Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Dislocation Configurations in Nanocrystalline FeMo Sintered Components

Metallurgical and Materials Transactions A, 2010
...Read more
Dislocation Configurations in Nanocrystalline FeMo Sintered Components PAOLO SCARDI, MIRCO D’INCAU, MATTEO LEONI, and ALESSANDRO FAIS Net-shape compaction of a nanocrystalline ball-milled commercial Fe-1.5 wt pct Mo powder was done via spark plasma sintering (SPS) and capacitor discharge sintering (CDS). A detailed microstructure analysis, performed by X-ray diffraction–whole powder pattern modeling (XRD-WPPM), shows that CDS, owing to the faster sintering conditions, retains a much finer and more uniform microstructure with dislocations uniformly distributed inside the nanocrys- talline grains. Conversely, SPS causes dislocations to pile up and extensive grain growth to occur, especially when high sintering temperatures are employed. DOI: 10.1007/s11661-009-9987-x Ó The Minerals, Metals & Materials Society and ASM International 2009 I. INTRODUCTION ONE of the challenges in metal powder sintering technology is how to transfer the peculiar properties of nanocrystalline materials to compacted components. This has become of special interest in recent years, in which extensive research has been devoted to the pro- duction of nanocrystalline, heavily deformed metal powders and their compaction by advanced sintering techniques. [14] The best processing conditions—highest (i.e. theoretical) density in the sintered component, while preserving as much as possible the fine grain size and high lattice defect density of the starting nanopowder—can hardly be achieved by conventional sintering. [2,5] Within the large family of field-assisted sintering techniques (FAST) or electric current activated sintering (ECAS), spark plasma sintering (SPS) has been studied in some detail. [6] Despite the good results reported in the literature, [7,8] sintering time is still on the order of minutes and the process is controlled in temperature, with an electronic feedback, thus allowing for a certain degree of thermal equilibrium to be reached. Despite undisputed improvements with respect to traditional techniques, the relatively high sintering temperatures required to reach high densities tend to anneal the microstructure, thus losing at least part of the properties of the starting powders. [1,2] Although suitable additions may be used to stabilize the nanostructure and to reduce the damaging effect of high temperature/long process time, [9] substantial improvements should probably involve a different approach. Key factors such as rapidity and effectiveness of the sintering process seem to be peculiar to another family of FASTs, known as single-pulse techniques, or electrodis- charge sintering (EDS). Electrodischarge sintering requires a single, short pulse of electromagnetic energy to sinter the powder compacts while applying mechanical pressure. [10] A recently developed EDS technique, named capacitor discharge sintering (CDS), is based on a single pulse of low voltage, high current electromagnetic energy, delivered in a short time (tens of milliseconds) to powders in a conducting mold previously loaded with mechanical pressure. Capacitor discharge sintering, com- pared to traditional EDS, [1114] employs low voltages (from 5 to 30 V) and is composed of two mutually coupled freely oscillating circuits with a solid-state switch on the primary circuit analogous to those used in capacitor discharge welding. The output voltage closely resembles another EDS named high energy-high rate (HEHR), [15,16] but the sintering/discharge time is only 20 to 30 ms in CDS compared to 3 seconds in HEHR. Preliminary studies have already proved the effectiveness of CDS in compacting several metal powders. [1719] The present article provides a comparison between SPS and CDS of a typical metal system, a Mo-stabilized iron alloy. Since the main differences between SPS and CDS components are in the microstructure, namely, the crystalline domain