Anda di halaman 1dari 22

Applied Composite Materials 11: 7798, 2004. 2004 Kluwer Academic Publishers. Printed in the Netherlands.

77

Application of Interlaminar Tests to Marine Composites. Relation between Glass Fibre/Polymer Interfaces and Interlaminar Properties of Marine Composites
CHRISTOPHE BALEY1, YVES GROHENS1, FRDRIC BUSNEL1 and PETER DAVIES2
1 Universit de Bretagne Sud, L2P, BP 92116, 56321, Lorient Cedex, France.

e-mail: {christophe.baley, yves.grohens, frederic.busnel}@univ-ubs.fr 2 IFREMER, Materials & Structures group (TMSI/RED/MS), BP70, 29280 Plouzan, France. e-mail: peter.davies@ifremer.fr (Received 14 October 2003; accepted 27 October 2003) Abstract. The need for improved performance and the development of new composite manufacturing methods require a better understanding of the role of interface phenomena in the mechanical behaviour of these materials. The inuence of the cure cycle on the bulk and surface properties of the matrix resin, and of composites based on polyester and epoxy resins reinforced with glass bres has been studied. While the mechanical properties of the epoxy vary with cure temperature the surface tension is not affected. The increase in interfacial shear strength and interlaminar shear strength with increased cure temperature cannot be simply explained by the wetting of the bres by the matrix. The importance of thermal stresses, generated at the interface by resin shrinkage and differences in thermal expansion, for the mechanical behaviour of the composite are demonstrated. Key words: glass bres, polyester, epoxy, surface energy, reversible adhesion energy, wettability, microbond test, interfacial shear strength, interlaminar shear strength.

1. Introduction The mechanical performance and durability of composite materials are mainly governed by three factors [1]: The strength and stiffness of the bres; The strength and chemical stability of the matrix; The efciency of the bonds and/or interactions between bres and matrix. The interface region ensures load transfer between bres and matrix. It is therefore not surprising that through-thickness properties in tension and shear are very sensitive to the quality of the bre/matrix interface.

78

CHRISTOPHE BALEY ET AL.

Composite boat-builders generally only look at macroscopic mechanical properties when they are qualifying or checking laminates. Interlaminar shear (short beam shear) tests are often performed as these indicate a potential weakness of the material. The stress state in such specimens is complex and the strength measured is strongly dependent on several parameters such as the matrix properties and the manufacturing conditions. The cure cycle applied can result in internal stresses and porosity and thus lead to increased risk of delamination [2]. In the present paper two materials will be studied, glass/polyester and glass/ epoxy, as these materials are frequently used in boat-building. Polyester resins are widely used in naval construction for small pleasure boats, shing and service vessels and military ships, while epoxies are generally limited to high performance craft such as racing yachts. The choice of laminating resin depends on many parameters including cost, manufacturing requirements (gel time, cure temperature, post-cure possibilities), compatibility with the reinforcement, mechanical performance and resistance to the marine environment. In the present paper the inuence of the cure cycle will be examined for these materials at different scale levels in order to study: The wetting of bres by the matrix in the liquid state; The shear strength of the bre/matrix interface; The interlaminar shear strength. Studies of the relationship between microscopic and macroscopic properties have been performed in the past on high performance composites, many references will be discussed below, but marine composites have received very little attention in this area. First, these relationships between microscopic and macroscopic properties will be described, then the observations made during the study and their validity for this type of material will be discussed.

2. Background 2.1. WETTING , ADHESION ENERGY AND INTERFACIAL SHEAR STRENGTH The measurement of the thermodynamic surface characteristics is important because these can be related to adherence (the quality of the interfacial bond) between bre and matrix. A proper wetting of the bres by the matrix is a necessary but not sufcient condition to obtain a good composite assembly [3]. Improvement of wetting enables the reversible adhesion energy to be increased and the number of defects in the interfacial zone to be reduced. Many workers have examined the relationship between the interfacial shear strength, obtained from micromechanics tests, and the reversible adhesion energy. The reversible adhesion energy may be determined either by wetting studies [47], or by inverse gas chromatography [813].

APPLICATION OF INTERLAMINAR TESTS TO MARINE COMPOSITES

79

The work of adhesion WA , or reversible adhesion energy, is a concept initially proposed by Harkins [14]. In a simple system where a liquid (L) adheres to a solid (S), the reversible work of adhesion is dened by: WA = S + L SL , (1)

where S , L and SL are, respectively, the surface tensions of the solid in air, of the liquid in air and of the solid/liquid interface. Another expression (YoungDupr) used to dene the work of adhesion is: WA = L (1 + cos ), (2)

where is the angle of wetting of the bres by the matrix. A commonly used approach to treating solid surface energies is that of expressing any surface tension (usually against air) as a sum of components due to dispersion forces ( D ) and polar (e.g., hydrogen bonding) forces ND [15, 16]: = D + ND. Similarly, WA can be written:
D ND WA = WA + WA .

(3)

(4)

Owens and Wendt [16] have suggested applying the geometric mean approach:
D D ND ND WA = 2(L S )1/2 + 2(L S )1/2 . g

(5)

The relation between the reversible adhesion energy and the shear strength of the bre/matrix interface has been studied by several authors [10, 11, 17, 18] who show that an increase in work of adhesion results in an increase in shear strength. Nardin and Schultz [10] proposed an adhesion pressure concept for composite materials. After studying a large number of polymer/bre couples they suggested that the reversible adhesion energy, measured by inverse gas chromatography, could be directly related to the interfacial shear strength measured by fragmentation tests by the following linear expression: = with Em Youngs modulus of the matrix, Ef Youngs modulus of the bres, WA reversible adhesion energy. is a distance independent of the system studied which is equal to about 0.5 nm. It corresponds to the intermolecular distance at equilibrium, centre to centre, of the molecules involved in interactions such as Van der Waals forces. Em Ef
1/2

WA

(6)

80

CHRISTOPHE BALEY ET AL.

