Sie sind auf Seite 1von 4

Available online at www.sciencedirect.

com

Scripta Materialia 58 (2008) 959962 www.elsevier.com/locate/scriptamat

Inuence of processing parameters on the bond toughness of roll-bonded aluminium strip


Wolz,b M. Homanb and M. Ferryb M.Z. Quadir,a,b,* Andre
a

Electron Microscope Unit, University Analytical Centre, Australian Microscopy and Microanalysis Research Facility (AMMRF), University of New South Wales, Kensington, Sydney, NSW 2052, Australia b ARC Centre of Excellence for Design in Light Metals, School of Materials Science and Engineering, University of New South Wales, Sydney, NSW 2052, Australia
Received 17 October 2007; revised 3 January 2008; accepted 15 January 2008 Available online 24 January 2008

The bond toughness of two adjacent aluminum sheets produced by roll bonding was investigated by measuring the critical strain energy release rate required to initiate a crack between them. It was found that higher rolling temperatures, higher rolling reduction and post-rolling recrystallization annealing improved the bond toughness. Higher rolling reduction also generated both simple and branched shear bands that helped to puncture the interfacial oxide layer and enhance bond toughness. 2008 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Keywords: Accumulative roll-bonding (ARB); Aluminum; Deformation structures; Shear bands; Recrystallization

Structural renement by severe plastic deformation (SPD) is a recent trend for fabricating advanced materials which produces a unique combination of properties. A large number of these techniques are under extensive investigation, with many of the outcomes of these works summarized in a recent review by Valiev and Langdon [1]. A promising method of SPD devised by Saito et al. [2] is accumulative roll-bonding (ARB), whereby stacked metal sheets are bonded during rolling to relatively high reductions (>50%). For a large number of roll-bonding cycles, ARB generates very high strains in the individual layers which generally results in substantial grain renement [1]. Since ARB is a relatively simple process using conventional rolling mills, it has been investigated on a large number of metals over the past few years [1]. A major factor in ascertaining the applicability of roll-bond systems is the bond strength of adjacent metal layers. The degree of metalmetal bonding during ARB has been considered by Saito and co-workers, based on the early work by Nicholas and Milner [3], where it was found that there is a threshold amount of rolling reduction, which is temperature dependent, below which a

* Corresponding author. Address: Electron Microscope Unit, University Analytical Centre, University of New South Wales, Kensington, Sydney, NSW 2052, Australia. Tel.: +61 2 9385 9752; fax: +61 2 9385 6400; e-mail: mzquadir@unsw.edu.au

sound bond is not possible. For temperatures less than half the homologous melting temperature, the critical strain is $50%; this also allows an accumulated strain during each ARB cycle without recrystallization. The recent expansion of ARB to a range of materials (see e.g. [47]) has prompted the present study on the inuence of rolling parameters on metalmetal bond quality. The aim of the present investigation is to determine the bond strength between two aluminium sheets using fracture principles [8]. Commercial purity aluminium plate was cold rolled to a 1 mm thick sheet and annealed for 40 min at 375 C to generate a fully recrystallized microstructure of grain size 70 30 lm throughout the sheet thickness. Samples of dimensions 105 32 1 mm3 were cleaned by ultrasonic vibration in an acetone bath and scrubbed using a stainless steel wire brush to remove any oxide. To maintain consistency between samples, wire brushing was carried out to generate evenly distributed scratches over the sample, as monitored using a low magnication optical microscope. Each sample was thoroughly dried and $25 mm of an end was painted using a dispersion of gypsum powder particles in water and dried at room temperature (Fig. 1a). The aim of this procedure is to prevent bonding in this region during subsequent rolling. The samples were stacked and rolled without lubrication in a two-high rolling mill (210 mm diameter rolls) using the variables shown in Table 1. An air circulating

1359-6462/$ - see front matter 2008 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved. doi:10.1016/j.scriptamat.2008.01.022