size and lattice defects density, an advanced X-ray diffraction (XRD) line profile analysis (LPA) method is used. Whole powder pattern modeling (WPPM) [2023] provides information on crystalline domain size and dispersion, and on the density and distribution of dislocations in the ferritic iron phase. Correlations with two basic macroscopic properties, density and microhardness, are discussed. II. EXPERIMENTAL Fe-1.5 wt pct Mo powders were ground at increasing times, for 8, 32, 64, 96, and 128 hours (Reference 24 for PAOLO SCARDI, Professor, and MIRCO D’INCAU and MATTEO LEONI, Doctors, are with the Department of Materials Engineering and Industrial Technologies, University of Trento, 38100 Trento, Italy. Contact e-mail: paolo.scardi@unitn.it ALESSANDRO FAIS, Doctor, is with the Department of Materials Science and Chemical Engineering, Turin Polytechnic, 10129 Torino, Italy. This article is based on a presentation given in the symposium ‘‘Neutron and X-Ray Studies of Advanced Materials,’’ which occurred February 15–19, 2009, during the TMS Annual Meeting in San Francisco, CA, under the auspices of TMS, TMS Structural Materials Division, TMS/ASM Mechanical Behavior of Materials Committee, TMS: Advanced Characterization, Testing, and Simulation Commit- tee, and TMS: Titanium Committee. METALLURGICAL AND MATERIALS TRANSACTIONS A
details). Milled powders were capacitor-discharge sin- tered into small disks of 2 g inside a graphite mold with metallic plungers/electrodes, which applied a nominal pressure of 250 MPa and with different specific energy inputs (SEIs) from 1.3 to 3.4 kJ/g. The SPS of the different powders was carried out at 600 °C, 650 °C, 700 °C, 750 °C, and 800 °C, with a heating rate to temperature of 50 °C/min (respectively 13, 14, 15, 16, and 17 minutes) in graphite punches and dies and a nominal pressure of 76 MPa. All samples where disks with a radius of 10 mm. The densities were evaluated with Archimede’s principle, while Vickers micro- hardness was obtained, after polishing, from the mean value of 8 to 12 hardness measurements done in two different laboratories with a force of 300 and of 50 gf and a hold time of 15 seconds. [17,25] To avoid including contributions from polishing, before XRD measurements, all specimens were chemi- cally thinned using a solution of nitric acid (4 pct)- orthophosphoric acid (80 pct). The XRD patterns were collected at each layer removal stage (approximately 20 lm), until no changes were observed after further removal. III. WHOLE POWDER PATTERN MODELING The observed (h) diffraction profile can be thought of as a convolution of instrumental (g) and specimen- related (f) profile components. [26] Fourier analysis is especially useful to deal with this case, as the complex problem of handling a convolution integral can be turned into the relatively simpler one of a Fourier integral: [20,21] Is; s hkl ð Þ/ X hkl w hkl Z C hkl L ð Þ exp 2piL s s hkl ð Þ ½ dL ½1 where s is the reciprocal space variable, and s hkl is s in Bragg condition, including possible shifts due to lattice defects (e.g., faulting). Since the latter may dif- ferently affect the various equivalent points in recipro- cal space, the observed powder diffraction profile results from the sum of Fourier integrals for all hkl equivalent combinations, with a weight w hkl . C hkl is the product of the Fourier transforms (FTs) of each profile component: C hkl ¼ T IP A S A F A D A APB ... ½2 where T IP groups all contributions to the g profile, and the other terms (A) refer to f components. The latter are known—in many cases, in closed analytical form—for several contributing effects, such as domain shape and size distributions (A S ), dislocations (A D ), planar defects (A F ), antiphase boundaries (A APB ), etc. Virtually any other contributing effect can be included by adding (multiplying in Eq. [2]) the corresponding FT. [2023] The g component can be handled by a fundamental parameters approach [27] or, more simply, through a parameterization of the experimental pattern of a suitable