Anda di halaman 1dari 13

Proceedings of the Institution of Mechanical Engineers, Part L: Journal of Materials Design and Applications http://pil.sagepub.

com/

Low coefficient of thermal expansion of thermoset composite materials


I Rhoney and R A Pethrick Proceedings of the Institution of Mechanical Engineers, Part L: Journal of Materials Design and Applications 2012 226: 76 originally published online 13 October 2011 DOI: 10.1177/1464420711424003 The online version of this article can be found at: http://pil.sagepub.com/content/226/1/76

Published by:
http://www.sagepublications.com

On behalf of:

Institution of Mechanical Engineers

Additional services and information for Proceedings of the Institution of Mechanical Engineers, Part L: Journal of Materials Design and Applications can be found at: Email Alerts: http://pil.sagepub.com/cgi/alerts Subscriptions: http://pil.sagepub.com/subscriptions Reprints: http://www.sagepub.com/journalsReprints.nav Permissions: http://www.sagepub.com/journalsPermissions.nav Citations: http://pil.sagepub.com/content/226/1/76.refs.html

>> Version of Record - Jan 6, 2012 OnlineFirst Version of Record - Oct 13, 2011 What is This?

Downloaded from pil.sagepub.com by guest on November 22, 2013

76

Low coefficient of thermal expansion of thermoset composite materials


I Rhoney and R A Pethrick* Department of Pure and Applied Chemistry, University of Strathclyde, Glasgow, UK
The manuscript was received on 14 March 2011 and was accepted after revision for publication on 30 August 2011. DOI: 10.1177/1464420711424003

Abstract: Selection of the correct resin filler combination is important in achieving materials which have a low coefficient of thermal expansion (CTE). A study of nanosilica-modified resin, in combination with silica fillers, allows low levels of CTE to be achieved. In this study, a novel curing agent, ytterbium triflate, is reported. This curing agent provides a stable catalyst system which can be used to create the viscous composite mixtures but has the facility of effective cure over a relatively narrow temperature range from 70  C to 100  C. A series of formulations were examined based on incorporation of fillers with nanoscale silica particles into either the pure resin or a resin which contained nanoscale functionalized silica particles. The filler incorporation leads to a significant increase in the glass transition temperature as determined by dynamic mechanical thermal analysis. The CTE was observed to be lower below Tg than above it. Changing the nanosilica particle size and distribution produced significant changes in the values. However, the CTE scaled according to the total silica content; and the values were in general lower than those calculated theoretically. The use of a highly functional o-cresol epoxy novolac demonstrated how increasing functionality raised the Tg and lowered the CTE, but the use of too high a post-cure temperature reversed this trend. Very good results were achieved using 3,4-epoxycyclohexylmethyl, 3,4 epoxycyclohexanecarboxylate, which has a low viscosity and allowed high levels of silica to be readily achieved. This article indicates how the adjustment of the epoxy selected to be used as the base material and the type of silica particles used allows the values of Tg and CTE to be modified in a composite material. Keywords: thermal expansion coefficient, ytterbium triflate catalysts, epoxy resin, silica fillers, nanocomposites, glass transition temperature, cure data

INTRODUCTION

lower limits of the Schapery theory, which is based on the simple rule of mixing c f  m 1  1

In a number of applications, it is desirable to have thermoset materials which have a low coefcient of thermal expansion coefcients (CTE). A composite, by its very nature, will have a CTE which reects the proportions of the resin and ller in the material [1]. Studies of a range of composites have shown that the CTE can be predicted to lie between the upper and

*Corresponding author: West Chem, Department of Pure and Applied Chemistry, University of Strathclyde, Thomas Graham Building, 295 Cathedral Street, Glasgow G1 1XL, UK. email: r.a.pethrick@strath.ac.uk
Proc. IMechE Vol. 226 Part L: J. Materials: Design and Applications

where c, f, and m are, respectively, the CTEs of the composite, ller, and matrix, respectively, and  the volume fraction of the ller [2]. The limits are dictated by the CTE of the ller, which has a value of 8.5 106 per  C for glass, and that of a typical thermoset resin which has a value of 90 106 [3, 4]. In practice, the values observed depend on a series of complex factors which relate to the density of the polymer matrix and the volume fraction of low-CTE llers which can be incorporated into the material. With the advent of nanoscale materials, the possibility arises in

Downloaded from pil.sagepub.com by guest on November 22, 2013

Low coefficient of thermal expansion of thermoset composite materials

77

exploring the way in which small particle llers may be used to achieve low values of CTE in composite structures. Deviations from the simple additivity relation equation (1) have long been recognized [5] and this article explores the effects of processing conditions on the observed values of CTE obtained with some novel nanocomposites. In previous papers, the effects of clay nanoparticles on the rheological and thermal mechanical properties of various epoxy resin composites have been reported [611]. The addition of low levels of clay platelets is found to have a signicant effect both on the rheology and thermal properties such as the glass transition temperature (Tg). The platelets are able to interact and signicant increase of the viscosity is observed with levels as low as a few weight percents. Similarly, the glass transition increased in certain cases by several tens of degrees. However, in a number of cases, no enhancement was observed and this was attributed to the clay platelets inhibiting the formation of a dense matrix. In a recent paper [11], we report the use of a novel curing system, ytterbium triate, which introduces the possibility of a pseudolatent type of cure for epoxy resins. The object of this article is to explore how it is possible to achieve materials with low CTE values by using various combinations of nanollers and cure regimes. 2 EXPERIMENTAL

materials were used as received. The epoxy resins were thoroughly degassed prior to mixing and after being mixed with the appropriate quantity of ller. The mixtures were in general stored overnight in a desiccator before catalysts were added and cure carried out. The cure was undertaken in an aluminium mould of dimensions; 10 10 2 cm3, which had been treated with a release agent and carried out in a vacuum oven which had been preheated to the required temperature. In order to avoid exotherms, certain materials were rst cured at a lower temperature and then post-cured at a higher temperature. The cured lms were then subjected to a series of thermal tests.

