You are on page 1of 10

Acta Materialia 53 (2005) 967976 www.actamat-journals.

com

Characterization of coppercementite nanocomposite produced by mechanical alloying


P.A. Carvalho
a,*
a

, I. Fonseca a, M.T. Marques b, J.B. Correia b, A. Almeida a, R. Vilar

cnico, Engenharia de Materiais, Av. Rovisco Pais, 1049-001 Lisboa, Portugal Instituto Superior Te b INETI-DMTP, Estrada do Pac o do Lumiar, 1649-038 Lisboa, Portugal Received 8 June 2004; received in revised form 26 October 2004; accepted 27 October 2004 Available online 21 November 2004

Abstract A copperiron carbide nanocomposite has been synthesized by high-energy milling of elemental powders (Cu69Fe23C8), followed by annealing at 873 K. Phase identication and microstructure characterization have been carried out by transmission electron microscopy and energy dispersive spectroscopy. The carbide phase found in the as-milled material has been identied as Fe3C and/or Fe7C3, but clearly only Fe3C was present after annealing. Overall, grain sizes ranged from 10 to 50 nm in the as-milled condition and from 30 to 160 nm after annealing, with the carbide phase presenting a higher growth rate than copper. Stacking faults and a dispersion of Cu nanoparticles (510 nm) have been detected in annealed cementite while copper grains exhibited twins on {1 1 1} planes. Cementite growth could be evaluated in terms of precipitate growth theory. The remarkable thermal stability of the copper matrix is proposed to be related to solute drag eects. 2004 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Keywords: Transmission electron microscopy; Mechanical alloying; Copper; Nanocomposite

1. Introduction Ultrane-grained copper is usually unstable when exposed to moderate temperatures as a result of recovery processes and grain growth. For instance, the 50% recrystallization temperature in 98% cold-rolled 99.999% pure Cu is as low as 360 K [1]. Thermal stability can, however, be expected to increase if elements with low diusivity and limited solubility are added to Cu in order to delay diusion and promote grain boundary pinning by second phase precipitates. The CuFeC system seems particularly interesting in this respect due to the low to moderate substitutional diusivities and restricted equilibrium solid solubilities. In fact, estimated diusion coecient values for Fe in Cu and for Cu in aFe at 873 K are,
Corresponding author. Tel.: +351 218 418 122; fax: +351 218 418 120. E-mail address: pac@ist.utl.pt (P.A. Carvalho).
*

17 respectively, D873 m2 s1 [2] and Fe in Cu 1:8 10 19 873 DCu in aFe 9:9 10 m2 s1 [3]. Furthermore, the equilibrium CuFe phase diagram predicts negligible mutual solid solubility at temperatures below 873 K [4] and at this temperature the maximum solubility of C in Cu is a few ppm [5]. Early work on mechanically alloyed (MA) Cu100x Fex mixtures, based essentially on X-ray diraction (XRD), pointed to the formation of face centred cubic (fcc) solid solutions for x < 60 (for a review see [6]). More recently, however, Mo ssbauer spectroscopy and X-ray absorption studies have shown that, for high Cu contents, Fe does not form a random solid solution with Cu, but instead clusters of iron-rich c-phase tend to develop in a copper-rich matrix [7,8]. Indeed, close compositions in both phases and the same crystalline structure produce deceivably simple fcc diraction patterns, which have been masking the true nanocrystalline nature of these pseudo-solid solutions. The addition of

1359-6454/$30.00 2004 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved. doi:10.1016/j.actamat.2004.10.042