Using this approach the load transfer between bre and matrix is considered to be perfectly linear and elastic. Pisanova and Mder [11] studied the bonds between different matrix resins and glass bres with and without treatments. The treatment of bres with coupling agents resulted in an improvement of the bre/matrix adherence which correlated well with the reversible adhesion energy. 2.2. RELATIONSHIP BETWEEN MICROMECHANICAL AND
MACROMECHANICAL TESTS

Several studies have described the sensitivity of different tests for the characterisation of interface properties by comparing results from micromechanics tests to macroscopic test values. For example, Park and Kim [19] showed that, for glass bres and a polyester resin, bre treatment with a coupling agent leads to an increase in the surface energy of the bres, and in the interlaminar shear strength and the mode II fracture toughness of the composite. For glass and carbon reinforced composites with increasing bre surface treatment levels, Mder [20] studied interfacial strength by pull-out tests and composite properties using transverse tension and interlaminar shear (compression on notched specimens and short beam shear tests). Herrera-Franco et al. [21] compared, for an epoxy resin and carbon bres with different surface treatments, micromechanics properties by fragmentation, microdroplet debonding and microindentation, and macroscopic properties of composites using tension on 45 specimens, Iosipescu shear, and interlaminar shear (short beam shear test). Keusch et al. [22, 23] studied the inuence of glass bre surface treatments on interfacial characteristics measured by pull out and the composite behaviour (with an epoxy matrix) measured by interlaminar shear (short beam shear test and notched compression) and transverse tension. Their results show a strong inuence of the glass bre surface treatment on properties and that the best treatment (an amino-silane coupling agent) produced superior properties in all tests. In all these studies when an improvement in shear strength of the bre/matrix bond was noted there was also an improvement in the macroscopic properties of the composite, although only rarely was the improvement directly proportional to the microscopic properties. In this type of study the composite properties depend on the quality of the interfacial bond but also on the matrix properties. 2.3. INFLUENCE OF THE CURE CYCLE ( ON MATRIX PROPERTIES ) In naval construction, polyester resins are the most common matrix materials but vinylesters, epoxies and phenolics are also used. These resins have different mechanical properties and several authors [24, 25] have noted the strong dependence of the delamination resistance of boat hull panels on the matrix properties. Manu-

APPLICATION OF INTERLAMINAR TESTS TO MARINE COMPOSITES

81

facturing conditions, cure kinetics and cure cycles all play an important role in the cross-linking of the matrix. The stability of the matrix may be further enhanced by a post-cure cycle to complete cross-linking [26]. An increase in cross-link density increases the mechanical stability and allows relaxation of the molecular network which can relax out internal stresses. This results in an increase in matrix failure strain and fracture toughness. Tucker [26] showed that a post-cure cycle applied to a glass/vinylester composite affects the mode I fracture energy, a result which is mainly caused by the improved toughness of the matrix. For an epoxy resin Albersen et al. [27] showed an inuence of cure temperature on the mechanical properties of the matrix. As cure temperature increased the Youngs modulus decreased while the failure strain increased. This unexpected result was examined further by NMR [28, 29]. The increase in cure temperature increases the cross-link density which increases mechanical losses by energy dissipation. The spatial scale of the molecular movements increases with increasing crosslink density, going from a local scale (restricted movements) to larger amplitude, co-operative rearrangements. The latter appear to increase the dissipation phenomena and reduce the modulus. The cooling rate also has an inuence on the composite properties due to the difference between thermal expansion coefcients of bre and matrix. Too rapid cooling can result in harmful residual thermal stresses. In the boat-building industry many structures such as hulls and decks are very large so post-cure is very difcult. Resins are formulated for the particular application and can be used with different cure cycles within limits dened by the suppliers. 2.4. MANUFACTURING TECHNOLOGY The main technologies under development for the boat-building industry are RTM (Resin Transfer Moulding) and infusion. In both cases a thermoset resin migrates through reinforcing bres placed in the mould. The presence of defects in structures alters their mechanical properties. The origins of these defects depends on either the constituents and cure cycle or on the impregnation method. In RTM, for example, the presence of voids and poorly impregnated zones is related to the ow of resin into a porous medium. The main difculty lies in controlling the rate of displacement of the resin front in the bres in the mould. The polymer ow must be studied on a microscopic scale as two types of ow exist: a viscous ow in the resin-rich zones and a capillary ow in the bre rich zones [3032]. If the ow rate in the bre rich zones is faster than that in the resin rich zones then macro-voids develop. If the ow is faster in the resin rich zones then micro-voids develop. Resin ow in a medium can be described by Darcys law and capillary pressure is estimated by [3336]: Pc = L cos , m (7)

82

CHRISTOPHE BALEY ET AL.

where L is the matrix surface tension, is the wetting angle of the bres by the matrix, m is a parameter (the hydraulic radius of the bre bed) which is the ratio of the capillary normal section to the wetted perimeter. m is given by: m=F with VF bre volume percentage, dF bre diameter, F a geometrical factor equal to 1 for unidirectional bres. There is thus a relation between matrix surface tension, bre/matrix wetting angle (wettability) and the development of voids during injection. In addition there is a relation between surface tension, wetting angle and adhesion energy, interfacial shear strength and composite interlaminar shear strength. The current paper will examine this last point. 3. Experimental 3.1. MATERIALS The bres used in the present study are E-glass with a textile/plastic nish compatible with polyester and epoxy resins. The reinforcement is a taffetas weave of 290 g/m2 surface weight supplied by Chomarat S.A., commonly used in naval construction. For bre debonding tests bres were taken from yarns in the weave. The resins are an isophthalic polyester (Norsodyne S 70361 TA with 1.5 wt% MEKP catalyst), and a DGEBA epoxy (Axson Epolam 2015, with 32% by weight aliphatic amine hardener). The cure temperatures applied are 65, 85, 105 and 120 C for the epoxy and 65 C for the polyester. Cure time at temperature is 14 hours in all cases. The cooling rate is controlled and constant at 10 C/hour. No tests were performed with a room temperature cure as the interfacial shear strength by debonding a micro-droplet requires the matrix to have reached a certain level of cross-linking. 3.2. SURFACE ENERGIES The liquid matrix surface properties were measured using the suspended droplet test [37]. The surface tensions of bres and solid matrix were determined by measurement of contact angles using different liquids (ethylene glycol, water, glycerol, di-iodomethane, tricresylphosphate) [38]. For the bres, contact angles are determined using measurement in the microscope [3942]. dF (1 VF ) 4 VF (8)

APPLICATION OF INTERLAMINAR TESTS TO MARINE COMPOSITES

83

Figure 1. Schematic diagram of the microbond test.

