You are on page 1of 6

MATERIALS FORUM VOLUME 29 - Published 2005 Edited by J.F. Nie and M.

Barnett Institute of Materials Engineering Australasia Ltd

376

MULTISCALE NANOSTRUCTURED ALUMINIUM ALLOYS: PROCESSING, MICROSTRUCTURE AND PROPERTIES


B.Q. Han, D. Witkin, E.J. Lavernia Department of Chemical Engineering and Materials Science University of California, Davis, CA 95616, USA

ABSTRACT
Strength in nanostructured metals is observed to increase with decreasing grain size, however, low ductility commonly exists in nanostructured materials. Available experimental and theoretical studies attribute the low ductility of nanostructured materials to the deficit of dislocation activity. Recent research has suggested that nanocrystalline metals with a bimodal distribution of grain sizes, where submicron grains are distributed in a nanostructured matrix, have high ductility with limited sacrifice in strength. We describe an approach to achieve this grain size distribution in which two Al 5083 powder fractions are blended then consolidated and extruded. The nanostructured-grained and submicrongrained fractions result from cryomilled 5083 Al powders and spray-atomized Al 5083 powders without milling, respectively. Microstructural characteristics of the as-extruded 5083 Al alloys with fractions of 15, 30 and 50 percent unmilled powders are examined. Inspection of microstructure reveals a strong bond between nanostructured and micron-grained regions. The tensile strength decrease as a function of volume fractions of unmilled powder follows the simple rule of mixture. Tensile ductility increases slightly with increasing volume fraction of unmilled powder, however, at higher volume fractions of 30% and 50%, there is significant ductility increase. Enhanced tensile ductility was attributed to the occurrence of crack bridging as well as delamination between nanostructured and coarse-grained regions during plastic deformation.

1. INTRODUCTION
Large increases in strength are commonly observed in nanostructured metals, a phenomenon that is explained in part by grain refinement and the well-known HallPetch relationship. However, low ductility is a serious deficiency in most of these metals [1]. A low ductility, and the associated loss of toughness, represents a serious drawback in many engineering applications. Among the various approaches that are currently being investigated to generate nanostructured materials, cryomilling affords unusual flexibility in microstructural design along with the ability to produce bulk samples [2]. As with other methods for producing nanostructured alloys, decreasing grain size is accompanied by a drastic loss of ductility [3]. In an earlier study, Tellkamp and co-workers reported on the mechanical properties of a nanostructured cryomilled 5083 Al alloy, with a yield strength of 334 MPa, ultimate strength of 462 MPa and an elongation of 8.4 % [2]. They suggested that the presence of coarse grains in the nanocrystalline matrix was responsible for the observed ductility, an observation that sparked the genesis of multi-scale materials. In this particular case, the coarse grains arose as a result of thermomechanical processing of the nanostructured powders, i.e., consolidation and extrusion. Under tensile loading, microcracks nucleated in nanocrystalline regions and propagated to ductile coarse-grain regions, where they were effectively arrested by a combination of crack blunting and crack bridging.

If a few coarse grains are added to a nanostructured matrix, the material will have increased ductility while maintaining high strength. On the basis of this consideration, and in an effort to elucidate the underlying mechanisms, several cryomilled 5083 Al alloys with different volume fractions of unmilled coarse-grain 5083 Al were produced. The same procedures were used that have been employed previously to make fully cryomilled Al alloys. The microstructural characteristics, mechanical behavior and deformation mechanisms of these materials were explored in the present study.

2. EXPERIMENTAL PROCEDURES
Spray-atomized 5083 Al alloy powders were mechanically milled in a liquid nitrogen slurry for eight hours to obtain the nanostructured grains in microstructure [6]. An illustration of cryomilling and the morphology of cryomilled Al powders after cryomilling for 8 hours are shown in Figures 1 (a) and (b), respectively. The cryomilled powder had a grain size of ~20 nm, as determined by X-ray diffraction and TEM. The bulk cryomilled 5083 Al alloys (Al4.2wt%Mg-0.67wt%Mn) blended with 15 %, 30 % and 50 % unmilled powders were produced following procedures similar to those used previously for nanocrystalline Al alloys [2, 4, 5]. The cryomilled powders were blended mechanically with different volume fractions of unmilled 5083 powders (15 %, 30 % and 50 %) in an inert atmosphere to achieve a uniform distribution of unmilled powders. The powders used for cryomilling and for the unmilled

