This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
0 correspond to the
Table 9-15. Aggregate average of elastic constants for a planar distribution. / p = 2
G*
=
13 > 0
= [/c*(l+v* T )-G* T ]
transversely isotropic (quasi-isotropic) case for a random in-plane distribution. The term
460
9 Elastic Properties of Composites
through E£. As was the case for aligned inclusions, the ellipsoid and cylindrical model predictions differ by less than 10% for aspect ratios greater than the limiting aspect ratio. Comparison of model predictions with experimental results is hampered by the lack of reports which include simultaneous measurements of elastic properties and the state of fiber orientation. Based on qualitative inspection, many workers assume a random planar distribution. This approach can be misleading since / p values ~0.3 can give the qualitative appearance of a random distribution (McGee and McCullough, 1981). Even those workers that report quantitative measurements of the state of orientation tend to limit the evaluation of elastic behavior to the Young's modulus. Camacho et al. (1990) examined discontinuous quartz and Kevlar 29 fibers embedded in a dicyclopentadiene Novalac dicyanate thermosetting resin. Composites were prepared by a process similar to paper processing to give mats of well-dispersed filaments. The mats and polymer were consolidated and cured in a compression mold to produce plaques approximately 3.2 mm (0.125 inches) thick. The individual filaments were 6.4 mm (quartz) and 12.7 mm (Kevlar 29) in length with diameters of 9.0 and 12.0 jam respectively. Fiber volume fractions were in the range 0.06 to 0.20. These workers measured the orientation distribution of the isolated filament in the plaques by a procedure described by Konicek (1987). These measurements showed that the filaments were uniformly distributed within the plane (/ p = 0) with a slight out-of-plane tilt of 5-10°. Camacho et al. (1990) used the Halpin-Tsai relationships to obtain the quantities P* of Table 9-15. Upon accounting for void fractions in the range of 7-15%, these workers re-
ported good agreement between predicted and measured elastic moduli in the sense that the differences generally fell within the experimental error of 10 to 20%. The dispersed filament aspect ratios of 710 (quartz) and 1,060 (Kevlar 29) place the systems within the "continuous fiber" range, with E£ = vfEf-\- vmEm. Accordingly, both the cylindrical model and ellipsoidal model are consistent with aspect ratios in this range. Kacir et al. (1978 b) provide a rare study in which measurements of nonrandom fiber orientations are concurrently reported with mechanical property evaluations. These workers used a special converging flow technique to obtain various orientations of strands of E-glass chopped into 3.2 mm (0.125 inches) and 6.35 mm (0.25 inches) length. Each strand contained 400 filaments of 10 jim diameter. Fiber orientation was determined by counting colored tracer fibers and counting X-ray photographs of nylon coated steel tracer fibers. The orientation was determined to be essentially planar with minimal out-of-plane tilting. Orientation distribution data was reported as the cumulative percent of aligned fibers in increments of 5°. These data can be used to generate the quantities
2-/ p
461
9.6 Relationship of Microstructure to Elastic Behavior of Discontinuous Fiber Systems
Table 9-16. Comparison of effective aspect ratios for cylindrical and ellipsoidal regions in various states of planar orientation. Orientation parameters
Aspect ratios
Young's modulus, (E± !> in GPa
Effective
a
4
/P
Filament
Cylinder
Ellipsoid
Cylinder a
Ellipsoid
Measured b
0.55 0.71 0.90 0.94
0.42 0.60 0.86 0.90
0.42 0.58 0.82 0.89
0.10 0.42 0.80 0.88
635 317 635 317
136 53 136 53
45 30 45 30
21.0 24.6 32.5 30.7
21.0 23.3 30.5 30.6
20.6 23.6 32.5 30.6
Carman and Reifsnider (1992);
b
Kacir et al. (1978 b), as reported by Carman and Reifsnider (1992).
As discussed in the previous section, Kacir et al. (1978 a) proposed that an effective aspect ratio be used rather than the filament aspect ratio to account for populations of close-packed bundles that would tend to behave as coherent reinforcing agents. Using the effective aspect ratios, Carman and Reifsnider (1992) obtained excellent agreement between predicted and measured values for <En>, based on cylindrical inclusion. Again, there is no compelling reason to require that the coherent region retain a cylindrical shape. The data of Kacir et al. (1978 b) can be reconciled in terms of effective ellipsoidal aspect ratios. The results obtained for effective cylindrical aspect ratios and effective ellipsoidal aspect ratios are compared in Table 9-16. Sheet molding compounds are a common form of discontinuous fiber composites exhibiting planar orientation. Sheet molding compounds are made by dropping 25 to 50 mm (1 to 2 inch) bundles (or "roving") containing about 400 glass filament onto a moving belt and impregnating the mat with an unsaturated polyester resin. Composites are fabricated by heating the material in a compression mold to cure the thermosetting resin. Although the sheet molding process is intended to generate nearly random in-plane orientation,
the belt movement and flow within the mold cavity can induce some orientations. Microscopic examination of low fiber content SMC-25 material (weight fraction = 0.25, i;f = 0.2) shows that the bundle structure is preserved. However, the bundle aspect ratio is approximately 200. As a consequence, the material can be approximated by a continuous fiber system. Using an average over elastic constants, McCullough et al. (1983) show that an orientation factor of / p ~ 0.2 and ae -> oo bring predicted values into good correspondence with measured values of {EX1} and (E22y. For the higher fiber content material SMC-65 (weight fraction = 0.65; vf = 0.5), the dense packing and overlapping of the rovings produce a grainlike structure of locally parallel filaments. Within a given plane these grainlike domains have small length-to-width ratios as compared to the original tow. These domains are connected by common filaments which are much longer than the length of the domain. McCullough etal. (1983) show that the experimental and predicted values of <£n> and (E2iy could be reconciled with / p = 0.3 and ae w 2. These workers propose that a value ae« 2 is consistent with the observed in-plane domains of denselypacked parallel filaments.
462
9 Elastic Properties of Composites
In summary, experimental evidence suggests that orientation averaging of the elastic constant array C captures the essential features of planar orientation distribution for systems with large filament aspect ratios. Comparison of orientation parameters / p and Fp with measured orientation distributions imply that Fp is dependent on / p so that only one independent orientation parameter is required to describe planar orientation. Comparison for other orientation distributions are needed to substantiate this simplification. With the exception of systems with isolated filaments, predictions based on the filament aspect ratio yield values for <£1X> which are larger than observed. Since lower bound results are used as the basis for both the ellipsoidal and cylindrical inclusion models, the predicted < £ n ) should be smaller than the observed values. As was the case for aligned systems, the notion of an effective aspect ratio can be employed to reconcile model predictions with experimental observation. The smaller value of the effective aspect ratio is rationalized in terms of close-packed bundles of filaments that appear to respond to loads and deformation as coherent units. At this time no model is available to predict effective aspect ratios. Accordingly, resort has been made to using the effective aspect ratio as an adjustable structural parameter. As shown in the previous section, the notion of the effective aspect ratio serves as a basis for establishing data reduction schemes that appear to provide systematic correlations. 9.6.3 Three-Dimensional Orientations
Three-dimensional states of orientation are associated with section thickness much larger than the lengths of the fibers. In this case, dimensional constraints no longer
confine the fibers to a plane. Composite structures of this type are usually produced by injection or transfer molding processes in which mixtures of polymer and chopped fibers are forced through gates and runners into a mold cavity in the shape of the desired part. The varying flow fields within the cavity carry the fibers into complex states of three-dimensional orientation. Converging flows through constricted regions tend to enhance fiber alignment; diverging flows from constricted regions tend to promote random orientations. For complex mold geometries, the state of orientation can vary with location. Even simple rectangular bars can exhibit spatially dependent orientations. The fibers near the surface of the cavity tend to align in the flow direction; fibers near the midsection approach random orientations. The thickness of the "skin" of nearly aligned fibers is dependent on flow-rates and the temperature of the mold surface. The development of processing models to predict the state of local fiber orientation induced in mold cavities with complex geometry remains a major challenge to the advancement of the technology of short-fiber composites. The milling actions encountered while forcing the polymer/fiber mixture into the mold cavity tend to produce dispersed filaments. Consequently, the aggregate averaging scheme may not be applicable to these materials. Injection and transfer molding processes require fibers of substantially smaller lengths and lower volume fractions of fibers than those used in SMC compounds. In addition, the mechanisms used to force the material into the mold cavity cause extensive fiber fragmentation resulting in a wide range of aspect ratios and end-shape geometries. These complex and varying structural features present a considerable challenge to both the development of micromechani-
9.6 Relationship of Microstructure to Elastic Behavior of Discontinuous Fiber Systems
cal models and the experimental characterization of the materials. Even in the ideal situation of spatially independent states of orientation, four orientation parameters are required to characterize three-dimensional states of orientation (fa9 Fa9 / p , Fp). As a consequence of fiber fragmentation, at least two aspect ratio parameters may be needed to describe the mean and spread of the aspect ratio distribution. These structural characterizations are difficult to conduct and as a consequence insufficient experimental data are available to test the validity of micromechanical models for discontinuous fibers in three-dimensional states of orientation. Semiempirical correlations with four to six adjustable parameters are expected to be successful but will give limited insights into the systematic development of optimum processes. Alternately, such semiempirical correlations may be the only means of relating processing conditons to the mechanical performance of short-fiber composites. In this application, the micromechanical model provides a rational basis for identifying the key structural parameters and the proper form for the correlations. In the remainder of this section, simplifying conditions will be introduced to facilitate the identification of significant trends and provide a framework for data reductions to establish process-property correlations. Attention will be directed to representative volume elements which are large enough to sample the local homogeneous elastic characteristics of the material but small enough to avoid significant spatial variations in the state of orientation. This approach has been employed by Bozarth et al. (1987) to model the skin-core behavior of injection molded bars. Fiber orientation distributions were measured for sections of the skin and the core. These structural parameters were used to predict
463
the elastic properties of the skin and core layers. These predicted properties were used as input to a traditional laminate analysis to predict the over-all behavior of the bar. Good correspondence was obtained between the observed and predicted longitudinal Young's modulus as well as the longitudinal and transverse coefficients of thermal expansion. Under the restriction of spatially independent fiber orientation, the general three-dimensional state of orientation within the representative volume element can be characterized by the relationships summarized in Table 9-12. Various averaging schemes may be generated from these relationships. The dispersed averaging scheme is obtained by replacing all Atj by Mij9 compliance averaging is obtained by replacing Atj (i = 1, 2, 3; j = 1, 2, 3), Su and Akk (k = 4, 5, 6) with (1/4) Skk; the Stj on the right-hand side are obtained from the predictions of a selected micromechanical model (e.g., ellipsoidal or cylindrical). Aggregate averaging is obtained by replacing (Atjy by
464
9 Elastic Properties of Composites
tribution could be encountered in rod-like extrusion processes. The reductions of the general Atj to the axial form are given in Table 9-17. This special distribution is transversely isotropic with respect to the 3 axis so that five independent elastic constants are required to describe the elastic behavior. For fa = l9 the relationships revert to the aligned case treated in Sec. 9.6.1. The quantities Ko and Go correspond to the bulk modulus and shear modulus of an isotropic material generated by a random three-dimensional distribution (i.e., fa = f^ = Fa = Fp = 0). Since an isotropic system can be described by two independent elastic constants, the Young's modulus and Poisson ratios for an isotropic three-dimensional random distribution are given by 1
Vn =
1 1 3Xn 9K« 3 Ko — 2 Go
(9-202 a)
(9-202 b)
The elastic behavior of random three-dimensional orientations is strongly dependent upon aspect ratio as reflected by the dependence of Ko, Go (or £ 0 , v0) on ££. This dependence of Eo and Go on aspect ratio is illustrated for a typical glass/polymer system in Figs. 9-14 and 9-15. For these examples, the aspect ratio dependence is based on an ellipsoidal model. The ellipsoidal aspect ratio dependence depicted in Figs. 9-15 and 9-16 can be converted to cylindrical aspect ratios by treating Eq. (9-199) as a scaling relationship relating ellipsoidal and cylindrical aspect ratios. These plots show that the change in moduli with respect to change in aspect ratio is largest in the region 1 < a < 10 and decreases for a > 10 to become vanishingly small for a > 100. In all cases, for a > 100 the moduli are essentially equivalent to the corresponding moduli for continuous fiber systems. The behavior of biased three-dimensional distributions for which fa ^ 0, / p ^ 0
Table 9-17. Aggregate average of elastic constants for an axial distribution. /. = j[3
+
f AG 0 Fa
4v*T2) - GTT] + 2 AG 0 Fa
15G o =
6(Gf
9.6 Relationship of Microstructure to Elastic Behavior of Discontinuous Fiber Systems 80 a=1 O=10
60 Ui
T3 O 2
0=100
J
40
20
>i
0.2
1 0.4 0.6 0.8 Volume fraction. Vi Figure 9-15. Elastic modulus of a polycarbonate/ glass composite with a three-dimensional random fiber orientation distribution as a function of fiber volume fraction and aspect ratio.
35 3 30
i\
CD
«/> 2 0
].
/
i
h «
0.2
0.4
0.6
0.8
bution (n((j)) = A cosk<}) (j)\ this condition
occurs at /p* = 2 - ^ 3 = 0.27. The quantity
I 5 0
dominates the behavior, <£*!> uniformly decreases with increasing fa and decreasing / p . The minimum value of <£*!> occurs at (/ p = 0, fa = 1) corresponding to E% for the aligned case. The maximum value for (^A17y occurs at (/ p = 0, fa = —1/2). This corresponds to the value of <£*2> f° r a random planar distribution; <£*2> decreases as / p and fa increase to the minimum at fa = i corresponding to Ef for an aligned system. The shear terms attain maximum values at an intermediate state of orientation. The quantity (A66), associated with
o=5 o=10 o=100
T25
465
1
= 0. For the assumed coske(9)
Volume fraction, Vf
Figure 9-16. Shear modulus of a polycarbonate/glass composite with a three-dimensional random fiber orientation distribution as a function of fiber volume fraction and aspect ratio.
requires quantitative values of fa and / p . However, general trends can be inferred from the relationships given in Table 9-11. The term <^ 33 > in Table 9-11 is independent of / p , but strongly dependent upon fa. Since the Young's modulus, <£33>, is dominated by this term, variations in fiber orientation in the 1-2 plane exert minimal influence on <£ 33 >. The quantity < i n ) of Table 9-10 takes on a maximum value at fa = —1/2 and / p = l, corresponding to the case for aligned inclusions, £J\ Since this term
form of the orientation distribution this occurs at f* = 1/2(5 - 3 y/2) = 0.38; G13 decreases from G13 (fa = 0.38; / p = 1) to the value for the planar random distribution (fa = - I/ 2 , /p = 0). The shear term {A44}, associated with
ject to
a= f
466
9 Elastic Properties of Composites
The off-diagonal elements
Table 9-18. Location of maxima and minima in three-dimensional state of orientation. Maxima
General Aggregate tensor model
(9-205)
= {Ct-Cm)
Applying the approximations used to simplify the expression for C* yields the following expressions for the components of the transversely isotropic tensor M:
km(oo) =
Ak
(l-x o )(A/c/GJ
/.
/P
-1/2 -1/2 1
1 0 0
1 1 -1/2
0 0 1
(A23>
-1/2 0.38 0.4-6.6
0.27 1 0
1 -1/2 -1/2
0 0 1
O*55> <^66>
0.4-0.6 0.38 -1/2
0 1 0.27
-1/2 -1/2 1
1 0 0
M 1 3 (a) = M 1 3 (oo)-
1 -
(AE/GJ(2v(HD-x0h4) t[l + (AE/Ga)HD] AC
M 13 (oo) =
(9-206 b)
13
(AE/GJHD l+(AE/GJHD/ (9-206 c)
(AE/Gm)(l-x0)Ak
M33(co) =
(l-xo)(Afc/GJ
M 6 6 (a) = - [Mt x (a) - Mt2 (a)] = M 6 6 (oo) 1 +
M 66 (OD)
2 Gm - M 6 6 (oo) [(2 - x 0 ) h2 - x0
M66(oo) =
2G m AG
1+
(AE/GJ(l-xo)h22 1+(AE/GJHD
/P
(9-203)
(9-204)
fa <^n>
C,*b,d = Cm + vt(v,Hf + »m<Mf > " 1 ) " 1
where the quantities <M f > y are obtained by replacing all Atj in Table 9-11, with the My. The Mi} are generated from the shape tensor E^ and elastic constants of the components.
(9-206 d)
(2-x 0 )AG M44(oo)
2Gm-M44(oo)(4x0h-h2) 2GmAG M 44 (oo) = (9-206 e) with AP = P f - P m , xo =
(9-206 a)
Minima
HD = (1 —xo)h1h2
1
2(1-vj
+ xoh4..
, and
467
9.6 Relationship of Microstructure to Elastic Behavior of Discontinuous Fiber Systems
Comparisons between the Young's moduli and shear moduli obtained under the dispersed and aggregate averaging schemes for axial orientations (/ p = 0) are given in Figs. 9-17 and 9-18. These comparisons are based on a glass/polycarbonate system with 0.4 volume fraction glass fiber and an ellipsoidal aspect ratio of 100. An ellipsoidal aspect ratio of 100 is equivalent to a continuous fiber and was selected to emphasize the difference between the two averaging schemes. The differences will decrease with decreasing aspect ratio and converge to equivalent predictions for a = \. The comparisons displayed in Figs. 917 a and 9-17 b show that the dispersed av-
eraging scheme yields consistently higher values for the Young's moduli. The largest difference in the axial Young's modulus, E 3 , occurs for a random planar orientation (fa = - 1/2, / p = 0). This difference decreases with increasing axial alignment and vanishes for an aligned system (fa = 1, / p = 0). The dispersed and aggregate averages of the in-plane Young's modulus ET = E2 are equivalent for both the planar random state (fa = -1/2, fp = 0) and the aligned state (/fl = l, / p = 0). However, at intermediate states of axial orientation, the dispersed average moduli are significantly larger than the aggregate averaged moduli. The Young's modulus for dispersed averaging for a random three dimensional state
35 30
Dispersed averaging Aggregate averaging
(a)
Dispersed averaging Aggregate averaging
(a)
\
to 3 i -0.5
0.5
-0.5
0.5
18 — CO
16
Dispersed averaging Aggregate averaging
(b) :
O
uf
12
LJlUS,
14
10 —
-0.5
0.5
Figure 9-17. (a) Axial modulus (E3) and (b) in-plane modulus (El=E2) predictions for aggregated and dispersed polycarbonate/glass systems; / p = 0.0, v{ = 0.4, aspect ratio = 100.
Dispersed averaging Aggregate averaging
1.5 -0.5
0.5
1
Figure 9-18. Shear modulus predictions, (a) G12 and (b) G13 = G23, for aggregated and dispersed polycarbonate/glass systems; fp = 0.0, vf = 0.4, aspect ratio = 100.
468
9 Elastic Properties of Composites
of orientation (/R = / p = 0) is about 25% larger than the aggregate averaged Young's modulus. The comparison of the dispersed averaging scheme and the aggregate averaging scheme for the shear moduli is shown in Fig. 9-18. Similar to the in-plane Young's modulus, the largest differences in the transverse modulus, GTT, occur for a random planar orientation (fa = —1/2, fp = 0), decreases with increasing axial orientation and vanish for an aligned system (/a = l, /„ = 0). The dispersed and aggregate averaged axial shear moduli, GLT, are equivalent for both the planar random state (fa = -1/2, / p = 0) and the axially aligned state (fa = 1, / p = 0). As expected from the previous discussion, the largest differences occur at the maximum at fa « 0.4 with the dispersed averaging result approximately 50% larger than the aggregate averaging result. These comparisons show that dispersed textures provide significantly enhanced reinforcing efficiencies over aggregated textures. Indeed, if the effective aspect ratio for clusters is used as a basis for predicting the aggregate elastic properties and the filament aspect ratio for dispersed elastic properties, the difference between the dispersed and aggregate textures is further increased.
9.6.4 Summary
Simplifications were introduced to provide a direct means of assessing the influence of various structural features on the elastic behavior of discontinuous fiber systems. The structural features treated were the state of orientation as specified by scaled orientation parameters (fa, Fa, / p , Fp), aspect ratios, and shape effects (associated with ellipsoidal and cylindrical inclu-
sions), as well as dispersed and aggregated textures. Comparison of the predictions from the dispersed averaging scheme and from the aggregate averaging scheme show that the dispersed textures provide significantly enhanced reinforcing efficiencies over aggregated textures. If the effective aspect ratio for clusters is used as a basis for predicting the aggregate elastic properties, and the filament aspect ratio for dispersed elastic properties, the difference between the dispersed and aggregated textures is further increased. The cylindrical shape of Carman and Reifsnider (1992) appears to be an appropriate model for the dispersed filaments of known cylindrical shapes. However, based on filament aspect ratios, both the cylindrical model and the ellipsoidal model yield predictions of EL which are considerably higher than those experimentally observed by several workers. Both the cylindrical and ellipsoidal models are based on lower bound formulations, and therefore these predictions should fall below the observed values. The incorporation of three-point and higher correlations will cause the predictions to give yet higher values, and hence their omission from the cylindrical and ellipsoidal models cannot be the source of discrepancy. Evidently, additional structural features, not considered in the formulation of the micromechanical models, must play a significant role in establishing elastic behavior. The notion of an effective aspect ratio based on clusters of filaments responding as a coherent reinforcing unit appears to provide a means for reconciling a wide range of experimental observations. At this time no model is available to predict effective aspect ratios. Accordingly, resort has been made to using the effective aspect ratio as an adjustable parameter which can
9.7 Summary
serve as a basis for establishing data reduction schemes that appear to provide systematic correlations. Experimental evidence suggests that orientation averaging of the elastic constant array C captures the essential features of planar orientation distribution for systems with large filament aspect ratios. Comparison of orientation parameters, / p and Fp, with measured orientation distributions imply that Fp is dependent of / p so that only one independent orientation parameter is required to describe planar orientation. Comparisons for other orientation distributions are needed to substantiate this simplification.
9.7 Summary The prediction of effective elastic moduli is a central problem underlying the development of quantitative structure-property relationships for a wide range of heterogeneous materials - e.g., composites, polycrystalline aggregates, and partially crystalline polymers. Indeed, as illustrated in Fig. 9-1, all of these material systems can be encountered at various levels in the structural hierarchy of composite materials. In this treatment, attention has been directed to the highest microstructural level consisting of reinforcing inclusions in a matrix phase. A rigorous prediction of effective elastic moduli requires a complete description of the microstructure. Rigorous treatment can be developed for woven or braided fiber architectures that can be competely described by periodic functions. However, most heterogeneous systems do not possess a precise deterministic microstructure. The stochastic nature of the microstructure of many important heterogeneous sys-
469
tems requires a statistical description of the microstructure in terms of distribution of shape and orientations. As illustrated in Fig. 9-2, the dominant geometrical features of the microstructure may be quantitatively described by orientation parameters which characterize the average state of orientation of the reinforcing agents and an effective aspect ratio which reflects the role of the average shape of the reinforcing agent. The effective aspect ratio can range from a value approaching zero for plate-like inclusions, to 1 for spherical inclusions, to a value approaching infinity for continuous fibers. Discontinuous fibers fall in a range between one and infinity. An essential step toward an understanding of the behavior of heterogeneous materials is the identification of the appropriate fundamental descriptions of the system which are susceptible to experimental characterization. It is apparent that the following information is essential input to any model to describe the behavior of composite materials: - the elastic properties of each of the components - the volume fraction concentration of each component In order to distinguish among the behaviors of widely different classes of composite materials (e.g., continuous fiber composites, particulate reinforced composites, and short-fiber composites), the following additional information is required: - a measure of the load transfer efficiency associated with various reinforcing geometries (e.g., the effective aspect ratio, - a measure of the state of fiber orientation (e.g., the orientation parameter, / ) In view of the complex and variable features of the internal microstructure, a rig-
470
9 Elastic Properties of Composites
orous treatment was abandoned in favor of the development of bounding models which attempt to capture the dominant features of the reinforcing geometry and the state of orientation. The bounding approach avoids the problems of specifying the explicit nature of the microstructure and the internal stress-strain field by employing variational principles to establish upper and lower bounds on the properties. The results from these bounding methods can provide practical guides to material behavior only if the upper and lower bounds are reasonably close together so as to bracket the properties to within experimental error. The Hashin-Shtrikman upper and lower bounds on longitudinal properties of continuous fiber composites are very close and converge to the simple results of the rule of mixtures. Unfortunately, the upper and lower bounds on transverse properties, shear moduli, and longitudinal properties of aligned discontinuous fiber composites are too far apart to be of much practical value. In light of experimental errors involved in characterizing composite materials and scatter in properties, it is prudent to base design parameters for composites on reasonably conservative estimates of the properties. The traditional Reuss model of Eq. (9-65) provides an ultraconservative estimate that would lead to inefficient use of expensive materials. The HashinShtrikman improved lower bound offers a more sensible yet adequate conservative estimate that incorporates the assurance that the observed properties of the material will be as good as or slightly better than the predicted values. Hence, it is recommended that the improved lower bound relationships (or the equivalent doubly embedded relationships or the composites cylindrical assemblage relationships) be used to predict transverse properties and shear mod-
uli of continuous fiber composites. These results are summarized in Table 9-5. Alternatively, these quantities may be computed by the Halpin-Tsai equation summarized in Table 9-6 a with the appropriate values of £p computed from the properties of the matrix as indicated by the relationships summarized in Table 9-6 b. Values for ET and vTT must be computed from the auxiliary relationships given in these tables. Particulate filled systems represent a higher level of complexity beyond continuous fiber composites. The discontinuities of particulate filled systems introduce significant fluctuations in the internal stressstrain fields which complicate the analysis. However, the ample spherical reinforcing geometry precludes a dependence on orientation so that consideration of this structural feature is not required for the analysis of these systems. Since these systems are isotropic, only two independent material constants (e.g., the bulk modulus K and the shear modulus G) are required to characterize elastic behavior. The improved Hashin-Shtrikman upper and lower bounds for K and G remain too far apart to bracket the behavior for particulate systems. However, the HashinShtrikman lower bound and the concentric spheres model adequately predict elastic behavior in the concentration range 0 < v{ < l / 3 and yield conservative estimates for vt > 1/3. The semi-empirical S-combining rule (McGee and McCullough, 1981) provides predictions in agreement with experimental data over a wide range of concentrations and for a variety of materials. This combining rule was developed by making use of the symmetry of the improved bounding expressions, with respect to the interchange of the role of the components for the continuous phase - the lower bound expression. The relationships devel-
9.7 Summary
oped from this approach are summarized in Table 9-8. The notion of a critical packing concentration, 0C, is used to designate transitions in phase continuity (or "contiguity"). When the volume fraction of either component is greater than this critical packing concentration, it is possible to unambiguously identify this phase as the continuous phase. Under these conditions of "zero" contiguity for the second component, the material system tends toward the behavior described by the bounding expressions in which the first (major) component is assigned the role of the continuum. The behavior of the material at intermediate concentrations is assumed to follow a smooth transition between the limiting behavior when either phase is identified as the continuum. An assignment of >c = 2/3 (corresponding to "random" close packing) appears to be an appropriate choice for the critical packing fraction. Discontinuous fiber reinforced materials represent a higher level of complexity than continuous fiber arid particulate systems. The processing conditions used to fabricate discontinuous fiber composites can cause fiber fragmentation resulting in a distribution of fiber lengths and a variety of end shapes which influence the capacity of the fiber to accept load transfer from the matrix. Varying flow fields carry the fibers into complicated states of orientation. In addition, two distinct textures have been observed for discontinuous fiber systems: dispersed and aggregated. A formulation capable of distinguishing between dispersed and aggregated microstructures in discontinuous fiber composites was presented. The dispersed microstructure was defined through the use of a structure function. It was shown that dispersed averaging and aggregate averaging were equivalent for the cases of spherical
471
inclusions and for perfectly aligned inclusions. The results predict that dispersed microstructures possess higher elastic properties than equivalent aggregated microstructures due to the more efficient reinforcement of a dispersed system. Most workers focus on the aggregate texture through the orientation averaging procedures employed to account for orientation effects. Aggregate averaging may be applicable to sheet-molding compounds which exhibit a domain microstructure, while the dispersed averaging technique is more suited to extruded, transfer molded and injection molded composites. A key issue in the analysis of discontinuous fiber composites is the identification of appropriate statistical parameters, representative of the microstructural features of the composite. The analysis reveals that the important statistical descriptors of the microstructure are obtained from appropriate moments of the fiber orientation distribution and the aspect ratio distribution. The scaled descriptors of the state of orientation (fa9 Ffl, / p , Fp) involve the second and fourth moments of the cosines of the angles 6 and 4>, or the first and second moments of 26 and 20. The descriptors of the aspect ratio distribution involve
472
9 Elastic Properties of Composites
geometries is suppressed with increasing inclusion aspect ratios. The cylindrical shape model of Carman and Reifsnider (1992) appears to be an appropriate model for dispersed filaments of known cylindrical shape. However, based on filament aspect ratios, both the cylindrical shape model and the ellipsoidal shape model yield predictions for longitudinal Young's modulus which are considerably larger than those experimentally observed by several workers. Both the cylindrical and ellipsoidal models are based on lower bound formulations, and therefore these predictions should fall below the observed values. Evidently, additional structural features, not treated in the formulation of these models, must play a significant role in establishing elastic behavior. The notion of an effective aspect ratio based on clusters of filaments responding as a coherent reinforcing unit appears to provide a means for reconciling a wide range of experimental observations. At this time no model is available to predict the effective aspect ratios for clusters. Finally, the general model for the effective elastic properties of discontinuous fiber composites can be reduced to the special cases of aligned fiber orientation, planar fiber orientation and three-dimensional orientation by appropriate assignments of the fiber orientation and fiber aspect ratio descriptors. Simplifications are introduced to provide a direct means of assessing the influence of various structural features on the elastic behavior of discontinuous fiber systems. The resulting models can be extended to predict the elastic behavior of partially crystalline fibers or matrices by assigning various crystalline morphologies the role of reinforcing inclusions in a matrix of "amorphous" material.
9.8 References Aboudi, J. (1983), Int. J. Solids Struct. 19, 693. Adams, D. K, Doner, D. R. (1967a), J. Compos. Mater. 1, 4. Adams, D. R, Doner, D. R. (1967b), J. Compos. Mater. 1, 52. Adams, D. R, Tsai, S. W. (1969), J. Compos. Mater. 3,6%. Adams, D. R, Doner, D. R., Thomas, R. L. (May 1967), Mechanical Behavior of Fiber-Reinforced Composte Materials, Air Force Technical Report AFML-TR-67-96. Ashton, J. W, Halpin, J. C , Petit, P. H. (1969), Primer on Composite Materials: Analysis. Lancaster, PA: Technomic. Berthelot, J. M. (1982), Fibre Sci. Technol. 17, 25. Berthelot, I M., Cupice, A., Maufras, J. M. (1978), Fibre Sci. Technol. 11, 367. Blumentritt, B. R, Vu, B. T., Cooper, S. L. (1974), Polym. Eng. Sci. 14, 633. Blumentritt, B. R, Vu, B. X, Cooper, S. L. (1975), Polym. Eng. Sci. 15, 482. Bozarth, J. M., Gillespie, J. W, Jr., McCullough, R. L. (1987), Polym. Compos. 8, 74. Budiansky, B. J. (1965), /. Mech. Phys. Solids 13, 223. Camacho, C. W, Tucker III, C. L., Yalvac, S., McGee, R. L. (1990), Polym. Compos. 11, 229. Carman, G. P., Reifsnider, K. L. (1992), Compos. Sci. Technol. 43, 137. Chou, T-W. (1992), in: Microstructural Design of Fiber Composites: Cahn, R. W., Davis, E. A., Ward, T. M. (Eds.). Cambridge: Cambridge University Press. Chou, T-W, Ko, R K. (1989), Textile Structural Composites, Composite Materials Series, Vol. 3. Amsterdam: Elsevier. Chou, T-W, Nomura, S., Taya, M. (1980), /. Compos. Mater. 14, 178. Christensen, R. M. (1979), Mechanics of Composite Materials: New York: Wiley. Christensen, R. M., Lo, K. H. (1979), J. Mech. Phys. Solids 27, 315. Cox, H. L. (1952), Br. J. Appl. Phys. 3, 72. Eduljee, R. R (1991), Ph.D. Thesis, University of Delaware, Newark, Delaware. Eduljee, R. R, McCullough, R. L. (1992), in: Computer Aided Design in Composite Materials Technology HI, Proc. 3rd Int. Conf. on Computer Aided Design in Composite Material Technology (CADCOMP 92). University of Delaware, Newark, Delaware, pp. 571-579. Eduljee, R. R, McCullough, R. L., Gillespie, J. W, Jr. (1992a), Polym. Eng. Sci., in press. Eduljee, R. R, McCullough, R. L., Gillespie, J. W, Jr. (1992b), Compos. Sci. Technol., in press. Eshelby, J. D. (1957), Proc. R. Soc. London A 241, 376.
9.8 References
Eshelby, J. D. (1961), in: Progress in Soil Mechanics, Vol. 2: Sheddon, N., Hill, R. (Eds.). Amsterdam: North-Holland, Chap. III. Foye, R. L. (1966 a), 10th National Symp. and Exhibit., San Diego, SAMPE 10, G-31. Foye, R. L. (1966 b), Quarterly Progress Reports Nos. 1, 2, AFML Cont. No. AF33/615-5150. Foye, R. L. (1972), /. Compos. Mater. 6, 193. Goldstein, H. (1957), Classical Mechanics. Reading, MA: Addison-Wesley. Halpin, J. C , Kardos, J. L. (1976), Polym. Eng. Sci. 16, 344. Halpin, J. C , Pagano, N. J. (1969), J. Compos. Mater. 3, 720. Halpin, J. C , Tsai, S. W. (June 1969), Air Force Technical Report AFML-TR-67-423. Hashin, Z. (1962), J. Appl Mech. 29, 143. Hashin, Z. (1965), 1 Mech. Phys. Solids 13, 119. Hashin, Z. (1983), /. Appl. Mech. 50, 481. Hashin, Z., Rosen, B. W. (1964), J. Appl. Mech., Trans. ASME 31, 233. Hashin, Z., Shtrikman, S. (1962a), J. Mech. Phys. Solids 10, 335. Hashin, Z., Shtrikman, S. (1962b), J. Mech. Phys. Solids 10, 343. Hashin, Z., Shtrikman, S. (1963), /. Mech. Phys. Solids 11, 127. Hermans, J. J. (1967), Proc. K. Ned. Akad. Wet. B 70(1), 1. Hermans, J. I, Hermans, P. H., Vermaas, D., Weidinger, A. (1946), Reel. Trav. Chim. Pays-Bas 65, 427. Hershey, A. V. (1954), /. Appl. Mech. 21, 236. Hill, R. J. (1963), /. Mech. Phys. Solids 11, 357. Hill, R. I (1964), /. Mech. Phys. Solids 12, 199. Hill, R. J. (1965 a), J. Mech. Phys. Solids 13, 189. Hill, R. J. (1965b), /. Mech. Phys. Solids 13, 213. Kacir, L., Ishai, O., Narkis, M. (1978 a), Polym. Eng. Sci. 18, 45. Kacir, L., Narkis, M., Ishai, O. (1978 b), Polym. Eng. Sci. 18, 234. Kardos, J. L. (1973), CRC Crit. Rev. Solid State Sci. 3, 419. Kelly, A., Tyson, W. (1965), /. Mech. Phys. Solids 13, 329. Kerner, E. H. (1956), Proc. Phys. Soc. London 69 B, 808. Konicek, T. S. (1987), MS Thesis, Department of Mechanical and Industrial Engineering, University of Illinois, Urbana, IL. Korringa, J. (1973), /. Math. Phys. 14, 509. Kroner, E. (1958), Z. Phys. 151, 104. Lewis, T. B., Nielsen, J. (1970), J. Appl. Polym. Sci. 14, 1449. McCullough, R. L. (1977), in: Treatise on Materials Science and Technology, Vol. 10, Part B, Properties of Solid Polymeric Materials. New York: Academic.
473
McCullough, R. L., Jarzebski, G. J., McGee, S. H. (1983), in: The Role of Polymer Matrix in the Processing and Structural Properties of Composite Materials: Seferis, J. C , Nicolais, L. (Eds.). New York: Plenum, p. 261. McGee, S. H., McCullough, R. L. (1981), Polym. Compos. 2, 149. Milton, G. W, Kohn, R. V. (1988), /. Mech. Phys. Solids 36, 597. Munson-McGee, S. H., McCullough, R. L. (1992), Polym. Eng. Sci., in press. Nomura, S., Chou, T-W. (1984), J. Appl. Mech. 51, 540. Paul, B. (1960), Trans. Metall. Soc. AIME 218, 36. Piggot, M. R. (1980), Load Bearing Fibre Composite. New York: Pergamon. Reuss, A. Z. (1929), Angew. Math. Mech. 9, 49. Rosen, B. W. (1964), Fiber Composite Materials. Papers presented at a Seminar of the American Society for Metals. Russel, W. B. (1973), /. Appl. Math. Phys. 24, 581. Russel, W. B., Acrivos, A. (1972), J. Appl. Math. Phys. 23, 434. Shaffer, B. W. (1964), AIAA J. 2, 348. Smith, J. C. (1974a), /. Res. Natl. Bur. Stand. 78A, 335. Smith, J. C. (1974b), / Res. Natl. Bur. Stand. 79A, 419. Smith, J. C. (1976), J. Res. Natl. Bur. Stand. 80 A, 45. Torquato, S. (1991), Appl. Mech. Rev. 44, 37. van der Poel, C. (1958), Rheol. Ada 1, 198. Voigt, W. (1910), Lehrbuch der Kristallphysik. Leipzig: Teubner. Walpole, L. J. (1966 a), J. Mech. Phys. Solids 14, 151. Walpole, L. I (1966 b), J. Mech. Phys. Solids 14, 289. Walpole, L. J. (1969), J. Mech. Phys. Solids 17, 235. Whitney, J. M. (1967), J. Compos. Mater. 1, 88. Whitney, J. M., Drzal, L. T. (1987), in: Toughened Composites, ASTM STP 937: Johnston, N. J. (Ed.). Philadelphia, PA: ASTM, p. 179. Whitney, J. M., Riley, M. B. (1966), AIAA J. 4, 1537. Wu, T. T. (1966), Int. J. Solids Struct. 2, 1. Wu, C. D., McCullough, R. L. (1977), in: Developments in Composite Materials: Holister, G. S. (Ed.). London: Applied Science.
General Reading Broutman, L. J. (Ed.) (1974), Fracture and Fatigue. New York: Academic Press. Chamis, C. C. (Ed.) (1974), Structural Design and Analysis, Parts I and II. New York: Academic Press. Chou, T-W, McCullough, R. L., Pipes, R. B. (1986), Scientific American 255(4), 192.
474
9 Elastic Properties of Composites
Hull,D. (1981), An Introduction to Composite Materials: Cahn, R. W, Thompson, M. W, Ward, I. M. (Eds.). Cambridge: Cambridge University Press. Jones, R. M. (1975), Mechanics of Composite Materials. Washington, DC: Scripta Book Co. Kelly, A., Macmillan, N. H. (1986), Strong Solids: 3rd ed. Oxford: Clarendon Press.
Lubin, G. (Ed.) (1982), Handbook of Composites. New York: Van Nostrand Reinhold. McCullough, R. L. (1971), Concepts of Fiber-Reinforced Composites. New York: Marcel Dekker, Inc. Tsai, S. W, Hahn, H. T. (1980), Introduction to Composite Materials. Westport, Connecticut: Technomic.
10 Inelastic Properties of Composites C. T. Sun School of Aeronautics and Astronautics, Purdue University, West Lafayette, IN, U.S.A.
List of Symbols and Abbreviations 10.1 Introduction 10.2 A 3-D Plasticity Model 10.3 Off-Axis Tension Test 10.4 One-Parameter Plasticity Model 10.5 AS4/PEEK Thermoplastic Composite 10.6 Metal-Matrix Composites 10.6.1 Boron/Aluminum 10.6.2 Titanium Matrix Composite 10.7 Overstress Viscoplasticity Model 10.8 Symmetric Laminates 10.9 Conclusion 10.10 Acknowledgements 10.11 References
Materials Science and Technology Copyright © WILEY-VCH Verlag GmbH & Co KGaA. Allrightsreserved.
