Sunteți pe pagina 1din 7

Materials Science and Engineering A 434 (2006) 2329

Effect of ball milling on simultaneous spark plasma synthesis and densication of TiCTiB2 composites
Antonio M. Locci a , Roberto Orr` a,b , Giacomo Cao a,b, , Zuhair A. Munir c u
Dipartimento di Ingegneria Chimica e Materiali, Centro Studi sulle Reazioni Autopropaganti (CESRA), Unit` di Ricerca del Consorzio Interuniversitario per la Scienza e Tecnologia dei Materiali (INSTM), a Universit` degli Studi di Cagliari, Piazza dArmi, 09123 Cagliari, Italy a b PROMEA Scarl, c/o Dipartimento di Fisica, Cittadella Universitaria di Monserrato, S.S. 554 bivio per Sestu, 09042 Monserrato (CA), Italy c Facility for Advanced Combustion Synthesis, Department of Chemical Engineering and Materials Science, University of California, Davis, CA 95616, USA Received 22 November 2005; received in revised form 6 June 2006; accepted 29 June 2006
a

Abstract The effect of ball milling pre-treatment of the starting Ti, B4 C and C powders mixture on the synthesis of dense TiC/TiB2 composite was investigated through the combined use of a SPEX mill and the spark plasma sintering (SPS) apparatus. When unmilled reactants were used, amorphous carbon was converted to TiC through a combustion reaction with elemental titanium, while the complete conversion is achieved through a relatively slow reaction between the residual Ti and B4 C. In contrast, powders ball milled for 24 h reacted completely under combustion regime. SPS product density was observed to increase from about 97% to 100% of the theoretical value, when the original powders were milled for 6 h, while further increase of the milling time led to a slight decrease in product density. Crystallite size of both boride and carbide phases decreased with milling down to about 50 nm after 24 h of mechanical treatment. The latter one also allowed for the formation of a material with relatively ner microstructure and more homogeneous phase distribution. 2006 Elsevier B.V. All rights reserved.
Keywords: Ball milling; Spark plasma sintering (SPS); Synthesis; Composites; Borides; Carbides

1. Introduction Ball milling (BM) is a solid-state powder process involving welding and fracturing of powder particles in a high-energy ball mill [1]. Originally developed to produce oxide-dispersion strengthened nickel and iron-base superalloys [2], BM has been shown to be capable of synthesizing a variety of equilibrium and non-equilibrium alloy phases starting from blended elemental or pre-alloyed powders. The non-equilibrium phases synthesized include solid solutions, metastable crystalline and quasi-crystalline phases, nanostructures and amorphous alloys [1]. In addition, it has been found that BM may increase powders reactivity [3,4]. This feature is associated to reactants interface formation, increase of internal and surface energy as well as

Corresponding author. Tel.: +39 070 6755058; fax: +39 070 6755057. E-mail address: cao@visnu.dicm.unica.it (G. Cao).

surface area, which all contribute to the so-called mechanical activation (MA) of powders [1,5]. In many cases, powders produced by taking advantage of high-energy mechanical milling are used directly as obtained for achieving desirable functions in several application elds such as electronics, cosmetics, optics, coatings and paints fabrications [6,7]. On the other hand, bulk nanostructured materials with tailored mechanical, physical, and chemical properties are needed for structural applications or as components in magnetic and electrical devices. These materials have been, for example, prepared by consolidation of mechanically activated or ball-milled powders [8]. Several techniques, such as hot-pressing [9,10], shock consolidation [1113], sintering with the application of ac currents [1416] or pulsed electric current [1721], have been employed for the sintering/densication step. In particular, when nanometric powders obtained by mechanical milling need to be densied in order to produce bulk bodies with desirable and superior properties, it can be very challenging to identify the

0921-5093/$ see front matter 2006 Elsevier B.V. All rights reserved. doi:10.1016/j.msea.2006.06.131

