Você está na página 1de 12

Structural Nanomaterials

Joanna R. Groza Jeffrey C. Gibeling


Department of Chemical Engineering and Materials Science, University of CaliforniaDavis, Davis, California, U.S.A.

INTRODUCTION By denition, structural materials are used for mechanical strength at room and elevated temperatures, to withstand cyclical loading and for wear and fracture resistance. Structural nanomaterials (with grain sizes less than 100 nm) are distinguished by unusually high strength, hardness, and wear resistance as well as good fatigue resistance, fracture toughness, and extensive high-temperature formability. However, commercial applications have only begun to tap the potential of high-strength parts (e.g., nearly 1 GPa in nano-Al alloys, known as GIGAS,[1] and high-strength nano-SiC springs). More commonly, they are used as thermal barrier coatings, or in friction-resistant and wear-resistant rotating parts (e.g., sleeves or bearings). Nanomaterials used for structural applications may have nanometer size in one dimension (1D), two dimensions (2D), or three dimensions (3D). Generally, 1D nanomaterials are wires, 2D refers to coatings and lms, and 3D solids are used to make bulk parts. In this entry, bulk or 3D materials are emphasized, but some 1D and 2D nanomaterials used for their mechanical properties are also presented.

FABRICATION OF NANOSTRUCTURED MATERIALS Processing The primary challenge in producing bulk nanostructures is to create and to maintain grain sizes less than 100 nm in defect-free, sizable, reproducible, and reliable parts. To date, the processing of raw nanomaterials (powders and thin lms) has progressed farther than the formation of bulk parts. Nanosize powders of metals, ceramics, polymers, and composites are now available, sometimes in tonnage quantities. All of these nanostructured powders may be used for processing 2D and 3D parts.

The most common techniques to produce 2D parts are thermal spraying and electrodeposition (ED). The former includes arc, ame, plasma, detonation gun, and high-velocity oxyfuel spraying to produce metal, alloy, ceramic, and composite coatings.[2] In the spraying process, the ne microstructure develops when molten or semimolten particles rapidly solidify as they are deposited onto a substrate. Electrodeposition techniques include both electroplating and electroforming, and are typically applied to a limited number of metals, although some ceramics and composites have been also obtained.[3] Although electroplating is typically used for coatings, electroforming is a cost-effective method for free-standing ultrathin foils or thick structures, wires, plates, and complex shapes. Grain size is controlled by enhanced nucleation by charge transfer at the electrode surface, high deposition rates, and the inhibition of grain growth by using additives. Electrodeposition can be used to control grain boundary type and to achieve nanocrystalline structures with a relatively narrow grain size distribution and with few pores. Electrodeposition and sputtering also may be used to produce nanolayered materials. Generally, the processing methods for bulk nanomaterials involve one or two steps. The more commonly used two-step methods consist of producing nanomaterials in a powder form and subsequently consolidating them to a nal 3D part. These methods offer the greatest versatility in the shape, size, and weight of the nal bulk parts. All powder processing methods have been applied to sinter nanostructured powders and many of them have achieved nal grain sizes less than 100 nm.[4] The challenge is to fully densify nanopowders without losing the initial metastable features (nanoscale grain size and, sometimes, metastable phases). The consolidation methods use a full range of temperature and pressure conditions from conventional pressureless sintering to high pressure and shock consolidation. Inert gas condensation (IGC) of powders followed by densication has been a widely used technique to process bulk nanocrystalline materials.

4205
2009 by Taylor & Francis Group, LLC

Soft Supramolecular

4206

Structural Nanomaterials

Most of these methods leave processing aws, including incomplete densication, poor interparticle bonding, cracks, or trapped gases. The one-step methods include severe plastic deformation (SPD), crystallization of amorphous bulk parts, and ED. Severe plastic deformation comprises variants such as equal-channel angular forming and high-pressure torsion straining. Generally, SPD results in grain sizes greater than 100 nm, which are known as ultrane-grained structures. However, smaller grain sizes (down to 20 nm) have occasionally been achieved.[5] Intense plastic deformation has also been used to produce in situ ber or lamellar composites. A novel intense deformation process based on ball milling with in situ densication (BMID) was recently reported by Koch.[6] The crystallization of bulk amorphous alloys to form nanophases occurs on controlled thermal processing.[7] The most challenging step is the formation of bulk amorphous precursors using slow cooling rates similar to conventional casting processes, to enable the fabrication of sizeable parts. This has been achieved in a few metal-based systems (Zr, Mg, and, partially, Al). Next, controlled direct or stepwise devitrication takes place at temperatures within the supercooled liquid region, or close to the onset of crystallization. The resultant structures are typically multiphase alloys formed by the precipitation of nanoscale metastable (metal-based or quasi-crystalline) or stable (intermetallic) phases in an amorphous matrix. The type and the size of nanocrystalline components vary with alloy chemistry and processing conditions.
Soft Supramolecular

Dependence of Properties on Processing Most processing methods result in some imperfections in the nal bulk part. In conventional materials, wellestablished standards control aw size and distribution, resulting in a negligible effect on mechanical properties. However, such aws have a more critical effect on the mechanical properties of nanomaterials, and are different in nature and size than in conventional materials. The most common imperfections in nanocrystalline bulk materials are pores, trapped gases, impurity contamination, cracks, imperfect bonding, surface defects, residual stresses, and texture. Although some defects are common to several techniques (e.g., contamination, surface defects, or pores), texture is typical in ED, residual stresses are most severe in SPD, and incomplete bonding and densication are observed in powder consolidation and thermal spraying. Defect characterization in nanomaterials is a relatively new eld, and appropriate guidelines are denitely required.

