

O

O N

O

O

DGEBA

TGDDM

TGAP

DDS

Figure 1. Chemical structures of DGEBA, TGDDM, TGAP and DDS monomers.

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(18)

For Peer Review Only

NH2

+

NH CH2 CH

OH

NH N CH2 CH

OH

CH2 CH CH2 OH

CH2 CH CH2 O

CH2 CH HO

1

2

3 O

+ +

O O

Figure 2. General reaction scheme for amine curing an epoxy resin.

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(19)

For Peer Review Only

0 2000 4000 6000 8000 10000 12000

-300 -200 -100 0 100 200 300 400

Temperature, T (oC)

Tensile Modulus, E (MPa) . TGDDM

50-50 TGAP

Figure 3. Plot of GIM predicted tensile modulus against temperature for TGDDM/DDS, TGAP/DDS and a 50-50 blend of the two at a strain rate of 0.00167 /s.

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(20)

For Peer Review Only

0 50 100 150 200 250

0 0.05 0.1 0.15 0.2 0.25 0.3

Strain, εεεε Compressive Stress, σσσσc (MPa) .

TGDDM 50-50 TGAP

Figure 4. Plot of GIM predicted stress-strain curves for TGDDM/DDS, TGAP/DDS and a 50-50 blend of the two at a strain rate of 0.00167 /s.

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(21)

For Peer Review Only

0 50 100 150 200 250

0 0.05 0.1 0.15 0.2 0.25 0.3

Strain, εεεε Compressive Stress, σσσσc (MPa) .

GIM Experiment

Figure 5. Plot of GIM predicted and experimental compressive stress-strain curves for a 50-50 blend of TGDDM/DDS and TGAP/DDS at a strain rate of 0.00167 /s and room temperature.

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(22)

For Peer Review Only

3600 4000 4400 4800 5200 5600

0.0001 0.001 0.01 0.1

Strain Rate (/s)

Modulus (MPa) .

170 190 210 230 250 270

Yield Stress (MPa) .

Experiment Modulus GIM Modulus

Experiment Yield Stress GIM Yield Stress

Figure 6. Plot of GIM predicted compressive modulus and yield stress for TGDDM/DDS as a function of strain rate at room temperature. Experimental points are shown for comparison purposes.

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(23)

For Peer Review Only

3400 3800 4200 4600 5000 5400

0.0001 0.001 0.01 0.1

Strain Rate (/s)

Modulus (MPa) .

170 180 190 200 210 220 230 240 250

Yield Stress (MPa) .

Experiment Modulus GIM Modulus

Experiment Yield Stress GIM Yield Stress

Figure 7. Plot of GIM predicted compressive modulus and yield stress for a 50-50 blend of TGDDM/DDS and TGAP/DDS as a function of strain rate at room temperature. Experimental points are shown for comparison purposes.

3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Références

Documents relatifs

Hallett, Effect of high through-thickness compressive stress on fibre direction tensile strength of carbon/epoxy composite laminates, Composites Science and Technology,

On obtient ainsi une forme affaiblie des assertions (i) et (ii) de l’´enonc´e : pour (i), on prouve seulement que la fonction est C m 0 en chaque variable ; pour (ii), on impose que

To answer this question, we will derive site-specific power laws relating the horizontal travel distance ∆L 0 to the unstable volume V, using a purely empirical approach, a

Therefore, in order to appropriately model the transport of dust and its associated optical properties and mass impacts, dust models must attempt to capture both the broad size range

Les personnes issues de l’immigration maghrébine qui déclarent une très faible importance relative accordée au travail, présentent des modèles d’identité très clivés selon

mortar dried (violet), humid (blue) and carbonated (green), unirradiated and irradiated in regards with 308. the mean of

The aim of this study is to evaluate the impact of post-curing temperatures on the mechanical properties of a unidirectional natural fiber reinforced composites (NFC)

CHF 58 million over 6 years, the Strategy relied on three priority domains: (i) Rural Income and Em- ployment through the creation of market and job opportunities for the rural