968

P.A. Carvalho et al. / Acta Materialia 53 (2005) 967976 Table 1 Raw materials Copper Size (lm) Purity (wt%) 44 < d < 149 99.9 Iron 44 99 Synthetic graphite 22 99.5

carbon to a copper-rich CuFe system seems most promising, as C is likely to be excluded from Cu and form iron carbides. This should delay diusion and consequently increase the thermal stability of the nanostructured system. In fact, mechanically alloying FeC mixtures is an ecient carbide production method, resulting in increasing fractions of FeyCz phases with increasing milling times (for a review see [9]). In addition, Mo ssbauer spectroscopy and XRD studies have shown that phase transformations occur on continued milling and subsequent heat-treatment in the FeC system. The transformation path is of the type: a-Fe ! amorphous=disordered carbide ! heating Fe3 C ! Fe7 C3 ! Fe3 C, with a suppression or extension of the amorphous stage determined by the impurities present, and with shorter transition times for higher carbon and energy contents [1013]. Preliminary characterization of the as-milled and annealed materials discussed in this paper has been carried out by XRD, scanning electron microscopy and microhardness tests [14]. The indexed reections of the asmilled material corresponded to copper and Fe7C3. This carbide indexation is in agreement with a Mo ssbauer spectroscopy and XRD study of a high-energy ballmilled Fe75C25 mixture [12]. After annealing at 873 K for 1 h, the carbide has clearly adopted the Fe3C structure. Average grain sizes determined with Scherrers equation for the as-milled condition were 13 nm for copper and 11 nm for the carbide phase. After annealing these values increased to 32 and 48 nm, respectively. The measured microhardness was 505 and 425 HV, before and after annealing, respectively. The MA material shows, therefore, a remarkable thermal stability. In eect, although annealing at high temperature tends to promote grain coarsening and a concurrent hardness decrease, the hardness level in the annealed condition still represents a signicant increase over pure as-milled copper (260 HV [15]). The purpose of this transmission electron microscopy (TEM) study was to conrm the phases present in the as-milled and annealed materials, to evaluate grain size and phase morphology, as well as to characterize structural defects and deformation structures, while aiming at an explanation for thermal stability.

In order to prevent oxidation, the container was lled with argon. The mill was operated at a rotating speed of 400 rpm for 50 h. The synthesized material was subsequently annealed at 873 K for 1 h in a tubular furnace under an argon ow. The material has been characterized by TEM using a Hitachi H-8100 instrument with a point resolution of 0.27 nm at 200 kV, equipped for energy dispersive spectroscopy (EDS). The samples for transmission electron microscopy were prepared by embedding the powder in an epoxy resin, followed by grinding, dimpling, and ion milling at 77 K to electron transparency. Standardless EDS analyses and microdiraction experiments have been performed with a 15 nm nominal probe diameter. Powder electron diraction patterns have been simulated for comparison with the electron microscopy image simulation (EMS) package program [16] using lattice parameters and atomic positions reported in [17,18].

3. Results and discussion Fig. 1 shows a bright/dark-eld set of images and corresponding diraction pattern obtained from the asmilled material. Analysis of the diraction rings enabled to establish the phases present as copper and Fe3C and/ or Fe7C3. Clear dierentiation between the two carbide phases was hindered by a low denition of the diraction rings (induced by the small scale of the microstructure) and by overlapping of major carbide rings with copper stronger ones. This last feature has also prevented a distinction between copper and iron carbide grains in dark-eld imaging. Nevertheless, dark-eld experiments have shown that the overall grain size lied between 10 and 50 nm. Furthermore, grain boundaries (GBs) were remarkably ill-dened and displayed a diuse contrast, characteristic of a high-energy nonequilibrium state. According to Valiev et al. [19], this diuse contrast results from the presence of an extremely distorted layer at the GBs. Twins could be observed in grains of the as-milled material. This result supports recent reports on twinning becoming an important deformation mechanism whenever few independent slip systems are available for general deformation, such as in nanocrystalline copper [20] and other nanocrystalline fcc systems [21,22]. Figs. 2 and 3 present bright/dark-eld sets of images and corresponding diraction patterns obtained from

2. Experimental procedures The nanocomposite has been prepared by mechanical alloying Cu, Fe and graphite powders in a Retsch PM400 planetary ball mill. The nominal composition of the mixture was 22.5 wt% Fe, 1.60 wt% C and bal. Cu. The size and purity of the elemental powders are presented in Table 1. The 250 ml stainless steel container was charged with a 20 g mixture of the elemental powders and 400 g of 10 mm diameter stainless steel balls.