24

A.M. Locci et al. / Materials Science and Engineering A 434 (2006) 2329

appropriate conditions during the consolidation process which are able to preserve the nanostructure. The above considerations are particularly valid for producing dense nanostructured refractory materials such as borides and carbides of transition metals. These materials have been recognized as candidate for advanced structural applications due to their exceptional hardness and stability at high temperature. Furthermore, the use of these ceramics in composites offers the advantages of enhanced fracture toughness [22,23]. However, because of their extremely high melting points, consolidation of these materials requires high temperatures and long holding times. These conditions adversely affect both microstructure, i.e. grain growth accompanies the consolidation, and cost effectiveness of the production process. Therefore, intensive research is needed for developing and optimizing processes aimed to the production of nanostructured dense refractory material, specically when the approach consisting of the combination of mechanical milling and subsequent consolidation is adopted. Recently, with the aim of synthesizing dense nanostructured TiC/TiB2 composites, the mechanical activation of elemental (Ti, C, and B) reactants followed by the eld-activated pressureassisted synthesis (FAPAS), which combines high ac current with uniaxial pressure, has been proposed [15]. Along these lines, in the present investigation the simultaneous synthesis and densication of TiC/TiB2 using the spark plasma sintering (SPS) apparatus was investigated starting from ball milled reactants (Ti, B4 C and C). While considering the same ceramic composite, i.e. TiC/TiB2 with a molar ratio equal to 1, as that one investigated by Lee et al. [15], it is worth noting that in this work the elemental boron is replaced by the relatively cheaper boron carbide. In addition, the SPS apparatus, which is characterized by a pulsed electric current instead of the ac current as in FAPAS, is employed in the proposed study. Specically, the effect of the milling time on the characteristics of powder reactants, SPS process dynamics, reactants conversion, relative density, and crystallite sizes of both ceramic constituents of the nal product, will be systematically investigated. The results are compared with those recently reported in the literature [24], which were obtained by the same method but on unmilled powders. 2. Experimental materials and methods Characteristics of the starting powders used in the present investigation are reported in Table 1. Powders were milled together (co-milled) in a stoichiometric ratio corresponding to
Table 1 Starting powders used in the present investigation Reactant Titanium (Ti) Boron carbide (B4 C) Carbon black (C) Vendor SigmaAldrich (366994) Alfa Aesar (40504) Alfa Aesar (39724) Particle size 325 mesh 17 m 0.042 m (average) Purity (%) 99.98 99.4 99.9+

the following reaction: 4Ti + B4 C + C 2TiC + 2TiB2 (1)

Mechanical activation experiments were conducted in a SPEX 8000 shaker mill (SPEX CertPrep, USA) by using stainless steel jars having an internal volume equal to about 65.8 cm3 (internal diameter, 3.8 cm; internal height, 5.8 cm). The jars were loaded with about 8 g of powders mixture and one stainless steel ball (weight, 8 g; diameter, 1 cm) so that the charge ratio, CR (ball to powders mass ratio) equals to about 1. In order to minimize oxidation, all powder handling and loading were performed inside a glove box lled with argon. The jars were then sealed and transferred to the mill. After each run, powders were removed and vials and balls were cleaned by using a slurry of silica and acetone with the aim to eliminate all residual powders of previous milling experiments. The range of ball milling time investigated, hereafter indicated as tBM , was 024 h. A spark plasma sintering (SPS) system (mod. 515S, Sumitomo Coal Mining Co. Ltd., Japan) was used to synthesize dense TiC/TiB2 ceramic composites. Details related to SPS apparatus are reported elsewhere [24]. The adopted pulse cycle consisted of 12 pulses on and 2 pulses off, for a total pulse cycle duration equal to about 46.2 ms, with the characteristic time of a single pulse equal to about 3.3 ms. About 4 g of mechanically activated powder mixture was rst cold-compacted inside a graphite die. With the aim to minimize heat losses by thermal radiation, the die was covered with a 2 mm thick layer of graphite felt (Atal s.r.l., Italy). The die was then placed inside the reaction chamber of the SPS apparatus and two cylindrical graphite blocks (diameter, 80 and 30 mm and both 40 mm high) were placed between the upper plunger and the upper electrode, as well as the lower plunger and the lower electrode. The blocks were made of AT101 graphite (Atal s.r.l., Italy) while the die and plungers were all made of E-940 graphite, supplied by Electrodes Inc. (USA). Additional experimental details can be found elsewhere [24]. The experiment is initiated with the application of a constant value of the integral mean electric current (I) equal to 1100 A and a mechanical pressure (P) of 20 MPa for 4 min (tSPS ). Temperatures were measured during synthesis by pyrometer Ircon Mirage OR 15-990 (Ircon, USA) and C-type thermocouple (Omega Engineering Inc., USA), which was inserted inside a small hole in the lateral surface of the graphite die. Temperature, applied current, voltage, mechanical load, and the vertical displacement of the lower electrode were measured in real time and recorded. The latter parameter can be regarded as the degree of powder compact densication. However, also thermal expansion of the sample as well as that of both electrodes, graphite blocks, spacers and plungers, contribute to the variation of this parameter. Such a contribution has been evaluated by performing a blank test, which consists in the application of same SPS conditions (current, mechanical load, time, etc.) to the ensemble die/plungers when powder compact is not present. It is worth noting that following this procedure, thermal expansion of processing powders is not considered. Moreover, it has