3.3. MATRIX CHARACTERISATION Different matrix properties were determined including: Glass transition temperature DSC (modulated differential scanning calorimetry, TA Instruments 2920), heating rate 5 C/minute. This enabled the state of matrix crosslinking to be checked after post-cure. Elastic modulus and tangent delta by DMA in exure (dynamic mechanical analysis, TA Instruments DMA 2980). Youngs modulus, stress and elongation at failure using a tensile test machine and a 50 mm gauge length extensometer, at least 10 dogbone specimens for each condition. Torsion shear modulus on parallel strips (Torsiomat from Prodemat S.A.). 3.4. MICROBOND PULL - OUT TECHNIQUE There are several micromechanics tests available to study the bre/matrix bond. The best-known are fragmentation, compression, microindentation and debonding [43]. The latter was used here, it consists of the debonding of a resin droplet from a single bre (Figure 1) [44]. Before the test each specimen is studied under an optical microscope to determine the geometry (bre diameter, microdroplet diameter, length of bond), and to check that the droplet is symmetrical and that there are no defects. In order for the debond force to be proportional to the bonded area the bonded length must be short (less than 250 m) [45]. After placing the resin microdroplets on the bres the specimens are heated to obtain the required degree of crosslinking. They are then placed on the test machine (MTS Synergie 1000 load cell 2 N capacity) and knife edges mounted on

84

CHRISTOPHE BALEY ET AL.

Figure 2. Typical load/displacement recording obtained from a valid microbond test.

a micrometer controlled base are adjusted under the microscope. During the test the debonding is observed using a magnifying binocular system. The tensile loading rate is 0.1 mm/min. The load-displacement plot is recorded (Figure 2) and shows a sudden interfacial failure followed by a roughly constant load due to friction of the droplet on the bre. The parameters which inuence the results are: The elastic properties of the matrix [46]; The bre diameter [46]; The bonded length [45, 47]; Residual thermal stresses [1, 45, 4749]; Interphase characteristics [5054]; Loading conditions (for example, knife edge opening) [55, 56]; The droplet geometry and in particular, the bre/matrix contact angle (meniscus) formed between bre and matrix [50]. Hodzic et al. [54] showed by nite element analysis that the contact angle plays a dominant role in the failure mechanism. The wetting angle depends on the surface tensions of each phase and therefore on the bre surface treatments.

3.5. INTERLAMINAR SHEAR STRENGTH AND SHORT BEAM SHEAR TEST The short beam shear test uses a rectangular unidirectional or laminate specimen (standard test methods ASTM D2344-84, NF EN 2377, NF ISO 4585, L 17-142 . . .). The advantage of this test is its simplicity; it represents a loading mode often encountered in composite structures loaded in exure and it can be performed on specimens taken from real structures for quality control purposes [57]. It can also

APPLICATION OF INTERLAMINAR TESTS TO MARINE COMPOSITES

85

Table I. Liquid matrix surface tensions, hanging drop test. L surface energies (mJ/m2 ) Polyester resin Epoxy resin 30.7 1.4 36.9 2.2

Table II. Total surface energy and its dispersive and polar components. Glass bre and matrix. D (mJ/m2 ) Glass bre Polyester. Cure at 65 C Epoxy. Cure at 65 C Epoxy. Cure at 85 C Epoxy. Cure at 105 C Epoxy. Cure at 120 C 27.5 1.4 30.9 0.5 30.1 1.2 31.8 0.9 28.6 0.8 32.0 0.5 ND (mJ/m2 ) 21.9 2.1 13.2 0.5 10.7 1.1 9.4 0.9 10.7 0.8 10.2 0.4 S (mJ/m2 ) 49.4 3.5 44.2 1 40.8 2.3 41.2 1.8 39.3 1.6 42.2 0.9

be employed in fatigue studies [58, 59]. Several studies have shown that the test is very sensitive to the quality of the bre/matrix interface [21, 60]. Analysis of the test [57, 61] indicates that the mode of loading (exure in the planes 13 and 23) is a combination of tension-compression and shear. There are stress concentrations at the supports and particularly below the central loading point. The shear stress distribution is not parabolic, analysis is thus more complex than the conventional strength of materials analysis would suggest. The failure mode [61] is strongly dependent on the ratio L/ h (length between supports over thickness) and the standard test methods suggest different ratios according to the type of material tested. The shear stresses dominate over the normal stresses when this ratio is 5 or less. 4. Results and Discussion 4.1. MATRIX SURFACE TENSIONS The liquid surface tension (Table I) is different to that measured in the solid state (Table II) for the two resins, the value is higher for the solids. This is due to the increase in density when the resin goes from liquid to solid. The surface energy is directly proportional to the cohesion energy of the material which is always greater for solids. This can also explain the difference between the epoxy and polyester: the contact angle difference, for the resin on the bre, between liquid and solid is 4 for the epoxy and 14 for the polyester, for volume shrinkage of 1% and 8%, respectively. Thus a large increase in density, as a result of a large shrinkage

86

CHRISTOPHE BALEY ET AL.

Table III. Properties of matrix resins as a function of cure cycle, measured by DSC and DMA. MDSC Tg ( C) Polyester. Cure at 65 C Epoxy. Cure at 65 C Epoxy. Cure at 85 C Epoxy. Cure at 105 C Epoxy. Cure at 120 C 89 83 91.5 95.5 94 DMA E (MPa) 3514 42 3063 105 2699 113 2790 95 2703 62 DMA Max. Tg 0.048 0.003 0.028 0.004 0.034 0.003 0.029 0.002 0.029 0.002 DMA T ( C) for Tg = max 91.5 87 2 99 1 103 2 103 1

during solidication, causes a large change in the interfacial energy and hence of the measured angles. For the epoxy, different cure cycles have been studied. Given the scatter in the results it is not possible to show the inuence of the post-cure on the surface tension. However, it has been shown that the room temperature density of an epoxy decreases when the crosslink density increases [62]. If the present measurements do not show these density changes it is either because they are too small to be detected by the wettability method or that they are compensated by surface molecular reorganizations. The latter may be linked to interfacial segregation of certain components of the resin or to specic orientations of some chemical groups with respect to the surface. The superposition of these effects may mask the correlation of surface energy with density. However, the important point to note is that the cure cycles do not modify the reversible bre/matrix adhesion energy. 4.2. PROPERTIES OF THE MATRIX AS A FUNCTION OF THE CURE CYCLE Different properties of the epoxy evolve with post-cure temperature: The glass transition temperature (Tg ) measured by DSC (Table III) and the temperature corresponding to a peak in tangent () measured by DMA (Table IV) increase with increasing cure temperature. This agrees with the hypothesis of an increase in crosslink density as cure temperature is increased. The Tg is a reliable indicator for crosslink density, unlike the sub-Tg transitions whose temperatures do not vary. Youngs modulus measured by DMA (Table III) and by tensile tests (Table IV, Figure 3) and the torsion shear modulus (Table IV) decrease with the increase in cure temperature. This is related to the amplitude of molecular movements as mentioned previously. The elongation and stress at failure increase with cure temperature (Table IV). The mechanical properties at failure are related to molecular mobility and to the -transition. However, contrary to modulus there is no universal relationship. Modifying the tensile loading rate may invert the results, for example.

