Вы находитесь на странице: 1из 8

Materials Science and Engineering A 528 (2011) 36013608

Contents lists available at ScienceDirect

Materials Science and Engineering A


journal homepage: www.elsevier.com/locate/msea

Inuence of strain rate on the characteristics of a magnesium alloy processed by high-pressure torsion
Pauline Serre a,1 , Roberto B. Figueiredo a,2 , Nong Gao a , Terence G. Langdon a,b,
a b

Materials Research Group, School of Engineering Sciences, University of Southampton, Southampton SO17 1BJ, UK Departments of Aerospace & Mechanical Engineering and Materials Science, University of Southern California, Los Angeles, CA 90089-1453, USA

a r t i c l e

i n f o

a b s t r a c t
A commercial magnesium AZ31 alloy was processed by high-pressure torsion (HPT) through 5 turns at 453 K using three different rotation rates. The results show that HPT introduces signicant grain renement to an average grain size of 0.5 m. The processed structures were heterogeneous in the centers of the disks and homogeneous at distances at and above 1 mm from the disk centers. The measured values of the microhardness were also reasonably homogeneous after HPT except only near the disk centers. The hardness values and the average grain sizes were correlated with the local strain rates by using the imposed rotation rates and estimates of the strains within the disks. The results demonstrate the hardness and the average grain size both exhibit a small dependence on the local strain rate. 2011 Elsevier B.V. All rights reserved.

Article history: Received 5 September 2010 Received in revised form 2 December 2010 Accepted 20 January 2011 Available online 26 January 2011 Keywords: Hardness High-pressure torsion Homogeneity Magnesium alloy Ultrane grains

1. Introduction The processing of metals through the application of severe plastic deformation (SPD) has become attractive as a procedure for introducing signicant grain renement in bulk solids and thereby reducing the grain size to the submicrometer (0.11.0 m) or even the nanometer (<100 nm) range [1]. Several SPD techniques are now available but the procedures receiving the most attention are equal-channel angular pressing (ECAP) [2] and high-pressure torsion (HPT) [3]. The experimental evidence available to date shows that the ultrane grains produced by HPT are generally smaller than those produced by ECAP [46]. The principles of HPT were described in earlier reports [7,8]. The sample, in the form of a thin disk, is held in place in a horizontal position between two massive anvils and it is then subjected to a high applied pressure, P, and concurrent torsional straining through rotation of the lower anvil. Numerous reports have been published documenting the variations in microstructure and properties attained in a wide range of materials when using different values

of the applied pressure and different numbers of rotational turns, N [3]. However, although there are some recent reports describing HPT experiments where testing was conducted using different rotational speeds [9,10], no systematic information is available on the inuence of the rotation speed on the microstructural characteristics produced in HPT disks. Accordingly, the present investigation was undertaken to provide this information for a magnesium AZ31 alloy. As will be demonstrated, the variation of strain within each disk, when combined with the speed of rotation of the lower anvil, provides a unique opportunity to incorporate the local strain rate into the analysis.

2. Experimental material and procedures The experiments were conducted using a commercial magnesium AZ31 alloy (Mg3 wt% Al1 wt% Zn). This material was selected because there are numerous reports of the development of heterogeneous microstructures in the processing of the AZ31 alloy by ECAP [1113] and therefore it is important to check the ability to achieve homogeneous microstructures when processing by HPT. The material was supplied by Timminco Corporation (now Applied Magnesium International), Aurora, CO, as extruded rods having diameters of 10 mm. Disks with thicknesses of 1.5 mm were cut from the rods and ground with abrasive papers to nal thicknesses of 0.81 mm. For processing by HPT, each disk was placed in a depression on the upper surface of the lower anvil of an HPT facility equipped with a small heating device in which heating elements can be

Corresponding author at: Department of Aerospace & Mechanical Engineering, University of Southern California, Los Angeles, CA 90089-1453, USA. Tel.: +1 213 740 0491; fax: +1 213 740 8071. E-mail address: langdon@usc.edu (T.G. Langdon). 1 On leave from Phelma - School of Engineering in Physics, Electronics and Materials, Grenoble INP Minatec, 3 Parvis Louis Nel, BP 257, 38016 Grenoble Cedex 1, France. 2 Now at Department of Metallurgical and Materials Engineering, Federal University of Minas Gerais, Belo Horizonte 31270-901, Brazil. 0921-5093/$ see front matter 2011 Elsevier B.V. All rights reserved. doi:10.1016/j.msea.2011.01.066

