Você está na página 1de 20

International Encyclopedia of Composites, Ed.: L. Nicolais Revised contribution by J.

Karger-Kocsis Tshwane University of Technology, Faculty of Mechanical Engineering and Built Environment, Department of Polymer Technology, Pretoria, 0001, Republic of South Africa, and Budapest University of Technology and Economics, Faculty of Mechanical Engineering, Department of Polymer Engineering, H-1111, Budapest, Hungary E-mail:karger@pt.bme.hu

Structure and Fracture Mechanics of Injection-Molded Composites The history of fiber-reinforced thermoplastic polymers began only a few decades ago, when industrial production of the reinforcing fibers (glass in 1935, carbon in 1959, aramid in 1971) and adequate matrix polymers (e.g. polyamide 6.6 in 1938 and polyethylene terephthalate in 1955) was started. Incorporation of discontinuous fibers into thermoplastics generally yields improvements in mechanical and thermal properties, for instance, stiffness, strength, dimensional stability, service temperature, resistance to creep and fatigue. These improvements are, however, connected with reduced strain (ductility) characteristics and pronounced anisotropy as a result of the structuring of the reinforcement in the molded parts. Fiber reinforcement is a way to make special or engineering thermoplastics from commodity or high volume thermoplastics, such as polypropylene (PP). Note that the criteria for engineering thermoplastics namely, continuous service temperature above 100C and tensile strength higher than 40 MPa [1] can also be met by several plastics without reinforcement. The relative high annual growth rate of fiber-reinforced composites compared to neat plastics is the result of a total or partial substitution of metallic and ceramic parts by injection-molded composite items manufactured from discontinuous fiber-reinforced polymers. To characterize these traditional construction materials that are being replaced by polymer composites, fracture mechanical methods are widely used. Therefore, it seems obvious that fracture mechanical approaches should be used for plastics and composites, too. This topic has been treated comprehensively in books and papers [2-9]. Fracture mechanics yield intrinsic or material parameters that can be reliably used for the design and construction of composite parts. The main advantage of this concept is that material parameters determined in different ways can be compared with one another directly. This is not the case with standardized test methods related to a given property, such as toughness. Toughness values derived by different standard methods can hardly be compared with one another because of differences in the loading conditions. 1

It is obvious that the mechanical performance of continuous-fiber- or fabric-reinforced polymers is superior to that of discontinuous-fiber versions. This disadvantage, which is attributed to restricted load transfer between the matrix and the fibers, is compensated by other benefits, i.e. by design freedom, easy processing via injection and extrusion moldings. Therefore, it is not surprising that the development of discontinuous fiber-reinforced thermoplastics is well reflected by a steady increase in the aspect ratio (length to diameter, l/d) of the fibers both in the parent granules and molded parts. The l/d ratio of short fiberreinforced thermoplastics (SFRTPs) produced by extrusion melt compounding technique was 20 earlier, nowadays it lies at 50. The next milestone in the development of SFRTPs was achieved by pultrusion and powder coating techniques, through which granule size fiber length was set. The related products are termed to long fiber-reinforced thermoplastics (LFRTPs). In their injection and compression moldable grades the initial aspect ratio of the discontinuous reinforcement (usually glass fiber) is 1000 and 2500, respectively. It is doubtless true that the microstructure of injection-molded composites strongly depends on the processing mode and its conditions. It is also well known that the mechanical properties of plastics depend on the testing conditions, especially frequency and temperature. Therefore, these aspects have to be considered when the fracture and failure performance of discontinuous fiber-reinforced thermoplastics are discussed [10]. The mechanical performance of discontinuous fiber-reinforced thermoplastics is affected by the followings (cf. Figure 1): 1. Composition and morphology 2. Type and amount of the reinforcement 3. Interface (or interphase) between matrix and reinforcement 4. Processing methods and conditions 5. Testing conditions There is a strong interrelation amongst items 1) to 5). For example both the matrix morphology and reinforcement structuring may be highly dependent on the processing methods as in the case of injection molding. On the other hand, the type and amount of the reinforcement dictate the selection of both suitable processing methods and conditions. Items 1) to 5) list some matrix-, reinforcement-, interface- processing- and testing-related factors and serve at the same time as an outline for this contribution. The mechanical tests are grouped into static and dynamic fracture with monotonic increasing load and static and dynamic (cyclic) fatigue measurements. The test results are interpreted based on fracture mechanical concepts. FIGURE 1 Factors influencing the fracture mechanical performance of discontinuous fiberreinforced thermoplastic composites

Development of Microstructure Changes in the molecular orientation and crystallization behavior in neat and matrix polymers of (!)"#T$s occur during processing. This is accompanied with fiber structuring (i.e. orientation and layering) in case of the reinforced grades. %lthough these changes are rather complex& the resulting microstructures can be explained by the viscoelasticity of the melt and by the melt flow fields evolved in the mold. The viscoelastic behavior of the melt depends on several parameters of the polymeric material (molecular weight and its distribution& main chain flexibility&conformation possibilities of the chain& etc) and the processing conditions (melt and mold temperatures& plunger speed) that affect the orientation and relaxation of the polymer. "or the flow field consisting of shear and elongational flows& processing conditions are not the only important factors' the mold construction (sprue& runner& gate& and cavity geometry inducing converging and diverging flow during processing) is also relevant. (t is widely accepted that fiber orientation in discontinuous fiber)reinforced thermoplastics can ade*uately be described by the model of Tadmor +11,& which involves the fountain or volcano effect discussed by #ose +12,. %ccording to this model& the fiber orientation pattern produced by in-ection molding can be approximated by a three)layer laminate structure. This is depicted schematically and as it loo.s in practice in "igure 2. (n the surface ( ) layers& fibers are oriented parallel to the mold filling direction (/"0). This is caused by the shear flow of the melt along the *uic.ly solidified layer at the mold wall. (n the central (C) layer& fibers adopt an orientation perpendicular to the /"0 in the plane of the molded pla*ue. This .ind of alignment is due to the elongational flow at the midplane of the cavity. "actors contributing to this elongational flow are diverging flow at the cavity entrance and the fountain effect described by #ose +12,. %n additional argument for the transverse fiber orientation in the C layer was found in the s*ueeze flow of the melt during the pac.ing stage. (n the literature& examples of a more complex layering of the particulate reinforcement can be found& as reviewed +11)12,. 3uite often a random fiber orientation can be produced in the solidified layer at the mold wall. (n the subsurface layer& however& fibers are aligned in the /"0 is a result of the shear flow evolved in this region. The splitting of the layers in this way yields a 4five)ply4 laminate structure. The fiber layering can be even more complicated& since particulate fillers tend to migrate toward the midplane of the molding& where flow speeds are higher +15,. This change& attributed to normal stress effects& again modifies the flow profile and thus the layering and orientation of the discontinuous reinforcement. FlGURE 2: "iber oriention resulting from in-ection molding (a) for 26 wt7 (819.2 vol 7) long :" reinforced polypropylene ($$)' (b) schematically. This picture illustrates position and designation of the compact tension (CT) specimens preferentially used. ;ote that the designation of the CT specimens considers the loading ) notching (longitudinal& ! or transverse& T) directions in respect to the /"0.