A.M. Locci et al. / Materials Science and Engineering A 434 (2006) 2329

25

been experimentally veried that temperature time-prole did not signicantly change in absence of powder compact. Thus, by subtracting the resulting displacement from that recorded in presence of powder sample, the real sample shrinkage () is obtained. After the synthesis process, the sample was rst allowed to cool and then removed from the die. The relative density of the product was determined by the Archimedes method using distilled water as wetting liquid. Phase identication and crystallite size determination of both milled powders and SPS products were made by using XRD analysis (mod. PW1830, Philips Analytical B.V., The Netherlands) with Cu K radiation and Ni lter. In particular, the analysis of SPS samples was performed on powdered sample in diffraction angle (2) range 20120 with a scan step and a scan time equal to 0.01 and 10 s, respectively. Crystallite size was determined by means of the HalderWagner method [25] from the line broadening of XRD peaks. The microstructure of end products was examined by scanning electron microscopy (SEM) (mod. S4000, Hitachi, Japan) and local phase composition was determined by energy dispersive X-rays spectroscopy (EDXS). Indentation method using a Zwick 3212 Hardness tester machine (Zwick & Co. GmbH, Germany) was employed to determine Vickers hardness of the SPS obtained products. The applied load was equal to 196.2 N while the dwell time was 18 s. Based on the Vickers indentation and by assuming a Palmqvist cracks geometry, fracture toughness (KIC ) was also estimated through the following formula [26]: KIC = 0.0319 P al1/2 (2)

Fig. 1. XRD patterns of blended 4Ti + B4 C + C powders as a function of milling time.

tional phases as a consequence of the milling. The TiC content clearly increased as tBM was increased up to 24 h. The apparent formation during milling of TiC before TiB2 is consistent with the higher diffusivity of carbon into Ti as compared to that of boron [27]. On the basis of these results, it is clear that within the experimental conditions considered in this work, no complete conversion of the starting reactants to the desired composite takes place during milling. This result was, on the other hand, achieved by Lee et al. [15] after 8 h milling of a Ti, B and C powders mixture by using a Fritsch planetary mill and a signicantly higher charge ratio (20). 3.2. SPS process dynamics and reactive system behavior The milled powders were processed in the SPS apparatus under optimized operating conditions (I = 1100 A, P = 20 MPa, tSPS = 4 min), which arise from a previous investigation of the simultaneous synthesis and densication of the composite TiC/TiB2 by SPS of commercial, unmilled powders [24]. Specically, under these conditions the complete conversion of reactants to the desired TiC/TiB2 composite with about 98% density was achieved. In Fig. 2, the XRD pattern of SPS product obtained from powders milled for 24 h is reported. It can be seen that the nal product consists of the desired TiC and TiB2 phases only. We now consider the SPS process dynamics when unmilled or co-milled powders were used. The sample shrinkage and temperature time proles recorded during the process for the case when I = 1100 A, P = 20 MPa and tSPS = 4 min, and corresponding to unmilled (tBM = 0 h) and milled (tBM = 24 h) powders, are shown in Fig. 3. First of all, it is clearly seen that milling does not have a marked effect on the temperature evolution during the process. On the other hand, the system behavior is signicantly different with respect to shrinkage () variation with time. In particular, the shrinkage prole related to the unmilled powders shows no signicant changes during the rst 15 s. An approximately linear

where P is the applied load, a the mean indentation half-diagonal length and l the crack length. 3. Results and discussion 3.1. Powder mixture Fig. 1 shows the variation of the XRD patterns of the powder mixture with milling time, tBM , for the conditions indicated in the Section 2. It should be noted that, since carbon was present in amorphous form, only Ti and B4 C were identied among the reactants by this analysis. With increasing milling time, the initially sharp diffraction peaks become broader and their absolute intensity decreases. This is due to powder crystallite size renement and internal strain increase induced by the mechanical treatment. In addition, these XRD results indicate that, within the detection limit of this analysis, additional phases are not formed at tBM up to 6 h, including any possible contaminants from the milling operation. First evidence of product formation was observed when the milling time was 12 h. However, progressive grain renement and increase of internal strain make the diffraction peaks broader and less intense with consequent overlapping. Beside the starting reactants, the TiC phase only was discernable but, for the reason above, it is not possible to exclude the formation of addi-

