You are on page 1of 19
Materials Transactions, JIM, Vol. 32, No. 1 (1991), pp. 110 19 OVERVIEW Strengthening Mechanisms of Metal Matrix Composites* Minoru Taya** Strengthening mechanisms of metal matrix composites (MCs) are discussed from the viewpoint of both macro and nano-level models. The macro-level models are law of mixtutes and shea lag models, which are reasonably adequate for explaining the strengthening of aligned fiber (continuous and short) MMCs. While the na-level models are based fon dislocations (punching and back stress) and are found to explain well the strengthening observed in particulate MMCs. I both macro-and nano-level models, the residual stresses induced due vo the CTE mismatch sein play anim portant role in influencing the yield stress of the composites. Keywords: metal matrix composites, yield stress, lw stress, work-hardening rate, aw of mixtures model ‘shear lag model, elastic peg model, back sires, punching of dislocations, CTE mismatch strain, Eshelby's method 1. Introduction ‘Metal matrix composites (MMCs) have been developed with the aim of using them in more demanding en- Vironments, which require higher specific mechanical properties and durability at high temperatures. Figure 1 illustrates the introduction years of various new high tem- perature material systems with engine applications in mind”. In this figure, the lower temperature MMCs (aluminum and magnesium alloy-based composites) are competitive with the high temperature polymer matrix composites (fiber reinforced plastics, in short FRP), while the high temperature MMCs are competitive with superalloys. Among the required mechanical properties, the yield stress (6,) and work-hardening rate (A) are con- sidered as important as the other mechanical properties: stiffness (E) and fracture toughness (K). For the higher the yield stress and work-hardening rate of a MMC are, the greater the capability of the MMC in sustaining a higher applied load, thus increasing the design flexibility, of a component made of such a MMC. The yield stress and work-hardening rate represent the second stage of the stress-strain curve of a MMC, see Fig. 2 where the first, second, and third stages of an aligned continuous fiber MMC are shown, as well as those of the unrein- forced metal and fiber. The existence of the third stage depends on the type of the matrix metal and fiber used; for example, a tungsten fiber/copper matrix composite possesses usually up to the third stage, while in a ceramic fiber/ aluminum matrix composite, only up to the second, * This overview is invited by the Mater. Trans., JIM Editorial Commitee Department of Materials Processing, Faculty of Engineer Tohoku Univesity, Sendai 980, Japan stage is likely to be observed. The mechanical behavior of MMCS both at room temperature and in use en- vironments has recently been summarized by Taya and Arsenault®, ‘The yield stress ¢, and work-hardening rate h of a com- 2 & & & A (te 3 uaaCoupgaitoe fee ciate LK “ —S -_ Invoductin to Engine Use Fig. 1 Use temperturesof materia systems vt introdution seas for engine applications”. ‘THREE STAGES OF THE ‘STRESS-STRAIN CURVE viele stress composite Vila strain of mate Strain stage I + both matic and tber remain elastic stage IL: matrix detorms plastcaly and tier remains alate TI + botn mate and fiber detorm pla sta aly Fig. 2 Strese-strain curves of a MMC, the uneeinforced matrix ‘metal and reinforcing fiber”. posite are characteristics with metal matrix composites, while in the other composite systems (ceramic matrix composite (CMC), and polymer matrix composite (PMC) or fiber reinforced plastics (FRP)) they are less ob- vious (in FRP) or do not exist in normal use conditions (CMC). Thus a number of studies of the yield stress and work-hardening have focused on the MMC system with the aim of modeling the strengthening mechanisms responsible for the increase in , and / over the unrein- forced matrix metal. The previous models for the strengthening mechanisms can be grouped into the following: (1) law of mixtures model, (2) shear lag model and (3) dislocation models. The order of the above models is arranged from simplest to more refined, which coincides with that of development of these models, from earliest to most re- cent, Dislocation models can be subgrouped to: (i) Orowan model, (fi) forest hardening model, (iii) elastic peg model and (iv) punching model. In this paper, the semi-details of the above models will be given in Section II, followed by concluding remarks in Section IIL II, Strengthening Models Among the three models for explaining strengthening, mechanisms of metal matrix composites, the first two, law of mixtures and shear lag models, were originally developed for prediction of composite stiffness, while the Minot Taya last ones, Orowan, forest hardening, elastic peg and punching models, were developed originally to analyze the dispersion hardening mechanism of two-phase metals. The dimension of the filler in the former models is that of fiber diameter of the order 1-10 jum, whereas i the latter models that is the diameter of dispersed par- ticles or precipitates, thus the order of 10-100 nm, Despite the difference in the filler size in the above models, we shall treat them as a composite strengthening model below. 1. Law of mixtures model Consider a composite which consists of fillets and matrix, and is subjected to the applied stress oo, Fig. 3 Let the stress, stiffness and volume of the i-th phase be denoted by ai, Ei, and Vj respectively where i= (com- posite), m (matrix) and f (filler). Then, the volume in- tegral of stress in the whole composite can be expressed as [eoe| etefaw where the subscript to integral sign { denotes the domain of the volume integral and dv is the volume element. If 0, Gn and @; are taken as the volume averaged values, then eq. (1) is reduced to Via.=Vnont Vios @ Dividing eq. (2) by V. and introducing the volume frac- tion of each phas EM, enn 0 0 14 Fig. 3A composite model: filles are embedded inthe matrix ‘subjected to applied sre Strengthening Mechanisms of Metal Matrix Composites a= Val Ve oh Ve cater @ we obtain ind Cra co) Equation (4) is a law of mixtures in terms of flow stress, which includes yield stress. The stress in a composite in eq. (1) can be decomposed into uniform (applied) stress om and disturbed stress as due to the existence of filles. ‘Thus, the volume integral on the left side of eq. (1) can be reduced to J e.av=| (rte. do oy Since the volume integral of the disturbed stress within the whole composite domain vanishes", the right side of eq. (5) is equal to Vaco, resulting in a=. © Hence, the applied stress can be considered “composite stress”, The accuracy of the composite stress a; given by eq. (4) depends on that of calculating om and a. The ac- curate evaluation of aq and o; involves three-dimentional stress analysis such as Eshelby’s method™®. If the geometry of the filler is simple, for example, a aligned continuous fiber composite which is axially loaded by oa, Fig. 4, then one can assume an iso-strain condition in that both matrix and fiber are subjected to the same strain, e. The averaged stress in each phase can then be approximated as a to 4 Fig. 4 t o4 ‘A continuous fiber composite model. 