#esults of numerous investigations carried out on in-ection)molded pla*ues of 1)2 mm thic.ness indicate that (cf. "igure 1) 1. <oth the fiber layering and alignment increase with fiber volume fraction (=f). 2. The absolute values of the fiber orientation are closely matched in the and C layers& and fiber orientation increases with =f. 1. The processing effects (melt& mold temperature& and in-ection speed) are of secondary importance compared with =f. These findings are for parts of normal thic.ness (1)2 mm) molded by a film gate +12,. "or thinner or thic.er items& which in addition involve other gate constructions& these statements are not always valid. FlGURE 3: >ffects of cavity thic.ness (<) and fiber volume fraction (=f) on the layering& planar orientation (fp) and mean fiber length of in-ection molded discontinuous fiber) reinforced composites /icrostructural investigations carried out on long glass fiber (!:") reinforced in-ection) molded thermoplastics showed significant analogies with short glass fiber ( :") composites +1?)1@,. (t was reported thatA 1. The relative thic.ness of the C layer increases with increasing aspect ratio. 2. "iber bunching and bundling may occur. 1. "iber bending can be evidenced. "iber bunching is connected with the pultrusion pelletizing process used for the production of !:")reinforced in-ection moldable composites. The local ordering of fibers during this process may result in bundles that move cooperatively and do not filamentize enough during molding. "iber bending& on the other hand& is an appearance of decreased structural stiffness due to the higher length. <oth of these effects reduce the effective aspect ratio& of the reinforcing fibers in the molded part +1?&1B,. The aspect ratio of the fibers in the molded item depends on material factors (especially = f)& mold geometry (e.g. runner and gate construction)& and processing parameters (e.g. in-ection speed). Cigher fiber loading shifts the aspect ratio distribution curve toward lower values as a result of increased fiberDfiber and fiberDwall interactions& which cause fracture. This effect is much less pronounced for !:" than for :" reinforced composites& provided that mold construction for the former system is ade*uate. (t is due to the preliminary orientation of the fibers during manufacturing. The aspect ratio distribution curve of the reinforcement may differ when various layers across the thic.ness of the molded part are considered +1?,. This is mainly due to fiber enrichment in the C layer& differences between bunching and filamentization in the and C layers& and effects of the flow field on fibers with different aspect ratios. "or the flow features and microstructural development in discontinuous fiber)reinforced thermoplastics& detailed information can be ta.en from the #ef. 19. 2

(t can be concluded that the microstructural parameters of reinforced in-ection)molded composites are fiber layering& fiber orientation (the two latter are commonly termed fiber structuring)& fiber volume fraction& and effective fiber aspect ratio and its distribution "Design" of Microstructure %mong the guidelines for processing of "#T$s and !"#T$s& priority is given to processing parameters and mold constructions that contribute to preserving the initial aspect ratio& that is& the fiber length of the reinforcement. %voiding fiber brea.age re*uires molding at minimal frictional heating. En a given reciprocating in-ection molding machine this can be achieved by slow screw rotation& low in-ection speed& low bac. pressure& and high barrel temperature. $rocessing of !:" reinforced thermoplastics is very similar to that of :" composites. (t is recommended& however& that a 16)26FC higher barrel temperature and a special 4low wor.4 screw be chosen. This screw is characterized by a long feed section with constant root and wide& deep flights. This section is followed by a low gradual compression zone without .neading or mixing elements' the screw ends in a constant)root metering section with flat flights. (n addition& certain aspects of mold construction have to be considered (short runners& large film or fan gates). ervice conditions for "#T$ composite parts often re*uire a given well)defined fiber structuring. "or in-ection)molded items& a new techni*ue called multiple live)feed in-ection molding was developed. (n this method& a pac.ing head is inserted between the mold and the head of the in-ection)molding machine. The melt flow& and thus fiber orientation in the pac.ing stage& can be modified accordingly by a programmable movement of the pistons of the pac.ing head that pressurizes the solidifying melt directly +26,. Computer aided design (C%0) is a new techni*ue that has been successfully applied to optimization of mold construction for molded parts. (n C%0 design of an in-ection)molded part& the first step is to visualize the wea. sites& that is& the .nit lines (supposing a runner and gate system). The next step is to change the position andDor type of runner and gate so that .nit lines do not evolve or& if this is impossible& are positioned where low stresses in the part during service can be predicted. The next phase is modelling the flow in the mold& subdivided into finite elements& and characterizing the melt flow patterns in these mold segments. "or the calculation of the flow patterns& rheological parameters& determined experimentally& are used. The flow modelling is repeated in several steps until optimized mold filling occurs. The aim during extrusion die design is to get the same material flow in all segments of the die resulting in smooth surfaced& warpage)free extrudates. "or the flow simulations different software pac.ages are available. Microstructural C aracteri!ation