26

A.M. Locci et al. / Materials Science and Engineering A 434 (2006) 2329

Fig. 2. XRD patterns of SPS reacted powders milled for 24 h (I = 1100 A, P = 20 MPa, tSPS = 4 min).

increase (to about 1.11.2 mm) followed by a steep increase (to 2.2 mm) takes place in the time range of 1570 s. The shrinkage then increases gradually to reach its maximum value (3.4 mm) at 240 s. A completely different behavior was observed when the starting powders were mechanically treated for 24 h. Fig. 3 clearly shows that an abrupt variation of the shrinkage up to approximately 3.4 mm takes place only after about 20 s from the beginning of the application of the pulsed electric current. The shrinkage then gradually increases up to 5.5 mm at about 190 s. It should be noted that the total variation of the sample shrinkage obtained in the case of milled powders is considerably larger than that observed in the case of unmilled powders. The evolution of the reaction synthesis given in Eq. (1) during SPS process was investigated in our previous work [24], using unmilled powders and I = 800 A. It was found that the most rapid and signicant sample shrinkage corresponded to

the carburization of Ti to form TiC, while TiB2 formation was completed subsequently. Since the current level considered in the present work is different, this result has been veried by analyzing the product obtained when unmilled powders are subjected to I = 1100 A, P = 20 MPa, for about 70 s, i.e. just few seconds after the time instant at which the rapid sample shrinkage takes place. The corresponding XRD analysis is reported in Fig. 4(a). It is seen that, a relatively high degree of conversion of starting reactants into one of the desired products, i.e. TiC, has been achieved. However, only small amounts of TiB2 have been obtained, while secondary phases such as TiB and Ti3 B4 , along with small amounts of reactants are present in the nal sample. Therefore, as reported in a previous work [24], the pulsed electric current needs to be applied longer, i.e. up to 4 min, in order to obtain the pure product. Different results arise from the XRD analysis performed on the SPS sample of powders milled for 24 h. Specically, in Fig. 4(b) the composition of sample obtained by interrupting the application of the current after only 20 s, i.e. immediately after the abrupt shrinkage variation occurs, is reported. It is apparent that complete conversion of the starting reactants to the desired phases TiC and TiB2 is achieved in this case. Therefore, the change of represents in this case an indication of the completeness of the synthesis reaction (1). It is interesting to note that the temperature of the die corresponding to the abrupt shrinkage is very low, i.e. equal to about 150 C, when powders milled for 24 h are processed, while it is about 850 C, if unmilled reactants are used as starting materials. Typically, rapid changes of the parameter during SPS processes may be associated to melting of reactants or products, or to the occurrence of combustion reactions, because they are both accompanied by a considerable sample shrinkage in presence of mechanical load. Since the recorded temperatures corresponding to the time at which the abrupt shrinkage takes place are in both cases (unmilled and mechanically treated pow-

Fig. 3. Sample shrinkage and temperature time proles of SPS outputs for the case of unmilled and milled (tBM = 24 h) powders (I = 1100 A, P = 20 MPa, tSPS = 4 min). (The asterisks indicate the time instants at which the electric current application was interrupted for kinetic mechanism investigations).

Fig. 4. XRD patterns of the SPS samples (I = 1100 A, P = 20 MPa) when starting from: (a) unmilled powders (tSPS = 70 s); (b) milled powders (tBM = 24 h, tSPS = 30 s).