476 478 479 481 482 482 485 485 486 488 490 492 492 492
476
10 Inelastic Properties of Composites
List of Symbols and Abbreviations atj a66 A d 44 Ex f(<Jij) F G 12 h H Hp k K m n {dN} [Qe] [<2ep] [Qep] Setj dWp Wp
coefficients that describe the amount of anisotropy parameter (function of temperature and strain rate) coefficient (function of temperature and strain rate) numerical value apparent elastic modulus of the off-axis yield function yield surface in-plane shear modulus laminate thickness effective overstress plastic modulus state variable viscoplastic parameter viscoplastic parameter coefficient (function of temperature and strain rate) incremental in-plane resultant force reduced elastic stiffness matrix elastic-plastic stiffness matrix transformed matrix of [Qep] elastic compliances increment of plastic work per unit volume rate of plastic work
y 12 sx dsfj dsp y de^ dsij deij de p ep etj ex 8P sex sp efj efj {de} {dep} {de0} {s1}
engineering shear strain elastic strain in loading direction plastic part of the incremental strain plastic strain increment in x, y-direction incremental strain in loading direction incremental strain elastic part of the incremental strain effective plastic strain increment total effective plastic strain increment total strain rate total axial strain rate axial plastic strain rate axial elastic strain rate effective plastic strain rate elastic part of the total strain rate plastic part of the total strain rate increment elastic strain (in matrix notation) incremental plastic strain (in matrix notation) incremental mid-plane strain vector initial strain
List of Symbols and Abbreviations
477
9 angle between x-axis (uniaxial loading direction) and fiber direction jx^-axis dA proportionality factor k proportionality factor v£ plastic Poisson ratio a effective stress ax applied uniaxial stress atj stresses that refer to the principal material direction {da} increment elastic stress (in matrix notation) {a1} initial stress
polyetheretherketone (PEEK) matrix reinforced by AS4 carbon fibers glass transition temperature titanium matrix reinforced by SCS 6 ceramic fibers
478
10 Inelastic Properties of Composites
10,1 Introduction Fiber composites exhibit physically nonlinear behavior. This nonlinearity is especially pronounced in metal-matrix and thermoplastic composites. A number of models have been used to describe the nonlinear stress-strain relationship. Hahn and Tsai (1973) employed a complementary elastic energy density function that contained a biquadratic term for in-plane shear stress. As a result, the strain-stress relation is linear in simple longitudinal and transverse extensions, and the shear strain is a third-order polynomial function in the shear stress, which is unsuitable for describing the nonlinear behavior exhibited by many composites. Assuming that fibers are linearly elastic, Sun et al. (1974) modeled the composite by employing nonlinear matrix layers alternating with effective linearly elastic fibrous layers. This approach is limited by the fact that the nonlinearity was obtained by including third order terms in strains in the stress-strain relations. The physical nonlinearity in fiber composites can also be considered a plastic behavior. Dvorak and Bahei-El-Din (1979, 1982) used a microchemical model consisting of elastic filaments of vanishing diameters and an elastic-plastic matrix. An elastic-plastic continuum model was developed for which the constitutive relations depend on the constituent properties, their volume fractions, and mutual constraints between the phases associated with the geometry of the microstructure. A few authors (Chamis and Sullivan, 1973; Adams, 1970; Foye, 1973; Lin et al., 1972) used finite element method to analyze a representative volume of a unidirectional composite to understand the elastic-plastic behavior of the composite.
Inherent in these micromechanical models is the need to know the constituent material properties, fiber array, and fiber matrix interfacial condition. As noted in Bahei-El-Din and Dvorak (1980), the insitu matrix property can be quite different from that of the bulk matrix material. Consequently, adjustment of the matrix property in the micromechanical model is often necessary to yield good results. Another difficulty encountered in formulating micromechanical models lies in the 3-D elastic constants of the fiber, which are extremely difficult to measure because of their small size. Recently, Johnson et al. (1988) indicated that the fiber-matrix interfacial bond in some metal-matrix composites is very weak as opposed to the perfect bond assumed in developing these micromechanical models. To overcome the aforementioned uncertainties present in the micromechanical level of the composite, an alternative approach is to regard the composite as a homogeneous elastic-plastic continuum and measure its macromechanical properties directly. This approach enables one to employ well documented classical plasticity theories and many existing finite element codes for structural analysis. For unidirectional composites, orthotropy exists in their plastic behavior under monotonic loading. Griffin et al. (1981) proposed a three-dimensional flow theory based on Hill's (1948) orthotropic yield criterion to describe the nonlinear behavior of a boron/epoxy composite. The present author and co-workers have recently employed a more general plastic potential function, which is a quadratic form of the stress components, to characterize the orthotropic plasticity in a boron-aluminum metal-matrix composite (Kenaga et al., 1987; Rizzi et al, 1987; Sun and J. K. Chen, 1987; Sun and J. L. Chen, 1989). A simpli-
10.2 A 3-D Plasticity Model
fied version of the plastic potential function with a single parameter proposed by Sun and Chen (1989) was successfully employed by Sun and Yoon (1989) to characterize the elastic-plastic behavior of AS4/ PEEK thermoplastic composite at various temperatures. This plasticity model is simple in form and can easily be obtained experimentally using off-axis specimens. Because of its simplicity, this testing procedure can easily be performed in elevated temperature environments. Many polymer matrix composites exhibit plastic behavior when they are loaded in the matrix dominant direction. A few models were proposed to describe the viscoplastic behavior of orthotropic materials. Most of them were modified from the classical viscoplasticity models of metallic materials, such as the overstress model (Malvern, 1951; Eisenberg and Yen, 1981) and Bodner and Partom's model (1975). The overstress concept was proposed by Malvern (1951) to explain the high strain rate effect on wave propagation in metallic bars. This overstress approach was discussed by Cristescu and Suliciu (1982) in several alternative forms of overstress function. Subsequently, Eisenberg and Yen (1981) incorporated an anisotropic hardening law in the overstress theory to describe the movement of the yield surface. Following a different approach, Bodner and Partorn (1975) developed a set of constitutive equations for elastic-viscoplastic behavior of isotropic materials which require neither a yield criterion nor the distinction between loading and unloading. The hypothesis of their model was partially motivated from dislocation dynamics which indicates that the dislocation velocity is a function of stress. To describe the viscoplastic effect in fiber reinforced composites, Gates and Sun (1989) and Yoon and Sun (1991) proposed
479
an elastic-viscoplastic constitutive model using a one-parameter plastic potential function in conjunction with the overstress concept. They verified the validity of the model using AS4/PEEK thermoplastic composite. Krempl and Hong (1989) introduced a simple lamination theory using an orthotropic viscoplastic theory based on overstress to describe the in-plane stressstrain relationship for metal matrix composites. In their model, at least 17 constants and two functions need to be determined experimentally. Additional viscoplasticity models include one by Ha and Springer (1989) and one by Robinson et al. (1987). Ha and Springer (1989) developed a rate-dependent model using viscoelastic and viscoplastic constitutive equations for each lamina. In their model, twenty-three material constants are needed for the viscoelastic part, and nine constants are needed for the viscoplastic part of the constitutive equation. The viscoplasticity model proposed by Robinson et al. (1987) involves the use of a potential function, flow law and evolutionary law. Four constants and one function need to be determined experimentally. The model has not been verified by experiments.
10.2 A 3-D Plasticity Model A yield criterion is assumed for the general 3-D fiber composite as
2a
480
10 Inelastic Properties of Composites
tions. The coefficients atj describe the amount of anisotropy in the inital plasticity. This expression for /(^y) satisfies the orthotropy condition. The values of atj can be determined from the experimental data. This criterion reduces to von Mises yield criterion for isotropic solids when the atj have the values
The effective plastic strain increment de p can be defined such that dWp = (Jrdsf-= ads9
(10-7)
Substitution of Eqs. (10-5) and (10-6) into (10-7) yields (10-8) and
= <*22 = %3 = 2/3
(10-2)
The Hill-type yield criterion for orthotropic materials (1948) is a special case of Eq. (10-1) when the function /(o^y) is assumed to be independent of the hydrostatic stress. Hill's yield function is obtained if = fl33 - « 1 1 - 0 22 2023=011 -022-033
(10-3)
2\da
Since the incremental strain de,-y is assumed to be small, we are able to decompose it linearly into the elastic part defy and the plastic part defy as
Consider a state of plane stress parallel to the x1-x2 plane. The plastic potential function reduces to = a11o1
= -^-dX
ij = 11,22,12
(10-4)
0(7:ij
where the superscript p denotes plasticity, and dX is a proportionality factor. In general, the yield function given by Eq. (10-1) implies neither the incompressibility of plastic strains nor the assumption that hydrostatic stresses result in no plastic deformation. The increment of plastic work per unit volume is given by dWp = Gijdefj = 2/d/l
(10-5)
Let the effective stress c be defined as (10-6)
(10-10)
e;y = defy + defy
2a 13 = 022-011 -033
By using the associated flow rule, the yield function is taken as the plastic potential function from which the incremental plastic strains can be derived as
(10-9)
) \ 5
<j +a22<j 22
22
(10-11) Without loss of generality, we further set 022 = 1-
The effective plastic strain increment can be derived using Eqs. (10-5)-(10-7) and the inversion of Eq. (10-3). It is explicitly written as
- 2a ±1
12
2a
(10-12) 66
From Eqs. (10-4) and (10-11), the plastic strain increments are obtained as (10-13) 12
1 0
o-
0 <0-22 2a 6 6 _ 1 ^12
481
10.3 Off-Axis Tension Test
10,3 Off-Axis Tension Test
Comparison of Eq. (10-19) with Eq. (10-8) leads to
The complete orthotropic plastic flow rule is defined if the parameters all9 a12, a66 and dX are determined. To determine dA, the effective stress-effective plastic strain relation must be established. This can be accomplished from the results of tension tests on off-axis specimens. Let the x-axis be the uniaxial loading direction which makes an angle 9 with the fiber direction xx-axis. The stresses referring to the material principal axes (xx and x2) are related to the applied uniaxial stress
dsp = dspx/h{9)
a 22 = sin2 9 ax
a = h{9)ax
(10-15)
where h(9) = , / f [alt cos4 9 + sin4 9+ 2
(10-16)
+ 2(a 12 + a 66 )sin 0cos 6>] 1/2
2
Using the coordinate transformation for strains, the plastic strain increment in the loading direction is given by 2 p dep = cos2 ede?, l i + sin 9ds ?? -
(10-17)
Substituting Eq. (10-13) into Eq. (10-17) and then using the relations (10-14), we obtain del = \h (9) ax dX
8p = spx/h(9)
(10-21)
The desired relation between a and i? can now be obtained from the experimentally obtained relation between ax and 8P. From Eqs. (10-15) and (10-21), it follows that ^L = h2(9)-^ Ui>
(10-22)
Ubx
(10-14)
Substitution of Eqs. (10-13) into Eqs. (10-10) and (10-5) yields
2
For monotonic loading, the above equation is integrable. Thus,
and
all = cos2 9 ax
— sin9cos9dyp12
(10-20)
(10-18)
In view of Eq. (10-15), we thus obtain (10-19)
Since the a-sp relation should be unique in monotonic loading for the given material, the parameters all9 a12 and a66 must be chosen so that the resulting a-sp relation is independent of 9. The above procedure can only determine the sum a12 + a 66 , as is evident from Eq. (10-16). To separate the two parameters, the plastic Poisson's ratio must also be used. Using the coordinate transformation law on strain components and Eqs. (10-13) and (10-14), we obtain the plastic strain increments in the x-direction and the y-direction as deP = [axl cos4 9 + sin 4 9 + (10-24) + 2(a12 +
a66)sin29cos29]dXax
and dep = [(1 + a±! - 2 a 66 ) sin2 9 cos2 9 + + a12 (sin4 9 + cos 4 9)] dX ax
(10-25)
respectively. The plastic Poisson's ratio is defined as
(1 +alt — 2 a 66 )sin 2 9 cos2 9 + a 12 (sin 4 9 + cos4 9) cos4 9 + sin 4 0 + 2(a 12 + a 66 )sin 2 9
(10-26)
482
10 Inelastic Properties of Composites
For 9 = 0°, 90°, and 45°, the plastic Poisson's ratios reduce to a12
-2(a12-a66) In the off-axis tension test, the transverse strain is measured, from which the transverse plastic strain e£ can be obtained, and, thus, the plastic Poisson's ratio. The additional data allow us to separate a12 and
10.4 One-Parameter Plasticity Model Numerous experimental data show that fiber composites behave linearly when loaded in the fiber direction. Thus, it is reasonable to assume that de?1=0
(10-28)
which leads to the condition fli
9 =
= [f (a\2 + 2a66a212)]1>2dX
respectively, and h(9) becomes (10-27)
# 1 1 =
and
(10-29)
'
As a result, we obtain a one-parameter plastic potential 2 / = a! 2 + 2fl66<7
2
(10-30)
2
From this plastic potential, the plastic strain increments are derived: (10-31) where y12 is an engineering shear strain. The corresponding effective stress and effective plastic strain increments reduce to 1/2
(10-32)
(10-33)
,.Q ^A\
h (9) = [f (sin4 9 + 2 a66 sin2 9 cos2 9)]1/2 This simplification results in a single parameter a66 to be determined. The procedure for determining the one-parameter plasticity model becomes very simple. In theory, only two off-axis specimens with different 0's are required to determine a66. Note that for 9 = 90°, h(9) = ^ 3 / 2 , and thus, the stress-strain relation for the 90°specimen is not affected by the choice of a66 and can be conveniently used to construct the master effective stress-effective plastic strain curve for the composite. Moreover, there is no need to measure the plastic Poisson's ratio to aid in determining the parameters as in the 3-parameter model. The plastic Poisson's ratio corresponding to the one-parameter plastic potential is given by (2a 6 6 -l)cos 2 6> sin2 9 + 2 a6fi cos2 9
(10-35)
As 9 -* 0°, the plastic Poisson's ratio approaches v0o = 1 - J -
(10-36)
-*66
10.5 AS4/PEEK Thermoplastic Composite The one-parameter plasticity model was employed by Sun and Yoon (1989) to characterize the plastic behavior of the AS4/ PEEK (APC-2) thermoplastic composite. Specimens with 9 = 0°, 15°, 30°, 45°, and 90° were tested. The stress-strain curves obtained at 24 °C (75 °F), 66 °C (150°F),
483
10.5 AS4/PEEK Thermoplastic Composite
121 °C (250°F)9 and 177 °C (350 °F) are presented in Figs. 10-1-10-4, respectively. Experimental results indicate that for the thermoplastic fiber composite there is no well defined yield point. In view of this, the master effective stress-effective plastic strain curve can be fitted as a power law, = A (a)n (
300 200 ' 30°
\
100
^ k
•
0
0.5 1.0 1.5 Strain in %
0° ,300
2.0
2.5
3.0
^z
"200
/
30°
!
M00 —•
-1.0 -0.5
0
2.0
-0.5 1.0 1.5 Strain in %
2.5
3.0
Figure 10-3. Stress-strain curves for AS4/PEEK offaxis specimens at 250 °F (121 °C). ( : Strain gage debonded but strength obtained.)
——Long.
0.5 1.0 1.5 Strain in %
2.0
/
0
,300
Moo
ID"
15°
\
n -1.0 -0.5
•
-Long.
Trans.-
^ 100 ^.30°
/
15°
45°
Figure 10-1. Stress-strain curves for AS4/PEEK offaxis specimens at 75 °F (24 °C). ( : Strain gage debonded but strength obtained.)
400
^200 .,15° (/)
:2oo
^90°
n -1.0 -0.5
O 300 Q_
0°
30°
/
45°^\ 1
/0°
400
1
0°
0°l
Trans. ^—
—— Long.
Trans.—
—^Long.
(10-37)
The parameter a 66 , coefficients A and n may be functions of temperature, strain rate and other environmental factors. The effective stress-effective plastic strain relations of all off-axis specimens at 24 °C (75 °F), 66 °C (150°F), 121 °C (250 °F), and
400
Trans.——
400
2.5
3.0
Figure 10-2. Stress-strain curves for AS4/PEEK offaxis specimens at 150°F (66 °C). ( : Strain gage debonded but strength obtained.)
—
^
^
15°.....
^
30°
\
0 45°, -1.0 -0.5
0.5 1.0 1.5 Strain in %
2.0
2.5
3.0
Figure 10-4. Stress-strain curves for AS4/PEEK offaxis specimens at 350 °F (177 °C). ( : Strain gage debonded but strength obtained.)
177 °C (350 °F) are shown in Fig. 10-5. It is seen that the effective stress-effective plastic strain data for all off-axis angles more or less collapse into one curve with a66 = 1.5 for all temperatures. This indicates that the orthotropy of plasticity in AS4/PEEK can be regarded as independent of temperature. The coefficients A and n can be determined from fitting the data into the power law of Eq. (10-37). The results, presented in Table 10-1, indicate that n = 7.0 is adequate for temperatures up to 121 °C (250 °F). It is
10 Inelastic Properties of Composites
75°F (24°C) 150°F (66°C) ._ 250°F |121°C) . 350°F (177°C)
Figure 10-5. Effective stress and effective plastic strain curves for AS4/PEEK offaxis specimens.
2.5
1.0 1.5 2.0 Effective plastic strain in %
3.0
Table 10-1. Coefficients A and n for AS4/PEEK. ^ ^ \ ^ ^ Temperature Coefficient
24 °C (75 °F)
66 °C (150°F)
121 °C (250°F)
177 °C (350 °F)
7.0 5.14xlO~ 18 -39.8
7.0 2.18 xlO~ 1 7 -38.4
7.0 1.70xl0~ 16 -36.3
2.9 7.06 xlO" 8 -13.5
^ ^ ^ ^
n A In (A)
also noted that coefficient A is related to temperature as
i
(0.036 T - 40.7 (Tin°C)
(1 38)
t
°
for temperatures below the glass transition point. At 177 °C (350 °F), which is above the T.G. (glass transition temperature), n = 2.9, and the value of A drops significantly. These data are presented in Fig. 10-6. The total incremental strain in the loading direction can be decomposed into the elastic part d&ex and the plastic part dex as dsx = del + dsx
In Eq. (10-40), sex is the elastic strain in the loading direction defined as
(10-39)
where Ex is the apparent elastic modulus of the off-axis specimen which can be obtained from the transform equation 1 _ 1
4
12
+ — sin 4 0 E~,
(10-42)
In the monotonic loading case Eq. (10-39) can be integrated to yield
Using Eqs. (10-15), (10-21) and (10-37), sx can be expressed as
(10-40)
(10-43)
X
X
I
485
10.6 Metal-Matrix Composites
Therefore, the stress-strain curve in monotonic loading can be predicted using the following equation.
1.0 Predicted Experimental • 75°F
o
'pO.8 u)
§0.6
sx = £ + eg = £ + [h (9)f ^x
+1
A < (10-44)
en in
^x
Figure 10-7 shows the total stress-strain curves for the off-axis specimens. Agreement between the experimental data and predictions according to the model is excellent. The value of a66 can be checked by the comparison of the plastic Poisson's ratios obtained from theory and experiment. Figure 10-8 shows the predicted plastic Pois-
° 02 Q_
0
15.0
30.0 45.0 60.0 8 in degrees
75.0
90.0
Figure 10-8. Measured and predicted plastic Poisson's ratios for AS4/PEEK.
son's ratio (with a66 = 1.5) and the experimental data up to 121 °C (250 °F). The present plasticity model seems to yield reasonably accurate plastic Poisson's ratios.
10.6 Metal-Matrix Composites
50
100
150
200
250
300
350
400
Temperature in °F
Figure 10-6. Logarithm of the coefficient A vs. temperature for AS4/PEEK composite.
Metal-matrix composites display pronounced stress-strain curves under off-axis loading. The one-parameter plasticity model was shown to be very successful in describing nonlinear behavior for boron/ aluminum by Sun and Chen (1989), and for SCS6/Ti-6-4 titanium matrix composite by Sun etal. (1990). It is important to note that in boron/aluminum the fiber/matrix interface bond is very good and the observed nonlinear behavior is originated from the matrix yielding. However, for the titanium matrix composite, the apparent plasticity is mostly due to fiber/matrix separation. 10.6.1 Boron/Aluminum
• 1.0
-.Measured Data Predicted Curve
1.5 2.0 Strain in %
2.5
3.0
Figure 10-7. Measured and predicted stress-strain curves for AS4/PEEK off-axis specimens at 75 °F (24 °C).
For a boron/aluminum composite, Sun and Chen (1989) found that a66 = 2.0 was able to bring all the off-axis plastic stressstrain curves together in the effective stress-effective plastic strain plot, see Fig. 10-9. This master curve was fitted into
486
10 Inelastic Properties of Composites
the power law given by Eq. (10-37) with the result: A=
J0.41 x 10 - 1 5 10.3 x l O " 1 0
a in MPa a in kpsi (10-45)
This master curve is also shown in Fig. 10-9. Using a66 = 2.0, the plastic Poisson's ratio vs. off-axis angle 6 is shown in Fig. 10-10. It is evident that experimental data fall on the predicted curve nicely except for the 10° off-axis case for which plasticity is not pronounced.
1
o
2 3 4 5 6 Effective plastic strain in %
7
Figure 10-9. Effective stress-effective plastic strain relation for boron/aluminum composite. Solid line is predicted with a66 = 2.0.
10.6.2 Titanium Matrix Composite Sun et al. (1990) investigated the nonlinear stress-strain behavior of a titanium matrix composite (SCS6/Ti-6-4 with 40% fiber volume) using off-axis specimens. The value of a66 was found to be 0.85. The master curve is shown in Fig. 10-11. This master curve can be fitted into the power law with
0
15
28
A=
0.134 xlO~ , a in MPa 0.325 x 10" 20 , a in kpsi
30 45 60 Fiber angle in degrees
75
90
Figure 10-10. Comparison of measured and predicted plastic Poisson's ratios for boron/aluminum.
(10-46) Although the off-axis plastic stress-strain curves do not collapse exactly into a single curve, the recovered total off-axis stressstrain curves match with the experimental curves very well, as shown in Fig. 10-12. Johnson et al. (1988) and Sun et al. (1990) found that in titanium matrix composites fiber could separate from the matrix under loading. Comparing the off-axis stressstrain curves for the SCS6/Ti-6-4 composite with those of the titanium matrix as shown in Fig. 10-13, it is evident that the apparent yielding is not caused by plasticity in the matrix. Sun etal. (1990) employed a micromechanical model to analyze the interfacial
100 • 90 *60 • 45 *30 0 15
0.6 0.5 0.4
50
r
0.3 0.2 0.1 n
0
0.1 0.2 0.3 0.4 Effective plastic strain in %
0.5
Figure 10-11. Effective stress-effective plastic strain relation for SCS 6/Ti-6-4 composite.
487
10.6 Metal-Matrix Composites
stresses. A fiber/matrix interfacial failure criterion was proposed for this composite as {oi2)2 + 2d66{GA12)2
= a2
(10-47)
where ai7 and oA7 are the interfacial stresses, and (10-48) as = 106.9 MPa (15.5 kpsi)
0.2
0.6
0.3 0.4 Strain in %
Figure 10-12. Comparison of the predicted and experimental off-axis stress-strain curves for SCS 6/Ti-6-4 composite.
/
°
Fiber- - /
/
/
15°
The values of d66 and
- • — Fiber-Matrix Separation
• 1.0
/ /
- * — Matrix Yielding
1.5
100 Matrix /
50
/
/
S 1.0
30° y
0
0.5
0
0.1
0.2 0.3 (U Strain in %
0.5
Figure 10-13. Comparison of Ti-6-4 matrix, SCS 6 fiber, and the composite off-axis stress-strain curves.
u
0
30 60 Fiber angle in degrees
90
Figure 10-14. Off-axis stresses corresponding to the onsets of fiber/matrix separation and matrix yielding in SCS 6/Ti-6-4 composite.
488
10 Inelastic Properties of Composites
10.7 Overstress Viscoplasticity Model The following development of an overstress viscoplasticity model follows that presented by Sun and Yoon (1991). Within the scope of small strain assumptions, the total strain rate is composed of an elastic part and a plastic part as
assuming that the effective stress-effective plastic strain relation is strain rate dependent. The effective plastic strain rate £p is assumed as a function of effective overstress if, i.e., (10-56) in which (10-57)
(10-49)
pP
The elastic components e^- are related to the stress rate as e^ = S^(JlV
(10-50)
where S£ are the elastic compliances. The plastic strain rate is derived from the associated flow rule as
where cr* is the quasistatic effective stress corresponding to vanishingly small strain rates. The overstress function
\(H/K)
1 o
if
<7>
if
(10-58)
where K and m are the viscoplasticity parameters. Consequently,
_ _„
(10-51) where k is a proportionality factor. The rate of plastic work can be expressed as
(10-59) if
<7<<J*
For off-axis loading in the x-direction, the total strain rate can be expressed as
The effective plastic strain rate ep is defined such that
sx = el +
Wp = alp
where the elastic and the plastic strain rates are given by
(10-53)
From the definition of a and Eqs. (10-51) to (10-53), we have (10-54) With Eqs. (10-52) and (10-53) we obtain l =
3
-j
(10-55)
The one-parameter plasticity model is combined with the overstress concept by
°x
fiP
(10-60)
(10-61)
and s =
(10-62)
respectively. In a monotonic loading with a constant strain rate, the effective stress is always higher than the quasistatic effective stress for a given strain. From Eqs. (10-59) through (10-62), the total axial strain rate
10.7 Overstress Viscoplasticity Model
can be expressed as (10-63)
If we let 1/m
(10-64) then the total strain rate is given by £x = — + P(
(10-65)
X
By solving the above nonlinear differential equation using a numerical program, the total stress-strain curves for different strain rates can be predicted. The above overstress-based viscoplasticity model was used by Yoon and Sun (1991) to describe the viscoplastic behavior of AS4/PEEK thermoplastic composite. From the stress-strain curves of off-axis specimens tested at a very slow strain rate of about 10" 6 s per second and a fast strain rate of about 10 ~ 3 e per second, the effective overstress-effective plastic strain rate relation can be established. The total
489
stress-strain curves of 15°, 30° and 45° offaxis coupon specimens for low and high strain rates at room temperature are shown in Fig. 10-15. The stress-strain relations obtained from the slow strain rate test were taken as quasistatic stress-strain relations. From the quasistatic curve and higher strain rate curve, the effective overstress and the effective plastic strain rate relationship for each off-axis angle are calculated and are shown in Fig. 10-16. Even though there is some scatter, the curves of all cases considered collapsed into one master curve which appears to be linear on the logarithmic plot. This means that the viscoplastic behavior of the composite can be characterized by the constant viscoplasticity parameters K and m. The parameters K and m for 24 °C (75 °F) were obtained by linear curve fitting on the log-log scale, and the resulting values are shown in Fig. 10-16. Thus, the overstress function H is given by Eq. (10-58) with m -0.37 K =196
(
MPa)
•.Experimental : Curve fitted with A and n 1.0 x10" 3 /sec. 1.0 *1(T 6 /sec.
15°
U*1Cr 3 /sec. l 3 0 o 1.0x1(T6/sec. / 0.7 x10"Vsec. 1.0*i0- 6 /sec.
1.2
1.6 2.0 Strain in %
7.U
2.8
Figure 10-15. Stress-strain curves for AS4/PEEK offaxis specimens for low and high strain rates at 75 °F (24 °C). 3.2
490
10 Inelastic Properties of Composites
£1000.0 2 ^ 100.0 in in O)
V>
10.8 Symmetric Laminates
•15° •45°
Elastic-plastic deformation in composite laminates can be formulated using classical laminated plate theory. Each lamina is assumed to be orthotropic elastic-plastic. In matrix notations, Eq. (10-10) can be written in the form
10.0
CD
Z 1.0
m 3 0.37
/C=196.A o f i f i = 1.5
o Qi
S
0.1 1 1010"' 10" 10" Effective plastic strain rate §[H) in s"1
{de} = {dse} + {de»}
Figure 10-16. Effective overstress vs. effective plastic strain rate at 75 °F (24 °C).
It is interesting to note that the value of a66 is the same as in the plasticity model. By solving the nonlinear differential equation (10-65) with the obtained viscoplasticity parameters, the strain rate-dependent stress-strain curve can be predicted. Figure 10-17 shows the total stress-strain curves of different strain rates, experimental and predicted, for off-axis specimens at 24 °C (75 °F). We can see good agreement between experimental data and predicted curves.
(10-66) T
where {ds} = {da 11 ,d£ 22 ,d'y 12 } . The incremental elastic stress-strain relations are given by (10_6?) where [Qe] is the reduced elastic stiffness matrix (Jones, 1975). To obtain the incremental stress-strain relations, we start from the yield surface defined as
= 2f(aij)-k(s») =
(10-68)
The consistency condition which imposes the restriction on the increment between d<Tij and dfc at work-hardening can be expressed as
dF(aipk) = 0
(10-69)
500 •
A : Experimental : Predicted with m and K
400 1.0*i0" 3 /sec. l 1 5 o 1.Q*10"6/sec. /
:AA
0.5
1.0
1.5 Strain in %
1.0*10-6/sec. 0.7 *10-V see 1.0 * 1 0"6/sec.
2.0
2.5
Figure 10-17. Measured and predicted (overstress model) stress-strain curves of AS4/PEEK off-axis specimens for different strain rates at 75 °F (24 °C). 3.0
10.8 Symmetric Laminates
Considering that the current stress components satisfy the yield criterion at the plastic stage, and substituting yield function F into Eq. (10-69), we obtain
491
where the elastic-plastic stiffness matrix is defined as [Qep] = [Qe] ~
do-70)
(10-77)
[Ge]
Since k = 2 / = 2
= lJL=
Hpa
(10-71)
where H p is the plastic modulus, i.e.,
Substitution of Eq. (10-67) into Eq. (10-70) yields 8/
2
Following the formulation of classical laminated plate theory, the incremental stress-strain relations given by Eq. (10-76) can be used to derive the relations between the plate incremental resultant forces and the mid-plane displacements. By using the proper coordinate transformation, this formulation can be achieved with an arbitrary coordinate system x-y. For a symmetric laminate, we have (10-78)
(10-73) .dAL
Using Eq. (10-8), we obtain from Eq. (10-73) the following. 074)
in which {dN} is the incremental in-plane resultant force vector, {de0} is the incremental mid-plane strain vector, and fc/2
8/ V
8/
Therefore, the incremental plastic strain can be expressed as {d e p } = dX
8/
(10-75)
[A]=
f [<2 ep ]dz
(10-79)
-fc/2
where h is the laminate thickness, and [Qcp] is the transformed matrix of [Qep]. The total stresses and strains can be obtained by accumulating stress and strain increments as (10-80)
8/ Y
(10-81)
8/
9^
P
8/
^V9W V f f i l ^
From Eqs. (10-67) and (10-75), the relations between stress increment and strain increment can be expressed as {da} = [Q*"] {de}
(10-76)
where {CJ1} and {e1}, respectively, are initial stresses and strains such as thermal residual stresses and strains. When residual stresses are present, it is necessary to include them in the plastic analysis. Using the above procedure, Sun and Yoon (1990) investigated the elastic-plastic
492
10 Inelastic Properties of Composites
behavior of an AS4/PEEK [± 45] 3s laminate at room temperature. Because of its high processing temperature (380 °C), the laminate contained significant thermal residual stresses at room temperature. The residual stresses in the laminas were obtained by measuring the warpage in a [04/904] unsymmetric laminate from which the effective stress-free temperature was estimated to be 252 °C. The thermal residual stresses were calculated using the elastic lamination theory. The result was ai11 = - 5 8 . 3 M P a
MPa
(10-82)
Figure 10-18 presents the comparison of the measured and predicted strains in the [±45] 3s AS4/PEEK laminate. The solid line is the prediction that includes thermal residual stresses, and the dotted line is the prediction that omits thermal residual stresses. It is evident that thermal residual stresses cannot be neglected in the plastic analysis of laminates.
10.9 Conclusion Thermoplastic and metal matrix composites display significant nonlinear inelastic behaviors. These orthotropic nonlinear stress-strain relations can be modeled very well by an orthotropic plasticity model which contains a single orthotropy parameter. The simplification of the model is achieved by assuming that plasticity in the longitudinal (fiber) direction is negligible. The one-parameter plasticity model and the overstress viscoplasticity model were verified experimentally to be suitable for the description of plastic and viscoplastic behaviors in the AS4/PEEK thermoplastic composite and boron/aluminum metal matrix composite. For AS4/PEEK composite, the orthotropy parameter a66 and the power index n can be regarded as constants for temperatures below the glass transition temperature. For some metal matrix composites with a weak fiber/matrix interface, the apparent yielding could be caused by debonding between the fiber and the surrounding matrix.
10.10 Acknowledgements
300 Experimental data • Predicted with residual stresses • Predicted without residual stresses
250
This work was supported by NASA Langley Research Center under Grant No. NAG-1-825 and by the National Science Foundation under Grant CDR 8803017 to the Engineering Research Center for Intelligent Manufacturing Systems, Purdue University, West Lafayette, Indiana.
10.11 References 0
0.4
0.8 1.2 1.6 Longitudinal Strain in %
2.0
Figure 10-18. Uniaxial stress-strain curves for AS4/ PEEK [±45] 3s laminate at room temperature.
Adams, D. F. (1970), /. Comp. Mats. 4, 310-328. Bahei-El-Din, Y. A., Dvorak, G.I (1980), /. Appl. Mech. 47, 827-832. Bodner, S. R., Partom, Y. (1975), J. Appl. Mech. 42, 385-389.
10.11 References
Chamis, C. C , Sullivan, T. L. (1973), Theoretical and Experimental Investigation of the Nonlinear Behavior of Boron Aluminum Composites. NASA-Lewis Research Center, Cleveland, OH: NASA TM X-68-205. Cristescu, N., Suliciu, I. (1982), Viscoplasticity. The Hague, Boston, London: Martinus Nijhoff Publishers. Dvorak, G. J., Bahei-El-Din, Y. A. (1979), /. Mech. Physics Sols. 27, 51-72. Dvorak, G. J., Bahei-El-Din, Y. A. (1982), /. Appl. Mech. 49, 327-335. Eisenberg, M. A., Yen, C. F. (1981), J. Appl. Mech. 48, 276-284. Foye, R. L. (1973), /. Comp. Mats. 7, 178-193. Gates, T. S., Sun, C. T. (1989), in: Proc. 30th AIAA Struct., Struct. Dynamics and Mats. Conf, Mobile, AL, pp. 845-851. Griffin, O. H., Kamat, M. P., Herakovich, C. T. (1981), /. Comp. Mats. 5, 543-560. Ha, S. K., Springer, G. S. (1989), /. Comp. Mats. 23, 1130-1158. Hahn, H. T., Tsai, S. W. (1973), J. Comp. Mats. 7, 102-118. Hill, R. (1948), Proc. the Royal Society, Math, and Phys. Sci. 193, No. 1033, 281-297. Johnson, W. S., Lubowinski, S. X, Highsmith, A. L., Brewer, W. D., Hoogstraten, C. A. (1988), Mechanical Characterization of SCS6/Ti-15-3 Metal Matrix Composites at Room Temperature. NASALangley Research Center, Hampton, VA: NASP Technical Memorandum 1014. Jones, R. M. (1975), Mechanics of Composite Materials. New York: McGraw-Hill Company. Kenaga, D., Doyle, J. F., Sun, C. T. (1987), /. Comp. Mats. 21, 516-531. Krempl, E., Hong, B. Z. (1989), Comp. Sci. & Tech. 35, 53-74. Lin, T. H., Salinas, D., Ito, Y M. (1972), /. Appl. Mech. 39, 321-326. Malvern, L. E. (1951), /. Appl. Mech. 18, 203-208. Rizzi, S. A., Leewood, A. R., Doyle, X F , Sun, C. T. (1987), /. Comp. Mats. 21, 734-749.
493
Robinson, D. N., Duffy, S. F , Ellis, X R. (1987), A Viscoplastic Constitutive Theory for Metal Matrix Composites at High Temperature, in: Thermal Stress, Material Deformation and Thermo-Mechanical Fatigue, Vol. 23: Sehitoglu, H., Zamrik, S. Y. (Eds.). New York: ASME, PVP. Sun, C. T., Chen, X K. (1987), J. Comp. Mats. 21, 969-985. Sun, C. T., Chen, X L. (1989), /. Comp. Mats. 23, 1009-1020. Sun, C.T., Yoon, K. X (1989), in: Proc. 7th Int. Conf. on Comp. Mats. (ICCM VII), Vol. 2, November 22-24. Guangzhou, China, pp. 185-191. Sun, C. T., Yoon, K. X (1990), in: Proc. 5th Japan-US Conf. on Comp. Mats., June 24-27. Tama City, Tokyo, Japan. Sun, C. T, Feng, W. H., Koh, S. L. (1974), Int. J. Eng. Sci. 12, 919-935. Sun, C.T., Chen, XL., Sha, G. T., Koop, W E . (1990), J. Comp. Mats. 24, 1029-1059. Yoon, K.X, Sun, C. T. (1991), /. Comp. Mats. 25, 1277-1296.
General Reading Christensen, R. M. (1979), Mechanics of Composite Materials. New York: Wiley. Dvorak, G. X (Ed.) (1991), Inelastic Deformation of Composite Materials. New York: Springer. Hill, R. (1950), The Mathematical Theory of Plasticity. London: Oxford University Press. Johnson, W, Mellor, P. M. (1962), Plasticity for Mechanical Engineers. Princeton, NJ: Van Nostrand. Malvern, L. E. (1969), Introduction to the Mechanics of a Continuous Medium. Englewood Cliffs, NJ: Prentice-Hall, Chap. 6. Whitney, X M., Daniel, I. M., Pipes, B. R. (1982), Experimental Mechanics of Fiber Reinforced Composite Materials. Brookfield Center, CT: The Society for Experimental Stress Analysis.
11 Strength of Fiber Composites Ran Y. Kim Research Institute, University of Dayton, Dayton, OH, U.S.A.
List of Symbols and Abbreviations 11.1 Introduction 11.2 Strength Analysis of Unidirectional Laminates 11.2.1 Longitudinal Tensile Strength 11.2.2 Longitudinal Compressive Strength 11.2.3 Transverse Tensile Strength 11.2.4 Shear Strength 11.3 Strength Analysis of Multidirectional Laminates ,11.3.1 Failure Modes 11.3.2 Failure Theories 11.3.2.1 Maximum Stress Theory 11.3.2.2 Maximum Strain Theory 11.3.2.3 Maximum Distortional Energy Theory 11.3.2.4 Quadratic Theory 11.3.3 First Ply Failure 11.3.4 Ultimate Failure 11.3.5 Free Edge Effect 11.3.6 Notched Strength 11.4 Test Methods 11.4.1 Constituent Materials 11.4.1.1 Fibers 11.4.1.2 Matrix 11.4.1.3 Interfacial Bond or Fiber-Matrix Adhesion 11.4.2 Laminates 11.4.2.1 Tension 11.4.2.2 Compression 11.4.2.3 Shear Tests 11.4.3 Statistical Treatment of Data 11.5 Conclusions 11.6 References
Materials Science and Technology Copyright © WILEY-VCH Verlag GmbH & Co KGaA. All rights reserved.
496 498 498 498 501 502 503 505 505 506 506 507 507 507 509 510 512 514 516 516 516 519 519 520 520 522 524 528 530 531
496
11 Strength of Fiber Composites
List of Symbols and Abbreviations A cross-sectional area of a fiber Aij9 Qij reduced laminate and ply stiffnesses a0 distance ahead of a hole b beam width b, W specimen width c half crack length of a center crack d, t specimen thickness df fiber diameter E Young's modulus Ef m, £ fy , my fiber, matrix Young's modulus, in longitudinal and transverse directions Es shear modulus ej9 e] laminate mid-plane total strains, ply thermal strains measured from the stressfree state FhFtj strength parameters (failure theory) Fx (x) cumulative distribution function Fx* dimensionless parameter /0 amplitude of fiber deflection G shear modulus G m , Gx matrix, composite shear modulus h laminate or beam thickness / sectional modulus of beam / y , J* percentage point for confidence interval K stress concentration factor Kms K in shear Kmy matrix stress concentration factor KT finite plate stress concentration factor Kj infinite plate stress concentration factor KT/Kj finite width correction factor L gage length L half wavelength of fiber buckling L(a, P) likelihood function /c critical length Nt,N? applied laminate stress resultants, due to thermal strains p, P applied load at failure PT probability Critical critical buckling load R radius of a hole in an infinite orthotropic plate S, S' positive a n d negative shear strength in longitudinal direction Sj interface or interlaminar shear strength ^n matrix shear strength Us testing system displacement Ut total displacement Vfm fiber, matrix volume fraction
List of Symbols and Abbreviations
w
X,X' X
Y,Y' z Z
r 3 3 £
c,f,m
width Weibull distribution random variable tensile and compressive strength in longitudinal direction static strength (random variable) fiber tensile strength interface strength matrix tensile strength tensile and compressive strength in transverse direction thickness coordinate ultimate strength in z-direction shape, scale parameter estimate of a, fi values of the confidence interval for a, /? gamma function ineffective length deflection at the mid-point of the specimen composite, fiber, matrix strain axial, transverse strain strain at failure strain at failure in longitudinal and transverse directions stress partitioning parameter ply angle mean value Poisson's ratio = R/(R + a0) = c/(c + a0)
far field stress, along the x-axis stress components in the material axis fiber, matrix strength applied stress shear stress <jf m in transverse direction variance ply residual stresses notched strength of the infinite width laminate average fiber strength average interfacial shear strength fraction of undisturbed stress value FPF IITRI LEFM PEEK SACMA
first ply failure Illinois Institute of Technology Research Institute linear elastic fracture mechanics poly(ether ether ketone) Suppliers of Advanced Composites and Materials Association
497
498
11 Strength of Fiber Composites
11.1 Introduction In this chapter, representative theories for predicting the strength of unidirectional laminae and multidirectional laminates and test methods for determining ultimate strength of composites are discussed. The discussion deals with continuous fiber reinforced resin matrix composites. Unidirectional strengths are discussed on the basis of micromechanics theories for predicting longitudinal, transverse, and shear strengths including the effect of fiber-matrix interface on composite strength. Representative failure theories for laminates are discussed for predicting the first-ply failure as well as the ultimate failure in conjunction with the classical laminated plate theory and compared with experimental results for a variety of material systems and laminates. The effects of free-edge stresses and stress concentration around a notch are also discussed. In the last section, representative test methods for determining strength of constituents, unidirectional and multidirectional laminates are described, including statistical treatment of test data for the evaluation of design allowables and reliability analysis.