377

additions were from the same spray-atomized batch. The various blends of powders were then canned, also under an inert atmosphere. The mixture compacts were vacuum degassed at a temperature of 673 K. Consolidation of the canned powders was completed by cold isostatic pressing at a pressure of ~ 400 MPa. To remove any remaining porosity and improve mechanical properties, the consolidated billets were extruded into a round bar with a diameter of 19.05 mm at a temperature of 823 K. Some of the extruded 5083 Al alloys were further swaged at a temperature of 723 K. The swaged 5083 Al alloys had a final diameter of 12.7 mm.

correspond to the unmilled powder particles, which were originally spherical and 440 m in diameter. In the 15% sample, the unmilled particles were elongated in the extrusion direction, but individual particles did not agglomerate or coalesce. In contrast, the 30% sample showed evidence that some unmilled powder particles coalesced, forming bands that were slightly thicker in the transverse direction. In the 50% sample, the coalescence is even broader. The unmilled regions are almost connected together by forming continuous bands.

(a)

(a)

(b) (b) Figure 1: (a) Illustration of cryomilling in a liquid nitrogen. (b) Morphology of cryomilled Al powders after cryomilling for 8 hours. The as-extruded materials were machined into flat dogbone tensile specimens along the extrusion direction, with a gauge length of 12 mm, width of 4 mm and thickness of 2 mm. Tension testing was performed using an INSTRON 8801 universal testing machine (Canton, MA). Tensile specimens were tested at a constant crosshead velocity of 0.012 mm/sec until failure, with direct measurement of the displacement of the gauge section by a dual-camera video extensometer. The fracture surface of the tensile specimens was studied using a scanning electron microscope (SEM) operated at 10 kV. Tensile tests of the as-swaged materials were conducted at ambient temperatures via using dog-bone tensile specimens with a gauge length of 19 mm and a diameter of 4.6 mm, in a load frame at a constant crosshead velocity of an initial strain rate of 0.016 mm/s.

(c)

(d)

3. RESULTS AND DISCUSSION


(e) The optical microstructure of the bimodal cryomilled 5083 Al alloys is shown in Figure 2 (a) - (f), revealing the dispersion of unmilled powders in the cryomilled matrix after extrusion processing. Lighter regions

378

800

600 True stress (MPa) Cryomilled 5083 Al extrusion . = 10-3 s-1 Vol. % of unmilled 0 0 15 30 0 0.01 True strain 0.02 0.03

400

(f)

200

Figure 3: Tensile strain-stress curves of the asextruded bimodal 5083 Al alloys. The fracture surfaces of tensile specimens were examined using SEM, as shown in Figure 4. Several distinctive characteristics were noted while examining the sample containing 30 % coarse grains. First was the existence of a poorly defined dimple morphology in the cryomilled regions, which has been noted in similar cryomilled Al alloys [4]. Second is the appearance of well-defined dimples associated with the coarse-grain regions, which are similar to those seen in conventional Al alloys. Necking occurred in the unmilled coarsegrain regions, indicating local ductile failure in these regions. Third are the areas of delamination at interfaces between the coarse-grain regions and the nanostructured regions. Since there is no interfacial debonding in the as-extruded microstructure as observed in TEM, the areas of delamination at interfaces must be formed during plastic deformation. It is noteworthy that delamination developed around the coarse-grained regions is greater in one side than in the other. This may suggest that microcracks nucleate in one side of the coarse-grained regions and propagate to the other side.