2.2 Coefficient of thermal expansion Measurements of CTE were carried out using a Mettler TMA 40 (Mettler-Toledo Limited, Leicester. A sample of dimension 1 1 cm2 was heated from 30  C to 250  C at 20  C/min, then cooled from 250  C to 30  C at 20  C/min, and then held at 30  C for 20 min, before being heated from 30  C to 250  C at 4  C/min. The rst and second runs were to relax the sample and erase any thermal history. The nal run involves slowly increasing the temperature and allows determination of the glass transition temperature Tg and the expansion coefcients above and below it. Before the sample was inserted in the instrument, the quartz probe was zeroed. The sample was placed on the sample stage and the probe lowered to sit on top of the sample. The thickness of the sample was recorded for future calculations. The supporting software allows analysis of the gradient at a given

2.1 Materials The characteristics of the resin and the llers are summarized in Table 1. The catalyst, ytterbium triate, was obtained from SigmaAldrich, UK. The
Table 1
Name

General characteristics of materials used in this study


Source Comment

Epoxy resins Bisphenol A Huntsman Bisphenol F Huntsman 3,4-Epoxycyclohexylmethyl, 3,4 Aldrich Chemicals epoxycyclohexanecarboxylate EEC Nanopox A610 Nanoresins AG Monomer EEC contains 40 wt% SiO2 20 nm particles Nanopox A510 Nanoresins AG Monomer DGEBF contains 40 wt% SiO2 20 nm particles Nanopox A410 Nanoresins AG Monomer DGEBA contains 40 wt% SiO2 20 nm particles Tris(4-hydroxyphenyl)methane Aldrich Chemicals triglycidyl ether (THP) YDCN-500-IP a novolac resin Kukdo Chemical Co.o-Cresol novolac EEW 190210 g/eq. Soft point 5054  C Silica Aerosil OX50 Aerosil 90 Aerosil 380 Glass flake GF100 nm Exfoliated clay Cloisite 30B Degussa AG Degussa AG Degussa AG Surface area 50 m2/g; size 40 nm; density 130 g/L Surface area 90 m2/g; size 20 nm; density 80 g/L Surface area 380 m2/g; size 7 nm; density 50 g/L

Glassflake Limited 100 nm thick Southern clay Montmorillonite modified with quaternary ammonium MT2EtOH

Proc. IMechE Vol. 226 Part L: J. Materials: Design and Applications


Downloaded from pil.sagepub.com by guest on November 22, 2013

78

I Rhoney and R A Pethrick

temperature. Where the graphs indicated a linear CTE, an average value was calculated below and above Tg. Where the plots showed a marked curvature, the values are quoted at a specic temperature. The temperature calibration of the equipment was checked using indium and lead. The unit of CTE is ppm/  C. The data reported are an average of a minimum of three experiments. Values were repeated until the deviation between individual measurements were less than 2 ppm/  C. 2.3 Dynamic mechanical thermal analysis

and 2  P1 ::!:A =k  P 1 :!:A =k 2 4

Using the above equations, it is possible to calculate the shear viscosity from the in-phase and quadrature coefcients. The data were recorded using Picolog software package. The gelation was taken as the point at which the viscosity of the material reaches a value of 104 Pa s. 3 RESULTS AND DISCUSSION

The method used has been previously described. The dynamic mechanical thermal analysis (DMTA) measurements were made using MKIII DMTA instrument (Polymer Laboratories, Church Stretton, UK). The samples were of dimensions 5 6 3 mm3 and cut from the original plaques using a diamond saw. Temperature scans were performed over a temperature range from 100  C to 230  C, at frequency 1 Hz, strain 4, and scanning rate 4  C/min. The samples were clamped with a torque of 40 Nm.

2.4 Strathclyde curometer The change in viscosity was measured as a function of time using the Strathclyde curometer [12, 13]. A small glass vial containing the sample was held in an oil bath. The temperatures of the oil bath and the sample were adjusted and maintained isothermally during the period of the cure. The paddle probe was lowered into the sample and operated at a frequency of 2 Hz and amplitude 500 mm. The sensor output is the amplitude and phase of the probe motion, which is directly related to the viscosity changes occurring in the sample. The amplitude of the motion of the probe which is coupled to the driver by a spring is damped due to viscosity. If P1(t) and P2(t) are the instantaneous displacements of the oscillatory driver and the response of the system damped by the curing liquid, then k P1 t P2 t :A : dP2 dt 2

where k is the spring constant coupling the driver to the probe,  the shear viscosity of the uid, and A the geometric factor related to the probe contact area. Separating real and imaginary parts of the complex probe motion, one obtains _1 P _2   P 1 :!:A =k 2 3