P.A. Carvalho et al. / Acta Materialia 53 (2005) 967976

969

Fig. 1. Bright/dark-eld set of images and corresponding diraction pattern obtained from the as-milled material. The dark-eld image was obtained using a small aperture that included part of the two rst diraction rings. A grain presenting twin domains is shown in the magnied inset taken from the dark-eld image.

the annealed material. As can be seen, the microstructure in the annealed condition, which has evolved from a 1050 nm dispersion of as-milled grains, is still severely fragmented. On the whole, the observed grain size ranged from 30 to 160 nm. GBs were still generally illdened and intensity variations in some of the smaller grains suggest high elastic strain (see for example grain 1 in Fig. 3). Furthermore, long-range strain elds could sometimes be detected in the vicinity of GBs (Fig. 4), emphasizing the non-relaxed condition of the material after annealing. In fact, these results are in agreement with reports on non-equilibrium GBs in metals processed by intense straining [23,24]. Stacking faults (indicated in Fig. 2) and twins (grain 2 in Fig. 3) were frequently observed. These types of structural defect [25] and any existent mosaicity contribute to justify the apparently smaller grain size values determined by XRD [14]. Furthermore, a closer inspection at the grain internal structure (magnied inset in Fig. 2) revealed 5 10 nm particles dispersed within some of the grains. Intense diracting behavior under dark-eld imaging demonstrated the crystalline nature of these nanoparticles. A radial intensity prole through a powder diraction pattern obtained from the annealed material is compared with simulations for Cu and Fe3C in Fig. 5.

As with the diraction patterns of the as-milled condition (see Fig. 1), the large number of spots arranged in rings indicates a random crystallographic orientation, i.e., there is no evidence of texture in the granular structure. This suggests that many of the GBs are large-angle boundaries. Grain growth and a certain degree of recrystallization resulted in a better denition of the diffraction rings in the annealed condition. This allowed a good matching between simulated and experimental patterns enabling to assign the carbide rings to Fe3C. In agreement with the XRD results obtained for the annealed material [14], no other carbides, namely Fe7C3, have been detected. This indicates that any Fe7C3 initially present suered a phase transformation. The use of dark-eld imaging to strictly distinguish between carbide and copper grains was still hindered after annealing by partial overlapping of the most intense reections (see Fig. 5). This means that it was not possible to determine an average grain size for each phase. Nevertheless, microdiraction experiments and EDS point analyses have been used instead to identify and characterize Cu and Fe3C crystallites in the annealed material. Figs. 6 and 7 present typical Cu and Fe3C grains used for microdiraction experiments. Fig. 8 shows the result of point analyses performed at the regions numbered from 1 to 6. Regions 1, 3 and 5

970

P.A. Carvalho et al. / Acta Materialia 53 (2005) 967976

Fig. 3. Bright/dark-eld set of images and corresponding diraction pattern where the reections used for dark-eld imaging are circled (annealed condition). Two grains are close to Bragg condition for the reections chosen. Grain 1 exhibits strain contrast, while grain 2 presents twin domains.

Fig. 4. Bright-eld image featuring long-range strain elds near a slightly curved grain-boundary (annealed condition). Both grains exhibit nanoparticles.