The most detailed defect characterization has been carried out for the aws specic to the two-step powder consolidation methods, particularly pores that result from incomplete densication.[8,9] At small grain sizes (less than 10 nm), a density shortfall may also originate from the high volume of low-density grain boundaries. Van Swygenhoven, Spaczet, and Caro.[10] calculated densities of 97% for a 10- to 12-nm grain size Ni and 95.2% for a 3.4-nm grain size Ni. Experiments conrmed a 97.6% density value measured by a high-precision Archimedes method in 10-nm grain size samples produced by IGC.[11] Processing aws in nanomaterials have an adverse effect on mechanical properties. The modulus of elasticity or Youngs modulus E is a reection of interatomic bonding and is largely dependent on defects. The low experimental E values reported in early studies of nanocrystalline materials have been unquestionably attributed to defects such pores, poor grain-to-grain bonding, or deciencies in measuring techniques.[9,12] More recent experimental measurements on low-defect nanomaterials demonstrate that grain size has a negligible effect on Youngs modulus.[13] Simulations also showed no variation in Youngs modulus down to a grain size of 10 nm.[14] Below this limit, when the slightly lower-density grain boundary volume becomes predominant, the modulus decreases with decreasing grain size. Some calculated values are as low as about 75% of the modulus of bulk metals at 36 nm grain size.[14,15] All other mechanical properties, including strength, hardness, and ductility, are sensitive to microstructure and are even more adversely affected by defects. Ample evidence exists for inferior strength and ductility values, particularly in tensile testing, because of pores and trapped gases, cracks, imperfect bonding, and surface defects. Agnew et al.[8] Sanders, Youngdahl, and Weertman.[9] Legros et al.[13] and Nieman, Weertman, and Siegel.[16] carefully documented the dependence of strength and ductility on processing aws in metals. He and Ma[17] characterized the variation of hardness with porosity in the Fe3Al intermetallics. Hardness values were reported to double by processing improvements to produce dense metals.[9] The reduction of the surface imperfections has been shown to increase the tensile strength by up to a factor 4 by preventing premature fracture.[16] Consequently, reduced size specimens (e.g., micro and miniature tensile and bend specimens) have been used to lower the defect probability.[13,18,19] The elimination of some defects by annealing can also increase the strength of metals[20,21] and intermetallics[2023] by decreasing porosity, relaxing the grain boundary structure toward equilibrium, or reducing internal stresses. As is detailed in Strength and Ductility, ductility is the property most affected by processing aws. The

2009 by Taylor & Francis Group, LLC

Structural Nanomaterials

4207

lack of good interparticle bonding is the major cause of brittle failure in consolidated samples tested in tension or bending.[9,24,25] As an example, 80-nm Fe3Al produced by shock wave consolidation of mechanically alloyed powders displayed brittle behavior in tension but a large plastic strain in compression (true plastic strain of 1.4).[26] Similar to strength, an improvement in ductility was reported after annealing at low temperatures without altering the grain size (e.g., an increase from 2 to 7.3% in 30-nm Cu) because of grain boundary relaxation.[27] Electrodeposition yields better ductilities than powder consolidation because of reduced porosity and good interparticle bonding. However, some ductility problems have been noted when using grain growth inhibitors, or because of texture development. Throughout this entry, an effort has been made to exclude materials with signicant processing deciencies, when these are reported. In conventional materials, it is acceptable to use average grain size to characterize mechanical properties. This approach is not appropriate in nanomaterials because of the size dependence of the deformation behavior. For instance, some large recrystallized 1 5 mm size grains enabled larger ductility values in nano-Cu vs. nano-Ni, which had fewer such recrystallized grains and was consequently brittle.[13] All processing methods result in a distribution of grain sizes. In nanomaterials, grain size measurements are commonly carried out by X-ray diffraction (XRD) and transmission electron microscopy (TEM). Transmission electron microscopy enables a direct measurement of grain sizes and grain size distribution. In contrast, XRD techniques provide only an average grain size, the value of which also depends on the specic method used [e.g., Scherrer, WarrenAverbach (WA), or WilliamsonHall].[28] Average TEM grain size and X-ray results are in good agreement when the grain size distribution is narrow and XRD corrections are made (Fig. 1A). Log-normal distributions of grain sizes are usually used. When grain size distribution is described by volume fractions, the peak value may be quite different from the TEM number fraction and XRD values (Fig. 1A). Volume fractions more realistically reect the effect of large grains, which may be predominant on mechanical properties.[21] The standard deviation of the volume fraction also affects the grain size distribution (Fig. 1B) and, consequently, the mechanical properties (Fig. 1C).

Fig. 1 Effects of grain size distributions on mechanical properties. Source: From Refs.[14,21]. (A) Comparison of grain size values given by TEM and XRD in IGC copper. The open circles are the relative number of grains in a given size range. The solid and dotted lines indicate the log-normal ts of the data points as numberfraction and volumefraction curves, respectively. (B) Effects of the standard deviation size Slnv on the grain size distribution and (C) on stressstrain curves for the same average grain size (d 20 nm). The IGC Cu is heavily twinned and the grain size value actually refers to the twin size. Source: MRS Bulletin and J. R. Weertman.

STRENGTH AND DUCTILITY The general description of the mechanical behavior in tension is given in Fig. 2.[29] Brittle materials are characterized by a linear elastic stress (s)strain (e)

response and fracture with no plastic deformation (Fig. 2A). Ductile materials yield beyond the elastic region and then fracture after a certain elongation (Fig. 2B). Fig. 2B denes the yield strength (sy), tensile strength (sUTS), and elongation to fracture (ef). In the

2009 by Taylor & Francis Group, LLC

Soft Supramolecular

4208

Structural Nanomaterials

Fig. 2 Stressstrain curves for brittle (A) and ductile (B) materials. The slope of the linear portion of the se plot denes the elastic modulus by the Hookes law (s Ee).

plastic region, work hardening is expressed using a true strain value e by: s K en 1

where K is a constant and n is the strain hardening exponent. Strain or work hardening is attributed to the generation and interaction of the dislocations. It controls the amount of uniform elongation at stresses larger than yield strength sy. A dimensional instability in tension, or necking, typically starts to develop at maximum load. Grain size renement results in the classical grain boundary strengthening (expressed in terms of yield strength sy but also valid for sUTS or hardness H), which is described by the HallPetch (HP) relationship:[29]
Soft Supramolecular

of dislocations emitted by grain boundaries (Fig. 3B). A third model developed by Meyers and Ashworth[29] is a variant of the dislocation emission model with localized plastic ow at a work-hardened grain boundary layer (Fig. 3C). Plastic deformation occurs when the applied stress exceeds the higher-ow stress of the grain boundary region. The resulting grain size dependence of the yield strength is the sum of d1/2 and d1 dependencies: sy so kd1=2 k0 d1 3

where k0 is a different constant from k. The second term is dominant at large grain sizes, whereas the third term becomes important at small grain sizes. All models assume conventional dislocation generation and mobility. Strength of Nanostructured Materials By extrapolating the grain size to the nanometer range, both strengthening and an increase in ductility were predicted.[30] Experimentally, this prediction has been observed for a number of materials (Fig. 4). According to the HP relationship, a 1000-fold reduction in grain size from microns to nanometers yields more than a 30-fold increase in strength, provided that the same dislocation mechanisms are

sy s0 k d1=2

where s0 is the friction stress to move dislocations, k is a material-dependent and temperature-dependent constant, and d is grain size. The HP relationship can be rationalized by two basic dislocation descriptions: 1) the dislocation pileup model, in which yielding occurs as a result of stress concentrations caused by pileups (Fig. 3A); and 2) the dislocation network model, in which the slip is because

Fig. 3 Dislocations models to explain the HP behavior developed by: (A) Cottrell (dislocation pileup); (B) Li (dislocations generated at ledges in grain boundaries); and (C) Meyers and Ashworth. Source: From Ref.[21].