%s stated before& the microstructural parameters are fiber layering and orientation& fiber volume fraction& and fiber aspect ratio and its distribution. "or the determination of fiber layering by imaging of polished sections or thin slices& light (reflective or transmission)& scanning electron microscopy ( >/)& and contact microradiography are preferred. "or fiber orientation& microwave& G)ray diffraction& sonic& and thermographic measurements can also be used. (n "igure 2a >/ micrographs ta.en from polished sections along the thic.ness)/"0 (z)x) and y)x planes are shown. The evaluation of fiber alignment and mean fiber orientation in a given plane is very time) consuming& as it involves determining the angle distribution under which fibers are aligned. (n this respect& image analysis offers the new possibility of getting information about not only in) plane but also spatial orientations +21,. "iber orientation can be described either by using mean orientation factors& such as Cermans +22,& Hrenchel +21,& and modified Cermans +22,& or by vectors +25,. The aspect ratio (since the diameter of the fibers is mainly constant& it can be replaced by fiber length) distribution curves are generally determined from microphotographs of the fibers ta.en after burning away the matrix. (n many cases the matrix polymer can also be removed by solvents. (nstead of histograms showing the relative fre*uency of fibers in a given length interwall& the use of envelope curves& either in differential or in integral form& is preferred. The above)mentioned microstructural parameters are 4integrated4 in a reinforcing effectiveness term (#). This term previously considered the effects of fiber structuring with respect to the loading direction and the fiber loading +2?,. This was extended later to include the aspect ratio and aspect ratio distribution +12&1?&1B,& and generalized in the formA
R = Trel & i f p & eff & iV f & i (
i

(l D d ) n & i l ) equ & i d (l D d ) m & i

(1)

where Trel,i is the relative thickness of the ith layer normalized to the sample thickness (B, see Figure 31.4), fpeff,i is the effective orientation in the i th layer calculated using the function of planar orientation (fp) vs. fp,eff introduced by Friedrich [26], =f&i is the fiber volume fraction in the ith layer, (l/d)equ,i is the equivalent aspect ratio in the ith layer, (l/d)m,i and (l/d)n,i are the mean mass- and number average aspect ratios in the ith layer, respectively [16]. It seems that the fracture mechanical response of discontinuous fiber-reinforced plastics can be appropriately related to this reinforcing effectiveness parameter [14]. Fracture Mechanics Detailed treatment of fracture mechanics is far beyond the scope of this article; it can be found in the literature [2-9]. Here only a brief overview is given. A fundamental aspect of fracture mechanics is that the onset of fracture depends not only on the applied stress but also on the size of intrinsic flaws that act as stress concentrators. The ?

presence of such flaws is the reason, for example, that the real tensile strength of solids, including polymers, is 1/10 of the theoretical value [4]. Such stress concentration sites are always present in neat and reinforced molded plastics, either as a result of processing ("notches" at the knit lines resulting from compressed air, bare fiber segments as a result of imperfect wetting by the matrix, voids caused by differences in the thermal expansion characteristics of the matrix and reinforcement etc.) or caused by use (scratches, damage by cutting or shaping, impacts etc.). The common effects of stress and flaw size are combined in linear elastic fracture mechanics (LEFM) which deals only with bodies that obey the Hookian law, that is, whose deformation is fully elastic in a term called stress intensity factor (K) or fracture toughness:
( K c =) K I = Y a1 D 2

(2)

where I 8 applied stress a 8 crac. size Y 8 geometrical correction factor ta.ing into consideration the finite size of the specimen used ( 8 tensile opening mode& mode ( According to the LEFM theory, fracture occurs when K I > KIc, where KIc, is a critical value of the stress intensity factor. This material parameter is also termed fracture toughness. From Eq. (2) it is obvious that KI is a stress-related fracture mechanical criterion. K I, can be determined from monotonic static loading measurements, for example according to ASTM E 399. In the case of monotonic dynamic loading, KIc can be computed by Eq. (2) as the slope of the plot Y against a-1/2. The other LEFM material parameter is an energy-related one that measures the energy required to extend the crack over a new surface unit. This term is denoted G Ic, and is called fracture energy, critical strain energy release rate, or specific crack extension force. The onset of fracture depends again on whether GI is less than or greater than GIc. Kc and Gc are interrelated by a function whose exact form depends on the stress state of the specimen (plane stress or plane strain). For S(L)FRTPs, G Ic is generally determined from high speed impact tests - using Charpy and Izod test set ups - through the method of Plati and Williams [27] or its derivatives (ISO 17281). The main criterion of LEFM, namely fully elastic deformation, is very severe for plastics that may undergo pronounced plastic deformation (yielding or tearing) during fracture. In this case, other approaches, also used originally for metals, were pursued for plastics: J-integral, crack opening displacement (COD), and essential work of fracture (EWF) [6,28]. These material parameters are included in plastic, elastoplastic, or postyield fracture mechanics (PYFM). JIc is an energy-related term connected with the onset of stable crack growth. For @

linear elastic bodies GIc = JIc. JIc can be determined for example by ASTM E813 and ISO/CD 28660. In addition, JIc can be deduced from high speed impact tests carried out on sharply notched Izod or Charpy specimens [28]. The fracture mechanical approach can also be applied for both static and dynamic (cycling) fatigue of cracked specimens. Static and dynamic fatigue means slow crack growth under subcritical stresses and stress amplitudes, respectively; that is, the stress intensity factor and its amplitude lie below KIc. The aim of both measurements is to establish crack extension characteristics with respect to the stress concentration at the crack tip as a function of either time (da/dt static fatigue loading) or number of fatigue cycles (da/dN dynamic fatigue). The latter characterization can be performed by the ASTM E 647 standard, originally developed for metals. For the determination of the critical values of fracture mechanical parameters related to the plane strain condition, the specimens used have to meet different size criteria. These can be taken from the corresponding standards; alterations to these standards for discontinuous fiberreinforced thermoplastics are summarized in Ref. 14. If these criteria are not met, critical values of KI, are denoted Kc, instead of KIc; the same designation is used also for Gc, and Jc. Fracture and Related Failure In both static and dynamic fracture measurements, breakdown is caused by monotonically increasing load. The only difference between them is related to the strain rate or frequency range; however, the threshold value is rather arbitrary. Measurements carried out below a cross-head speed v of 1 m/min are referred to as static, whereas impact measurements with a striker speed above 1 m/s are referred to as dynamic fracture tests. tatic !oading EFFECT OF MICROSTRUCTURE. Fiber loading. Fiber reinforcement may affect fracture toughness in different ways. It can be improved, worsened, or held at a constant level by fiber incorporation, depending on the matrix of the composite [14]. The run of K c, as a function of Vf can hardly be predicted, because of competitive micromechanisms that either increase or decrease Kc. Nevertheless, discontinuous fiber reinforcement is always the right tool to increase the fracture toughness of low molecular weight polymers prone for brittle fracture. The effects of reinforcement-matrix bond quality and of matrix toughening are worth mentioning here. Improving the coupling between fiber and matrix is not necessarily beneficial for Kc. Strong bonding may hinder the deformability of the composite so that K c tends to decrease [29]. This is in accordance with the Hahn-Rosenfield equation [30]. This equation explicitly shows that Kc does not depend solely on strength but also depends on ductility parameters: 1D 2 Kc = ( E B B L) (1) where B