(g) Figure 2: Optical micrographs of cryomilled plus 15% unmilled 5083Al (a, b), cryomilled plus 30% unmilled 5083Al (c, d) and cryomilled plus 50% unmilled 5083 Al (e, f). Extrusion directions are indicated on each figure. (g) The TEM microstructure of the as-extruded 5083 Al alloys with 50 % of unmilled 5083 Al in a longitudinal section. The number of coarse grains with a grain size of approximately 1000 nm increases with increasing volume fraction of unmilled powder. In addition, the unmilled coarse grains are distributed from discretely to continuously along the extrusion direction as the volume fraction is increased from 15 to 50 %. When viewed normal to the extrusion direction (longitudinal sections in Figure 2 (g)), coarse-grain bands extend in the extrusion direction. The mean grain size in TEM measured from approximately 200 grains is 207 nm for the fully-cryomilled 5083 Al. Equiaxed grains are distributed uniformly in the nanostructured regions of the as-extruded alloy. Inspection of microstructure reveals a strong bond between nanostructured and micron-grained regions. The tensile behavior of the as-extruded bimodal alloys is shown in Figure 3. The fully cryomilled 5083 Al alloy showed a high yield strength, followed by a brief work hardening region, but the specimen failed abruptly without necking. The cryomilled 5083 Al alloys with 15 % and 30 % coarse-grain fraction showed a slightly lower strength and improved ductility, as manifested by the appearance of a longer work hardening region and a flow-stress plateau after the ultimate tensile strength. Obvious necking does not occur, but a Lders band was observed in the middle of the gauge section and final failure occurred at one end of the Lders band in the bimodal alloys. It appears that the Lders band formed immediately after the maximum load, as manifested by the appearance of stress-drop periods. In the subsequent deformation, intense plastic deformation was localized in the vicinity of the Lders band.

Figure 4: Fracture surface of the as-extruded 5083 Al alloy containing 30% coarse-grained 5083 Al. The tensile behavior of alloy samples in the as-swaged condition was similar to samples in the as-extruded condition. Figure 5 (a) shows the stress-strain response of samples with different coarse-grain volume fractions, while Figure 5 (b) summarizes the effect of coarse-grain volume fraction on the as-swaged cryomilled 5083 Al alloys. For the materials with low additions (15 % or 30 %) of unmilled powder, the strength decreases and ductility increases were approximately linear. However, addition of 50 % unmilled powder caused a nonlinear strength decrease

379

and ductility increases. The as-swaged alloy processed from 100% cryomilled powders exhibited a yield strength and ultimate tensile strength similar to those of the as-extruded alloy, but the elongation in the former alloy was slightly improved. After the UTS was achieved, a brief stress drop occurred, followed by abrupt failure. For the as-swaged alloy with 15 % coarse-grain fraction, the yield strength was slightly reduced and reached a value of 684 MPa. Compared to the all-cryomilled extrusion, there was a slight increase in both elongation and uniform strain, although the ultimate tensile strength was virtually the same. The as-swaged alloy with 30 % coarse-grain fraction exhibited a linear decrease in yield strength and ultimate tensile strength, and a linear increase in elongation. The alloy containing 50% unmilled powder showed a significant decrease in strength, accompanied by a marked increase in elongation.
800

3.1 Tensile Behavior The brief period of initial work hardening in the cryomilled 5083 Al alloys resembles behavior observed in other Al-Mg alloys and pure Al [7, 8]. In these earlier studies, the behavior was attributed to multiplication of dislocations. Dislocation activity is expected to be difficult in cryomilled Al alloys given the complex microstructure, which includes supersaturated Mg solute, segregation of solute and impurities to grain boundaries, and the possible existence of nanoscale oxide/nitride dispersoids in the matrix [2, 4, 6]. However, dislocation multiplication over short distances might be possible under sufficiently high-applied stresses (well above the yield stress for conventional aluminium). The grain size is a few hundred nanometers, which exceeds the critical grain size of several tens of nanometers considered a minimum for the viability of the dislocation activity [9]. The number of dislocations in a pileup is directly proportional to the length of the pileup. However, in the grain size regime of the fully cryomilled nanostructured 5083 Al alloy, the number of dislocations in a pileup is expected to be small. If only the number of dislocations is considered, a low stress concentration at the front of pileups is expected (tip = n). If this stress exceeds the critical stress (c) required to move the leading dislocation through the boundary, the leading dislocation will overcome obstacles and initiate plastic deformation [10]. Therefore, smaller grain sizes result in larger yield strengths. When the leading dislocation starts to move and propagates the plastic deformation into adjacent grains, the density of dislocations in the nanostructured region increases, causing work hardening. However, the work-hardening region due to dislocation glide is limited by the formation of microcracks in the nanostructured regions under the high stress level. In order to provide substantial flow, the stress concentration due to dislocation glide has to be accommodated. The alloys in the present study were designed to have a hybrid microstructure consisting of relatively hard and soft regions. With the addition of a small fraction of unmilled coarse-grained powder to the cryomilled 5083 Al alloy, the majority of the resulting coarse grains are dispersed discretely within the cryomilled nanostructured matrix. The coarse-grain regions are elongated and distributed parallel to the extrusion direction. Because the unmilled coarse-grained regions consist of larger grain sizes and lack strengthening impurities, these regions are much softer than the nanocrystalline regions. Microhardness measurements performed on the two discrete regions indicated that the nanocrystalline region was approximately 50 percent harder than the coarse-grained regions [11]. During tensile tests, dislocations build up in both nanostructured and coarse-grain regions. Because of the large grain size and fewer obstacles to slip, dislocations are more easily relaxed by slip in the coarse-grain regions than in the nanostructured regions.