APPLICATION OF INTERLAMINAR TESTS TO MARINE COMPOSITES

87

Table IV. Mechanical characteristics of the matrix resins from tensile and torsion tests. Tensile Em (MPa) Polyester. Cure at 65 C Epoxy. Cure at 65 C Epoxy. Cure at 85 C Epoxy. Cure at 105 C Epoxy. Cure at 120 C 3797 110 3097 86 2797 43 2716 56 2835 37 Tensile A (%) 1.2 0.1 1.6 0.2 3.0 0.9 3.4 0.9 3.7 0.7 Tensile (MPa) 46 3 45 4 58 7 60 10 66 6 Torsion Gm (MPa) 1396 80 1193 53 1036 46 1015 57 1001 61

Figure 3. Tensile test results versus cure temperature.

4.3. STUDY OF FIBRE / MATRIX CONTACT ANGLES AND ASPECT RATIO OF


MICRODROPLETS WITH RESPECT TO CURE CYCLE

Before debonding the contact angles between the glass bre and the solid matrix together with the microdroplet geometry are measured. These angles depend on the state of the matrix (solid or liquid) (Table V). For the polyester, during the passage from liquid to solid the angle increases. This is again a result of the large volume shrinkage during curing (about 8%). For the epoxy on the other hand, the angle decreases with cure temperature. The volume shrinkage is about 1% for the epoxy during cure. After placing the resin on the bres the specimens are put in an oven; the evolution in surface tension with temperature before gel explains part of the change in wetting angles. Some workers leave the specimens to gel at room

88

CHRISTOPHE BALEY ET AL.

Table V. Contact angles between glass bre and resin, and aspect ratio of microdroplets (bonded length/diameter). contact angles with glass bre Liquid polyester Solid polyester. Cure at 65 C Liquid epoxy Solid epoxy. Cure at 65 C Solid epoxy. Cure at 85 C Solid epoxy. Cure at 105 C Solid epoxy. Cure at 120 C 16.3 2.3 25.7 2.1 19.4 2.0 15.5 2.5 14.0 1.1 13.3 2.5 13.3 1.4 Ratio embedded length on microdrop diameter 1.63 0.09 1.52 0.04 1.57 0.17 1.50 0.14 1.50 0.13 1.70 0.12 1.68 0.10

temperature before cure to avoid evaporation of certain hardeners, which could result in stoichiometry changes and hence changes in matrix properties [63]. This precaution was not taken following talks with the resin supplier and preliminary tests. However, a local variation in stoichiometry is possible on account of the differences in surface tension between the base epoxy and the hardener in the resin formulation. The preferential absorption of one of the components, dependent on the temperature, may also explain the change in contact angles measured here. This phenomenon could result in the creation of an interphase whose properties could be quite different from those of the matrix resin. The values of aspect ratio (Figure 1) indicate a good reproducibility of the microdroplet geometry (Table V). 4.4. DEBONDING TESTS . ANALYSIS OF FAILURE AND FRICTION OF
MICRODROPLETS ON THE FIBRE VERSUS MATERIAL AND THERMAL CYCLE

The stresses in the microdroplet test have been studied in detail by FE analysis [21, 50, 54, 55, 6467], by photo-elasticity [68], and by Raman spectroscopy [49]. These studies show that the loading is complex (the shear stresses are not constant along the bre/matrix interface) and that the residual thermal stresses are not negligible. In order to analyse the results from these tests several approaches have been used in previous work, to determine: The apparent interface shear stress [44] assuming a uniform stress distribution along the interface. This value is obtained from the mean shear stresses measured or by a linear regression of the plot of debond load versus bonded bre/matrix interface area. The shear stress is the slope of that plot (Figure 4). The maximum shear stress in the region where cracking starts, taking account of the non-uniform stress state and residual stresses [48, 65].

APPLICATION OF INTERLAMINAR TESTS TO MARINE COMPOSITES

89

Figure 4. Microbond test. Plot of load at debond versus embedded area.

The normal interface stress or adhesive pressure [69]; this approach was developed to correlate test results with work of adhesion from micromechanics tests. Critical interface fracture energy GIc [45, 47, 64, 70, 71]. In the present study the apparent shear stress and the ultimate shear stress have been determined. The latter (maximum value) is obtained from the expressions proposed by Zhandarov et al. [48]. (This model takes into account the non-uniformity of the shear stress distribution along the interface caused by both reaction of the elastic matrix to external load and residual thermal stresses.) app Le Le rf + Ef (m f ) T tanh tanh(Le ) 2 2 8Gm , 2 = 2 Ef df ln(2 Rm ) df ult = , (9) (10)

90

CHRISTOPHE BALEY ET AL.

Table VI. Micromechanics debonding tests, interfacial shear stresses, apparent and ultimate. i average of tests (MPa) Polyester. Curing at 65 C Epoxy. Curing at 65 C Epoxy. Curing at 85 C Epoxy. Curing at 120 C 15.1 4.5 29.4 3.6 39.0 6.3 47.4 9.3 i from slope of regression (MPa) 15.7 2.9 29.3 2.4 37.6 3.7 48.2 5.1 70.9 14.7 75.7 9.7 114.6 11.1 134.7 19.1 iultime (MPa)

where m and f are the coefcients of thermal expansion of the bre and matrix, respectively. T is the variation of temperature, df is the bre diameter, Ef is the bre Youngs modulus, Gm is the matrix shear modulus, and Rm is the matrix radius. is a quantity reciprocal to the ineffective length. In contrast to app , the ultimate adhesion strength ult characterises the intensity of physical interactions at the interface and does not depend on the specimen geometry or on residual stresses or mechanical properties of the components, i.e. it should characterise the fundamental adhesion. There is a good agreement between the ult experimental data obtained by different micromechanical techniques for a given polymer-bre pair [48]. Thus, it allows ult to be used as a parameter characterizing the strength of adhesion bonding and compare it with the thermodynamic work of adhesion for the same system. Table VI shows mean and ultimate shear stresses from the debonding tests, between 20 and 30 valid tests were made to obtain each value. The mean interface stresses calculated by two methods (mean of measured stresses and linear regression method) are very similar. The ultimate shear stresses are much higher but are similar to published values [48, 65]. The use of an epoxy resin leads to better adherence to these glass bres than polyester, this has been observed previously. For the epoxy, the interface strength increases with cure temperature. This may be explained by: Crosslink density/different network organisation; Residual thermal stresses; Chemical bond creation between bre and matrix; Increased matrix failure strain and reduced Youngs modulus. Examination of the debonded droplets (Figure 5) shows that while debonding is controlled by the propagation of an interfacial crack the initiation occurs in mode I loading in the matrix [47] and the matrix fracture toughness plays an important role [63]; The evolution of the bre/matrix wetting angle. A low angle enables stress concentrations to be limited.