11.2 Strength Analysis of Unidirectional Laminates A unidirectional laminate consists of continuous fibers embedded in a resin matrix and is a basic building block of structural laminates. The fibers are bonded to the matrix at the interface. The fibers are very stiff and strong compared with the matrix. Thus, a unidirectional composite displays a large degree of anisotropy in its properties. The unidirectional composite is very strong in the fiber direction but is
weak in the direction perpendicular to the fibers. It is also important to note that longitudinal tensile response is mainly governed by the fiber properties, whereas the transverse and shear responses are strongly dependent upon the matrix and the interfacial bond. Prediction of unidirectional composites on the basis of constituent properties has not been as successful as prediction of their stiffness. Consequently, a number of micromechanics models for prediction of unidirectional strengths are available in the literature. An important role of the micromechanics analysis is to establish how these macroscopic properties can be controlled by the geometric and material properties of the constituents. In this section, longitudinal tensile and compressive strengths, transverse tensile and compressive strengths, and shear strength are discussed in continuous fiber reinforced resin matrix composites following a brief description of failure modes. 11.2.1 Longitudinal Tensile Strength
Fiber break and interface debonding are two major failure modes in a unidirectional composite in a longitudinal tensile loading state, and differ widely, depending on the properties of the constituent materials and fiber defects. In most advanced composites, such as boron and graphite fiber with polymer matrix, the strain to failure is greater in the matrix than in the fibers. Therefore, fibers which contain flaws or defects can break before the interface or matrix fails. Representative fracture surfaces of the unidirectional composites loaded under tension along the fiber direction are illustrated in Fig. 11-1 (Chamis, 1974). The crack created by a fiber break grows along a path varying with matrix, fiber-matrix interface, or adjacent fibers as the load increases. If bonding is strong, the
11.2 Strength Analysis of Unidirectional Laminates
Axial crack or ||/interf acial failure
Filament pullout
.Fiber break
499
where subscripts c, f, and m refer to composites, fiber and matrix, respectively. The use of force equilibrium along the fiber direction results in the following expression for composite longitudinal tensile strength, X (11-2)
Figure 11-1. Longitudinal tensile failure modes, (a) Brittle, (b) brittle with fiber pullout, (c) irregular (Chamis, 1967).
when the fiber controls the composite strength, and (11-3)
crack grows into the matrix resulting in a fairly smooth surface across the section. With a weak bond, the crack is more likely to lead to interfacial debonding and extensive fiber pullout. These failure mechanisms occur under static as well as fatigue loading. A variety of analytical models in the literature are available for evaluating the longitudinal tensile strength. We will present two of the most popular approaches, the rule-of-mixture model and the statistical model. The rule-of-mixture model is simple and reasonably accurate for estimation of composite strengths from the constituent properties. Furthermore for resin matrix composites reinforced with stiff fibers such as carbon and boron fibers, fiber strength and its volume fraction are only required for calculation of composite strength. This is very convenient for material development, quality control, material screening and selection, etc. This model also assumes that the constituents (fiber and matrix) carry loads proportional to their stiffness as schematically shown in Fig. 11-2; the composite fails when the average fiber strength is reached, and the ply is in a state of constant strain, i.e., 8 = 8
f
=
8n
(11-1)
when the matrix controls the composite strength. The symbols V and E are volume fraction and Young's modulus, respectively. Equations (11-2) and (11-3) are known as the rule of mixture equations. Although these equations are crude approximations to the longitudinal tensile strength of a unidirectional laminate, these equations display the significant contribution of the fiber strength to the ply strength. Tensile strengths of all resin matrix composites reinforced with carbon, boron and glass fibers are controlled by fibers, and
Fiber
Matrix
Strain
Figure 11-2. Typical stress-strain relations of fiber, matrix and composite. The composite failure strain is close to the fiber failure strain. The matrix is nonlinear above the fiber failure strain (Tsai and Hahn, 1980).
500
11 Strength of Fiber Composites
thus are calculated by Eq. (11-2). For the cases where EJEf <£ 1 Eq. (11-2) becomes X = Vt*f
(11-4)
Fiber strength can be determined by either the fiber bundle strength or the single filament strength. It is noted that the fiber bundle strength is generally smaller than that of the single filament strength. Experiments show that most composite materials exhibit large scatter in strengths. The degree of scatter may depend mainly on flaws or cracks in fibers which were formed in the course of fiber manufacturing as well as in composite processing. Since the size and distribution of the above defects are generally statistical in nature, many investigators adopted a statistical approach that can describe the prediction of the longitudinal strength. Rosen (1964) treats the composite failure based on the statistical distribution of flaws or imperfections in the fibers. The fibers containing critical flaws will fail first at the stress level below the average fiber strength. The accumulation of such fiber failures causes composite failure when the remaining unbroken fibers are unable to withstand the applied load. In this treatment of the tensile failure, Rosen introduced the ineffective fiber length concept. The ineffective length is defined as the portion of a broken fiber which is ineffective as reinforcement at the stress below the average stress of unbroken fibers. Outside this ineffective length, the broken fiber is still stressed equally with adjacent fibers due to the load transfer in the matrix. The most probable failure strength X for the composite was calculated by using Eq. (11-5) derived by Rosen (1964) X= V{(oLp5e)~1/p
(11-5)
where V{ is the fiber volume fraction, a and B are the statistical constants and are de-
termined by experiment for the different lengths of fibers, 3 is the ineffective length, and e is the base of natural logarithms. The statistical constants a and /? can be determined from test data on average strength as a function of fiber length. Assuming fiber strength is characterized by the Weibull distribution function, the average fiber strength is given by (11-6) where L is the gage length of the specimen and F is the gamma function. For a straight line log-log plot of <7f versus L, /? is estimated by log (L 2 /L t )
(11-7)
log((T fl /(J f2 )
Subscripts 1 and 2 are two stress levels, and a can be calculated from Eq. (11-6) after estimating /?. The ineffective length 3 is determined by a one-dimensional shear-lag theory. Rosen derived the following equation for the ineffective length, 3
V<1/2Gm cosh - 1
J
2(1-.
(11-8)
where df is fiber diameter, Vf fiber volume fraction, Ef fiber modulus, Gm matrix shear modulus, and cp the fraction of the undisturbed stress value below which the fiber is considered to be ineffective. The determination of q> value can be found in the reference: Rosen (1964). These ineffective lengths are relatively small. The predicted composite failure stress is plotted in Fig. 11-3 for the range of ineffective lengths of 1 to 100 fiber diameters. Experimental investigation shows that the typical failure of the high fiber content
11.2 Strength Analysis of Unidirectional Laminates 10 000
5 000
1000 10 Ineffective Length Ratio, 6/df
100
Figure 11-3. Statistical model of glass fiber-plastic composite failure stress (Rosen, 1964).
composite closely resembles the suggested failure model, characterized by random fracture of the fibers prior to ultimate failure of the composite. This model predicts considerably higher strength than the experimental data obtained from E-glass/ epoxy (Chamis, 1967). 11.2.2 Longitudinal Compressive Strength
In longitudinal compression, unlike tension, the matrix plays a significant role in compressive strength. The matrix provides lateral support for the fibers to carry compressive load without buckling. Without such matrix support fibers cannot resist compressive load. Fiber microbuckling and shear failure are considered to be two major failure mechanisms in longitudinal compression. There is also constituent debonding followed by fiber microbuckling, and ply separation or vertical splitting by transverse tension. Fiber microbuckling with an elastic matrix usually occurs in composites with low fiber volume content. In most composites in which the fiber volume fraction is greater than 0.5, fiber microbuckling, matrix yielding and constituent debonding occur.
501
Experimental investigation indicates that in most cases the microbuckling failure occurs at a considerably lower stress than the anticipated compressive strength. The compressive strength with shear failure mode approaches that for longitudinal tension in graphite, boron and s-glass fiber reinforced composites. Rosen (1964) proposed an analytical model for compressive strength of fibrous composites based on the microbuckling mechanism. Two possible failure modes considered in this analysis are the extension mode and shear mode. In the extension mode the fibers may buckle in opposite directions (symmetric) in adjacent fibers, and the major deformation of the matrix is an extension in the direction perpendicular to the fibers. In the shear mode the fibers buckle in phase with one another, so that the matrix deformation between fibers is due primarily to a shear. The condition for instability is given by equating the strain energy change to the work done by the external loads during buckling. The results for the compressive strength, X\ for the extension mode are given by X' = 2 F f {[F f £ m £ f /3(l - V,)]}1'2
(11-9)
The result for the shear mode is given by ' = Gm(l-7f)
(11-10)
As shown in Eqs. (11-9) and (11-10), the matrix shear modulus is an important factor in compressive strength. Hahn also derived a similar expression that the upper bound of compressive strength becomes equal to the effective shear modulus of the composite (Hahn and Williams, 1986). He notes that a correction should be made for the nonlinearity in shear of a unidirectional composite in order to improve prediction. The compressive failure of the composite is caused by a local shear failure,
502
11 Strength of Fiber Composites
cussed earlier, internal defects, such as voids, can also have an important effect on transverse strength property. In transverse tension or compression the load sharing by the matrix is of the same order of magnitude as that of the fiber. Since the matrix is much weaker than the fiber, the matrix will fail first and lead to the failure of the composite. The composite stress is related to the average stress in the constituents, and the transverse strength is given by
and the resulting compressive strength is X' =
(11-11)
1+(TI/0/L)/(S/GX)
where L is half wavelength of fiber buckling, / 0 the amplitude of fiber deflection, S the shear strength of the composite, and Gx the shear modulus of the composite. The compressive strength also can be calculated by the rule-of-mixture method as shown in tension in a previous section. Although the rule-of-mixture equation is a rough approximation for the longitudinal compressive strength, some experimental data indicate that the equation might predict reasonable results.
Y=Vfafy+Vm
(11-12)
where subscript "y" denotes the transverse direction. The transverse matrix strength and transverse fiber strength are related to the stress partitioning parameter, rjy as follows (Tsai and Hahn, 1980)
11.2.3 Transverse Tensile Strength
Depending upon the matrix properties and interface bond, the unidirectional composites would fail in the following modes under transverse tensile load as illustrated in Fig. 11-4: (1) matrix tensile failure, and (2) mixture of matrix tensile failure and fiber-matrix debonding. For the case where bond strength and fiber transverse strength are high, the failure surface tends to run primarily through the matrix (Fig. 11-4 a). When the fiber-matrix bond strength is low, debonding occurs and the fracture surface includes many debonded resin surfaces (Fig. 11-4 b). When fiber transverse strength is low, and matrix and bond strength are high, transverse fiber failure may occur (Fig. 11-4 c). As dis-
J
fy
(11-13)
where 0 < rjy < 1. By substituting Eq. (11-13) into (11-12), the average transverse stress of the composite can be expressed as follows y = [l + Vf(l/rjy- l)]ermy
(11-14)
The stress amy in the matrix is not uniformly distributed, but its maximum value, (o-my)max, occurs at the fiber-matrix interface. Therefore when the (crmy)max reaches the matrix tensile strength, X m , at the interface, the failure occurs; that is, ((7my)max ~
(11-15)
I—^-e~^—i Matrix failure
Bond failure
Fiber failure
(a)
(b)
(c)
Figure 11-4. Typical transverse tensile failure modes, (a) Matrix failure, (b) bond failure, (c) fiber failure.
11.2 Strength Analysis of Unidirectional Laminates
It is noted that if the interface strength X{ is smaller than Xm, X{ should be used in place of Xm. Introducing a stress concentration factor Kmy defined by Kmy
=
((Jm
y)max
( 1 1 1 6 )
Tsai (1968) and Adams et al. (1967) used finite difference to solve the plane elasticity problem for obtaining the stress concentration factor. The results of stress concentration from the analysis are shown in Fig. 11-5 (Adams et al., 1967). This is because the fibers and matrix have different transverse stiffness properties. The results indicate that the transverse stress in the material is significantly higher than the applied transverse stress. The magnitude of the stress concentration factor increases as fiber volume increases. The results also show that the stress concentration factor increases as the ratio Ef/Emy increases. This occurs when fiber volume fraction Vl increases as shown in Fig. 11-5. Experimental results indicated that only qualitative agreement exists between experiment and theory. The major reasons for this discrepancy are as follows. The distribution of
o o 3.4 LL
c 3.0 - circular filaments o in a square array "o 2.6 S 2.2 o o 1.8
503
fiber in a real composite is greatly different from that of a regular square array. Because of the progressive nature of composite failure, local failure at some region of the specimen prior to the final failure will redistribute the stress, and the elastic analysis is no longer valid. Nevertheless, analysis of this type provides a valuable insight into factors affecting composite strength. By combining Eqs. (11-14), (11-15) and (11-16), we obtain the transverse tensile strength of the composite, Y K my
(11-17)
An exact determination of rjy is difficult mainly because the behavior of the matrix is nonlinear near failure. If the matrix is linear elastic up to failure, then the factor [1 + Vf(l/rjy - i)]/Kmy is known to be less than unity and decreases with increasing fiber volume fraction. Therefore, the transverse tensile strength will always be less than the matrix tensile strength, and the difference increases with V{. On the other hand, if Kmy becomes close to unity, Y can be greater than the matrix tensile strength. Figure 11-6 shows the experimental evidence for the observation just described (Tsai and Hahn, 1980). The fracture surface in transverse tension is normal to the loading as shown in Fig. 11-4. However, in transverse compression, it is approximately 45° to the loading. The transverse compressive strength in graphite/epoxy composites is four to six times as high as the transverse tensile strength for advanced composites.
S 1.4 in 1.0 1
2 4 6 810 20 4060100 200 400 1000 Constituent Stiffness Ratio, Ef/Em
Figure 11-5. Matrix transverse stress concentration factor for various composite systems (Adams et al., 1967).
11.2.4 Shear Strength
A composite could fail in the following modes under in-plane shear stress: (a) matrix shear failure, (b) constituent debond-
504
11 Strength of Fiber Composites
pears to be the interfacial bond strength. In other words, the interlaminar strength can be increased only by improving interfacial bond strength. Voids can also have a considerable effect on the shear strength as illustrated in Fig. 11-9 (Brelant and Petker, 1970). The voids cause severe internal stress concentration in the material.
Figure 11-6. The composite-to-matrix strength ratio in transverse tension decreases as the matrix strength increases. Note that the composite strength can be higher than the matrix strength. The data are for E-glass and S-glass/epoxy composites (Tsai and Hahn, 1980).
ing, and (c) matrix shear failure and constituent debonding. The mechanisms of shear failure are similar to those of transverse tension failure. The shear strength S can be studied using an equation similar to Eq. (11-14)
- circular filaments in a square array
2
4 6 810 20 4060100 200 400 1000 Shear Modulus Ratio, Gx/Gm
Figure 11-7. Stress concentration factor subjected to in-plane shear loading (Adams et al., 1967).
(11-18) IDU "
D Q_
B
a
I
C
£ 100-
B / B
cn c (U
B jL
(/)
J"
g 50-
0 0
B
B
B^/H
B/
Resin
where Sm and Kms are the matrix shear strength and the matrix stress concentration factor in shear, respectively. Analytical results (Adams et al., 1967) on the stress concentration factor shown in Fig. 11-7 are similar to the transverse case discussed in the previous section. Again, if the interface shear strength Sj is smaller than Sm, Sj should be used in place of Sm. The significance of resin strength on interlaminar shear strength is illustrated in Fig. 11-8 (Brelant and Petker, 1970). Interlaminar strength increases as resin tensile strength until about 80 MPa and levels off afterward. The limitation factor for this ap-
B
20 40 60 Interlaminar Shear Strength in MPa
80
Figure 11-8. Interlaminar shear strength versus resin tensile strength (Brelant and Petker, 1970).
11.3 Strength Analysis of Multidirectional Laminates
505
tion. The analytical results are compared with experimental results obtained from a number of laminates. The effect of delamination and notch on the ultimate strength of laminates is given. 11.3.1 Failure Modes
5
10 Void Content in %
Figure 11-9. Effect of void on interlaminar shear strength (Brelant and Petker, 1970).
11.3 Strength Analysis of Multidirectional Laminates Most composite laminates for structural application consist of multidirectional layers to meet the structural integrity requirement. In-plane strength analysis of a multidirectional laminate is of great importance in design of the laminates as well as development of material systems. In this section we first describe briefly the damage modes, in terms of transverse crack and delamination for selected laminates, and present some of the failure theories currently employed in composite strength analysis. The prediction of the first-ply failure and the ultimate tensile strength of multidirectional laminates is described as an engineering approxima-
Fiber reinforced polymer matrix composites such as graphite/epoxy, boron/epoxy, and glass/epoxy exhibit complex damage mechanisms under static and fatigue loading because of anisotropic characteristics in their strength and stiffness. Instead of a predominant single crack, often observed in most isotropic brittle materials, extensive damage throughout the specimen usually accompanies static and fatigue failure in composites. Basic failure mechanisms in a multidirectional laminate are matrix cracking, delamination, fiber breakage, and interfacial debonding. Any combination of these causes damage, which may result in reduction of strength and stiffness. Damage varies widely, depending on material properties, lamination (including stacking sequence), type of loading, etc. Unidirectional laminates with the fiber along the loading direction do not show any matrix cracking in the polymer matrix composite, since the strain to failure for the matrix is much larger than that of the fiber (Fig. 11-2). However matrix cracking occurs in multidirectional laminates prior to ultimate failure. To illustrate matrix cracking, consider the matrix of a [0/90/±45] s graphite/epoxy laminate subjected to uniaxial tension. Successive transverse cracks in the respective off-axis layer are expected as the applied load to the laminate increases. The first cracking occurs in the 90° layers; with greater load, the number of cracks increases and is initially confined to the 90° layers. As the load increases, new
506
11 Strength of Fiber Composites
cracks occur in the adjacent 45° layers, appearing at the tip of the 90° cracks and extending to the interface of the + 45°/-45° layers. Subsequently, the number of cracks increases with the load level until final laminate failure (Kim, 1980). However, some laminates reach a crack density limit, after which no new cracks occur before final failure, despite additional loading (Reifsnider et al., 1979). The crack-density limit for a given layer varies with its thickness and constraints but appears to be independent of laminate type (Wang et al., 1980; Garret and Bailey, 1977; Flaggs and Kural, 1982; Reifsnider et al., 1979). The 0° layers in a multidirectional laminate are also susceptible to cracking along the fiber. In composites, free-edge delamination is caused by interlaminar stresses localized around the free edge under in-plane loading (Pagano, 1978). Among the interlaminar stresses, tensile normal stress is a significant factor in most delaminations (Pagano, 1978; Kim and Soni, 1984). The laminate stacking sequence determines whether interlaminar normal stress produces tension or compression at the free edges. For example, a quasi-isotropic laminate with two stacking sequences, [0/90/±45]s and [0/±45/90] s , is subjected to uniaxial tensile loading. The latter stacking sequence produces tensile normal stress along the free edges of a coupon, whereas the former produces compression. Consequently, the [0/±45/90] s laminate shows extensive delamination under tension before final failure but shows no delamination under compression. In addition to interlaminar tensile stress, other mechanisms such as transverse cracking and interlaminar shear stresses appear to be significant in the onset and growth of delamination under fatigue loading (in some cases).
11.3.2 Failure Theories
There exist a number of failure theories available in composite strength analysis, although it is important to have a unique failure theory that can be predict the strength of laminated composites accurately. We will discuss failure theories that are widely used in composite strength prediction. 11.3.2.1 Maximum Stress Theory
Maximum stress theory assumes that failure of the structural element subjected to a combined state of stress occurs when loading has reached such a value that one of the principal stresses becomes equal to the uniaxial strength in tension or compression. If one of the following three principal stresses reaches the uniaxial strength, then the failure occurs: <jx = X , if (j x > 0 or
ay = X
=X'
if (Jx < 0
if
f
if
where X9X' = tensile and compressive strength in longitudinal direction, X Y' = tensile and compressive strength in transverse direction, and S, S" = positive and negative shear strength in longitudinal direction. From orthotropic symmetry condition, shear strength is independent of sign. There are five independent failure modes as shown in Eq. (11-19). Since failure processes in composite laminates are highly interacting among the stress components and more complicated than the simple mode, the agreement between test result and theory is poor for certain combinations of stresses. Figure 11-10 represents the maximum stress theory for two dimensional conditions expressed by Eq. (11-19).
11.3 Strength Analysis of Multidirectional Laminates
507
parameters. The relations between the strength parameters, F , . . . , AT, and actual strength values, X, Y and S, can be determined simply by applying individual uniaxial loadings, i.e.:
X'
Y'
Figure 11-10. Maximum stress criteria for an orthotropic ply in stress space.
G + H = 1/X\
F + H = 1/72,
F + G = 1/Z2,
2iV = l/S 2 ,
where Z is the ultimate strength in the z-direction. Consequently,
11.3.2.2 Maximum Strain Theory
1
This is very similar to the maximum stress theory. This theory assumes that failure of the structural element subjected to a combined state of stresses occurs when the maximum value of the principal strains equals the value of the strain at failure (e*) in uniaxial tension or compression. If one of the following conditions is met, then the failure occurs:
A
1 1
1
2F = 2 +
T
1 Z^
1
1
(11-22)
Y*~Y2
Assuming that Y = Z and plane stress conditions (cz = erxz =
X' or Y
X2 Y'
X2
xy
~
(11-23)
11.3.2.3 Maximum Distortional Energy Theory
Equation (11-23) is known as the Tsai-Hill failure criterion. A comparison with experimental results (and with the two previous failure criteria) is shown in Fig. 11-11 (Tsai, 1968). Clearly the Tsai-Hill failure criterion seems to agree very well with experimental results obtained for glass/epoxy.
Hill (1950) has formulated the von Mises criterion for orthotropic materials as
11.3.2.4 Quadratic Theory
(11-20)
S
F (
This failure criterion proposed by Tsai and Wu (1971), has been published in numerous reports and journal articles. The mathematical formulation of this model was originally presented by Tsai (1968). Details on experimental procedures, optimal experimental measurements, and mathematical formulations are given by Tsai and Wu (1971) and Wu (1972) and it is
508
11 Strength of Fiber Composites
To ensure that the failure surface will intercept each stress axis, the magnitude of the Fxy term is constrained by the following inequality F
F
—F
> i
(11-27)
In view of the difficulty in experimental determination of Fxy, Tsai introduced a dimensionless parameter (11-28)
30 45 60 Fiber Orientation, d
Figure 11-11. Comparison of maximum stress, maximum strain and Tsai-Hill failure criteria (Tsai, 1968).
recommended to refer to these in order to become more familiar with this model. Tsai-Wu (1971) failure theory for anisotropic materials in stress space is given by FtjWj + Fta^l
(11-24)
where Ftj and Ft are strength parameters associated with the failure theory, and o{ and a • are the stress components in the material axis. When the problem is limited to orthotropic materials under the plane stress state, Eq. (11-24) assumes the following form
Tsai assigned — 1/2 to the value of Fx* based on a generalization of the von Mises yield criterion for isotropic materials. Details can be found in: Tsai and Hahn (1980). The quadratic failure criterion takes into account the interaction among the stress components and offers a great convenience in operation. This failure criterion is widely accepted and is used in the composite community. Figure 11-12 shows the failure envelope constructed by the quadratic failure criterion in stress space. Strength tests on off-axis and angle-ply laminates offer useful information for the evaluation of the failure theory and failure mechanisms. Figures 11-13 and 11-14 show comparison of Tsai-Wu quadratic failure theory with experimental results obtained from the off-axis laminates and angle-ply laminates, respectively (Kim, 1981 a).
- 2 + 2Fxy<7x<7y + Fss<7s2 + L P ff + F a = 1 (11-25) + F,
Fxxa2x+FT
The strength parameters are determined from material strength and are given by 1 XX
yy
xx" I YY"
J
'
x
y
1 ss
ss"
J s
1
I
X
x'
1
l
y
r
l
l
s
(11-26) Figure 11-12. Failure envelope constructed by the quadratic failure criterion.
509
11.3 Strength Analysis of Multidirectional Laminates
Gr/Ep As/3501
1600
1 600
Gr/Ep As/3501 Angle ply
Off-axis
F12*=-0.5
F 12 =-0.5 1200
1200
800
800
400
400
.£
i
i
15 30 9 in Degree
15 30 6 in Degree o -400
-400
-800
-800
-1200
-1200
-1600L
-1600 L
45
60
75
90
Figure 11-13. Tensile and compressive strength of offaxis laminates as a function of fiber orientation.
Figure 11-14. Tensile and compressive strength of angle-ply laminates as a function of ply angles.
11.3.3 First Ply Failure
chanical stresses and curing residual stresses. The analytical model proposed by Pagano and Hahn (1976) can be used to calculate the residual stresses resulting from cure. The basic assumption of this model is that the material behavior is elastic. Thus the total ply stresses, af, are related to the applied laminate stress resultants, Ni9 and ply residual stresses, of, and given by Ot-QtjAjSNt + of (11-29)
Because of the inherent inhomogeneity, the ultimate failure of multidirectional composite laminates is invariably preceded by the failure of weaker plies. The first ply failure (FPF) is defined as the first occurrence of a crack within a laminate subjected to a loading. The applied stress at the moment of the FPF depends on the type of laminate, constituent ply properties, constraint by the adjacent plies, and residual stress present within the laminate. The first ply failure can be predicted by applying a failure theory in conjunction with calculation of stresses in the ply of interest. The stresses consist of the me-
where Qtj and Atj are reduced ply stiffnesses and laminate stiffnesses, respectively. Note that the superscript — 1 denotes matrix inversion.
510
11 Strength of Fiber Composites
The ply residual stresses, of, within a laminate can be calculated by (11-30)
of = QtJ(eJ - e]) where
ment for first ply failure. Wet specimens contain approximately 1.4% of moisture by weight. 11.3.4 Ultimate Failure
e°j = mid-plane total strains measured from the stress-free state, e] = ply thermal strains measured from the stress-free state. The laminate thermal strains e? then follow from the classical laminated plate theory as e°i=A^N?
(11-31)
N? are the equivalent stress resultants due to the thermal strains, i.e., -h/2
Que]dz
(11-32)
where h is the laminate thickness and z the thickness coordinate. Upon substitution of Eq. (11-31) into Eq. (11-32), the residual stresses of in each ply are obtained as follows of = QijA^1
JVfeT - Qtje]
(11-33)
Equation (11-33) gives the residual stress components in the ply of interest. By applying an appropriate failure theory with the total stresses calculated by Eq. (11-29), we can calculate the first ply failure stress for a laminate. Table 11-1 shows comparison of theory with experiTable 11-1. First ply-failure stresses in MPa (Kim
and Hahn, 1979). Dry
Laminate
Theory Experiment [0/903]s [±45/90 2 ] s [0/± 45/90],
119 79 153
159 73 174
Wet Theory Experiment 139 84 197
168 69 205
When the applied load exceeds the FPF load, the laminate may or may not carry the additional load. Most laminates, especially those containing the 0° ply under tension, have a capability to carry the load beyond the FPF and reach ultimate failure. The ultimate failure or last-ply failure will be examined in this section for applied tensile loading. Most laminates reveal a progressive type of failure as described in Sec. 11.3.1 when subjected to an applied stress. In view of this progressive failure in multidirectional laminates, a number of approaches have been reported to develop a reliable methodology to predict the in-plane tensile strength. These can be divided into two major approaches that are with and without considering the damage occurred prior to the last failure. The approaches with consideration of damage can be divided into again two groups, depending upon complete elimination of the damaged ply or employing damage functions in calculation of strength. One of the main obstacles in accepting the damage function approach in strength prediction is complexity in the calculation process. Therefore, success of this approach appears to remain in question until one has a proper understanding of the damage process and its influence on the strength change. In this section we will discuss the simple approach without considering the damage within the laminate as examples. Figures 11-15 and 11-16 show analytical prediction of the failure stress of the respective layers for the laminate family of [0 2 /± 0]s, [0/90/± 6]s and [0/+ 9/90]s as a function of
11.3 Strength Analysis of Multidirectional Laminates
T300/5208 Gr/Ep [O 2 /+0] s
uoo 1200
a
1000
Q_
•£ 800 en c
600 400 200 0
0
15
30 45 60 8 in Degree
75
90
Figure 11-15. Comparison of calculated in-plane tensile strength with experimental. The solid line and dotted line indicate the predicted strengths of 0° and 90° plies, respectively.
1000
T300/5208 Gr/Ep o[O/±0/9O]s MO/90/10],
800 600
200
0
15
30 45 60 6 in Degree
75
90
Figure 11-16. Comparison of calculated in-plane tensile strength with experiment. Predicted 90°, off-axis and 0° plies are shown for two stacking sequences.
angle 9 (Kim, 1981 b). The dotted line and solid line indicate the predicted strength for the first ply failure and the last ply failure, respectively. Various symbols in Figs. 11-15 and 11-16 are experimental results which represent the mean of 10 to 11 specimens for a given laminate. In the [02/ ± 0]s laminate the experimental results compare fairly well with the analytical results of the ultimate failure (0° layer failure
511
in this case). Analysis gives a slightly conservative prediction in all cases except for 0 = 10° and 15°. Analytical prediction shows almost identical strength for the 0° and off-axis layers up to approximately 30°, and thereafter the difference increases as angle 9 increases as shown in Fig. 11-15. This implies that the laminates for 9 < 30° will fail without revealing progressive failure; that is, no off-axis ply failure will precede the final failure. On the other hand, off-axis ply failure will be anticipated before final failure when 9 is greater than 30°. The theory also appears to underestimate the off-axis ply (± 9) failure for the family of the [0 2 /± 0]g laminate for 9 > 30°. In spite of the numerous cracks extending to the entire off-axis layers in the [O 2 /±0] s and [0 2 /± 0]s laminates, the experimentally-determined strengths are greater than those of prediction. This implies that the transverse cracks in the off-axis ply which is greater than 60° do not influence the tensile strength of this laminate group. Experimentally-determined strength is significantly different between two families of the [0/90/ ±9]s and [0/+ 0/9O]s laminates for 9 < 45° in spite of identical prediction as shown in Fig. 11-16. For those laminates which did not show any delamination until final failure, the experimental results compared fairly well with the analytical results of the 0° layer. Experimentally-determined strength for the laminates experienced in delamination is markedly smaller than the predicted strength. Quadratic failure theory in conjunction with laminated plate theory has been applied to predict the in-plane tensile strength of multidirectional composite laminates. In analytical prediction, the ultimate failure of a laminate is assumed to occur upon the failure of the strongest ply in the laminate by ignoring the contribution of interlaminar stresses and transverse cracks preced-
512
11 Strength of Fiber Composites
ing final failure. The experimental results obtained from various types of laminates of T300/5208 graphite/epoxy were found to be in good agreement with prediction with consistent exceptions. The exceptions are that all the laminates show delamination at the free edges, which substantially reduces the load-carrying capability. The experimental result indicates that the effect of transverse cracking on the strength appears to be insignificant for those laminates considered in this study. This approach has the advantage of simplicity and quickness in approximation of the in-plane tensile strength of multidirectional laminates. 11.3.5 Free Edge Effect Delamination has long been recognized as one of the important failure modes exhibited in composite laminates. The growth of delamination under load may result in a reduction of strength and stiffness of a laminate. Such a detrimental effect of delamination on laminates has been demonstrated in: O'Brian (1981), Kim (1983).
Analysis has shown the existence of interlaminar stresses along the free-edge region of composite laminates in the presence of an in-plane uniaxial loading (Pipes and Pagano, 1970; Pagano and Rybicki, 1974; Pagano and Soni, 1983). Figure 11-17 illustrates the state of stress at the free edge region of a laminate under uniaxial loading. Of the interlaminar stresses, tensile normal stress plays a significant role in initiation and propagation of delamination. The nature (tensile or compressive) and determination of the interlaminar normal stress has received extensive analysis. The strength reduction due to delamination varies, depending upon the delaminated area and type of loading for a given laminate. Table 11-2 shows the tensile strength of a quasi-isotropic laminate with four different stacking sequences. The first three stacking sequences underwent considerable delamination before ultimate failure, but the fourth stacking sequence, [0/90/± 45]s, failed without exhibiting delamination. The ultimate strength of the undelaminated [0/90/±45] s is greater than the strength of all three that had delami-
Free Edge Interface
Figure 11-17. Models for studying free-edge effects laminate geometry.
513
11.3 Strength Analysis of Multidirectional Laminates
Stacking sequence
Unreinforced
Reinforceda
Delamina- Ultimate tion thresh- strength old level (MPa) (MPa)
Ultimate strength
[0/±45/90]s 360 [±45/0/90]8 270 [+45/90/0], 370 [0/90/+45]s No delamination
446 431 498 573
(MPa) 577 572 576 585
No delamination occurred in reinforced specimens.
1 000 900
Prediction: Experiment; unreinforced A(O/9O/i0]s O[O/i0/9O]s reinforced #(O/±0/9O]s
\ O
800 o Q_ IE
700
JZ
\ \ A
_
C
6ue
nated. These specimens were also tested with a reinforced free edge to prevent delamination. None of the reinforced-specimen free edges exhibited delamination. The strength of the first three laminates with reinforcement was substantially increased and was practically equal to the strength of the [0/90/±45] s laminate. In other words, all the reinforced laminates show practically the same strength. Figure 11-18 shows the effect of delamination on the in-plane tensile strength of the [O/±0/9O]s and [0/90/±0] 8 laminates as a function of the angle 6 (Kim, 1981b). The solid line indicates the analytical prediction under the assumption that ultimate failure of the laminate occurs upon the strongest ply failure for a given externally applied stress resultant, ignoring the damages incurred in the course of loading. To construct the solid line, the in-plane stress components in each constituent ply for a given laminate were calculated using laminated plate theory. The quadratic failure criterion is then applied to the respective ply. In Fig. 11-18 the triangles and circles are the observed strengths for the [0/90/ ±0] s and [0/ + 6/90]s laminates, respectively. For 9 less than 45°, the [O/±0/9O]s laminates
revealed substantial delamination and failed at a much lower stress than the [0/90/ ±0] 8 laminates. However, with the free-edge reinforcement (solid circle in Fig. 11-18), the strengths increased substantially and were almost the same as those of the [0/90/±0] s laminate. It is also noted that for those laminates which did not delaminate until final failure, the experimenal results agree fairly well with the prediction. The strength reduction appears to be more pronounced in those laminates that do not contain 0° plies. Figure 11-19 shows the predicted strain response of the boron/epoxy [ ± 30]s angle-ply laminate along with the experimentally observed response (Pipes et al., 1973). The prediction of the strain response agrees quite well with the observed response, but the observed strength was approximately onehalf of the predicted strength. This discrepancy between prediction and observation of the strength is mainly attributed to the presence of a high interlaminar shear
Str
Table 11-2. Tensile strength of quasi-isotropic laminate with four different stacking sequences.
O
600
-
500
-
#
\ \ \
1 0 \a
400
a
^
^
O
300 1
15
1
1
30 45 60 6 in degrees
1
I
75
90
Figure 11-18. Comparison of prediction and experiment for ultimate failure for quasi-isotropic laminates of T300 graphite/5208 epoxy laminates.
514
11 Strength of Fiber Composites
NARMCO 5505 B0R0N-EP0XY 130° COUPON Predicted Experimental
Figure 11-19. Comparison of prediction and experiment for strength and strain for the [±30] s laminate of boron/epoxy (Pipes etal., 1973).
0.0077 I i
0.002
0.004
0.006 0.008 0.010 c x , Strain
0.012
0.0U
stress at the +30/ —30 interface. A similar strength reduction was found for the graphite/epoxy [±30/90] s laminate (Kim, 1986). The effect of delamination can be much more severe under compressive loading than under tensile loading. Buckling is the major failure mode of a delaminated laminate under applied compression. When a laminated composite with delamination is subjected to compressive loads, the delaminated region buckles, as illustrated in Fig. 11-20 (Whitcomb, 1984). This buckling causes a high interlaminar stress concentration at the delamination front (crack tip); and under increased load, the buckled area increases to a critical size which can lead to a loss of global stability or total collapse of the plate. This occurs usually at load levels far below the undamaged compressive strength of the material or stability of the plate. There are some
Figure 11-20. Illustration of through width buckling of a composite plate (Whitcomb, 1984).
0.016
analytical models available to predict the buckling growth and the effect of various parameters on the delamination growth (Chai, 1981; Whitcomb, 1984; Donaldson, 1987). The models are based on the fracture criterion or on the strain energy release rate. However, because of the complexity of the problem, due to factors such as geometric shape of the delaminated region, thickness of the buckled laminate, fracture criterion, the highly nonlinear character of deformation, type of laminate (including stacking sequence), one-dimensional or two-dimensional growth, multiple delamination, stiffness and toughness of the laminate, etc., the models cannot be considered to be validated at this time. 11.3.6 Notched Strength
The presence of a discontinuity or notch such as a hole or a crack in a laminated structure introduces stress concentration around the notch. This stress concentration can be high, which results in initial localized failure. Problems of stress concentration in an anisotropic plate are much more complicated than in an isotropic plate. We consider an anisotropic homogeneous plate of any shape which is weakened by a circular hole and deformed by
11.3 Strength Analysis of Multidirectional Laminates
Figure 11-21. Plate containing a circular hole subjected to an axial tension.
515
size effect in anisotropic materials. The problem of the notch size effect has been studied by many investigators. Two approaches used are concepts of linear elastic fracture mechanics (LEFM) (Waddaoups etal., 1971; Cruse, 1973) and stress criterion (Whitney and Nuismer, 1974). In advanced composites, the model developed by Whitney and Nuismer is widely used for predicting the strength of laminated composites containing a circular hole or straight crack using stress criteria rather than LEFM. Consider a hole of radius R in an infinite orthotropic plate, as shown in Fig. 11-21. If a uniform stress a is applied parallel to the x-axis at infinity, then the stress, o"x, along the y-axis can be approximated by (Patterson, 1974)
f - ,,,-34, forces acting on the edges as shown in Fig. 11-21. When the opening is small in comparison with the plate size, then the problem becomes simple. We can assume the plate as being infinite and will disregard the effect of the finite width. Unfortu-
? =1 +
where Kj is the orthotropic stress concentration factor for an infinite width plate as determined from the following relationship ^ 2 2 ~~
L
n
nately, in most cases, the plate width is finite and finite width correction is required. In anisotropic materials the problems for finite width have not been fully studied, and many investigators have been using the correction for the isotropic case, which may result in some errors. Unlike isotropic material, experiment shows that the fracture strength of notched anisotropic plates is dependent upon the notch sizes, even though the stress concentration at the edge of the notch is independent of the notch size. This is called notch
2A 66
(11-35)
where A(j are the in-plane laminate stiffness as determined from laminated plate theory. Subscripts 1 and 2 denote the directions parallel and perpendicular to the applied stress at infinity, respectively. The average stress criterion introduced by Whithney and Nuismer (1974) has been widely accepted and used for calculation of the notched strength of composite laminates. Figure 11-22 illustrates this failure theory. The average stress criterion assumes failure to occur when the average value of <JX over some distance, a 0 , ahead of
516
11 Strength of Fiber Composites
and circles are experimental data. In data reduction the notch strength for an infinite plate is obtained from the experimental notched strength aN using the relationship
d
Figure 11-22. Stress distribution around a circular hole.
where KT/Kj is finite width correction factor. Approximate expressions for these factors can be found in refs. (Patterson, 1974; Paris and Sih, 1965) and are given by = T/
the hole first reaches the unnotched tensile strength of the laminate; that is, when \
-
(11-39)
T
2 + (l-2R/w)3 3(l-2i?/w)
for circular hole and KT/K? = v/(w/7ic)tan(7ic/w)
R+a0
J ox(y,0)dy = a0
(11-36)
Using this criterion with Eq. (11-34) results in the notched to unnotched strength ratio
(11-37) where a0) and O"N is the notched strength of the infinite width laminate. For the case of center crack with crack length 2 c, the strength ratio is given by (Whitney and Nuismer, 1974)
(11-40)
(11-41)
for center crack. Equations (11-37) and (11-38) predict the proper trend of observed strength behavior, that is, decreasing strength of a laminate with increasing hole size. It is clear that the utility of the model would be greatly increased if the characteristic distance a0 is shown to remain constant for all laminates of a particular materials system. Otherwise, the utility of the model decreases considerably. Tan and Kim (1991) observed that in general the characteristic distance remains constant unless the predominant failure mode changes from fiber dominant mode to matrix dominant mode and vice versa.
(11-38)
11.4 Test Methods £2 = c/(c + a0) Figures 11-23 and 11-24 show the results of Eqs. (11-37) and (11-38) compared with experimental results obtained from a quasiisotropic laminate of graphite/epoxy (Whitney and Nuismer, 1974). The solid line represents prediction for a0 = 3.81 mm
11.4.1 Constituent Materials 11.4.1.1 Fibers
Test methods for determination of tensile strength and tensile modulus of reinforcing fibers will be described. Accurate measurement of tensile properties of rein-
11.4 Test Methods
517
(0/±45/90) 2 S T300/5208
Figure 11-23. Comparison of failure theory and experimental results for circular holes in [0/±45/90]2s laminate (Whitney and Nuismer, 1974). 5.0 7.5 Hole Radius R in mm
10.0
12.5
1.0 (0iA5/90]2S T300/5208 0.8 -
0.6 0.4 --
0.2 -
\
-c=±=
o0 = 3.8 mm
J
U-2C-*J
1° 2.5
" """"—•
-
+~ y 1
1
1
5.0 7.5 10.0 Crack Half Length C in mm
forcing fibers is useful from the standpoint of product development, quality control, materials selection, and design and stress analysis of composite materials. Test methods commonly applied to measure tensile properties of fibers are single filament test and resin impregnated yarn or strand test. The details of these test methods are given in ASTM Standards D 3379-75 for single filament test and D 4018-81 for yarn test. Sometimes fiber strength is calculated from the value of unidirectional composite laminates using Eq. (11-4) in Sec. 11.2.1.
1 12.5
Figure 11-24. Comparison of failure theory and experimental results for center cracks in [0/±45/90]2s laminate (Whitney and Nuismer, 1974).