8 o A substitution of eq. (7) into (4) provides the composite stiffness E, that is expressed by the law of mixtures: Ex=tnE nt CE ®) ‘The first stage of the stress-strain curve in Fig. 2 was ac- tually obtained from eq. (8). While the second and third stages were constructed using eq. (4) and the stress-strain curves of the matrix and fiber. The above procedure of calculating the composite flow stress may be adquate for the case of a aligned continuous fiber composite, as evidenced in Fig. 5. In Fig. 5, the experimental data of the tensile yield stress of tungsten fiber/copper matrix composite are shown as filled circle and the prediction based on the law of mixtures model, eq. (4) by solid line. However, this is not the case with a short fiber or par- ticulate composite, Fig. 3. where the assumption of iso- strain is not valid, and the state of stress within each place is not necessarily uniform. The input in the law of mixtures formula, eq. (4) ate the volume fractions of the matrix cq and filler ¢; and the average stress in each phase. Thus, other data related to the microstructure of the composite such as shape and its distribution in space are not used as input in the formula. Hence for a given set of constituent properties and c, (i=m and f), eq. (4) predicts the same flow stress regardless of the filler being 2000 1500 1000 exp 500 iment aiction 10 20 30 40 50 60 70 80 Fiber Volume Fraction, V6) Fig. 5 Tensile yield stress vs fiber volume fraction for continuous tungsten fiber/copper matrix composite: experiment (filed ctces) and the prediction based on the law of mixtures (line) 4 Minoru Taya a continuous fiber or particulate. This is indeed incor- rect. The flow stress and stiffness predicted by the law of mixtures model give rise to the upper bound for the cor- rect values®", This defect with the law of mixtures model can be corrected if one can use more refined models; shear lag and Eshelby's models, which require additional input such as the fiber aspect ratio. The former model will be discussed in the next subsection and the latter will be used in conjunction with elastic peg and dislocation punching models in the later subsections. 2. Shear lag model The shear lag model was developed first by Cox", and later used extensively by a number of researchers, for ex- ample”, and itis best applied to an aligned short fiber composite loaded along the fiber axis, Fig. 6°. Unlike the law of mixtures model where only the volume average (in macroscopic sense) was accounted for, in the shear lag model the load transfer at the matrix-fiber interface is focused on. The shear lag model provides a tool to calculate the average stress o and oy in the law of mix- tures model by focusing on the local stress field in and around the interface. In the original Cox model the load transfer at the interface was considered only for the side surface of the fiber, over the length /, see Fig. 6(b), while in the model modified later by Nardone and Prewo", the load transfer at the fiber-end over the length d, Fig. 6(b) wwas also accounted for. We will state in the following the calculation procedure for the average stresses in the matrix o» and in the fiber o; based on the shear lag model. matrix EEE 6 ._4__| trtt $=. wo Fig. 6 Shear lag mode! (a) an aligned short fiber composite subjected applied Consider a representative short fiber of circular cross section with length / and diameter d which is located in a unit cell of similar circular cross section with length L and diameter D, Fig. 6. This unit cell is assumed to be loaded by the applied strain ¢ along the fiber axis, z-axis. ‘The local coordinates attached to this fiber is z (along the fiber axis) and r (along the radial direction) with the origin at the center of the short fiber. First imagine that the unit cell without a short fiber was subjected to the ap- plied strain e which gives rise to uniform displacement » in the unit cell (or the whole composite domain), e=dv/dz, then a representative short fiber is added to the unit cell, leading to non-uniform displacement field. Let the displacements in the fiber and in the matrix on the boundary of the unit cell be denoted by u and », respec- tively, Fig. 6(c). It is natural to assume that the difference in these displacements, «—v is proportional to the shear stress at the matrix-fiber-interface, tp. The equilibrium of force in the z-direction on the volume element with length dz, Fig. 6(¢), leads to do, _4t aod where g is a unknown constant, which will be determined later by appropriate boundary conditions. In the fiber which is assumed to deform elastically, one-dimensional Hooke’s law holds: u-0) 0 du an a 0) <« > <* « [sed] u (6) nit cll use forthe shear las analysis and (e stress and displacement fed in euilibsium’® Strengthening Mechanisms of Metal Matrix Composites s In the above equations, a; is the axial stress in the fiber and assumed to be uniform along the r-direction, but a function of z. In the following, we first treat the case where both the matrix and fiber deform elastically to calculate the composite stiffness, E., since this is the original Cox model, then will discuss the case of the plastically deforming matrix to predict the composite yield stress, oe ‘Taking the derivative of eq. (9) with respect to z, which is then combined with eq. (10) leads to fo; ar qepaee ay The general solution to eq. (11) is given by Eye, cosh fe+C: sinh z «a where p=\e 03) Unknown integration constants C; and C; will be deter- mined from the boundary conditions: 6/00 (constant) at z=1/2 and da;/dz=0 at 2=0, thus, the fiber stress 0 is obtained explicitly as ay ‘Then, the average stress in the fiber, &; can be calculated as Hew 2-1) em @ : w(r=D/2)=0, see Fig. 6(c). The force equilibrium along the z-direction at r=r and r=d/2 leads to 2) =Ee | 1+ as) a9) sroir( 2) where + is the shear stress at r, which is related to y by a7 Ga and where Gp is the shear modulus of the matrix, and eq. (16) was used to drive the last term in eq. (17). Integrating eq. (17) with respect to r from r=d/2 (the matrix-fiber interface) to D/2 (the boundary of the unit cal) leads to fa (2 26m, ma vmw cas) From eqs. (1) and (18), g can be solved as 8G 9 ~ d?In(D/d) Hence, f is explicitly obtained as 2d2 [GalEr (GalEr ea in (D/d) In the above equations, the diameter ratio of the unit cell to the fiber D/d must be estimated properly. Morimoto et al. investigated experimentally the relation between the aspect ratio of the unit cell L/D and that of short fiber //d: L ( 1 ) =k where k is a proportionality factor and its value ranges 0.7-1.0 for 10-20% vol. SiC whisker/aluminum com- posite. Kelly and Street™ used a crude approximation to relate D/d to known input such as the volume fraction of fiber ¢,. Under this crude assumption; L=I and short fibers being arrayed in a hexagonal way, they estimated Didas alee) The average normal stress