APPLICATION OF INTERLAMINAR TESTS TO MARINE COMPOSITES

91

Figure 5. Example of micro-droplet after debonding.

In order to understand the inuence of residual stresses an approximate analysis of the microdroplet friction after debonding can be performed. After debonding the friction is governed by the friction coefcient and the interfacial pressure P [71]. For a glass bre and a thermoset matrix Piggott [71] estimated the value of the friction coefcient to be between 0.35 and 1.8. This coefcient represents the physical and mechanical interactions between bre and matrix. The interfacial pressure P originates from the matrix shrinkage during crosslinking and the residual stresses developed during cooling. The thermal expansion coefcient of the matrix is much greater than that of the glass bres (polyester m = 75 106 / C, epoxy m = 70 106 / C, glass bre f = 5 106 / C). For a transversely isotropic bre in an innite isotropic matrix, the interfacial pressure P due to a temperature variation T is given by [7274]: P = (1 + vlf )(m f ) T , 2 (1 vtf )/Etf + (1 + vm )/Em 2vlf /Etf (11)

where v is Poissons coefcient, the thermal expansion coefcient, E Youngs modulus, with f for bre, m for matrix, l for longitudinal and t for transverse. After debonding, the friction coefcient can be estimated from the measured load and the contact area. The simple linear expression linking friction stress, friction coefcient and the interfacial pressure P often used is [71, 7476]: Frict = P . (12)

92

CHRISTOPHE BALEY ET AL.

Table VII. Micromechanics debonding test. Study of friction after debonding. Reference Polyester/glass Epoxy/glass Epoxy/glass Epoxy/glass Cure temperature ( C) 65 65 85 120 Friction stress (MPa) 4.4 0.7 7.7 1.4 15.3 8.3 21.7 8.6 Radial stress at the interface (MPa) 10.6 8.8 10.6 15.65 Friction coefcient 0.41 0.07 0.88 0.16 1.44 0.78 1.39 0.55

Table VIII. Short beam shear test. Interlaminar shear strength of composites, reinforced by 290 g/m2 weave, Vf = 35%. Matrix Polyester 150 h at room temperature (20 C) Polyester. Cure at 65 C Epoxy 150 h at room temperature (20 C) Epoxy. Cure at 65 C Epoxy. Cure at 85 C Epoxy. Cure at 105 C Epoxy. Cure at 120 C Tg (MDSC) ( C) 90 65 77 80 95 86 ILSS (MPa) 18.6 0.7 23.8 1.5 25.3 0.9 30.7 1.3 32.2 1.5 32.6 0.9 36.5 1.8

The measured friction stresses, calculated interfacial pressures (Equation (12)) and estimated friction coefcients (Equation (13)) are presented in Table VII. For the epoxy, the friction stress increases with cure temperature; this conrms the existence of residual thermal stresses. The scatter is quite high but the mechanisms are complex and the specimens are very small. Figure 5 shows in particular that a thin polymer layer remains on the bre. The calculated interfacial pressure is only an estimation, the assumption of an innitely long specimen is of course not respected. 4.5. INTERLAMINAR SHEAR STRENGTH AS A FUNCTION OF THE MATRIX
TYPE AND CURE CYCLE

Interlaminar shear strengths were measured, at room temperature, using the short beam shear test, on composites of woven E-glass bres. The samples were made by press moulding in order to maintain a constant bre volume fraction of 35%. Different cure cycles were then applied (Table VIII). Before testing the specimen edges were polished and inspected under the optical microscope to check that porosity was acceptably low (<1.5%). After testing samples were taken from specimens for DSC analysis.

APPLICATION OF INTERLAMINAR TESTS TO MARINE COMPOSITES

93

Table IX. For a curing temperature to 65 C and for two matrix resins. Total adhesion work WA , interfacial shear strength and interlaminar shear strength. Glasspolyester WA according to Equation (2) WA according to Equation (5) IFSS Interfacial shear i (MPa) ILSS Interlaminar shear (MPa) 60.2 3.1 92.3 4.2 15.7 2.9 23.8 1.4 Glassepoxy 71.7 4.7 88.2 5.6 29.3 2.4 30.7 1.3

For the glass/polyester and glass/epoxy the cure enables crosslink density to be increased and leads to improved interlaminar shear strength (Table VIII). The shear strength values depend on matrix ductility (Table IV) and on the quality of bre/matrix bonding (Table VI). Both these characteristics are improved by the cure cycle. The strength of the glass/epoxy composites is superior to that of the glass/polyester. The Tg values measured on the epoxy composites increase with cure temperature but are generally a few degrees lower than those measured on the resins (Table III). There are many possible reasons for this difference. The presence of bres may result in heterogeneous cross-linking, local temperature variations or interphase development with modied properties. Further work is needed to clarify this point. For a woven glass reinforced polyester for naval construction, Smith [24] indicated an ILSS mean value of 23.5 MPa for a bre content of 34% by volume, very similar to the value obtained here (23.8 MPa) (Table VIII). The use of glass bres with other constructions (plain or satin weave, for example) can lead to higher values.

4.6. RELATIONSHIPS BETWEEN REVERSIBLE WORK OF ADHESION , DEBOND TESTS AND INTERLAMINAR SHEAR STRENGTH , MICRO - MACRO Table IX summarises the results from this study. The glass/epoxy shows higher values of interfacial and interlaminar shear strengths than the glass/polyester. For the reversible energy of adhesion however, the epoxy shows higher values for the liquid matrix (Equation 2) but lower values for the solid matrix (Equation 5). In both cases scatter is quite high. However, the work of adhesion between two solids is not a direct method but requires calibrated liquids, while in the liquid state there are difculties related to the oven so it would be necessary to know the surface tension and wetting angle at temperature to determine the real adhesion energy. The measurements of wetting angle as a function of cure temperature (Table V) show that there is a signicant effect.