The single filament test method has been widely used to determine tensile strength, modulus, and strain to failure on a variety of fibers. The single fiber drawn from yarn or tow is placed on a paper mounting tab as shown in Fig. 11-25 (ASTM, 1987). The fiber must be handled with special care to avoid any visible damage. Both ends of the gage length are bonded to the paper mound using an adhesive (sometimes wax). The specimen gage is recommended to be greater than 2000 times the fiber diameter and 25 mm long in most cases. After grip-
518
11 Strength of Fiber Composites This section burned or cut away after gripping in test machine
Cement or wax
50 Test specimen .Grip area
A
,40-
/ E
Width
/
.£ 30-
/
5: 5
Grip area
20-
Figure 11-25. Tab showing typical specimen mounting method (ASTM, 1987).
a. t
10y
ping the specimen in the testing machine, the middle section of the paper mounting tab is cut away. The tensile strength and tensile modulus are calculated from the following equations:
y
A
/
/
*1
2.5 )U /P System Cc mpliance S 0
1 10
20 30 Gage Length in mm
50
Figure 11-26. System compliance calibration method (ASTM, 1987).
Tensile strength .
(11-42)
Tensile modulus Ef = LIA (Ut/P - UJP)...
(11-43)
where P = applied load at failure, A = cross-sectional area of fiber L = gage length, Ut = total displacement, Us = testing system displacement. The cross-sectional area can be determined by a variety of methods such as laser diffraction, microscopy, etc. The testing system compliance UJP must be determined experimentally by using several different gage lengths as shown in Fig. 11-26. Instead of the specimen displacement, the specimen strain can be directly measured with considerable accuracy using a laser extensometer. The fiber strength determined by this method varies widely dependent upon the gage length as illustrated in Fig. 11-27 (Kim, 1986). Impregnated yarn test method is widely used by many fiber producers. The test
1 10 Gage Length in mm
100
Figure 11-27. Tensile strength versus gage length for AS4 single fibers. Solid line represents linear fit of data.
specimen is prepared by impregnating the sample yarn or tow with suitable binder materials. The purpose of the impregnating resin is to provide yarn, roving or tow, when cured, with sufficient mechanical strength to produce a rigid test specimen capable of sustaining uniform loading of the individual filaments in the specimen. If the impregnating resin-to-fiber ratio is ex-
11. 4 Test Methods
W = G = L = R =
O
Specimen Dimensions in mm
CD O O O O
cessively high or low, erratic test results will be obtained. Tension test specimens shall consist of whole-number multiples of the yarn, tow, etc. The total length of the specimen is usually 330 mm. The specimen is usually tested by either a special castresin end tab with gage length of 152 mm or no end tabs. Figure 11-28 (ASTM, 1987) illustrates the test specimen with end tabs. The gage length of the specimen is 250 mm. The specimen strain is usually measured by an extensometer. Sometimes the crosshead displacement is used for calculation of elastic modulus. In this case the modulus and strength are calculated by Eqs. (11-42) and (11-43), respectively. The cross-sectional area (A) in this case is given by dividing weight per unit length of dry yarn or tow (g/m) by density of the dry yarn or tow (g/m3).
519
Figure 11-29. Tensile test specimen of neat resin.
strain can be obtained using a strain gage mounted perpendicular to the loading axis. Special care for strain measurement for plastics and composites is given in Sec. 11.4.2.1.
-330 mm-152 mm-
11.4.1.3 Interfacial Bond or Fiber-Matrix Adhesion
16 mm
t
1 End tab
Figure 11-28. Tensile yarn test specimen with castresin tabs (ASTM, 1987).
11.41.2 Matrix The tensile strength, modulus, and failure strain can be determined by using flat panel or die cast specimens as shown in Fig. 11-29 (ASTM, 1987). The details of the test method are given in ASTM D 638M-84 for a thick panel (1.0-10.0 mm) and D 63872 for a thin panel (< 1.0 mm). Because of the high degree of strain rate dependency on the loading rate, test results must be carefully interpreted for application. The strain can be measured using either a strain gage or an extensometer. If Poisson's ratio is desirable, the transverse
Several test methods are available for measuring the strength of bond at the interface. The most common test in measuring interfacial bond strength is the fiber pullout test. This fiber pullout method is extensively used for glass fibers but found to be very difficult to use for graphite fibers because of their mainly brittle nature (they usually break before pullout). In order to overcome this difficulty, Drzal et al. (1980) developed an indirect test method using a single fiber imbedded in a resin dogboneshaped specimen as illustrated in Fig. 11-30. A tensile force is applied to the specimen and is then transmitted to the fiber by shear at the fiber-matrix interface. As shown in Fig. 11-31 a the fiber axial stress rises from the ends of the fiber until the fiber fracture stress oi is reached. At that point the fiber will fracture at some point
520
11 Strength of Fiber Composites -A
11.4.2 Laminates 11.4.2.1 Tension
L Section A-A
Figure 11-30. Schematic diagram of single fiber interfacial shear strength specimen (Drzal et al, 1980).
where the fiber stress is a maximum depending on fiber defect and probability. Continued application of stress to the specimen will result in repetition of fracture until all remaining fiber lengths are equal to or less than the critical length (Fig. 11-31 c). This lc or critical length can then be measured and an average interfacial shear strength (T) calculated according to the relationship (11-44) The specimen is examined under a polarizing microscope to measure the critical length. The details of the method can be found in ref. (Drzal et al., 1980). More extensive discussions of the interfacial strength behavior can be found in Chap. 6 of this Volume. !a)
The uniaxial tension test is the most fundamental method for the determination of data such as material specification, screening, research and development, and design of structural components. For unidirectional and woven composites, we can measure £ x , Ey, vxy, X and Y. For multidirectional symmetric laminates, we can measure El9 v 12 , and tensile strength of laminates. Straight-sided, constant cross section specimens with tabs bonded to the ends are widely accepted. Figure 11-32 shows the specimen geometry and dimensions recommended by ASTM D 3039-76. End tabs that are made of fabric or cross-ply E-glass/epoxy and aluminum plates are widely used. The 90° specimen is often tested without end tabs. The length of the tab is determined by the adhesive shear strength and tensile strength of the composite. It should range between 25 and 38 mm. Tab thickness varies with specimen thickness and can range between 1.5 and 2.5 mm. Any high elongation adhesive that will meet the environmental conditions is recommended. The bond surface must be prepared by sanding and cleaning with an appropriate solvent such as acetone. A diamond impregnated saw with water cooling is recommended for cutting the specimen. This should prevent rough or un-
<J1 • -
\. (b)
a2
\
Ic)
a3
^-
a3>a2>
Figure 11-31. Fiber fracture and stress distribution in an interfacial shear strength specimen, (a) At low stress, (b) after fiber break at defects, and (c) after critical length is reached (Drzal et al., 1980).
11.4 Test Methods
T b
" 5-10 deg. taper Figure 11-32. Tensile coupon.
even cut edges and surface scratches. Specimen edges should be parallel to within 0.125 mm. The following is the general test procedure. Width and thickness are measured at several places, and the minimum values are used to calculate the cross-sectional area. Apply load to the specimen through a set of wedge section grips in order to provide sufficient lateral pressure to prevent slippage. A serrated grip surface is desirable (always keep the serrated surface clean). Alignment of the specimen can be checked by strain gages mounted as shown in Fig. 11-33. A constant strain rate of 16.7 10" 6 to 33.7- K T S " " 1 is recommended. A constant crosshead speed or loading rate corresponding to the strain rate is also acceptable. Strain measurement during test can be done using either an extensometer with a gage length of 12.5 or 25 mm or electrical resistance strain gages having 350 or 120 ohm with a gage length in the range of 3.3 to 6.5 mm. It is suggested that manufacturer's recommendations be followed for strain gage mounting technique and selection of adhesive. How-
ever, precaution must be taken for probable stiffening of the test specimen by the strain gage and adhesive. This often occurs in certain composites such as the low-density carbon-phenolic composites. Misalignment of the strain gage axis to the specimen axis is a more serious problem with anisotropic material than the conventional isotropic material. Low voltage for the strain gage circuit is desirable to minimize heat generation. One volt is recommended. Flexural Test This is a quality control and material specifications test. It is not a data generation test. This test is used to determine the outer fiber tensile strength and Young's modulus of homogeneous composite and polymeric materials. For flexural testing of multidirectional composite laminates, the strength and stiffness interpretations are not as simple because of the complex state of stress in the specimen. The most common three-point and four-point flexural tests are illustrated in Figs. 11-34 and 11-35, respectively. The three-point test is
•L/2
p/2
p/2
Figure 11-34. Three-point flexural test. p/2
p/2
-L/2—H 1 1
1
2i
i
Figure 11-33. Specimen alignment, gage 3 is on the back. | [ e 3 - ( e x + e2)/2]/e11 < 0.05 and \(e1-s2)/e11 < 0.05.
521
p/2
p/2
Figure 11-35. Four-point flexural test.
522
11 Strength of Fiber Composites
principally designed for materials that break at relatively small deflection. The strength and modulus for this test are given by X =
3PL 2bh2
(11-45)
E =
PL3 4bh38
(11-46)
where 3 is deflection at the mid-point of the specimen. The four-point test is primarily designed for materials that undergo large deflection. The strength and modulus for this test are given by X = E =
3PL 4bh2 11PL 3
64bh35
(11-47) (11-48)
In flexure test the specimen should be deflected until failure occurs. Deflection should not exceed 10 percent of the span length. Correction factors for large deflection will be needed. A cylindrical surface is used for loading and supporting noses. The diameter of the nose is usually greater than 6.4 mm to minimize local indentation and stress concentration. In many cases of advanced composite testing, the indentation due to the nose cannot be completely eliminated. The span-to-depth ratio (L/h) depends on the ratio of the tensile to interlaminar shear strengths and a guideline for suitable L/h ratio will be given in Sec. 11.4.2.6. Recommended ratios are 16, 23, 40 and 60 for composite laminates. When a specimen is tested at low span-todepth ratio, shear deflection influences the apparent modulus. The 60:1 ratio is recommended for modulus determination.
11.4.2.2 Compression
Compressive elastic constants and strengths are determined by this test. It is well known that it is exceedingly difficult to determine accurate, reliable, and reproducible compressive strengths for unidirectional composites. Because of the high anisotropy of these materials, compressive strength is sensitive to the test method and procedure employed. A slightly eccentric load will cause premature buckling rather than the intrinsic compressive failure. A number of test fixtures have been developed to measure true compressive strength by various investigators. As a result, a number of techniques have evolved and are described in the literature. The method of load application, either direct end loading or shear loading such as ASTM D3410, significantly influences the compressive strength and the failure mode. Direct end loading causes end crushing and vertical splitting in orthotropic materials. This vertical splitting occurs at a stress considerably lower than the compressive strength of the materials system. To avoid these undesirable features, shear loading has been commonly employed and offers the most attractive and feasible approach to determine composite compression strength. The most widely-used test method in the composite is ASTM D 3410, which uses either a Celanese fixture or an IITRI test fixture. These fixtures are designed for a specimen gage length less than 12.6 mm and are primarily used for the 0°, 90°, and cross-ply laminates. Figure 11-36 illustrates the Celanese fixture and specimen. Recommended specimen thicknesses for various materials are: Boron Graphite Glass
1.5-2.0 mm 1.5-3.0 mm 3.2-4.0 mm
11.4 Test Methods
523
Fixture
Dimensions in mm: a = 3.99 b = 6.35 c = 57.15 d = 12.7 e = 3.18 f = 63.5
TEdge front
Euler's column buckling formula can be used for the specimen thickness calculation. For pinned-end column with length L critical
= n2EI/L2
(11-49)
For a fixed-end column the buckling load is four times that for a pinned-end. The test fixture is placed between two platens, and the compressive load is applied. One of the platens normally has a spherical seat to minimize the eccentricity in the load introduction. The short gage length of this test has several undesirable features. Probable end effects for measuring elastic constants and strengths will be encountered. This test method cannot be applied to laminates with off-axis angles and cannot be used for tension-compression fatigue-testing. The failure normally occurs at the gripping area rather than the gage section. These specimens (Celanese and IITRI) frequently fail prematurely, due to an instability arising from misalignment of the specimen with respect to the loading axis. To overcome this instability, SACMA (Sup-
Figure 11-36. IITRI compression test fixture with specimen.
pliers of Advanced Composites and Materials Association) advocates a modified ASTM D 695 method where the specimen has a gage length of 5 mm. End effects are of significant concern with such a short gage length, and furthermore the instability still remains a problem. In addition to these drawbacks, a majority of the specimens tested by these techniques usually fail within the gripped section rather than the gage section. For the above reasons, the ultimate failure stress determined in these tests may not represent the true compressive strength of the material system. Recently, a promising test method has been developed using a mini-sandwich specimen (Crasto and Kim, 1990). A sandwich was constructed in which the conventional honeycomb core was replaced with a core of neat resin, similar to the matrix of the composite system to be tested. The specimen geometry and other dimensions (except for thickness) are in accordance with ASTM D3410. Specimens are tested in the IITRI compression test fixture. The compressive stress in the
524
11 Strength of Fiber Composites
composite skins is calculated by applying the rule of mixtures and given by Xf = — (
(11-50)
where V is volume fraction, o stress, e strain, and E elastic modulus. The subscripts "a", "1", and "2" represent the entire specimen, skin (composite), and core, respectively. The results of sandwich specimens and ASTMD3410 (all-composite) are compared in Table 11-3. The compressive strengths obtained from the sandwich specimen are significantly greater than those determined from the ASTM specimen and comparable to the tensile strengths. The most significant result of this test method is that these mini-sandwich specimens display no evidence of premature failure caused by instability of the test specimen, and specimen failure is confined to the gage section in most cases. The details of this method can be found in ref. (Crasto and Kim, 1990).
Table 11-3. Compressive strength of sandwich and all-composite specimens. Material system S-glass/1034 AS-4/3501-6 AS-4/PEEK
Fiber volume
Minisandwich (MPa)
All composite (MPa)
60 62 57
2284 2020 1573
1400 1280 1100
where p is applied load at failure, b specimen width, d specimen thickness, as shear stress, and ax and e2 axial and transverse strains, respectively. The classical laminated theory indicates that a combined state of stress rather than a pure shear state of stress exists in the specimen. The results of this specimen compared well with the results of the cylindrical torsion specimen that has the pure shear. This method has the advantage of utilizing the relatively inexpensive straight-sided tensile coupon and conventional test method without any special fixture. A nonlinear stress-strain relation usually occurs in shear. Interpretation of shear strength requires care and judgement.
11.4.2.3 Shear Tests
There are a number of test methods available for the in-plane shear modulus, in-plane shear strength, and interlaminar shear strength. [±45] s Coupon This is a simple test that follows the same procedure as the tensile test. Figure 11-37 shows the specimen and the locations of the two strain gages. From the measured longitudinal and transverse strain data, we can deduce the shear strength (S) and shear modulus (£s), as follows: S
=p/2bd
(11-51) (11-52)
Rail Shear
A flat rectangular plate is tested in a two-rail or three-rail fixture. The latter is shown in Fig. 11-38. An analysis of the load and strain gages data is as follows:
P 2A
(11-53)
= 2 8'45
(11-54) (11-55)
Strain gages 1 and 2 Figure 11-37. [±45] 2s specimen for in-plane shear test.
11.4 Test Methods
525
Center rail slides through guide
Dimensions in mm: a = 203.20
—
Bolts
b = 152.40 c = 1 84.15 d=
31.75
e =
12.70
f =
44.45
g =
31.75
/ =
19.05
} =
9.53 dia
Figure 11-38. Three-rail shear test fixture.
where A = b h, b is the length, and h the thickness. Special recommendation on the rail shear: drill oversized holes to prevent stress risers when bolts are tightened. Abrasive paper or cloth should be adhered to the rails for improved gripping. Recommendation on the specimen: holes should be drilled and reamed. Use carbide tipped drill bit. Holes should be larger than the bolts. Apply torque of 7 to 70 Nm to each bolt. Tighten bolts evenly. A fixed pattern of tightening should follow finger tightening. The final level should be reached in two or three stages. On testing; unless otherwise specified, use a loading rate of 1 to 1.5 mm/min. Apply preload and release to align the heads and rails. Alignment can be improved by using a spherical seat between the load head and the center rail. For unidirectional laminates, the fiber orientation can be parallel or perpendicular to the longitudinal axis of the rails. The perpendicular orientation produces higher shear
strength than the parallel orientation. Failure in the parallel orientation often starts in the corners of the rail caused by local stress concentration. Iosipescu Test
The specimen is loaded to produce a shear stress acting between the notch ends with zero bending moment and zero normal stress. The shear test is achieved by application of two counteracting moments produced by two force couples as illustrated in Fig. 11-39 (Adams et al., 1990). Under such a loading configuration, a state of constant shear stress is induced through the middle section of the test specimen and given by P Wt
(11-56)
where t is specimen thickness. The shear loading is achieved by a shear fixture
526
11 Strength of Fiber Composites
to three decimal figures:
P
(7^ = 0.970(7! (7yy = 0.030 *! (7^ = 0.171(7!
(11-58)
For the shear failure, the following conditions must be satisfied if the combined stress interaction is not taken into account: Figure 11-39. Schematic of the loading fixture for the Iosipescu shear test.
a
S X (11-59)
which prevents the rotation of the ends of the specimen while undergoing shear loading. A number of different specimen designs and test fixtures have been used in the Iosipescu test. 10° Off-Axis Test The 10° off-axis unidirectional tension coupon has been used for the in-plane shear characterization of continuous fiber reinforced composite materials. This test method was proposed by Chamis and Sinclair (1976), and later adopted as an ASTM standard test method for in-plane shear. The state of the stress in the off-axis tension coupon is given by the following transformation equations which are derived from force equilibrium considerations: axx = GX cos2 9 ayy = ax sin 2 9 avv = a, sin 2 9
(11-57)
al denotes applied stress, axx,
yy
These inequalities are true in most continuous fiber reinforced composite materials. Furthermore, in most advanced composite materials, the computed longitudinal and transverse stresses are approximately 60 percent and 40 percent of their respective failure stresses when the computed shear stress reaches its shear strength. Since the contribution of longitudinal stress and transverse stress to the shear failure is so small, the use of the 10° off-axis specimen is justifiable. The effect of the combined stress interaction on the shear failure appears to be negligible. Short Beam Shear This test is for estimating the interlaminar shear strength only. The test is shown in Fig. 11-34 or 11-35 with an appropriate span-to-depth ratio to produce interlaminar failure instead of tensile failure. The analysis for this test for the interlaminar shear strength Sj is:
= 3P/4bh
(11-60)
where P is applied load, and b and h are beam width and thickness, respectively. This test is not suitable for design data generation because failure often occurs at a location other than the expected neutral
527
11.4 Test Methods
plane. It is therefore important to know the actual location of the fracture and the failure mode. The span-to-depth ratio should be selected so as to induce interlaminar shear fracture mode. The ratio between the maximum outer fiber stress and the interlaminar stress as a function of the span-todepth ratio is shown in Fig. 11-40. The selection of span-to-depth ratio above the straight line for a given strength ratio would lead to shear failure. For example, for the strength ratio 10, the span-to-depth ratio for interlaminar shear failure should be smaller than 5 and 10 for three-point bend and four-point bend, respectively.
30 3pt/ .20 4-pt^ 10
0
5
10 L/h
15
20
Figure 11-40. Strength ratio vs. span-to-depth ratio.
Double Notch Shear A double notch specimen for interlaminar testing is shown in Fig. 11-41. The two notches are cut to half thickness in each face and are spaced a distance L apart along the length of the specimen. Under either applied tension or compression, shear stress is distributed along the central plane section between the two notches and given by 1
(11-61)
WL
where W is specimen width. The specimen geometry must be chosen which guarantees that only tangential stresses will act and that failure due to the interlaminar shear will take place. It is well recognized that there exists a considerable stress concentration around the region of the notch tip. The variation of shear stress, as, between the notches is shown in Fig. 11-42. The shear stress ratio crs/Sl varies with the value of 9 as given by (Markham and Dawson, 1975) '2
(11-62)
Figure 11-41. Schematic of the double notch specimen for interlaminar shear test, t is the specimen thickness.
where t is thickness, G shear modulus, and E Young's modulus. The stress concentration factor decreases with 9 value and reaches to roughly unity at 9 = 0.5. In addition, the bending moment (M = P t/2) produces normal stress at the notch tip which in turn significantly influences the measured value of shear strength. This normal stress is tension for applied tensile load, while it is compression for compressive load. The specimen may fail due to the tensile normal stress rather than the interlaminar shear when tested under tension. On the other hand, the specimen may fail due to the shear stress when tested under compression. This is the main explanation for the experimental observation that the failure load in compression is almost two times greater than in tension. Whitney (1989) performed a stress analysis on the double notch specimen and suggested that
11 Strength of Fiber Composites
Figure 11-42. Variation of interlaminar shear stress along the center plane between notches.
a relatively small value of L/t is desirable to assure that the interlaminar stresses are distributed over majority of the specimen, rather than being isolated to a region near the notch tip. For a large value of L/t, a pure tension field results in a major portion of the specimen.
Xx, X2, • • •, Xn is a random sample of size n (data) from the Weibull distribution, W(a,j3), the likelihood function, L(a,/J), is given by: L{pL9P)= ft/(*,-,a, jB)
(11-65)
11.4.3 Statistical Treatment of Data For static strength testing of composite materials, the two-parameter Weibull distribution is used widely for failure estimation. A random variable, X (static strength), has a Weibull distribution with shape parameter a and scale parameter /? if the cumulative distribution function, Fx (x\ is:
x>0
(11-63)
or its density function, fx (x), is given by
The maximum likelihood estimations of a and ft maximize L (a, /?) and therefore it is necessary to solve the equations: 8 — L(ocJ) = O (11-66) 6a and 8 — L(ot,P) = O
(11-67)
That is, a and /? are solutions of the maximum likelihood equations:
£ xflnx,.
£ lnx,i= 1
(11-64) A maximum likelihood estimation method is applied for the parameter estimation. If
= I"1 v *T
=o
(11-68)
(11-69)
529
11.4 Test Methods
The symbol " A " denotes the estimate of the parameters. The Weibull parameters a and P are related with mean, \i, and variance,
^.^{r(. + ?)-r»(.+I)}(«-7i, An efficient iterative technique for obtaining a, the solution off (A) = 0 [Eq. (11-68)], is the Newton-Raphson method, in which the (j + l)th successive approximation, &j+1 to &j, is given by:
f'(&j)
(11-72)
Design Allowables A statistical procedure will be described for obtaining material allowables for use in evaluating composite material capabilities in structural design. The allowable is a value determined from a specified probability of survival with a 95% confidence in the assertion. A-basis (-allowables) and B-basis (-allowables) are defined by 99% and 90% probabilities of survival. The confidence intervals for a and /? can be computed from Tables 11-4 and 11-5 given by Thoman et al. (1969), that is, =7
(11-73)
for a and
Table 11-4. Percentage points Iy so that PT{d/a
5 6 7 8 9 10 12 14 16 18 20 22 24 26 28 30 40 50
for /?. In the Tables, n and y are sample size and confidence level. Using the values of the confidence interval, aL for a in Eq. (11-73) and fih for /? in
0.95
0.98
2.277 2.030 1.861 1.747 1.665 1.602 1.513 1.452 1.406 1.356 1.343 1.320 1.301 1.284 1.269 1.257 1.211 1.182
2.779 2.436 2.183 2.015 1.896 1.807 1.682 1.597 1.535 1.487 1.449 1.418 1.392 1.370 1.351 1.334 1.273 1.235
3.518 3.067 2.640 2.377 2.199 2.070 1.894 1.777 1.693 1.630 1.579 1.538 1.504 1.475 1.450 1.429 1.351 1.301
Table 11-5. Percentage points I* so that
\ n 5.. 6 7. 8 Q
10 12 14 16 18 70
(11-74)
0.90
99
?4 96 78
30 40 50
y
0.90
0.95
0.98
1.107 0.939 0.829 0.751 0.691 0 644 0.572 0.520 0.480 0.447 0.421 0.398 0.379 0.362 0.347 0.334 0.285 0.253
1.582 1.291 1.120 1.003 0.917 0.851 0.752 0.681 0.627 0.584 0.549 0.519 0.494 0.472 0.453 0.435 0.371 0.328
\ .. 0.772 0 666 0 598 0 547 0 507 0 475 .. 0.425 .. 0.389 .. 0.360 0.338 0.318 . .0.302 . 0.288 ,.. 0.276 ... 0.265 ... 0.256 ... 0.222 ... 0.195
530
11 Strength of Fiber Composites
Eq. (11-74), A-basis and B-basis can be determined from the following equations ln
is = ^L [in ( ^ ] 1 M L
B-basis
(11-75) (11-76)
Example The tensile strength data for the [0/ ± 45/ 90]s laminate of graphite/epoxy in MPa are 497, 499, 528, 532, 533, 547, 555, 562, 565, 566, 567, 569, 575, 578, 586, 591, 592, 596, 597, 598, 607, 610, 613, 618, 622, 630, 632, 637, and 644. The following estimates are obtained using the Newton-Raphson approximation: dc= 18.01 £=598 To find the 95% confidence intervals for a and /?, let a 0 95 = 1.342 (n = 29) and let I* 95 = 0.34 (n = 29) from Tables 11-4 and 11-5. Therefore, 95% confidence intervals for a and /?, are a
18.01 = 13.42 1.342 598 = 587 exp (0.34/18.01)
With these values A-basis and B-basis can be calculated from Eqs. (11-75) and (11-76) and are 417 MPa and 496 MPa, respectively.
11.5 Conclusions The prediction of the static strength of unidirectional composite laminates from the properties of their constituents is not reliable in many cases because there are many unknown variables which influence the strength. These variables range from
constituent properties and inherent material flaws to fabrication processes. The formulas derived from micromechanics, however, provide valuable insight in identifying important variables, and their influence on the strength, and give direction to material improvement. There exists a serious need for more analytical and experimental work to improve the accuracy of the predictions and in the characterization of the constituents and fiber-matrix interface properties. The prediction of strength for multidirectional laminates after first ply failure is complicated by a multitude of independent and interacting damage mechanisms of fiber breaks, matrix cracks, delamination and crack propagation. These complicated failure mechanisms with successive ply failures make it difficult to calculate the stress redistribution in the laminate at the moment of ultimate failure. Consequently, the stresses calculated on the basis of the intact laminate significantly differ, in general, from the actual state of stresses at failure, and this results in an unreliable prediction of strength in many cases. More experimental data must be generated to fully verify the existing failure criteria under a wide range of combined loading. Thus, strength prediction may be regarded as a useful tool for material characterization and design, which is a relatively simple method of estimating the load carrying-capacity of a structural component. An effort must be made to develop a methodology to quantify the degree of degradation caused by progressive failure in a multidirectional composite laminate. Finally, we need more systematic research to develop methodology that can integrate micromechanics and macromechanics and predict the load-carrying capacity of a structure from the properties of the constituents.
11.6 References
11.6 References Adams, D. F. (1990), Polymer Composites 11, 287. Adams, D. R, Doner, D. R., Thomas, R. L. (1967), AFML-TR-67-96: Wright-Patterson Air Force Base (Ohio). ASTM (1987), ASTM Standards and Literature References for Composite Materials. Philadelphia (PA): ASTM. Brelant, S., Petker, I. (1970), Mechanics of Composite Materials: Wendt, F. W, Liebowitz, H., Perrone, N. (Eds.). Oxford: Pergamon Press. Chai, H., Babcock, C. D., Knauss, W. G. (1981), Int. J. Solids and Struct. 17, 1069-1083. Chamis, C. C. (1967), DMSMD Rep. No. 9. Cleveland, Ohio: Case Western Reserve University. Chamis, C. C. (1974), Composite Materials Vol. 6: Plueddemann, E. P. (Ed.). New York: Academic Press, pp. 31-74. Chamis, C. C , Sinclair, J. H. (1976), NASA TN D8215. Houston (TX): NASA. Crasto, A., Kim, R. Y. (1990), International SAMPE Technical Conference 22, 264-277. Cruse, T. A. (1973), J. Compos. Mater. 7, 218. Donaldson, S. L. (1987), Compos. Sci. and Technol. 28, 34. Drzal, L. T, Rich, M. J., Camping, J. D., Park, W. J. (1980), 35 th Annual Technical Conference: Society of Plastic Industries: p. 20-c. Flaggs, D. L., Kural, M. H. (1982), J. Compos. Mater. 16, 103-116. Garrett, K. W, Bailey, J. E. (1977), J. Mater. Sci. 12, 157. Hahn, H. T, Williams, J. G. (1986), Composite Materials: Testing and Design, ASTM STP 893, 115 — 139. Hill, R. (1950), The Mathematical Theory of Plasticity. Oxford: Oxford University Press. Kim, R. Y. (1980), Third International Conference on Composite Materials: Paris (France): p. 1015. Kim, R. Y. (1981a), ASTM STP 734, 91-108. Kim, R. Y (1981b), Technical Report UDR-TR-81-84. Dayton (Ohio): University of Dayton Research Institute. Kim, R. Y. (1983), 28th National SAMPE Symposium and Exhibition: Anheim (CA): pp. 200-209. Kim, R. Y. (1986), Unpublished data. Kim,R. Y, Hahn, H. T. (1979), /. Compos. Mater. 13, 3 Kim, R. Y, Soni, S. R. (1984), /. Compos. Mater. 18, 70. Markham, M. R, Dawson, D. (1975), Composite, July, 173. O'Brien, T. K. (1981), NASA Technical Memorandum 81940. Houston (TX): NASA. Pagano, N. J. (1978), Int. J. Solids and Struct. 14, 385. Pagano, N. J., Hahn, H. T (1976), ASTM STP 617, 317-329.
531
Pagano, N. J., Rybicki, E. F. (1974), J. Compos. Mater. 8, 214. Pagano, N. J., Soni, S. R. (1983), Int. J. Solids and Struct. 19(3), 207. Paris, P. C , Sih, G. C. (1965), ASTM STP 381, 84113. Parvizi, A., Bailey, J. E. (1978), J. Mater. Sci. 13, 2131. Patterson, R. E. (1974), Stress Concentration Factors. New York: Wiley, pp. 110-111. Pipes, R. B., Pagano, N. J. (1970), J. Compos. Mater. 4, 538. Pipes, R. B., Kaminski, B. E., Pagano, N. J. (1973), ASTM STP 521, 218. Reifsnider, K. L., Henneke, E. G., Stinchcomb, W. W. (1979), AFML-TR-76-81 Part IV: Wright-Patterson Air Force Base (Ohio). Rosen, B. W. (1964), J. Am. Inst. Aero. Astron. 2, 198. Tan, S. C , Kim, R. Y. (1991), ASTM STP 1120, 414-427. Thoman, D. R., Bain, L. J., Antle, C. E. (1969), Technometrics II, 445-450. Tsai, S. W. (1968), Fundamental Aspects of Fiber Reinforced Plastic Composites: Schwartz, R. T, Schwartz, H. S. (Eds.). New York: Wiley Interscience, pp. 3-11. Tsai, S. W, Hahn, H. T. (1980), Introduction to Composite Materials. Lancaster (PA): Technomic, Chap. 7. Tsai, S. W, Wu, E. M. (1971), /. Compos. Mater. 5, 58-80. Waddoups, M. E., Eisenmann, J. R., Kaminski, B. E. (1971), J. Compos. Mater. 5, 446. Wang, A. S. D., Crossman, F. W (1980), / Compos. Mater. Suppl, 71. Whitcomb, J. D. (1984), NASA Technical Memorandum 86301. Houston (TX): NASA. Whitney, J. M. (1989), 4th Technical Conference, American Society for Composites. Lancaster (PA): Technomic. Whitney, J. M., Nuismer, R. J. (1974), /. Compos. Mater. 8, 254. Wu, E. M. (1972), J. Compos. Mater. 5, All.
General Reading Ashbee, K. H. G. (1989), Fundamental Principles of Fiber Reinforced Composites. Lancaster (PA): Technomic. ASM International (1987), Engineered Materials Handbook, Vol. 1, Composites. Metal Park (OH): ASM International. Broutman, L. J. (Ed.) (1974), Composite Materials, Vol. 5, Fracture and Fatigue. New York (NY): Academic Press. Datoo, M. H. (1991), Mechanics of Fiberous Composites. London: Elsevier Applied Science.
532
11 Strength of Fiber Composites
Hahn, H. T, Zweben, C. (1989), Mechanical Behavior and Properties of Composite Materials in Delaware Composite Design Encyclopedia, Vol. 1. Lancaster (PA): Technomic. Jones, R. M. (1975), Mechanics of Composite Materials. Washington (DC): Scripta Book. Tsai, S. W. (1988), Composite Design, 4th Ed. Dayton (OH): Think Composites.
Vinson, I R., Sierakowskie, R. L. (1987), The Behavior of Structures Composed of Composite Materials. Dordrecht: Kluwer. Whitney, I M., Daniel, I. M., Pipes, R. B. (1982), Experimental Mechanics of Fiber Reinforced Composites. Brookfield Center (CT): Society of Experimental Mechanics.
12 Fracture of Fiber Composites Leif A. Carlsson Department of Mechanical Engineering, Florida Atlantic University, Boca Raton, FL, U.S.A.
List of Symbols and Abbreviations 12.1 Introduction 12.2 Basic Fracture Mechanics Concepts 12.2.1 Stress-Intensity Factors for Isotropic Solids 12.2.2 Stress-Intensity Factors for Anisotropic Solids 12.2.3 Strain-Energy Release Rate 12.2.4 Mixed-Mode Fracture Mechanics 12.2.5 Cracks at Bimaterial Interfaces 12.2.6 Cracks in Bonded Structures 12.3 Fracture Specimens 12.3.1 Double Cantilever Beam Specimen 12.3.2 End-Notched Flexure Specimen 12.3.3 Mixed-Mode Fracture Specimens 12.3.4 Test Specimens for Interleaved Composites 12.4 Fracture Data 12.4.1 Mode I Fracture Data 12.4.2 Mode II Fracture Data 12.4.3 Mixed-Mode Fracture Data 12.4.4 Mode I and Mode II Fracture Data for Interleaved Composites 12.5 Acknowledgements 12.6 References
Materials Science and Technology Copyright © WILEY-VCH Verlag GmbH & Co KGaA. All rights reserved.
534 536 537 538 539 544 548 556 559 561 561 562 564 566 572 573 575 576 578 580 580
534
12 Fracture of Fiber Composites
List of Symbols and Abbreviations a semi-crack length A crack area 5a crack extension c distance between point of load application and midspan C compliance Cij stiffness elements Cf modified shear-compliance E Young's modulus; axial modulus of beam E1,E2 Young's modulus in fiber direction, in transverse direction F c ,F d normal forces to close crack FH,FV horizontal and vertical crack-closure forces FT crack-closure force applied in the z-direction G strain-energy release rate G shear modulus Gc work required to create a unit crack area Gc fracture toughness Gio GIIC fracture toughnesses for mode I, II loadings GIC (init.) initiation fracture toughness GIC (prop.) steady-state fracture toughness GI? G n , G in strain-energy release rate components for mode I, II, and III loadings G*, G^ asymptotic energy-release rates G12 shear modulus in 1-2 plane h thickness H potential energy; experimentally determined parameter hl9h2 distances from top (1) and bottom (2) surfaces / moment of inertia of a cross-section k shear correction-factor Kl9KU9Km stress intensity factors for mode I, II, and III loadings K±,K2 bimaterial interfacial stress-intensity factors ^ic> ^IIC stress intensity factors L half span length M moment Ml9 M 2 , M 3 , Ml9 Mu moments (local quantities) M* moment P applied load Pc critical load Pt upward force Pu P2, P3 axial loads Pl9 Pu P components for mode I, II loadings r radial distance from the crack-tip Tc, Td tangential forces to close crack u x-axis displacement uc,ud tangential nodal displacements
List of Symbols and Abbreviations
elastic-strain energy stored in a body y-axis displacement normal nodal displacements specimen width work supplied by movement of external forces material-dependent factor complex stress function
u v w W Y Z a yxy
V
angle shear strain displacement at the critical load size of a crack-tip element oscillation index strains in x-, y-, z-directions angle Poisson's ratio thickness ratio stress stress along x, y, and z-axes yield strength of matrix far-field tensile stress shear stresses far-field shear stress angle Airy stress-function; Westergaad function; angle; mode mixity angle; dimensionless function Laplacian operator
ASTM CLS DCB ELS ENF GMPS IDCB IENF ksi LMPS LVDT MMB NDE PEEK PTFE TPI TSI
American Society for Testing and Materials cracked-lap shear double cantilever beam end-load split end-notch flexure global mode partitioning scheme interleaved DCB specimen interleaved ENF specimen 1000 lb/sq. inch local mode partitioning scheme linear variable differential transformer mixed-mode bending nondestructive evaluation poly(ether-ether-ketone) poly(tetrafluoroethylene) thermoplastic interleaf thermoset interleaf
<5C A 8 X'
V'
Z
9 V
G
Gx, Gy,
Gz
G
y
Xxr
TXZ,
CO
Xy z
535
536
12 Fracture of Fiber Composites
12.1 Introduction With the development of improved manufacturing methods, composite materials have become the materials of choice whenever high stiffness and strength at minimum weight are desired. Consequently, many advanced structural applications such as airplane wings, fuselages, and aerospace structures now involve composites. Although advanced nondestructive evaluation (NDE) techniques applicable for detecting flaws and defects in composite structures have been developed (ASTM STP696, 1979; Pipes and Blake, 1990), there is still a possibility that material deficiencies in the form of preexisting flaws may pass inspection. Another threat of great concern for composite structures is service-induced damage, where low velocity, hard-object impact damage is most important (ASTM STP936, 1987; Masters, 1985; Guynn and O'Brien, 1985; Bostaph and Elber, 1982). This type of damage may be caused by dropped tools on an airplane structure or runway rocks and debris hitting the airframe during take-off or landing. There are many possible types of damage in composites as discussed by Beaumont and Schultz (1990 a, b). They list fiber fracture and debonding, fiber pullout, matrix microcracking and matrix cracking parallel to the fibers as failure processes in composites. On a macroscopic level, damage can be grouped into fiber breakages, matrix cracks parallel to the fibers, and "delamination" or "interlaminar cracks", which represent cracks between the plies in a composite laminate (Wilkins, 1983). The delamination mode of failure has received considerable attention in the literature, (e.g., ASTM STP 876, 1985; Bascom et al., 1980; O'Brien, 1982; Russell and Street,
1982; Wang, 1983; Wilkins etal., 1982), because delaminations may significantly reduce the compressive load-bearing capability of a composite structure and may grow during cyclic loading (Martin and Murri, 1990; Mohlin et al., 1985; O'Brien, 1982; Trethewey, 1988; Wilkins, 1982). Fracture mechanics was introduced after World War II to analyze fracture, which occasionally occurred at low stresses, in high strength steel structures (Broek, 1982). Fracture mechanics is a science developed to analyze the tendency for a preexisting crack in a structure to grow as a result of applied external loads. Due to the preexistence of a crack, the strength of the structure may be decreased and, depending on the crack size, may fall below the designed limit load. Under these conditions, high loads may cause crack propagation and structural collapse. The assessment of defect criticality on structural performance is one of the applications of fracture mechanics. Another common situation where fracture mechanics, in combination with NDE techniques, has been found extremely useful is in the prediction of the maximum crack size that can be allowed in a structure. Furthermore, fracture mechanics may be employed to formulate laws on crack extension with an elapsed number of load cycles (fatigue). This information enables structural inspection schedules to be defined for monitoring the actual growth of the crack. The field of fracture mechanics traditionally involves applied mechanics and materials science. Applied mechanics relates external loads applied to a flawed, or cracked, structural component to crack-tip stress fields and elastic deformations of the material in the vicinity of the crack tip. By contrast, materials science views fracture mechanics as a means of characterizing the fracture resistance of a material on a mi-
12.2 Basic Fracture Mechanics Concepts
crosopic scale. In this manner, material parameters of importance to crack resistance can be isolated and materials with improved fracture toughness can be devised. Both aspects of fracture mechanics have provided improved analytical capabilities and damage tolerant structural materials. Most work on fracture mechanics is related to isotropic materials, such as metals or polymers. Several textbooks (e.g., Broek, 1982; Ewalds and Wanhill, 1983; Hellan, 1984; Kinloch and Young, 1983; Knott, 1973) dealing with fracture in isotropic materials have been published. Fracture mechanics of anisotropic materials has been treated to a much lesser extent, although basic solutions to plane-crack problems are available (Sih et al., 1965; Sih and Liebowitz, 1968). It is the objective of this chapter to review some important developments in linear elastic fracture mechanics of composite materials. Because of the linear elastic and self-similar crack growth restrictions, the analysis and experiments presented are limited to crack propagation parallel to the fibers, such as matrix cracking and delamination. Crack propagation perpendicular to the fibers is not considered. Both applied mechanics and materials science aspects of fracture will be emphasized. The chapter is organized in three
537
major sections. In the first section, basic fracture mechanics concepts are reviewed. The second section deals with the analysis of some contemporary fracture test specimens employed for composite materials. The final section considers experimental fracture toughness data for composite materials and also discusses methods to improve the fracture resistance of composites. Only high-performance continuous fiber composites are considered in this review, because these materials are employed in highly stressed structures. Fracture of short fiber composites is considered in, for example, Mandell et al. (1981) and Friedrich (1989).
12.2 Basic Fracture Mechanics Concepts In this section we will introduce the concepts of crack-tip stress intensity factors and the elastic-strain energy release rate. Relations between these fracture parameters will be discussed for the three basic modes of crack surface displacement illustrated in Fig. 12-1. A common coordinate system has the x-axis normal to the crack contour, the y-axis perpendicular to the crack plane, and the z-axis directed along the edge of the crack, see Fig. 12-1. Stress
Figure 12-1. Modes of crack loading and definition of stress components: (a) Mode I (opening), (b) mode II (forward shear), (c) mode III (tearing).