In this article, cure of the epoxy resins was achieved using a cationic catalysed cure system based on ytterbium (Yb) triate (11). This cure system is very efcient at high temperatures, while at low temperatures, the activated mixture can be stable for many hours/months. A catalysed sample was stored in a refrigerator for a period of 6 months and it was found that there was no increase in its viscosity or loss of its reactivity. A preliminary study of the effects of temperature and catalyst concentration on the cure characteristics of one of the epoxy EEC (Fig. 1). The in-phase displacement reects the motion of the probe as this is directly coupled to the spring. The out-of-phase displacement measures the phase shift of the driver against the probe and is a reection of the energy dissipation in the spring which is itself a measure of the energy dissipation in the uid. Increase in the temperature of the cure produces a shift of the peak in the out-of-phase displacement and a coincidental drop in the in-phase displacement at shorter times. The drop in the in-phase displacement is a reection of the increase in the viscosity of the media and the peak in the out-of-phase displacement a consequence of the growing viscoelastic nature of the media. Gelation of the material corresponds to a point close to the peak in the out-of-phase displacement and this can be used to assess the effect of temperature on the cure. The point at which the in-phase displacement approaches the baseline coincides with vitrication of the material. The breadth of the peak reects the time taken to transform from a gel to a glassy state. A plot of the variation of the gelation time against reciprocal temperature is shown in Fig. 2. The plot demonstrates a simple Arrhenius type of behaviour and the activation energy was calculated from the slope of the curve and found to be 101 3 kJ/ mol. A study was undertaken in which the concentration of the Yb triate was varied from 0.08 to 0.15 phr and it was found that the activation energy remained essentially constant.

Proc. IMechE Vol. 226 Part L: J. Materials: Design and Applications


Downloaded from pil.sagepub.com by guest on November 22, 2013

Low coefficient of thermal expansion of thermoset composite materials

79

2.5x103 2.0x103 In Phase Displacement 1.5x103 1.0x103 5.0x102 0.0 0 2000 4000 6000 8000 10000 12000

250

200 Out of Phase Displacement

150

100

50

0 14000

Time (secs)

Fig. 1

Plot of the in phase {left axis} and out of phase displacement {right axis} as a function of time for cure reactions carried out at: (#) 100  C, () 110  C, () 120  C, and (n) 130  C. The units are arbitrary as they related to the spring used

148.41316

Ln (Time (mins))

54.59815

20.08554

7.38906

2.71828 0.0023 0.0024 0.0025 1/Temperature 0.0026 0.0027

Fig. 2

Plot of gelation time against reciprocal temperature (K1)

3.1 Incorporation of platelet fillers Equation (1) is purely based on volume fraction and it may be argued that nanoscale platelet llers may in principle be able to reduce the CTE to a greater extent that would be predicted on the basis of their volume fractions. Two types of platelet llers were studied: organically modied clay Cloisite 30B and glass akes with nanoscale thickness. The primary concern addressed in this study was whether the ller has any effect on the cure of the resin. 3.2 Organically modified clay-containing material For this study, Cloisite 30B was selected as this has been found in previous studies [611] to be one of

the most easily dispersed organically modied clays. Cloisite 30B contains methyltallow dihydroxyethyl ammonium ion and has platelets which are 1 nm thick and 1000 nm long. The epoxy EEC was selected for study and a mixture containing 3 wt% of Cloisite 30B was prepared and sonicated for a period of 30 min to achieve optimum exfoliation of the clay platelets. The sample was then allowed to equilibrate overnight. To this dispersion 0.10-phr Yb triate dissolved in ethanol was added and the mixture stirred at 80  C and degassed. A sample was then analysed in the curometer and it was observed that no signicant reaction had occurred after 22 h at 100  C. Attempts to achieve cure were carried out at higher temperatures but were unsuccessful and the system showed evidence of degradation occurring and the emission of acrid fumes. The clay is modied with a quaternary ammonium salt. The degradation temperature of Cloisite 30B is 174  C, which would imply that under the conditions used for the cure, no loss of the quaternary ammonium slat would be expected. Thermal degradation of these organic modiers has been shown to produce aldehydes, carboxylic acids, and amine-containing residues, and evolve ammonia and carbon dioxide. A study of the degradation products was not undertaken but it was clear that the presence of the Yb triate was able to catalyse the degradation at temperatures of the order of 130  C over a period of time. Increasing the catalyst level to 0.20 phr allowed partial cure to be achieved in 23 h. Cure was also attempted using nanomodied resins and while some degree was achieved in all cases, signicantly longer gelation times were observed compared with the pure resins. It was concluded that the
Proc. IMechE Vol. 226 Part L: J. Materials: Design and Applications

Downloaded from pil.sagepub.com by guest on November 22, 2013

80

I Rhoney and R A Pethrick

2.0x103 300 Temp raised to 140C


In Phase Displacement Out of Phase Displacement

1.5x103

250 200 Temp raised to 160C

1.0x103

150 100

5.0x102

50 0

0.0 0 10000 20000 30000 40000 50000 Time (secs) 60000 70000

Fig. 3

Plot of the variation of the in phase (top curve) and out of phase displacements (bottom curve) for a mixture catalysed with 0.3 phr of Yb and containing 3 wt% Cloisite 30B