3602

P. Serre et al. / Materials Science and Engineering A 528 (2011) 36013608

placed around the lower and upper anvils. The two anvils were then brought together by raising the lower anvil so that the disk was contained between depressions on either face of the two anvils, the heating elements were placed in position and the temperature was raised to 453 K. The temperature was continuously monitored and controlled by a thermocouple placed above the disk in the center of the upper anvil. All disks were processed under quasi-constrained conditions [3] using an applied pressure of P = 6.0 GPa and a total of N = 5 turns. In order to evaluate the inuence of the processing speed, individual disks were processed by HPT using rotational speeds for the lower anvil of 0.5, 1 and 2 rotations per minute. During processing, the measured variation in temperature was not greater than 5 degrees. After processing, the disks were marked to keep track of the top and bottom surfaces. This is important because there is very recent evidence of unusual macroscopic shearing patterns in HPT processing [1416] and these patterns may vary depending upon the precise plane within the HPT disk. Following HPT, the processed disks were mounted in cold resin and carefully ground with abrasive paper in order to remove layers of 0.1 mm in thickness from the top surfaces. The samples were then polished on cloth with diamond paste and nal polishing was carried out using a colloidal silica solution. All microstructural observations were recorded on these polished sections close to the upper surface of each disk. The polished surfaces were etched using an acetic picral solution (5 mL of acetic acid, 6 g of picric acid, 10 mL of water and 100 mL of ethanol) in order to reveal the grain boundaries. An optical microscope was used to observe the surfaces under polarized light. Representative images of the grain structures were recorded at different locations on the samples and for each image the distance was measured between the center of the image and the center of the disk. Measurements were taken within each image to record the longest axis of each grain and a minimum of 100 grains was used to provide quantitative information on the distribution of grain sizes within each area. In addition, similar sets of images and measurements were also recorded for disks of the unprocessed alloy. Following these measurements, the etched surfaces were repolished to a mirror-like nish and measurements of the Vickers microhardness, H v, were taken using a Matsuzawa Seiki MHT1 microhardness tester equipped with a Vickers indenter. These measurements used a load of 200 gf and a dwell time of 15 s. The locations of the indentations were chosen in order to determine the distributions of hardness along the disk diameters and over a one-quarter segment of each disk surface. For measurements of the hardness along the diameters of the disks, two sets of parallel indentations were recorded across the disk surfaces through the centers of the disks and the hardness values were then determined as the average of four separate indentations each equally distant from the selected point. For measurements of hardness over the one-quarter segments of the disks, a grid of indentations was established with each indentation separated from the adjacent indentations by a distance of 0.3 mm. The hardness value at each indentation was then plotted directly in order to provide a colorcoded visual display of the hardness distribution within the quarter segment. 3. Experimental results 3.1. Grain structure after HPT Fig. 1 shows two representative images of the grain structure in the as-received material. In the central region of the rod, the structure consisted of a distribution of ne grains having sizes within the range of 545 m as shown in Fig. 1(a). Around the periphery of the rod, up to a distance of 0.4 mm from the edge, there were

Fig. 1. Grain structure in the as-received material (a) in the central region and (b) near the edge of the rod.

coarser grains and signicant twinning as shown in Fig. 1(b). Measurements within this narrow region showed there was a mean separation between twins of 7 m and an average twin thickness of 2.5 m. Signicant grain renement was introduced by the HPT processing. Detailed inspection over wide areas showed the microstructure generally consisted of ultrane grains having sizes of <1 m but some larger grains were present in some areas, especially near the centers of the disks, with sizes exceeding 1 m. Fig. 2 shows representative images of the grain structure in a disk processed by HPT with a rotation rate of 1 rpm (a) close to the center of the disk and (b) at a radius, r, of 2 mm from the center where there is a uniform distribution of grains. 3.2. Grain size measurements The grain size and the grain size distribution were determined at selected areas in the as-received material and in the disks processed by HPT. Fig. 3 shows the average grain size plotted as a function of the distance from the centers of the unprocessed disk and a disk processed by HPT through 5 turns at a rotational speed of 1 rpm, respectively. This plot demonstrates the signicant grain renement introduced by HPT and shows also there is a small decrease to an average grain size of 0.5 m with increasing distance from the center of the disk after processing by HPT. This result is consistent with several earlier reports showing lower hardness values,