E 8 > modulus b& Jb 8 tensile strength and elongation at brea.& respectively L 8 proportionality constant including strain hardening& in length dimension It is well known that toughening of the matrix results in increased ductility; however, this is at the cost of stiffness and strength. Thus matrix toughening may also be connected with deterioration in Kc. The course of Jc as a function of Vf depends on the corresponding Kc - Vf and E - Vf functions: K2 (2) J c ( = Gc ) = C E provided the plane stress condition satisfies the LEFM theory. If the K c increment due to the square function overcompensates for the increment in E modulus, J c increases; when it does not, the opposite tendency becomes evident. It should be noted here that J lc values can be found scarcely for discontinuous fiber-reinforced thermoplastics in the literature [14,28]. In spite of the very complex fracture mechanical response to fiber loading, the following conclusions can be drawn: 1. An increase in Kc from fiber loading is more probable the higher the E modulus and the lower the ductility of the unfilled matrix (brittle, low molecular weight, degraded polymers, especially polycondenzates). 2. For ductile materials a relative increase in both K c and Jc can be achieved by using fibers of higher aspect ratio (e.g. LGF instead of SGF). 3. The effects of matrix toughness and fiber-matrix bonding are hardly predictable. For relative improvements in Kc and Jc, the strength and ductility characteristics of the matrix have to be balanced by the reinforcement. Fiber structuring. The layering and orientation of the fibers in injection-molded items were already shown in connection with Figures 2 and 3. On the fracture surface of the specimens, fibers lying parallel or longitudinal to the crack plane (L fibers) can clearly be distinguished from those oriented perpendicular or transverse to it (T fibers) (Figure 4). FIGURE "A "racture surface at the razor notch of 26 wt 7 (819.2 vol 7) :" reinforced in-ection)molded $$. ((n this T)! type CT specimen& ! fibers can be found on the surface& whereas T fibers in the central layer& as indicated' cf. "ig. 2. #azor blade notch is mar.ed by arrow.) It is doubtless true that the anisotropic structuring of the fibers yields different fracture mechanical values when specimens with various notch directions (T and L; see Fig. 2b) are tested [14,17,26,29]. The load bearing capacity of T fibers aligned in the load direction is considerably higher than that of the L fibers, which have practically no reinforcing effect.

Therefore, the fracture mechanical response depends on the relative thickness of the layers containing T and L fibers, respectively. The degree of fiber orientation in these layers is also important. T fibers completely aligned in the load direction guarantee the best stress transfer and thus the greatest reinforcement. Fiber misalignment along the load direction necessarily reduces the overall reinforcing effect. Friedrich introduced an effective fiber orientation term that takes this fact into account cf. Figure 5 [26]. FIGURE #: #elationships between the effective (fp&eff) and planar fiber orientations (fp) considering the actual mechanical loading direction Many investigations carried out on SGF and LGF composites (e.g., Refs. 14,16,17,18) have indicated that the anisotropy in the mechanical response of the LGF-reinforced systems is not very pronounced, in spite of the fact that the three-ply laminate structure caused by the injection molding still exists. This observation suggests the important role of the aspect ratio. Fiber aspect ratio. The influence of filler shape on fracture toughness at a given filler loading strongly depends on the matrix characteristics [29]. However, with increasing fiber aspect ratio Kc always increases, at least above a given threshold l/d. This is connected with an increase in the loadability of the discontinuous-fiber-reinforced composites, since their strength increases with increasing aspect ratio [31]. The l/d ratio can be increased either by using longer fibers of the same diameter or by using smaller diameter fibers of the same length. Reinforcing with small diameter fibers is beneficial only in a given aspect ratio range. This is due to the fact that the deleterious effect of stress concentration at the fiber ends should be compensated for by toughness-enhancing effects, such as increasing interface, improved stress interaction between fibers, and enhanced crack path [32]. A decrease in the critical fiber length for smaller diameter fibers is expected to lead to an increase in the clack path length, which promotes fiber pullout and reduces fiber fracture [14]. A definite answer on the effect of fiber aspect ratio distribution cannot be given. There are, however, several indications [26,32] that use of reinforcing fibers with different aspect ratios as a result of varying diameters can be beneficial for fracture mechanical characteristics.

EFFECTS OF TESTING CONDITIONS. Temperature. The fracture toughness, measured at low cross-head speed, decreases as a function of temperature T for both matrix and its fiberreinforced versions [16,29]. A steeper decrease in the related plot can always be found in the vicinity of the glass transition temperature (Tg) of the matrix. Here the enhanced molecular mobility of the matrix induces a change in deformation mode from ductile to viscous. So, T g is always the upper threshold for the applicability of the LEFM theory. K c values calculated 16

according to Eq. (1) for temperatures above T g no longer have meaning for fracture toughness; they can be treated only as trend data of a mechanical property thus defined. Cross-head speed. At v = 1000 mm/min, the trend of K c with T is basically different from that at v = 1 mm/min. Kc values start at a relatively low level in the subambient temperature range before they increase at the Tg [29, 33, 34]. This is attributed to a clear change in the stress state of the samples (plane strain to plane stress) and related failure manner (brittle to viscous as a result of adiabatic heating at the crack tip).