A.M. Locci et al. / Materials Science and Engineering A 434 (2006) 2329

27

ders) much lower than the melting point of Ti (1639 C), i.e. the low-melting-point-component, the presence of molten phases can be excluded. On the other hand, combustion reactions are likely to take place. It should be mentioned that an analogous behavior (fast reaction and abrupt shrinkage occurring at low temperature) was already observed for instance during the synthesis and simultaneous densication of MoSi2 by SPS, when starting from mechanically activated elemental reactants [18]. Some considerations related to the combustion behavior displayed by the system under investigation during the SPS process can be examined by taking advantage of recent results obtained when starting from unmilled powders [24]. It was suggested that the formation of the composite is due to the following reaction path: Ti + C TiC 5Ti + B4 C 4TiB + TiC 16TiB + B4 C 5Ti3 B4 + TiC 3Ti3 B4 + B4 C 8TiB2 + TiC (3) (4) (5) (6)

contained into the jar. Thus, it is expected that when relatively shorter milling times are considered, the resulting powders had not reached a uniform activation level and thus, may also behave like unmilled powders. It also worth noting that, in spite of the possibility of the occurrence of the two observed regimes in the transition zone of the milling time interval, SPS products obtained after 4 min and related to powders milled for the same time interval, not only do not display differences in their composition, but also nal densities and crystallite sizes characteristics are very close, regardless of the dynamic behavior observed during the process. This result maybe related to the fact that the sample, after the rapid shrinkage changes, is still subjected to the sintering conditions for relatively long time, during which the nal characteristics of the product are dened. 3.3. Product characteristics Fig. 5 reports the effect of milling time on the density of nal product obtained by SPS. All experiments refers to the same SPS operating conditions, i.e. I = 1000 A, P = 20 MPa, tSPS = 4 min. It may be seen that as milling time, tBM is increased, the density rst increases, reaches a maximum value and then slightly decreases. In particular, nal product density increases from about 97% to about 100% of the theoretical density (4.687 g/cm3 ) when tBM was 6 h. A further increase of the parameter tBM led to a slight decrease in product density, which becomes about 99.5% when tBM = 24 h. As milling time increases, interfacial area between reactants increases and powder size decreases. These phenomena induced by mechanical treatment enhance mass transport by diffusion and, consequently, synthesis and sintering processes. However, as seen in Fig. 1, some products are also formed as a consequence of mechanical treatment. The presence of such refractory phases, i.e. carbides and borides, at the interfaces between reactants particles, can hinder the sintering process. From this point of view, the mechanical treatment of the starting powders could lead to SPS products characterized by relatively lower density. It is also

Based on the experimental results reported in the cited work [24] and in the present investigation when considering unmilled powders, it is postulated that the direct carburization of elemental Ti by amorphous carbon, i.e. Eq. (3), is the rst step taking place during the spark plasma synthesis of TiC/TiB2 , and corresponds to the most considerable shrinkage change. This step occurs presumably as a combustion reaction because the temperature reached at this point was relatively low and, thus, the formation of molten phases can be excluded. In addition, Eqs. (4)(6) are expected to be more likely responsible of the subsequent TiB2 formation, as well as of the intermediate phases, i.e. TiB and Ti3 B4 , detected before the synthesis reaction was completed. It is worth noting that, during the occurrence of reactions (4)(6), no drastic variation of the parameter are observed (cf. Fig. 3), thus they most probably did not evolve in the combustion mode but through a gradual solidsolid diffusion mechanism. On the other hand, when reactants mechanically activated for 24 h were used, a different reacting system behavior is observed. In fact, the examination of sample shrinkage (cf. Fig. 3) in parallel with product composition (cf. Fig. 4), allows us to conclude that the entire reaction path (3)(6) evolves in this case under combustion regime, thus giving rise to complete conversion of reactants to the desired composite. Thus far, the effect of mechanical activation on the dynamic behavior of SPS process is examined only for powders milled for 24 h. For relatively shorter time intervals, i.e. 6 and 12 h, the two synthesis mechanism discussed above are both randomly observed regardless of the same MA and SPS conditions. Since this dual behavior was never displayed either when starting from unmilled powders or when powders were mechanically treated for 24 h, the existence of a transition zone, where the two kinetic mechanisms are possible, may be postulated. Although further investigations are needed to better clarify this issue, this behavior may be related to the heterogeneous nature of the ball milling treatment. Milling consists of a series of collision events, each of them involving only a portion of the total powder mixture

Fig. 5. Effect of milling time on the density of the obtained SPS samples (I = 1100 A, P = 20 MPa, tSPS = 4 min, theoretical density = 4.678 g/cm3 ).