in the matrix, dm can be calculated as a en 2) Fn=Ene 3) Then, the composite normal stress g. can be estimated by using the law of the mixtures formula, eq. (4) where oi and 9; are replaced by dq and 6 respectively. a= Cnt 016% ey [A substitution of eqs. (15) and (23) into (24 leads to the formula to predict the Young's modulus of the com- posite, E. E 5) 1mopeetee, {1-2} BIR It is clear from eqs. (8) (the law of mixtures model) and (25) (the shear lag model) that in the shear lag model, the ‘composite stiffness is a function of the fiber aspect ratio Id, while it is not in the law of mixture model. Nardone and Prewo™ extended the above shear lag model to predict the yield stress of a MMC where the matrix is assumed to deform plastically and the fiber re- mains elastic. By using eq. (9), we integrate it with respect to z from z=//2 to z=z, to obtain z ormevt ans (5-=) 2d de e9 where o» is the normal stress in the matrix adjacent to fiber-ends. The average fiber stress d; can be calculated similar to eq. (15). ol d= 004 a @n 6 Minoru Taya When the matrix yields, one can assume (28) where Gay is the tensile yield stress of the matrix, and the average stress in the matrix Gq is assumed to be equal to Guy. Upon substitution of d-and dy so calculated into the law of mixtures formula, eq. (4), the composite yield stress a is now obtained as 9) w {Leavitayer} The distribution of the fiber stress g¢ along the z-direc- tion based on the elastic analysis (Cox model) and the plastic analysis (Nardone and Prewo model) are shown schematically by the solid and dashed lines, respectively in Fig. 7. It is noted in Fig. 7 that the stress at fiber-ends based on the elastic analysis, a0 may vary depending on the bonding condition and the nature of the stress concen- tration there, but it is approximated here as g., the ap- plied stress of the composite. It is also noted in the figure that for larger applied stress a., the fiber stress at the center of the fiber, , (E:/E,) can exceed the breaking stress of the fiber, op. Then the solid curve should be bounded by ais if the fiber does not fracture, However, for the larger applied stress, the matrix yielding is more likely to take place, thus, to the first order approxima- tion, one can set in eq. (26). ores) 30) From eq. (30), one can calculate the critical fiber length, J. at and above which the fiber can break: 1 oe where 7909/2 from eq. (28), or in the case of weaker interfacial strength, the value of z be set equal to the weaker interfacial strength. ‘The shear lag model has also been applied successfully to the case of creep problems of MCs", since it is not a en em(tof) Ee, Fig. 7 The distribution of fiber stress: elastic analysis (sli ine) nd plastic analysis (dashed line)”. restricted to linear problems. The shear lag model, however, tends to give rise to poor estimate for the mechanical properties of a short fiber composite with small fiber aspect ratio. Because in the case of small fiber aspect ratios such as particulate composite, the load transfer in terms of shear stress at the matrix-fiber inter- face becomes limited due to the shorter interfacial length and the stress field in the matrix being three-dimensional, departure from the assumption used in the shear lag model. In this case, models which can analyze such a three-dimensional stress field in a particulate composite, should be used. Eshelby’s model is best suited for the analysis of a particulate composite, which will be dis- cussed in the next subsection in conjunction with elastic peg and dislocation punching models. 3. Dislocation models The dislocation models were originally constructed to predict the second stage of the stress-strain curve of dispersion-hardened alloys!" the yield stress. and work-hardening rate h. It seems that the existing models, for strengthening of dispersion hardened alloys can be grouped into: (i) Orowan model", (ii) forest harden- ing model’, (ii) elastic peg model" and (iv) punch- ing model""'®-"", We will state these models below. () Orowan model The yield stress in shear of a dispersion-hardened alloy normally set equal (o the Orowan stress", Fig. 8, Gb R where G and bis the shear modulus and the magnitude of Burgers vector of the matrix metal, respectively and R is the radius of curvature to which a dislocation is bent and a constant a ranges from 0.5 to 1.0. Ashby'"" calculated +, based on a more rigorous mode! to account for the effect of additional microstructural data associated with single crystal dispersion-hardened alloys; number of par- ticles intersecting unit area of slip plane N., the type of dislocation (edge and screw), and particle diameter d. Ashby’s formula for r, is given by % wy @2y ieee for edge disk a ee Fig glg tors titoaion t Ses te Fig. 8 Orowan modet™, Strengthening Mechanisms of Metal Matrix Composites, 7 where G4) ris the inner-cut off radius, usually taken as 2b, and vis Poisson’s ratio of the matrix metal. The above formulas. suggest that x, increases with a decrease in the average spacing between particles, and an increase in particle size, The effect of the reduction in particle size on ty however, must be carefully assessed in eqs. (33) and (34): for a constant volume fraction of particles, the alloy with the smaller sized particles gives rise toa higher value of s, than that with the larger sized particles, since in the former alloy, D becomes smaller and A’ decreases logarithmically, resulting in the inerease in z,.It is noted here that the increase in particle volume fraction ¢; results in the reduction in D in eq (33), thus the increase in s,. The above formulas were obtained for a single crystal loaded in shear. One can convert the above results. to those for a polyerystalline metal loaded in tension 0 calculate the tensile yield stress ¢, which is roughly est mated as as) 6, where a constant & ranges from 2 (single crystal) to 3.1 (Taylor factor for polycrystalline metals). @) Forest hardening model Ebeling and Ashby"”” showed that most of the ex: perimental data at intermediate (a few percent) to large plastic strains obey the following shear stress/shear strain relation: [bow ren 406 G6) where t and y are the applied shear stress and plastic, strain, respectively, G and b are the shear modulus and Burgers vector of the matrix metal ¢; and d are the volume fraction and mean diameter of particles, respec- tively and 7 is a dimensionless constant equal to 0.240.04. Work-hardening stress tt, is proportional to Vy i.e., departure from the linear work-hardening, predicted by the elastic peg model, Ashby"? attempted to explain this square root dependance on shear strain by proposing the relaxation mechanism of the large misfit strain at the matrix-particle interface, which increases with increasing plastic shear strain. The relaxation mechanism considered by Ashby is the punching of dislocation loops along the secondary slip planes, which upon interaction with those dislocation lines on the primary slip plane contributes to the hardening of a dispersion-hardened alloy. Ashby considered a hard particle