FAILURE BEHAVIOR. It is obvious that big differences due to testing conditions are related to substantial changes in microscopic breakdown events. The microscopic failure mechanisms occurring in S(L)FRTPs are shown schematically in Figure 6. They can be grouped into matrix-related (crazing and shear yielding) and fiber-related (fiber fracture, pullout, and debonding) events. For longer-fiber-reinforced injection-molded composites, the latter can be extended by fiber bridging (unbroken fibers connect the crack sides) and by cleavage and slip of fibers within bundles or rovings during debonding and pullout. It should be noted here that the relative orientation of the fibers (L and T) strongly influences the relative probability of the individual fiber-related energy absorption mechanisms [14,29,35]. FIGURE $: "ailure mechanisms for discontinuous)fiber)reinforced thermoplastic at a microscopic level Based on fractographic results, the characteristic failure modes of the composites can be summarized in failure maps. In such maps the dominant matrix- and fiber-related breakdown processes are indicated as a function of the testing conditions T and v [18,35,36]. Failure maps not only give guidance for selecting composites for particular circumstances but also suggest methods of increasing the toughness. In the patent 1iterature one can find abundant examples and ideas for such improvement.

Dynamic Loading EFFECTS OF MICROSTRUCTURE. Fiber loading. The plot of dynamic fracture toughness as a function of Vf can be as complex as that for the static case. Generally the dynamic fracture energy (Gd) decreases with increasing fiber loading. This effect may or may not be compensated for by the increasing E modulus with respect to the resulting K d. Since this effect of the E modulus is very closely matched to the static one, the plot of K d as a function of Vf depends mostly on the Gd - Vf function of Eq. (4). The effect of V f on Kd is demonstrated in Figure 7 for SGF- and LGF-reinforced composites. It can be concluded that 11

increasing the aspect ratio of the reinforcement yields an improvement in K d. The Gd - Vf function, on the other hand, should depend on the matrix toughness and fiber-matrix bond quality. FIGURE %: Change in the static (Hc) and dynamic (Hd) fracture toughness as a function of fiber loading (=f) at ambient temperature for :") and !:")reinforced in-ection)molded $$ (Hc determined on CT specimens at v 8 1 mmDmin& whereas H d deduced from lzod measurements) At this point attention has to be paid to the internal flaws induced by specimen machining (sawing, cutting, polishing, and the like). The presence of such flaws causes trouble during curve fitting and a very big scatter in G d, especially for long fiber-reinforced systems. However, if failure of the specimens occurs not at the notch introduced but near to it means that cutting introduced a flaw (either by debonding or by intrabundle fiber cleavage) that acted as a stress concentrator site. In some cases this threshold or latent notch value was found to be = 0.5-0.6 mm [37]. Fiber aspect ratio. This effect is surprisingly pronounced if one compares K c and Kd for SGFand LGF-filled composites (see Fig. 7). This suggests some differences in the load transfer during static and dynamic measurements that should be clarified. EFFECTS OF TESTING CONDITIONS. Gd and Kd values derived from Izod and Charpy measurements are very closely matched. This is due to an analogous stress state during impact, which is carried out at practically the same deformation rate. The plot of K d and Gd as a function of T can be calculated from Eq. (4) provided that the E(T) and G d(T) or Kd(T) functions are known at the given frequency. Kd generally increases with decreasing T as a result of the increase in E modulus, whereas G d remains practically constant. The plots of both fracture mechanical parameters depend on the temperature range investigated; maxima and minima in the plots can also be found. Their appearance is attributed to primary and secondary relaxation transitions of the matrix and its components [4,5,38]. FAlLURE BEHAVIOR. It has to be emphasized that dynamic failure mechanisms are the same as those shown and discussed with respect to static loading. Although failure mapping has not been performed for dynamic measurements, the following findings are expected: 1. The frequency embrittlement of the matrix promotes brittle matrix cracking, the onset of which depends on the frequency-dependent T g. Among the matrix-related failure mechanisms, crazing is more common than shear yielding.

12

2. Among the fiber-related failure events, fiber pullout and fracture tend to dominate. Their relative proportions depend not only on the testing conditions but also on the fiber-matrix bonding. Fatigue and Related Failure S(L)FRTP parts are widely used in fields in which constant and cyclic subcritical loading occur. The response to long-term constant loading (static fatigue) is often called either stress corrosion (SCC) or environmental stress corrosion cracking (ESC). In this case, crack propagation is induced in different environments at the condition K 0 < KIc, where K0 denotes the initial stress intensity factor. The result of this kind of measurement is either the stress corrosion threshold (KI,SCC) or the crack growth rate (da/dt vs. KI), or both. KI,SCC means a threshold KI, value below which no crack growth takes place in a given environment. In dynamic or cyclic fatigue, the crack growth rate per cycle is established as a function of the stress intensity factor amplitude (K). A threshold value K th, which is connected with the onset of fatigue growth, can be read from the fatigue crack propagation (FCP) curve. The characteristics of static and dynamic fatigue with respect to the measurements and results to be discussed are summarized in Figure 8. FIGURE 8: The fatigue behavior of injection-molded composites. Static Fatigue EFFECT OF MICROSTRUCTURE. When Kc, from static fracture, increases with V f, one can expect a similar trend in the resistance to SCC for the given composite. Incorporation of a rubbery impact modifier in the matrix results in further improvement in SCC resistance of the related composite. Fractographic analysis supports the conclusion that this is due to better wetting of the fibers and thus better protection of them against acidic attack. On the other hand, when Kc decreases with Vf, an acceleration in SCC growth can be predicted in relation to the corresponding matrix. The da/dt-KI curves, or at least their segments in double logarithmic representation, can be approximated by straight lines. This indicates the validity of the Paris-Erdogan relationship (often called the Paris power law [3,4,39]):
da = K Im dt

(5)