Fig. 2. Bright/dark-eld set of images of the annealed material and corresponding diraction pattern where the reections used for darkeld imaging are circled. The grain sizes range from 30 to 160 nm. The magnied inset of the dark-eld image reveals the internal structure of the grains, which show a dispersion of bright 510 nm particles (indicated by arrows). Intense diracting behavior under dark-eld imaging demonstrates the crystalline nature of these nanoparticles. Stacking faults were observed in some of the grains (also indicated by arrows).

have been assigned to Fe3C grains. Since in Cu-rich regions 2, 4 and 6, grain contours are poorly dened, it is probable that more than one grain has been probed and

these analyses have to be interpreted with care. Nonetheless, the results suggest the presence of some Fe in Cu grains. Although microdiraction and EDS results have been obtained from relatively large grains, some general characteristics can be inferred. In fact, Cu grain size (<80 nm) is typically smaller than the one attained by Fe3C (<160 nm), suggesting a higher growth rate for the latter phase. A mottled contrast in some Cu grains (see grain 1 in Fig. 6 and regions 2, 4 and 6 in Fig. 8) indicates a high degree of elastic deformation. Fe3C grains, on the other hand, generally displayed a more uniform contrast (see Fig. 7), showing that this phase underwent a higher degree of recrystallization. This is corroborated by the fact that copper grains often presented corrugated and concave surfaces, while Fe3C grains exhibited smoother

P.A. Carvalho et al. / Acta Materialia 53 (2005) 967976

971

Fig. 5. Radial intensity prole through a powder diraction pattern and simulations for Cu and Fe3C.

Fig. 6. Bright-eld images of Cu grains used for microdiraction experiments in the annealed material. Grains 5 and 6 exhibit twins on {1 1 1} planes.

and convex surfaces. Microdiraction enabled to conrm that neighboring grains were generally severely disoriented, separated by large-angle GBs. Stacking faults

were present in Fe3C grains (see grain 4 in Fig. 7) while twins could be found on 1 1 1 planes of Cu (see grains 5 and 6 in Fig. 6). Since twins are typical deformation

972

P.A. Carvalho et al. / Acta Materialia 53 (2005) 967976

Fig. 7. Bright-eld images of Fe3C grains used for microdiraction experiments in the annealed material. Grain 4 exhibits stacking faults on planes close to (0 2 0).

Fig. 8. Bright-eld image of the annealed material region where EDS point analyses were performed. The associated table presents the atomic fraction of Fe and Cu contents found in the numbered regions. Regions 1, 3 and 5 were assigned to Fe3C grains. Cu-rich regions 2, 4 and 6 do not have well-dened grain contours and it is probable that more than one grain has been probed.

P.A. Carvalho et al. / Acta Materialia 53 (2005) 967976

973

Fig. 9. Bright-eld image of a Fe3C grain with stacking faults and an internal dispersion of nanoparticles. Magnied insets reveal 0 2 0 Fe3C lattice fringes. Bending of these fringes around the nanoparticles point to the presence of strain-elds. At the bottom inset some fringes terminate at the nanoparticle suggesting that the elastic constraint may have been reduced by dislocation formation (white arrow at a terminating fringe).

structures in milled copper [26], and have been observed in the as-milled condition, some twin domains present in the annealed material may have had a mechanical origin being subsequently extended during grain growth. Numerous nanoparticles have been found encapsulated in Fe3C (see grains 1 and 2 in Fig. 7). Since EDS analyses 1, 3 and 5 were carried out on well dened and relatively large Fe3C grains (Fig. 8), the measured Cu content resulted from Cu nanoparticles encapsulated within the grains. The presence of Cu nanoparticles in Fe3C does in eect explain an apparent excessive presence of Fe3C grains in an alloy with about 30% volume of carbide. Frequently, only a few nanoparticles existent in a grain displayed fringe contrast (probably Moire see magnied inset of grain 2 in Fig. 7). This conrms their crystalline nature and reveals that no preferred orientation relation exists between nanoparticles and encapsulating grains. The bright-eld image in Fig. 9, presents a Fe3C grain with nearly edge-on stacking faults and an internal dispersion of nanoparticles. The Fe3C lattice fringes observable at the insets resulted from the inclusion of small g reections by the objective aperture. According to their spacing, 0.33 nm, these fringes may correspond to either {1 0 1} planes or more probably to {0 2 0} planes (d1 0 1 = 0.337 nm and d0 2 0 = 0.336 nm [18]). In eect, stacking faults on 0 2 0 planes have been observed in other cementite grains (see grain 4 in Fig. 7, where the fault fringes are nearly perpendicular to g0 2 0) and have been reported previously [27]. The observable fringe bending around the nanoparticles results from the presence of strain-elds. Such strain may have originated