538
12 Fracture of Fiber Composites
components defined according to this doordinate system are also illustrated in Fig. 12-1. Subsequent to this section, mixed-mode fracture will be discussed, that is, situations where combinations of the basic modes illustrated in Fig. 12-1 prevail. Following this, cracks between dissimilar elastic solids and cracks in bonded structures will be discussed. 12.2.1 Stress-Intensity Factors for Isotropic Solids
As an introduction, the stress field for a mode I loaded crack in an isotropic solid will be discussed. Consider a cracked body in Fig. 12-2 subject to mode I loading. For plane problems, the equilibrium equations in the absence of body forces are 9TV
9x
+
dy
(12-1 a)
•=0
(12-1 b)
dx
With u and v denoting x- and y-axis displacements, the expressions for the strains are
_dv "y~d~y xy
(12-2b)
_ du dy
dv dx
(12-2c)
The equilibrium equations are satisfied if x= | | a
>
=
(12-4a)
8V
(12-4 b) (12-4c)
8x6);
where i// = ifr (x, y) is the Airy stress function (Timoshenko and Goodier, 1970). Combination of Eqs. (12-2), (12-3), and (12-4) yields, after differentiation twice,
8>
8V
9x4+ or
8x28
8V 8/
4
V iA = 0
l/£ = [ -v/E 0
- v/E \/E
(12-6)
where "Re" and "Im" denote the real and imaginary parts of a complex quantity, respectively, and Z and Z are defined by
0
where E and G are Young's modulus and the shear modulus, respectively, and v is Poisson's ratio; G = £/[2(l + v)].
(12-5)
Equation (12-5) is the biharmonic equation of elasticity theory. As will be shown later, a similar equation may be derived for plane problems in anisotropic solids. As shown by Timoshenko and Goodier (1970) many plane problems in linear elasticity can be solved by finding a stress function i// that satisfies Eq. (12-5). The stress components obtained from Eqs. (12-4) must satisfy the particular boundary conditions of the problem. For the crack problem under consideration, Fig. 12-2, a complex stress function Z (z), where z — y + i y and i = -v/— 1, has been shown to provide a solution (Broek, 1982). The Westergaard (Ewalds and Wanhill, 1989) function is \j/ = Re Z + y Im Z
The constitutive relation is
'
az dZ
(12-7 a) (12-7 b)
12.2 Basic Fracture Mechanics Concepts
539
t t t t t t t t t
I I I I I I I I I It can be shown that Eq. (12-6) satisfies the biharmonic Eq. (12-5). The stress components obtained from Eqs. (12-4) are -ylmZ' ay = ReZ + ylmZ'
(12-8 a) (12-8 b)
^ v =
(12-8 c)
- y
where Z' = dZ/dz. For the cracked plate shown in Fig. 122, the stress function is (Broek, 1982)
f(z) Z(z) =
(12-9)
where / (z) is a real function that attains a constant value at the origin. In the limit (12-10)
Figure 12-2. Mode I crack under equal biaxial stress.
For plane stress oz = 0, and for plane strain oz = v (ax +ffy).KY is called the "stress intensity factor" for mode I loading. As r -• 0 the stresses become unbounded. KY is thus a measure of the stress singularity at the crack tip. For tension a at infinity, K{ is proportional to a. From dimensional arguments, KY must also be proportional to the square root of crack length. Most commonly KY is given by (12-12) Equations for the displacements can also be determined, see Broek (1982). Similar procedures can be used to achieve stress solutions for mode II and mode III crack problems (Broek, 1982; Ewalds and Wanhill, 1989).
/ 2 7TZ
In polar coordinates, z = re 10 , the asymptotic crack-tip stresses can be calculated from Eqs. (12-8) and (12-10) . e . 30
1 — sin - sin — 3 2 (12-11 a) KY
9 . 09 .. 3 0 cos -I 1 + sin - sin 2 2 (12-11 b)
. 9 9 36 K ^ ^ sin-cos-cos — 2 n r i l l
(12-11 c)
12.2.2 Stress-Intensity Factors for Anisotropic Solids The complex stress potential theory thoroughly described by Lekhnitskii (1981) is used to find stress solutions (Sih et al., 1965; Sih and Liebowitz, 1968) for cracked anisotropic solids. The off-axis constitutive relation for an orthotropic solid under plane stress is (12-13)
540
12 Fracture of Fiber Composites
where the overbars denote off-axis properties. For orthotropic materials with the principal coordinate directions in the x ysystem (on-axis), the shear coupling terms S16 = S26 = 0, and
where the elements btj are
s -J- s -
0
^11 — r- '
b22
r-
"21
^11 ^33 ~~ ^13
(12-14)
(12-19b)
bl2 = -
S33
'12
where E1 and E2 are the principal Young's moduli in the fiber (1) and transverse (2) directions, respectively, G12 is the shear modulus in the 1 —2 plane, and v12 and v 21 are the major and minor Poisson's ratios. Relations between on-axis and off-axis properties are provided by Jones (1975), Lekhnitskii (1981), and Tsai and Hahn (1980). Plane Strain For thick sections (large dimension along the z-axis), plane strain conditions are generally assumed
(12-19a)
S33
°12 —
s -J- s - J -
b26
c _ e »°33 °13 (
(12-19d) (12-19e) °33
S 3 3 -i
(12-19f)
For the case where the xy-system coincides with the principal material axes (onaxis)
(12-20a)
(12-15)
= lyz = 7x2 =
(12-20b)
For this situation, a normal stress oz will develop in the thickness direction:
The constants btj are given by
^=_(S^ax
b.j^Sij-^j^1
+ S23Gy + S36xxy)
(12_16)
33
b6e = S66
where expressions for S 13 , S 23 , and S36 are provided by Lekhnitskii (1981). For the on-axis situation — v3 1
— V13
(12-19c)
(ij = l,2) (i=j = 6)
(12-21 a) (12-21 b)
As recognized by Sih et al. (1965) and Sih and Liebowitz (1968), Eqs. (12-13) and (12-18) are of the same form, and the solution of a plane-stress problem can thus be transformed to the corresponding plane strain problem by replacing Stj with btj.
(12-17) Stress Analysis for Plane Problems The off-axis constitutive relation for plane strain is:
With the stress function \\i (x, y) defined in exactly the same manner as for isotropic materials, Eqs. (12-4), and using the consti-
12.2 Basic Fracture Mechanics Concepts
tutive relation, Eq. (12-13), the governing differential equation for two-dimensional anisotropic materials is obtained as S ^2
8x4
b± 4= b2. The stress function thus be expressed as
541 may
(12-26)
' 8x3 8y
where zx = x + fi1 y, z2 = x + jn2 y9 and \//1 '8x 2 8y 2
-
(12-22)
'8x8y 3
It can be easily verified that this equation reduces to the biharmonic equation, V4i/f = 0, for isotropic materials. The solutions of Eq. (12-22) are most conveniently expressed in terms of the complex variable (Lekhnitskii, 1981): (12-23) z = x jay where ji = a + i b, where a and b are real numbers and i = ^/ — 1. ja is obtained from the characteristic equation =0
(12-24)
This can be readily verified by recognizing that I,IV
and \j/2 are analytical functions of z± and z 2 , respectively. Because the solutions appear in terms of conjugate pairs, it is possible to express the stress function as (A(x,};) = 2Re[iA1(z1) + ^ 2 (z 2 )]
(12-27)
For isotropic materials both roots ji1 and \i2 are equal to i (Williams, 1989) and the stress function becomes (12-28)
iA (x, y) = 2 Re [z
It is common to introduce two new functions (j> and F A,l,
A,I,
dZl'
dz,
(12-29)
Combination of Eqs. (12-4), (12-28), and (12-29) yields ux =
(12-30 a)
a =
(12-30b)
(12-25 a)
IV
(12-25 b)
and so on. Substitution of these results into Eq. (12-22) yields Eq. (12-24). It has been shown (Lekhnitskii, 1981) that the four roots to Eq. (12-24) are either complex or purely imaginary and occur as two conjugate pairs:
(12-30c) Integration of the strain-displacement relations [Eqs. (12-2)] yields the displacement components u and v: (12-31 a)
]
(12-31 b)
where
+
where overbars here denote the complex conjugate (e.g., ju3 = a± —ibj and al9 a2, bl9 b2 are positive real numbers and
}
ri2
_
(12-32)
542
12 Fracture of Fiber Composites
is satisfied provided that Thus, the plane problem of anisotropic elasticity is reduced to identifying the two 82w 62w functions <j>{z^) and F(z2) that satisfy the boundary conditions specified. Longitudinal Shear Longitudinal shear of cylindrical bodies possessing at least one plane of symmetry normal to the generators, which are parallel to the z-axis, see Fig. 12-3, has been treated by Sih et al. (1965) and Sih and Liebowitz (1968). This mode of loading corresponds to the action of loads directed along the generators of the cylinder and has been denoted "antiplane" loading (Sih et al, 1965; Sih and Liebowitz, 1968). The elastic displacements, strains and shear stresses in the cylinder are given by
(12-37) 82w
^8/
=0
The solution of Eq. (12-37) is again conveniently expressed in terms of the complex variable z = x + \i y. \i is obtained from the characteristic equation C
ii2 A-1C
u A- C
—0
M?-^
Because w must be real, it follows that the two roots \x 5 and ji6 are (12-39) where
(12-40)
~
45
->2
•-7^^
•^45
-^44
U = V
= o,
: = 7Xy
8X = £ T
yz
=
w -= w(x,y) = 0
C44y yz + Q 5 yxz
(12-33) (12-34 a) (12-34 b)
where the C{j are the stiffness elements (Lekhnitskii, 1981) and the shear strains are 9w
9w
8?
(12-36)
dy
w(x, y) = U3 (x + M5 y) + U3 (x + fi5 y) = = 2Re[t73(z3)]
(12-41)
For convenience, a new function W(x,y) is introduced:
(12-35)
[cf. Eq. (12-33)]. The equilibrium condition dx
The displacement w can be expressed in an arbitrary function of the complex variable z 3 , that is,
W'(z3) = i
- C
Generators
?xz
=
\
-2Re[pi6W(z 2lm[W(z3)]
Ic c - r2
/^44L55
Figure 12-3. Longitudinal shear of a cylindrical body.
5
^
(12-42)
With this definition, the stresses and displacement become 2Re[Pr (z3)]
ay
2
3)]
(12-43 a) (12-43 b) (12-43 c)
^45
In contrast to the plane problem previously discussed, longitudinal shear requires only one function, W(z3), that satisfies the boundary conditions prescribed.
543
12.2 Basic Fracture Mechanics Concepts
Solutions for Crack Problems
II —
For a central crack in an infinite sheet as shown in Fig. 12-2, but with the origin of the x j/-system placed at the center of the crack, any point is defined as = rsin6
(12-44)
As shown previously, the stresses and the displacements are given by three functions:
KTTI =
T
yz
na
(12-46b)
na
(12-46c)
Based on the above stress functions, Sih etal. (Sih etal., 1965; Sih and Liebowitz, 1968) derived expressions for the stresses corresponding to mode I, mode II, and mode III loadings of single crack geometries. For mode I loading the near tip stresses are ax=
(12-47 a)
sin 6) F'(z2) =
O(r1/2) "(cos 9 + fi2sm9)
(12-47 b)
(12-45 b) Hi V^/cos 9 + ji2 sin 9
0{r112) ' (cos 9 -\- ji6 sin 9)
Hi
(12-45c)
The constants X{ are dependent on the stress intensity factors Kl9 Ku, Km (Sih etal, 1965; Sih and Liebowitz, 1968), which for single crack geometries loaded with the remote stresses of, %™y, and T*Z, are (12-46 a)
Figure 12-4. Anisotropic plate with crack at an angle cc to the loading direction.
(12-47c) '2nr
\_Hi~Hi "•
^/cos 9 + fi2 sin 0/ Expressions for stresses and displacements for this and the other modes of crack loading are provided by Sih et al. (1965) and Sih and Liebowitz (1968). Note that the authors defined stress intensity factors without including n inside the square roots in Eqs. (12-46). Here n is included in the definition of the stress intensity factors in order to retain the conventions of isotropic fracture. Inspection of the near crack-tip stress field shows that, similar to isotropic crack problems, the stress singularity is of the order of r~1/2. As an illustration, consider the anisotropic plate in Fig. 12-4 loaded by a remote
544
12 Fracture of Fiber Composites
tensile stress of magnitude p. The crack front may be considered as subjected to a far-field tensile stress of magnitude aO0=psin2a
(12-48)
and a far-field shear stress of magnitude Too
=p sin a cos a
(12-49)
The stress intensity factors KY and Kn become KY =/7sin 2 a v / 7ra
(12-50 a)
Ku = p sin a cos a ^ / n a
(12-50b)
where a is the semi-crack length. Notice that the stress intensity factors are the same for plane stress and plane strain. Wu (1967 and 1974) was probably the first to experimentally study fracture of anisotropic materials. Balsa wood and glass/ polyester plates with cracks aligned with the principal material directions (along the fiber directions) were fractured under combined tensile and shear loading. Figure 12-5 shows critical stress for onset of crack propagation plotted versus critical crack length (crack length at the onset of fast crack growth) for a glass/polyester composite. The linearity of the data with a slope of approximately —1/2 is experimen-
4.0
0.2
tal support that the critical stress intensity factors defined according to the concepts above [Eqs. (12-46)] are indeed constants.
0.3 0.4 0.5 0.6 0.8 1.0 1.5 Critical Half Crack Length, a c (in.)
Figure 12-5. Critical stress versus critical crack length at onset of crack propagation in glass/polyester composite (Wu, 1974).
12.2.3 Strain-Energy Release Rate In the application of fracture mechanics to anisotropic materials it has become customary to consider the rate of input of work (or strain-energy release rate) into the fracture process. The strain-energy release rate, G, is a quantity based on energy considerations. It is mathematically well defined and may be physically measurable in experiments. The energy approach, which stems from the original Griffith treatment (Griffith, 1921), is based on a thermodynamic criterion for fracture. One considers the mechanical work and strain energy of the system available for crack growth on one hand, and the energy consumed to extend an existing crack on the other hand. In brittle systems, Griffith found that the dominant source of energy consumption during crack propagation was the generation of additional crack surface energy. In composites, numerous energy absorbing micromechanisms contribute to the macroscopic fracture work: (i) formation of the fracture surfaces of the main crack, (ii) plastic deformation of the matrix in the cracktip region, (iii) microcracking and secondary cracks, (iv) crack bridging by fibers resulting in peeling and fracture. The strain energy release rate, G, is conveniently defined in terms of a potential energy, H, which quantifies the energy available for the formation of new crack surfaces: H= W- U
(12-51)
where W is the work supplied by the movement of the external forces, and 17 is the elastic-strain energy stored in the body.
12.2 Basic Fracture Mechanics Concepts
545
Many fracture specimens employed for experimental fracture characterization are of constant width, w. The amount of new crack surface is 5^4 = w 8a, where 5a is the crack extension that is measured in fracture testing. The energy release rate becomes w 9a Figure 12-6. Cracked plate of thickness w, with load P and displacement u.
The energy release rate becomes
where dA is the increase in crack area. If Gc is the work required to create a unit crack area, a criterion for crack growth may be stated as 5H > Gc 5 A
(12-53 a)
or equivalently, G>GC
(12-53 b)
Critical conditions occur when the net energy supplied just balances the energy required, that is, bH = Gc5A
(12-54)
The equilibrium becomes unstable when the net energy supplied exceeds the work of fracture (crack growth resistance)
8H>Gn8A
(12-55)
Unstable crack growth is thus expected to occur when 8G
As discussed elsewhere (Broek, 1982; Carlsson and Pipes, 1987), it is possible to determine G from the compliance of a cracked body. The compliance, C, Fig. 12-6, is defined as C=^
(12-52)
(12-56)
if Gc is independent of crack extension.
(12-57)
(12-58)
where P is the applied load and u is the displacement of the point of load application, see Fig. 12-6. The energy release rate becomes (12-59) G= — — 2w da Equation (12-59) is very convenient for analytical and experimental fracture studies of composites because it relies on the global specimen compliance. The alternative stress intensity approach previously discussed requires quite elaborate stress analysis. For many fracture specimens the compliance C can be derived from strength of materials principles. Alternatively, C can be measured experimentally at various crack lengths and the C versus a relation can be differentiated numerically to obtain the energy release rate G from Eq. (12-59). This is essentially the basis for experimental compliance calibration methods. Fracture Mode Separation As for the stress intensity factors Kl9 Klu and Km, it is possible to separate G into
546
12 Fracture of Fiber Composites
K2
three components: G = GY + G n + Gm
Theoretically, the mode separation is based on Irwin's contention that if the crack extends by a small amount, Aa, the energy absorbed in the process is equal to the work required to close the crack to its original length (Irwin, 1958). For the three basic modes shown in Fig. 12-1, this method (12-61 a) = lim —— Aaf oyx (Aa — r, 0) t; (r, n) dr i, Aa^o2Aa 1
G n = lim —-— | txy(Aa
(12-61 b) — r,O)u(r,n)dr
Aa -> 0 2. Afl o !
Aa
(12"61C)
G i n = lim — - J T yz (Aa-r,0)w(r,7i)dr where r is the radial distance from the crack tip, cy,xxy, and iyz are normal and shear stresses ahead of the crack tip, and v, w, and w are the relative opening and sliding displacements between points on the crack faces behind the crack tip. As pointed out by Sih et al. (1965) and Sih and Liebowitz (1968), calculation of the components Gl5 Gn, and Gm of the energy release rate, according to Eqs. (12-61) is somewhat academic because cracks in anisotropic materials do not generally extend in a planar fashion. If the crack plane, however, coincides with the direction of minimum crack resistance, the direction of crack propagation will be collinear with the original crack, and the computed values of G,, G n , Gm are physically meaningful. Substitution of the crack-tip stresses and displacements into Eqs. (12-61) yields relationships between energy release rate components and stress intensity factors (Sih et al., 1965; Sih and Liebowitz, 1968): ~2'
(12-62b)
(12-60)
(12-62 a)
KUI Im [C 45 + JX6 C 44 ]
(12-62c)
Equations (12-62) were derived under the assumption that each fracture mode can be analyzed separately, that is, in the absence of the other two modes (Sih et al., 1965; Sih and Liebowitz, 1968). If more than one mode is present at the crack tip, cross product terms will appear in the expressions for the components of the energy release rate. For orthotropic materials with the crack plane coinciding with one of the planes of material symmetry, the three fracture modes are independent and Eqs. (12-62) become 1/2
G, = Kt
'11 °22
(12-63 a)
1/2
11/2
H/2
(12-63 b) G
m =
r
W2
(12-63c)
where C 4 4 and C55 are shear stiffnesses defined by = Z
CAr4ryy
(12-64 a) (12-64b)
Equations (12-63 a) and (12-63 b) for Gx and Gn are defined for plane stress. For plane strain, substitute btj given by Eqs. (12-21) into Eqs. (12-63). It may be readily verified that Eq. (12-63 a) applied to a mode I crack in an
547
12.2 Basic Fracture Mechanics Concepts
Table 12-1. Material properties of graphite-PEEK and glass-epoxy composites (Salpekar et al., 1988).
isotropic materials yields G,= G,=
E (l-v2)K,2
(plane stress)
(12-65 a)
(plane strain) (12-65 b)
where E is Young's modulus and v is Poisson's ratio. Similar results are obtained for mode II and mode III. These are the classical relations between stress intensities and energy release rates for isotropic materials (Broek, 1982). To illustrate the difference between plane stress and plane strain in an orthotropic material, Eq. (12-63 b) is applied to a highly anisotropic graphite-PEEK composite and a moderately anisotropic glass-epoxy composite under pure mode II loading (Salpekar, 1988). Material properties for the two composites are provided in Table 12-1. In the calculations of the plane stress and plane strain constants it is assumed that: v12 = v 23 = v 13 , G12 = G 13 , and E2 = E3. The material has the fiber direction aligned with the x-axis as shown in Fig. 12-1. The ratio between the energy release rates, Gn (plane stress)-G n (plane strain), is 1.016 for graphite-PEEK and 1.023 for glass-epoxy. This example shows that plane stress and plane strain yield similar results, and that the difference between plane stress and plane strain diminishes with increasing anisotropy (EJE2) ratio.
(a)
(b)
Material AS4-PEEK S2-SP 250-epoxy
Et
E3
G 13
(GPa)
(GPa)
(GPa)
146 43.5
10.3 17.2
4.62 4.14
0.37 0.25
Finite-Element Crack-Closure Method The finite-element method is widely used to analyze fracture mechanics problems. A commonly employed technique to separate the components Gl9 Gn, and Gm of the energy release rate has been proposed by Rybicki and Kanninen (1977). This technique is called "the crack-closure method" and constitutes a numerical solution of Eqs. (12-61). Figure 12-7 illustrates the crack-closure method applied to a finite-element mesh of a mixed mode I and mode II problem. A certain crack extension Aa may be introduced in the finite-element mesh by releasing duplicate nodes at the crack tip. The resulting deformations are the relative opening and sliding crack-tip deformations. By applying nodal forces to the released nodes in two orthogonal directions, it is possible to close the crack tip to its original state. The products of crack-tip nodal displacements and forces enable the work of fracture to be evaluated numeri-
(c)
Figure 12-7. The crack-closure technique in finite elements. FH and F v are horizontal and vertical forces applied to close the crack, (a) Original configuration, (b) configuration after release of crack-tip nodes, (c) the released nodes are brought back to their initial position by application of nodal forces.
548
12 Fracture of Fiber Composites
cally. For the plane situation illustrated in Fig. 12-7, GY and Gn become =
FvAv 2wAa 2 wAa
(12-66a) (12-66 b)
where FH and Fy are the horizontal and vertical crack-closure forces and Aw and At; are the horizontal and vertical increments of displacements required to bring the released nodes to their original positions. Notice that the far-field load applied is constant in the steps illustrated in Figs. 12-7 a through c. For mode III, Gm is similarly obtained as FTAw 2wAa
(12-67)
where JFT is the magnitude of the crackclosure forces applied in the z-direction (Fig. 12-1), and Aw is the z-directional incremental displacement required to close the crack. Implicit in the derivation of Eqs. (12-66) and (12-67) is the requirement that the elements enclosing the crack-tip are identical in size. Consequently, Aa is equal to the distance between adjacent nodes. This constraint can be relaxed, but Eqs. (12-66) and (12-67) must be modified to evaluate the crack closure forces accurately, see Rybicki and Kanninen (1977) for details. Further details on the finite-element crack-closure technique as applied to anisotropic fracture problems can be found in Wang and Crossman (1980) and Raju (1987).
It should be observed that the released nodes are assumed to displace in opposite directions with the same magnitude (antisymmetry), see Fig. 12-8. This is the situation encountered in symmetric crack geometries and specimens. For asymmetric crack geometries or cracks between dissimilar materials, the displacements of the released nodes may no longer be of equal magnitude. 12.2.4 Mixed-Mode Fracture Mechanics Cracks in isotropic materials tend to grow under mode I conditions (Sih and Liebowitz, 1968). Due to the presence of weak planes between the layers of a composite laminate, however, delaminations often grow under the influence of a mixed mode stress field. In a mixed mode, both normal and shear stresses act across the interface ahead of the crack tip, and both opening and sliding displacements occur on the crack faces behind the crack tip. In two-dimensional crack problems involving mixed mode I and mode II loading, fracture toughness characterization must account for the relative amounts of mode I and mode II loadings. Most analysis of mixed-mode crack problems employ finite elements to compute the energy-release rate components based on nodal forces and displacements [e.g., Eqs. (12-66) and (12-67)]. In order to obtain accurate results with finite-element solutions to crack problems, highly refined meshes around the crack tip are required. For each new crack geometry or change in material
Figure 12-8. Displacements behind the crack tip in symmetric specimens; vl = v2 and w1 = u 2 .
12.2 Basic Fracture Mechanics Concepts
549
M1
Figure 12-9. Beam-type fracture specimen subjected to bending moments.
M o = M1 + M2
properties, a new finite-element solution is necessary. To circumvent this inconvenience, analytical schemes for fracturemode separation in beamlike specimens have been recently proposed by Williams (1988 and 1989), Suo and Hutchinson (1990) and Suo (1990 b).
Global-Mode Partitioning Method Williams (1988 and 1989) proposed a scheme for determining G and the components Gj and Gn for fracture specimens without the need of a detailed analysis of the crack-tip stress and displacment fields. This analysis is termed the "global mode partitioning scheme" (GMPS). A beamlike specimen of total thickness 2 h with a crack located a distance h1 from the top surface and a distance h2 from the bottom surface is considered (see Fig. 12-9). Initially, it is assumed that only bending moments are acting on the cracked beam, see Fig. 12-9. The definition of the energy
release rate, Eq. (12-57) is _ 1 w \ da
(12-68)
da /
To calculate the work performed by the applied moments, consider the cross-sections AB and CD at crack length a and a + 5a, respectively, in Fig. 12-10. The moments will perform work when rotating the cross-sections. If the crack tip is originally at 0 on AB and extends to 0' on CD, the changes in slopes in the upper and lower beams are, according to Fig. 12-10, (upper beam)
da
d0 o da
da
8a (lower beam)
Consequently,
d
\da
da da
5a
(12-69)
Crack
Figure 12-10. Rotations of cross-sections before and after crack growth.
550
12 Fracture of Fiber Composites
From classical theory of beam bending (Gere and Timoshenko, 1984), the magnitude of the change in slope (beam curvature) is given by d(f) _M
(12-70)
dx " " £ /
where E is the axial modulus of the beam and / is the moment of inertia of the crosssection. For the beam shown in Figs. 12-9 and 12-10 M1 + M 2 da
~d~a~ d^2 da
w{2hf 12
(12-71 a)
wh\ 12
(12-71 b)
wh\
Mi
=
(12-71 c)
Ex 12
Introducing a dimensionless parameter 3
Z = h1/{2h) and I = wh /12,
M2
1
M2f
(12-72)
The strain energy per unit length stored in a beam is given by Gere and Timoshenko (1984) as dU dx
Mz :IEI
(12-73)
For the beam under consideration, the rate of strain energy change is (i? 74 a ) dU da or
M\ :IE1I1
dU da
1
leE^
J
M\ ' 2E1I2
YM\
- (M, + M 2 ) 2
(12-75)
This equation allows for the computation of G based only on local (crack tip) quantities (M\ and M2) and eliminates the need to compute the global specimen compliance. Axial loading of the beam is also considered by Williams (1988 and 1989). If, instead of the moments M x and M 2 , axial loads Px and P2 act on the neutral plane of each beam in Figs. 12-9 and 12-10, a similar analysis as above yields G=
4w2E1h
the rate of
work change is dW "da"
Combination of Eqs. (12-68), (12-72), and (12-74) yields
(M1 + •M2f 2 £ Jo
Ml
-a3
- ( M , 4 •M,)21
(12-74 b)
Notice that the load Po = Px + P2 is applied from the at a distance h0 = hx + h2 bottom surface in order to maintain moment balance. In structural applications of composite materials, cracks will most likely propagate under mixed-mode stress fields. The most commonly studied fracture mode in composite materials is delamination, or interlaminar fracture, where crack propagation occurs in the resin-rich region between the plies in a composite laminate (Wilkins et al., 1982). Studies of delamination propagation under various modes have shown that the critical values of G are significantly different for the various fracture modes, (e.g., Hashemi, 1990; Johnson and Mangalgiri, 1987; Jordan and Bradley, 1987; Russell and Street, 1985). This is illustrated by the data in Table 12-2, which lists critical values of G determined in pure mode I, mode II, and mixed-mode I/II delamination tests. It is, therefore, necessary to sep-
551
12.2 Basic Fracture Mechanics Concepts
Table 12-2. Critical values of the strain-energy release rate in kJ/m 2 for toughened graphite-epoxy composites (Jordan and Bradley, 1987). Mode Mode I 20% mode II 43% mode II Mode II
AS4-3502
T6T145-F155
T6T145-HX205
T6T145-F185
0.189 0.264
0.520 0.525 0.548 1.27
0.455 0.796 0.789 1.05
2.21
0.570
arate (partition) the total G into its components, Gl9 Gn, and Gm. For the plane situation considered Gm = 0
(12-77 a)
Mn =
M2 + M1 1 +a
(12-77 b)
When both modes are present, a suitable mixed mode fracture criterion has to be established by testing. This will be discussed further in Sec. 12.4.3. Williams (1988 and 1989) states the following condition for pure mode II loading:
(12-80b)
Substitution of these expressions in Eq. (12-75) gives
m
Gu
2.44
(1 + a)
Mf
a—
_aMj)2
1
(12-81 a)
r(i 3 MS
c
16w£lJ
r -M 2 ) 2
#2
da
da
that is, equal curvatures of the beams. This condition will be fulfilled if a moment Mn is applied to the upper beam and a moment oc Mu on the lower beam so that
EJ2
(12-81 b)
(pure mode II)
or a =
Lzi
Axial loads, according to Williams (1988 and 1989), correspond to pure shear (mode II). A similar analysis as above applied to the axially loaded situation yields Pi
(12-78)
Pure mode I, according to Williams (1988 and 1989), corresponds to opposite moments of the same magnitude, that is, — M, on the upper beam, and MY on the lower beam, see Fig. 12-11. Superposition of the two cases in Fig. 12-11 gives
=
Pi +
(12-82 a)
Pu
P2 = {~r®Pl + Pll
(12-82b)
n
aM T
±J (a) Mode II
(12-79 a) M2 = a M n H
(12-79 b)
or
(b) Mode I
M7-
(12-80 a)
Figure 12-11. Moments corresponding to (a) pure mode II, and (b) pure mode I.
552
12 Fracture of Fiber Composites
For the ENF test, Fig. 12-12b,
and substitution into Eq. (12-76) yields G, = 0
(12-83 a)
G,,=
(12-83 b)
(12-86 a) £ = 0.5, a = 1.0
(12-86b)
and from Eqs. (12-81) Examples
G, = 0
The global mode partitioning scheme will be applied on some commonly employed fracture specimens, namely the double cantilever beam (DCB), the endnotch flexure (ENF) and the end-loaded split (ELS) specimens illustrated in Fig. 12-12. These specimens were introduced to characterize delamination fracture by Bascom et al. (1980) and Wilkins et al. (1982) (DCB specimen), Russell and Street (1982) (ENF specimen), and Hashemi et al. (1990) (ELS specimen). For the DCB test shown in Fig. 12-12 a t= -M2=
-Pa
(12-84a)
(12-87 a) 2
G,,=
3P a
2
(12-87 b)
which agrees with results derived from classical beam theory (Russell and Street, 1982). As pointed out by Williams (1989), the GMPS yields Gl = 0 even if the crack is located asymmetrically (h1 # h2). For the ELS test, shown in Fig. 12-12 c, M± =
(12-88 a)
-Pa
(12-88 b)
M2 =
Substitution of Eqs. (12-88) into Eq. (12-75) yields
(12-84b)
= 0.5, a = 1.0
Substitution into Eqs. (12-81) yields GT =
P2a2 wEI
(12-85 a) (12-85 b)
Equations (12-81) yield G, =
Notice that this analysis yields Gu = 0 even if the crack is located asymmetrically
P/2
ENF (Mode II) (a)
i
h h
•pt DCB (Mode I) (b)
(Pa a)2
(12-90 a) (12-90b)
P/2 P
1
\ h2
ELS (Mixed Mode) (c)
Figure 12-12. Fracture specimens for composite materials, (a) ENF specimen (mode II), (b) DCB specimen (mode I), (c) ELS specimen (mixed mode).
12.2 Basic Fracture Mechanics Concepts
553
Local Mode Partitioning Method
1.0
Figure 12-13. Mode mix (Gj/G) versus thickness ratio [h1/(2h)] for ELS specimen.
Consequently, both fracture modes are acting on the crack front. The fraction of the opening mode is 1
(12-91)
This equation shows that the fraction of GY is only dependent on the thickness ratio ?; = hj(2h). From an experimental point of view this is an advantage, because the ratio between the two modes does not vary with crack extension. The ratio Gx/G is plotted versus £ in Fig. 12-13. It is observed that the ELS test is pure mode I for near top surface cracks, and pure mode II as the crack plane approaches the bottom surface.
Alternative analyses of stress intensity factors and strain energy release rates for mixed mode I and mode II problems have been presented by Suo and Hutchinson (1990) and Suo (1990 b). Suo and Hutchinson (1990) deal with isotropic interfacial fracture, while Suo (1990 b) deals with cracked orthotropic materials, pertinent for this chapter. The superposition scheme shown in Fig. 12-14 is employed to construct the solution for an edge dislocation (introduced to represent the crack tip singularity) embedded in an infinite strip of orthotropic material. A solution without singularity in the strip, obtained by the Airy stress function i//{x,y), see Sec. 12.2.2 and Suo (1990), is superposed onto the solution for a strip containing an edge dislocation to account for tractions on the strip boundaries. The solution employs the complex potential method of Lekhnitskii (1981), see also Sec. 12.2.2, and Fourier integral equations. The analysis provides analytic expressions for the stress intensity factors with only one parameter undetermined. This parameter is obtained from numerical solutions to integral equations, and the results are tabulated by Suo (1990 b). The results are, despite the involved mathematical procedure, relatively
-2-*p3 M
Figure 12-14. Superposition scheme for reduction of variables.
554
12 Fracture of Fiber Composites
simple to apply, as will be shown by analysis of some fracture specimens. This analysis is termed the "local mode partitioning scheme" (LMPS). Analysis For a crack in an orthotropic body, where the crack plane is one of the principal planes of symmetry and the crack front is in a principal direction, the crack tip stresses are given asymptotically by (12-92 a) (12-92b) [cf. Eq. (12-47 b)]. The energy-release rate obtained from the crack-tip stresses and the displacements behind the crack tip [see Eqs. (12-61)] is
Kf
(12-93)
where b1± is given by Eq. (12-19 a), and n and X are material parameters given by
same for the beams in Fig. 12-14 a and 12-14 c. Static equilibrium of forces and moments yields M* = M -\-P(H +•h)/2 P = Pi- -CtP3- -C2M
(12-95 a) (12-95 b)
M = Mt -C3M
(12-95c)
where
22
Q)/2
(12-94b)
with '66 Q =
(12-94c)
These results are consistent with those derived by Sih et al. (1965), except that the stress intensities are defined including y/n, as in Eqs. (12-46). With consideration of the problem shown in Fig. 12-14, it is recognized that the beam shown in Fig. 12-14 b contains no crack. Consequently, the stress intensity factors and energy release rates must be the
6<
1
r C3 = Cl,
n=-
(12-96)
The solution is now focused on the beam shown in Fig. 12-14c. For this beam the energy release rate is
where (12-98 a)
A=
(12-98 b)
12
(12-94 a) y
3
Al
Hi+n)
(12-98 c)
where |y| < n/2 and h < H {t] < 1). By solving the two expressions for G, Eqs. (12-93) and (12-97), Suo (1990) found COS CO +
(12-99 a)
(12-99 b)
555
12.2 Basic Fracture Mechanics Concepts
where the dimensionless function co depends only on rj and Q, and not on A, that is, co = (D(Y],Q). The magnitude of co is restricted by: 0 < CD < n/2 to recover KY > 0 and Ku > 0 for the case P > 0 and M = 0. The function co {r\, Q) can be determined exactly for a symmetric DCB fracture specimen, Fig. 12-12 a with h = H, 17 = 1, by considering a special loading: P = 0 (P1 = p3 = M 3 = 0), and M = 1 (Suo, 1990 b). For this case Eqs. (12-98) give A = 1/14,
1 = 1/24,
Table 12-3. Values of a) (rj, Q) indegrees (Suo, 1990 b). \
-0.5
Q
1
2
3
4
52.1 50.9 49.1
52.2 51.1 49.1
52.2 51.1 49.1
52.3 51.7 49.1
\
51.0 51.7 50.4 50.4 49.1 49.1
0 0.5 1
Substitution of Eq. (12-101) into (12-102) yields M2
G,=
sin y = V 3/7
0
Eq.
(12-103)
with / = 1/24, and M = Pa/w9 GY becomes Because Kn = 0 for this specimen, Eq. (12-99 b) gives cos(co 4- y) = 0 with the solution co = 90° - sin" x y/3/7 « 49.1°. Consequently, co(l,g) = 49.1°, independent of g. Numerical results for the general case co(rj9g) are presented in Table 12-3. If the (weak) dependency of co{rj,g) on g is neglected (Suo and Hutchinson, 1990; Suo, 1990 b), co is approximated by co « 52.1 — 3f|
(in degrees)
(12-100)
G, =
12 P2 a2
(12-104)
h3w2E1
which is the classical beam theory result for the DCB specimen [cf. Eq. (12-85 a)]. For the ENF specimen (Fig. 12-12 b) Pi =P2 = Pi = 0
(12-105 a)
M j = — M 2 = P a/(4 w)
(12-105 b) (12-105 c)
Eqs. (12-95) yield
P = -C2M3/h= M =M1-C3M3
Examples The DCB, ENF, and ELS specimens shown in Fig. 12-12 are considered. For the DCB specimen, Fig. 12-12a, co(l9g) = 49.1°, from above. The angle co + y required for Eq. (12-99 a) is co + y = 49.1° + sin" x ^ 3 / 7 = 90° Consequently,
= M1-2C3M1 C2 = 3/4,
The strain energy release rate Gx = G is (for plane stress) given by Eq. (12-93)
C 3 = 1/8
(12-106 c)
Consequently, P = - 3MJ(2h), M = 3 Mi/4. Substitution into Eqs. (12-99) with
Ka =
- 4.2422 A
-
1/8
M1
(12-107 b)
The strain energy release rate Gu is (plane stress)
nK\ (12 102)
(12-106a) (12-106b)
K, = 0 (12-101)
&
-2C2MJh =
9 Mf
9P2a2 16WE!?!3
(12-108)
556
12 Fracture of Fiber Composites
which is the classical beam theory result, Eq. (12-87 b). The mixed mode ELS specimen, Fig. 12-12 c, may be analyzed in a similar manner as the DCB and ENF specimens. This has been performed by Hashemi et al. (1990), who also compared the GMPS with the LMPS. The results are shown in Table 12-4. Inspection of the results in Table 12-4 shows that the global and local mode partitioning schemes predict widely different results, except when the crack is in the center of the beam thickness. For example, for bending of a thin arm, hj(2h) -> 0, the LMPS predicts about 38% mode II while the GMPS predicts 100% mode I behavior for this configuration. The discrepancy between these methods is currently the subject of investigation by Williams (1990) and Hutchinson (1990). The global mode partitioning scheme appears to produce more consistent data from an experimental point of view (Hashemi, 1990). On the other hand, Hutchinson (1990) has provided examples on crack geometries where a beam analysis would not be adequate. It appears beyond doubt that the approach by Suo and Hutchinson (Suo, 1990 b; Suo and Hutchinson, 1990) is correct within the limits of linear elasticity. 12.2.5 Cracks at Bimaterial Interfaces
Figure 12-15 shows a bimaterial plate with a central crack subjected to uniform tension. Initially both materials are as-
t t t t t Material 1
Material 2
I I I I I Figure 12-15. Bimaterial plate with an interfacial crack.
sumed to be isotropic. Because the materials are joined at the interface (y = 0) shear stresses xxy develop along the bondline to maintain continuity in displacements along the interface, and this creates a local mixed mix at the crack tip. Analyses of the stress state near the crack tip (Williams, 1959; England, 1965; Erdogan, 1965; Erdogan and Gupta, 1971) show that the stresses share the inverse square-root singularity of homogeneous fracture [Eqs. (12-11) and (12-47)]. In addition, the stresses exhibit an oscillatory behavior as the crack tip is approached. The stresses can be expressed in a compact complex form as (Hutchinson, 1989):
oy + i r ^
+
l52)^
(12-109)
where K1 and K2 are the bimaterial interfacial stress intensity factors. Because rie = cos (s In r) + i sin (e In r), the stresses <jy and T will change sign infinitely often as
Table 12-4. GJG ratios for ELS specimens of various thickness ratios [hj(2h)] computed by global (Williams, 1988 and 1989) and local (Suo, 1990 b) mode partitioning schemes (Hashemi et al., 1990).
hj(2h)
0
0.2
0.4
0.5
0.6
0.8
1.0
GJG (global) GJG (local)
1.00 0.623
0.992 0.611
0.824 0.588
0.571 0.571
0.292 0.549
0.031 0.481
0 0.377
12.2 Basic Fracture Mechanics Concepts
r -• 0. Consequently, s is called the oscillation index (Hutchinson, 1989) and is given by (12-110) where /? is (12-111) where subscripts 1 and 2 refer to materials 1 and 2 in Fig. 12-15, x = (3- v)/(l + v) for plane stress and 3-4 v for plane strain (v is Poisson's ratio), and F = GJG29 where G is the shear modulus (Suo and Hutchinson, 1990; Dundurs, 1969). For plane strain the relative crack face displacements 8{ = u{(r, 6 = n)-u{(r,0 (u1 = u and u2 = v)
= -n)
are given by Hutchinson (1989): \-v\
Sy-idx =
l-v\
+ 2 i e) cosh (7i e) (12-112)
For plane stress, replace the factors (1 - vf)/E{ in Eq. (12-112) with l/£ s . Consequently, the displacements also show oscillatory behavior as the crack tip is approached. In fact, Eq. (12-112) indicates crack face interpenetration behind the crack tip, which is physically inadmissible. As pointed out by several investigators, (England, 1965; Erdogan, 1965; Erdogan and Gupta, 1971; Hutchinson, 1989; Rice, 1988; Williams, 1959), the contact region is generally exceedingly small compared to the plastic zone or fracture-process zone, and may therefore be disregarded. K1 and K 2 , however, cannot be directly linked to the remote normal and shear stresses and
557
the dimensional form of a complex quantity (K1, K2) will cause the relative proportion of Kx and K2 to change when the units of stress and crack length are changed (Hutchinson, 1989; Rice, 1988). In order to circumvent these complicating features of interfacial fracture mechanics, Hutchinson (1989) and Suo and Hutchinson (1989) suggest that a rational approach is to take /? = 0 = e in the analysis and application of interfacial fracture mechanics. With e = 0, ri£ = l in Eqs. (12-109) and (12-112), the conventional meanings of K± and K2 will be retained. An interfacial crack between dissimilar anisotropic materials has been pursued by many authors (Ma and Luo; Raju et al., 1988; Sun and Manoharon, 1987; Suo, 1990; Ting, 1986 and 1990; Wang, 1983; Willis, 1971). These studies provide useful information on the structure of the near-tip fields, analogies with the isotropic-bimaterial interface problem, the oscillation index, stress intensity factors and displacement jumps across the crack faces. As mentioned in Sec. 12.2.3, it has become customary to approach damage tolerance of composite structures by computing the energy release rate, G, and subsequently separate it into its components. The most commonly applied approach utilizes finiteelement modeling of the cracked geometry and fracture-mode separation based on the crack-closure method discussed in Sec. 12.2.3. As will be discussed in this section, however, the analysis of cracks in composite structures is greatly complicated by the bimaterial nature of many crack geometries. A delamination, for example, is likely to exist between plies of different fiber orientation angles, which disrupts the symmetry in material properties across the crack plane and represents a "bimaterial" crack problem, even if the principal material properties of the plies are identical.