The crack growth kinetics depend on both the microstructure and the environment [29,40,41]. Note that the final breakdown of the specimens (usually CT type) does not necessarily occur at the Kc derived from static loading measurement. In SCC tests, KI at the specimen breakdown may be higher than K c when a material with rather ductile behavior is investigated. This is a consequence of the very low frequency of this kind of measurement, which generates a process or damage zone that is fundamentally 11

different from that found in static fracture. Alterations can be observed in both the size of the zone and the related failure mechanisms therein. It can be concluded that final breakdown of a specimen with ductile features (either due to the material or due to the stress state) takes place near the static Kmax value. This value can be determined by Eq. (1) using the maximum load. Such response can be observed for composites either in air or in surrounding media that plastify their matrix. Final failure of the specimen may also occur at K I < Kc as a result of aggressive attack by a given environment. If the initial stress intensity factor K 0 is very small but higher than KI,SCC, the surrounding medium can penetrate deeply into the specimen, causing failure events (e.g. multiple fiber breakage, surface degradation) that decrease the SCC resistance (diffusion-assisted SCC). Therefore, it is highly reasonable to always indicate K0. For composites with a ductile matrix or one that is "ductilized" during the measurement, the crack tip at its advance can hardly be resolved. Instead of a sharp crack, a well-evolved damage zone can be observed. The propagation of this zone as a whole can be treated and adequately described by the crack layer theory (e.g. [42]). Changes in the time to failure curves due to microstructural and environmental effects are very similar to those discussed above with respect to SCC growth [40,41]. However, final failure may occur at a K c which is either independent or dependent of K 0. The latter case suggests that failure depends additionally on K0, that is, on the corrosion loading history of the specimen. This corrosion loading history is rather complex, since it involves the immersion time and all changes in both the structure and the stress state that are caused by the diffusion and penetration of the environment. This observation means that Eq. (5) no longer holds, since da/dt depends in addition on K 0; therefore, talking about a material parameter according to the fracture mechanical concept is very questionable. EFFECT OF ENVIRONMENT. Both the da/dt-KI, and the K0-time (t) curves depend strongly on the environment. They can be grouped by whether their degradative attack relates mostly to the matrix, to the fibers, or to the fiber-matrix interface. In addition, the SCC response is highly affected by the corrosion loading history. FAILURE. First impressions about the failure mode can be got from the surface appearance of the broken specimens. A nearly planar fracture surface indicates fiber degradation, which mostly occurs in acidic environments [43,44]. A zig-zag fracture surface path, reflecting the fiber alignment, on the other hand, illustrates matrix and matrix-fiber interface attacks. The failure micromechanisms during SCC usually agree with those in static fracture; however, their relative occurrence varies considerably with the aggressive nature of the environment and the K0 of the test. Dynamic Fatigue 12

Since flaws are always present in S(L)FRTPs, one can conclude that the main constituent of durability is the propagation of such flaws rather than their initiation and development. In this case, the results of FCP measurements are relevant and important from the point of view of construction. EFFECT OF MICROSTRUCTURE. Fiber loading. It should be emphasized that FCP resistance due to fiber loading changes along with the K c-Vf function. Increasing fracture toughness thus correlates with improved FCP resistance (Fig. 9), while decreasing K c with Vf yields FCP acceleration [14,29]. Figure 9 again shows the validity of the Paris power law (Eq. 6) for the stable FCP range:
da = ( K ) m d!

(?)

FIGURE &A Changes in "C$ behavior due to microstructural parameters& schematically The exponential term m of Eq. (6) generally increases with V f (Figure 10). This change is accompanied by a shift of K th, toward higher K values; that is, fiber incorporation enhances the threshold limit below which no crack growth takes place. The same trend in K th can be observed when results achieved on L- and T-cracked specimens are compared. This is due to a change in the stress state of the specimen as it approaches the plane strain condition as a result of increasing fiber loading and load direction aligned fiber structuring. The onset of unstable crack growth (final fast fracture) occurs near either K c or Kmax, just as it does in static fatigue. This upper limit and the stable FCP range itself are strongly affected by the viscoelasticity of the material under the given testing condition (e.g., crack tip heating effects [45]). FIGURE 1'A $aris range in the "C$ curves of $$ and its :") and !:")reinforced grades at different fiber volume fractions (indicated in vol.7) Fiber structuring. L-T specimens, because of the higher quantity of T fibers oriented in the load direction (see Fig. 2b), exhibit higher FCP resistance than T-L-cracked ones. This finding is again analogous to the results of static loadings. The effect of fiber structuring becomes more and more pronounced with increasing Vf for SGF than for LGF reinforcement. This supports the statement made before in connection with monotonic loading, namely, that incorporation of LGF diminishes the mechanical anisotropy. Fiber aspect ratio. When the FCP curves of SGF- and LGF-reinforced PPs are compared a clear improvement in FCP resistance can be seen as a result of the use of fibers of higher 15

aspect ratio (Figure 9). The relative improvement diminishes with increasing V f. FCP enhancement of LGF reinforced systems was attributed to the evolution of a more extended damage zone and longer debonding and pullout routes [46]. The effects of the above microstructural parameters on the FCP behavior are summarized schematically in Figure 9. For the microstructural interpretation of the FCP response, a qualitative [47] and a quantitative model [16,18,29] exist. The latter is based on the reinforcing effectiveness (R) and microstructural efficiency (M) concepts [16,26]. Figure 11 shows the difference in the FCP rates between SGF- and LGF-reinforced PP composites as a function of the M values. One can clearly see that with increasing M the crack rate, at a given K value, decreases markedly [18].

FIGURE 11: "C$ rates of :") and !:")reinforced $$ grades as a function of the microstructural efficiency (/) at a given stress intensity factor amplitude ( H81.5 /$am1D2)

EFFECTS OF TESTING CONDITIONS. Information on the effects of external testing conditions (waveform, frequency [47], main load, temperature, environment, etc.) on the FCP response of SFRTPs is very limited. These effects seem to be highly material-dependent; therefore, general conclusions cannot be drawn. FAILURE. Analogies between the fracture mechanica1 responses given for fracture and for fatigue suggest that the individual failure mechanisms at the microscopic level are the same (see Fig. 6). The first step in failure is again debonding at the fiber ends. This is followed by pullout and further debonding for T- and L-oriented fibers, respectively. The fatigue crack path in Figure 12 demonstrates this. In this picture, stress concentration at the fiber ends is clearly perceptible. In the damage zone, in addition, a stress concentration field was developed and preserved by matrix deformation. The matrix underwent crazing and shear yielding, initiated by fiber debonding and pullout processes. FIGURE 12: "atigue crac. profile on the surface of a CT specimen of 26 wt 7 (8 B.1 vol 7) :" reinforced $$A ( ) behind the crac. tip' (B) before the crac. tip. (Crac. direction from left to right.) Among the discrepancies between fracture and fatigue, the size of the damage zone and the sharpness of the crack have to be mentioned; the zone is smaller and the crack is sharper during fatigue [33]. Further differences between the failure modes are connected with crack tip heating and accompanying changes in both matrix- and fiber-related micromechanisms caused by cyclic loading. At the onset of stable FCP, the matrix fails in a semibrittle manner with multiple 1?