from dierential contraction during cooling from 873 K to room temperature. This can be justied by the difference in volumetric thermal expansion coecients, since aCu is 4.95 105 K1 [28], while aFe3 C is 4.1 105 K1 for temperatures above the Curie temperature (480 K) and under 1.8 105 K1 for lower temperatures [18]. Although the lack of preferred orientation relation points to incoherent interfaces between nanoparticles and encapsulating grains (ruling out lattice mist strains), some fringes terminate at the nanoparticles. This indicates that elastic constraint induced by thermal mist may have been relieved by dislocation formation as a consequence of the anisotropic thermal expansion behavior of Fe3C [18]. The clearly convex surfaces of Fe3C grains (see Figs. 6 and 9) imply an undergoing growth process, which is not so evident in the morphology of copper grains. In case of a supersaturated copper matrix, growth of the carbide phase takes place under the so-called precipitate growth regime until supersaturation drops close to its equilibrium value (%0 at.% Fe), at which time a coarsening regime can take over. In to order to discuss the mechanisms behind the observed growth it seems necessary to access the relevance of both regimes. Diusion-controlled growth of precipitates in a supersaturated matrix is described by the solution to the diusion eld equation [29]: r kDt
1=2

where r is the precipitate thickness, k is a function of supersaturation, D the volume diusivity and t the growth period considered. Given that interstitial diu-

974

P.A. Carvalho et al. / Acta Materialia 53 (2005) 967976

sion of carbon should not be a limiting factor, data on CuFe milled mixtures can be used to discuss growth of Fe3C in nanostructured copper: a solubility of nearly 12.5 at.% Fe in Cu-rich regions of a milled Cu80Fe20 mixture can be inferred from the fraction of CuFe to CuCu bonds in the Cu rst-coordination sphere [8]. As the equilibrium Fe-rich and Cu-rich phases have negligible mutual solubilities, such metastable solubility represents a supersaturation of about 0.25. As proposed by Aaron et al. [30] for spherical precipitates and low to moderate supersaturations, the stationary interface condition is the best approximation to the cumbersome analytical solution. In this case, k % 0.6, which for 17 D873 m2 s1 [2] and an initial radius Fe in Cu 1:8 10 of 10 nm results in r % 190 nm. For growth occurring after the matrix composition dropped close to the equilibrium value, the process should be discussed in terms of Ostwald ripening using the usual growth kinetics following LifshitzSlyozovWagner (LSW) theory [31,32]. The average radius r of coarsening particles after a heat treatment carried out for a period t at a temperature T is given by [29]: r kt1=3 ; 1

where k is the coarsening constant directly related to diusivity and equilibrium solubility. The LSW theory has been used previously to describe coarsening of Fe particles in a copper matrix [2,3336] and it has also been successfully tted to growth observed in CuCr and CuNb processed via mechanical alloying [37]. Considering again that interstitial diusion of carbon should not be a limiting factor, results on aFe coarsening in a copper matrix can be used to discuss the growth of Fe3C in nanostructured copper. In fact, Monzen and Kita [2] have investigated the coarsening of spherical aFe particles with no preferential orientation relation to the polycrystalline Cu matrix in diluted CuFe alloys. However, these authors have found that the reduced equilibrium solubility and low diusivity of Fe in Cu result in negligible growth of aFe after 1 h at 873 K, with k873 = 1.3 1029 m3 s1 [2]. For signicant volume fractions of the coarsening phase (/) a correction factor needs to be introduced [38]: a 10-fold increase of the coarsening constant can be expected for / around 0.30 (present 28 case). Even so, k 873 m3 s1 , which after /0:3 % 1 10 1 h at 873 K results in an average radius of only 11 nm for an initial radius of 10 nm and negligible growth for an initial radius of 80 nm. The previous analysis shows that a coarsening process under the thermal cycle used would not account for any signicant carbide growth. On the other hand, in view of the observations (30 nm < 2r < 160 nm), the average grain size calculated from a straight forward application of precipitate growth theory (2r % 380 nm)