600 True stress (MPa)

400

Cryomilled 5083 Al alloys Unmilled 5083 Al

200

0% 15 % 30 % 50 % 0 0.02 True strain 0.04 0.06

(a)
800 10

8 600 Strength (MPa) 400


0.2

Strain (%)

UTS f u 4

200 2

0 0 10 20 30 Volume (%) 40 50

(b) Figure 5: (a) Tensile strain-stress curves and (b) volume fraction dependence of strength and ductility of as-swaged bimodal 5083 Al alloys. Nevertheless, there was no significant change in the uniform strain of the as-swaged alloy containing 50 % unmilled powder (Fig. 5 (b)). The uniform strain had a weak linear dependence on the volume fraction of unmilled powders. The significant increase in ductility of the 50 % coarse-grain alloy was attributed to the extended saturated flow stress after reaching the UTS.

380

In terms of load transfer, the nanocrystalline regions sustain most of applied stress and only a small part of the load is transferred to the softer coarse-grained regions. This produces a slight decrease in yield strength. In addition, stress concentrations in the nanostructured regions may be relaxed by transferring local loads to the softer coarse-grained regions, causing flow in the softer coarse-grained regions [12]. These coarse-grained regions are more compliant, not only because of their larger grains, but also because they lack the impurities (nitrogen, oxygen, carbon, iron, etc.) normally associated with cryomilled powders [2, 13]. The absence of impurities is likely to facilitate glide in the coarse-grained, causing increased flow during tensile tests. As the volume fraction of coarse-grained regions is increased, the extrusion-induced flow deformation leads to more contact between coarse-grain regions and in some cases, coalescence of adjacent particles. The coarse-grained regions become semi-continuous and interpenetrating, resembling the nanostructured regions in distribution and morphology. Consequently, in the 50-50 blended material, both coarse-grained and nanocrystalline regions carry the applied loads. The coarse-grained regions will undoubtedly yield at lower stress levels, causing flow and relaxation of local stresses, thereby decreasing stress levels in the nanostructured regions, which continue to deform elastically. As work hardening initiates in the coarsegrained regions, the nanostructured regions will eventually experience plastic flow, although at much higher stress levels. Thus, the observed decrease in yield strength of the bimodal Al alloys with increasing coarse-grain fractions is attributed to the load transfer to softer constituents. The relationship between yield strength (0.2) and volume fraction (Vu) of unmilled powders can be approximated by a simple rule of mixtures, 0.2 = (1 Vu)c + Vuu, where c and u are the yield stress of 5083 Al alloys processed with 100 % cryomilled powders and 100 % unmilled powders, respectively. The work-hardening region after yielding is brief in the all-cryomilled 5083 Al sample whereas it is increased in the bimodal alloys. Although the uniform deformation has a weak dependence on the volume fraction of coarse grains, the work hardening exponent in the work-hardening region decreases with increasing volume coarse-grained fraction. After the maximum stress is reached, the stress typically drops, and the work-hardening rate diminishes. The slight drop in stress suggests a plastic instability, possibly associated with the formation of necking and/or cavities. The decreased work-hardening exponent may depress damage tolerance by limiting the volume of stress concentration ahead of a crack. Therefore, stress drop after the ultimate tensile strength as well as the Lders effect is observed in the tensile deformation of bimodal 5083 Al alloys. A similar stress drop phenomenon has been reported in other mechanically-alloyed (MA) fine-grained Al alloys [4, 14-16]. For example, in an as-extruded MA