94

CHRISTOPHE BALEY ET AL.

The expression proposed by Nardin [10] to relate reversible adhesion energy (determined by inverse gas chromatography), to the interfacial shear strength (which he measured by fragmentation) (Equation 9) is not directly applicable to the study of the inuence of the cure temperature of epoxy (Table IX). In fact increasing the cure temperature causes a reduction in Youngs modulus of the matrix and an increase in the interfacial shear stress. These opposing evolutions show how complex it is to analyse the response of the interface. It is possible to substitute the mean shear stress in expression (9) proposed by Nardin, by the ultimate value. For epoxies the calculated adhesion energy value is increased compared to the values measured by wetting or chromatography. Pisanova et al. [65] consider that this approach accounts for the contribution of acid-base interactions and chemical bonds to the work of adhesion. The reversible adhesion energy concept is insufcient to estimate the interfacial shear strength [65] because: Physical interactions at the interface are not accounted for (possibility of interdiffusion between resin and sizing not included); No account is taken of chemical bonding between bres and matrix; No account is taken of fabrication conditions.

5. Conclusion This paper describes a study whose aim was to examine the inuence of interface properties on the global properties of marine composites. Very little work of this type has been performed on such materials and the results show the considerable difculties associated with such an exercise. Nevertheless some conclusions may be made. First, it is apparent that the cure temperature has a critical inuence on the performance of a glass/epoxy composite. The modulus of the epoxy matrix decreases with crosslink density, as shown previously. Second, the limitations of models based on reversible adhesion energy, Wa , are shown and interface shear strength measurements were not possible on room temperature cured materials. On the other hand a good correlation was found between interfacial shear strength measured by microdroplet debonding and interlaminar shear strength of post-cured composites. The glass/epoxy material showed improved performance compared to the glass/polyester for all the properties examined here (reversible adhesion energy, interfacial shear strength, interlaminar shear strength). However for marine applications further work is needed to examine the inuence of the marine environment, as long term durability is an essential element of resin selection. The behaviour of the interface dominates the behaviour of composites under critical loading and will continue to be the focus of research studies.

APPLICATION OF INTERLAMINAR TESTS TO MARINE COMPOSITES

95

References
1. Biro, D. A., Mclean, P. and Deslandes, Y., Application of the Microbond Technique: Characterization of Carbon Fiber-epoxy Interfaces, Polymer Engineering and Science 37(17), 1991, 12501256. Baley, C., Davies, P., Grohens, Y. and Dolto, G., Application of Interlaminar Tests to Marine Composites. A Literature Review, submitted to Applied Composite Materials, 2003. Ramanathan, T., Bismarck, A., Schultz, E. and Subramanian, K., Investigation of the Inuence of Acidic and Basic Surface Groups on Carbon Fibres on the Interfacial Shear Strength in an Epoxy Matrix by Means of Single-bre Pull-out Test, Composite Science and Technology 61, 2001, 599605. Drzal, L. T., Sugiura, N. and Hook, D., The Role of Chemical Bonding and Surface Topography in Adhesion between Carbon Fibers and Epoxy Matrices, Composite Interfaces 4, 1997, 337354. Jacobasch, H. J., Grundke, K., Uhlmann, P., Simon, F. and Mder, E., Comparison of Surfacechemical Methods for Characterizing Fiber-epoxy Resin Composites, Composite Interfaces 3, 1996, 293320. Liu, F. P., Wolcott, M. P., Gardner, D. J. and Rials, T. G., Characterization of the Interface between Cellulosic Fibers and a Thermoplastic Matrix, Composite Interfaces 2, 1994, 419432. Wagner, H. D., Gallis, H. E. and Wiesel, E., Study of the Interface in Kevlar 49-epoxy Composites by Means of Microbond and Fragmentation Tests: Effects of Materials and Testing Variables, Journal of Material Science 28, 1993, 22382244. Nardin, M., Asloun, E. M. and Schultz, J., Physico-chemical Interactions between Carbon Fibers and PEEK, in Controlled Interfaces in Composite Material, H. Ishida (ed.), Elsevier, New York, 1990, pp. 285293. Nardin, M. and Schultz, J., Relationship between Fibre-matrix Adhesion and the Interfacial Shear Strength in Polymer-based Composites, Composite Interfaces 1, 1993, 172 192. Nardin, M. and Schultz, J., Relations between Work of Adhesion and Equilibrium Interatomic Distance at the Interface, Langmuir 12, 1996, 42384242. Pisanova, E. and Mder, E., Acid-base Interactions and Covalent Bonding at a Fibre-matrix Interface: Contribution to the Work of Adhesion and Measured Adhesion Strength, Journal of Adhesion Science Technology 14(3), 2000, 415436. Shen, W. and Parker, I. H., Surface Composition and Surface Energetics of Various Eucalypt Pulps, Cellulose 6, 1999, 4155. Rield, B. and Kamdem, P. D., Estimation of the Dispersive Component of Surface Energy of Polymer-grafted Lignocellulosic Fibers with Inverse Gaz Chromatography, Journal of Adhesion Science Technology 6(9), 1992, 10531067. Harkins, W. D., Surface Energy and the Orientation of Molecules in Surfaces as Revealed by Surface Energy Relations, Z. Phys. Chem. 139, 1928, 647691. Fowkes, F. M., Additivity of Intermolecular Forces at Interfaces. 1. Determination of the Contribution to Surface and Interfacial Tensions of Dispersion Forces in Various Liquids, Journal of Physical Chemistry 67, 1963, 25382541. Owens, D. K. and Wendt, R. C., Journal of Applied Polymer Science 13, 1969, 1741. Mder, E., Grundke, K., Jacobash, H. J. and Wachinger, G., Surface, Interphase and Composite Property Relations in Fibre-reinforced Polymers, Composites 25(7), 1994, 739 744. Bismarck, A., Richter, D., Wuertz, C. and Springer, J., Basic and Acidic Surface Oxides on Carbon Fiber and Their Inuence on the Expected Adhesion to Polyamide, Colloids Surfaces 159A, 1999, 341350.

2. 3.

4.

5.

6.

7.

8.

9.

10. 11.

12. 13.