clay was effectively inhibiting the cure. Since the Yb salt has to dissociate for the active site to initiate polymerization, adsorption at anionic sites on the clay would effectively inhibit polymerization. To test this hypothesis, the catalyst level was increased to 0.3 phr and the gelation time reduced to 4.6 h, but only after the temperature had been successively raised to 140  C and then 160  C, (Fig. 3). The in-phase displacement is the top curve, which falls with increase in time and the out-of-phase displacement increases and then falls with increasing time. 3.3 Vermiculite A sample of 2.0 wt% of vermiculite (Phyllosilicate) in EEC was prepared and stirred with a high-speed mixer (Ultra Turrax) for 1 h. Vermiculite is known to be easily exfoliated and interestingly satisfactory cure was achieved. This implies that the inhibition effects observed are peculiar to the Cloisite clays. 3.4 Nano-glass fake In principle, glass will not have the same ability to bind cations as clay and hence nano-glass akes should not exhibit the problems encountered with Cloisite. A mixture of 6.4 wt% glass ake was stirred at 100  C and allowed to stand overnight. The suspension was unstable and it was clear that precipitation of the glass was slowly occurring. The Yb catalyst was added but the mixture rapidly became viscous; but cure was only achieved after 6 h. It is apparent that if the mixture has a high initial viscosity as a consequence of the ller addition, this inhibits the subsequent cure. This effect can be attributed to the nature of the polymerization process in which diffusion of the oxirane to the
Proc. IMechE Vol. 226 Part L: J. Materials: Design and Applications

active site is required for propagation of the polymerization. If the platelets interact with the catalyst, then the polymerization will be inhibited. 3.5 Samples containing silica 0.5 mm To achieve the loading levels required by equation (1) to reduce the CTE to a low level, it will be necessary to increase the ller level signicantly above those used above. To maintain a low viscosity for processing, the ller must be in a near-spherical form. A mixture was prepared which contained 20 g EEC and 10 g of silica 0.5 mm. The large silica particles in the low-viscosity resin tend to separate out. To prevent sedimentation, 2 g Aerosil 90 was added. This signicantly increases the viscosity, suppresses sedimentation, and increases the total silica content to 37.8 wt%. This mixture has a gelation time of 7.3 min. Addition of silica does not lead to the inhibition effects observed with the platelet materials. Examination of the curometer curves for the cure of materials which contained silica 0.5 mm and with added nanosize silica, Fig. 4, indicates that the cure is inuenced by the viscosity of the media. The initial rise in the out-of-phase displacement is similar to that for the simple resin, Fig. 1; however, after the peak, it drops rapidly, indicating a rapid rise in viscosity after gelation and the system is rapidly transformed to a glass state. The glassy state is indicated as the point at which the in-phase displacement becomes close to zero. 3.6 Addition of nanosilica Two possible routes are available for the incorporation of nanosilica, either as in situ created

Downloaded from pil.sagepub.com by guest on November 22, 2013

Low coefficient of thermal expansion of thermoset composite materials

81

2.0x103

350 300

In Phase Displacement

250 200 150

1.0x103

5.0x102

100 50

0.0 0 0 200 400 600 800 1000

Time (secs)

Fig. 4

Cure curves for an EEC silica 0.5 mm with 2 g Aerosil 90 cured at 100  C with 0.1 phr of Yb catalyst

nanodispersion and available as a modied resin or as preformed particles which may then be dispersed in a normal resin. 3.7 Modified resins Ranges of modied resins are now available which contain up to 60 wt% of nanosilica particles which have been created in situ and have viscosities comparable to the unmodied resins. Since these modied resins contain up to 60 wt% of nanosilica, they provide a route for the increase of silica llers and hence lowering of the CTE. A study of Nanopox A610, which is the EEC modied with 20 nm silica particles was undertaken. As with the study with the pure resin, measurements were made at various temperatures between 100  C and 160  C and a linear plot similar to that shown in Fig. 2 was obtained. The activation energy was also investigated as a function of the catalysts loading between 0.08 and 0.15 phr and found to be independent of loading. A value of the activation energy of 76 4 kJ/mol was obtained, which is signicantly lower than the that of the pure resin 101 3 kJ/mol. Clearly, the Nanopox catalyses the reaction. The in situ creation of the nanoparticles involves growth using a solgel process. The size of the particles is controlled by the incorporation of an organically modied silane and this allows the introduction of functional groups into the surface of the particles. Aiding the development of the interface with the resin incorporation of oxirane functionalities is desirable and these may in part account for the higher reactivity of these modied resins compared with their normal counterparts.

3.8 Addition of nanosilica A range of nanosilica llers are available and can be added to the resin to further increase the silica loading. The silica investigated in this study was Aerosil OX50. The silica in this form has nominally the same size as that in the particle-modied resin but will require to be dispersed. The cure curves obtained were however markedly different (Fig. 5). Whereas in the case of the in situ generated nanosilica particles, the out-of-phase displacement drops rapidly after passing through the peak, in the case of the Aerosil OX50, there is a much slower drop, implying that gelation is followed by a slow transition to a vitried form. The functionality of in situ grown nanoparticle surface can generate a rigid matrix by further reaction with the resin; in the case of the Aerosil OX50 particles, no such reactions appear to be possible and the matrix slowly moves to a completely cured form. It was however noted that increasing the Aerosil OX50 content signicantly decreases the time to gelation, Table 2.