28

A.M. Locci et al. / Materials Science and Engineering A 434 (2006) 2329

Fig. 6. Effect of milling time on the crystallite sizes of TiC and TiB2 in the obtained SPS samples (I = 1100 A, P = 20 MPa, tSPS = 4 min).

possible that increased milling contributed to the formation of agglomerates, which make the sintering more difcult. The changes of average crystallite sizes of product phases formed during SPS with milling time are shown in Fig. 6. With increasing milling time, crystallite sizes markedly decrease initially and then decrease at a lower rate down to 50 nm after mechanically treating the starting powders for 24 h. The observed gradual crystallite size renement in dense product obtained by SPS as the ball milling time of reactants powders increases, is consistent with the results reported in the literature for analogous systematic investigations [13,15,19]. Indeed, ball milling leads to grain renements in the starting powders and the formation of interfaces between reactants. These features may induce a relative increase of nucleation sites, as compared to the case of unmilled powders, thus favoring nucleation with respect to grain growth. Moreover, by considering the starting mixture to be processed by SPS, the presence of nanostructured reaction products formed during milling, whose amount increases as mechanical treatment proceeds, also contributes to the formation of a material with grain size in the nanoscale range. This behavior was clearly observed in the literature during the synthesis of dense nanometric MoSi2 through mechanical and eld activation [18]. The effect of milling on the product microstructure is shown in Fig. 7(a) and (b) where, respectively, the SEM back-scattered micrographs of SPS samples obtained from unmilled powders and milled powders (tBM = 6 h) are shown. It may be seen that mechanical activation resulted in a ner microstructure and a more homogeneous phase distribution. Vickers hardness and fracture toughness were determined for samples obtained starting from unmilled powders and those synthesized by using 6 h milled reactants, which led to fully dense samples (cf. Fig. 5). It was found that Vickers hardness values for the samples obtained starting from unmilled and 6 h milled powders were 18.5 1.0 and 19.5 0.8 GPa, respectively. For the same samples, fracture toughness values were 6.94 0.49 and 7.73 0.44 MPa m1/2 , respectively. Thus, mechanical activation

Fig. 7. SEM back-scattered micrograph of SPS end-product when starting from: (a) unmilled; (b) ball milled (tBM = 24 h) powders (I = 1100 A, P = 20 MPa, tSPS = 4 min).

seems to improve mechanical properties of SPS product, especially regarding fracture toughness. It can probably due to higher density, microstructure homogeneity as well as ner crystallite size obtained when starting from milled reactants powders. It should be noted that the values obtained here are very similar to, and in some cases higher, than the best results reported in the literature for TiC/TiB2 dense composites having the same stoichiometry. In particular, hardness and fracture toughness of the synthesized SPS material are signicantly higher than the values reported by Lee et al. [15] and Zhang et al. [28], when comparing samples obtained starting from unmilled powders. The observed differences may be explained on the basis of the higher product density, i.e. greater than 98% of theoretical value, achieved during the present investigation. The density of the samples reported by Lee et al. [15] and Zhang et al. [28] are less than 96.5% and about 91%, respectively. On the other hand, Yuan et al. [29] have produced TiC/TiB2 composites characterized by very high density (99.8%) and hardness (19.6 GPa). However, fracture toughness was 5.53 MPa m1/2 , i.e. signicantly lower than the value obtained in this work. The comparison suggests that through the SPS method is possible to obtain dense