P. Serre et al. / Materials Science and Engineering A 528 (2011) 36013608

3603

Fig. 2. Grain structure after processing by HPT for 5 turns at a rotation rate of 1 rpm (a) in the central region and (b) at a distance of 2 mm from the center.

and therefore larger grain sizes, in the centers of disks processed by HPT [5,7,1725]. To provide a direct evaluation of the level of homogeneity within the grain size population, the grains were divided into separate groups each composed of a 5 m increment. Fig. 4(a) shows the

Fig. 4. Distribution of grain sizes at different positions within the materials (a) in the as-received condition and (b) after processing by HPT for 5 turns using a rotation rate of 1 rpm.

Fig. 3. Average grain size as a function of the distance from the center in the asreceived condition and after processing by HPT for 5 turns using a rotation rate of 1 rpm.

area fraction of each group as a function of the range of grain sizes for the as-received material and a similar histogram is shown in Fig. 4(b) for the disk processed by HPT at a rotational speed of 1 rpm. In Fig. 4(a) the two separate plots correspond to grain size measurements recorded in the vicinity of the center of the rod at a radius of r = 0 mm and within the rod at a distance of r = 2 mm from the center. These results show that there is no preferential range of grain size corresponding to a majority of the area fraction and therefore the grain size distribution in the as-received material is heterogeneous with grain sizes predominantly in the range from 5 to 45 m. Although some grains smaller than 5 m were observed in the as-received material, these grains occupied an area fraction of <1%. By contrast, Fig. 4(b) shows the distribution of grain size in the material processed by HPT at 1 rpm with separate distributions plotted for the center of the risk and at radial distances of 1, 2, 3 and 4 mm from the center. It is apparent that the structure is heterogeneous in the center of the disk with grain sizes ranging, essentially in a uniform distribution, from 0.5 to 5.5 m. However, the structure is homogeneous at distances at and above 1 mm from the center with a majority of the grains lying in the range from 0.5 to 1.0 m. The distributions show also that the area fractions of grains in the range from 1.0 to 1.5 m decrease with increasing

3604

P. Serre et al. / Materials Science and Engineering A 528 (2011) 36013608

Fig. 5. Distribution of grain sizes at (a) the center of the disk and (b) at a distance of 2 mm from the center for disks processed by HPT through 5 turns using rotation rates of 0.5, 1 and 2 rpm.

distance from the center whereas the trend occurs in the opposite sense for grains smaller than 0.5 m. Although Figs. 3 and 4(b) provide information only on the disk tested at a rotational speed of 1 rpm, very similar distributions were also recorded for the disks processed at 0.5 and 2 rpm. The signicance of the rotational speed is illustrated directly in Fig. 5 where the grain size distributions are shown after processing by HPT at the three different rotational speeds for areas located (a) at the centers of the disks and (b) at 2 mm from the centers. These distributions conrm that, for all three rates of rotation, the structures are heterogeneous in the centers and homogeneous at 2 mm from the centers. For the measurements at a distance of 2 mm from the centers, it is apparent also that the area fractions of the smallest grains, in the range below 0.5 m, increase with increasing rotation speed. This is reasonable because less time is available for any grain growth at the faster rotation rates. 3.3. Measurements of the Vickers microhardness Fig. 6 shows the distributions of the microhardness values plotted as a function of the position on each disk for the samples processed by HPT for 5 turns at rotational speeds of (a) 0.5, (b) 1 and (c) 2 rpm, respectively: each plot includes also the measured