line profile standard (e.g., LaB 6 SRM 660a distributed by NIST; e.g., see Reference 23 for details), so it is in any case a known quantity. In the present case of study, the phase of interest is a bcc metal (a-Fe) for which we can expect the main contribution to the f profile to come from the fine size of crystalline domains (A S ) and dislocations (A D ). For the ‘‘size’’ effect, we can assume that the approximately equiaxial domains can be represented by spherical crystallites whose diameters follow a lognormal distribution with mean l and variance r. [2022] Concerning dislocations, the Krivoglaz–Wilkens theory [2830] is still the most appro- priate one to model the effect of a system of dislocations with an average density q (assuming equally a populated 1 10 111 h i slip system) and an effective outer cut-off radius R e . [29,30] Expressions for the FT can be evaluated both for screw and for edge dislocations, for which it is possible to calculate the corresponding average ‘‘‘visi- bility’’ effect, or contrast factor, [2832] which represents the anisotropic line broadening effect of the line defect and of the crystalline medium. Diffraction peak profiles can then be modeled by adjusting (by nonlinear least squares) just a few micro- structural parameters, which in this case are q, R e , the dislocation character (f E , edge/screw fraction), l, and r of the domain size distribution. Additional parameters are used to reproduce a background, unit cell param- eters, and some common aberrations of the powder geometry. Peak intensities are usually considered as free parameters, but they can also be constrained according to a structural model, as in the Rietveld method. [33] If this is the case, WPPM also provides for a quantitative phase analysis, of interest in phase mixtures. For further details and a survey of WPPM applica- tions, see Reference 23 and the references therein. IV. RESULTS AND DISCUSSION Figure 1 shows an example of WPPM for a FeMo sintered component. In this specific case (CDS 64 hour), Fig. 1—WPPM results for CDS 64-h sintered component: experi- mental data (dot), model (line), and their difference (residual, line below). The inset shows a logarithmic plot, with indications of the most intense lines of the c-Fe minority fraction (arrows). METALLURGICAL AND MATERIALS TRANSACTIONS A
Dislocation Configurations in Nanocrystalline FeMo Sintered Components PAOLO SCARDI, MIRCO D’INCAU, MATTEO LEONI, and ALESSANDRO FAIS Net-shape compaction of a nanocrystalline ball-milled commercial Fe-1.5 wt pct Mo powder was done via spark plasma sintering (SPS) and capacitor discharge sintering (CDS). A detailed microstructure analysis, performed by X-ray diffraction–whole powder pattern modeling (XRD-WPPM), shows that CDS, owing to the faster sintering conditions, retains a much finer and more uniform microstructure with dislocations uniformly distributed inside the nanocrystalline grains. Conversely, SPS causes dislocations to pile up and extensive grain growth to occur, especially when high sintering temperatures are employed. DOI: 10.1007/s11661-009-9987-x Ó The Minerals, Metals & Materials Society and ASM International 2009 I. INTRODUCTION ONE of the challenges in metal powder sintering technology is how to transfer the peculiar properties of nanocrystalline materials to compacted components. This has become of special interest in recent years, in which extensive research has been devoted to the production of nanocrystalline, heavily deformed metal powders and their compaction by advanced sintering techniques.[1–4] The best processing conditions—highest (i.e. theoretical) density in the sintered component, while preserving as much as possible the fine grain size and high lattice defect density of the starting nanopowder—can hardly be achieved by conventional sintering.[2,5] Within the large family of field-assisted sintering techniques (FAST) or electric current activated sintering (ECAS), spark plasma sintering (SPS) has been studied in some detail.