embedded in a matrix metal subjected to tensile stress T which results in a homogeneous shear strain y in the x-direction, Fig. 9a). Then an initially spherical particle deforms to an ellipsoid with the longer axis along, the loading direction. If the hard particle is removed from the ellipsoidal hole from the matrix, while keeping the tensile stress T, then the shape of the ellipsoidal hole would become more elongated, shown as solid line in Fig. 9(b) along the loading direction, the x/-axis. In order to bring back to the ellipsoidal shape in a loaded disper- sion-hardened alloy before the removal of the particle, fone must apply a set of displacements, uj (along the x/- axis) and 1 (along the x-axis), Fig. 9(b). This set of displacement field is equivalent to punching out m disloca- tion loops with Burgers vector & along four directions, T SL x 6) cy} cnTensTIAL” xy Loors vacancy" t50°8 ° Fig, 9. Ashby model: (a hard particle embedded in soft matrix sub- jected to tensile stress 7, (b) displacements u, 1 applied to elipsoid {bring back tots orginal spherical shape, and (c) dislocation loops needed to produce the configuration of (bj 8 Minoru Taya Fig. 9(¢). The displacements ui and wi at the matrix- particle interface take the following maximum values: (Wn (WS) en Thus, along each direction, the following equation should hold: vd 4 nl (8) The total number of loops per unit volume of a single crystal, Nr, is the total number per particle 4n, times the number of particles per unit volume N,=6¢;/ zd?, and is written as @9) ‘The number of loops intersecting a primary slip plane, 'N,, can be estimated by multiplying Ne by the probability of a secondary loop intersecting the primary slip plane, times the average diameter of loops, Hence 3) (x). fe GG) () ‘The interaction of a straight gliding dislocation with a prismatic loop has been studied analytically by Kroupa”. Kroupa showed that with many different orientations of loops, the average force exerted by one loop, opposing the glide of a straight dislocation along the primary slip plane is approximated as 0.25Gb". Then, the total force opposing the motion of unit length of dislocation due to all loops is the number of loops per unit length of distoca- tion, times 0.25Gb" is set equal to the stress increment above the initial yield stress, (r—t,) times b, resulting in N (40) an where z; is the initial yield stress in shear and can be given by eq. (33). The work-hardening stress r~ z, predicted by eq. (41) agreed well with the experimental data of single crystal Cu-SiO; strained to three different levels: y=5%, 10% and 15%, Fig. 10, where the prediction is shown by the solid line. (3) Elastic peg model As (0 the work hardening of dispersion alloys, one ‘must first refer to Fisher et al.‘ who considered disloca- tion loops left around each particle, Fig. 11 and used a magnetic analog (0 evaluate the stress state, Using this result which was originally designed for application to the initial yielding of dispersion alloys'”, Ashby has shown that the work hardenig stress tis written as 3el!*NGb @ (42 where Nis the number of dislocation loops with Burgers Theoretical slope approx. 0.24 Experimental slope 0.26:0.08 Dey o0n03 Be © e001 008s ° 108 (ee 10° Vbayia Fig. 10. The prediction based on the Ashby mode! (sold line) com: pared with the experimental results of single erystal Cu-SiO, strained to various shear strains, Fig. 11 The elastic pep model of Fisher ea. vector b around each particle, and a is the radius of the pare, N ean be tld tote plats sear tes shown by Ashby’. Referring to Fig. 12, one obtains b where dis the diameter ofthe partie. With cg (2), 6. (econ be sped .=6e!!’Gy. 3) a) ‘Strengthening Mechanisms of Metal Matrix Composites 9 = shosr Strain Shear Strain —+— Fig. 12. The single slip of the Fisher et al. model (bean be modified 10 ‘multiple slips @)". Equation (44) is easier to use than that eq. (42) since the former includes only macroscopic parameters, cy and y, while in the latter, the number of dislocation loops N must also be estimated. It is clear from eq. (44) that the work-hardening rate h=dn)/dy is a function of only the volume fraction of particle, ci, but not the details of the dislocation arrangement. Tanaka and Mori” derived based on a continuum model the linear dependence of ft ‘on ¢ in contrast with the above model where the dependence of ft on cy is not linear, but c}!?, Namely, in the Ashby's analysis, ¢/=(d/da)? was used where dy is the mean spacing beetween particles, resulting in eq. (44). While they pointed out that c; should be equal to (d/dz)", resulting in the linear dependence of x, on c,. Ashby col- lected the data of the work-hardening stress of various dispersion-hardening alloys (mostly single crystals) to conclude that the slope of log A-log ¢r plot ranges be- tween I and 1.3. This suggests that the linear dependence of t, on ¢; obtained by the Tanaka-Mori model appears to be more reasonable than the Fisher er al. model. The concentric dislocation loops proposed by Fisher et al. were not observed in dispersion hardened alloys where instead tangled dislocation loops have been observed. The forest dislocation model proposed by Ashby cannot explain the Bauschinger effect observed in dispersion-hardened alloys” and MMCs"”. The Bauschinger effect can be quantified in terms of the difference between the flow stress in tension (a) and that in compression (a}), Aci=o}—!, as schematically shown in Fig. 13°". Ag; was observed to increase with plastic strain e, linearly up to some strain. The relation- ship between Aa; and ep cannot be explained without the hardening due to “back stress", elastic particles resisting plastically deforming matrix metal With the above in mind, Tanaka and Mori‘ proposed ‘an elastic peg model based on contimum mechanics. In the Tanaka and Mori model the dislocation loops to represent the misfit strain between the clastic particle and plastically deforming matrix are smeared out to Fig. 13. Bauschinger effect of a MMC in terms ofthe diference in the flow stresses in tensile nd compressive loadings! become “transformation strain’” defined inside the particle. Thus the problem becomes tractable by use of Eshelby's equivalent inclusion method™", In Eshelby's method, fillers are represented by ellipsoidal _in- homogeneity, thus the fillers other than sphere can easily be treated: disk, short and long fibers. In the following, we shall state the Tanaka and Mori model to predict the yield stress o, and work-hardening rate h of MMCs, Consider a MMC which consists of ellipsoidal fillers (Q) with stiffness tensor Cy and metal matrix (D—2) with stiffness tensor Cy and uniform plastic strain ef in the matrix, and is subjected to the applied stress 0, Fig. 14a), where a and e} are given by a}=©, 0, a0, 0, 0, 0, 0) (1/2, =1/2, 1, 0, 0, OVep as) Here 1 denotes transpose, and the order of stress and strain components in eq. (45) is: 11, 22, 33, 23, 31, 12. ‘The domain of the MMC is denoted by D. In Fig. 14(a), af denote prescribed transformation (cigenstrain) such as the thermal expansion misfit strain, but in the origi- nal Tanaka-Mori model, af=0. The reason for assum- ing the shape of filler as ellipsoid is to facilitate the computation using Eshelby’s method. We shall derive the formula to predict the stress-plastic strain relation by using the energy balance equation: su=-80 where U is the total potential energy of the composite and 6U is the change in U due to the change in plastic strain de}. 