3.9 Combination of nanosilica and modified nanosilica containing resin To achieve the maximum silica loading, it would be desirable to combine the nanomodied resin with the nanosilica particles. A series of mixtures of A610 and OX50 were investigated (Table 3). Increasing the temperature from 100  C to 130  C leads to the decrease in the cure time from 170 to 8 min. Increasing the volume of Aerosil OX50 in the blend, while reducing the volume of Nanopox A610
Proc. IMechE Vol. 226 Part L: J. Materials: Design and Applications

Downloaded from pil.sagepub.com by guest on November 22, 2013

Out of Phase Displacement

1.5x103

82

I Rhoney and R A Pethrick

2.0x103 100 Out of Phase Displacement


Gelation time (min) 86 24 17

1.5x103 In Phase Displacement

1.0x103 50 5.0x102 0 0 0 1000 2000 Time (secs) 3000 4000

Fig. 5

Cure curves for an EEC plus 40 wt% Aerosil OX50 with 0.1 phr Yb and cured at 100  C

Table 2
Concentration EEC (wt%) 71.38 59.95 54.52

Variation of the gelation time with increasing Aerosil OX50 content


Concentration OX50 (wt%) 28.55 39.98 45.43 Total silica (wt%) 28.55 39.98 45.43 Yb triflate (phr) 0.10 0.10 0.10

leads to a decrease in the gelation time. While the Nanopox A610 would be expected to promote the transition from gelation to vitrication, the Aerosil OX50 raises the base viscosity, which will enhance the exothermic effects of the reaction leading to a shortening of the gelation time. A surprising result was the observation of a reduction in the gelation time with a decrease in the Yb triate level. The cationic cure of epoxy resins involves a chain-like process and as such lowering the catalyst level will decrease the initiation level, but by analogy with free radical propagation process can lead to an increase in the chain length of the species created, which will raise the rate at which the viscosity increases. The increase in viscosity will increase the exotherm and hence shorten the gelation time. These observations emphasize the different ways in which the nanosilica particles interact with resin and are involved in the total cure process. The addition of the Yb triate in ethanol required the mixture to be degassed; however, when the Yb catalyst is added using methyoxymethanol, the viscosity was lowered and did not require degassing and the gelation time changes was approximately halved by changing to methoxymethanol, which is a better solvent for the catalyst.
Proc. IMechE Vol. 226 Part L: J. Materials: Design and Applications

3.10 Influence of the resin matrix The thermal expansion of the resin makes a signicant contribution to the total CTE. A series of different resin systems were investigated, which included o-cresol novolac epoxy, bisphenol A, and N,N diglycidyl-4-glycidylaniline; the latter being used to inuence the cross-link density of the matrix. As with EEC, the addition of Aerosil OX50 leads to a reduction of the cure time (Table 4). The o-cresol novolac has a functionality of 2.7 and in principle is capable of forming a highly cross-linked matrix. While this system showed normal gelation and vitrication behaviour, it was found that the degree of cure was low and that post-cure was required in most cases. This observation is consistent with the idea that the highly branched epoxy resin can form a gel relatively easily but that completion of the reaction may be a relatively slow process and require additional energy. The Aerosil OX50 produced the expected acceleration of the gelation as did the use of Nanopox A610. The latter being a low molecular weight bifunctional epoxy acts a little like a reactive diluent, lowers the viscosity and aids the reaction and achieves a higher degree of cure. The Nanopox A410 which has a bisphenol A base resin was blended with the novolac

Downloaded from pil.sagepub.com by guest on November 22, 2013

Low coefficient of thermal expansion of thermoset composite materials

83

Table 3

Variation of gelation time with composition and temperature for Nanopox A610 OX50 blends and variation of the catalyst level at different temperatures
Concentration OX50 (wt%) 11.11 14.28 10.91 11.11 14.28 10.91 Total silica (wt%) 55.53 57.12 55.53 55.53 57.12 57.12 Yb triflate (phr) 0.10 0.10 0.080 0.080 0.080 0.080 Cure temperature ( C) 100 100 100 130 130 160 Gelation time (min) 170 154 113 8 32 5

Concentration A610 (wt%) 44.42 42.84 44.42 44.42 42.84 42.84

Table 4
Additive

Comparison of gelation times with the addition of OX50 with variation of resin type
Silica content (wt%) 23.06 Cure temperature ( C) 100 120 140 140 120 140 Yb (phr) 0.1 0.1 0.1 0.1 0.1 0.1 0.3 0.1 0.2 0.4 0.5 0.1 0.3 0.3 Cure time (min) 684 218 130 44 130 37 30 1125 180 40 38 744 57 42

o-Cresol novolac resin

Aerosil OX50 Nanopox A610

o-Cresol novolac resin Nanopox A410 50/50 wt% mixture Aerosil OX50 46.59 120 Aerosil OX50 19.92 100 Aerosil OX50 19.92 120 Aerosil OX50 19.92 120 Aerosil OX50 19.92 120 Aerosil OX50 35.98 120 Aerosil OX50 40.67 120 Aerosil OX50 42.70 120

as a 50/50 wt% mixture to reduce the viscosity and reduce the requirement for post-cure. The data obtained are presented in Table 4. When the Aerosil OX50 is added to a 50/50 mixture of the novolac and Nanopox A410 bisphenol A resin, the cure is inhibited and the gel time increases dramatically. However, increase in the level of catalysts from 0.12 to 0.3 phr reduces the cure time to values, which are comparable to those found in the blend without the Aerosil OX50. This inhibition effect is particularly marked since addition of the Aerosil OX50 signicantly reduces the total amount of resin present in the mixture. 4 SUMMARY OF CURE DATA