Fig. 6. Distribution of the Vickers microhardness. H v, along the diameter of disks processed by HPT through 5 turns using rotation rates of (a) 0.5, (b) 1 and (c) 2 rpm: the distribution of hardness in the as-received material is also shown.

microhardness distribution in the as-received material. All of these plots show there is a very signicant increase in hardness when processing by HPT and this is consistent with the renement in grain size documented earlier. The plots show also the hardness distributions are reasonably homogeneous over the diameters of the disks except only for slightly lower values of hardness near the disk centers. Close examination of the plots in Fig. 6 shows that the average values of H v near the outer edges of the disks increase with

P. Serre et al. / Materials Science and Engineering A 528 (2011) 36013608

3605

Fig. 7. Color-coded maps showing the distributions of hardness of one-quarter of the disks after processing by HPT through 5 turns using rotation rates of (a) 0.5, (b) 1 and (c) 2 rpm: the signicance of the colors is indicated in the color key on the right. (For interpretation of the references to color in this gure legend, the reader is referred to the web version of the article.)

rotational speed from H v 85 at 0.5 rpm to H v 90 at a speed of 2 rpm. The overall homogeneity in hardness may be displayed more directly by plotting color-coded contour maps as shown in Fig. 7 for one-quarter of each disk surface for rotational speeds of (a) 0.5, (b) 1 and (c) 2 rpm, respectively: in these plots, the hardness values were recorded following a recti-linear grid pattern over a one-quarter segment of each disk and the datum points are plotted using an arbitrary orthogonal XY coordinate system wherein the points labeled (0,0) correspond to the centers of each disk. These maps show there is a good homogeneous microstructure in all three disks after HPT processing with all microhardness values lying within the limited range of H v 8090 but with slightly lower values of H v in the vicinity of the disk centers. 4. Discussion 4.1. The nature of microstructural renement The rst important conclusion from these experiments is that HPT processing of the magnesium AZ31 alloy at a temperature of 453 K leads to exceptional grain renement. This temperature for HPT processing is within the range of temperatures generally used for processing by ECAP but the grain sizes attained in HPT processing are very much smaller than in ECAP. For example, a recent detailed tabulation of the results obtained when processing the AZ31 alloy by ECAP [26] shows that, although using similar processing temperatures, the nal average grain sizes are generally in the range of a few microns in ECAP whereas the present experiments give a nal grain size of 0.5 m when using the same alloy and a similar processing temperature. This conrms that, as in earlier experiments [46], processing by HPT is more effective in producing an ultrane grain structure by comparison with ECAP. The second important conclusion relates to a comparison of the grains sizes produced when processing by HPT and ECAP. The recent tabulation of structural evolution in magnesium alloys when processing by ECAP, including the AZ31 alloy [26], shows that homogeneous grain structures are achieved in ECAP only when the processing is continued through a relatively large number of passes. For example, a heterogeneous grain structure was reported in AZ31 after processing through 4 passes of ECAP but there was an evolution into a homogeneous structure after processing through 8

passes, with similar results obtained both when pressing at 423 K in the presence of a back-pressure and when pressing at 473 K without a back pressure [27]. The nature of the grain renement in ECAP processing is consistent with reports of the occurrence of dynamic recrystallization in magnesium alloys deformed at temperatures similar to the ECAP processing temperature [28] where grains are nucleated along the original grain boundaries and along twin boundaries in a necklacelike pattern due to the activation of slip on both the basal and non-basal slip systems [29]. It follows, therefore, that the rening of grains by SPD processes is very different in magnesium alloys by comparison with f.c.c. metals where there is no shortage of active slip systems. Whereas in f.c.c. metals the grains are rened by the formation of arrays of subgrains and the subsequent evolution of these low-angle boundaries in to boundaries having high angles of misorientation [30], in magnesium alloys there is a formation of necklace arrays of smaller grains along the original grain boundaries and the gradual expansion of these arrays to ll the grain interiors [26,31]. Therefore, in ECAP the microstructure can evolve, with sufcient numbers of ECAP passes, into a homogeneous microstructure whereas in HPT the high applied pressures and the large imposed strain throughout most of the disk leads to a homogeneous distribution of grain sizes in the early stages of HPT except only for any heterogeneities that are retained initially in the centers of the discs. The subsequent evolution in microstructure in the centers of the disks, as demonstrated in earlier reports [10,22,24,3235], is consistent with the strain gradient plasticity modeling developed to explain microstructural evolution in HPT [36]. These observations demonstrate that the heterogeneous distribution of grain sizes is transient in nature and a homogeneous grain structure will develop after relatively large amounts of deformation where this is attained more readily when processing by HPT by comparison with ECAP. The evolution in microstructure can be readily illustrated as in Fig. 8 where distributions are plotted for three different conditions: for the as-received unprocessed material and after HPT processing at the centers of the disks and at a distance of 2 mm from the centers. In Fig. 8, the as-received data is the sum of the measurements recorded at the r = 0 mm and r = 2 mm positions shown in Fig. 4(a) and, since the grain sizes were reasonably similar for all three rotational speeds as shown in Fig. 5(b), the data for the HPT disks were obtained using the summation of experimental data for the three different processing conditions. In HPT the equivalent von Mises