2009 by Taylor & Francis Group, LLC

Structural Nanomaterials

4209

Fig. 4 Strengthductility values of nanostructured materials (lled symbols) compared to conventional counterparts (open symbols: circlesdispersion-strengthened Al alloys; trianglescold-worked steels).

operational. Experimentally, strength or hardness was found to increase by a factor of 510 by reducing the grain size from microns down to the 30- to 50-nm level.[9,25,31] Similar high-strength values for nanocrystalline metals and alloys were observed in compression, or by hardness measurements, in which processing defects are less detrimental than in tension. Yield strengths greater than 1 GPa were found in Pd (40 50 nm), and yield strengths close to 1 GPa were found in Cu (20 nm).[9] Ball-milled and densied Fe (24 nm) had a true compressive strength of approximately 3 GPa,[24] whereas IGC-processed 21-nm Ni failed in compression at more than 2.1 GPa.[21] Often the spread of experimental data is quite large, but reasonable HP correlations at the smallest grain sizes have been revealed (i.e., stronger materials at smaller grain size) (Fig. 5). Deviations from HP strengthening, where the slope decreases or the curve attens, also have been reported, usually at grain sizes smaller than about 30 nm (Fig. 5B and C).[24,32] An inverse HP effect, in which strength decreases at the smallest grain sizes, has been reported in alloys such as NiP.[33] This effect was most often reported in specimens annealed to increase the initial small grain size in metals or intermetallics.[20,22] Generally, these results must be treated with caution because either processing artifacts may not have been completely eliminated, or other processes (e.g., precipitation of a second phase in NiP alloys or in mechanically alloyed Fe) may have interfered with the measurements.[24,33] Various mechanisms have been proposed to rationalize the deviations from HP strengthening, such as the breakdown of dislocation pileups, lack of

Fig. 5 Grain size dependence of mechanical properties in nanocrystalline metals: (A) Cu (IGC) Source: From Refs.[9,21,31].); (B) Fe (ball-milled/densied) Source: From Ref.[24].; and (C) Ni (electrodeposited) Source: From Ref.[32]. The dotted lines are the HP extrapolations of coarse grain data Source: Materials and Metallurgical Transactions and J. R. Weertman.

dislocation activity, and increased contributions from diffusional deformation mechanisms (creep and grain boundary sliding).[14,21,34,35] Cheng, Spencer, and Milligan [34] proposed a deformation mechanism map showing the transition from the dislocation mechanisms to intergranular processes. Transmission electron microscopy studies have conrmed the lack of dislocation structure below a critical grain size of 10 30 nm.[13,34]

2009 by Taylor & Francis Group, LLC

Soft Supramolecular

4210

Structural Nanomaterials

To explain the HP behavior, a model based on dislocation generation from grain boundary sources was recently proposed by Cheng, Spencer, and Milligan.[34] The model also predicts the observed tension/compression asymmetry based on the pressure dependence of dislocation self-energy during bowout. Composite models considered grain interior and grain boundary properties, and, sometimes, triple junction lines and quadruple nodes.[21] The width of grain boundary or adjacent area is still under debate, but the maximum yield strength was found at similar grain size value (about 10 nm). Composite models that are more consistent with microstructural results by accounting for a dispersion in grain sizes have been developed.[28,35] These models combine deformation mechanisms that are active in various grain size ranges. Masumara, Hazzledine, and Pande[35] considered a combination of dislocation glide for grains larger than a critical size d and Coble creep below d . In the model developed by Mitra et al.[28] the critical grain size d differentiates grains that undergo plastic deformation (d > d ) from grains that remain elastic (d < d ).[21] The overall calculation was based on the plastic deformation of grains in an elastic matrix that used the mechanics of non-deformable inclusions. The average grain size was given by a volume averaging of the grains, similar to previous work.[35] For the same average grain size, the volume fractions are shown to be greatly inuenced by the standard deviation (Fig. 1B). In turn, a larger standard deviation (with more small-size grains deformed by creep) results in softer behavior (Fig. 1C).
Soft Supramolecular

increase in mechanically alloyed Zr55Al10Cu30Ni5 reinforced with 35-nm W particles obeyed the rule of mixtures.[38] In contrast, ber and lamellar composites (multilayers) display hardness values far in excess of the rule of mixture values because of the restriction of dislocation propagation by the second layer.[39]

Ductility of Nanostructured Materials Ductility is dened as the ability to plastically deform prior to fracture. Therefore it is a critical property for both the performance and the manufacturing of nanostructured materials.[18] Ductility is measured by percent elongation before failure (e.g., elongation to failure ef in Fig. 2B), or percent reduction of area after necking in tension. With the decrease of grain size to nanometer levels, increases in ductility have been predicted.[18,29,30] According to Cobles theory, the signicantly larger interfacial area enhances deformation by diffusional transport and grain boundary sliding. This enhanced sliding at the more numerous grain boundaries should result in an increased deformation in metals even at room temperature. The most dramatical change of ductility was predicted for normally brittle materials such as intermetallic compounds.[18] The prediction is based on conventional material behavior, in which a ne grain size decreases the stress required to nucleate microcracks as compared with that required to propagate them. If crack propagation occurs only at high stress values, the ne-grained materials may undergo plastic strain and work hardening before failure. In other words, in tension testing (Fig. 2B), a material displays ductile behavior if plastic deformation occurs without the onset of catastrophical failure. Chan[40] calculated a critical grain size (e.g., 2.8 mm in NiAl) below which crack growth occurs in a stable manner to render the intermetallics ductile. Experimentally, such behavior was observed in the now-classical results for NiAl alloys, which become ductile at a grain sizes below 20 mm. A high ductility (e.g., greater than 11.5% in compression) was reported in NiAl with submicron grain size.[41] Ductility of more than 5% in compression was reported by Kimura[42] in TiAl (grain size less than 22 nm) crystallized from an amorphous precursor. However, this behavior was not reproduced in other intermetallics. In metals and alloys, ductilities lower than in conventional materials have been reported.[18,43] Despite the efforts to improve the processing and testing, it is difcult to differentiate intrinsic from aw-limited ductilities. When the grain size is reduced to less than 25 nm, ductility values in tension or bending decrease from about 50% in conventional materials to 12%.[18]