Al-4Mg-1.5O-0.8C alloy [14], characterized by a complex network of fine dislocation cells, the stress drop was attributed to the collective movement of large numbers of mobile dislocations that were previously pinned. The subsequent flow stress plateau was attributed to the reduced resistance to glide of unpinned dislocations. 3.2 Enhanced Tensile Ductility The concept of ductile-phase toughening brittle materials is well-established, and includes ductile inclusions of particles, fibers or layers [17-23]. By shielding the crack tip from applied stress, higher levels of remote stresses can be applied to a ductilebrittle composite material before fracture-critical conditions are reached. Fiber-reinforced composites generally promote greater levels of toughening than particulate-reinforced composites, and ductile-layer reinforced laminates lead to the greatest enhancement of toughening. Metal matrix composites usually have lower fracture toughness and low ductility than unreinforced matrix materials although they possess higher specific stiffness and strength. Laminated metal composites with ductile layers are developed for the purpose of improvement of the ductility and toughness [22, 23]. For instance, the fracture toughness of a laminated aluminium alloy composite with alternate layers of 15 % SiC particulate-reinforced 7093 Al composites and monolithic 3003 Al-Mn alloy is 79 % higher than that of the control 15 % SiC particulatereinforced 7093 Al composites. Several toughening mechanisms have been proposed to explain ductile-phase toughening of composite microstructures, such as crack bridging, crack blunting, crack deflection, stress distribution of crack tip, crack front convolution and local plane stress deformation, etc. [22, 24]. The microstructure of the present bimodal alloys is analogous to short-fiber metal matrix composites, in that the coarse-grain bands are distributed parallel to the extrusion direction. The bands tend to be discontinuous when the volume fraction of coarse gains is low, and continuous when the volume fraction of coarse gains is high. However, because the composition of the constituents is nearly the same, there is no mismatch in thermal expansion coefficient between the two regions in the bimodal cryomilled Al alloys. Consequently, the strength difference between two regions arises mainly from the grain size difference and Orowan strengthening difference. On the basis of microstructural characteristics observed in the present study, the concept of crack blunting and delamination is modified to explain the enhanced ductility in the bimodal alloys with coarse-grains. In the cryomilled Al alloys with coarse-grain bands, microcracks are expected to nucleate in the nanostructured regions and propagate along grain boundaries. When a microcrack meets a coarse-grain band, the band will retard propagation by blunting the crack and by delamination of interfaces between

381

coarse-grain and fine-grain regions. When more dislocations are emitted into the coarse grain, a new slip surface may be formed, eventually leading to necking and cavitation within the coarse-grain bands. Finally, dimples on the coarse-grained regions and delamination at interfaces will be generated on fracture surface. For crack bridging to occur, the bridging of ductile phase ligaments must have sufficient ductility to resist fracture at or ahead of the advanced crack tip. Bridging ligaments will restrict crack openings and undergo plastic stretching, thus promoting shielding of the crack tip. An interesting aspect of the process is the constraint of the ductile bands within the harder constituent, which may have implications for the kinetics of the multi-step crack growth process. If delamination occurs in the vicinity of one end part of the discontinuous coarse-grain bands, the final failure of the bands will not produce dimples. The delamination at interfaces and the necking deformation of ductile coarse-grain regions will cause significant energy loss, resulting in an enhanced tensile ductility.

to Dr. Z. Lee and Prof. S. Nutt for their contribution to the projects.