558
12 Fracture of Fiber Composites
Sun and Manoharan (1987) were perhaps the first to illustrate the complications that may arise in the strain-energy release rate analysis of cracks between orthotropic materials of different ply orientation. For two orthotropic plates joined together (Fig. 12-15), Sun and Manoharan (1987) calculated the energy release rate components GY and Gu for remote tensile and shear loadings: (Aa^-[2e CI.r(l+4e-2 t. F lim — ± 47\L2 cosh \^7^ (71 e) Aa -> 0 \ a Aa -> 0 \ a
Aa
C 11 = — \ —
-i2e
— -F! lim a 4 |_2cosh(7i£) " Aa -> 0 - is lim — Aa-
where i = 1,2 for tensile and shear loading, respectively. The constants Cl9 C2, Fl9 F2 are functions of the elastic constants of the materials and the magnitude of the applied loads, e is the oscillatory index, which is a function of the orthotropic elastic constants only. As the crack-closure distance (Aa) approaches zero, the exponential functions have no definite limits. This means that in the finite-element analysis, the computed values of Gj and Gn will depend on crack-tip element size and will not converge with decreasing element size. If the components Gx and Gn are added the result is 4 cosh (n e)
(12-114)
The total energy release rate will thus not oscillate as the crack closure distance approaches zero. To further investigate the influence of the oscillating singularity on the components of the energy-release rate, Sun and Manoharan (1987) performed
finite element studies of a bimaterial plate using different crack increments Aa. The results for tensile and shear loading are presented in Fig. 12-16. The results show that the total energy release rate remains unchanged as the size of the crack-tip elements decreases, which is indicative of a convergent numerical solution. The components Gr and Gn, however, do not converge with decreased crack increment. At larger crack increments (or crack-tip element sizes) constant values of Gj and Gn are approached. The nonconvergent behavior is apparently due to the oscillatory stress singularity. Raju et al. (1988) studied convergence of finite element solutions for free-edge delamination problems where the delamination crack is located between different ply angles (-35°/90°). They found that the total energy release rate converged to a constant value with decreasing size of the crack-tip elements. However, the individual components Gj and Gn did not con-
o.oo
0.06
Figure 12-16. Strain-energy release rate (arbitrary units) for bimaterial plate with interfacial crack under remote tension (top), and remote shear (bottom). Symbols •, •, and A represent G, GI? and Gn respectively (computed by finite element analysis). Solid line represents closed-form analytical solution (Sun and Manoharan, 1987).
12.2 Basic Fracture Mechanics Concepts
verge with decreasing element size. Raju et al. (1988) similarly showed that the oscillatory component of the singularity is the cause of the nonconvergent behavior of Gx and Gu. To further study the accuracy of the results they incorporated a thin (0.1 mm) layer of resin between the plies and located the crack centrally within this layer. Because the crack exists in a homogeneous (and isotropic) material there is no oscillatory component of the singularity, that is, the classical singularity of the type r -i/2 j s r e c o v e r e c j Strain energy release rates for the "bare interface" crack and the crack within the resin layer were computed by finite-element analysis (Raju et al., 1988) and are shown over a wide range of cracktip element sizes, A, normalized with the composite ply thickness, /z, in Fig. 12-17. Comparison of the results shows that the "bare interface" models converge to the resin layer case with increased crack-tip element size. Raju et al. (1988) recommend a minimum size of the crack-tip elements in the range h/4 to h/2. Based on the insight provided by the works discussed above, however, the partitioning of the energy release rate into tensile and shear components is associated with fundamental difficulties for bimaterial crack problems, and a stress intensity approach seems better justified on physical grounds.
559
12.2.6 Cracks in Bonded Structures
A recently introduced concept to improve the delamination resistance of laminated composites is to bond "interleave", thin layers of ductile polymer films to the plies in the laminate (Evans and Masters, 1987; Masters, 1987 a and b). As a consequence of the improved delamination toughness of the material, the impact strength has drastically improved (Evans and Masters, 1987; Masters, 1987 a and b). The reason for the improved delamination toughness appears to be that the ductile resin film allows for more expansion of the crack-tip yield zone between the plies (Aksoy and Carlsson, 1991). In composites without interleaves, the rigid elastic layers provide severe constraints to the expansion of the plastic and process zones around the crack tip (Bradley and Cohen, 1985; Friedrich et al., 1989; Hunston, 1984). The presence of an interlayer, however, alters the local stress field, and may complicate the data reduction procedure for the fracture toughness (Aksoy and Carlsson, 1992; Sela et al, 1989; Suo and Hutchinson, 1989). There are two main reasons for the alteration of the stress field. First, the crack commonly propagates at the film-composite interface, see Fig. 1218, which renders the situation of a bimate-
0.06
0.05
Results of the resin layer model "bare" interface model / Total
QU-—o-o
0.04
0.03
0.02
0.01
Mode III
0.00
log (A/h)
Figure 12-17. Strain-energy release rates for interfacial crack and for a crack within a thin resin layer between the plies in a composite laminate as a function of crack-tip element size (Raju et al., 1988). e0: applied axial strain, t: total laminate thickness, and E1: major Young's modulus of a ply.
560
12 Fracture of Fiber Composites
where p = V(l - a)/(l - ,
(12-117)
in which a = •
Figure 12-18. Interface crack in bonded structure.
rial fracture problem with an oscillatingstress singularity as discussed in Sec. 12.2.5. Second, the asymmetry of the geometry of interleaved specimens may lead to a local mixed-mode crack loading situation, even if the far-field loading is symmetric. The local mixed-mode phenomenon has been termed "phase-shifting" of the stress field by Suo and Hutchinson (1989). The local stress field for test specimens with isotropic constituents has been investigated (Suo and Hutchinson, 1989). The authors considered a semi-infinite crack problem and derived asymptotic (layer thickness -> 0) expressions for the near-tip stresses and displacement fields based on an integral equation formulation for an edge dislocation located at the interface. The far field loading is expressed by the stress intensity factors K{ and Kn that would be present in a homogeneous specimen (without an interlayer) K
= /Cj + 1 Ku
(12-118)
P is given by Eq. (12-111), and r = GJG2 where G is the shear modulus, x is defined in connection with Eq. (12-111). The function co (a, jS) is obtained from a numerical solution of the integral equations for the interface crack problem and is tabulated by Suo and Hutchinson (1989). The interfacial stress intensity is thus scaled by a factor, /?, and phase shifted by an angle, co, with respect to the far field stress intensities, see Fig. 12-19 (notice that the factor tj~iB has magnitude of unity). For material combinations where ft vanishes (and thus s = 0), the stress intensities K1 and K2 can be interpreted in terms of the crack opening and forward shearing modes of crack propagation, and Eq. (12-116) becomes ,.*
K1 + \K2 =
iKu)eic
The phase shift o>, see Fig. 12-19, means that, for example, a mode I specimen with an interlayer will not remain mode I at the crack tip. As shown by Suo and Hutchinson (1989), however, the amount of shift will not be large.
(lz-lljj
For the interface crack shown in Fig. 1218, the interfacial stress intensities are denoted by Kx and K2. The stresses oy and Txy are given by Eq. (12-109). By energy and dimensional arguments, Suo and Hutchinson (1989) derived a relationship between the far field and interfacial stress intensities (12-116)
Figure 12-19. Relation between far field and interfacial stress intensities.
12.3 Fracture Specimens
It is expected that analysis of crack propagation in interleaved composite specimens with anisotropic or orthotropic constituents would also be complicated by the issues discussed above. To date, no analytical results have been presented for thin bondlines between anisotropic or orthotropic materials, but numerical studies provide some insight, see Sec. 12.3.4.
12.3 Fracture Specimens In this section various contemporary fracture specimens employed for composite materials will be described. Most of the attention in the evaluation of damage tolerance of advanced laminated composite structures has been devoted to delamination failure, because in these structures there are no reinforcing fibers holding the plies together. In particular, hard object impact loading has been a much studied loading case for quality assurance of composite structures; this loading case typically results in delamination (ASTM STP 936, 1987; Bostaph and Elber, 1982; Dost et al, 1991; Guynn and O'Brien, 1985; Masters, 1985, 1987a and 1987b; Wilkins, 1983). The description is limited to plane fracture specimens (mode I, mode II, and mixed mode I and II). A discussion of mode III specimens for composite materials is found in Gillespie Jr. and Carlsson (1990). Mode I, mode II, and mixed mode composite fracture specimens are also discussed by Carlsson and Pipes (1987), Friedrich (1989), Gillespie Jr. and Carlsson (1990), and Wilson and Carlsson (1991). These references provide information on specimen design, testing and data reduction methodology. The description presented herein will therefore be relatively brief.
561
Because the strain energy release rate, G, is experimentally accessible for many of the beam-like fracture specimens employed for fracture characterization of composite materials, most fracture studies focus on this parameter. The presentation in this section essentially follows this convention. 12.3.1 Double Cantilever Beam Specimen
The double cantilever beam (DCB) specimen was discussed in Sec. 12.2.4. The geometry and loading of the DCB specimen is shown schematically in Fig. 12-12. A more comprehensive illustration is provided in Fig. 12-20. Commonly a 24 ply unidirectional composite is split in the center, end tabs are attached, and the specimen is loaded in a testing machine with a sensitive load cell. The precrack is achieved by placing a thin inert film, typically 1 5 75 |im in thickness at the mid-plane of the laminate prior to processing. The choice of material and thickness for the insert film has been the subject of an ASTM committee on interlaminar fracture testing (ASTM D30.06) and several publications, for example, Martin (1988). Procedures for load tab attachment and specimen preconditioning are discussed by Carlsson and Pipes (1987) and Gillespie Jr. and Carlsson (1990). The crack may be grown continuously, as suggested in the preliminary ASTM procedure (Proposed Standard Test
Figure 12-20. DCB test geometry.
562
12 Fracture of Fiber Composites
Method, ASTM, 1990), while the load-displacement response is recorded and crack length is monitored by inspection of the specimen edge. Alternatively, the crack may be grown incrementally while monitoring several loading-unloading increments (Wilkins et al., 1982). This procedure has the advantage over continuous loading of providing opportunity for more precise crack-length measurements (since the cross-head is stopped). Furthermore, information on crack initiation mechanisms is obtained for the entire load history, see also Ozdil and Carlsson (1992 a).
expressions for the energy-release rate of the DCB specimen discussed previously in Sec. 12.2.3 [Eqs. (12-85 a) and (12-104)] are strictly valid only for infinitely long crack lengths. End effects tend to decay at a slow rate in orthotropic specimens (Horgan, 1982; Carlsson et al., 1986 c) and long precrack lengths may be necessary to achieve satisfactory agreement with the infinite beam solutions. Finite-element calculations presented by Suo et al. (1991) and reviewed by Hutchinson and Suo (1991) yield the following calibration for finite geometry DCB specimens,
DCB Data Reduction
+
ASTM (Proposed Standard Test Method, ASTM, 1990) recommends adoption of the experimental compliance calibration method, based on the approach originally suggested by Berry (1963). In this approach the beam compliance, C, is expressed as
where G* is the asymptotic energy release rate for DCB specimens given by Eq. (12-104), 7 is a material-dependent factor, Y(Q) = 0.667 + 0.149 ( e - 1 ) -0.013(e-l)2
C=—
(12-120)
where a is the crack length, and the exponent n and factor H are experimentally determined parameters. Substitution of Eq. (12-120) in Eq. (12-59) yields the critical value of the strain energy release rate associated with a given crack length, nPG(a)6e(a) 2wa
(12-121)
in which Pc and Sc are the critical load and the displacement at the critical load, and w is the specimen width, see Fig. 12-20. The empirical basis for the Berry method is necessitated by the fact that the loaded beam ends are elastically built into the remainder of the specimen and that interlaminar shear deformations may affect the results for the finite geometry specimens typically employed in actual testing. The analytical
(12-122)
(12-123)
and Q and X are defined in Eqs. (12-94b and c). For plane stress k=^
(12-124a)
2G 12
(12-124b)
For DCB specimens with A1/4 a/h > 1 and 0 < Q < 5, Eq. (12-123) is accurate within 1% (SuoetaL, 1991). 12.3.2 End-Notched Flexure Specimen
Several mode II fracture specimens have been suggested for interlaminar fracture testing of composites, see Gillespie Jr. and Carlsson (1990). The most widely adopted mode II fracture specimen, however, is the so-callled end-notched flexure (ENF) specimen illustrated in Fig. 12-12. The ENF ge-
12.3 Fracture Specimens
ometry was originally proposed for mode II fracture studies of wood by Barrett and Foshi (1977), and was adopted for delamination studies of composites by Russell and Street (1982). The ENF specimen employs a 24 ply unidirectional composite with a starter crack that is achieved by placing a thin inert film, 7 to 25 jim in thickness at the center of the layup. Because the ENF specimen is inherently unstable (Carlsson et al., 1986 a), precracking becomes an important issue. Initially, the proposed ASTM standard (Test Procedure for the ENF Specimen, ASTM D 30.02, 1987) suggested mode I precracking, but more recent experimental evidence and physical arguments (O'Brien et al., 1987) favor mode II precracks or very thin films with no precracking over the tensile precrack. The issue is still under investigation (Minutes of the ASTM D 30.06 subcommittee). The fracture test uses a conventional 3point flexure test fixture with a total span, 2 L, of 50.8 mm between the outer specimen support pins, see Fig. 12-12. The crack length, a, is 25.4 mm (a/L = 0.5) so that the crack tip is positioned midway between the outer support pin and midspan-load point. The beam deflection, 5, should be monitored at the midspan (beneath the load point) by an extensometer or a LVDT (linear variable differential transformer) or from the cross-head displacement if corrections for machine compliance are made. Further experimental details are provided in the preliminary ASTM test procedure (Test Procedure for the ENF Specimen, ASTM D 30.02, 1987) and by Carlsson and Pipes (1987) and Gillespie Jr. and Carlsson (1990). After the crack propagates and is arrested, the fracture test is completed. The compliance and critical load are employed in the data reduction for the fracture toughness and should be recorded.
563
ENF Data Reduction The data reduction scheme presented here is essentially the one recommended by the preliminary ASTM standard (Test Procedure for the ENF Specimen, ASTM D 30.02,1987). The ENF beam compliance is given by the following expression (Carlsson et al., 1986 a; Carlsson and Pipes, 1987; Gillespie Jr. and Carlsson, 1990), 1.2 L +0.9 a 4whG 13
(12-125)
where L is the half span length, a is the crack length, E1 is the axial Young's modulus of the beam, w is the beam width, h is the semi-thickness of the beam (Fig. 12-12) and G13 is the interlaminar shear modulus. Substitution of Eq. (12-125) in the expression for G, Eq. (12-59), yields after some algebraic manipulations, the mode II fracture toughness, 9 a2 Pc2 (C - Cs*) 4 w L 3 [ l + 1.5 (a/L) 3 ]
(12-126)
in which Pc is the critical load for propagating the crack and Cf is a modified shear compliance given by C.* = -
0.6a-0.2L 3 /a 2
20whG 13
(12-127)
For experimental fracture toughness determination, C in Eq. (12-126) represents the experimentally determined ENF compliance. Equation (12-126) represents an expression valid for the finite width geometries employed in fracture testing (Salpekar, 1988). The expressions (12-87b) and (12-108) are infinite beam solutions strictly valid only for long crack lengths and beams with high shear modulus G 13 . An alternative expression for G that applies to finite geometry orthotropic ENF
564
12 Fracture of Fiber Composites
specimens (Hutchinson and Suo, 1991) is (12-128) where G^ is the asymptotic energy release rate for the ENF specimen, Eq. (12-108), and Y(Q) = 0206 + 0.078 (Q - 1) -0.008(£-l)2
shearing displacements of the crack flanks. As argued by Hutchinson and Suo (1991), systems with /? + 0 may be analyzed by taking /J = 0, because p is generally small (<1) and has negligible influence on fracture parameters. A convenient measure of the mode mixity at the crack tip is given by the angle ij/ (Hutchinson and Suo, 1991):
(12-129)
where X and Q are defined in Eqs. (12-94 a and c) and are given for plane stress in terms of engineering constant in Eqs. (12-124). Alternative data reduction schemes based on compliance calibration for the fracture toughness evaluation using the ENF specimen (Test Procedure for the ENF Specimen, ASTM STP 936,1987, Carlsson etal., 1986 b; Gillespie Jr. and Carlsson, 1990; O'Brien, 1986) have been examined, but they demand more work from the experimenter than the procedure outlined herein, and appear to produce more scatter in the data (minutes of the ASTM, D 30.02.02 Task Group). The ASTM task group does not recommend compliance calibration. 12.3.3 Mixed-Mode Fracture Specimens As mentioned in Sec. 12.2.4, cracks in orthotropic fiber-reinforced materials are often trapped in mixed-mode-stress fields. Furthermore, the toughness for a delamination or an interface depends on the mode of loading. For situations when the delamination is located between similar isotropic or orthotropic materials, or for bimaterial interfacial cracks with the oscillation index s = 0 (and thus /? = 0), see Sec. 12.2.5, the interface stress intensity factors Kx and K2 measure the amplitudes of the tensile and shear stress singularities respectively, and the relative opening and
(12-130) where the interfacial stress intensity factors Kx and K2 are replaced by the conventional Kx and Ku. Thus, for pure mode I problems \jj = 0, while pure mode II is represented by = ± 90°, with the sign given by the direction of shear. Many studies, for example, Hashemi etal. (1990), Johnson and Mangalgiri (1987), Jordan and Bradley (1986), Jurf and Pipes (1982), Leichti and Chai (1992), and Wang and Suo (1990) have shown that the fracture toughness, Gc, depends on the mode of loading: Gc = Gc(iA)
(12-131)
These studies show that both the delamination and interfacial toughness generally increase with increasing amount of mode II loading, that is, with increased magnitude of \j/. This topic will be discussed further in Sec. 12.4.3. Consequently, delamination and debonding fracture characterizations require mixed-mode fracture testing. Several fracture specimens have been designed for the study of mixed-mode delamination fracture of composites. Figure 12-21 shows some of the more popular mixedmode fracture specimens. The earliest mixed mode composite specimen is the cracked-lap shear (CLS) specimen introduced by Wilkins et al. (1982). The CLS specimen (Fig. 12-21 a) was a candidate for a standard ASTM mixed-mode composite
12.3 Fracture Specimens
(a)
(c)
(d)
fun mi i
(e)
(f)
Figure 12-21. Mixed-mode fracture specimens for composite materials: (a) cracked-lap shear (CLS), (b) modified Arcan, (c) asymmetric DCB, (d) mixed-mode flexure, (e) end-loaded split (ELS), and (f) mixed mode bending (MMB).
test specimen, but it suffers from testing difficulties, geometric nonlinearities, and a narrow range of mode mixities, \j/9 and is, therefore, not a popular specimen anymore. Jurf and Pipes (1982) modified the Arcan specimen (Arcan et al., 1978) to produce mixed-mode fracture data by using a single-edge notch specimen in place of the original V-notched specimen, see Fig. 12-21 b. The asymmetric DCB specimen shown in Fig. 12-21 c was introduced by Bradley and Cohen (1985). This specimen, however, demands a complex loading arrangement. The mixed-mode flexure specimen, Fig. 12-21 d, devised by Russell and Street (1985) is easier to test than the previously mentioned specimens but does not allow much variation in the mode mixity. The
565
end-loaded split (ELS) specimen devised by Hashemi et al. (1990), Fig. 12-21e, was discussed at some length in Sec. 12.2.4. It was demonstrated, based on global-mode partitioning, that the mode I fraction, GYIG, could be varied between zero and unity (Fig. 12-13) by varying the ratio of the thickness of the loaded beam to the total beam thickness. Notice, however, that the local-mode partitioning predicts a ratio GJG that varies only between 0.38 and 0.62 for the full range of thickness ratios, see Table 12-4. Besides the limited range of mode mixities realized for most specimens discussed above, some of the specimens suffer also from having a varying mode mixity as the crack grows. A specimen that does not possess these limitations is the recent mixedmode bending (MMB) test suggested by Crews and Reeder (1988), see Fig. 12-21 f. The MMB test is currently a candidate for becoming an ASTM standard because of simplicity of testing and the wide range of mode mixities. Figure 12-22 depicts the geometry parameters of the specimen. The MMB test is a superposition of the DCB and ENF tests discussed previously. As shown in Fig. 12-22, the loading lever will add an opening load to the mid-span loaded ENF specimen. The distance, c, between the point of load application and the
, Loading lever
Figure 12-22. Geometry and loading of the MMB test.
566
12 Fracture of Fiber Composites
mid-span, see Fig. 12-22, determines the ratio of the downward force, Pc, to upward force, Pt, and hence the mode mixity. Pure mode II corresponds to c = 0 with the ratio GJGn increasing with increased distance c. Crews and Reeder (1988) derived the following asymptotic beam theory expressions for Gl5 Gn, and G 3a2P2 GT = 4w . 2,h3, , L„2 £i „ (3c-L)2
(12-132a)
9 a2 P2 Gu = 16w2h3L2E1
(12-132b)
G =
3 a2 P2
16w2h3L2E1 + 3(c + L)2]
L)2 [4(3c-L)2
(12-132c)
where the geometry symbols are defined in Fig. 12-22, and Ex is the longitudinal beam modulus. The mode mixity is Gj _ 4 / 3 c -
(12-133)
Thus, Gi/Gn is a function only of load position, c, and half-span length, L. For c < L/3, Eq. (12-133) is invalid because crack-face contact may occur that is not accounted for in the beam analysis. The asymptotic beam expressions above are strictly valid for long crack lengths and beams with high shear modulus G 13 . Crews and Reeder (1988) provide modified expressions for GY and Gn that are quite accurate for the finite geometries employed in actual fracture testing
3P 2 (3c-L) 2 4w2h3L2E1
2a
10G 13 16 w2
1
(12-134a)
, h2E, (12-134b) 5G 13,
where (12-135) Inspection of Eqs. (12-134) reveals that the ratio Gl/Gu is slightly dependent on crack length. Detailed analysis of a typical MMB test presented by Crews and Reeder (1988), however, shows that within the practical range of crack lengths (a = 25 to 45 mm), the Gl/Gll ratio displays only 5% variation. The design of the MMB apparatus is straightforward for stiff beams with low toughness values. Ductile thermoplastic composites, however, tend to fracture at relatively large displacements, and rotation of the lever arm of the original test apparatus introduced a geometrically nonlinear axial load in the beam that reduced the strain-energy release rate of the specimen (Reeder and Crews, 1991). The beam analysis presented above is valid only for small displacements and significant errors may appear if fracture toughness data recorded in the nonlinear range are reduced based on the above linear small displacement analysis. Reeder and Crews (1991) recently presented a redesigned MMB apparatus that possesses negligible geometrical nonlinearities even for extremely tough composites. It is thus recommended that mixed-mode fracture testing be performed using the redesigned MMB apparatus, see Reeder and Crews (1991). 12.3.4 Test Specimens for Interleaved Composites
As discussed in Sec. 12.2.6, interleaved composites (Fig. 12-18) have aroused considerable interest because of their improved delamination resistance. Delamination test specimens for interleaved composites are similar to the DCB and ENF specimens employed for mode I and
12.3 Fracture Specimens
mode II fracture testing of homogeneous composites, see Sees. 12.3.1 and 12.3.2. Figure 12-23 shows the geometry of the interleaved DCB and ENF specimens, here denoted by "IDCB" and "IENF", respectively. An asymptotic solution for bonded fracture specimens with isotropic constituents by Suo and Hutchinson (1989) was discussed in Sec. 12.2.6. For interleaved specimens with moderately thick interlayers or orthotropic constituents, no closed-form solution for the crack-tip fields exists, but it is reasonable to expect that the presence of an interlayer in a composite laminate will influence crack-tip field parameters in a manner similar to that in bonded specimens with isotropic constituents. The influence of a compliant interlayer on the compliance and energy-release rate of composite fracture specimens has been investigated by Sela et al. (1989), Ozdil and Carlsson (1992 b) and Carlsson and Aksoy
567
(1991). In the investigations of the IDCB and IENF specimens presented by Ozdil and Carlsson (1992 b) and Carlsson and Aksoy (1991), particular emphasis was placed on the detection of any mixed-mode effects. In the next section we will review recent finite element and beam theory analyses of these specimens. IDCB Specimen Ozdil and Carlsson (1992 b) present finite element analysis of the IDCB specimen illustrated in Fig. 12-23 a. Material properties of the composite are representative for graphite-epoxy composites: E = 149 GPa
E2 = E3 = 9.4 GPa
v 2 , = 0.54 v12 = v 13 = G12 = G 13 = 7.5 GPa G 23 = 3.1 GPa Material properties of the interlayers listed in Table 12-5 are typical for two commer-
Piono hinge Insert film (Starter crack) Precrack y—Crack extension Interleaf film
>—Composite yf
(a)
B
L
C
IP/2
\
Figure 12-23. Interleaved fracture specimens, (a) IDCB specimen, (b) IENF
(b)
specimen.
568
12 Fracture of Fiber Composites
Table 12-5. Mechanical properties of thermoplastic and thermoset interleaves. Property
Units
Thermoplastic
Thermoset
E v
GPa
4.57 0.40
2.45 0.38
dally available interleaves made from thermoplastic and thermoset polymers. These materials are denoted TPI and TSI, respectively. The IDCB specimen was modeled in plane strain using four-node isoparametric elements. Two configurations of the crack plane were considered, one (symmetric) where the crack is located centrally within the interleaf, and the other (asymmetric) where an interfacial crack is located between the interleaf and the composite. As will be discussed in Sec. 12.4.4, fracture of interleaved composites commonly occurs between the interleaf and composite. Therefore, the analysis of the asymmetric specimen may be more pertinent for the actual fracture process. In the finite-element analyses, a homogeneous DCB specimen (no interleaf), and IDCB specimens with 16, 40, 80, and 128 (im thick thermoplastic and thermoset interleaves were considered. The total specimen length and thickness were kept constant at 120 and 3.4 mm irrespective of the interlayer thickness. The crack length was 20 mm for all cases and the specimen was loaded by 1 N (1 mm wide specimen). Although the analysis is limited to specific loading and specimen dimensions, the results should be of some general character. Changes in the magnitude of load and specimen dimensions will only alter the magnitude of the stresses and displacements, but not the character of the distributions.
Compliance Compliance, C, of the IDCB specimen was calculated from the vertical deflection, <5, of the points of load application, divided by the applied load, P. Table 12-6 lists IDCB compliance as a function of interleaf thickness for the symmetric and asymmetric crack configurations, Ozdil and Carlsson (1992 b). For interleaf thicknesses up to about 16 (im, the presence of an interlayer does not significantly affect the IDCB compliance. For thicker interleaves, however, the increased compliance is due to the decreased thickness of the composite plies (total beam thickness is constant). The asymmetry has a negligible influence on the IDCB compliance. Slightly larger compliance values were obtained for the more compliant TSI. Strain-Energy Release Rate Components of the strain-energy-release rate, GY and Gn, were computed for the IDCB specimen using the finite element crack closure method outlined in Sec.
Table 12-6.
Compliance C of IDCB specimen. a = 20 mm, w = \ mm. Interleaf thickness (urn)
Crack configuration a
C (mm/N) TPI
TSI O.128Ob
0
S
16 40 80 128
S S S S
0.1294 0.1320 0.1362 0.1416
0.1297 0.1325 0.1373 0.1431
16 40 80 128
A A A A
0.1294 0.1319 0.1361 0.1414
0.1297 0.1325 0.1372 0.1429
a
A: asymmetric, S: symmetric; b without interleaf; TPI: thermoplastic interleaf, TSI: thermoset interleaf.
12.3 Fracture Specimens
12.2.3. Equations (12-66) were modified to incorporate the asymmetric crack configuration according to
2Aa
(Fcvc-Fdvd)
(12-136a)
(Tcuc-Tdud)
(12-136b)
where Fc and Fd, and Tc and Td are the normal and tangential forces, respectively, required to close the crack tip (bring nodes c and d together). Similarly, the quantities vc, vd and uc ud are the normal and tangential nodal displacements (crack closure distances). For the symmetric crack configuration: Fc = Fd, vc = — vd and Tc = Td = 0. Substitution in Eq. (12-136b) yields Gn - 0, which shows that the symmetric crack configuration is pure mode I, and G = Gx. As
569
discussed previously, the asymmetric configuration is inherently mixed mode and the decomposition of G into tensile and shear components, G, and Gn, may suffer from convergence problems due to the bimaterial oscillatory singularity. Therefore, the results for GY and Gn presented herein should be viewed as qualitative. The total energy-release rate, however, is a well-defined physical quantity. Table 12-7 presents components of the energy-release rate for symmetric and asymmetric IDCB specimens. Inspection of the results in Table 12-7 reveals that the energy-release rate is only slightly affected by the presence of thin interlayers. However, for thicker interleaves, the energy-release rate increases significantly due to the addition of a compliant interlayer. The asymmetric configuration provides somewhat larger energy-release rate than the
Table 12-7. Strain-energy release rate components for symmetric (S) and asymmetric (A) IDCB specimens. a = 20 mm, P = 1 N/mm. Interleaf thickness (^im)
Interleaf material a
Crack configuration
Gl (J/m2)
Gu (J/m2)
G (J/m2)
None
S
7.72
0
7.72
16 40 80 128
TPI TPI TPI TPI
S S S S
8.16 8.29 8.49 8.82
0 0 0 0
8.16 8.29 8.49 8.82
16 40 80 128
TSI TSI TSI TSI
S S S S
8.27 8.36 8.56 8.90
0 0 0 0
8.27 8.36 8.56 8.90
16 40 80 128
TPI TPI TPI TPI
A A A A
8.49 8.67 8.97 9.30
0.09 0.13 0.17 0.33
8.58 8.80 9.14 9.63
16 40 80 128
TSI TSI TSI TSI
A A A A
8.45 8.65 8.99 9.35
0.06 0.12 0.20 035
8.51 8.77 9.19 9.70
0
TPI: thermoplastic interleaf, TSI: thermoset interleaf.
570
12 Fracture of Fiber Composites
symmetric one. Furthermore, the asymmetric geometry is mixed mode, although the magnitude of Gn is very small for thin interlayers. Stress-Intensity Factor The mode I stress-intensity factor Kl9 was determined for the symmetric IDCB geometry. Kx was calculated from the global energy-release rate and relations between KY and GI? provided in Sec. 12.2, see Ozdil and Carlsson (1992 b). Table 12-8 lists calculated Kx values for the various symmetric IDCB specimens considered. It is observed that the intensity of stress is much reduced by the presence of a compliant layer, and that KY is almost independent of layer thickness. The TSI reduced KY more than the TPI because of its lower Young's modulus, see Table 12-5. Similar features have been observed in adhesively bonded specimens (Crews Jr. et al., 1988; Sela et al., 1989; Wang, 1978). The reduction of KY due to the compliant bondline material has been termed "elastic shielding" by Hutchinson and Suo (1991). The apparent fracture toughness of bonded specimens, Klc, may thus increase as a result of the shielding effect.
Table 12-8. Stress intensity factor for homogeneous and interleaved symmetric DCB specimens, a = 20 mm, P = \ N/mm. Interleaf thickness
Xj(kNm" 3/2 )
TPI 368.1l
0 16 40 80 128 a
TSI
210.6 212.5 215.0 219.1
154.0 154.6 156.5 159.7
Without interlayer; TPI: thermoplastic interleaf, TSI: thermoset interleaf.
The mixed mode stress intensity factors, Kx and K 2 , for the asymmetric IDCB configuration have not been determined. In principle, K1 and K2 could be determined from the stress field ahead of the crack, but this procedure requires an extremely refined finite-element mesh. Based on the analysis of bonded fracture specimens with isotropic constituents by Suo and Hutchinson (1989) it is reasonable to expect that K1 and K2 would diminish in the IDCB specimens with orthotropic adherends as a result of the compliant interlayer. Data Reduction for Fracture Toughness of IDCB Specimen The results presented for the IDCB specimen show that this specimen essentially remains pure mode I in the presence of a thin compliant interlayer. Furthermore, as shown in the previous section, a thin interlayer was found not to change the compliance and energy release rate very much, as compared to the homogeneous specimen. Compliance and energy-release rate are global quantities that are dominated by the stiffness of the composite material in the cantilever beams. Consequently, compliance calibration procedures for determination of GIC from experimental fracture data, derived for homogeneous and symmetric DCB specimens, are adequate also for the IDCB specimen with thin interleaves. At larger interleaf thicknesses, however, mixed mode effects may influence the results. IENF Specimen The IENF specimen, Fig. 12-23 b, has been investigated with both shear deformation beam and finite-elements theory by Carlsson and Aksoy (1991). Below, we will review these results.
571
12.3 Fracture Specimens
Beam Theory Analysis With the geometry symbols in Fig. 12-23b, the IENF compliance is Q
L (^ll)BC , M ^ 5 5 ) B C
2w/c
6w
12w
(12-137)
where L, w, and k are the half span length, beam width, and shear correction factor (k = 5/6), respectively. D'11 and A%5 are elements of the inverted bending and shear stiffness matrices, defined by Carlsson and Aksoy (1991). The subscripts AB and BC denote delaminated and intact beam sections, see Fig. 12-23 b. The energy release rate is given by G =
8w 2
(12-138)
Notice that, due to the approximations involved, the beam analysis makes no distinction between symmetric and asymmetric crack configuration. Finite-Element Analysis A two-dimensional plane strain finite element analysis of the IENF geometry was performed by Carlsson and Aksoy (1991) in order to numerically calculate compliance and energy-release rate components. Stress-intensity factors were not considered by the authors. As for the IDCB specimen, both symmetric and asymmetric (bimaterial) crack configurations were investigated. The material properties of the composite and thermoset interleaf are identical to those provided in Table 12-5 for the IDCB specimen discussed in the previous section. Homogeneous and interleaved specimens with 12.7, 25.4, 76.2, and 127 \xm thick interleaves were considered. For all cases,
the crack length (a) was 25.4 mm and the semispan length (L) was 50.8 mm, (a/L = 0.5). The total beam thickness was kept constant at 3.43 mm for all specimens, and the beam width (w) was 25.4 mm. Compliance of the IENF specimen was determined from the vertical beam deflection, (5, at the beam center, divided by the applied load, P. Strain energy release rate components GY and Gu were calculated using the finite element crack closure method, Eqs. (12-136). In all symmetric and asymmetric configurations considered, no crack opening was found which shows that the IENF geometry is pure mode II. Table 12-9 provides compliance and energy release rate for symmetric and asymmetric IENF specimens with 12.7 and 25.4 jim thick interlayers. The asymmetry of the crack has negligible influence on compliance and energy release rate for this range of interleaf thicknesses. Table 12-10 provides a comparison between finite-element (symmetric crack location) and beam theory results, Eqs. (12-137) and (12-138), for the IENF compliance and energy release rate. Finite element analysis gives slightly larger compliance values than the beam theory. Also the strain-energy release rates predicted by the two approaches are in close agreement. Consequently, the more convenient beam theory approach could be employed for ex-
Table 12-9. Compliance and energy-release rate for IENF specimens with symmetric (S) and asymmetric (A) (interfacial) crack location; a = 25A mm, L = 50.8 mm, w = 25.4 mm, P = 4.44 N. Interleaf thickness (Mm) 12.7 25.4
C[(m/N)x 10"5] S 4.190 4.222
A 4.189 4.220
Gu[(J/m 2 ) x l 0 " 2 ] S
A
1.320 1.330
1.330 1.360
572
12 Fracture of Fiber Composites
Table 12-10. Compliance and energy-release rate for IENF specimen (thermoset interleaf, symmetric crack location) calculated from beam theory (BT) and finite elements (FE); a = 25 A mm, L = 50.8 mm, w = 25.4 mm, p = 4.44N. Interleaf
cFE
thickness (urn)
[(m/N)x 10~5]
0 25.4 76.2 127
4.157 4.222 4.330 4.431
CBT
4.137 4.153 4.193 4.239
GFE
GBT
[(J/m 2 ) x l 0 " 2 ] 1.260 1.330 1.530 1.730
1.230 1.340 1.620 1.720
perimental evaluation of mode II fracture toughness. Data Reduction for Fracture Toughness of IENF Specimen Inspection of Table 12-10 reveals that the compliance and energy release rate of the interleaved ENF specimen may be significantly larger than for the homogeneous specimen. Crack growth is inherently unstable for the ENF and IENF specimens (Carlsson and Pipes, 1987; Aksoy and Carlsson, 1992) and fracture toughness evaluation is based on the compliance and critical load at one crack length, as opposed to the compliance calibration method recommended for the IDCB specimen. Aksoy and Carlsson (1992) outline the recommended data reduction methodology for the IENF specimen. The appropriate composite modulus E1 is obtained by iterating Eq. (12-137) until the calculated compliance agrees with the experimentally measured. The modulus E± so obtained is then substituted into Eq. (12-138) together with the critical load for crack propagation, Pc, to obtain the fracture toughness. For further details, see Aksoy and Carlsson (1992).
12.4 Fracture Data The experimental fracture characterization of composites will be discussed in this section. The description is limited to crack propagation parallel to the fibers in homogeneous orthotropic beam specimens and fracture of interleaved composites. With the above restrictions, the experimental determination of fracture toughness appears straightforward, but experimenters have found that the interlaminar fracture toughness is extremely sensitive to numerous experimental parameters. Figure 12-24 illustrates micromechanisms that may influence the resistance of a growing crack. The fracture energy measured in a test could be related to the following mechanisms: formation of the fracture surface of the main crack, plastic dissipation due to plastic-zone formation in the matrix around the crack tip and translation of the plastic zone, crack bridging by intact fibers left behind the crack front, and formation of microcracks and secondary cracks that may parallel the main crack. Several of these mechanisms develop with crack extension. For example, fiber briding tends to build up and increase with crack propagation and a resistance curve (K-curve) is necessary to describe the fracture process. Crack tip plasticity may also increase with crack extension. FihprQ Mbers \
Fiber breakage
^ Main crack
Side
1
••>
>r^2
Matrix
Bridging fibers
A
s
Plastic zone
Figure 12-24. Illustration of energy absorbing mechanisms in interlaminar fracture (Friedrich et al., 1989).
12.4 Fracture Data
The size of the crack-tip plastic zone is sensitive to the volume fraction of matrix material because the fibers generally are linear elastic and constrain the development of crack-tip plastic zones. Significant increases of the plastic-zone size and fracture toughness are achieved by interleafing, as will be discussed later in this section. In addition, the matrix may display viscoelastic and viscoplastic effects, in which case the fracture resistance is deformationrate and temperature dependent. Other complications that may influence the i^-curve for composites are the methods of forming a precrack during processing of the composite and the subsequent extension of the precrack prior to the actual fracture testing (precracking). Figure 12-25 illustrates the situation at the crack tip after processing of the composite. The starter film is placed between the plies at the desired crack plane during layup of the composite. Typically, the insert is a 7-25 jim thick PTFE or polyimide film coated with a release agent. As illustrated in Fig. 12-25, a wedge-shaped resin zone is formed between the plies ahead of the insert film, and fracture testing where the crack is propagated directly from the tip of the starter film may lead to artifically large toughness values because of the locally high resin content (Martin, 1988; Murri and O'Brien, 1985). In order to minimize the influence of the insert film, its thickness should be kept at a minimum. ASTM recommends a maximum film thickness of 13 pm for the DCB test (minutes of the ASTM Task Group). On the other hand, fiber bridging, Fig. 12-26, may develop during mode I loading already at short crack lengths and, therefore, the precracking distance should be kept at a minimum. In fact, ASTM recommends no precracking for the DCB test (minutes of the ASTM Task Group). Fiber bridging does not oc-
Starter film 10-50 [im thick
573
Resin-rich region 200-500 \im in length
Figure 12-25. Starter film and resin-rich region in composite laminate.
cur in multidirectional composite laminates and fracture toughness values incorporating contributions from fiber bridging should thus not be used in an assessment of the damage tolerance of composite structures. This discussion brings up yet another issue, namely the appropriate mode of loading during precracking. Initially, mode I precracking was suggested for most fracture tests, but the extensive fiber bridging that occurs in some system has shifted the attention to shear-driven precracks or thin films without precracks (Meeting between the ASTM Task Group D 30.02.02, European Group on Fracture, and Japan Industrial Standards in 1989; minutes of the ASTM Task Group). The issue is under investigation by standard issuing organizations.