Two-phase alloys and nanocomposites The addition of a second phase produces greater mechanical strength in nanomaterials just as in conventional materials. Bulk materials made of quasi-crystalline or icosahedral phases (less than 50 nm) in an Al matrix (510 nm; AlCr or MnLn, e.g., Al93Mn5Co2) exhibited strength values of 550850 MPa with 525% elongations.[1] High strength and high ductilities also have been found in the case of nanoparticles reinforcing amorphous matrices in Mg-based, Al-based, and Zrbased alloys.[7] The nanosize particles prevent fracture because of improved interfacial bonding, which favors load transmission from one phase to the other. Tensile strength doubled in ED NiSiC composites by reducing matrix grain size to 1015 nm. SiC particles (400 nm) up to about 2 vol.% changed the fracture from brittle in nanocrystalline Ni to ductile (e > 2%) in NiSiC composites.[36] However, only a modest increase in hardness was observed by the addition of 5 vol.% SiO2 particles (20 nm) to 60 nm of SPD copper (microhardness of 2800 vs. 2600).[37] The hardness

2009 by Taylor & Francis Group, LLC

Structural Nanomaterials

4211

Fig. 6 Tensile elongation vs. grain size in nanocrystalline Zn and Cu. Source: C. C. Koch.

Little tensile ductility (less than 2%), or completely brittle behavior is commonly reported in metals even after careful processing (e.g., in IGC-processed 28-nm Ni,[13] ED-processed 27.5- to 30-nm Ni,[44] and IGCprocessed 27-nm Cu,[19]). An illustration of the variation of ductility with grain size is provided in Fig. 6 for two nanocrystalline metals. It is noteworthy that higher ductility is observed in the hcp metal (Zn) rather than fcc (Cu), which is contrary to the dislocation-based plastic behavior at conventional grain sizes. A few truly ductile nanomaterials (ef > 5%, as a structural design criterion) are shown in Figs. 4 and 6. If the dislocation generation and interactions are similar to the coarse-grained materials, strain hardening can be expected in nanosize materials. Strain hardening controls the amount of uniform elongation at stresses larger than yield strength. Generally, hardening was observed, sometimes, to be of a signicant value and at a large plastic strain such as in a 60-nm SPD Cu[37] or in tension or compression in other metals.[43] However, a different behavior with little or no strain hardening and shear banding has also been reported in nanomaterials, mostly bcc metals and alloys (Fe, FeCu, and Ta), but also in fcc and hcp and intermetallics.[25,43,45] Perfectly plastic behavior with no strain hardening is typical for materials in which initial plastic ow is difcult (e.g., amorphous or highly strained metals). In nanomaterials, one assumption is that the grain boundaries introduce sufcient defects similar to amorphous materials such that unstable plastic ow is initiated as soon as the yield

strength is reached at a local stress concentration.[9] Crack growth will occur, resulting in failure without prior macroscopical ductility.[24] Brittleness may be observed if the stress to activate surface cracks is lower than the stress to activate FrankRead dislocation sources and no crack blunting mechanism is in operation.[31] Calculations show that tensile stresses at which brittle fracture occurs indeed correspond to the stresses required for brittle fracture in materials with micron-size aws. The other source of low ductility may be the absence of effective mechanisms for plastic deformation, mainly the lack of dislocation activity in grains below a size of about 30 nm.[13,34] The lack of noticeable dislocations is consistent with the brittle fracture exhibited by a 28-nm Ni specimen.[13] Zimmerman et al.[36] also reported higher ductilities in 40- vs. 14nm grain size ED Ni. Grain boundary sliding and grain rotation were shown to be the deformation mechanisms in some nanomaterials that exhibited ductile behavior such as electrodeposited Ni or 10-nm sputtered gold lms.[45,46] Indirect evidence of grain boundary sliding is provided by the strain rate increase as the grain size becomes smaller and there is no change in the total amount of grain boundary deformation.[14] Strain rate shows no grain size dependence at larger grain sizes (greater than 10 nm). Molecular dynamics simulations showed grain boundary sliding responsible for inverse HP behavior even at low temperature and without any porosity.[14,15]

2009 by Taylor & Francis Group, LLC

Soft Supramolecular

4212

Structural Nanomaterials

Two-phase nanomaterials Second-phase particles increase the ow stress on local shear planes, thus resulting in more uniform deformation and higher elongation. High elongations (525%) are exhibited by Al-based alloys with nanoscale icosahedral quasi-crystalline phases in a ductile fcc Al matrix (obtained by extrusion).[1,7] A homogeneous dispersion of nanosize crystalline particles in an amorphous matrix is able to suppress local shear deformation in Mg-based and Zr-based alloys.[7] In some Zr-based or Al-based alloys, the precipitation of nanoscale particles induces considerable strengthening but decreases ductility, as compared with a fully amorphous alloy.[7] When the precipitate amount is greater than 50%, the alloys become brittle. However, second-phase particles in metal-based alloys seem to have a different effect in nanosize matrices than in conventional dispersion-strengthened alloys. For instance, an addition of 0.7 vol.% SiC to a nano-Ni matrix increased the ductility to 3.4% as compared with 0.2% in pure 14-nm Ni.[36]

Fracture toughness was found to decrease with a reduction in grain size below 1 mm in TiAl, whereas it remained constant down at 12 MPa m1/2 to $40 nm, and then decreased to about 5 MPa m1/2 in FeAl.[23,51] Similarly, very low values of fracture toughness (2 MPa m1/2) were reported in Fe45% Al and Al3Ti.[52] Because grain boundary diffusion and sliding can occur in nanocrystalline solids at room temperature, these processes may control toughness at low crack propagation rates. This suggestion is supported by the observation of Mirshams et al.[53] of the R-curve fracture behavior in nanocrystalline nickel produced by ED, in which Kc increases with the growth of a crack in a slow, stable manner. As with other mechanical properties, these differences in behavior may be partially attributable to the different defect densities associated with the various processing techniques. Additional data are needed to more fully distinguish between intrinsic grain size effects and the inuence of processing on fracture toughness.