3606

P. Serre et al. / Materials Science and Engineering A 528 (2011) 36013608

Fig. 8. Distributions of the area fractions occupied by groups of grains with different sizes for the as-received material and after processing by HPT through 5 turns in the central region of the disk and at a distance of 2 mm from the center.

Fig. 9. Values of the Vickers microhardness plotted as a function of the equivalent strain in disks processed by HPT through 5 turns using rotation rates of 0.5, 1 and 2 rpm.

strain, , may be estimated using the relationship [3739] = 2N r h 3 (1)

where h is the height (or thickness) of the disc. With reference to Fig. 8, there is no strain in the as-received material so that = 0 and, using Eq. (1), the strains are estimated as 15 in the center of the disk assuming representative radii in the central region of 50250 m and 45 at a distance of 2 mm from the center. It is evident from Fig. 8 that the initial coarse and heterogeneous grain size distribution in the as-received material at = 0 evolves into a ne but nevertheless heterogeneous grain structure at 15 but there is additional evolution into an ultrane grain structure which is essentially homogeneous at 45. This analysis demonstrates, therefore, that it should be relatively easy to achieve homogeneous structures in magnesium alloys when processing by HPT. 4.2. The evolution of hardness in HPT processing Although Section 4.1 concentrated on the evolution of the grain size distribution in HPT, there is also a similar evolution in the measurements of the Vickers microhardness. Again, the amount of deformation at different radii within the disks is readily estimated using Eq. (1). It was suggested in very early HPT experiments that the variations in hardness across HPT disks may be readily correlated at every point within the disks by plotting the hardness values against the equivalent strain [19] and this same approach was used recently to normalize hardness data in HPT processing for a number of pure metals [40]. Using this approach, the values of the Vickers microhardness plotted as a function of distance from the centers of the disks in Fig. 6 may be replotted as a function of the equivalent strain as shown in Fig. 9 where the hardness values were obtained by averaging the two sets of measurements recorded on either side and equidistant from the disk centers. It is apparent from Fig. 9 that all points for each disk fall on a reasonable line and there is an initial increase in hardness at the lowest equivalent strains but the hardness values for each disk tend to saturate at equivalent strains higher than 20. Despite this agreement, the results show also that slightly different hardness values occur at the same strain within the different disks and, in addition, the saturation hardness values at the highest equivalent strains increase slightly but signicantly with increasing rotational speed. Although this latter increase is very small, representing a