The above study indicates how changes in the initiator level can be used to change the gelation and vitrication times in a controlled manner. The other feature emerging from the study is the impact which the surface modication incorporated in the in situ growth of the silica particles in the Nanopox resins has on the cure process. The particles are able to catalyse the reaction, lower the activation energy, and shorten the gelation time. These particles also signicantly shorten the time between gelation and

vitrication and lead to asymmetric out-of-phase displacement curves. The functionalized nanosilica particles are sufciently well dispersed to add an insignicant amount to the viscosity of the base resin. In contrast, the similarly sized nanosilica particles interact sufciently in a strong manner during the dispersions leading to a very signicant increase in the viscosity. The enhancement in the viscosity arises from the formation of chains of particles at a nanolevel. The increased viscosity slows the propagation reaction and increases the time between gelation and vitrication. The use of a small amount of methyoxymethanol as a dispersion solvent for the catalyst aids the reaction but also helps dispersion of the nanosilica and lowers the viscosity. The incorporation of platelet-type llers effectively inhibits the cure of this system. A mixture of 1005 Nanopox A610 with Aerosil OX50 which had a total silica content of 44.44 wt% was observed to undergo cure without the addition of Yb triate, which implies that there is the possibility of surface-activated catalyst of the epoxy auto-catalytic process. The initiation of the auto-catalytic process suggests that the surface contains active hydroxyl groups. The reduced activation for cure is consistent with this observation.
Proc. IMechE Vol. 226 Part L: J. Materials: Design and Applications

Downloaded from pil.sagepub.com by guest on November 22, 2013

84

I Rhoney and R A Pethrick

Table 5
EEC (wt%) 49.9 100 0 0 A610 (wt%) 29.9 0 59.9 0

Coefficients of thermal expansion for EEC and Nanopox A610


Silica (wt%) 19.97 0 39.9 100 CTE below Tg 54 86 2 48.5 1.5 0.55 CTE above Tg 169 228 20 136 11 CTE calculated 70 86 53 0.55 Tg ( C) 169 162 200

Nanosilica (wt%) 19.97 0 39.9 100

Table 6
EEC wt% 0 0 0 71.38 55.56 54.52 59.96 0 0 0 0 0 0 0 0 0 0

Variation of the coefficients of expansion (ppm/ C) for EEC Nanopox A610 Aerosil OX50
Nanopox A610 (wt%) 46.13 44.42 42.84 0 0 0 0 46.13 42.83 44.42 37.49 49.98 42.84 42.84 42.84 42.84 42.84 Aerosil OX50 (wt%) 23.07 29.6 28.56 28.55 44.44 45.43 39.98 23.07 28.55 25.91 37.49 16.66 28.56 28.56 28.56 28.56 28.56 Nanosilica (wt%) 30.76 25.9 28.56 0 0 0 0 30.76 28.55 29.62 24.99 33.3 28.56 28.56 28.56 28.56 28.56 Silica (wt%) 53.82 55.53 57.12 28.55 44.44 45.43 39.98 53.82 57.1 55.53 62.48 49.98 57.12 57.12 57.12 57.12 57.12 CTE below Tg 39 31 31 58 46 46 51 35 38 36 31 55 35 38 37 35 29 CTE above Tg 98 111 100 134 96 119 119 109 111 104 99 129 109 130 111 110 117 CTE calculated 40.9 39.4 38 63 49 48 53 41 38 39 33 44 38 38 38 38 38 Tg ( C) 217 218 223

189 151 144

DESIGN OF COMPOSITES WITH LOW CTEs

On the basis of equation (1), it is desirable to achieve the maximum level of silica in the material and also to use a resin matrix which itself has a low value of CTE. The formulations summarized above were subjected to DMTA and thermomechanical analyses to determine the glass transition temperature Tg and the CTE, respectively. The values of Tg were obtained from the location of the peak in tan  which is typically 10 above the Tg, as dened by the onset of modulus reduction obtained from the E data. The CTEs were determined as the average values below Tg and above Tg or where the plots were curved; a value at a specic temperature is presented. 5.1 EEC Nanopox A610 The Nanopox A610 contains 40 wt% 20 nm SiO2 particles, which have been produced in situ using a sol gel process. Blends of the EEC and Nanopox A610 were examined and the data obtained are presented in Table 5. For the 100 wt% resin, the values are obtained from an average of three separate samples, whereas the 100 wt% Nanopox A610 were obtained using seven samples.
Proc. IMechE Vol. 226 Part L: J. Materials: Design and Applications

The observed CTE below the Tg for the Nanopox A610 is lower than that predicted on the basis of equation (1). The Nanopox A610 consists of solgel particles which are surface functionalized and hence it would be expected that during the cure reaction, the silica particles will be able to act as cross-links in the matrix. The incorporation of the functionalized silica increases Tg from a value of 162  C for the base EEC resin to a value of 200  C for the combined resin with a total loading of 40 wt% silica. Aerosil OX50 contains 40 nm hydrophilic silica particles, has a low surface area of 50 m2/g, and does not readily form agglomerates. The Aerosil OX50 was incorporated into both the Nanopox A610 and the EEC in an attempt to increase the total silica loading, and the data are presented in Table 6. Number of formulations were repeated and the cure conditions varied slightly. At the 57 wt% total silica loading, the CTE below Tg is 31 ppm/  C, which is signicantly lower than the calculated values. Increasing the total silica content has decreased the CTE to a value which is 0.35 of that of the resin. The Aerosil OX50 is produced by an aqueous solgel process and will have OH surface functionality, but this will be different from the functionalities on the particles in the Nanopox A610. However, the Aerosil OX50 appears to be achieving reductions in the CTE of the resin which are