4Q is the energy dissipation due to the plas- tic work in the matrix given by 6Q where of is the yield stress of the matrix, and is assumed to be constant. The change in the total potential energy due to de’ is given by"; 46) 1ena$de, «“ Fig. 14 The elastic pee model based on continuum mechanics! ‘93 embeded inthe mat ©. -| oirtonses av ou: (48) where a is the stress disturbance due to the existence of fillers which deform only elastically, giving the misfit strain inside the fillers, which can be defined only in the fillers. As far as the stress disturbance is concerned, the problem of Fig. 14(a) is equivalent to that of Fig. 14(b),, where the uniform plastic strain in the matrix e{) was removed from the matrix, and ~e! is now given in the do- main of the fillers. From eqs. (45)-(48), one can arrive at 9) $4 Gag (oon) where o) and oy are the stress disturbance that is solved from Fig. 14(b) and (o;)~o4:) can be expressed as a linear function of plastic strain e,. Then, solving for ay in eq. (49), we obtain the stress (ao)-plastic strain (e,) rela- tion of a MMC given by Go=0,+ hey (50) where o, and /h are the yield stress and work-hardening rate of the MMC. /his related to the tangent modulus, Er, Minoru Taya > eP) clos MUCK Bi) ) elastic fillers with stifiess Cy) and CTE mismatch strain with stifines Cand uniform plastic strain e () can be converted 0 equivalent elastic problem Fig, 1S. The stress-strain curves of model MMC and unreinforced ‘matrix metal, Strengthening Mechanisms of Metal Matrix Composites = = ~Prsicton 200} Sts, ofMPa Tens Tensile Stain (%) Fig. 16 The stees-strain curves of particulate MMCs: the predictions ‘by the continuum elastic peg mode (dashed lines) and the experiment (solid lines). the slope of the second stage of the stress (a) total strain (6) relation, Fig. 15, by Er = E Both a, and h are functions of the stiffness constants of the matrix and filler, the shape and volume fraction of the filler (or fiber). Tanaka and Mori‘ compared the prediction based on their continuum elastic peg model with the experimental data of the stress-strain curves of single crystals of copper containing BeO, Al;O; and SiO; particles, Fig. 16. As far as the initial part of the stress- strain curves (i.e. at small plastic strains) are concerned, the prediction agreed well with the experiment. They also, compared the prediction with the experimental data of tungsten fiber/copper matrix (W/Cu) composites with various volume fractions of fibers; c.=0.043-0.38, Fig. 17, where the predictions by the Tanaka-Mori model and the experimental data are shown by dash-dot and dashed lines, respectively. Also shown as solid lines in Fig. 17 is the prediction by the Tanaka-Mori model modified for finite volume fractions of fiber, It is clear from Fig. 17 that the original Tanaka-Mori model is valid only for small volume fractions of fibers whereas the modified Tanaka-Mori model can predict the yield stress and work-hardening rate of the composite well for wide ranges of cr. The effect of filler shape on the work-harden- ing rate was examined analytically by using the con- tinuum elastic peg model"? and some of the results are shown in Fig. 18 where the hardening-rate normal- ized by matrix Young’s modulus, /t/Eq, is plotted as a function of the stiffness ratio of the matrix to filler E;/Em. Figure 18 indicates that the disc-shaped filler loaded along the direction parallel to the dise plane gives h on 600} 509} 8 Tensile Stress, af MPa 8 200} 100 fs0083 0 GTO “Tensile Strain (96) a3 04 05 Fig. 17 The stresstrain curves of W/Cu composite: experiment ‘(Gashed lines), the prediction based on the orginal Tanaka Mori ‘model (dash-dot lines) and the Tanaka-Mori model modified for finite filer volume fractions (slid lines)": Work Hardening Rote, 10° hiE a er a ) ‘Young's Moduli Ratio, &/E Fig. 18 The hardening rate & normalized by the matrix Youns's ‘modulus £, as function of the Young's moduli ratio of the mattix, {0 filler for various shapes of fille! rise to the most effective hardening, followed by the nee- die (fiber)-shaped filler loaded along the needle-axis, the disc-shaped filler loaded along the disc-axis. Spherical filler turns out to be the least effective. The effect of fiber n 2 Minore Taya 2000) | 1nd=80 jg 1800- 2 ‘ig=10 i 5 1000) 3 500] ~ oe om 12 Tonsile Strain (%) Fig. 19. The effect of fiber aspect ratio / don the stress-strain curve of ‘short SiC fiber/Al composite predicted by the continuum clastic pe model™, aspect ratio (I/d) on the yield stress and work-hardening rate of a short fiber MMC was also examined by using the continuum elastic peg model and the results are il- lustrated for the case of SiC fiber /aluminum alloy matrix composites in terms of the stress-strain curves in Fig. 19%, It follows from Fig. 19 that the larger fiber aspect ratio has a significant effect in increasing not only stifiness (Young’s modulus), but also yield stress and work-hardening rate of the composite. The above continuum elastic peg model assumed an in- itially stress-free state in a composite. This is not true with most of composites which must be processed by ex- periencing temperature change and /or mechanical work ‘The temperature change is inherent in processing: from the processing temperature to the temperature at testing (mostly room temperature), i.e. the change being negative, AT'<0. The coefficient of thermal expansion of the matrix (aq) is different from that of filler (a); in a typical MMC system, Aa=ai~ay<0, resulting in a misfit strain defined in fillers upon cooling to room tem- perature, aif=erd, where d, is Kronecker’s delta and er=AaAT>0. This misfit strain induces the internal stress field, what is called as thermal residual stress. In a unidirectional MMC system, this thermal residual stress at room temperature gives rise to tension in the matrix and compression in the fiber, Hence, this thermal residual stress field is expected to reduce the tensile yield stress a} while increasing the compressive yield of the MMC, a5. The prediction of a and a5 of a MMC can be © Exp Theory 00 (of one 50 ° on oz Vo Fig, 20. The effect of thermal residual stress on the difference in the Yield stresses in tension (o}) and compression (@;) a8 a function of SiC whisker volume fraction, V2 made by using the continuum clastic peg model if it is ‘modified to account for thermal residual stress”. This modification involves merely the replacement —e!) by aj —e' in the analytical model of Fig. 14(b), and the ther- ‘mal residual stress so calculated will be substituted in the second term on the right side of eg. (49). Thus, if the ther- ‘mal residual stress is dilatational as in the case of particle MMCs, then the effect of residual stress on a, becomes ze10, for (6—0u) in eq. (49) is reduced to zero. This does not mean that the CTE misfit strain has no influence on gy. Instead it still has an effect on @, due to the in- crease in the dislocation density in the matrix metal, which will be discussed in subsection 11.3.