[6] Despite the good results reported in the literature,[7,8] sintering time is still on the order of minutes and the process is controlled in temperature, with an electronic feedback, thus allowing for a certain degree of thermal equilibrium to be reached. Despite undisputed improvements with respect to traditional techniques, the relatively high sintering temperatures required to reach high densities tend to anneal the microstructure, thus losing at least part of the properties of the starting powders.[1,2] Although suitable additions may be used to stabilize the nanostructure and to reduce the damaging effect of high temperature/long process PAOLO SCARDI, Professor, and MIRCO D’INCAU and MATTEO LEONI, Doctors, are with the Department of Materials Engineering and Industrial Technologies, University of Trento, 38100 Trento, Italy. Contact e-mail: paolo.scardi@unitn.it ALESSANDRO FAIS, Doctor, is with the Department of Materials Science and Chemical Engineering, Turin Polytechnic, 10129 Torino, Italy. This article is based on a presentation given in the symposium ‘‘Neutron and X-Ray Studies of Advanced Materials,’’ which occurred February 15–19, 2009, during the TMS Annual Meeting in San Francisco, CA, under the auspices of TMS, TMS Structural Materials Division, TMS/ASM Mechanical Behavior of Materials Committee, TMS: Advanced Characterization, Testing, and Simulation Committee, and TMS: Titanium Committee. METALLURGICAL AND MATERIALS TRANSACTIONS A time,[9] substantial improvements should probably involve a different approach. Key factors such as rapidity and effectiveness of the sintering process seem to be peculiar to another family of FASTs, known as single-pulse techniques, or electrodischarge sintering (EDS). Electrodischarge sintering requires a single, short pulse of electromagnetic energy to sinter the powder compacts while applying mechanical pressure.[10] A recently developed EDS technique, named capacitor discharge sintering (CDS), is based on a single pulse of low voltage, high current electromagnetic energy, delivered in a short time (tens of milliseconds) to powders in a conducting mold previously loaded with mechanical pressure. Capacitor discharge sintering, compared to traditional EDS,[11–14] employs low voltages (from 5 to 30 V) and is composed of two mutually coupled freely oscillating circuits with a solid-state switch on the primary circuit analogous to those used in capacitor discharge welding. The output voltage closely resembles another EDS named high energy-high rate (HEHR),[15,16] but the sintering/discharge time is only 20 to 30 ms in CDS compared to 3 seconds in HEHR. Preliminary studies have already proved the effectiveness of CDS in compacting several metal powders.[17–19] The present article provides a comparison between SPS and CDS of a typical metal system, a Mo-stabilized iron alloy. Since the main differences between SPS and CDS components are in the microstructure, namely, the crystalline domain size and lattice defects density, an advanced X-ray diffraction (XRD) line profile analysis (LPA) method is used. Whole powder pattern modeling (WPPM)[20–23] provides information on crystalline domain size and dispersion, and on the density and distribution of dislocations in the ferritic iron phase. Correlations with two basic macroscopic properties, density and microhardness, are discussed. II. EXPERIMENTAL Fe-1.5 wt pct Mo powders were ground at increasing times, for 8, 32, 64, 96, and 128 hours (Reference 24 for details). Milled powders were capacitor-discharge sintered into small disks of 2 g inside a graphite mold with metallic plungers/electrodes, which applied a nominal pressure of 250 MPa and with different specific energy inputs (SEIs) from 1.3 to 3.4 kJ/g. The SPS of the different powders was carried out at 600 °C, 650 °C, 700 °C, 750 °C, and 800 °C, with a heating rate to temperature of 50 °C/min (respectively 13, 14, 15, 16, and 17 minutes) in graphite punches and dies and a nominal pressure of 76 MPa. All samples where disks with a radius of 10 mm. The densities were evaluated with Archimede’s principle, while Vickers microhardness was obtained, after polishing, from the mean value of 8 to 12 hardness measurements done in two different laboratories with a force of 300 and of 50 gf and a hold time of 15 seconds.