change from H v 85 at 0.5 rpm to H v 90 at 2 rpm, the change is larger than the estimated error bars inherent in the hardness measurements and, furthermore, the same trend is visible for all datum points at equivalent strains above 45. Therefore, this small variation is considered a reliable indication of the overall behavior when processing by HPT. 4.3. The inuence of the local strain rate There has been no attempt to date to correlate hardness and grain size measurements in HPT with the strain rate imposed during the processing operation and instead earlier studies focused exclusively on the signicance of applied pressure and the total number of revolutions in torsional straining [3]. However, it is reasonable to anticipate the local strain rate may play a signicant role because of a potential inuence on the advent of nonbasal slip systems and, accordingly, on the ability to achieve dynamic recrystallization. A fundamental characteristic of the HPT process is the heterogeneity of the strain imposed throughout the disc during torsional straining, where this strain may be directly estimated from Eq. (1). In practice, the level of strain increases with increasing distance from the center of the disk but the total time of processing remains constant for the whole disk. It follows, therefore, that the strain , equivalent to /t where t is the time, is proportional to rate, the local strain and therefore it reaches a maximum at the edge of the disk. It also follows that the local strain rate at any selected distance from the center of the disk is dependent upon the rotation rate. Considering a constant rate of rotation during HPT processing, the local strain rate may be estimated at any point using the relationship = 2N r th 3 (2)

where t corresponds to the total time for the processing operation. Considering the three different rotation rates used in these experiments, Fig. 10 shows a plot of the microhardness values against the estimated local strain rate for all three disks tested at the different rotational speeds. Despite some scatter, all datum points fall essentially on the same line and the separation into discrete lines for the three different rotation speeds, as evident in Fig. 9 at the higher equivalent strains, is now removed. Furthermore, the slope suggests a proportionality of the form 0.02 Hv (3)

P. Serre et al. / Materials Science and Engineering A 528 (2011) 36013608

3607

Fig. 10. Values of the Vickers microhardness plotted as a function of the local strain rate for disks processed by HPT through 5 turns at different rotation rates.

which implies a minor, but not insignicant, dependence on the speed of rotation in the HPT processing operation. It is important to note that the values of hardness at different positions on the same disk and on different disks processed at different rotational speeds all follow the same general trend. This shows that the hardness values measured on HPT disks are dependent not only upon the local strain but also upon the local strain rate. The area fraction occupied by the nest grains increases both with the rotation rate as evident in Fig. 5(b) and with increasing distance from the center of the disk as evident in Fig. 4(b) thereby suggesting that ner grain structures may be attained at faster strain rates. In order to check the variation of grain size with strain rate, Fig. 11 plots the average grain size measured at different positions on the disks experiencing different strain rates as a function of the local strain rate. Datum points are shown in Fig. 11 for the r = 2 mm position at the three different rotation speeds and for the three additional positions of 1, 3 and 4 mm at a rate of 1 rpm. Again, all points fall on a single line such that the grain size decreases slightly with increasing strain rate. These results suggest a proportionality of the form 0.1 d (4)

where again there is a small, but not insignicant, dependence on strain rate. It is important to note that Eqs. (3) and (4) are mutually consistent because the results in Fig. 11 imply an increasing hardness with increasing strain rate since ner grain sizes are expected to display higher strengths. In order to explain these results, it is instructive to examine other data available in the literature. Experiments on the AZ31 alloy using ECAP with a channel angle of 120 showed that the size of the newly formed grains decreased with increasing strain rate [41] where this is consistent with Eq. (4). Other results suggest that the area fraction occupied by the ne grains may decrease with increasing strain rate [42]. These trends are readily predicted by the mechanism developed earlier for grain renement in magnesium alloys [26,31] because at low strains the rening process is incomplete so that a relatively large volume of the structure remains unrened. In the present investigation, the area fraction occupied by the nest grains increases with increasing strain rate and this is attributed to the large amount of deformation imposed by HPT which in practice is sufcient to rene the whole structure. Therefore the size of the newly formed grains plays a major role in HPT processing in the nal distribution of the individual grain sizes. It should be noted that no other information is available at present on the effect of the imposed strain rate on the microstructural characteristics in HPT but a very early study examined the inuence of pressing speed on microstructural development in pure Al and an Al1% Mg alloy when processing by ECAP [43]. Using transmission electron microscopy (TEM), the results from this study concluded that the pressing speed has no inuence on the measured ultrane grain sizes, at least over pressing speeds spanning three orders of magnitude from 102 to 10 mm s1 . This earlier result is not inconsistent with the results obtained using HPT processing when it is noted that the grain size produced in SPD processing exhibits only a very minor dependence on strain rate as shown in Eq. (4) and, in addition, the earlier grain size measurements were taken directly from the TEM images where the error bars are necessarily fairly large. Finally, it is important to address the question of whether texture may play a role in microstructural evolution in these different samples. An earlier report described the evolution of texture in extruded pure magnesium processed by HPT and it was shown that the initial texture evolved rapidly into a steady-state condition that was characteristic of shear deformation [44]. Since the present investigation used a magnesium alloy in an extruded condition and all samples were processed through 5 turns, it is anticipated that the textures for all samples were in similar steady-state conditions so that the inuence of texture was negligible. This is consistent also with a recent investigation of the AZ31 alloy processed by ECAP where it was shown that the texture has no inuence on the size of the new grains produced during structural renement [45]. 5. Summary and conclusions 1. A magnesium AZ31 alloy was processed by high-pressure torsion (HPT) through 5 turns at a temperature of 453 K using rotation rates of 0.5, 1 and 2 rpm. 2. Processing by HPT gives signicant grain renement to an average size of 0.5 m with heterogeneous structures in the centers of the disks and homogeneous structures at distances at and above 1 mm from the disk centers. 3. The distributions of hardness over the disk surfaces are reasonably homogeneous after processing by HPT except only near the disk centers where there are slightly lower hardness values. 4. Since the local strain varies throughout the disk and the experiments used three different rotation speeds, it is possible to correlate both the measured values of the Vickers microhard-