(4). Arsenault and Taya” investigated experimentally and analytically the effect of thermal residual stress of the yield stressed of SiC whisker/6061 Al matrix (SiC/Al) composites. The results are summarized in Fig. 20 where the difference of the yield stresses in compression and tension, a)—a5 is plotted as a function of volume fraction of whisker, Ve. ‘The prediction based on the continuum elastic peg model ‘modified for finite Vand thermal residual stress, and the experiment are shown by solid line and open circle, re- spectively, indicating a good agreement between them. (4) Punching model In modeling for the strengthening of MMCs, one tends to assume that the mechanical properties of the matrix, metal in a MMC are the same as those of an unreinforced metal. This is not a correct assumption as evidenced from Fig. 21 where (a) and (b) denote transmissing electron microcrope(TEM) photographs of the unreinforced 6061 Alminum and the matrix domain of a 20% SiC par- Strengthening Mechanisms of Metal Matsix Composites 1B @ Fig. 21 ticulate/6061 Alminum (SiCp/Al) composite, respec- tively", Figure 21 demonstrates that a much higher den- sity of dislocations exists in the matrix in a composite than in the unreinforced metal The thermal residual stress induced by CTE misfit strain af in a MMC was discussed in the preceding subsection. In these models which can account for CTE misfit strain, the thermal residual stress was calculated by assuming that the CTE misht strain at the matrix-filer interface induces elastic strain in a MMC. This elastic analysis leads to the high stress field in and around fillers, which must be relaxed in real MMCs. The critical stage for the relaxation at the nano-level is the generation of dislocations at a particle-matrix interface’. For a large dispersoid such as an equiaxed filler, short fiber‘ and spherial particulate", a simple approach of dislocation punching can be taken. Arsenaullt and Shi"® studied the punching of disloca- tion loops from equiaxed particulate in @ particulate MMC mainly from a geometrical point of view. Their model is shown in Fig. 22 where the dimension of an equiaxed particulate is given by f, f: and f. The average dislocation density in the matrix predicted by Arsenault and Shi is given by Acier Bie) 7 (52) where er is the CTE misfit strain defined previously, 1 is the smallest value of 1), f: and f; and A is a geometrical factor taking the following values depending on the shape of filler: 4. for one dimension, very small compared with other two f, 8 for platelet of aspect ratio 2 10. for short fiber of aspect ratio 2 12 for equiaxed particulate: 4=1=45 (53) TEM micrographs ofthe matrix meal: () unreinforced 6061 Al and (b) 20% c, SiC/6061 o Once A is calculated, the increase in the flow stress (Acre) due to CTE misfit strain may be estimated by (54) where y is a constant of order of 1, Gand b are the shear modulus and Burgers vector of the matrix metal, respec- tively. Taya and Mori” studied the case of a short fiber MMC with an emphasis on the preferred direction of punching. In the Taya-Mori model, the punching direc tion is assummed to be along the fiber axis, Fig. 23, since the stress concentration at fiber ends is the largest, thus requiring its relaxation by punching. It is noted that the CTE misfit strain before punching in Fig. 23(a) is Sacte=~GbV 5 f | PUNCHED bistoc. Fig. 22. The analytical model used by Arsenault and Shi for punching ‘of primate dsiocation loops to relax the CTE mist atthe matrix particulate interface™ ry Minors Taya Xs fo 3 to) Fig. 23. CTE misfit strain at the matrixshort fiber interface (a) is relaxed by punching of prismatic dislocation loops along the fiber (5) axis)" represented by eigenstrain e74,!" where d, is Kronecker’s delta, whereas the eigenstrains after punching (Fig. 23(b)) are given by e 00 [om s} ea 000 00 0 oo 0 in Q (56) where %, and Q; are the domain of a short fiber and punched-out region including the short fiber, respec- tively. cis one half of the fiber length and e the punching distance. The punching distance can be calculated from the condition SICIAI Composite 4T=-200 k, £=0.001 Fiber Aspect Ratio, lid Fig, 24 The punching distance along the fber axis c* normalized by ‘one half fiber length © a8 a function of fiber aspet ratio U/ a" 67) where U is the total potential energy of a MMC (Fig, 23(b)) and Wis the total dissipation energy, that is plastic work required for the motion of the dislocation loops during punching, The total potential energy of the MMC can be calculated by using the energy related theorems of Eshelby’s egivalent inclusion method®""” and the final form is given in Ref. (17). Wis obtained per unit volume en where f isthe fiber aspect ratio equal to ¢/a, and kis the friction stress of the matrix metal. Taya and Mori exam- ined the effect of several parameters on the punching distance c’. Figure 24 shows the relation between the punching distance e* normalized by c and fiber aspect ratio ¢/a. It can be concluded from Fig, 24 that the punch- ing along the fiber axis direction would become difficult as the fiber aspect ratio inerease, and no punching is pos- sible for the fiber aspect ratio larger than a certain value: in this case such a critical aspect ratio, £. is about 27 What would happen ifthe fiber aspect ratio exceeds f.? In the case of longer fibers with > f., the punching direc- tion is expected to be along the transverse direction". The formulation for a long fiber is basically the same as that for a short fiber, except for the eigenstrains in the punched-out stage are now given by (58) PoO0 00 0} fora, (59) 00 & 0 fo} for 2: (60) 0 Strengthening Mechanisms of Metal Matrix Composites 15 ° Fig. 25 y tt a 44 fo The analytical model for CTE mist strain used by Taya et al for the ease of a spherical particulate MMC {@) before punching, (b) ater punching and (c) with dislocation loops rearranged ina spheccally symmetric manner where @ and a’ are the radius of long fiber and that of the puched-out region including the fiber. The fiber axix is again taken along the x-axis. For 50% c; AlLOs~y fiber / 5086 Al composite with AT=—200 K°", the punch- ing distance a” divided by the fiber radius a is calculated as about 2.0, thus the punched-out region covers the majority of the matrix domain. The case of spherical particulate was recently studied both analytically and experimentally by Taya et al." The analytical model before and after punching is shown in Fig. 25(a) and (b), respectively, where a, a” and a” are the radius of the spherical particulate, the punching distance and the radius of the punched-out domain (of spherical shape). The CTE misfit can be represented as isotropic ‘eigenstrain as in the case of a short fiber. By the punching of dislocation loops, the misfit strain is now smeared out in the punched out domain (93) with the corresponding cigenstrain: 1) Then, we introduce a spherical particulate (2) within each domain @ which will disturb the internal stress field due to inclusion . In an actual composite, the in teraction between particulates and that between the free surface and particulates can be estimated easily if we can convert the three-phase problem, Fig. 26(a), to the two- phase problem, Fig. 26(b) where punched-out domains are smeared-out as uniform matrix domain (D— 9). The detailed procedure for computing punching distance is again based on the Eshelby method and given in Refs. (18)G0)30). The average dislocation density in the matrix p can be estimated as follow: first the total number of prismatic, dislocation loops per particulate, ns, is estimated as mybA=aV 2) where b is the Burgers vector, Ay is the area of the loop and equal to na*, and AV is the volume change due to CTE mismatch and approximately equal to 4xa°er/3. Hence sae: o 3) ‘The total number of prismatic dislocation loops per unit volume, Nr, is obtained as the product of the total number of the loops per particulate (rx) and the number of particulates per unit volume, N,=¢,/ (4a /3). Hence Bercy (64) nab ‘Thus, the average dislocation density in the matrix p is ob- tained as the product of Ny and the length of a loop a” divided by the volume fraction of the matrix; Fis. 26 Three-phase problem associated in the punching mode (a) ‘can be reduced to two-phase problem (b) ifthe punched-out domains are smeared out in the matrix 6 Minoru Taya @ o Fig. 27_The punching model used by Luly fora disk-shaped MMC: {@) before punching (b) after punching (5) If the results of Arsenault and Shi for the case of equiax- ed particulates eqs. (52) and (53) are compared with those of Taya ef al., eq. (65), the prediction of the dislocation density based on the Arsenault and Shi model is more than twice as large as that on the Taya er al. model, since a” >a, It should be noted here that the length of a disloca- tion in the punched-out stage of the Taya er al. model for one-and two-dimensional punchings'"*” and of the Arsenault-Shi model for three-dimensional punching re- mains the same as that before punching, whereas in the Taya et al, model for three-dimensional punching, the length of a dislocation in the punched-out stage a” is larger than that before punching, a. Lulay® recently studied analytically the punching problem for a disk-shaped particulate MMC by assuming that the CTE mismatch strain represented by the primatic loops of Fig, 27(a) is relaxed by the punching of the loops along the disk plane (x, x axis), Fig. 27(b). The cigenstrains after the punching, e*? in Q, and e7* in @:, are given by eqs. (59) and (60) except that (a/a’)’ be replaced by (c/¢’}. ‘The punching distance is determined by a similar condi- tion to eq. (57). U can be calculated in a straightforward manner using the Eshelby method, if the eigenstrains such as those in eqs. (59) and (60) and eq. (61) are given after punching. To justify such simple forms of cigenstrains, Mori has taken the following method! Since punching is physically achieved by the glide motion of dislocations, a pure plastic shear strain e}(@ troduced in the punched region, 0:—2). The analytical form of ef is determined so that the punched-out disloca- tions exist only at the boundary of @. Further, the impo- tent eigenstrain ef* which produces no stress" is in- troduced in ®:. It has been shown that the combination of the preexisting eigenstrain (e,4,), ¢3, and e* leads to the total eigenstrain given in eqs. (59) and (60) and ea 6). Since ef is determined, W can be easily calculated and is expressed in a logarithmic function of the punching distance. W is also proportional to the yield stress of the matrix. Then, using eq. (57), the punching distance is determined. The residual stress in , after punching is calculated, once the punching distance is determined The average density of dislocations which act as Forest dislocations is evaluated, With these factors, one can es- timate an increase in flow stress due to punching. The increase in the flow stress of a MMC predicted by the punching model should be added to that of the unrein- forced matrix metal. Thus, the yield stress of the matrix metal a; that will be used to predict the yield stress of a MMC is the sum of the yield stress of the unreinforced metal of and Agcre given by eq. (54): = 09+ doce (66) 0, given eq. (66) should replace of in eq. (49) where (@s5~ 01) contains not only the internal (back) stress as a result of elastic fillers resisting the plastically deforming, matrix due to loading, but also the thermal residual stress asa result of punched-out dislocations. Taya et al. used the above two-step modeling, punch- ing and elastic peg models to study the strengthening of particulate MMCs. In order to see the effect of CTE misfit strain clearly, they conducted quenching of SiCp/6061 Al composites, which were melt casted, followed by extrusion and solution-treated at tempera- ture 803 K, into several low temperatures: 293, 273, 209 and 83K. Immediately after the quenching, the mechanical tests were conducted to measure the stress- strain curves at those temperatures that MMC specimens were quenched to, from which the yield stresses at 0.2% offset strain were obtained for the unreinforced matrix and MMC specimens. The results of the measured yield stresses (open circles) are plotted as a function of the tem- perature drop by quenching in Fig. 28 where (a) and (b) denote the case of 10% cy and 20% c, composites, respec- tively. Also shown in Fig. 28 are the yield stresses predicted based on the punching and elastic peg models. ‘A good agreement between the prediction and experi- ment is obtained, particularly for 10% c. composite. This is because in the case of 10% c; composite, the punching distance is not t00 large so that the punching-out regions may not overlap with adjacent ones, Fig. 29 (b), which is one of the assumptions used in the Taya ef al. model. On the other hand, in the case of 20% ¢; composite, the overlapping of punched-out regions with the adjacent ‘ones can take place, Fig. 29(a). To see the contribution of two strengthening mechanisms: punching (Agcie) and back stress (Ao), the analytical results of Agcre and Ags are plotted in Fig. 30. It can be concluded from Fig. 30 that the strengthing due to the punched-out dislocations generated by CTE misfit strain is dominant over that due to back stress. This is why the previous models, which did not account for the punching by CTE misfit strain, cannot explain fully the strengthening observed in SiC/ Al composites. In other words, the previous models underestimate the yield stress, as demonstrated in Fig. 31°”. In Fig 31 the dash, dash-dot and solid lines are the Strengthening Mechanisms of Metal Matrix Composites 0 20 © é £35 108 Voline Faction 2 an) g ¢ pane oe © 150] gs i © Eo 3 a. : 10 Vane Fn : secre 2 0 “* B09 -o- o9 é z = &. go e 3 é an so 55000 Gs ~—«70D~~«S Temperature Orop, ATIK 50089 ead aS 00 730400 Temperatore Orop, ATIK Bs 7 é& é€ 28 Veeck wei] g i. —— 5 ao i Q Blk oe $ g wl i io F pe 7 Ey : sve vine i £50 oop é cre i =e qe ; i 8 ol 0 soo sso 600 6s «700 ~«750”~—« “Temperature Drop, ATIK Fig. 28 Composite yield streses of SiC particulate/6061 Al com: posites asa function of temperature drop by quenching: (a) ¢-=10% (©) 6;=20% where filled and open ctcles denote the prediction based fon the Taya e a. model" and the experimental data Fig. 29. The analytical model for punching used by Taya er a. (a) the case of extended punching where some of the punched-out domains are overlapped, and (b) the less extended punching! 