[17,25] To avoid including contributions from polishing, before XRD measurements, all specimens were chemically thinned using a solution of nitric acid (4 pct)orthophosphoric acid (80 pct). The XRD patterns were collected at each layer removal stage (approximately 20 lm), until no changes were observed after further removal. III. WHOLE POWDER PATTERN MODELING The observed (h) diffraction profile can be thought of as a convolution of instrumental (g) and specimenrelated (f) profile components.[26] Fourier analysis is especially useful to deal with this case, as the complex problem of handling a convolution integral can be turned into the relatively simpler one of a Fourier integral:[20,21] Z X Iðs; shkl Þ / whkl Chkl ðLÞ exp½2piLðs  shkl ÞdL ½1 distributed by NIST; e.g., see Reference 23 for details), so it is in any case a known quantity. In the present case of study, the phase of interest is a bcc metal (a-Fe) for which we can expect the main contribution to the f profile to come from the fine size of crystalline domains (AS) and dislocations (AD). For the ‘‘size’’ effect, we can assume that the approximately equiaxial domains can be represented by spherical crystallites whose diameters follow a lognormal distribution with mean l and variance r.[20–22] Concerning dislocations, the Krivoglaz–Wilkens theory[28–30] is still the most appropriate one to model the effect of a system of dislocations with an average density q (assuming equally a populated   110 h111i slip system) and an effective outer cut-off radius Re.[29,30] Expressions for the FT can be evaluated both for screw and for edge dislocations, for which it is possible to calculate the corresponding average ‘‘‘visibility’’ effect, or contrast factor,[28–32] which represents the anisotropic line broadening effect of the line defect and of the crystalline medium. Diffraction peak profiles can then be modeled by adjusting (by nonlinear least squares) just a few microstructural parameters, which in this case are q, Re, the dislocation character (fE, edge/screw fraction), l, and r of the domain size distribution. Additional parameters are used to reproduce a background, unit cell parameters, and some common aberrations of the powder geometry. Peak intensities are usually considered as free parameters, but they can also be constrained according to a structural model, as in the Rietveld method.[33] If this is the case, WPPM also provides for a quantitative phase analysis, of interest in phase mixtures. For further details and a survey of WPPM applications, see Reference 23 and the references therein. IV. RESULTS AND DISCUSSION hkl where s is the reciprocal space variable, and shkl is s in Bragg condition, including possible shifts due to lattice defects (e.g., faulting). Since the latter may differently affect the various equivalent points in reciprocal space, the observed powder diffraction profile results from the sum of Fourier integrals for all hkl equivalent combinations, with a weight whkl. Chkl is the product of the Fourier transforms (FTs) of each profile component: Chkl ¼ TIP AS AF AD AAPB  . . . Figure 1 shows an example of WPPM for a FeMo sintered component. In this specific case (CDS 64 hour), ½2 where TIP groups all contributions to the g profile, and the other terms (A) refer to f components. The latter are known—in many cases, in closed analytical form—for several contributing effects, such as domain shape and size distributions (AS), dislocations (AD), planar defects (AF), antiphase boundaries (AAPB), etc. Virtually any other contributing effect can be included by adding (multiplying in Eq. [2]) the corresponding FT.