Bridged fibers
Figure 12-26. Schematic of fiber bridging in mode I loading.
12.4.1 Mode I Fracture Data A comprehensive review of mode I interlaminar fracture testing and data has been presented by Davies and Benzeggagh
574
12 Fracture of Fiber Composites
(1989). In this section only some illustrative experimental fracture data generated in mode I testing of polymer matrix composites will be presented. R-Curve Effect As discussed previously in this section, pull-out and fracture of bridged fibers behind the crack front (Figs. 12-24 and 12-26) is a common occurrence in mode I fracture testing. At the first loading increment (Aa = 0), when the delamination grows from the tip of the thin film insert starter crack, Fig. 12-25, there is no fiber bridging and the crack flank displacements behind the crack tip are at a minimum. As the crack grows, the crack surfaces become more and more separated and bridged fibers may fracture or become pulled out from the matrix, which causes the apparent fracture toughness to increase. With further crack extension, a steady-state toughness is usually reached corresponding to an equilibrium number of bridged fibers per unit crack area, see Fig. 12-27. The #-curve illustrated in Fig. 12-27 may be characterized by an initiation value, GIC (init.) corresponding to crack extension after precracking from the film starter, and a steady-state propagation
Graphite/ rubber toughened epoxy
60
70
80
90
Crack length (mm)
Figure 12-27. incurve describing mode I fracture resistance of a graphite-rubber toughened epoxy composite.
value, GIC(prop.). As demonstrated by Hashemi et al. (1990), however, GIC(init.) is not a unique value for a composite system, because GIC(init.) depends on the amount of precracking beyond the tip of the insert film. To obtain the full #-curve, Hashemi et al. (1990) showed that the extent of precracking (crack growth beyond the tip of the insert film) should be kept at a minimum. To propagate the delamination directly from the starter crack may lead to an erroneous overestimate of the initiation toughness, GIC (init.), due to a resin-rich region at the tip of the film insert (see Fig. 12-25). The increase in fracture toughness with crack extension illustrated in Fig. 1227, may (besides fiber bridging) also be due to an increase in the degree of plastic deformation and microcracking around the crack tip with crack extension. Temperature and Rate Effects Because polymer matrices generally are viscoelastic or viscoplastic, changes in temperature and strain rate will influence the fracture toughness. Shivakumar and Crews (1987) proposed that the contribution of the plastic energy dissipation to the fracture toughness, GIC, is directly proportional to the height of the crack-tip plastic zone. Precracking leads to the formation of a plastic zone while fracture testing on precracked specimens leads to translation of the plastic zone. The volume of the newly formed plastically deformed material should therefore be proportional to the dimension of the plastic zone perpendicular to the crack plane ("plastic-zone height"). To examine rate and temperature effects on fracture toughness, we will thus consider the factors contributing to the plastic-zone height. The height of the mode I crack-tip plastic zone in the absence of fiber bridging is for plane strain (Ozdil and
575
12.4 Fracture Data 10'
Carlsson, 1992 c) (12-139) in which K1 is the mode I stress intensity factor and oy is the yield strength of the matrix. It should be pointed out that Eq. (12-139) is valid only for the situation when the plastic zone is fully contained in the resinrich layer between plies in a composite. As this layer is thin, 8-32 jim in thickness (Bradley and Cohen, 1985; Friedrich et al, 1989; Hunston, 1984), severe constraints may be imposed on the plastic zone development by the rigid fiber-reinforced layers, which will limit plastic zone height, and thus the fracture toughness (Bradley and Cohen, 1985; Friedrich et al., 1989; Hunston, 1984). The plastic zone model should thus be used with caution, but is useful for the interpretation of rate and temperature effects on fracture toughness. For a viscoelastic matrix, the fracture toughness Klc generally decreases strongly with strain rate while the yield strength increases, see, for example, Friedrich et al. (1991). A higher strain rate is thus expected to lead to a smaller plastic zone height in the composite and hence diminished fracture toughness, GIC. Increased temperature will soften the polymer and reduce the yield strength. The influence of temperature on Klc is usually less pronounced and Eq. (12-139) would predict an increased plastic zone height and fracture toughness GIC with temperature. Figures 12-28 and 12-29, obtained from Friedrich et al. (1989), Smiley and Pipes (1987 a), and Hashemi et al. (1990) provide examples on rate and temperature dependence of GIC for graphite fiber composites with thermoset and thermoplastic matrices. The trends of the data are consistent with the plastic zone arguments above.
/Graphite/PEEK
i
^, •i-i-
o,0-i
Graphite/Epoxy
HO
10
ICf
IO"V
IO'° 10"' yCT (m/sec)
Figure 12-28. Mode I fracture toughness, GIC, versus crack tip opening velocity for graphite-epoxy and graphite-PEEK composites (Friedrich et al., 1989; Smiley and Pipes, 1987).
80
100
120
140
Temperature, C
Figure 12-29. Mode I fracture toughness, GIC (prop.) versus temperature for a graphite-PEEK composite (Hashemi et al., 1990).
12.4.2 Mode II Fracture Data Mode II interlaminar fracture has been the subject of many investigations (Carlsson etal., 1986 b; Hashemi et al., 1990; Martin and Murri, 1990; Murri and O'Brien, 1985; O'Brien et al., 1987; Russell and Street, 1982 and 1985; Trethewey et al., 1988). It has generally been observed that the mode II fracture toughness, GIIC, is larger than GIC for composites with brittle resins, while increased matrix ductility makes GIC approach GIIC. The micromechanisms responsible for the fracture tough-
576
12 Fracture of Fiber Composites
G
..C<
__o_^, o trcP—
3
E
3
V 50mm
•
o"
2 / i 1r j L_I 1
0
O a p = 63mm
o V 68mm
GIIC (init.)
A ap
=
82mm
I
I
i
i
I
10
20
30
40
50
60
Increment of crack propagation, Aa (mm) Figure 12-30. R-curve describing mode II fracture resistance of a graphite-PEEK composite (Hashemi etal., 1990).
10.00
/Graphite/PEEK
E 100 pox Graphite/Epoxy0.10
0.01
ioy
I07
icr
u CT (m/sec)
Figure 12-31. Mode II fracture toughness, GIIC (init.), versus crack tip sliding velocity for graphite-epoxy and graphite-PEEK composites (Friedrich et al., 1989; Smiley and Pipes, 1987).
20
40
60
80 100 120 140
Temperature, T (°C) Figure 12-32. Mode II interlaminar fracture toughness, GIIC (prop.), versus temperature for a graphitePEEK composite (Hashemi et al., 1990).
ness were discussed early in Sec. 12.4 and are illustrated in Fig. 12-24. Plastic deformation around the crack tip appears to be the major energy-absorbing mechanism because fiber bridging is much less pronounced in shear (mode II) than in tension (mode I) (Hashemi etal., 1990). Consequently, the plastic zone argument presented above for mode I fracture testing should be appropriate for explaining the mode II fracture toughness. Figure 12-30 shows a mode II R-curve for a graphite-PEEK composite (Hashemi et al., 1990). The various symbols represent various precracking lengths and the unique R-curve observed indicates that fiber bridging is not a major factor in mode II fracture of this composite. The primary mechanism explaining the rising portion of the R-curve is crack-tip plasticity that develops in the intense crack tip shear field and evolves with crack growth. It should be pointed out that PEEK is an engineering thermoplastic with extraordinary toughness. For a more brittle matrix such a pronounced incurve is not expected. Rate effects on mode II fracture were examined by Friedrich et al. (1989), Carlsson etal. (1986b), and Smiley and Pipes (1987b). Figure 12-31 displays GIIC(init.) versus crack tip sliding velocity. Similarly to the mode I fracture toughness, Fig. 12-28, GIIC drops substantially with rate at high crack tip sliding velocities. The effect of temperature on GIIC of a graphite-PEEK composite is illustrated in Fig. 12-32. Similar to GIC, Fig. 12-29, GIIC increases with increasing temperature. 12.4.3 Mixed-Mode Fracture Data This section will address data compiled for fracture due to combinations of tensile (mode I) and shear (mode II) stresses ahead of the crack tip. The micromechanisms dis-
577
12.4 Fracture Data
cussed early in Sec. 12.4 and the plastic zone argument detailed in Sec. 12.4.1 should be applicable also for mixed mode fracture. As discussed in Sec. 12.3.3, a complete fracture characterization of composites requires mixed mode testing. This was first recognized by Wu (1967), who studied intralaminar fracture of balsa wood and a glass-polyester composite, and formulated a "best fit"-type fracture criterion in terms of stress intensity factors: K ic
(12-140)
K IIC
More customary in interlaminar fracture characterization of composites is to use the components of the strain-energy release rate and to employ the following fracture criterion (Russell and Street, 1985)
ic,
G
=1
(12-141)
IIC,
where the exponents m and n are determined by curve-fitting of experimental data. Johnson and Mangalgiri (1987) collected data on mixed-mode interlaminar fracture toughness for composites with a wide variety of resin toughnesses. As shown in Fig. 12-33, a straight line, corresponding to m = n = 1 in Eq. (12-141) provides a good fit to the experimental data, that is, -^- + - ^ = 1
2.0
0.4
0.8
1.2
G,|,kJ/m
1.6
2.0
2
Figure 12-33. Mixed mode interlaminar fracture toughness data for various graphite fiber composites. Only the matrix material is indicated in the figure (Johnson and Mangalgiri, 1987).
with caution because GIC is rarely equal to GIIC. More recent mixed-mode fracture studies of graphite-PEEK composites have been presented by Reeder and Crews (1991) and Hashemi et al. (1990). Figure 12-34 shows delamination toughnesses of a graphite-PEEK composite determined from the MMB test. The results in Fig. 12-34 support the validity of the linear mixed-mode criterion, Eq. (12-142). The characterization of mixed-mode fracture of unidirectional composites is obscured by the i^-curve effects, see Fig. 12-35. The difference between GIC(init.) and GIC(prop.) in Fig. 12-35 can be attributed
(12-142)
It was also found that the ratio Glc/Gnc increases from about 0.1 to about 1.0 as the resin ductility increases. For ductile resins with G I C « GIIC, the total strain-energy release rate controls the propagation of a crack under pure or mixed-mode conditions. Generally, however, the total-fracture-energy approximation should be used
G, (kJ/m2)
Figure 12-34. Mixed mode interlaminar fracture toughness of a graphite-PEEK composite (Reeder and Crews, 1991).
12 Fracture of Fiber Composites
578
AS4/PEEK
Propagation
ness in a similar manner as the mode I and mode II fracture toughnesses discussed previously. Except at room temperature (20 °C), the data in Fig. 12-36 are reasonably approximated by the linear mixedmode criterion, Eq. (12-142). 12.4.4 Mode I and Mode II Fracture Data for Interleaved Composites
I
2 G,,,(kJ/m2)
Figure 12-35. Mixed mode interlaminar fracture toughness data for a graphite-PEEK composite (Hashemi et al., 1990).
to various degrees of crack-tip plasticity and fiber bridging, see the earlier discussion in this section. The data on mixedmode fracture toughness for initiation toughness presented in Fig. 12-35 are somewhat larger than those presented for the graphite-PEEK system in Figs. 12-33 and 12-34, especially at high Gl/Gll ratios. This may be due to differences in materials processing and precracking methods. Figure 12-36 shows that temperature influences the mixed-mode fracture tough-
-
AS4/PEEK TCC) • 130 D 60 O 20
\v
2 3
I 9
0
1
2
3
4
5
G,, (prop.), (kJ/m*) Figure 12-36. Mixed mode interlaminar fracture toughness data for a graphite-PEEK composite at various temperatures (Hashemi et al., 1990).
Fracture toughness characterization of interleaved composites has been reported in several studies (Aksoy and Carlsson, 1991 and 1992; Evans and Masters, 1987; Masters, 1987 a and b; Ozdil and Carlsson, 1992 a; Sela et al., 1989). All these investigations report tremendous elevations of the mode I and mode II fracture toughnesses, GIC and GIIC. In this section we will briefly review the experimental results of Aksoy and Carlsson (1990) and Ozdil and Carlsson (1992 a). The mode I and mode II fracture investigations (Aksoy and Carlsson, 1992; Ozdil and Carlsson, 1992 a) employed a unidirectional graphite-epoxy composite with a large range of thermoplastic (TPI) and thermoset (TSI) interleaf thicknesses. The mode I (IDCB) and mode II (IENF) fracture specimens, Fig. 12-23, used in the experimental fracture studies, were of similar geometry and material properties as those analyzed in Sec. 12.3.4. The mode I fracture investigation covered interleaf thicknesses between 16 and 129 jim for TPI and 59 and 214 jam for TSI. The corresponding TPI and TSI thickness ranges for the mode II fracture investigation (Aksoy and Carlsson, 1992) are 12.7 to 127 and 43 to 429 fim. It should be pointed out that several of the interleaved fracture specimens suffered from poor adhesion, which preempted the potentially large fracture toughness. In this section, we will limit the presentation to specimens with adequate
12.4 Fracture Data
579
5.0
2.0r
4.0 -
Thermoplastic interleaf (Film E)
3.0
20
40 Interleaf
60
80
I00
I20
I40
thickness,/i.m
Figure 12-37. Mode I fracture toughness versus interleaf thickness for a graphite-epoxy composite interleaved with thermoplastic (TPI) and thermoset (TSI) interleaves (Ozdil and Carlsson, 1992 a).
adhesion. For a more complete discussion, see references (Aksoy and Carlsson, 1992; Ozdil and Carlsson, 1992 a). Data reduction for mode I fracture toughness evaluation followed the empirical compliance calibration procedure outlined for DCB specimens by ASTM (proposed standard test method), while mode II fracture toughness evaluation was based on beam theory, see Sec. 12.3.4 and Aksoy and Carlsson (1992). Figure 12-37 shows GIC plotted versus interleaf thickness. The TPI enhanced the fracture toughness about five-fold over the baseline value at a thickness of only 16 jam. Thicker interleaves did not provide any further elevation of GIC. The 59 \xm TSI provided about 30% elevation of GIC compared to the baseline. Thicker TSI were investigated, but the specimens failed at low loads due to adhesion problems. It should be noted that the thermoset (FM 300) interleaved composites investigated by Sela et al. (1989) displayed large enhancements of GIC, which indicates that thermoset interleaves also have toughening potential. Thick thermoset interleaves may, however, be required. The thinnest
0.00
0.05 0.10 Interleaf thickness, mm
0.15
4.0
E
U)
in (D
C JZ
Thermoset interleaf (FM300I)
b) 0 0.0
0.1 0.2 0.3 0.4 Interleaf thickness, mm
0.5
Figure 12-38. Mode II fracture toughness versus interleaf thickness for an interleaved graphite-epoxy composite (Aksoy and Carlsson, 1992). (a) TPI, (b) TSI.
FM 300 interlayer studied by Sela et al. (1989) was 100 pm. Microscopic investigation of the fracture process (Ozdil and Carlsson, 1992 a) indicated that the large toughness of the IDCB specimens is associated with a significant degree of plastic deformation of the interleaf. This observation supports the plastic
580
12 Fracture of Fiber Composites
energy dissipation arguments for the toughness discussed in Sec. 12.4. Detailed estimates of the plastic zone height in interleaved DCB specimens presented by Ozdil and Carlsson (1992 c) support the experimentally observed trends in Fig. 12-37. Figure 12-38 shows GIIC plotted versus interleaf thickness. Notice the small interleaf thickness range and the rapid development of GIIC for TPI, Fig. 12-38 a. The extremely large GIIC value at 127 jum film thickness is largely attributed to fracture of fibers bridging the crack surfaces. The development of GIIC versus film thickness for the TSI is less rapid than that for the TPI, see Fig. 12-38 b, although large GIIC values can be achieved at large TSI thicknesses. Inspection of the fracture process under optical and scanning electron microscopes (Aksoy and Carlsson, 1992) revealed plastic deformation and limited microcrakking in the TPI and significant microcrakking in the TSI. Detailed estimates of the plastic zone height in interleaved ENF specimens presented by Aksoy and Carlsson (1991) are in agreement with the toughness versus interleaf thickness trends shown in Fig. 12-38.
12.5 Acknowledgements This work was supported in part by the National Science Foundation (Grant MSM-88-22229), and by the Department of Mechanical Engineering, Florida Atlantic University (FAU). Part of the review builds on my previous cooperation with graduate students of FAU and Dr. John W. Gillespie Jr. and Dr. R. Byron Pipes of the University of Delaware, as well as several graduate students I was fortunate enough to work with during may stay at the University of Delaware. In addition, I would like to acknowledge cooperation with Prof.
Klaus Friedrich and Dr. Joseph KargerKocsis, both of the University of Kaiserslautern. More recently, I have had much fruitful communication with Prof. John W. Hutchinson of Harvard University. This chapter would not have been possible without the assistance from Rosemarie Chiucchi, Susan Cury, and Evelyn Hoffman, who typed and retyped many iterations of the manuscript. The graphics work by Mark Deshon of the University of Delaware and Shawn Pennell and Sherri Vonhartman of FAU is gratefully acknowledged.
12.6 References Aksoy, A., Carlsson, L. A. (1991), Eng. Fracture Mech. 39, 525. Aksoy, A., Carlsson, L. A. (1992), Compos. Sci. Tech. 43, 55. Arcan, M., Hashin, S., Voloshin, A. (1978), Exp. Mech. 18, 141. ASTM STP696 (1979). Philadelphia, PA: Am. Soc. Test. Mat. ASTM STP876 (1985). Philadelphia, PA: Am. Soc. Test. Mat. ASTM STP936 (1987). Philadelphia, PA: Am. Soc. Test. Mat. Barrett, J. D. Foshi, R. O. (1977), Eng. Fracture Mech. 9, 371. Bascom, W. D., Bitner, J. L., Mouton, R. J., Siebert, A. R. (1980), Composites 11, 9. Beaumont, P. W. R., Schultz, J. M. (1990a), in: Delaware Composites Design Encyclopedia, Vol. 4: (Eds.). Westport, CT: Technomic, p. 21. Beaumont, P. W. R., Schultz, J. M. (1990b), in: Delaware Composites design Encyclopedia, Vol. 4: (Eds.). Westport, CT: Technomic, p. 39. Berry, J. P. (1963), J. Appl. Phys. 34, 62. Bostaph, G. M., Elber, W. (1982), paper presented at: AS ME Winter Conference. Bradley, W. L., Cohen, R. N. (1985), in: Delamination and Debonding of Materials: ASTM STP876. Philadelphia, PA: Am. Soc. Test. Mat., p. 389. Broek, D. (1982), Elementary Engineering Fracture Mechanics. Dordrecht: Martinus Nijhoff. Carlsson, L. A., Aksoy, A. (1991), Int. J. Fracture 52, 67. Carlsson, L. A., Pipes, R. B. (1987), Experimental Characterization of Advanced Composite Materials. Englewood Cliffs, NJ: Prentice-Hall. Carlsson, L. A., Gillespie, Jr., J. W, Pipes, R. B. (1986 a), J. Compos. Mater. 20, 594.
12.6 References
Carlsson, L. A., Gillespie, Jr., J. W, Trethewey, B. R. (1986 b), /. Reinf. Plast. Compos. 5, 170. Carlsson, L. A., Sindelar, P., Nilsson, S. (1986c), Compos. Sci. Tech. 26, 307. Crews, Jr., J. H., Reeder, J. R. (1988), NASA TM 100662. Crews, Jr., J. H., Shivakumar, K. N., Raju, I. S. (1988), in: Adhesively Bonded Joints: Testing, Analysis and Design: ASTM STP981. Philadelphia, PA: Am. Soc. Test. Mat., p. 119. Davies, P., Benzeggagh, M. L. (1989), in: Application of Fracture Mechanics to Composite Materials. Friedrich, K. (Ed.). Amsterdam, New York: Elsevier, p. 81. Dost, E. E, Ilcewicz, L. B., Avery, W. B., Coxon, B. R. (1991), Composite Materials: Fatigue and Fracture, 3rd Vol. ASTM STP 1110. Philadelphia, PA: Am. Soc. Test. Mat., p. 476. Dundurs, J. (1969), /. Appl. Mech. 36, 650. England, A. H. (1965), /. Appl. Mech. 32, 400. Erdogan, F. (1965), J. Appl. Mech. 32, 403. Erdogan, F , Gupta, G. D. (1971), Int. J. Solids Struct. 7, 1089. Evans, R. E., Masters, J. E. (1987), in: Toughened Composites: ASTM STP 937. Philadelphia, PA: Am. Soc. Test. Mat. Ewalds, H. L., Wanhill, R. J. H. (1989), Fracture Mechanics. London: Edward Arnold. Friedrich, K. (Ed.) (1989), Application of Fracture Mechanics to Composite Materials. Amsterdam, New York: Elsevier. Friedrich, K., Carlsson, L. A., Smiley, A. J., Walter, R., Gillespie, Jr., J. W. (1989), J. Mater. Sci. 24, 3387. Friedrich, K., Carlsson, L. A., Gillespie, Jr., J. W, Karger-Kocsis, J. (1991), in: Thermoplastic Composite Materials: Carlsson, L. A. (Ed.). Amsterdam, New York: Elsevier, p. 233. Gere, J. M., Timoshenko, S. P. (1984), Mechanics of Materials, 2nd ed. PWS Engineering. Gillespie, Jr., J. W, Carlsson, L. A. (1990), in: Delaware Composites Design Encyclopedia, Vol. 6. Westport, CT: Technomic, p. 111. Griffith, A. A. (1921), Phil Trans. Roy. Soc. of London A 221, 163. Guynn, E. G., O'Brien, T. L. (1985), in: Proceedings AIAA/ASME/ASCE/AHS 26th Structures, Structural Dynamics and Materials Conference (AIAA85-0646), Orlando, FL. Hashemi, S., Kinloch, A. I , Williams, J. G. (1990), J. Compos. Mater. 24, 918. Hellan, K. (1984), Introduction to Fracture Mechanics. New York: McGraw-Hill. Horgan, C. O. (1982), J. Compos. Mater. 16, 411. Hunston, D. L. (1984), Compos. Tech. Rev. 4, 176. Hutchinson, J. W (1989), Report MECH-139, Division of Applied Sciences, Harvard University, MA. Hutchinson, J. W. (1990), personal communication. Hutchinson, J. W, Suo, Z. (1991), Adv. Appl Mech. 28.
581
Irwin, G. R. (1958), in: Handbuch der Physik, Vol. 6: Flugge, (Ed.). Berlin, Heidelberg: Springer, p. 551. Jones, R. M. (1975), Mechanics of Composite Materials. New York: McGraw-Hill. Johnson, W. S., Mangalgiri, P. D. (1987), in: Toughered Composites. ASTM STP 875. Philadelphia, PA: Am. Soc. Test. Mat., p. 295. Jordan, W. M., Bradley, W. L. (1987), in: Toughened Composites. ASTM STP 937. Philadelphia, PA: Am. Soc. Test. Mat., p. 95. Jurf, R. A., Pipes, R. B. (1982), J. Compos. Mater. 16, 386. Kinloch, A. I , Young, R. J. (1983), Fracture Behavior of Polymers. Amsterdam, New York: Elsevier. Knott, J. F. (1973), Fundamentals of Fracture Mechanics. London, Washington: Butterworths. Leichti, K. M., Chai, Y.-S. (1992), /. Appl. Mech. 59, 295. Lekhnitskii, S. G. (1981), Theory of Elasticity for an Anisotropic Body. Moscow: MIR Publishers. Ma, C.-C, Luo, J.-X, Theoretical Analysis of the Interfacial Crack for General Anisotropic Material, Part I: Theory, to be published. Mandell, J. F., Darwish, A. Y, McGarry, F J. (1981), Test Methods and Design Allowables for Fibrous Composites. ASTM STP 734. Philadelphia, PA: Am. Soc. Test. Mat., p. 73. Martin, R. H. (1988), in: Proc. Am. Soc. Composites: Westport, CT: Technomic, p. 688. Martin, R. H., Murri, G. B. (1990), in: Composite Materials: Testing and Design. ASTM STP 1059. Philadelphia, PA: Am. Soc. Test. Mat. Masters, J. E. (1985), Presented at: ASTM Symposium on Fractography of Modern Engineering Materials. Nashville, TN. Masters, J. E. (1987 a), Final Report for Period Sept. 1984-June 1987, AFWAL/MLBC. Wright Patterson Air Force Base, Ohio. Masters, J. E. (1987b), Key Eng. Mater. 37, 317. Mohlin, T, Blom, A. F , Carlsson, L. A., Gustavsson, A. I. (1985), in: Delamination and Debonding of Materials. ASTM STP 876. Philadelphia, PA: Am. Soc. Test. Mat., p. 168. Murri, G. B., O'Brien, T. K. (1985), in: AIAA-85-0647, Proceedings of the 26th AIAA/ASME/ASCE/ASH Conference on Structures, Structural Dynamics and Materials, Orlando, FL. O'Brien, T. K. (1982), in: Damage in Composite Materials. ASTM STP 775. Philadelphia, PA: Am. Soc. Test. Mat., p. 140. O'Brien, X K., Murri, G. B., Salpekar, S. A. (1987), NASA TM 89157. Ozdil, F , Carlsson, L. A. (1992 a), J Compos. Mater. 26, 432. Ozdil, F , Carlsson, L. A. (1992b), Eng. Fracture Mech. 41, 475. Ozdil, F , Carlsson, L. A. (1992c), Eng. Fracture Mech. 41, 645.
582
12 Fracture of Fiber Composites
Pipes, R. B., Blake, Jr., R. A. (1990), in: Delaware Composites Design Encyclopedia, Vol. 6. Westerville, CT: Technomic, p. 55. Proposed Standard Test Method for Mode I Interlaminar Fracture Toughness of Continuous Fiber Reinforced Composite Materials, ASTM (1990). Philadelphia, PA: Am. Soc. Test. Mat. Raju, I. S. (1987), Eng. Fracture Mech. 28, 251. Raju, I. S., Crews, Jr., J. H., Aminpour, M. A. (1988), Eng. Fracture Mech. 30, 383. Reeder, J. R., Crews, Jr., J. H. (1991), in: Proc. ICCM-8: Springer, G. S. (Ed.). Honolulu, HI. Rice, J. R. (1988), /. Appl. Mech. 55, 98. Russell, A. J., Street, K. N. (1982), in: Progress in Science and Engineering of Composites: Hayashi, T., Kawata, K., Umekawa, S. (Eds.). Tokyo: ICCM-IV, p. 279. Russell, A. X, Street, K. N. (1985), in: Delamination and Debonding of Materials. ASTM STP876. Philadelphia, PA: Am. Soc. Test. Mat., p. 349. Rybicki, E. R, Kanninen, M. F. (1977), Eng. Fracture Mech. 9, 931. Salpekar, S. A., Raju, I. S., O'Brien, T. K. (1988), /. Compos. Tech. Res. 10, 133. Sela, N., Banks-Sills, L., Ishai, O. (1989), Eng. Fracture Mech. 32, 533. Sela, N., Ishai, O., Banks-Sills, L. (1989), Composites 20, 257. Sih, G. C , Liebowitz, H. (1968), in: Mathematical Theories of Brittle Fracture, Fracture, Vol. 2: Liebowitz, H. (Ed.). London: Academic Press. Sih, G. C , Paris, P. C , Irwin, G. R. (1965), Int. J. Fracture Mech. 1, 189. Shivakumar, K. N., Crews, Jr., J. H. (1987), Eng. Fracture Mech. 28, 319. Smiley, A. I, Pipes, R. B. (1987 a), J. Compos. Mater. 21, 670. Smiley, A. X, Pipes, R. B. (1987b), Compos. Set Tech. 29, 1. Sun, C. T, Manoharan, M. G. (1987), in: Am. Soc. for Composites. Technomic, p. 49. Suo, Z. (1990a), Proc. R. Soc. London A 427, 331. Suo, Z. (1990b), /. Appl. Mech. 57, 627. Suo, Z., Hutchinson, J. W. (1989), Mater. Sci. and Eng. A 107, 135. Suo, Z., Hutchinson, J. W. (1990), Int. J. Fracture 43, 1. Suo, Z., Bao, G., Fan, R, Wang, T. C. (1991), Int. J. Solids Struct. 28, 235. Test Procedure for the End Notched Flexure (ENF) Specimen (1987), ASTM D 30.02 Round Robin. Philadelphia, PA: Am. Soc. Test. Mat. Timoshenko, S. P., Goodier, J. N. (1970), Theory of Elasticity, 3rd ed. New York: McGraw-Hill. Ting, T. C. T. (1986), Int. J. Solids Struct. 22, 965. Ting, T. C. T. (1980), / Mech. Phys. Solids 38, 505.
Trethewey, B. R., Gillespie, Jr., J. W., Carlsson, L. A. (1988), /. Compos. Mater. 22, 459. Tsai, S. W, Hahn, H. T. (1980), Introduction to Composite Materials. Westerville, CT: Technomic. Wang, S. S. (1983), /. Compos. Mater. 17, 210. Wang, S. S., Choi, I. (1983), NASA CR 172269 and 172270. Wang, A. S. D., Crossman, F. W (1980), /. Compos. Mater. Suppl. 14, 71. Wang, S. S., Mandell, J. R, McGarry, R J. (1987), Int. J. Fracture 14, 39. Wang, J. S., Suo, Z. (1990), Acta Met. 38, 1279. Wilkins, D. J. (1983), in: Proc. AGARD meeting on Characterization, Analysis and Significance of Defects in Composite Materials, Proceeding No. 355, London. Wilkins, D. X, Eisenmann, X R., Camin, R. A., Margolis, W S., Benson, R. A. (1982), in: Damage in Composite Materials. ASTM STP 775. Philadelphia, PA: Am. Soc. Test. Mat., p. 168. Williams, M. L. (1959), Bull. Seismol. Soc. America 49, 199. Williams, J. G. (1988), Int. J. Fracture 36, 101. Williams, X G. (1989), in: Application of Fracture Mechanics to Composite Materials: Friedrich, K. (Ed.). Amsterdam, New York: Elsevier, p. 3. Williams, X G. (1990), personal communication. Willis, X R. (1971), J. Mech. Phys. Solids 19, 353. Wilson, D., Carlsson, L. A. (1991), in: Physical Methods of Chemistry: Rossiter, B. W (Eds.). New York: Wiley, p. 139. Wu, E. M. (1967), J. Appl. Mech. 34, 967. Wu, E. M. (1974), in: Strength and Fracture of Composite Materials, Fatigue and Fracture, Vol. 5: Broutman, L. X (Ed.). London: Academic Press.
General Reading Application of Fracture Mechanics to Composite Materials (1989): Friedrich, K. H. (Ed.). Amsterdam: Elsevier. Delamination and Debonding of Materials: ASTM STP 876 (1985). Philadelphia, PA: Am. Soc. Test. Mat. Ewalds, H. L., Wanhill, R. H. X (1989), Fracture Mechanics. London: Edward Arnold. Hutchinson, X W, Suo, Z. (1991), Mixed Mode Cracking in Layered Materials, in: Adv. Appl. Mech. 28. Lekhnitskii, S. G. (1981), Theory of Elasticity for an Anisotropic Body. Moscow: MIR Publishers. Mathematical Theories of Brittle Fracture, in: Fracture, Vol. 2 (1968): Liebowitz, H. (Ed.). New York: Academic Press.
13 Fatigue of Fiber Composites Ramesh Talreja School of Aerospace Engineering, Georgia Institute of Technology, Atlanta, GA, U.S.A.
List of Symbols and Abbreviations 13.1 Introduction 13.2 Short Fiber Composites 13.3 Long and Continuous Fiber Composites 13.3.1 Unidirectional Composites 13.3.1.1 Loading Parallel to Fibers 13.3.1.2 Loading Inclined to Fibers 13.3.2 Laminates 13.3.2.1 Angle Ply Laminates 13.3.2.2 Cross Ply Laminates 13.3.2.3 General Laminates 13.3.3 Woven Fabric Composites 13.4 Conclusion 13.5 References
Materials Science and Technology Copyright © WILEY-VCH Verlag GmbH & Co KGaA. All rights reserved.
584 585 585 592 592 593 597 599 599 601 603 605 605 606
584
13 Fatigue of Fiber Composites
List of Symbols and Abbreviations a, c A, m E fc (c) Fc (c) ge (9) Gd {9) K Kc Xj Kth / N JVf V{
crack length constants Young's modulus density function for crack length cumulative distribution function for crack length density function for crack orientation cumulative distribution function for crack orientation stress intensity factor stress intensity factor, critical value stress intensity factor, mode I stress intensity factor, threshold value length of sampling line number of cycles number of cycles to failure volume fraction of fibers
8C gm ed b £f L e max 9 a amax (T U T S
composite failure strain fatigue limit of matrix fatigue limit, fiber debonding fatigue limit m a x i m u m strain crack orientation stress maximum stress ultimate tensile stress
CDS CSM SMC UTS
characteristic damage state chopped strand mats sheet molding compounds ultimate tensile strength
585
13.2 Short Fiber Composites
13.1 Introduction Fatigue of single-phase materials, e.g. metals and polymers, is a classical field. For fiber composites, i.e. composites in which fibers play a significant role in determining the properties, the process of fatigue is far from fully understood. Indeed, this field may be considered in its infancy since the basic concepts concerning how to characterize and describe the fatigue process are still being developed. It is often the case with fatigue work that test data are presented without satisfactory interpretations and at best some remarks are presented concerning fatigue life of a material being better or worse than another. Since for fiber composites numerous possibilities exist for combining the constituents, the need for rational interpretations of the basic fatigue characteristics is critical. This exposition will not attempt to conduct an exhaustive survey of fatigue of fiber composites. Rather, selected results will be presented and discussed to bring out the essential characteristics of the fatigue process in this class of materials. Since the fiber geometry has a major influence on the process the treatment will be divided into sections pertaining to short fiber composites, long and continuous fiber composites, laminates and woven fabric composites. Only polymeric matrices will be treated. The reader is referred to a number of reviews for supplementary information (Boiler, 1969; Owen, 1974; Hahn 1979; Stinchcomb and Reifsnider, 1979; Harris, 1986; Konur and Matthews, 1989).
polyester was reported by Owen and associates in the period 1967-79 which has been summarized in Owen (1982). The material consisted of 50 mm glass multifilament strands randomly arranged in a plane and surrounded by a polyester resin. Flat plates of the composite were made with wet lay-up technique. Fiber strands could be observed by optical microscopy with transmitted light in thin plates. Combining this with observations of polished edges of specimens the mechanisms of debonding and resin cracking were detected. Debonding of strands from matrix was the first mechanism observed in static as well as fatigue loading. The first group of strands to debond were those with their fiber axes at approximately normal to the applied load direction. With increased load level or number of cycles of a maximum load debonding of strands deviating from the normal to the load axis occurred. The debond cracks extended into resin rich areas causing resin cracking which grew with further load cycling. Figure 13-1 140 120 100 CO Q_
Separation
"o. 60
I 40
A substantial amount of work on fatigue of chopped strand mat reinforced
Debonding
20
0.1
13.2 Short Fiber Composites
Cracking v ^
1
10
102
103 104 Cycles
105
106
Figure 13-1. Stress amplitude versus cycles for polyester resin reinforced with chopped strand mats of glass indicating the onset of debonding and cracking under cyclic loading with zero mean stress (from Owen etal. 1969).
586
13 Fatigue of Fiber Composites
shows when debonding, cracking and separation of specimen occurred at various stress amplitudes in a cyclic loading with zero mean stress (Owen et al., 1969). It is noted that the stress threshold below which no debonding occurs may be usable as a conservative estimate of the fatigue limit. It was thought that improvement in the polyester resin ductility might delay the onset and development of fatigue damage in the reinforced resin. A flexibilizing agent was added to the resin to increase its strain to failure. The static strain to failure of the composite showed, however, only slight improvement, see Fig. 13-2 (Owen and Rose, 1970). The elastic modulus of the resin decreased with increasing flexibilizer content causing the composite modulus also to decrease. However, it turned out that the strain for onset of debonding remained unaffected at about 0.3% under static loading irrespective of the flexibilizer 100r
#10 CD
CO
C.S.M. and fabric laminates
0
20
40
Flexible resin addition (wt %)
Figure 13-2. Strain to failure of polyester resin with polypropylene maleate adipate flexibilizer and of its composites with chopped strand mats (CSM) and with fabric (from Owen and Rose, 1970).
content. More importantly, the stress-cycles plots for onset of debonding, e.g. that in Fig. 13-1, for all composites tested, with varying contents of flexibilizer in resin, superimposed on a single curve (or band) when strain was plotted instead of stress. Also, the threshold strain for debonding in fatigue, referred to as fatigue limit for transverse fiber debonding, came out to be a fairly constant value of 0.12%. As far as resin cracking is concerned, Owen and Rose (1970) found that no significant delay in its occurrence was observed with the addition of flexibilizer in resin. For high contents of the flexibilizer, however, it was found that cracks appeared at the ends of strands that were aligned with the loading axis slightly before initiation of transverse fiber debonding. An investigation of fatigue damage in sheet molding compounds (SMC) was reported by Wang et al. (1986) (see also Wang et al., 1987). The material investigated is denoted SMC-R50 and consists of chopped glass fibers of length 25.4 mm, calcium carbonate filler and polyester resin. The fibers are randomly oriented in the plane of the sheet and constitute 50 per cent by weight of the composite. The main difference from the material of the work described above is that the fibers are about half as short here and are more evenly spread rather than being clumped together as thick strands. The observations of damage were made by taking replicas of the specimen surface periodically during the cyclic tensile loading. The replicas were coated by sputtering and examined in a scanning electron microscope. Observations showed that in a resin-rich area microcracks formed nearly normal to the loading direction. In a fiber-dominant region with fibers aligned with the loading direction, cracks appeared between fibers,
587
13.2 Short Fiber Composites
also normal to the loading direction and were limited in length by the interfiber spacing. Debonding of fiber strands roughly normal to the loading axis was also seen. Typical micrographs showing the cracks are reproduced in Fig. 13-3. All crack types were grouped together for the purpose of statistical descriptions. This was achieved by a crack sampling procedure where cracks intersecting a straight line parallel to the loading axis were counted. The crack size and orientation were represented by the parameters c and 6, see Fig. 13-4. From the sampling statistics, distributions of the two parameters were derived assuming Weibull-type functions. The cumulative distribution function for crack length, denoted FC(Q, and its density function/c (C) are shown in Fig. 13-5 for a fatigue loading with the maximum stress amax at 60% of the ultimate tensile stress aUTS. For the same loading the cumulative distribution and density functions for crack orientation, Ge (6) and ge (9), respectively, are shown in Fig. 13-6. The main observations of interest concerning the microcrack geometry are that most cracks remain less than 1 mm in length during fatigue and that their orientations are mainly within 30° of the normal to the loading direction. The multitude of microcracks within the volume of a SMC degrade the elastic constants. Figure 13-7 shows the longitudinal Young's modulus versus the number of cycles applied. The predicted curve in the figure is calculated by a so-called self-consistent estimation scheme using the measured crack statistics (Wang et al., 1986). The preferential cracking resulting from the applied uniaxial loading leads to anisotropy in the initially isotropic SMC material. This necessitates calculation of more elastic constants than those needed to characterize the virgin state of the material.
(a)
(b)
Figure 13-3. Microcracking in a short fiber reinforced SMC-R50 composite, (a) Cracks in a resinrich region and (b) cracks in a fiber-dominant region with fibers oriented at an angle to the loading direction (from Wang et al. 1987).
ACT
ACT
Figure 13-4. (a) Specimen of a short fiber composite with a sampling line of length / parallel to the axial direction, (b) A crack intersecting the sampling line is described by length 2c and angle 6 (from Wang et al. 1987).
588
13 Fatigue of Fiber Composites
40
Fitted Exptl curves data
35
<W=0-6o" UTS
30 _
D
---
O A
25 -
o
Cycles
J.Q — Q — ^—D3-D
20 15
o
5000 Cycles 3000 1000 100
— —
10
C5)
O-C0D"O<3> O ' O O - - u u -
10 5
5000 1000 100 10
-o
.
w
I
l
1 2 Crack length (mm)
Crack length (mm)
Figure 13-5. The cumulative distribution function FG(c) and the density function fc(c) of crack length at various cycles with maximum stress
by weight of glass or graphite fibers. The matrix materials were engineering thermoplastics, including semicrystalline Nylon 66, amorphous polycarbonate, polysulfone, amorphous polyamideimide and semicrystalline polyphenylene sulfide. The fibers were reported to seldom exceed 1 mm in length and have typical length of 0.2 mm, giving a typical length to diameter aspect ratio of 20.
Studies related to fatigue of injection molded reinforced thermoplastics have been reported by Mandell et al. (1982) and (1985). These authors have studied growth of a precut crack under cyclic tensile loading. The plane of the precut crack was normal to the dominant fiber direction which was also the direction of loading. The materials tested were commercial injection molding compounds containing 30 to 40% 40 35
Fitted curves
I 30
Exptl data n 5000 Cycles o 1000
ou a
max=0-6auTs
40 ^ ^ ^
•g 2 5
E
CO
_^^K
"S 30 o
""* # '*s fc#
CO &
^15
Z
20
•••
— 5 0 0 0 Cycles 3000 1000
100 ^^w
10
">-v^
10 a> 10
5 TT/4
Crack orientation (rad)
TT/2
TT/4
Crack orientation (rad)
Figure 13-6. The cumulative distribution function G9 (6) and the density function ge (9) of crack orientation at various cycles with maximum stress on . of 60% of the ultimate tensile stress al]TS for a short fiber SMC-R50 composite (from Wang et al., 1987).