14. 15.

16. 17.

18.

96
19.

CHRISTOPHE BALEY ET AL.

Park, S. J. and Kim, T. J., Studies on Surface Energetics of Glass Fabrics in a Unsaturated Polyester Matrix System: Effect of Sizing Treatment on Glass Fabrics, Journal of Applied Polymer Science 80, 2001, 14391445. 20. Mder, E., Grundke, K., Jacobash, H. J. and Wachinger, G., Surface, Interphase and Composite Property Relations in Fibre-reinforced Polymers, Composites 25(7), 1994, 739744. 21. Herrera-Franco, P., Wu, W. L., Madhukar, M. and Drazl, L. T., Comtempory Methods for Measurement of Fiber-matrix Interfacial Shear Strength, in 46th Annual Conference, Composite Institute, Session 14B/1-6, The society of the Plastics Industry, 1991. 22. Keusch, S., Queck, H. and Gliesche, K., Inuence of Glass Fibre/Epoxy Resin Interface on Static Mechanical Properties of Unidirectional Composites and on Fatigue Performance of Cross Ply Composites, Composites 29A, 1998, 701705. 23. Keusch, S. and Haessler, R., Inuence of Surface Treatment of Glass Fibres on the Dynamic Mechanical Properties of Resin Composites, Composites 30A, 1999, 9971002. 24. Smith, C. S., Design of Marine Structures in Composite Materials, Elsevier Applied Science, London, 1990. 25. Compston, P., Jar, P. Y. B. and Davies, P. I., Matrix Effect on the Static and Dynamic Interlaminar Fracture Toughness of Glass-bre Marine Composites, Composites 29B, 1998, 505516. 26. Tucker, R., Compston, P. and Jar, P. Y. B., The Effect of Post-cure Duration on the Mode I Interlaminar Fracture Toughness of Glass-bre Reinforced Vinylester, Composites 32A, 2001, 129134. 27. Albersen, H. and Peters, P., The Inuence of Fibre Surface Treatment on Interphase Properties in CFRP, in ECCM-5 European Conference on Composite Materials Bordeaux, A. R. Bunsell, J. F. Jamet and A. Massiah (eds), 1992, pp. 391396. 28. Eustache, R. P., Etude par RMN 13C haute rsolution dans les solides de la structure chimique des rsines polyester et de la dynamique molculaire dans les rsines poxy et polyester, Thse de doctorat, Universit Paris VI, France, 1990. 29. Espuche, E., Galy, J., Gerard, J. F., Pascault, J. P. and Sautereau, H., Inuence of Crosslinked Density and Chain Flexibility on the Mechanical Properties of Model Epoxy Networks, Macromolecular Symposium 93, 1995, 107115. 30. Labat, L., Etude des dfauts de type vides pour la matrise du procd RTM, PhD thesis, Universit du Havre, France, 2001. 31. Labat, L., Brard, J., Pillut-Lesavre, S. and Bouquet, G., Void Fraction Prevision in LCM Parts, European Physical Journal Applied Physics 16(2), 2002, 269279. 32. Lee, C.-L. and Wei, K.-H., Effect of Material and Process Variables on the Performance of Resin-transfer-molded Epoxy Fabric Composites, Journal of Apllied Polymer Science 77, 2000, 21492155. 33. Luo, Y., Verpoest, I., Hoes, K., Vanheule, M., Sol, H. and Cardon, A., Permeability Measurement of Textile Reinforcements with Several Test Fluids, Composites 32A(10), 2001, 14971504. 34. Lekakou, C. and Bader, M. G., Mathematical Modelling of Macro- and Micro-inltration in Resin Transfer Moulding (RTM), Composites 29A, 1998, 2937. 35. Sevostianov, I. B., Verijenko, V. E., von Klemperer, C. J. and Chevallereau, B., Mathematical Model of Stress Formation During Vacuum Resin Infusion Process, Composites 30B, 1999, 513521. 36. Amico, S. and Lekakou, C., An Experimental Study of the Permeability and Capillary Pressure in Resin-transfer Moulding, Composites Science and Technology 61, 2001, 1945 1959. 37. Gunde, R., Kumar, A., Lehnert-Batar, S., Mder, R. and Windhad, E. J., Measurement of the Surface and Interfacial Tension from Maximum Volume of a Pendant Drop, Journal of Colloid and Interface Science 244, 2001, 113122.

APPLICATION OF INTERLAMINAR TESTS TO MARINE COMPOSITES

97

38. 39.

40. 41.

42.

43. 44. 45.

46. 47.

48. 49.

50.

51.

52.

53.

54.

55. 56. 57.