[20–23] The g component can be handled by a fundamental parameters approach[27] or, more simply, through a parameterization of the experimental pattern of a suitable line profile standard (e.g., LaB6 SRM 660a Fig. 1—WPPM results for CDS 64-h sintered component: experimental data (dot), model (line), and their difference (residual, line below). The inset shows a logarithmic plot, with indications of the most intense lines of the c-Fe minority fraction (arrows). METALLURGICAL AND MATERIALS TRANSACTIONS A as in the other cases discussed subsequently, a minor fraction of c-Fe (arrows) appears together with the main a-Fe phase. Minor amounts of iron oxide (magnetite) were also found in several sintered components. As WPPM supports the analysis of all phases, allowing also for a quantification of the relative fractions, microstructural results for the main phase of interest, a-Fe, are nearly unaffected by the presence of other phases. The main results of WPPM are summarized in Figure 2, where average dislocation density ((a) and (b)) and mean domain size ((c) and (d)) are shown as a function of sintering temperature (T in °C) and SEI (in J/g), respectively, for SPS and CDS components. As a general trend, dislocation density tends to decrease with the SPS temperature, the diminution being more evident for 96- and 128-hour powders sintered above 750 °C. Correspondingly, the domain size increases, with a marked growth above 750 °C. The domain sizes for SPS at 800 °C are of several hundreds of nanometers, close to, or possibly above, values reliably accessible to XRD LPA. The overall result is that the nanostructure is preserved up to sintering temperatures of 750 °C, with the powders ball milled for 128 hours preserving a proportionally higher dislocation density/smaller domain size than the powders ground for shorter time. Above 750 °C, extensive grain growth and defect annealing takes place. Capacitor discharge sintering also causes a marked decrease in dislocation density, the difference between different powders decreasing for increasing SEI. The main difference with respect to SPS seems to be the domain size, never exceeding 100 to 150 nm, even for the highest SEI employed. According to this information, CDS seems to better preserve the nanostructure than SPS. Electron microscopy supports this hypothesis. As shown in Figure 3, CDS gives a uniform, finegrained (<1 lm) microstructure, whereas SPS clearly shows a coarser microstructure. The bimodality of the SPS grain size distribution is also evident, with scattered regions of fine grains (submicron size) separated by larger grains (5 to 10 lm). This feature of the SPS has already been observed[34,35] and can partly be controlled by varying (increasing) the pressure. As a potential drawback of the CDS, a minor but clearly measurable fraction of austenitic phase is formed. As shown in Figure 4, the amount of austenite tends to increase for increasing SEI, and the effect is more visible for the most extensively ball-milled pow- (a) (b) (c) (d) Fig. 2—WPPM results. (a) and (b) Average dislocation density and (c) and (d) mean domain size as a function of sintering temperature or SEI, respectively, for SPS or CDS components obtained from powders ball milled for 128 h (square), 96 h (circle), 64 h (rhomb), and 32 h (triangle). Lines are drawn just to drive the eye. Data points at T = 0 and SEI = 0 refer to the ball-milled powders, before sintering. METALLURGICAL AND MATERIALS TRANSACTIONS A (a) Fig. 3—ESEM micrographs of (a) SPS 96 h and (b) CDS 96 h components. (b) Fig. 5—Correlation between microhardness and density for (a) SPS and (b) CDS components: Hv 0.05 (full symbol) and Hv 0.3 (open symbol). The dashed line highlights the corner region of high microhardness/large density. Refer to the text for discussion. Fig. 4—Residual austenite in CDS components as a function of SEI. Trends refer to the four starting powders, ground for different times (32, 64, 96, and 128 h). ders, which have a higher excess energy. The SPS components are made of ferritic phase only, with traces of iron oxide, with the exception of the components made of 128-hour powder sintered at 750 °C and 800 °C, where iron