appears excessive. The dierence between the experimental grain sizes and this calculated value suggests that Fe supersaturation in Cu-rich regions is lower in asmilled Cu69Fe23C8 than in as-milled Cu80Fe20 [8]. This agrees with the fact that the present material exhibits a strong tendency to form carbides since early stages of milling [14]. In eect, as Cu (metastable) solubility in carbide phases is expected to be lower than in fcc iron, a higher volume fraction of purer Cu-rich phase is probable for the as-milled CuFeC system. However, following this reasoning, only a comparably much lower supersaturation of 0.01 (0.5 at.% Fe) could have resulted in a more reasonable average radius of 80 nm by application of precipitate growth theory. Alternatively, the presence of a delaying mechanism could have slowed down carbide growth, preventing r to attain the higher expected value. In fact, a substantially unnished precipitation process would be supported by the apparent presence of Fe in Cu grains after the thermal cycle (see Fig. 8). A Fe7C3 ! Fe3C phase transformation involving recrystallization could be responsible for such a growth delay. Furthermore, although carbon diusion is not expected to be a limiting factor, the location of C in the as-milled material is of relevance to this discussion. Indeed, a Raman spectroscopy study has shown that CC bonds exist in CuC milled mixtures [39]. This suggests carbon does not tend to form interstitial solid solutions with Cu but rather appears in small clusters inside the Cu matrix [39]. In this case, besides Fe rejection by the supersaturated matrix, dissolution of C clusters would be required and could contribute to delay carbide growth. One is tempted to consider that growth in a nanostructured material may benet from GB diusion. In fact, Harrison [40] has shown that for non-negligible bulk diusivity, the diusion elds of neighboring GBs (short-circuits) start to overlap and contributing for an overall eective diusivity when the bulk diusivity penetration depth, (Dt)1/2 is larger than the average grain size. This is the case in the material studied, because (Dt)1/2 is estimated to be 250 nm, whereas the average carbide grain size is: 10 nm < 2r < 50 nm in the as-milled condition and 30 nm < 2r < 160 nm in the annealed one. Furthermore, the contribution to diusivity is very ecient for (Dt)1/2 higher then 10 times the grain size [41]. As a result, GB diusion may have signicantly increased the rate of carbide precipitation. This strengthens the hypothesis of a delaying mechanism since precipitate growth based on volume diusion already predicts growth in excess. Regarding the comparatively higher stability of the copper grains, it is important to mention that GB pinning by Fe3C precipitates has not been detected, although a dispersion of second phase grains among copper may have contributed to retard growth of this phase. On the other hand, annealing at 873 K is