fractures. Therefore, the mean pullout length here is relatively small. In addition, this multiple matrix fracture yields a stress field that promotes fiber breakage. At the end of stable FCP, the ductility of the matrix increases considerably; crazed arrays and viscous torn matrix parts can also be evident. Among the fiber-related mechanisms, pullout dominates-however; with an increased average length [48,49]. On the FCP curves of many S(L)FRTPs, a stable delayed crack growth region can also be resolved just before stable acceleration crack growth occurs (see Fig. 13a). This is connected with an evolution and stabilization of the damage zone. This is the right place to call the attention to differences in cyclic and static fatigue responses. Under static conditions the stable crack deceleration occurs at much higher actual stress intensity factor compared to the cyclic one (see Fig. 13b). This is due to the formation of a more extended equilibrium damage zone the formation of which was supported by the fact that the apparent frequency of static fatigue is lower than the cyclic one [18,48]. Under apparent frequency the reciprocal value of the time causing the specimen fracture is meant. FIGURE 13" table crac. deceleration ranges registered during cyclic (a) and static fatigue (b) for :") and !:")reinforced $$ grades with different fiber volume contents The fiber-related micromechanisms strongly depend on the relative fiber angle to the crack plane [29,35,50]. In addition, crack closure may also appear, especially in T-fiber regions with higher fatigue crack lengths. This decelerates the FCP rate and is connected with a mixed mode (modes I and II) stress state. It would be highly reasonable to construct fatigue failure maps analogous to those for fracture. Because experimental results are lacking, however, mapping cannot be performed yet. Design Aspects It has been shown above that the fracture mechanical characterization of injection-molded thermoplastics and their composites contributes to a better understanding of their performance in different loading situations. Although in many cases real plane strain fracture parameters can hardly be deduced because the mean thickness of injection-molded parts (3-6 mm) does not reach the required one, their measured values can be used for design and construction. For this purpose, however, an adequate fracture mechanical characterization method should be selected. The proper choice of method when a fracture-resistant part is to be constructed depends on the behavior of the material. For composites with high stiffness and low ductility, which fail by brittle fracture, stress-related terms (Kc) are used, whereas for those of high ductility, energyrelated fracture mechanical terms (Gc , Jc, COD and EWF) should be preferred. In the case of a part that is to withstand fatigue loading, the first question to be answered is whether failure occurs mainly by crack initiation or by crack propagation. When the cycles to 1@

failure at different stresses are plotted for a given material, the response curve can be divided into two regions: crack initiation and crack propagation (Fig. 14). It was shown above that for SGF and LGF reinforced injection-molded composites, crack propagation is of basic importance. In this case, therefore, threshold values (KI,SSC, Kth) derived from static and dynamic fatigue measurements can be used for construction purposes.

FIGURE 1": 0esign factors for fatigue)resistant composites depending on whether crac. initiation or crac. propagation dominates the failure behavior.

(ummar) This article has dealt with microstructural development& microstructure)related fracture& and fatigue performance at static and dynamic loadings and corresponding failure behavior of short and long discontinuous fiber)reinforced in-ection)molded composites. The response of these systems to different loading conditions was treated by fracture mechanical concepts using stress) and energy)related terms. The accompanying failure was analyzed by fractography and grouped into matrix)and fiber)related events. (n addition& the changes related to different loading situations were demonstrated and discussed. %ttempts were made to determine general trends in microstructural development& fracture mechanical response& and connected failure behavior& and to summarize them schematically. %nalogies and discrepancies between fracture and fatigue were emphasized and discussed. "inally& recommendations were given for design of in-ection)molded thermoplastic matrix composite parts under increasing (fracture) and alternating (fatigue) load. The literature cited offers further detail on this topic.

+ ee also /olding& $olymer (n-ection KKK, L. Harger)Hocsis References 1. !. %. Mtrac.i& Int# $%l&m# $r%ce''#( 2& 1 (19B@). 2. C. C. Hausch& $%l&mer )racture( pringer& <erlin& 19@B. 1. #. N. Certzberg and L. # /anson& )ati*ue %f En*ineerin* $la'tic'( %cademic $ress& ;ew Oor.& 19B6. 2. %. L. Hinloch and #. L. Ooung& )racture Be+a,i%r %f $%l&mer'& %pplied cience $ubl.& <ar.ing& Mnited Hingdom& 19B1. 5. L. :. Nilliams& )racture -ec+anic' %f $%l&mer'& >llis CorwoodDL. Niley& Chichester& 19B2. ?. %. :. %t.ins and O.)N. /ai& Ela'tic and $la'tic )racture" -etal'( $%l&mer'( Ceramic'( C%mp%'ite'( Bi%l%*ical -aterial'( >llis CorwoodDL. Niley& Chichester' 19B5. 1B