oxides reach 3 pct and austenite is about 2 pct. It is interesting to integrate the information provided by XRD-LPA with two basic properties of the sintered components: density and microhardness. Data are shown in Figure 5 for (a) SPS and (b) CDS components, using two indentation loads. It is quite evident that CDS components reach higher densities and higher microhardness (upper-right corner), the latter being slightly sensitive to variations in the load. On the contrary, SPS components show a remarkable difference in microhardness, depending both on the density and on the applied load. These results are coherent with the WPPM results and with the morphological observations. Capacitor discharge sintering provides a much finer and more uniform microstructure than SPS, and the latter causes extensive annealing when high temperatures (>750 °C) are employed. Whole powder pattern modeling provides additional information, useful for a deeper understanding of the two different sintering processes. In fact, in addition to domain size and dislocation density, WPPM also METALLURGICAL AND MATERIALS TRANSACTIONS A provides the effective outer cut-off radius, which can be conveniently combined with the dislocation pffiffiffi density in the so-called Wilkens parameter, M ¼ Re q. The latter is shown in Figure 6 for SPS (open symbols) and CDS (full symbols) components, as a function of the dislocation density. It is worth noting that, in this plot, movement from right to left (i.e., for decreasing dislocation density) corresponds to components sintered at increasing SEI (CDS) and increasing T (SPS). The value of M tends to decrease after the SPS process, whereas it increases markedly after CDS. According to the theory of the line broadening effects of dislocations,[28–30] M values close to (and below) unity are observed when dislocations are strongly interacting, as in dipole configurations or in dislocation walls: this is the case for the SPS components, as the result of near thermal equilibrium process conditions, and could be the result of a tendency of dislocations to pileup and reach lower energy configurations, as in all temperature-activated sintering processes. A Wilkens parameter well above unity, instead, may result from randomly dispersed dislocations. This is the case observed in CDS components, where the fast sintering conditions push the system very far from equilibrium, with the dislocations remaining ‘‘frozen’’ in a random distribution inside the crystalline domains. It is therefore concluded that, although the average dislocation density after SPS and CDS is similar, the distribution of the dislocations is probably different, and might contribute to an explanation of the higher microhardness observed for CDS components with respect to SPS components with similar density and mean domain size. This observation confirms analogous results obtained in a sensibly different system such as pure copper[19] and is still the object of ongoing research activity. pffiffiffi Fig. 6—Correlation between the Wilkens parameter (M ¼ Re q) and average dislocation density (q) for CDS (full symbols) and SPS (open symbols) components obtained from nanocrystalline FeMo powders ground for different times: 128 h (square), 96 h (circle), and 64 h (triangle). Full symbols in gray on the right side (after the x-axis break) refer to the starting powders. Lines are drawn just to highlight the overall trend. METALLURGICAL AND MATERIALS TRANSACTIONS A As a final remark, this article shows that XRD LPA, following a WPPM approach, provides detailed information on the microstructure of sintered components from nanocrystalline metal powders. Information is given on the domain size as well as on the density and possible arrangement of line defects. The same approach can be used for any type of crystalline material, as the applicability of the model has recently been extended to systems with any crystal symmetry,[31,32] and can also account for other microstructural features and lattice defects that produce measurable effects on the diffraction line profile.