P.A. Carvalho et al. / Acta Materialia 53 (2005) 967976

975

expected to have increased the ease of migration of iron, carbon and impurities to grain boundaries. This kind of solute migration can induce GB pinning by a solute drag eect, which is known to stagnate growth and increase the thermal stability of nanostructured materials [42]. The copper nanoparticles in cementite grains can be remnants of an initial nanocrystalline structure. Indeed, as the diusivity of Cu in Fe3C (estimated by Fourlaris 21 for 832 K, D832 m2 s1 [43]) is much Cu in Fe3 C 2:6 10 lower than that of Fe in Cu, any iron-rich clusters present in Cu grains tend to dissolve away towards developing Fe3C grains, whereas copper-rich clusters inside carbide grains may tend to remain there. Alternatively, some nanoparticles can result from encapsulation by the fast growing carbide. Precipitation of Cu particles in cementite cannot however be ruled out as it is known to occur in copper steels [4346]. In any case, inclusion of Cu particles in cementite favors its growth and is detrimental to Cu grain growth contributing for its stabilization. In summary, a coarsening mechanism would not be able to account for the observed carbide growth. On the other hand, the application of precipitate growth theory seems to result in excessive growth rates (especially if GB diusion and Cu encapsulation are taken into account). It is therefore proposed that the growth process is not solely diusion-controlled, i.e., there is a delaying mechanism hindering the fast growth of the carbide phase. Since GB pinning by Fe3C precipitates has not been detected, the comparatively higher stability of the copper grains may result from a solute drag eect, with a contribution from the Cu nanoparticles encapsulation inside carbide grains. The work carried out shows that MA in the Cu FeC system followed by heat-treatment involves complex growth mechanisms and therefore deserves more attention. In the particular mixture studied, the various factors aecting the kinetics of Fe3C grain growth from a supersaturated Cu matrix need to be thoroughly analyzed. Namely, the delaying inuences of a Fe7C3 ! Fe3C phase transformation and of a possible presence of C clusters in the copper matrix need to be assessed. On the other hand, the eect of GB vs. volume diusion on growth rate and the role of Cu encapsulation must be evaluated. The mechanisms behind Cu encapsulation in cementite need to be further scrutinized in order to establish whether the nanoparticles present are remnants of the initial nanocrystalline structure, result from encapsulation by developing Fe3C grains or even from Cu precipitation in cementite. Furthermore, the solute drag eect at Cu grain boundaries, which seems responsible for the thermal stability of this phase, needs to be investigated. Evaluation of partially annealed states and of other compositions and annealing temperatures is thus

required to fully comprehend this nanostructured system.

4. Conclusions The carbide phase present in the as-milled material has been identied as Fe3C and/or Fe7C3, while only Fe3C could be found after annealing, in agreement with the XRD results previously reported. Overall, grain sizes ranged from 10 to 50 nm in the as-milled condition and from 30 to 160 nm after annealing, with the carbide phase presenting a higher growth rate than copper, which has proven to be remarkably stable under the annealing conditions used. Non-equilibrium features observed at the grain boundaries indicate that the material is in a high-energy non-relaxed condition even after annealing. Stacking faults were present in Fe3C, whereas copper grains exhibited twins on {1 1 1} planes. A dispersion of Cu nanoparticles (510 nm) has been detected within annealed cementite; their presence was attributed to the low diusivity and solubility of Cu in Fe3C. The thermal stability of the copper phase is proposed to be related to solute drag eects, while the growth behavior of cementite can be accountable by precipitation from an Fe supersaturated Cu matrix. Nevertheless, a straight forward application of precipitate growth theory results in excessive growth rates, suggesting that a delaying mechanism is aecting carbide growth. Acknowledgement ncia e The authors acknowledge Fundac a o para a Cie a Tecnologia, projectPOCTI/CTM/40892/2001, for nancial support. References
[1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] Decker BF, Harker D. Trans AIME 1959;188:887. Monzen R, Kita K. Phil Mag Let 2002;82:373. Lazarev VA, Golikov M. Fiz Metallov Metalloved 1970;29:598. Alloy phase diagrams, ASM handbook, vol. 3. Materials Park (OH): American Society for Metals; 1992, p. 110. pez GA, Mittemeijer EJ. Scripta Mater 2004;51:1. Lo Ma E, Atzmon M. Mater Chem Phys 1995;39:249. Principi G, Spataru T, Gupta R, Enzo S, Kuncse V, Filoti G. J Alloy Compd 2001;326:188. Wei S, Yan W, Li Y, Liu W, Fan J, Zhang X. Physica B 2001;305:135. Suryanarayana C. Prog Mater Sci 2001;46:1. Wang GM, Champbell SJ, Calka A, Kaczmarek WA. Nanostruct Mater 1995;6:389. Campbell SJ, Wang GM, Calka A, Kaczmarek WA. Mater Sci Eng 1997;A226A228:75. Tokumitzu K, Umemoto M. J Metastable Nanocrystall Mater 2001;10:183. Umemoto M, Liu ZG, Masuyama K, Tsuchiya K. Scripta Mater 2001;45:391.