FATIGUE PROPERTIES The study of fatigue is normally divided into two important aspects: 1) deformation under cyclical loading conditions leading to crack initiation; and 2) the rate of growth of a preexisting aw or crack. Only a few studies of fatigue of nanocrystalline or ultranegrained materials, dealing mainly with cyclical deformation, have been published to date. As with other mechanical properties, these studies suggest that the processing method is important. Nanocrystalline copper prepared by IGC and compaction revealed no cyclical hardening or softening during tensiontension loading.[54] However, some creep strain superimposed on the cyclical response was evident. Of particular note is the observation of shear bands extending over many grain diameters on the surface of the fatigued specimens. In addition, the material was observed to protrude from the surface much like the extrusions observed in conventional copper alloys. This nding suggests that some mechanism of strain localization on a scale larger than the grain size is operating. Similar surface observations were reported for SPD-processed ultrane-grained copper.[55] On a strainlife basis, this ultrane-grained copper was less fatigue-resistant than conventional materials. However, the mechanical response of this material reveals a decreasing stress amplitude with an increasing number of cycles or cyclical softening, as expected for a predeformed metal. A similar material was more fatigue-resistant than conventional copper when compared on a stresslife or SN basis.[56] Surface markings have also been observed in a yttria-stabilized zirconia ceramic, although in this case the strain localization is most likely

FRACTURE AND FAILURE Fracture toughness (Kc) is dened as the value of a stress intensity factor at a crack tip to induce crack propagation and failure as given by:[29] p Kc Y sf a
Soft Supramolecular

where Y is a dimensionless geometry factor, sf is the applied stress at fracture, and a is the crack length. Usually, a high fracture toughness value is an indication of ductile behavior because work is needed to produce local plastic strain at the crack tip. There is some expectation that decreasing grain size would also result in increased Kc. This effect was expected to be particularly benecial for normally brittle materials such as intermetallic compounds. However, Armstrong[47] showed that there is a limit when the crack tip plastic zone size is approximately equal to the grain size, below which the toughness should decrease as the grain size decreases. Only limited measurements of fracture toughness of nanomaterials are available in the literature, owing in part to the stringent specimen size requirements.[29] The available data suggest that, in general, the anticipated improvement in fracture toughness with decreasing grain size has not been realized. Rather, experiments and modeling indicate that the fracture resistance in both metals and intermetallics appears to be limited by the nucleation of nanovoids at grain boundaries and intergranular crack propagation.[4850]

2009 by Taylor & Francis Group, LLC

Structural Nanomaterials

4213

because of diffusional deformation processes rather than dislocation processes.[57] It is not yet clear whether any of our understanding of fatigue in conventional materials can be extended to the nanometer grain size level. However, it is important to recognize that the features observed in fatigue are normally on the order of 1 mm in scale, hence are clearly coarser than the nanocrystalline grain size. Additional fundamental studies are needed.

where AC is a constant, s is the stress, G is the shear modulus, b is the Burgers vector, d is the grain boundary thickness, Db is the grain boundary diffusivity, k is the Boltzmanns constant, and T is the temperature. The Coble mechanism is favored at high stresses and low temperatures, even down to room temperatures in nanocrystalline materials. An alternative deformation mechanism involves grain boundary sliding, accommodated by the diffusion of vacancies in the grain boundaries, and is described by a similar equation:[29] _ E AGBS   dDb Gb b 3 s 2 kT d G 6

WEAR The processes involved in tribological conditions are heavy local deformation, work hardening, cyclical loading, and fracture, collectively resulting in material removal. Nanocrystalline materials have been used for wear resistance in bulk (e.g., cutting tools), or for surface applications (coatings) because of their high hardness. The materials for cutting tools are carbides (WC or (Fe, Mo, Cr, V)xC) in soft matrices (Co, Fe, or intermetallics), but most materials used commercially have grain sizes in the submicron range. Wear resistance has been shown to scale with hardness in metal (Ni and NiP), composites (NiSiC), and lamellar coatings (e.g., CuNi).[25] The friction coefcient in sliding and rubbing wear was found to decrease by a factor of 23 by decreasing grain size (e.g., in sputtered Al layers).[3,25] The scratch indentation of 10 wt.% TiO2 (32 nm) in an epoxy matrix was 33% smaller than for pure epoxy.[58]

ELEVATED TEMPERATURE BEHAVIOR At elevated temperatures, creep deformation under sustained loads is of importance in structural materials. Whereas a ne grain size is benecial in promoting high strength at low temperatures, creep resistance normally becomes worse as the grain size decreases. Rather than deforming by dislocation motion, the principle mechanisms of creep deformation in nanomaterials involve diffusional processes because of enhanced high diffusivity at grain boundary and triple junction paths. Specic creep mechanisms are characterized by the dependence of creep rate on stress, temperature, and grain size. Temperature dependence is expressed through the activation energy for the diffusion of vacancies. For nanocrystalline materials, the most commonly reported mechanisms involve vacancy transport along grain boundaries. This process is known as Coble creep, and the strain rate is represented by an equation of the form:[29]   dDb Gb b 3 s _ E AC kT d G 5

Grain boundary sliding is the mechanism to control superplastic deformation, in which ne-grained materials exhibit extended ductilities of up to several thousands of percent. One key question is whether nanocrystalline solids exhibit superplasticity at low temperatures and high strain rates. If so, this would enable these materials, especially otherwise brittle ceramics and intermetallics, to be formed readily into near-net shapes. Finally, we note that the above two equations differ primarily in the dependence of strain rate on grain size and stress. Thus by measuring these two dependencies, it is possible to distinguish which mechanism controls the creep deformation. A variety of pure metallic, intermetallic, and ceramic materials have been creep-tested in nanocrystalline form.[59] In addition to the usual processing problems (porosity and suitably large specimens), the interpretation of the creep results is hampered by possible grain growth during testing, testing under constant load rather than constant stress, and often limited amounts of creep strain that may not represent a steady state to which the above equations apply. In spite of these difculties, a number of common ideas have emerged from the available data. Tests of nanocrystalline pure copper and nickel produced by ED appear to demonstrate Coble creep, even at room temperature.[6062] However, there is some evidence that grain boundary sliding controls the deformation of Ni with the nest grain sizes (6 nm).[61] In addition, when the grain boundaries are of low angle rather than of high angle as in materials prepared by IGC,[63] the creep rates are orders of magnitude slower than those predicted by the Coble creep equation. The creep of nanophase NiP and FeBSi alloys (approximately 25 nm in grain size) produced by crystallizing amorphous alloy ribbons also follows the Coble creep equation, although absolute comparisons cannot be made in the former case because of a lack of diffusivity data.[64,65] In a similar manner, nanocrystalline ceramics such as monoclinical zirconia and yttria-stabilized zirconia