[23] ACKNOWLEDGMENT The authors thank Dr. M. Zadra for the precious support in the preparation of SPS components. REFERENCES 1. H.J. Fecht: in Nanostructured Materials and Composites Prepared by Solid State Processing, C.C. Koch, ed., Noyes Publications, Norwick, NY, 2002, pp. 73–113. 2. J.R. Groza: in Nanostructured Materials and Composites Prepared by Solid State Processing, C.C. Koch, ed., Noyes Publications, Norwick, NY, 2002, pp. 115–78. 3. C. Suryanarayana: Prog. Mater. Sci., 2001, vol. 46, pp. 1–184. 4. R.Z. Valiev, R.K. Islamgaliev, and I.V. Alexandrov: Prog. Mater. Sci., 2000, vol. 45, p. 103. 5. C.C. Koch: J. Mater. Sci., 2007, vol. 42, pp. 1403–14. 6. J. Rawers: Nanostruct. Mater., 1999, vol. 11 (4), pp. 513–22. 7. K.Y. Sastry, L. Froyen, J. Vleugels, O. Van der Biest, R. Schattevoy, and J. Hennicke: Rev. Adv. Mater. Sci., 2004, vol. 8 (1), pp. 27–32. 8. K.Y. Sastry, L. Froyen, J. Vleugels, O. Van der Biest, R. Schattevoy, and K. Hummert: Mater. Sci. Forum, 2006, vols. 519–521, pp. 1409–14. 9. S. Libardi, M. Leoni, L. Facchini, M. D’Incau, P. Scardi, and A. Molinari: Mater. Sci. Eng. A, 2007, vols. 445–446, pp. 244–50. 10. R. Orrù, R. Licheri, A.M. Locci, A. Cincotti, and G. Cao: Mater. Sci. Eng. R, 2009, vol. 63, pp. 127–288. 11. S. Clyens, S.T.S. Al-Hassani, and W. Johnson: Int. J. Mech. Sci., 1976, vol. 18 (1), pp. 37–44. 12. T. Alp, S.T.S. Al-Hassani, and W. Johnson: J. Eng. Mater. Technol., 1985, vol. 107, p. 109. 13. W. Johnson, S. Clyens, and S.T.S. Al-Hassani: Metallurgia Met. Form., 1976, vol. 43, p. 82. 14. M. Shakery, S.T.S. Al-Hassani, and T.J. Davis: Powder Metall. Int., 1979, vol. 11 (3), pp. 120–25. 15. D.R. Ervin, D.L. Bourell, C. Persad, and L. Rabenberg: Mater. Sci. Eng. A, 1988, vol. 102, pp. 25–30. 16. S.L. Raghunathan, C. Persad, D.L. Bourell, and H.L. Marcus: Mater. Sci. Eng. A, 1991, vol. 131, pp. 243–53. 17. A. Fais: Ph.D. Thesis, Turin Polytechnic, Turin, 2008. 18. A. Fais and G. Maizza: J. Mater. Proc. Technol., 2008, vol. 202, pp. 70–75. 19. A. Fais and P. Scardi: Z. Kristallogr. Suppl., 2008, vol. 27, pp. 37–44. 20. P. Scardi and M. Leoni: Acta Cryst. A, 2001, vol. 57, pp. 604–13. 21. P. Scardi and M. Leoni: Acta Cryst. A, 1002, vol. 58, pp. 190–200. 22. M. Leoni and P. Scardi: J. Appl. Cryst., 2004, vol. 37, pp. 629–34. 23. P. Scardi: in Powder Diffraction: Theory and Practice, R.E. Dinnebier and S.J.L. Billinge, eds., The Royal Society of Chemistry, Cambridge, United Kingdom, 2008, pp. 376–413. 24. M. D’Incau, M. Leoni, and P. Scardi: J. Mater. Res., 2007, vol. 22, pp. 1744–53. 25. M. D’Incau: Ph.D. Thesis, University of Trento, Trento, 2008. 26. H.P. Klug and L.E. Alexander: X-ray Diffraction Procedures for Polycrystalline and Amorphous Materials, Wiley, New York, NY, 1974. 27. R.W. Cheary and A. Coelho: J. Appl. Cryst., 1992, vol. 25, pp. 109–21. 28. M.A. Krivoglaz and K.P. Ryaboshapka: Fiz. Metall., 1963, vol. 15, pp. 18–28. 29. M. Wilkens: Phys. Status Solidi A, 1970, vol. 2, pp. 359–70. 30. M. Wilkens: Fundamental Aspects of Dislocation Theory, J.A. Simmons, R. de Wit, and R. Bullough, eds., National Bureau of Standards (US) Special Publication No. 317, NBS, Washington, DC, vol. II, pp. 1195–1221. 31. M. Leoni, J. Martinez-Garcia, and P. Scardi: J. Appl. Cryst., 2007, vol. 40, pp. 719–24. 32. J. Martinez-Garcia, M. Leoni, and P. Scardi: Acta Cryst. A, 2009, vol. 65, pp. 109–19. 33. R.A. Young: The Rietveld Method, Oxford University Press, Oxford, United Kingdom, 1993. 34. S. Kawamura and H. Kuwano: J. Jpn. Soc. Powder Metall., 2003, vol. 50 (7), pp. 545–50. 35. S. Kawaura, H. Kuwano, Y. Takeda, S. Takahashi, and H. Kaga: J. Jpn. Soc. Powder Metall., 2003, vol. 50 (12), pp. 1052–56. METALLURGICAL AND MATERIALS TRANSACTIONS A
Keep reading this paper — and 50 million others — with a free Academia account
Used by leading Academics
Kazuki Morita
The University of Tokyo
Suseendran Jayachandran
IMEC
Estela Blaisten-Barojas
George Mason University
Kenneth Vecchio
University of California, San Diego