Downloaded from pil.sagepub.com by guest on November 22, 2013

Low coefficient of thermal expansion of thermoset composite materials

85

Table 7
EEC (wt%) 83.19 0 EEC (wt%) 79.94 41.61 0

Variation of the coefficients of expansion (ppm/ C) for EEC Nanopox A610 Aerosil 200 and Aerosil 90
Nanopox A610 (wt%) 0 54.5 A610 (wt%) 0 24.97 52.14 Aerosil 200 (wt%) 16.64 9.08 Aerosil 90 (wt%) 19.98 16.64 13.03 Nanosilica (wt%) 0 36.33 Nanosilica (wt%) 0 16.64 34.76 Silica (wt%) 16.64 45.41 Silica (wt%) 19.98 33.29 49.79 CTE below Tg 50 48 CTE below Tg 64 65 63 CTE above Tg 171 146 CTE above Tg 176 147 146 CTE calculated 73.45 48.29 CTE calculated 70.5 58.9 46.2 Tg ( C) 164 199 Tg ( C) 162 165 173

comparable to the in situ dispersed Nanopox A610 resin. A number of formulations were prepared in which the EEC and the Nanopox A610 were combined in various proportions. A series of samples with a total silica content of 57.12 wt% were prepared and slightly different processing conditions used. While there are variations in the values observed, an averaged value of 35 ppm/  C is lower than the theoretical prediction; however, the values of Tg were low. The sample with the highest silica content, 62.48 wt%, showed the lowest value of CTE at 31 ppm/  C, but surprisingly Tg was not as high as that observed with lower levels of silica. It is possible that at the highest loadings, the ability for the matrix to be completely cross-linked will become limited and this will be reected in lower values of Tg. Two alternative hydrophilic Aerosil samples were studied; Aerosil 200, which has a particle size of 12 nm but a surface area of 200 m2/g, is a grade used for controlling the rheology of uids and Aerosil 90 which has a particle size of 20 nm and a surface area of 90 m2/g. The Aerosil 200 has a tight particle distribution size similar to that of Aerosil OX50 but with a slightly larger mean diameter, whereas Aerosil 90 has both a large mean diameter and a broader particle size distribution. The data obtained are presented in Table 7. The Aerosil 200 with the Nanopox A610 gives a value which is close to the theoretically predicted value, while with pure EEC, it gives a signicant lowering of the CTE. The high surface area and the hydrophilic functionality of the surface may allow favourable packing to be achieved. While Aerosil 200 showed a signicant enhancement of the Tg at 45 wt% loading, the CTE was not signicantly lower than the theoretically predicted value. Using the large size and broad particle distribution, Aerosil 90 gives little enhancement of Tg and relatively high values of the CTE. This implies that the particles are not forming a signicantly improved network or packing together effectively to enhance the composite structure. If the cross-link density in the resin matrix is a controlling factor, then the use of a highly functional

epoxy, N,N diglycidyl-4-glydylaniline, has the potential of increasing the cross-link density. An addition of 0.36 wt% of this tri-functional epoxy to the resin formed from Nanopox A610 and Aerosil OX50 produced CTE values 33 ppm/  C below Tg and 99 ppm/  C above Tg and a Tg of 208  C at a total silica loading of 56.98 wt%. Rather than increasing Tg, the value is slightly lowered and however, the CTE above Tg is slightly reduced. 5.2 o-Cresol novolac epoxy resin The use of the trifunctional epoxy clearly does not achieve a signicant reduction in the CTE of the matrix. o-Cresol novolac epoxy resin is typically a 2.7 functional material and hence has the potential of producing a highly cross-linked matrix. Values of CTE and Tg are presented in Table 8. The value of the Tg for the o-cresol epoxy novolac varied signicantly with cure conditions. Post curing the pure resin at 130  C causes an increase from an initial low value of 86  C to a value of 146  C and on further increasing the post-cure temperature, a value of 185  C was achieved. Increasing Tg causes a corresponding decrease in the CTE below and above Tg, except that the nal post-cure appears to have increased the CTE. It has previously been observed that heating epoxy resins above 150  C can lead to ether bond scission, and in this case, a reduction in the degree of cross-linking with a commensurate increase in the CTE [14, 15]. A sample with Aerosil OX50 having a total silica content of 23.06 wt% had a Tg of 184  C, and CTE values 53 and 152 ppm/  C, respectively, below and above Tg. This material was very viscous and it was difcult to achieve higher silica loadings. To reduce the viscosity and increase the silica content, Nanopox A510 was blended as a 10 wt% component with the novolac and this increased the silica level to 33.31 wt%. The Tg of this blend was 174  C but there was a slight reduction in the CTE values 45 and 130 ppm/  C, respectively, below and above Tg. Addition of Aerosil OX50
Proc. IMechE Vol. 226 Part L: J. Materials: Design and Applications

Downloaded from pil.sagepub.com by guest on November 22, 2013

86

I Rhoney and R A Pethrick

Table 8
o-Cresol epoxy novolac 99.9 99.9 99.9 76.86 49.96 74.93 74.93 24.98 41.64 49.96 45.42 45.44

Variation of the coefficients of expansion (ppm/  C) for o-cresol epoxy novolac Nanopox A610 Aerosil OX50
Nanopox A610 (wt%) 0 0 0 0 29.98 14.99 14.99 44.97 24.98 29.98 27.25 27.26 Aerosil OX50 (wt%) 0 0 0 23.06 0 0 0 0 16.66 0 9.08 9.08 Nanosilica (wt%) 0 0 0 0 19.98 10 10 29.98 16.66 19.98 18.17 18.18 Silica (wt%) 0 0 0 23.06 19.98 10 10 29.98 33.31 19.98 27.25 27.25 CTE below Tg 103 69 65 53 55 56 58 54 45 55 51 51 CTE above Tg 246 163 180 152 143 150 153 138 130 155 149 137 CTE calculated 65 65 65 50 59 62 62 56 49 59 54 54