@. ;. :. /cCrum& C. $. <uc.ley& and C. M. <uc.nall& $rinciple' %f $%l&mer En*ineerin*& Exford Mniversity $ress& Exford& 19BB. B. H. "riedrich& >d.& pplicati%n of )racture -ec+anic' t% C%mp%'ite -aterial'( >lsevier& %msterdam& 19B9. 9. !. %. Carlsson& >d.& T+erm%pla'tic C%mp%'ite -aterial'( >lsevier& %msterdam& 1991. 16. L. Harger)Hocsis( PReinforced polymer blends in D. R. Paul and C. B. Bucknall, Eds., Polymer Blends, Volume 2: Performance, J. Wiley, N.Y., 2000, p. 395. 11. Q# Tadmor& L. ppl# $%l&m# .ci#( 1B& 1@51 (19@2). 12. N. #ose& !ature( 191& 222 (19?1). 11. /. L. "ol.es& .+%rt )ibre Reinf%rced T+erm%pla'tic'( #esearch tudies $ress& Chichester& 19B2. 12. L. Harger)Hocsis& PMicrostructure and fracture mechanical performance of short fibre reinforced thermoplastics in #ef. B& p. 1B9. 15. #. $. Cegler& :. /ennig& and C. chmauch& d,# $%l,m# Tec+n%l#( @( 1 (19B@). 1?. L. Harger)Hocsis and H. "riedrich& C%mp%'# .ci# Tec+n%l#( 12( 291 (19BB). 1@. 0. >. pahr& H. "riedrich& L. /. chultz& and #. . <ailey& J# -ater# .ci#( 25& 222@ (1996). 1B. L. Harger)Hocsis& Microstructural aspects of fracture in polypropylene and in its filled, chopped fiber and fiber mat reinforced composites in J. Karger-Kocsis, Ed., Polypropylene - Structure, Blends and Composites, Chapter 4, Vol.3, Chapman and Hall, London, 1995, p. 142. 19. 0. N. Clegg and %. # Collyer& >ds.& -ec+anical $r%pertie' %f Reinf%rced T+em%pla'tic'( >lsevier %pplied cience. !ondon& 19B?& especially Chapters 5 (L. !. Nhite)& ? (!. %. :oettler)& and 9 (#. C. <urton and /. L. "ol.es). 26. :. Halay and /. L. <evis& P%pplication of shear)controlled orientation in in-ection molding of isotactic polypropyleneR in J. Karger-Kocsis, Ed., Polypropylene: An A-Z Reference, Kluwer Academic, Dordrecht, 1999, p. 38. 21. :. "ischer and $. >yerer& $%l&m# C%mp%'#( 9& 29@ (19BB). 22. $. C. Cermans& C%ntributi%n t% t+e $+&'ic' %f Cellul%'e )ibers& >lsevier& ;ew Oor.& 192?. 21. C. Hrenchel& )ibre Reinf%rcement( %.ademis. "orlag& Copenhagen& 19?2. 22. #. <. $ipes& #. !. /cCullough& and 0. :. Taggart& $%l&m# C%mp%'#( 1& 12 (19B2). 25. . :. %dvani and C. !. Tuc.er& J# R+e%l#( 11& @51 (19B@). 2?. H. "riedrich& C%mp%'# .ci# Tec+n%l#( 22& 21 (19B5). 2@. >. $lati and L. :. Niliams& $%l&m# En*# .ci( 15& 2@6 (19@5). 2B. N. :rellmann and . eidler (>ds.)& $%l&mer Te'tin*& Canser& /unich& 266@ 29. H. "riedrich and L. Harger)Hocsis& 4"racture and fatigue of unfilled and reinforced polyamides and polyesters&4 in L. /. chultz and . "a.irov& >ds.& .%lid .tate Be+a,i%r %f Linear $%l&e'ter' and $%l&amide'( $rentice)Call& >nglewood Cliffs& ;L& 19B9& p. 229. 16. :. T. Cahn and # #. #osenfield& .%urce' %f )racture T%u*+ne''" T+e Relati%n bet/een KIc and t+e 0rdinar& Ten'ile $r%pertie' %f -etal'( % T/ T$ 123( %merican ociety for Testing and /aterials& $hiladelphia& 19?B& p. 5. 11. 0. Cull& n Intr%ducti%n t% C%mp%'ite -aterial'( Cambridge Mniversity $ress& Cambridge& 19B1. 12. ;. ato& T. Hurauchi& . ato& and 0. Hamigaito& J# C%mp%'# -ater( 33( 456 (19BB). 11. H. "riedrich& !. %. Carlsson& L. N. :illespie& and L. Harger)Hocsis& PFracture of thermoplastic compositesR in #ef. 9& p. 211. 12. L. Harger)Hocsis and H. "riedrich& $la't# Rubb# $r%ce''# ppl#( @& 78 (19B@).

19

15. H. "riedrich and L. Harger)Hocsis& 4"ractography and failure mechanisms of unfilled and short fiber reinforced semicrystalline thermoplastics&4 in %. C. #oulin)/oloney& >d.& )ract%*rap+& and )ailure -ec+ani'm' %f $%l&mer' and C%mp%'ite'( >lsevier %pplied cience& !ondon& 19B9& p. 21@. 1?. L. Harger)Hocsis and H. "riedrich& L. -ater# .ci#( 22( 92@ (19B@). 1@. L. Harger)Hocsis& J# ppl# $%l&m# .ci#& 25& 1595 (1992). 1B. L. Harger)Hocsis and =. ;. Huleznev& $%l&mer( 21& ?99 (19B2). 19. $. C. $aris and ". >rdogan& J# Ba'ic En*#( B5& 52B (19?1). 26. H. "riedrich& L. -ater# .ci#( 2?& 1292 (19B1). 21. C. =oss& # 0olgopols.y& and H. "riedrich& $la't# Rubb# $r%ce''# ppl#( B& @9 (19B@). 22. %. Chudnovs.y& 4Crac. !ayer Theory&4 ;% % Contractor #eport 1@2?12& Case Nestern #eserve Mniversity& Cleveland& 19B2. 21. C. !hymn and L. /. chultz& $%l&m# En*# .ci#( 22& 16?2 (19B2). 22. L. ;. $rice& 4 tress corrosion crac.ing in glass reinforced composites&4 in %. C. #oulin)/oloney& >d.& )ract%*rap+& and )ailure -ec+ani'm' %f $%l&mer' and C%mp%'ite'( >lsevier %pplied cience& !ondon& 19B9& p. 295. 25. #. N. !ang and L. %. /anson& J# -ater# .ci#( 22& 15@? (19B@). 2?. L. Harger)Hocsis and H. "riedrich& C%mp%'ite'( 19& 165 (19BB). 2@. #. N. !ang& L. %. /anson& and #. N. Certzberg& P"atigue crac. propagation in short) glass)fiber)reinforced ;ylon ?.?A >ffect of fre*uencyR in L. C. eferis and !. ;icolais& >ds.& T+e R%le %f t+e $%l&meric -atri9 in t+e $r%ce''in* and .tructural $r%pertie' %f C%mp%'ite -aterial'( $lenum& ;ew Oor.& 19B1& p. 1@@. 2B. L. Harger)Hocsis& H. "riedrich& and R. S. Bailey, Adv. Composite Mater., 1, 103 (1991). 29. L. Harger)Hocsis& #. Nalter& and H. "riedrich& L. $%l&m# En*# B( 221 (19BB). 56. #. N. !ang& L. # /anson& and #. N. Certzberg& J# -ater# .ci#( 22& 2615 (19B@).

26

Você também pode gostar