2009 by Taylor & Francis Group, LLC

Soft Supramolecular

4214

Structural Nanomaterials

exhibit creep deformation at elevated temperatures with stress dependencies of 1.7 and 1.4, respectively.[66,67] These values are difcult to interpret, but the grain size dependence in the former case clearly suggests that Coble creep is the operating mechanism. Superplastic-like behavior has been observed in nanocrystalline nickel[68] and copper,[69] which exhibited strains up to 5100% during cold rolling of 28-nm grain size material at room temperature. The intermetallic Ni3Al has also been shown to undergo superplastic deformation, with a stable grain structure because of the ordered nature of this material.[70] Perhaps the greatest interest is in the potential of nanocrystalline ceramics to deform superplastically. An evidence of such behavior has been presented by a number of authors for TiO2, yttria-stabilized zirconia, and MgOAl2O3 spinels.[71] At elevated temperatures, grain growth inevitably occurs and may even contribute to achieving high elongations. Two-Phase Nanomaterials Current experimental evidence suggests that multiphase nanocrystalline materials may exhibit superior ductility, especially at elevated temperatures. For example, Al alloys with nanoscale quasi-crystalline icosahedral (I) phases retain their good mechanical strength up to high temperatures (350 MPa at 473 K and 200 MPa at 573 K).[72] This high strength is stable on long exposure to elevated temperatures (1000 hr at 573 K) because of the high microstructural stability of the I phase. In Mg-based alloys (e.g., MgAlGa), ductility increases with temperature, and superplastic behavior was reported.[73] However, much like the ceramic materials, the microstructure had coarsened.

enhance the interface processes to contribute to plasticity. A better understanding of the deformation mechanisms as a function of crystal structure, stacking fault energy, grain size, and distribution is yet to be developed. Additional data are needed to more fully distinguish between intrinsic grain size effects and the inuence of processing on mechanical properties. Simulations of deformation behavior at larger grain sizes, in bcc, hcp, or complex crystal structures, are desirable. A wider application of structural nanomaterials requires a more reliable mechanical property characterization and a more aggressive shift from the concept to the design and fabrication of optimum parts with unique properties at low cost.

ACKNOWLEDGMENTS NSF and ARO support for the work on nanomaterials is gratefully acknowledged. We are grateful to colleagues who provided their preprints and to Crystal Merril for her help with gures.

REFERENCES
1. Inoue, A.; Kimura, M. Development of highstrength Al-based alloys by synthesis of new multicomponent quasicrystals. MRS Symp. Proc. 1999, 553, 495506. 2. Lau, M.L.; Lavernia, E.J. Thermal Spray Processing of Nanocrystalline Materials. In Nanostructured Materials; Koch, C.C., Ed.; Noyes Publications: Norwich, NY, 2002; 5172. 3. Erb, U.; Aust, K.T.; Palumbo, G. Electrodeposited Nanocrystalline Materials. In Nanostructured Materials; Koch, C.C., Ed.; Noyes Publications: Norwich, NY, 2002; 179222. 4. Groza, J.R. Nanocrystalline Powder Consolidation Methods. In Nanostructured Materials; Koch, C.C., Ed.; Noyes Publications: Norwich, NY, 2002; 115178. 5. Valiev, R.S. Structure and mechanical properties of ultrane-grained metals. Mater. Sci. Eng., A Struct. Mater.: Prop. Microstruct. Process. 1997, 234236, 5966. 6. Zhang, X.; Wang, H.; Scattergood, R.O.; Narayan, J.; Koch, C.C.; Sergueeva, A.V.; Mukherjee, A.K. Tensile elongation (110%) observed in ultrane-grained Zn at room temperature. Appl. Phys. Lett. 2002, 81, 823825. 7. Eckert, J. Structure Formation and Mechanical Behavior of Two-Phase Nanostructured Materials. In Nanostructured Materials; Koch, C.C., Ed.; Noyes Publications: Norwich, NY, 2002; 423526. 8. Agnew, S.R.; Elliott, B.R.; Youngdahl, C.J.; Hemker, K.J.; Weertman, J.R. Microstructure and mechanical

Soft Supramolecular

CONCLUSION The production and study of structural nanomaterials have challenged the conventional description of microstructural features and mechanical properties. New demands have surfaced to carefully monitor grain sizes and distributions, dislocation structures, as well as other defects (pores, cracks, impurities, and surface and internal stresses). The full characterization of the property dependence on processing is yet to be performed. The fabrication of sizable, reproducible, and reliable parts is an additional challenge driven by the high strength values achieved so far. The conventional dislocation-based strengthening and yielding mechanisms show limitations, which have fostered an exploration of new mechanisms specic to the nanoscale. The inherently large fractions of atoms in non-equilibrium positions (at grain boundaries)

2009 by Taylor & Francis Group, LLC

Structural Nanomaterials

4215

9.

10.

11.

12.

13.

14.

15.

16.

17.

18.

19.

20.

21.

22.

23.

24.

25. 26.