976

P.A. Carvalho et al. / Acta Materialia 53 (2005) 967976 [29] Martin JW, Doherty RD, Cantor B. Stability of microstructure in metallic materials. Cambridge: Cambridge University Press; 1976. p. 67. [30] Aaron HB, Fainstein D, Kotler GR. J Appl Phys 1970;41:4404. [31] Lifshitz IM, Slyozov VV. J Phys Chem Solids 1961;19:35. [32] Wagner C. Z Electrochem 1961;65:581. [33] Fujii T, Kato M, Mori T. Trans JIM 1991;32:229. [34] Kita K, Monzen K. Scripta Mater 2000;43:1039. [35] Kita K, Monzen K. Trans JIM 2001;65:223. [36] Monzen R, Tada T, Seo T, Higashimine. Mater Lett 2004;58:2007. [37] Morris MA, Morris DG. Mater Sci Eng A 1989;111:115. [38] Ardell AJ. Acta Metall 1972;20:61. [39] Marques MT, Correia JB, Conde O. Scripta Mater 2004;50: 963. [40] Harrison LG. Trans Faraday Soc 1961;57:1191. [41] Le Claire AD, Rabinovitch A. J Phys 1983;C16:2087. [42] Liu KW, Mu cklich F. Acta Mater 2001;49:395. [43] Fourlaris G, Baker AJ, Papdimitriou GD. Acta Mater 1996;44:4791. [44] Fourlaris G, Baker AJ, Papdimitriou GD. Acta Metall Mater 1995;43:2589. [45] Fourlaris G, Baker AJ, Papdimitriou GD. Acta Metall Mater 1995;43:4421. [46] Fourlaris G, Baker AJ. J Phys IV 1997;C5:389.

[14] Correia JB, Marques MT. Adv Materials Forum II 2004;455 456:501. nez MJ, Matteazzi P. [15] Marques MT, Correia JB, Criado JM, Dia Key Eng Materials 2002;230232:652. [16] Stadelmann PA. Ultramicroscopy 1987;21:131. [17] Straumanis ME, Yu LS. Acta Crystallogr 1969;A25:676. [18] Wood IG, Vocadlo L, Knight KS, Dobson DP, Marshall WG, Price GD, et al. J Appl Cryst 2004;37:82. [19] Valiev RZ, Korznikov AV, Mulyukov RR. Mater Sci Eng 1993;A168:141. [20] Liao XZ, Zhao YH, Srinivasan SG, Zhu YT, Valiev RZ, Gunderov DV. Appl Phys Lett 2004;84:592. [21] Liao XZ, Zhou F, Lavernia EJ, He DW, Zhu YT. Appl Phys Lett 2003;83:5062. [22] Ro sner H, Markmann J, Weissmu ller J. Phil Mag Lett 2004;84:321. [23] Valiev RZ, Gertsman VY, Kaibyshev OA. Phys Status Solidi 1986;97A:97. [24] Horita Z, Smith DJ, Nemoto M, Valiev RZ, Langdon TG. J Mater Res 1988;13:446. [25] Gayle FW, Biancaniello FS. Nanostruct Mater 1995;6:429. [26] Huang JY, Wu YK, Ye HQ. Acta Mater 1996;44:1211. [27] Nishiyama Z, Koreeda A, Katagiri S. Trans JIM 1964;5:115. [28] Gray DE, editor. American institute of physics handbook. 3rd ed. New York: McGraw-Hill; 1972.

You might also like