13.2 Short Fiber Composites
Predicted
growth with local failure zones ahead of the crack. The band of width (or height) h about the crack plane is suggested by the authors to be a characteristic of the material microstructure and independent of the loading mode. The authors further suggest that only processes within this band govern crack growth (i.e. a plastic zone-like concept). The crack growth rate is taken to be given by the Paris law:
—
Exptl data D O 102 103 104 Number of cycles
105
Figure 13-7. Change in the longitudinal Young's modulus of a short fiber SMC-R50 composite under fatigue at the maximum stress of 60% of the ultimate tensile stress (from Wang et al., 1987).
Optical and scanning electron microscopy were done to study the mode of crack growth. This mode was described as a fiber avoidance mode. Although differences were found from material to material, the dominant feature of the crack growth process was described as a main crack growing in a zigzag pattern avoiding cutting through fibers and instead finding local paths of lower resistance along the fibers and between the fibers. Ahead of the main crack tip local failure zones were found. These zones contained isolated crazes or isolated shear yielding. It was thought that these zones coalesce and merge into the main crack. The schematic in Fig. 13-8 depicts the fiber avoidance mode of crack Isolated yield zones or craze / cracks
___-Ar_
Fibers
589
Main crack tip zone
Figure 13-8. Schematic of crack zone in fiber avoidance mode of crack growth (from Mandell et al., 1985).
(13-1) where a is the crack length, N is the number of cycles, K is the stress intensity factor and A and m are constants. For a zigzagging crack the measurement of crack length is not simple. The zigzag pattern observed on the specimen surface does not necessarily hold through the specimen thickness. The authors give no details concerning this and it therefore appears that they use the projected length on the extended initial crack plane measured on the specimen surface. In any case, their crack growth data are reproduced in Figs. 13-9 and 13-10. It is noted that the crack growth exponent increases from the value of 4 for unreinforced matrix to 7 (for nylon 6,6) or 8 (for polysulfone). Assuming the Paris law to hold until the critical crack length to failure, the fatigue life can be obtained by integrating Eq. (13-1). The slope of the <7max — N{ line on a log-log scale would then be the inverse of the exponent in the Paris law. Figure 13-11 shows the Gmax — Nf data for unreinforced and reinforced polysulfone with the corresponding slopes indicated. The deviation at high stresses is attributed to general yielding of the materials. From the work of Mandell et al. (1982, 1985) summarized above it may be concluded that for sufficiently small fibers a
590
13 Fatigue of Fiber Composites
Kc (MNrrr3/2) -1 p Material a Unreinforced 4.1 A 4 0 % Glass 9.3 -2 O40% Carbon 10.9 o -3
E -4
-5
-6
-7
o
? -0.6
-0.4
0 0.2 0.4 0.6 Log (maximum Klt MNnr3/2)
-0.2
crack emanating in a resin-rich area may grow under cyclic loading, avoiding fibers in its path by zigzagging around them, until a critical size is attained beyond which further growth is unstable and failure reKc (MNnr 3 / 2 )
Material
D Unreinforced A 4 0 % Glass o 4 0 % Carbon
K th (MNrrr 3/2 )
1.37 2.35 2.70
5.82 11.7 17.3*
/
A
Some non colinear crack growth
A/
7
• a
E
CD •D
s
AA /
-6
m =7
o
-7
? , -0.2
0
0.2 0.4 0.6 0.8 Log (maximum Kx, MNm372)
1.0
Figure 13-10. Fatigue crack growth data for unreinforced and injection molded fiber reinforced nylon 6,6 (from Mandell et al., 1985).
1,2
0.8
1.0
Figure 13-9. Fatigue crack growth data for unreinforced and injection molded fiber reinforced polysulfone (from Mandell et al., 1985).
suits. The question remains as to what role the fibers play other than being in the way of a growing matrix crack and thereby slowing down its rate of growth. Based on some fracture toughness considerations Mandell et al. (1985) suggest that the ultimate tensile strength (UTS) of a composite, determined by the fiber strength, geometry and volume fraction, determine also the fatigue strength. The schematic of relationship between fatigue crack growth and fatigue life, suggested by Mandell et al. (1985), is reproduced in Fig. 13-12. The difficulties related to characterization of fatigue crack growth data in composites are illustrated by a systemic investigation on composites with spherical fillers reported by Gadkaree and Salee (1983). They studied fatigue crack growth in cantilever beam specimens of a thermoplastic bisphenol A-terephalate/isophthalate copolyester filled with a fly-ash filler with mean diameter of 20 jim. Standard fracture mechanics techniques of compliance calibration of crack length were used. Tests were carried out with composites
591
13.2 Short Fiber Composites 2.5
£ 2.0^
W
40 % Carbon 40 % Glass 20 % Glass
OO,
Yielding Fracture
1.5 Figure 13-11. Gm&x-Nf data
1.0
for unnotched polysulfone in unreinforced and various injected molded fiber reinforced conditions (from Mandell et al., 1985).
3 4 5 Log (Cycles to fail, Nf)
containing 10, 20, 30 and 40% by weight of filler. The crack growth data for unfilled resin are shown in Fig. 13-13 and the Paris law constants derived from best-fit are also indicated. Note that the exponent value is
fairly close to the value of 4 generally reported for polymers. The data for composite with 10% by weight of filler are shown in Fig. 13-14. The exponent here takes the value 8.78, which is not far from the values of 7 and 8 found by Mandell etal. (1985) for their injection molded composites. The data for composites with
Glass Graphite i
•a
m= 8
3
-2.60
^~
r
Corr.coeff .=• 0.9698 JO
=A(AK) f t
— -2.80 - d / V
y
o Log /Cmax
-3.00 - -3.20
o^o 7
3
^
7°
-3.40 -
/1
-3.60 -
LogA/f
Figure 13-12. Schematic of relationship between fatigue crack growth andtfmax—N{behavior of injection molded fiber reinforced composites (from Mandell etal., 1985).
-3.80 2.30
5.57108 • 10' 14 3.97990
/
i
2.50 LogAK
2.70
Figure 13-13. Fatigue crack growth data for unfilled resin (from Gadkaree and Salee, 1983).
592
13 Fatigue of Fiber Composites 4 2 - = 2.42795x10" 2 4 (A/<) 8 ' 7 7 9 4 5 dA/
-3.60
Corr. coeff.= 0.976
-3.50
u
o/ / /o
oo
B " 4 - 10
a) -3.70
3 -4.60 -3.90 -5.10 / -5.60 2.0
2.1
2.2 LogAK
-4.10 2.20
2.3
Figure 13-14. Fatigue crack growth data for resin with 10% by weight filler (from Gadkaree and Salee, 1983).
20 and 30% by weight of filler are shown next in Figs. 13-15 and 13-16. The data here refuse to be characterized by the Paris law. Finally a comparison of the data for unfilled resin and the composite with 40% by weight of filler are shown in Fig. 13-17. Once again, the composite data show lower growth rates but cannot be described by the Paris law. Apparently, there are certain mean interfiber or interparticle spacings beyond which crack growth becomes erratic and cannot be described by the stress intensity parameter.
2.22
2.24 LogAK
2.25
2.28
Figure 13-15. Fatigue crack growth data for resin with 20% by weight filler (from Gadkaree and Salee, 1983). -3.40 r" o
o
-3.80 £ -4.20 3 -4.60
_ \
°.oo
•v °\ .'/
-5.00 I-
-5.40 2.20
2.25
2.30 LogAK
i
I
2.35
2.40
Figure 13-16. Fatigue crack growth data for resin with 30% by weight filler (from Gadkaree and Salee, 1983). -2.60
13.3 Long and Continuous Fiber Composites
o .o>°
-3.10
cf°
Unfilled
B -3.60 -o o ^oo o
13.3.1 Unidirectional Composites In this section composites with aligned fibers only will be considered. Laminates with plies of aligned fibers and woven fabric composites will be treated in the subsequent sections. Mechanisms of fatigue damage in aligned fiber composites depend significantly on the mode of loading, i.e. whether
O)
o
-3 -4.10
/ Filled
-4.60
-5.10
2.38
i
2.48
2.58
2.68
Log A K
Figure 13-17. Comparison of fatigue crack growth data for unfilled resin and for resin with 40% by weight filler (from Gadkaree and Salee, 1983).
13.3 Long and Continuous Fiber Composites
it is tensile, compressive, shear or combinations of these. The mechanisms are usually first studied in individual loading modes and effects on these due to presence of the other loading modes are then investigated. Due to difficulties involved in fatigue testing, most studies reported have been for uniaxial tensile loading. The testing techniques for other loading modes are still being developed. The mechanisms of damage will be discussed separately for tensile loading parallel to and inclined to the fiber axis. 13.3.1.1 Loading Parallel to Fibers
The mechanisms of damage may be separated as those governed by fibers and those determined by the matrix properties. Since fibers are usually much stronger than matrix fiber damage occurs at high load levels, restricted mainly within the scatterband of the composite failure load. At first application of the maximum load, fibers fail at weak points which are dispersed throughout the volume of composite. An individual fiber failure leads also to failure of the fiber/matrix interface. This interface is usually strong in polymeric matrix composites and the resulting debond zone is therefore small. The cavity-like zone created by a fiber failure and the associated debond raises the local stress on the neighboring fibers, one or more of which will also fail if the local stress exceeds the strength. In a perfectly brittle composite the local stress redistributions caused by fiber and interfacial failures at the first load maximum in a cyclic load would be achieved instantaneously and would remain unchanged at subsequent load applications, assuming no frictional energy dissipation to take place. In principle, therefore, no fatigue failure is possible in perfectly brittle composites with no friction
593
mechanism. However, neat polymeric resins are ductile and although their ductility is subdued by stiff fibers under loading parallel to fibers, some local flow around broken fibers would occur. This flow would allow changes in the local stress state with repeated load application. Thus, failure of more fibers in the vicinity of fibers broken in the first load application would occur. For relatively brittle thermosetting resins, e.g. polyester and epoxy, the very limited local flow combined with the statistical nature of fiber failure suggests that the fiber failure process may, for all practical purposes, be assumed non-progressive. In other words, a given site of broken fiber would not nucleate more broken fibers at a well-defined rate (per cycles of load). Instead, failure would occur at random from a broken fiber site which happens to find sufficient weak fibers in its neighborhood to form a core of fiber failures large enough to grow unstably at the applied maximum load (see Fig. 13-18). Let us now consider fatigue damage at cyclic loads with a maximum load at sufficiently low value to cause essentially no failure of fibers. The fibers would now cyclically deform and the matrix would be cycled between the strain limits allowed by the fibers. This would result in a straincontrolled fatigue of the matrix. The polymeric matrix would initiate cracks if the strain range is sufficient for the purpose. If cracks initiate, they would grow until the crack fronts encounter fibers (see Fig. 1319). The strain in the composite at the maximum cyclic load below which the matrix cracks remain arrested by fibers would then be the fatigue limit of the composite. If the maximum cyclic load subjects the matrix to higher strain than the fatigue limit, the matrix cracks will either fail fibers at their fronts or bypass fibers by
594
13 Fatigue of Fiber Composites N =1
N =Ni
N =Nf
Jxrndii Figure 13-18. A schematic describing the non-progressive fiber failure in fatigue of unidirectional composites under loading parallel to fibers.
debonding the interfaces, in which case, fibers could fail by pullout from the matrix. A matrix crack is shown schematically during its growth process in Fig. 13-20. The final failure (separation) is expected to result from unstable growth of a matrix crack. The mechanisms of fatigue damage described above are based on logical deductions rather than direct observations. Techniques of direct observations, e.g. microscopy, are less easily performed on polymeric matrix composites than on metals and the observations are often intractable due to the complexity of the operative mechanisms. To facilitate interpretation of the fatigue behavior, therefore, a
conceptual framework in the form of a fatigue life diagram has been proposed (Talreja, 1981). The diagram for tension-tension loading of unidirectional fiber composites is shown in Fig. 13-21. The axes of the diagram are strain and logarithm of the load cycles to failure. Although fatigue testing is done under controlled load, the variable on the vertical axis of the diagram is the maximum strain attained in the first load cycle. The significance of the maximum strain is that this quantity represents the state of damage reached in the first load cycle and it is reasonable to expect that any progression of damage in subsequent load cycles will be determined by this state of damage. Furthermore, the two
o O ( o o o c G O o c# ( fO (w? o o o \o c o c I/O o o §1 D O o
i ki
W°
U
Matrix cracks
Figure 13-19. Matrix cracking in unidirectional composites arrested by fibers such that the crack planes are confined to resin-rich areas.
13.3 Long and Continuous Fiber Composites
I
1 1
11i 1 I n
;•
Ii
Pi
i
595
\ Crack front (A/ = A / f )
Crack front (N =A/ 1 )
P
Fiber - bridged matrix crack
extreme states in fatigue, i.e. the static failure and the fatigue limit, are given generically in terms of strain. The static failure occurs at the strain to failure of fibers irrespective of the fiber volume fraction and the fatigue limit is governed by the matrix which is undergoing strain-controlled fatigue, as noted above. The scatter-band located horizontally about the composite failure strain represents the assumed non-progressive fiber failure mechanism likely to occur in composites with thermosetting resins, as discussed above. The sloping scatter-band stands for the progressive cracking of matrix and the associated fiber/matrix interfacial failure. The role of fibers in this mechanism is to suppress the rate of matrix crack growth by bridging the crack planes. The location of the scatter-band, i.e., where it deviates from the fiber failure scatter band, is governed by the effectiveness of the fiber bridging mechanism. The lower bound to the fatigue life diagram is the fatigue limit, which, as discussed above, is a matrix property affected by the geometric configuration in which the fibers are distributed in the cross-sectional plane. Experimental data (for a glass-epoxy composite) seem to indicate that the fatigue limit of composite is practically the same as that of the neat resin (Dharan, 1975). In any case, the fatigue
Figure 13-20. Matrix cracking in unidirectional composites aided by fiber/matrix debonding, fiber pullout and fiber failure.
limit of the neat resin would be the lowest limit and therefore it may be useful to operate with this quantity for design purposes. In order to appreciate the usefulness of the fatigue life diagram consider two unidirectional composites with the same matrix but different fibers. Let the fibers be of relatively low stiffness in one case and of relatively high stiffness in the other. Figure 13-22 illustrates schematically stress-strain curves for the two cases. The composite failure strain, denoted by ec and given by fiber failure strain, is different in the two cases while the fatigue limit, denoted by Fibre breakage, interfacial debonding
Matrix cracking, interfacial shear failure
Y///////////////////////A V
log/V
Figure 13-21. Fatigue life diagram of unidirectional composites in tension-tension fatigue under loading parallel to fibers (from Talreja 1981).
596
13 Fatigue of Fiber Composites
Fibre
I
(a) Figure 13-22. Stress-strain characteristics of unidirectional composites: (a) low stiffness fibers, (b) high stiffness fibers.
8m, would remain the same in the two cases if it is assumed to be equal to the neat resin fatigue limit. The two strain values, am and ec, are further apart in the first case than in the second. Indeed, for fibers of very high stiffness the composite failure strain may even be less than the fatigue limit strain. To illustrate the significance of fiber stiffness (and the resulting failure strain) on fatigue behavior, data for a glass-epoxy composite are shown in Fig. 13-23. Note first that fatigue lives for composites of three different volume fractions fall within the same scatter-band, while three different and separate scatter-bands result when the usual stress-life plots are made (Dharan, 1975). The fatigue limit of neat resin was found by Dharan (1975) to be at 0.6%
Figure 13-23. Fatigue life diagram of glass-epoxy under loading parallel to fibers; data from Dharan (1975).
strain, which is the value set in the fatigue life diagram. Remarkably different behavior is displayed by a graphite-epoxy composite whose strain to failure is only about 0.5%. Assuming that the fatigue limit of epoxy is at least 0.6% strain, the fatigue life diagram takes the shape shown in Fig. 13-24. The sloping scatter-band of progressive matrix and interfacial damage disappears, as the fatigue limit of the resin becomes unreachable. Thus, a logical interpretation of the fatigue damage tolerance of ultra-high modulus graphite fiber composites emerges. Figure 13-25 shows the fatigue life diagram of a high modulus graphite-epoxy composite. The horizontal scatter-band (limited by 5% and 95% failure probability levels) is unusually wide here, presumably due to poor quality of fibers and/ or composite, and the fatigue limit is once again set at 0.6% for epoxy. The apparent fatigue damage tolerance may now be interpreted differently from the behavior displayed by data in Fig. 13-24. Data for a lower modulus graphite/epoxy composite are shown plotted in the fatigue life diagram, Fig. 13-26. Note here that the sloping scatter-band is further out to the right than that for glass-epoxy composites, Fig. 13-23. This is due to more effective fiber bridging action by graphite fibers than is possible by the less stiff glass fibers.
13.3 Long and Continuous Fiber Composites
597
13.3.1.2 Loading Inclined to Fibers
0.006
•->
;
T
•
>v k • •
•
0.004
0.002 0
4
8
log N
Figure 13-24. Fatigue life diagram of an ultra-high modulus graphite-epoxy under loading parallel to fibers; data from Sturgeon (1973).
0.010
0.006
y/////////////////////////
0.002 0
4 log N
Figure 13-25. Fatigue life diagram of high modulus graphite-epoxy under loading parallel to fibers; data from Awerbuch and Hahn (1977).
When loading is not parallel tofibersthe fiber/matrix interface is subjected to normal and shear stresses. The interfacial region (interphase) initiates cracks under these stresses which then grow on repeated load application aided by irreversible dissipative processes at crack fronts. The mixed-mode crack growth in interfacial region becomes the primary progressive fatigue damage mechanism. The role of fibers under loading inclined to fibers is only significant for small inclination angles (probably less than 5°). As the inclination angle increases the normal stress in fibers falls rapidly. At the same time, normal and shear stresses in the interfacial region increase. The openingmode contribution to the mixed-mode crack growth also increases with increasing inclination angle. Since the crack growth resistance is expectedly less for opening-mode growth, the threshold strain for crack growth will decrease with increasing inclination angle. Consequently, the fatigue limit will decrease as the inclination angle increases. The above-mentioned considerations are reflected in the schematic fatigue life diagram, Fig. 13-27. The fatigue life dia-
Figure 13-26. Fatigue life diagram of a lower modulus graphite-epoxy under loading parallel to fibers; data from Curtis (1987).
598
13 Fatigue of Fiber Composites
_I_I_I_I_I_I_I
(0°< 6 <90°)
Mixed-mode matrix and interfacial damage
Figure 13-27. Fatigue life diagram for unidirectional composites under loading inclined to fibers. Dottedline diagram is for loading parallel to fibers.
Transverse fibre debonding
gram for inclination angle 9 = 0°, i.e. for loading parallel to fibers, is shown as a reference by dashed lines. For an inclination angle of more than a few degrees, where fiber failure ceases to be a governing mechanism, the sloping scatter-band starts at the static failure strain and ends at the fatigue limit. The horizontal scatter-band of fiber failure is lost and, consequently, the mechanism for delaying the progressive damage is also not effective. In the limiting condition of inclination angle, 0 = 90°, the opening-mode crack growth, described as transverse fiber debonding, is the predominant mechanism of damage. Test data for a glass-epoxy unidirectional composite subjected to tension-tension fatigue at various inclination angles have been reported by Hashin and Rotem (1973). Their data, reported in terms of the maximum stress, have been replotted with maximum first-cycle strain in Fig. 13-28. The lines drawn through the data-points have been started at the static failure strain. The fatigue limit of epoxy, reported by Dharan (1975), is shown as a reference fatigue limit for 0 = 0°. The fatigue limit for transverse fiber debonding was reported by Owen and Rose (1970) for various
composites of reinforced polyester to be at 0.12% strain. The data for glass-epoxy shown in Fig. 13-28 suggest that this fatigue limit is approximately at the same value. In Fig. 13-29 the strains at fatigue limits for various inclination angles are plotted. The strain values are taken from the fa^/deg
0.008
°5
> = 5°)
+
•
0.006
10 30 60
/ / / / /
/ / /v / / / / / X
= 10°) 0.004
v O
cc(0 = 30°) 0.002 ec(d = 60°)
•^—A—_
/ / / / / / / / /, 1
0
1
2
1
1
4 log N
6
Figure 13-28. Fatigue life diagram of a unidirectional glass-epoxy under loading inclined to fibers; data from Hashin and Rotem (1973).
13.3 Long and Continuous Fiber Composites
0.006
0.004
0.002 -
30
60
90
0/deg
Figure 13-29. Variation of the fatigue limit ef#1# with the inclination angle, 6, based on data of Hashin and Rotem (1973).
599
In order to achieve rational and efficient design procedures for laminates under fatigue the first step is to generate basic understanding of the damage development process. In the following this aspect will be reviewed and the role of damage mechanisms will be interpreted with the aid of fatigue life diagrams. For the sake of clarity and systematic description, laminates of angle ply and cross ply types will be considered before discussing more general laminates. 13.3.2.1 Angle Ply Laminates
tigue life diagrams, Fig. 13-28. A smooth curve drawn through the points is made to pass through the reference values at the two extremes, i.e. 6 = 0° and 0 = 90°, at 0.6% and 0.12%, respectively. 13.3.2 Laminates A primary motivation for constructing laminates is to achieve desired combination and distribution of properties. As far as the elastic response is concerned, wellestablished plate theories exist for making assessment of the stiffness properties. A number of strength criteria have been proposed and, although a single criterion has not been found to be universally applicable, sufficient basis exists for evaluation, at least a preliminary one, for selection of laminate configurations. For long-term performance of laminates, in particular under cyclic loads, the situation is far from satisfactory. A selection of laminate configuration on the basis of fatigue performance cannot yet be made. The design approach, instead, has been to select laminate configurations on the basis of stiffness and strength considerations and then to assure that these properties do not degrade unacceptably under fatigue.
This class of laminates is widely used in filament wound structures. It is also suited for systematic studies of the so-called constraint effect in laminates. For illustration of this effect, consider the fatigue damage of unidirectional composites under loading inclined to fibers. As discussed above, the predominant damage mechanism here is the mixed-mode cracking in the fiber/ matrix interfacial region. Indeed, failure will occur when a single crack grown from a preexisting flaw reaches a critical size, at which stage further growth becomes unstable, see Fig. 13-30 a. Now consider a unidirectional ply placed in an angle ply laminate subjected to loading along the symmetry axis, see Fig. 13-30b. Under the effect of the stresses in the plane of a ply mixed-mode cracking will also occur here. However, growth of a crack here will be under constraint of the adjacent ply (or plies) and failure of the laminate will not depend on the length of the crack since the constraining ply (or plies) will locally (i.e. at crack fronts) carry the load shed by the cracking ply. If fibers in a constraining ply sustain the additional load, which acts over a distance from the crack planes, called the shear-lag distance, more cracks can form further away. This multiple
600
13 Fatigue of Fiber Composites
(a)
- - Initial flaw Fatigue crack
(b)
- Matrix crack Delamination
Figure 13-30. Comparison of fatigue damage mechanisms in unidirectional and angle ply laminates, (a) Cracking from initial flaw in a unidirectional ply under loading inclined to fibers, (b) Multiple matrix cracking in one ply of an angle ply laminate under the constraint of an adjacent ply of opposite orientation. The intraply cracks cause delaminations along the crack fronts.
cracking process has been treated by Aveston et al. (1971) and since then by many others. In angle ply laminates loaded along a symmetry axis the multiple cracking process takes place in plies of both orientations since the plies of one orientation apply constraint to the plies of the other ori-
entation and simultaneously crack under the constraint of plies of that orientation. Under cyclic tensile loads of a constant amplitude the multiple cracking process in angle ply laminate progresses causing increase of the crack number density. If no other damage mechanism is initiated the multiple intraply cracking process will stop when saturation spacings between cracks in plies of both orientations have been reached. However, the intraply crack fronts can locally debond the ply interfaces if the debond energy is exceeded by the local strain energy release rate. The delaminations thus caused (see Fig. 13-3Ob) can grow under cyclic loads. The stress states in plies in the delamination regions will then change and may cause fiber failures and thereby failure of laminate. The delaminations can also merge together by growth and cause laminate failure by separating plies from each other sufficiently such that linkage of intralaminar cracks throughout the thickness is achieved. The process of damage in angle ply laminates has not been quantitatively analyzed for predicting fatigue life. It can, however, be appreciated, from description of the damage process given above, that the constraint effect (determined by ply thickness, orientation and stiffness) governs initiation, progression and eventual saturation of intraply cracking. The associated local delamination process is also affected by the ply constraint since the surface displacements of the intraply cracks are subject to this constraint. For given ply material and stacking sequence the constraint effect can be made to vary by varying the off-axis angle of angle ply laminates. Fatigue life data for angle ply laminates of glass-epoxy with the off-axis angle (denoted 6 in Fig. 13-30) of various values between 30° and 60° were reported by Rotem and Hashin (1976). The fatigue limits tak-
13.3 Long and Continuous Fiber Composites
en from their data have been plotted as strains (by dividing their stress by their reported elastic moduli) and are shown in Fig. 13-31. Also shown in the figure are fatigue limit data, connected by dotted line, for unidirectional composites of same material under loading inclined to fibers. Comparison of the two sets of data gives a good illustration of the constraint effect discussed above. A significant improvement in the fatigue performance due to the constraint effect (or simply the fiber architecture) is thus displayed. (The nature of variation of the fatigue limit of angle ply laminates of glass-epoxy, shown in Fig. 13-31, has been found to be slightly different by a recent study, yet to be published. The improvement over unidirectional composites remains of about the same magnitude.)
0.006 -
0.004
0.002
30
60 0/deg
Figure 13-31. Variation of the fatigue limit s{l of an angle ply laminate of glass-epoxy. Fatigue limits are taken from the data reported by Rotem and Hashin (1976). The dotted line reproduces fatigue limit variation for unidirectional composites, Fig. 13-28.
13.3.2.2 Cross Ply Laminates This class of laminates, although of less practical application than the angle ply laminates, has been the subject of extensive analytical, numerical and experimental studies. The reason is that the multiple transverse cracking in these laminates is
601
constrained by longitudinal plies that remain uncracked in most cases on loading in longitudinal direction, while in angle ply laminates both sets of off-axis plies crack simultaneously. Thus, the multiple transverse cracking can be studied under the constant constraint of the longitudinal plies. The constraint can, however, only be varied by varying the ratio of thickness of the two sets of plies, if the same ply material is used throughout the thickness. The studies related to the cracking behavior of cross ply laminates have been concerned with three aspects: the initiation of cracking, also called the first ply failure, the development, i.e. multiplication, of cracking, and the changes (reduction) of stiffness properties due to cracking. A recent stateof-the-art review of this subject is given in Han and Hahn (1989). Most studies reviewed there have treated monotonic tensile stressing of cross ply laminates along longitudinal plies. A thorough study of tensile fatigue of cross ply laminates of graphite-epoxy has been reported by Jamison et al. (1984). The evolution of fatigue damage was described by separating it in three stages, the reflection of which was shown by the modulus reduction curve, Fig. 13-32, where the three stages are marked as I, II and III. Stage I consists of transverse crack formation which progresses with the cycles applied until a fairly constant spacing between cracks is attained. Figure 13-33 shows the transverse crack density with cycles and the associated reduction in the longitudinal Young's modulus of the laminate. Damage in Stage II is exemplified by an X-ray radiograph shown in Fig. 13-34. The horizontal lines are the transverse cracks in 90° plies and the vertical lines represent longitudinal cracks between fibers in 0° plies. These cracks initiate in Stage I but are fewer in number and short-
lus
602
0.99
o 0.97
13 Fatigue of Fiber Composites
M.
c 0.95 o 0.93 ized
§
0.91
P3
o
0.89 0.87
5
10
15 20 25 30 35 40 45 Thousands of cycles
Figure 13-32. Development of damage in (0, 902)s laminate of graphite-epoxy reflected as Stages I, II and III in the modulus reduction curve (from Jamison et al., 1984).
1.00 r Normalized secant modulus
Transverse crack density
0
10
cut at various places along a delamination it was found that a delamination was formed by diversion of a longitudinal crack into the interlaminar plane. Stage III of damage was found to be dominated by growth and coalescence of the delaminations and fiber failures in the delaminated regions. The delamination growth was primarily in the longitudinal direction. The coalescence of such delaminations formed strips of width equal to the spacing between adjacent longitudinal cracks. Failure of such strips by fiber breakage in them induced the laminate failure. It may be argued that if the maximum strain attained in the first application of a tensile load on a given cross ply laminate does not exceed the threshold strain for multiple transverse cracking, subsequent application of the same load will not cause fatigue failure (in, say, 106 cycles). Thus, the strain to onset of multiple cracking
-
20 30 40 Thousands of cycles
f:
Figure 13-33. Development of transverse crack density and modulus under fatigue of (0, 902)s graphiteepoxy laminate (from Jamison et al., 1984).
er in length. The initiation of these cracks is attributed to the tensile stress normal to fibers resulting from the difference of the Poisson's ratio in the two ply directions. Subsequent to the appearance of the transverse and longitudinal cracks delaminations of roughly elliptical shape were found located around the crossover points of the two sets of cracks. These delaminations appear as shadow-like regions in the X-ray radiographs, e.g. Fig. 13-34. By scanning electron microscopy of sections
Figure 13-34. X-ray radiograph showing transverse cracks (horizontal lines), longitudinal cracks (vertical lines) and local delaminations (shadow-like zones) in (0, 902)s graphite-epoxy laminate under fatigue, Stage II (from Jamison et ah, 1984).
13.3 Long and Continuous Fiber Composites
may be taken as a good estimate of the fatigue limit. The static (or one cycle) failure limit is given by the fiber failure strain, which is also the composite failure strain. A fatigue life diagram of a cross ply laminate will thus have these two strains (i.e. strain to multiple transverse cracking and strain to fiber failure) as the two limits. Figure 13-35 shows an example of the fatigue life behavior of a cross ply laminate. The data taken from Grimes (1977) are interpreted on the basis of the fatigue
Debonding in 90° plies.delamination
603
life diagram concept. The fatigue limit is placed at the strain of 0.43% which was the strain at which the plot of longitudinal stress versus transverse strain showed a proportionality limit indicating onset of (multiple) transverse cracking. A scatter band is placed horizontally about the static failure strain to indicate presence of the nonprogressive fiber failure mechanism discussed above in connection with tensile fatigue of unidirectional composites under loading parallel to fibers. In a cross ply laminate this mechanism will also exist due to presence of the 0° plies stressed parallel to fibers. 13.3.2.3 General Laminates
0.008
0.004
0
Figure 13-35. Fatigue life diagram for a cross ply laminate of graphite-epoxy; data from Grimes (1977). emc, the strain to multiple cracking, is the fatigue limit.
I - matrix cracking
Numerous laminate configurations can be obtained by combining the unidirectional, angle ply and cross ply arrangements. A number of these have been studied by various workers. A summary of an extensive range of investigations has been reported by Reifsnider et al. (1983). Figure 13-36 shows the schematic description of
Controlling damage modes during fatigue life 3 - delamination
2 - crack coupling - interfaciat debonding Percent of life
5 - fracture
4 - fiber breaking
100
Figure 13-36. Development of damage in composite laminates (from Reifsnider etal., 1983).
604
13 Fatigue of Fiber Composites
the controlling damage mechanisms suggested by them for a wide class of laminates. The initial stage of damage development consists of multiple matrix cracking along fibers in plies inclined to the applied load direction. The nature of this cracking process is as that described above for angle ply laminates and cross ply laminates. The saturation of the intraply cracking process in individual plies has been found to lead to a cracking state which is a generic property of the given laminate and is independent of the loading amplitude or in general the path of loading. This state has been called the characteristic damage state (CDS) and signifies termination of the initial stage of damage development. Further damage on continued cycling occurs by local debonding of interply bonds at the fronts of the intraply cracks. This causes delamination which grows in the interlaminar planes leading to coalescence of adjacent delamination zones and stress enhancement in the separated plies. This enhances fiber breakage and induces instability of the damage developments leading to final failure. The complexity of the damage development process has not yet allowed a quanti1.8
[
Graphite / epoxy
T 800 / 5245
1.5
1.2
•
•o c
D Unidirectional 0.9
tative description of the rates of the process. A pragmatic approach at present is therefore to search for trends in the fatigue lives induced by changes of ply orientations from a basic configuration. As an example of this consider three laminate configurations: a undirectional, (0, ±45) s and (0, ±45, 90)s all of the same graphiteepoxy. Fatigue life data for these laminates obtained under tension-tension loading have been kindly given to the author by Dr. P. T. Curtis of Royal Aerospace Establishment, U.K. In Fig. 13-37 the data have been plotted together on axes of the maximum strain in first cycle and logarithm of the number of cycles to failure. The fatigue life diagram has been drawn in light of the data for the unidirectional composite. The usual characteristic features of the diagram discussed above are present. It is of interest to see that fatigue lives for (0, ±45) s and (0, ±45, 90)s laminates fall within the lower part of the sloping scatterband of the unidirectional composite. This suggests that at relatively low strains the progressive damage mechanism of matrix cracking and the associated interfacial failure in unidirectional composites dominate and grovern the fatigue life. At higher strains the deviation in fatigue lives of the two laminates from the sloping scatterband of the unidirectional composite may now be interpreted as due to the delamination growth in the laminates causing enhanced stresses in the 0° plies and the consequent lower fatigue lives.
o (0,±45)s • (0, ±45,90 )s
0.6 log N
Figure 13-37. Fatigue data (P.T. Curtis, Royal Aerospace Establishment, U.K.) of three graphite-epoxy composites. The drawn fatigue life diagram is for unidirectional composite only.
605
13.4 Conclusion
13.3.3 Woven Fabric Composites
Woven fabric composites are a useful class of composites in particular for structures with large thickness. Little work on fatigue behavior of these composites has, however, been reported. A systematic study has been conducted by Schulte et al. (1987), which will be summarized below. Schulte et al. (1987) tested 8-harness satin weave of graphite fibers impregnated by an epoxy resin. Laminates of this material were made by stacking the prepregs and curing them. For comparison continuous fiber cross ply laminates of the same material were made. Development of damage in the laminates under tension-tension fatigue was studied by conducting X-ray radiography intermittently during fatigue and by post-failure scanning electron microscopy of the interlaminar regions. The development of damage in the initial stage showed basically the same features as have been observed in cross ply laminates. Cracks were found to initiate in the warp (transverse to loading) direction along the fibers. On continued loading longitudinal cracks appeared between fibers in the fill direction in the undulation regions where the fill fibers cross over the warp fibers. The crack densities increased with load cycling and a uniformly distributed pattern of orthogonal cracks appeared to develop. On further cycling delaminations were observed confined primarily to the undulation regions. Thus a uniform distribution of delamination areas in the interlaminar planes developed. Toward the end of fatigue life fiber bundles were found to fail in the undulation regions. A progressive weakening of the laminate strength then led to final failure. Figure 13-38 shows the Young's modulus change E/Eo with fatigue life for the woven fabric composite. Data for the
o Continuous fibre
1.0
N f = 799700 Cyc UJ
Matrix cracking
'O-O-O
I
• 8 - harness satin fabric \
0.8
N f = 763475 Cyc Delamination at interlaced regions
55 0.7
20
40 60 Fatigue life, %
80
100
Figure 13-38. Change in the longitudinal Young's modulus of a woven fabric composite and a cross ply laminate of graphite-epoxy under fatigue; data from Schulte etal. (1987).
straight fiber cross ply laminate are also plotted in the figure for comparison. The differences in the two systems appear beyond about 50% of fatigue life suggesting that the delaminations at the undulating regions in the woven fabric laminate are responsible for the extra reduction of stiffness. The fatigue limit of the woven fabric laminate was found at 0.62% strain as against the value of 0.85% for the straight fiber cross ply laminate. Although the reasons for this difference have not been explored, it is likely that the cause of this may lie in the reduced constraint to transverse cracking provided by the curved fibers in the woven fabric composite.
13.4 Conclusion This chapter has been a selected exposition of the fatigue behavior of fiber reinforced composites. Of major focus have been the mechanisms of damage - multitudes of cracks of various size, shape and orientation - operating in composites with different fiber reinforcement configura-
606
13 Fatigue of Fiber Composites
tions. For short fiber composites matrix crack growth characteristics dependent on the fiber size have been described. The role of fiber/matrix debonding has also been discussed. For long and continuous fiber composites the fiber orientation with respect to the loading direction has been seen as the major factor governing initiation and growth of cracking. The role of fibers has been described as the primary loadbearing constituent and as a constraining element in growth of matrix cracks. The method of plotting fatigue life diagrams as a means of interpreting the governing mechanisms of fatigue failure has been discussed. Various examples have been given that elucidate the use of these diagrams. A particular detail of practical use is the fatigue limit whose variation with fiber orientation and configuration has been described. A large emphasis has recently been placed on toughened polymers, e.g. modified epoxies and some thermoplastics. Fatigue behavior of composites of these polymers has not been treated here. Many interesting features have already emerged and more are being reported as work on these composites continues.
13.5 References Aveston, I , Cooper, G. A., Kelly, A. (1971), in: Conference on the Properties of Fibre Composites, National Physical Laboratory. Guildford, Surrey: IPC Science and Technology Press, pp. 15-26. Boiler, K.H. (1969), in: Composite Materials: Testing and Design, ASTM STP 460, Philadelphia: American Society for Testing and Materials, pp. 217235. Dharan, C.K.H. (1975), in: Fatigue of Composite Materials, ASTM STP 569, Philadelphia: American Society for Testing and Materials, pp. 171188. Gadkaree, K. P., Salee, G. (1983), Polymer Composite 4, 19.
Grimes, G. C. (1977), in: Composite Materials: Testing and Design, ASTM STP 617. Philadelphia: American Society for Testing and Materials, pp. 106-119. Hahn, H. T. (1979), in: Composite Materials: Testing and Design, ASTM STP 674: Tsai, S.W. (Ed.). Philadelphia: American Society for Testing and Materials, pp. 383-417. Han, Y.-M., Hahn, H.T. (1989), Ply Cracking in Composite Laminates, CMTC-8937. University Park, Pennsylvania: Composite Manufacturing Technology Center. Harris, B. (1986), Engineering Composite Materials. London: Institute of Metals. Hashin, Z., Rotem, A. (1973), J. Comp. Mats. 7, 448. Jamison, R. D., Schulte, K., Reifsnider, K. L., Stinchcomb, W.W. (1984), in: Effects of Defects in Composite Materials, ASTM STP 836. Philadelphia: American Society for Testing and Materials, pp. 21-55. Konur, O., Matthews, EL. (1989), Composites 20, 317. Mandell, I E , Huang, D.D., McGarry, F.J. (1982), in: Short Fiber Reinforced Composite Materials, ASTM STP 772: Sanders, B.A. (Ed.). Philadelphia: American Society for Testing and Materials, pp. 3-32. Mandell, IF., McGarry, E l , Li, C.G. (1985), in: High Modulus Fiber Composites in Ground Transportation and High Volume Application, ASTM STP 873: Wilson, D. W. (Ed.). Philadelphia: American Society for Testing and Materials, pp. 36-50. Owen, M.I (1974), in: Composite Materials, Volume 5, Fatigue and Fracture: Broutman, L.I (Ed.). New York: Academic Press, pp. 341-369. Owen, M.I (1982), in: Short Fiber Reinforced Composite Materials, ASTM STP 772: Sanders, B.A. (Ed.). Philadelphia: American Society for Testing and Materials, pp. 64-84. Owen, M.I, Rose, R.G. (1970), Modem Plastics 47, 130. Owen, M.J., Smith, T.R., Dukes, R. (1969), Plastics and Polymers 37, 227. Reifsnider, K. L., Henneke, E. G., Stinchcomb, W. W., Duke, I C . (1983), in: Mechanics of Composite Materials, Recent Advances: Hashin, Z., Herakovich, C.T. (Eds.). New York: Pergamon, pp. 399-420. Rotem, A., Hashin, Z. (1976), AIAA J. 14, 868. Schulte, K., Reese, E., Chou, T.-W. (1987), in: Sixth International Conference on Composite Materials, Second European Conference on Composite Materials, Volume 4: Matthews, F.L., Buskell, N.C.R., Hodgkinson, IM., Morton, I (Eds.). London: Elsevier, pp. 89-99. Stinchcomb, W.W, Reifsnider, K.L. (1979), in: Fatigue Mechanisms: ASTM STP 675: Fong, IT. (Ed.). Philadelphia: American Society for Testing and Materials, pp. 762-787. Talreja, R. (1981), Proc. R. Soc. Lond. A378, 461.
13.5 References
Wang, S.S., Chim, E.S.-ML, Suemasu, H. (1986), J. Appl. Mech. 53, 339. Wang, S.S., Suemasu, H., Chim, E.S.M. (1987), /. Comp. Mats. 21, 1084.
General Reading Fong, J. T. (Ed.) (1979), Fatigue Mechanisms, ASTM STP 675. Philadelphia: American Society for Testing and Materials.
607
Hahn, H. T. (Ed.) (1986), Composite Materials: Fatigue and Fracture, ASTM STP 907. Philadelphia: American Society for Testing and Materials. Lagace, P. A. (Ed.) (1989), Composite Materials: Fatigue and Fracture, 2nd Conf, ASTM STP 1012. Philadelphia: American Society for Testing and Materials. Lauraitis, K. N. (Ed.) (1981), Fatigue of Fibrous Composite Materials, ASTM STP 723. Philadelphia: American Society for Testing and Materials. Reifsnider, K. L. (Ed.) (1991), Fatigue of Composite Materials, in: Composite Materials Series, Vol. 4: Pipes, R. B. (Ed.). Amsterdam: Elsevier. Talreja, R. (1987), Fatigue of Composite Materials. Lancaster, PA: Technomic Publishing Co.