500580 ao 80780 Temperature Drop, ATK Fig. 30. Contribution to the yield stress increase by two strengthening mechanisms, the punched-out dislocation due 10 CTE mismatch (Gare) and back stress (80): (8) c/= 10% and () e-=2096"" predictions based on the shear lag type model", the elastic peg model and that accounting for the residual stress”, respectively. The discrepany between the prediction based on the previous composite models and a number of experimental data becomes wider toward annealed and high purity aluminum matrix. This is quite reasonable, for these models did not account for Accre, which is enhanced toward the case of annealed and high purity aluminum matrix where the punching would presumably be more extensive. By using the above punching models, Lulay®” summa- rized the effect of various types of filler and the loading direction relative to the filler reference axis. The types of filler examined by Lulay are disk with two different ratios of the diameter to thickness; 1/5 and 1/2, short fiber with two different fiber aspect ratios; 5 and 10, and sphere. Figure 32(a) and (b) show the analytical results on the relation between the composite (SiCp/6061 Al) tensile field stress and reinforcement volume fraction, cr for the case of the loading along the x-axis (filler major axis), and the x-axis (perpendicular to the filler major axis), respectively. These results are obtained based on the assumptions that MMCs were quenched with tempera- ture drop of 500 K and the yield stress of the MMCs were measured immediately upon quenching. Figure 32 sug- gests that the MMC with the disk shaped filler of aspect ratio 1/5 loaded along the x-axis (perpendiculer to the 8 Minoru Taya Spr ° [—Eshelby Type Model shelby Type Model ‘nit Residual Stress Models /# sic./al6061-76 © SiC /A16061-T6 © SiCZIAIT 100 Anneated © SiGr/AI6061 Annesieg Experiments! § 14 SIC/A17091 Annesled ¢ 3 6 co a Fig. 31. The composite yield stresses (.) normalized by the yield stress ofthe unreinforced matrix (9) as a funtion of flber aspect ratio (A) experimental results (symbols) and the predictions by the previous models (lines) which did not account for the punched-ovt Aistocations! Song, on /MPa Fig, 32 Composite yield strength (¢,) as a function of particulate volume percent (4) for various types of filler (a) loading along the syeirection, (6) along the x, x, deco disk plane) gives rise to the largest tensile yield stress, whereas the tensile yield stress (o..) of short fiber of aspect ratio 10 turns out to be the smallest. oy. of the MMC with short fiber of aspect ratio 10 becomes zero for 20.1 for the temperature drop by quenching of S00 K. This is because the magnitude of residual stress (¢1)— 011) becomes so large as to off-set the strengthing effect due t0 the punching and back stress. In reality, there must be a time lag between the quenching and measurment of oy, allowing the relaxation of the residual stress and thus leading to positive 0. rather than zero. III. Concluding Remarks Various models for strengthening of MMCs are review ed, from which the following conclusions can be drawn: ()_ The major strengthening mechanisms for MCs are the punching of prismatic dislocation loops and the resulting residual stress field due to CTE mismatch strain, and back stress due to the resistance of elastic fillers against plastically deforming matrix metal. (2). The strengthening depends highly on the shape of fillers, and the loading direction relative to the filler geometry. Acknowledgments ‘The present author is thankful to Alcan international Lid, for its financial support on this subject area and to Professor T. Mori of Tokyo Institute of Technology for his valuable comments on the punching model. REFERENCES (1) M, Taya: Tetsu-to-Hagané, 24 (1988), 172 (@)M. Taya and RJ. Arsenault: Metal Matrix Composies: Ther ‘momechanical Behavior, Pergamon Press, (1989). (@) 1. D. Eshelby: Proc. Roy. Soe. 241 (1957), 376. () T. Mura: Micromechanics of Defects in Solids, 2nd ed. Martinus Nijhoft Publishers, (1989). (8) A. Kelly: Strong Solids, 2nd ed. Oxford University Press, (1973), chapter 5 (6 VC. Nardon and K. Prewo: Seripta Metal, 20 (1986), 43, () H. Fukuda: Recent Advances in composites in the United States 4nd Japan, ASTM STP 864 e8, by 1. R. Vinson and M. Taya, ‘American Society for Testing and Materials (1985), pp. 5-15, () T. Morimoto, T. Yamaoka, H. Lilholt and M. Taya: J. Eng, Mater, Tech, 10 (1989), 7. (9) A, Kelly and K. N. Steet: Proc, Rey, Sos, Lond., A328 (1972), 261. (40) A. Kelly and W. R. Tyson: J. Mech. Phys. Solids, 13 (1968), 329, (11) L: M. Brown and W. M. Stobbs: Phil. Mag., 23 (1971), 1185 (12) K. Tanaka and T. Moris Acta Met., 18 (1970), 931 (13) 4. C. Fisher, E. W. Hast and R. H, Pry: Acta Met, 1 (1953) 336 (14) M. F, Ashby: Oxide Dispersion Strengthening, eds. G. S, Ansell et fal, Gordon and Breach Science Pub, (1966), pp. 143-205, (15) E. Orowan: Internal Stresses and Fatigue in Metals, ed. by G. M Rassweiler and W. L. Grube, Elsevier Sei. Pub (1959), pp. $9-80, (06) R. J. Arsenault and N. Shi: Mater. Sei Eng., 1 (1986), 175. (1) M, Taya and 7. Mori: Acta Met., 35 (1987), 15S (08) M. Taya, K. E, Lulay and D. 4. Lloyd: Acta Met., 39 (1991), 73 (09) R. Ebeling and M. F. Ashby: Phil. Mag., 13 (1966), 805, (20) F. Kroupa: Phil. Mag., 7(1962), 783. {@1) 4.4, Alkinson, L, M. Brown and W. M. Stobbs: Phil. Mag., 30 (0974, 1287 (€2) M, Taya, KE. Lulay, K, Wakashima and D. J. Lloyd: Mater. Si Strengthening Mechanisms of Metal Matrix Composites Eng. A, (190), 108. (23) T. Mori and K’ Tanaka: Acta Met., 24 (1973), S71 (24) LM. Brown: Acta Met. 21 (1973), 879, (25) K. Wakashima, M. Otsuka and S. Umekawa: J. Comp. Mater.,8 (1974, 391 (26) A. Daimaru and M. Taya: Progress in Science and Engineering of Composites, T. Hayashi et a, eds, Proc, ICCM-4, Tokyo, (1982), pp. 1089-1106. (27) RJ. Arsenault and M. Taya: Acta Met, 38 (1987), 651 (@8) K. Tanaka, K. Wakashima and T. Mori 1. Mech, Phys. Solids, 21 973), 207 (29) K. Wakashima, S. Kurihara and S. Umekawa: J. Japan Soe ‘Comp Mater. 2 (1976), 1 (G0) M, Taya, KE. Lulay and D. J. Lloyd: Proc. Prof. M. E. Fine ‘Symp. TMS, 1990, Oct. 2-5 (1) K. E. Lulay: PRD Dissertation, Universiy of Washington, June 150, (62) T. Mor: Private commucation, (G3) M. Taya and R. J. Arsenault: Scripta Metal, 21 (1989), 3.

You might also like