Tg ( C) 86 146 185 184 184 165 180 184 174 184 186 196

Fig. 6

Variation of the CTE vs wt% silica. Uncertainty in the values of CTE 2 ppm/ C which is typically the size of the symbol

increases the silica content to 50.72 wt%, leading to a reduction of the Tg to 117  C but did achieve a reduction of the CTE values to 37 and 121 ppm/  C, respectively, below and above Tg. Addition of Nanopox A610 to the novolac produced a Tg value of 184  C at 29.98 wt% loading and CTE values of 54 and 138 ppm/  C below and above Tg. A 50/50 wt% blend of novolac and Nanopox A510, which is the bisphenol matrix with added Aerosil OX50, gave CTE values CTE 50 and 160 ppm/  C below and above Tg at 20 wt% silica loading and 34 and 123 ppm/  C below and above Tg at 46.5 wt% silica. The Tg of the latter material was 175  C. Surprisingly, the novolac resin did not
Proc. IMechE Vol. 226 Part L: J. Materials: Design and Applications

produce signicantly lower values of CTE than those obtained using the EEC cure system. 6 CONCLUSIONS The study of a range of different polymer systems containing a combination of in situ dispersed nanoparticles and added nanollers indicates that equation (1) is broadly followed (Fig. 6). While the values for Aerosil 90, which is the broad distribution material, are above the theoretical line, the narrow distributions of nanollers Aerosil OX50 and Aerosil 200 produce values which are below the theoretical line.

Downloaded from pil.sagepub.com by guest on November 22, 2013

Low coefficient of thermal expansion of thermoset composite materials

87

The best results are obtained when the low surface area Aerosil OX50 is combined with the nanolled resin, Nanopox A610. This blend of nanoparticles and reactive functionality produces the highest values of Tg and the lowest values of CTE. This material is relatively easy to mix and cure and has the potential of being used as a composite material for applications that require low CTE. The o-cresol epoxy novolac, although in principle able to achieve a higher cross-link density and by implication a lower CTE above Tg, showed that improvements were only obtained by careful selection of the ller size and also post-cure temperature. FUNDING Support form Samsung is gratefully acknowledged for part of this research. Authors 2011 REFERENCES

7 8

10

11 1 Wong, C. P. and Bollampally, R. S. Thermal conductivity, elastic modulus, and coefficient of thermal expansion of polymer composites filled with ceramic particles for electronic packaging. J. Appl. Polym. Sci., 1999, 74, 33963403. 2 Schapery, R. A. Thermal expansion coefficients of composite materials based on energy principles. J. Compos. Mater., 1968, 2, 380404. 3 Varshneya, A. K. Fundamentals of inorganic glasses, 2006, p. 682 (Society of Glass Technology, Sheffield, UK). 4 Brandrup, J., Immergut, E. H., Grulke, E. A., Abe, A., and Bloch, D. R. (Eds) Polymer handbook, fourth edition, 2006 (John Wiley & Sons, New York). 5 Sideridis, E., Kytopoulos, V. N., Papadopoulos, G. A., and Bourkas, G. D. The effect of low-filler volume fraction on the elastic modulus and thermal

12 13

14

15

expansion coefficient of particulate composites simulated by a multiphase model. J. Appl. Polym. Sci., 2009, 111(1), 203216. Pethrick, R. A., Miller, C., and Rhoney, I. Influence of nanosilica particles on the cure and physical properties of an epoxy thermoset resin. Polym. Int., 2010, 59, 236241. Pethrick, R. A., Ishaque, S., and Rhoney, I. Novel epoxy nanocomposite materials. Plast. Rubber Compos., 2010, 39(3/4/5), 165170. Ingram, S. E., Liggat, J. J., and Pethrick, R. A. Properties of epoxy nanoclay system based on diaminodiphenyl sulfone and diglycidyl ether of bisphenol F: influence of post cure and structure of amine and epoxy. Polym. Int., 2007, 56, 10291034. Ingram, S. E., Pethrick, R. A., and Liggat, J. J. Effects of organically modified clay loading on rate and extent of cure in an epoxy nanocomposite system. Polym. Int., 2008, 57, 12061214. Ingram, S. E., Rhoney, I., Liggat, J. J., Hudson, N. E., and Pethrick, R. A. Some factors influencing exfoliation and physical property enhancement in nanoclay epoxy resins based on diglycidyl ethers of bisphenol A and F. J. Appl. Polym. Sci., 2007, 106, 519. n, A., and Serra, Arasa, M., Pethrick, R. A., Manteco A. New thermosetting nanocomposites repaired from diglycidyl ether of bisphenol and g-valerolactone initiated by rare earth triflate initiators. Eur. Polym. J., 2010, 46, 513. Pethrick, R. A. Rheological measurements, 1998 (Chapman and Hall, London). Affrossman, S., Collins, A., Hayward, D., Trottier, E., and Pethrick, R. A. A versatile method of characterizing cure in filled reactive polymer systems. J. Oil Colour Chem. Assoc., 1989, 72, 452454. Maxwell, I. D., Pethrick, R. A., and Datta, P. K. Thermal modification of amine cured epoxyresins - thermal and mechanical studies. Br. Polym. J., 1981, 13, 103106. Maxwell, I. D. and Pethrick, R. A. Low-temperature rearrangement of amine cured epoxy-resins. Polym. Degrad. Stab., 1983, 5(4), 275301.

Proc. IMechE Vol. 226 Part L: J. Materials: Design and Applications


Downloaded from pil.sagepub.com by guest on November 22, 2013

Anda mungkin juga menyukai