behavior of nanocrystalline metals. Mater. Sci. Eng., A Struct. Mater.: Prop. Microstruct. Process. 2000, 285, 391396. Sanders, P.G.; Youngdahl, C.J.; Weertman, J.R. The strength of nanocrystalline metals with and without aws. Mater. Sci. Eng., A Struct. Mater.: Prop. Microstruct. Process. 1997, 234236, 7782. Van Swygenhoven, H.; Spaczer, M.; Caro, A. Characterization of the microstructure of nanophase Ni: A molecular dynamics simulation study. Nanostruct. Mater. 1999, 12, 629632. Sanders, P.G.; Eastman, J.A.; Weertman, J.R. Pore distribution in nanocrystalline metals by small-angle neutron scattering. Acta Mater. 1998, 46, 41954202. Krstic, V.; Erb, U.; Palumbo, G. Effect of porosity on Youngs modulus of nanocrystalline materials. Scr. Mater. 1993, 29, 15011504. Legros, M.; Elliott, B.R.; Rittner, M.N.; Weertman, J.R.; Hemker, K.J. Microsample tensile testing of nanocrystalline metals. Philos. Mag., A 2000, 80, 10171026. Weertman, J.R.; Farkas, D.; Hemker, K.; Kung, H.; Mayo, M.; Mitra, R.; Van Swygenhoven, H. Structure and mechanical behavior of bulk nanocrystalline materials. MRS Bull. 1999, 24 (2), 4450. Schitz, J.; Di Tolla, F.D.; Jacobsen, K.W. Softening of nanocrystalline metals at very small grain sizes. Nature 1998, 391, 561563. Nieman, G.W.; Weertman, J.R.; Siegel, R.W. Mechanical behavior of nanocrystalline Cu and Pd. J. Mater. Res. 1991, 6, 10121027. He, L.; Ma, E. Full-density nanocrystalline Fe29Al 2Cr intermetallic from mechanically milled powders. J. Mater. Res. 1996, 11, 7280. Koch, C.C.; Morris, D.G.; Lu, K.; Inoue, A. Ductility of nanostructured materials. MRS Bull. 1999, 24 (2), 5458. Gertsman, V.Y.; Hoffmann, M.; Gleiter, H.; Birringer, R. The study of grain size dependence of yield stress of Cu for a wide grain size range. Acta Mater. 1994, 42, 35393544. Chokshi, A.H.; Rosen, A.; Karch, J.; Gleiter, H. On the validity of the HallPetch relationship in nanocrystalline materials. Scr. Metall. 1989, 23, 16791684. Weertman, J.R. Mechanical Behavior of Nanocrystalline Metals. In Nanostructured Materials; Koch, C.C., Ed.; Noyes Publications: Norwich, NY, 2002; 397421. Chang, H.; Alstetter, C.J.; Averback, R.S. Characteristics of nanophase TiAl produced by inert gas condensation. J. Mater. Res. 1992, 7, 29622970. Morris-Munoz, M.A.; Dodge, A.; Morris, D.G. Structure, strength and toughness of nanocrystalline FeAl. Nanostruct. Mater. 1999, 11, 873885. Malow, T.R.; Koch, C.C. Mechanical properties, ductility, and grain size of nanocrystalline iron produced by mechanical attrition. Metall. Mater. Trans. 1998, 29A, 22852295. Morris, D.G. Mechanical Behavior of Nanostructured Materials; Trans. Tech. Publ. Ltd.: Zurich, 1998. Jain, M.; Christman, T. Synthesis, processing, and deformation of bulk nanophase Fe28Al2Cr intermetallic. Acta Metall. Mater. 1994, 42, 19011911.

27. Gunther, B.; Baalman, A.; Weiss, H. Preparation and mechanical properties of ultrane grained metals. MRS Symp. Proc. 1990, 195, 611615. 28. Mitra, R.; Ungar, T.; Morita, T.; Sanders, P.G.; Weertman, J.R. Assessment of Grain Size Distributions in Nanocrystalline Cu and Their Effect on Mechanical Behavior. In Advanced Materials for the 21st Century; Chung, Y-W., Dunand, D.C., Liaw, P.K., Olson, G.B., Eds.; TMS: Warrendale, 1999; 553564. 29. Meyers, M.A.; Chawla, K.K. Mechanical Behavior of Materials; Prentice Hall: Upper Saddle River, NJ, 1999. 30. Gleiter, H. Nanocrystalline materials. Prog. Mater. Sci. 1989, 33, 224315. 31. Sanders, P.G.; Eastman, J.A.; Weertman, J.R. Elastic and tensile behavior of nanocrystalline copper and palladium. Acta Mater. 1997, 45, 40194025. 32. Erb, U. Electrodeposited nanocrystals. Nanostruct. Mater. 1995, 6, 533538. 33. Lu, K.; Wei, W.D.; Wang, J.T. Microhardness and fracture properties of nanocrystalline NiP alloy. Scr. Metall. Mater. 1990, 24, 23192323. 34. Cheng, S.; Spencer, J.A.; Milligan, W.W. Strength and tension/compression asymmetry in nanostructured and ultrane-grained materials. Acta Mater. 2002. submitted for publication. 35. Masumara, R.A.; Hazzledine, P.M.; Pande, C.S. Yield stress of ne grained materials. Acta Mater. 1998, 46, 45274534. 36. Zimmerman, A.F.; Palumbo, G.; Aust, K.T.; Erb, U. Mechanical properties of nickel silicon carbide nanocomposites. Mater. Sci. Eng., A Struct. Mater.: Prop. Microstruct. Process. 2002, 328, 137146. 37. Alexandrov, I.V.; Zhu, Y.T.; Lowe, T.C.; Islamgaliev, R.K.; Valiev, R.Z. Microstructures and properties of nanocomposites obtained through SPTS consolidation of powders. Metall. Mater. Trans. 1998, 29A, 22522260. 38. Eckert, J.; Kubler, A.; Schultz, L. Mechanically alloyed Zr55Al10Cu30Ni5 metallic glass composites containing nanocrystalline W particles. J. Appl. Phys. 1999, 85, 71127119. 39. Baker, S.; Nix, W.D. Mechanical properties of compositionally modulated AuNi thin lmsNanoindentation and microcantilever deection experiments. J. Mater. Res. 1994, 9, 31313145. 40. Chan, K.S. Theoretical analysis of grain size effects on tensile ductility. Scr. Metall. Mater. 1990, 24, 17251730. 41. Dymek, S.; Dollar, M.; Hwang, S.J.; Nash, P. Deformation mechanisms and ductility of mechanically alloyed NiAl. Mater. Sci. Eng., A Struct. Mater.: Prop. Microstruct. Process. 1992, 152, 160165. 42. Kimura, H. High-strength intermetallic TiAl synthesized via high-temperature crystallization of an amorphous alloy. Philos. Mag., A 1996, 73, 723737. 43. Koch, C.C. Ductility in nanostructured and ultra negrained materials: Recent evidence for optimism. J. Metastable Nanocryst. Mater. 2003, 18, 9. 44. Ebrahimi, F.; Ahmed, Z.; Morgan, K. Effect of grain size distribution on tensile properties of electrodeposited nanocrystalline nickel. MRS Symp. Proc. 2001, 634, B2.7.1B2.7.6.