A.M. Locci et al. / Materials Science and Engineering A 434 (2006) 2329

29

TiC/TiB2 composite characterized by superior fracture toughness properties. By comparing TiC/TiB2 composite sample synthesized starting from milled powders, Lee et al. [15] obtained hardness value close to 21 GPa for milling time equal to 10 h. However, when tBM was set to 6 h, the authors reported about 18 GPa as hardness, which is lower than the one achieved in this work for the same milling time. Since, Lee et al. [15] did not report fracture toughness data, comparison about this mechanical property is not possible. It also worth noting that the mechanical pressure applied in this work is lower than those ones considered by other authors. In addition, except for Zhang et al. [28], who adopted a very fast technique, the synthesis time (4 min) needed to obtain by SPS the results reported in this work, is shorter than the one required by competitive methods proposed by other authors. Acknowledgements The nancial support of MIUR-PRIN (2002) and PRISMAINSTM (2003), Italy as well as NAMAMET project (NMP3-CT2004-001470), EU is gratefully acknowledged. The contribution given by Mrs. Gloria Arangino when performing some experimental trials is also gratefully acknowledged. The support of this work by the Army Research Ofce (ARO) to one of us (Z.A.M) is acknowledged. References
[1] C. Suryanarayana, Prog. Mater Sci. 46 (2001) 1184. [2] J.S. Benjamin, Metall. Mater. Trans. A 1 (1970) 29432951. [3] F. Charlot, E. Gaffet, B. Zeghmati, F. Bernard, J.C. Niepce, Mater. Sci. Eng. A262 (1999) 279288. [4] V. Gauthier, C. Josse, F. Bernard, E. Gaffet, J.P. Larpin, Mater. Sci. Eng. A265 (1999) 117128. [5] M.K. Beyer, H. Clausen-Schaumann, Chem. Rev. 105 (2005) 29212948. [6] D.L. Feldheim, Metal Nanoparticles: Synthesis, Characterization & Applications, Marcel Dekker Publisher, 2001.

[7] G. Schmid, Nanoparticles: From Theory to Application, John Wiley & Sons, 2004. [8] D.L. Zhang, Prog. Mater. Sci. 49 (2004) 537560. [9] M. Krasnowski, T. Kulik, Scripta Mater. 48 (2003) 14891494. [10] Y. Zheng, T. Yamasaki, T. Mitamura, M. Terasawa, T. Fukami, J. Jpn. Inst. Met. 67 (2003) 7478. [11] M. Jain, T. Christman, Acta Mater. Metall. 42 (1994) 19011911. [12] G.E. Korth, R.L. Williamson, Metall. Mater. Trans. A 26 (1995) 25712578. [13] T. Chen, J.M. Hampikian, N.N. Thadhani, Acta Mater. 47 (1999) 25672579. [14] F. Bernard, F. Charlot, E. Gaffet, Z.A. Munir, J. Am. Ceram. Soc. 84 (2001) 910914. [15] J.W. Lee, Z.A. Munir, M. Ohyanagi, Mater. Sci. Eng. A325 (2002) 221227. [16] E.M. Heian, S.K. Khalsa, J.W. Lee, Z.A. Munir, T. Yamamoto, M. Ohyanagi, J. Am. Ceram. Soc. 87 (2004) 779783. [17] M.S. El-Eskandarany, M. Omori, T.J. Konno, K. Sumiyama, T. Hirai, K. Suzuki, Metall. Mater. Trans. A 29 (1998) 19731981. [18] R. Orr` , J.N. Woolman, G. Cao, Z.A. Munir, J. Mater. Res. 16 (2001) u 14391448. [19] J.W. Lee, Z.A. Munir, M. Shibuya, M. Ohyanagi, J. Am. Ceram. Soc. 84 (2001) 12091216. [20] Z.M. Sun, Q. Wang, H. Hashimoto, S. Tada, T. Abe, Intermetallics 11 (2003) 6369. [21] S. Paris, E. Gaffet, F. Bernard, Z.A. Munir, Scripta Mater. 50 (2004) 691696. [22] K. Upadhya, J.M. Yang, W.P. Hoffman, Am. Ceram. Soc. Bull. 76 (1997) 5156. [23] S.R. Levine, E.J. Opila, M.C. Halbig, J.D. Kiser, M. Singh, J.A. Salem, J. Eur. Ceram. Soc. 22 (2002) 27572767. [24] A.M. Locci, R. Orr` , G. Cao, Z.A. Munir, J. Am. Ceram. Soc. 89 (2006) u 848855. [25] N.C. Halder, C.N.J. Wagner, Acta Crystallogr. 20 (1966) 312313. [26] D.K. Shetty, I.G. Wright, P.N. Mincer, A.H. Clauer, J. Mater. Sci. 20 (1985) 18731882. [27] A.L. Lasker, J.L. Bocquet, G. Brebec, C. Monty, Diffusion in Materials, Kluwer, 1990. [28] X.H. Zhang, C.C. Zhu, W. Qu, X.D. He, V.L. Kvanin, Compos. Sci. Technol. 62 (2002) 20372041. [29] R. Yuan, Z. Fu, W. Wang, H. Wang, Int. J. Self-Propag. High-Temp. Synth. 10 (2001) 435450.

S-ar putea să vă placă și