Fig. 11. Values of the average grain size plotted as a function of the local strain rate at different distances from the centers of disks processed by HPT at selected rotation rates.

3608

P. Serre et al. / Materials Science and Engineering A 528 (2011) 36013608 [17] H. Jiang, Y.T. Zhu, D.P. Butt, I.V. Alexandrov, T.C. Lowe, Mater. Sci. Eng. A290 (2000) 128. [18] A.P. Zhilyaev, S. Lee, G.V. Nurislamova, R.Z. Valiev, T.G. Langdon, Scripta Mater. 44 (2001) 2753. [19] A. Vorhauer, R. Pippan, Scripta Mater. 51 (2004) 921. [20] A.P. Zhilyaev, K. Oh-ishi, T.G. Langdon, T.R. McNelley, Mater. Sci. Eng. A410 (2005) 277. [21] Z. Horita, T.G. Langdon, Mater. Sci. Eng. A410411 (2005) 422. [22] C. Xu, Z. Horita, T.G. Langdon, Acta Mater. 56 (2008) 5168. [23] K. Edalati, T. Fujioka, Z. Horita, Mater. Sci. Eng. A497 (2008) 168. [24] C. Xu, Z. Horita, T.G. Langdon, J. Mater. Sci. 43 (2008) 7286. [25] A. Loucif, R.B. Figueiredo, T. Baudin, F. Brisset, T.G. Langdon, Mater. Sci. Eng. A527 (2010) 4864. [26] R.B. Figueiredo, T.G. Langdon, J. Mater. Sci. 45 (2010) 4827. [27] K. Xia, J.T. Wang, X. Wu, G. Chen, M. Gurvan, Mater. Sci. Eng. A410411 (2005) 324. [28] S.E. Ion, F.J. Humphreys, S.H. White, Acta Metall. Mater. 30 (1982) 1909. [29] A. Galiyev, R. Kaibyshev, G. Gottstein, Acta Mater. 49 (2001) 1199. [30] T.G. Langdon, Mater. Sci. Eng. A462 (2007) 3. [31] R.B. Figueiredo, T.G. Langdon, J. Mater. Sci. 44 (2009) 4758. [32] C. Xu, Z. Horita, T.G. Langdon, Acta Mater. 55 (2007) 203. [33] M. Kawasaki, B. Ahn, T.G. Langdon, J. Mater. Sci. 45 (2010) 4583. [34] Z.C. Duan, X.Z. Liao, M. Kawasaki, R.B. Figueiredo, T.G. Langdon, J. Mater. Sci. 45 (2010) 4621. [35] C. Xu, Z. Horita, T.G. Langdon, Mater. Trans. 51 (2010) 2. [36] Y. Estrin, A. Molotnikov, C.H.J. Davies, R. Lapovok, J. Mech. Phys. Solids 56 (2008) 1186. [37] R.Z. Valiev, Yu.V. Ivanisenko, E.F. Rauch, B. Baudelet, Acta Mater. 44 (1996) 4705. [38] F. Wetscher, A. Vorhauer, R. Stock, R. Pippan, Mater. Sci. Eng. A387389 (2004) 809. [39] F. Wetscher, R. Pippan, S. Sturm, F. Kauffmann, C. Scheu, G. Dehm, Metall. Mater. Trans. 37A (2006) 1963. [40] K. Edalati, Z. Horita, Mater. Trans. 51 (2010) 1051. [41] S.X. Ding, C.P. Chang, P.W. Kao, Metall. Mater. Trans. 40A (2009) 415. [42] I.A. Maksoud, H. Ahmed, J. Rdel, Mater. Sci. Eng. A504 (2009) 40. [43] P.B. Berbon, M. Furukawa, Z. Horita, M. Nemoto, T.G. Langdon, Metall. Mater. Trans. 30A (1999) 1989. [44] B.J. Bonarski, E. Schaer, B. Mingler, W. Skrotzki, B. Mikulowski, M.J. Zehetbauer, J. Mater. Sci. 43 (2008) 7513. [45] J.A. del Valle, O.A. Ruano, Mater. Sci. Eng. A487 (2008) 473.