2009 by Taylor & Francis Group, LLC

Soft Supramolecular

4216

Structural Nanomaterials

45. Wang, N.; Wang, Z.; Aust, K.T.; Erb, U. Room temperature creep behavior of nanocrystalline nickel produced by an electrodeposition technique. Mater. Sci. Eng., A Struct. Mater.: Prop. Microstruct. Process. 1997, 237, 150158. 46. Ke, M.; Hackney, S.A.; Milligan, W.W.; Aifantis, E.C. Observation and measurement of grain rotation and plastic strain in nanostructured metal thin lm. Nanostruct. Mater. 1995, 5, 689697. 47. Armstrong, R.W. The (cleavage) strength of pre-cracked polycrystals. Eng. Fract. Mech. 1987, 28, 529538. 48. Farkas, D.; Van Swygenhoven, H.; Derlet, P.M. Intergranular fracture in nanocrystalline metals. Phys. Rev., B 2002, 6606, 101 49. Xiao, C.H.; Mirshams, R.A.; Whang, S.H.; Yin, W.M. Tensile behavior and fracture in nickel and carbon doped nanocrystalline nickel. Mater. Sci. Eng., A Struct. Mater.: Prop. Microstruct. Process. 2001, 301, 3543. 50. Gan, Y.; Zhou, B.L. Effect of grain size on the fracture toughness of nanocrystalline FeMoSiB. Scr. Mater. 2001, 45, 625630. 51. Morris, M.A.; Leboeuf, M. Grain-size renement of gamma-TiAl alloysEffect on mechanical properties. Mater. Sci. Eng. 1997, 224, 111. 52. Varin, R.A.; Zbroniec, L.; Czujko, T.; Song, Y.K. Fracture toughness of intermetallic compacts consolidated from nanocrystalline powders. Mater. Sci. Eng., A Struct. Mater.: Prop. Microstruct. Process. 2001, 300, 111. 53. Mirshams, R.A.; Mao, C.H.; Whang, S.H.; Yin, W.M. R-curve characterization of the fracture toughness of nanocrystalline nickel thin sheets. Mater. Sci. Eng., A Struct. Mater.: Prop. Microstruct. Process. 2001, 315, 2127. 54. Whitney, A.B.; Sanders, P.G.; Weertman, J.R.; Eastman, J.A. Fatigue of nanocrystalline copper. Scr. Metall. Mater. 1995, 33, 20252030. 55. Agnew, S.R.; Weertman, J.R. Cyclic softening of ultrane grain copper. Mater. Sci. Eng., A Struct. Mater.: Prop. Microstruct. Process. 1998, 244, 145153. 56. Agnew, S.R.; Vinogradov, A.Y.; Hashimoto, S.; Weertman, J.R. Overview of fatigue performance of Cu processed by severe plastic deformation. J. Electron. Mater. 1999, 28, 10381044. 57. Yan, D.S.; Zheng, Y.S.; Gao, L.; Zhu, C.F.; Wang, X.W.; Bai, C.L.; Xu, L.; Li, M.Q. Localized superplastic deformation of nanocrystalline 3Y-TZP ceramics under cyclic tensile fatigue at ambient temperature. J. Mater. Sci. 1998, 33, 27192723. 58. Ng, C.B.; Schadler, L.S.; Siegel, R.W. Synthesis and mechanical properties of TiO2epoxy nanocomposites. Nanostruct. Mater. 1999, 12, 507510. 59. Mohamed, F.A.; Li, Y. Creep and superplasticity in nanocrystalline materials: current understanding and future prospects. Mater. Sci. Eng. 2001, 298, 115. 60. Cai, B.; Kong, Q.P.; Lu, K.; Lu, L. Low temperature creep of nanocrystalline pure copper. Mater. Sci. Eng. 2000, 286, 188192. 61. Wang, N.; Wang, Z.R.; Aust, K.T.; Erb, U. Room temperature creep behavior of nanocrystalline nickel

62.

63.

64. 65.

66.

67.

68.

69.

70.

71.

72.

73.

produced by an electrodeposition technique. Mater. Sci. Eng. 1997, 237, 150158. Yin, W.M.; Whang, S.H.; Mirshams, R.; Xiao, C.H. Creep behavior of nanocrystalline nickel at 290 and 373 K. Mater. Sci. Eng. 2001, 301, 1822. Sanders, P.; Rittner, G.M.; Kiedaisch, E.; Weertman, J.R.; Kung, H.; Lu, Y.C. Creep of nanocrystalline Cu, Pd, and AlZr. Nanostruct. Mater. 1997, 9, 433440. Xiao, M.L.; Kong, Q.P. Creep behaviour of a nanocrystalline FeBSi alloy. Scr. Mater. 1997, 36, 299303. Kong, Q.P.; Cai, B.; Xiao, M.L. High temperature creep of two nanocrystalline alloys. Mater. Sci. Eng. 1997, 234, 9193. Roddy, M.J.; Cannon, W.R.; Skandan, G.; Hahn, H. Creep behavior of nanocrystalline monoclinic ZrO2. J. Eur. Ceram. Soc. 2002, 22, 26572662. Gutierrez-Mora, F.; Dominguez-Rodriguez, A.; Jimenez-Melendo, M.; Chaim, R.; Hefetz, M. Creep of nanocrystalline Y-SZP ceramics. Nanostruct. Mater. 1999, 11, 531537. McFadden, S.X.; Zhiyaev, A.P.; Mishra, R.S.; Mukherjee, A.K. Observations of low-temperature superplasticity in electrodeposited ultrane grained nickel. Mater. Lett. 2000, 45, 345349. Lu, L.; Sui, M.L.; Lu, K. Superplasticity extensibility and deformation mechanism of a nanocrystalline copper sample. Adv. Eng. Mater. 2001, 3, 663667. McFadden, S.X.; Valiev, R.Z.; Mukherjee, A.K. Superplasticity in nanocrystalline Ni3Al. Mater. Sci. Eng., A Struct. Mater.: Prop. Microstruct. Process. 2001, 319, 849853. Mayo, M.J. High and low temperature superplasticity in nanocrystalline materials. Nanostruct. Mater. 1997, 9, 717726. Inoue, A.; Kimura, H.M.; Kita, K. New Horizons in Quasicrystals; Goldman, A.I., Sordelet, D.J., Thiel, P.A., Dubois, J.M., Eds.; World Scientic: Singapore, 1997; 256 pp. Uoya, A.; Shibata, T.; Higashi, K.; Inoue, A.; Masumoto, T. Superplastic deformation characteristics and constitution equation in rapidly solidied MgAl Ga alloy. J. Mater. Res. 1996, 11, 27312737.

Soft Supramolecular

BIBLIOGRAPHY
1. Ebrahimi, F.; Zhai, Q.; Kong, D. Deformation and fracture of electrodeposited copper. Scr. Mater. 1998, 39, 315321. 2. Karimpoor, A.A.; Erb, U.; Aust, K.T.; Wang, Z.; Palumbo, G. Tensile properties of bulk nanocrystalline hexagonal Co electrodeposits. Mater. Sci. Forum 2002, 386388, 415420. 3. Valiev, R.Z.; Alexandrov, I.V.; Zhu, Y.Y.; Lowe, T.C. Paradox of strength and ductility in metals processed by severe plastic deformation. J. Mater. Res. 2002, 17, 58. 4. Wang, Y.M.; Ma, E.; Chen, M.W. Enhanced tensile ductility and toughness in nanostructured Cu. Appl. Phys. Lett. 2002, 80, 23952397.

2009 by Taylor & Francis Group, LLC

Você também pode gostar