Wu, S., Experimental Methods for Contact Angle and Interfacial Tensions, in Polymer Interface and Adhesion, Marcel Dekker Inc., New York, 1982, pp. 257278. Carroll, B. J., The Accurate Measurement of Contact Angle, Phase Contact Areas, Drop Volume, and Laplace Excess Pressure in Drop-on-ber Systems, Journal of Colloid and Interface Science 57(3), 1976, 488495. Yamaki, J. I. and Katayama, Y., New Method of Determining Contact Angle between Monolament and Liquid, Journal of Applied Polymer Science 19, 1975, 28972909. Song, B., Bismarck, A., Tahhan, R. and Springer, J., A Generalized Drop Length-height Method for Determination of Contact Angle in Drop-on-bre Systems, Journal of Colloid and Interface Science 197, 1998, 6877. Rebouillat, S., Letellier, B. and Steffenino, B., Wettability of Single Fibres beyond the Contact Angle Approach, International Journal of Adhesion and Adhesives 19, 1999, 303 314. Piggott, M. R., Why Interface Testing by Single-bre Methods Can Be Misleading?, Composite Science and Technologie 57, 1997, 965974. Miller, B., Muri, P. and Rebenfeld, L., A Microbond Method for Determination of the Shear Strength of a Fiber/Resin Interface, Composite Science and Technology 28, 1987, 1732. Liu, C. H. and Nairn, J. A., Analytical and Experimental Methods for a Fractures Mechanics Interpretation of Microbond Test Including the Effects of Friction and Thermal Stresses, International Journal of Adhesion and Adhesives 19, 1999, 5970. Venkatakrishnaiah, S. and Dharani, L. R., Interfacial Stresses in a Microbond Pull-out Specimen, European Journal of Mechanics/Solids 13(3), 1994, 311325. Scheer, R. J. and Nairn, J. A., A Comparison of Several Fracture Mechanics Methods for Measuring Interfacial Toughness with Microbond Tests, Journal of Adhesion 53, 1995, 4568. Zhandarov, S. F. and Pisanova, E. V., The Local Bond Strength and Its Determination by Fragmentation and Pull-out Tests, Composites Science and Technology 57, 1997, 957964. Day, J. R., Marquez, M., Verpoest, I. and Jones, F., in Interfacial Phenomena in Composite Materials 91, Leuven, Belgique, 1991, Butterworth Heinemann (ed.), London, 1991, pp. 7376. Ash, J. T., Cross, W. M., Svalstad, D., Kellar, J. J. and Kjerengtroen, L., Finite Element Evaluation of the Microbond Test: Meniscus Effect, Interphase Region, and Vise Angle, Composites Science and Technology 63, 2003, 641651. Moon, C. K., The Effect of Interfacial Microstructure on the Interfacial Strength of Glass Fiber/Polypropylene Resin Composites, Journal of Applied Polymer Science 54, 1994, 7382. Wu, H. F. and Claypool, C. M., An Analytical Approach of the Microbond Test Method Used in Characterizing the Fibre-matrix Interface, Journal of Material Science Letters 10, 1991, 260262. Nishiyabu, K., Yokoyama, A. and Hamada, H., Assessment of the Interfacial Properties on Stress Transfer in Composite by a Numerical Approach, Composites Science and Technology 57, 1997, 975983. Hodzic, A., Kalyanasundaram, S., Lowe, A. E. and Stachurski, Z. H., Geometric Considerations in the Experimental and Finite Element Analyses of the Microdroplet Tests, in 12th International Conference on Composite Materials, Paris, 1999, 9 pages. Day, R. J. and Cauich Rodrigez, J. V., Investigation of the Micromecanics of the Microbond Test, Composites Science and Technology 58, 1998, 907914. Hann, L. P. and Hirt, D. E., Simulating the Microbond Technique with Macrodropplets, Composites Science and Technology 54, 1995, 423430. Roselli, F. and Santare, M. H., Comparison of the Short Beam Shear (SBS) and Interlaminar Shear Device (ISD) Tests, Composites 28A, 1997, 587594.

98
58.

CHRISTOPHE BALEY ET AL.

59.

60.

61. 62.

63. 64.

65.

66.

67.

68. 69. 70. 71.

72. 73. 74. 75. 76.

Roudet, F., Comportement en exion trois points avec cisaillement prponderant de composites verre/epoxyde unidirectionnels: sous chargement monotone et cyclique, Thse de doctorat, Universit des Sciences et Technologies, Lille, 1998. Williams, J. C., Yurgartis, S. W. and Moosbrugger, J. C., Interlaminar Shear Fatigue Damage Evolution of 2-D Carbone-carbone Composites, Journal of Composite Materials 30(7), 1996, 785799. Hoecker, F., Friedrich, K., Blumberg, H. and Karger-Kocsis, J., Effects of Fiber/Matrix Adhesion on Off-axis Mechanical Reponse in Carbone-ber/Epoxy-resin Composites, Composites Science and Technology 54, 1995, 317327. Bagueri, M., The Three-points Bending Test, in Critical Review Proceedings of the 8th JapanUS Conference on Composite Materials, Technomic Publ. (ed.), Baltimore, 1998, pp. 683692. Pang, K. P. and Gillham, J. K., Anomalous Behaviour of Cured Epoxy Resins: Density at Room Temperature versus Time and Temperature of Cure, Journal of Applied Polymer Science 37, 1989, 19691991. Zinck, P., Wagner, H. D., Salmon, L. and Gerard, J. F., Are Microcomposites Realistic Models of the Fibre/Matrix Interface? I. Micromechanical Modelling, Polymer 42, 2001, 54015413. Kessler, H., Schller, T., Beckert, W. and Lauke, B., A Fracture-mechanics Model of the Microbond Test with Interface Friction, Composites Science and Technology 59, 1999, 22312242. Pisanova, E., Zhandarov, S., Mader, E., Ahmad, I. and Young, R. J., Three Techniques of Interfacial Bond Strength Estimation from Direct Observation of Crack Initiation and Propagation in Polymer-bre Systems, Composites 32A, 2001, 435443. Hodzic, A., Kalyanasundaram, S., Kim, J. K., Lowe, A. E. and Stachurski, Z. H., Application of Nano-indentation, Nano-scratch ang Single Fibre Tests in Investigation of Interphases in Composite Materials, Micron 32, 2001, 765775. Schller, T., Beckert, W. and Lauke, B., A Finite Element Model to Include Interfacial Roughness into Simulations of Micromechanical Tests, Computational Materials Science 15, 1999, 357366. Herrera-Franco, P. J. and Drzal, L. T., Comparison of Methods for the Measurement of Fibre/Matrix Adhesion in Composites, Composites 23, 1992, 227. Pisanova, E., Zhandarov, S. and Mder, E., How Can Adhesion Be Determined from Micromechanical Tests?, Composites 32A, 2001, 425434. Penn, L. S. and Lee, S. M., Interpretation of Experimental Results in the Single Pull-out Filament Test, Journal of Composites Technology and Research 11(1), 1989, 2330. Piggott, M. R., Sanadi, A., Chua, P. S. and Andison, D., Mechanical Interactions in the Interfacial Region of Fiber Reinforced Thermosets, in Composites Interfaces, H. Ishida and J. L. Koenig (eds), Elsevier Science publishing, 1986, pp. 109121. Nairn, J. A., Thermoelastic Analysis of Residual Stresses in Unidirectional, High-Performance Composites, Polymer Composites 6(2), 1985, 123130. DiLandro, L., Dibenedetto, A. T. and Groeger, J., The Effect of Fiber-matrix Transfer on the Strength of Fiber-reinforced Composite Materials, Polymer Composites 9(3), 1988, 209221. DiLandro, L. and Pegoraro, M., Evaluation of Residual Stresses and Adhesion in Polymer Composites, Composites 27A, 1996, 847853. Chua, P. S. and Pigott, M. R., The Glass Fibre-polymer Interfac: III Pressure and Coefcient of Friction, Composites Science and Technology 22, 1985, 185196. Dzral, L. T., Herrera-Franco, P. and Ho, H., Test Methods for Measurement of Fiber-matrix Adhesion, in Comprehensive Composite Materials 5, A. Kelley and C. Zweben (eds), Elsevier Science, 2000.

Anda mungkin juga menyukai