REFERENCES
1. 2. C. C. Koch, D. G. Morris, K. Lu, and A. Inoue: MRS Bulletin, 1999, vol. February, pp. 54-58. V. L. Tellkamp, A. Melmed, and E. J. Lavernia: Metall. Mater. Trans. A, 2001, vol. 32A, pp. 23352343. B. Q. Han, F. A. Mohamed, and E. J. Lavernia: J. Mater. Sci., 2003, vol. 38, pp. 3319-3324. B. Q. Han, Z. Lee, S. R. Nutt, E. J. Lavernia, and F. A. Mohamed: Metall. Mater. Trans. A, 2003, vol. 34A, pp. 603-613. D. Witkin, Z. Lee, R. Rodriguez, S. Nutt, and E. Lavernia: Scri. mater., 2003, vol. 49, pp. 297-302. F. Zhou, X. Z. Liao, Y. T. Zhu, S. Dallek, and E. J. Lavernia: Acta Materialia, 2003, vol. 51, pp. 2777. M. Verdier, Y. Brechet, and P. Guyot: Acta Mater., 1999, vol. 47, pp. 127-134. D. Kuhlmann-Wilsdorf: Phil. Mag. A, 1999, vol. 79, pp. 955-1008. T. G. Nieh, and J. Wadsworth: Scrip. metall. mater., 1991, vol. 25, pp. 955-958. J. C. M. Li, and Y. T. Chou: Metall. Trans., 1970, vol. 1, pp. 1145-1159. Z. Lee, D. B. Witkin, E. J. Lavernia, and S. R. Nutt: in MRS Proc., eds., Materials Research Society, 2003, pp. I1.7.1-6. Z. Lee, R. Rodriguez, R. W. Hayes, E. J. Lavernia, and S. R. Nutt: Metall. Mater. Trans. A, 2003, vol. 34A, pp. 1473-1481. F. Zhou, J. Lee, S. Dallek, and E. J. Lavernia: J. Mater. Res., 2001, vol. 16, pp. 3451-3458. Y. W. Kim, and L. R. Bidwell: Scri. metall., 1982, vol. 16, pp. 799-802. T. Mukai, K. Ishikawa, and K. Higashi: Mate. Sci. Eng., 1995, vol. A204, pp. 12-18. H. R. Last, and R. K. Garrett: Metall. Mater. Trans. A, 1996, vol. 27A, pp. 737-745. L. S. Sigl, P. A. Mataga, B. J. Dalgleish, R. M. McMeeking, and A. G. Evans: Acta Metall., 1988, vol. 36, pp. 945-953. P. A. Mataga: Acta Metall., 1989, vol. 37, p. 3349. M. Bannister, H. Shercliff, G. Bao, F. Zok, and M. F. Ashby: Acta Metall., 1992, vol. 40, p. 1531. M. Hoffman, B. Fiedler, T. Emmel, H. Prielipp, N. Claussen, D. Gross, and J. Rodel: Acta Metall., 1997, vol. 45, pp. 3609-3623. W. O. Soboyejo, K. T. V. Rao, S. M. L. Sastry, and R. O. Ritchie: Metall. Trans. A, 1993, vol. 24A, pp. 585. D. R. Lesuer, C. K. Syn, O. D. Sherby, J. Wadsworth, J. J. Lewandowski, and J. W.H. Hunt: Inter. Mater. Rev., 1996, vol. 41, pp. 169-197. A. B. Pandey, B. S. Majumdar, and D. B. Miracle: Acta Mater., 2001, vol. 49, pp. 405-417. W. Soboyejo: Mechanical properties of engineered materials, Marcel Dekker, Inc., New York, 2003.

3. 4.

5. 6. 7. 8. 9.

4. CONCLUSIONS
Cryomilled 5083 Al alloys blended with coarse-grain volume fractions of 15 %, 30 % and 50 % were produced by consolidation of a blended mixtures of cryomilled and unmilled powders. Microstructural characteristics and mechanical properties were investigated to assess the possible combinations of strength and ductility. A bimodal grain size was achieved in the as-extruded cryomilled 5083 Al alloys with a grain size of 200 nm in nanostructured regions and 1 m in coarse-grained regions. A tensile yield strength of 700 MPa was obtained in the fully cryomilled 5083 Al alloy, which compares favorably with the yield strength of conventional 5083 (145 MPa). As the volume fraction of coarse grains was increased, strength decreased slightly, but ductility increased. An enhanced tensile elongation associated with the occurrence of a Lders band was observed in bimodal cryomilled 5083 Al alloys. Enhanced tensile ductility was attributed to the occurrence of crack bridging as well as delamination during the plastic deformation. The results of the present study indicate a process route for the design of microstructures with selected distributions of grain sizes. When the strengthening achieved by grain refinement is coupled with the ductility afforded by incorporation of coarser grains, one can achieve combinations of strength, ductility, and toughness that were previously not possible. A challenge for future work will involve identifying microstructures and process parameters to optimize alloy performance.

10. 11.

12.

13. 14. 15. 16. 17.

18. 19. 20.

21.

22.

23.

ACKNOWLEDGMENTS
24. Support from the Office of Naval Research (grant numbers N00014-03-1-0149 and N00014-03-C-0163) is gratefully acknowledged. Appreciation is extended

You might also like