ness and the average grain sizes with estimates of the local strain rates. These correlations reveal that, with increasing strain rate, there is both a small increase in hardness and a small decrease in the average grain size. Acknowledgements One of the authors acknowledges support in the form of a summer fellowship from Phelma (PS). This work was supported in part by the National Science Foundation of the United States under Grant No. DMR-0855009. References
[1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] R.Z. Valiev, R.K. Islamgaliev, I.V. Alexandrov, Prog. Mater. Sci. 45 (2000) 103. R.Z. Valiev, T.G. Langdon, Prog. Mater. Sci. 51 (2006) 881. A.P. Zhilyaev, T.G. Langdon, Prog. Mater. Sci. 53 (2008) 893. A.P. Zhilyaev, B.K. Kim, G.V. Nurislamova, M.D. Bar, J.A. Szpunar, T.G. Langdon, Scripta Mater. 46 (2002) 575. A.P. Zhilyaev, G.V. Nurislamova, B.K. Kim, M.D. Bar, J.A. Szpunar, T.G. Langdon, Acta Mater. 51 (2003) 753. A.P. Zhilyaev, B.K. Kim, J.A. Szpunar, M.D. Bar, T.G. Langdon, Mater. Sci. Eng. A381 (2005) 377. G. Sakai, Z. Horita, T.G. Langdon, Mater. Sci. Eng. A393 (2005) 344. M. Kawasaki, T.G. Langdon, Mater. Sci. Eng. A498 (2008) 341. K. Edalati, Z. Horita, T.G. Langdon, Scripta Mater. 60 (2009) 9. M. Kawasaki, B. Ahn, T.G. Langdon, Acta Mater. 58 (2010) 919. M. Jane cek, M. Popov, M.G. Krieger, R.J. Hellmig, Y. Estrin, Mater. Sci. Eng. A462 (2007) 116. R. Lapovok, Y. Estrin, M.V. Popov, T.G. Langdon, Adv. Eng. Mater. 10 (2008) 429. S.H. Kang, Y.S. Lee, J.H. Lee, J. Mater. Process. Technol. 201 (2008) 436. Y. Cao, Y.B. Wang, S.N. Alhajeri, X.Z. Liao, W.L. Zheng, S.P. Ringer, T.G. Langdon, Y.T. Zhu, J. Mater. Sci. 45 (2010) 765. Y. Cao, M. Kawasaki, Y.B. Wang, S.N. Alhajeri, X.Z. Liao, W.L. Zheng, S.P. Ringer, Y.T. Zhu, T.G. Langdon, J. Mater. Sci. 45 (2010) 4545. Y.Z. Tian, X.H. An, S.D. Wu, Z.F. Zhang, R.B. Figueiredo, N. Gao, T.G. Langdon, Scripta Mater. 63 (2010) 65.

Вам также может понравиться