960

M. Z. Quadir et al. / Scripta Materialia 58 (2008) 959962

where Pmax is the load at which the crack starts to propagate, E is Youngs modulus of aluminium (70.6 GPa [8]), h and B are dimensional parameters shown in Figure 1b, and c is the crack length which can be obtained using the loading compliance of the specimen:  3 3=2 Eh BQ 2 c 8 where Q is the compliance of the sample prior to crack propagation, as shown in Figure 1c. The inuence of rolling temperature, for a given roll speed, on Gc (measure of toughness) is shown in Figure 2. While a broadening of the data occurs at high rolling temperatures, there is a denite increase in bond strength with rolling temperature. However, Figure 2 does not show clearly the eect of peripheral roll speed on bond strength. The importance of this parameter cannot be ignored, as it is clear that reduction of rolling temperature and increasing the roll speed results in the weakest bonding. For example, a roll speed of 20 m per minute (mpm) results in insignicant bonding after rolling at 300 C and only weak bonding is achieved at 350 C. Therefore, there might be a threshold value of roll speed, which cannot be dened from the data set, above which considerable bonding does not occur. A slower speed allows longer time of contact under pressure to establish bonding. Figure 3 shows toughness as a function of strain (measured by percentage of reduction) after rolling at 5 mpm and 300 C. For low reductions (up to 40%), the bond toughness is low but increases signicantly for reductions greater than 47%, although the data at high reductions is quite scattered, as indicated by the error bars. The eect of annealing after roll-bonding on bond toughness was also investigated. Figure 4 shows the eect of annealing time at 350 C on toughness for samples rolled to 50% reduction at 300 C. It can be seen that the maximum bond toughness is achieved after annealing for 20 min at this temperature. Roll-bonding was used to understand as a solid state diusion process, whereby atoms of the two adjacent surfaces diuse into each other under pressure [9]. The bond quality is inuenced by a number of interdependent parameters such as temperature, pressure (determined by the degree of reduction) and contact time (roll speed). Prior to roll bonding, sample surfaces need to be clean and in good contact. However, this is not easily achieved in aluminium alloys after heating to elevated temperatures in air (300450 C in this work), due to the increased rate of surface oxidation. Nevertheless, bonding occurs by fracturing of the oxide layers to leave fresh metal surfaces during rolling since the oxide has a much lower ductility than aluminium. It was found that the oxide also fractured during rolling due to the operation of two kinds of shear banding: (i) simple shear banding that breaks the oxide at single points at the interface and permits the metalmetal contact (Fig. 5a); and (ii) branched shear banding fragments the oxide layers (Figs. 5b and 6). Shear band branching was reported by Lee and Duggan [10] in rolled brass, and this phenomenon was found to be the cause of the river line structures that form on rolled sheet. Branched

Figure 1. (a) Surface preparation of the aluminium strips prior to rollbonding: part of the strip is coated with plaster paint with the other part remaining in the as-brushed condition. (b) Method of tensile testing to measure fracture toughness along the bonding line. (c) Typical loaddisplacement curve for a sample roll-bonded at 5 mpm at 400 C.

furnace was used for heating the samples for 5 min prior to rolling and also for subsequent annealing. For testing the bond strength, as-rolled samples were cut into strips of width 10 mm and length 80 mm, with the plastercoated region bent apart by 90 at the bondedunbonded interface (Fig. 1b). Since some deviation of both the target rolling reduction and nal sample dimensions is to be expected (i.e. 10 80 mm2), individual sample dimensions were used for calculating bond toughness. Figure 1b shows the testing procedure: each sample was clamped and loaded at a constant crosshead speed of 2 mm min1 in an Instron 1185 mechanical testing machine equipped with a 100 N load cell. Figure 1c shows a typical forcedisplacement curve for a sample rolled at 400 C in which crack initiation occurs at the maximum load, after which it does not increase any more. The loading condition used in this work is similar to the double-cantilever beam specimen, enabling the strain energy release rate, G, during fracture to be calculated [8]. The critical value strain energy release rate, equivalent to the toughness of the bond then be determined as Gc
2 12P 2 max c 3 Eh B

M. Z. Quadir et al. / Scripta Materialia 58 (2008) 959962 Table 1. Rolling variables used in the investigation Variable Peripheral roll speed (mpm) Rolling temperature (C) Nominal rolling reduction (%) Post-rolling annealing time at 350 C (min) Range 5, 10, 15, 20 300, 350, 400, 450 30, 40, 50, 60, 70 5, 10, 20, 40 Constant parameters

961

50% reduction 50% reduction Rolling temperature = 300 C at 5 mpm 50% reduction at 300 C at 5 mpm

Figure 2. Inuence of both rolling temperature and roll speed on bond toughness (samples rolled to 50% reduction).

Figure 3. Inuence of rolling reduction on bond toughness (samples rolled at 5 mpm and 300 C).

Figure 5. SEM ICC micrographs of a rolling directionnormal direction (RDND) section of the bonded interface after rolling to 70% reduction at 300 C. The oxide layer is broken in (a) by the operation of shear band and in (b) by shear band branching. The substructure in the adjacent areas is ne and equiaxed, and dierent from the remaining bulk materials having smeared microstructures.

Figure 4. Inuence of time during annealing at 350 C on bond toughness (samples rolled at 5 mpm at 300 C to 50% reduction).

Figure 6. SEM ICC micrograph of an RDND section of the bonded interface after rolling to 70% reduction at 300 C followed by annealing for 10 min at 350 C. The materials in between the oxide fragments are recrystallized as well as the bulk aluminium.

shear banding allows material to ow into the in-between gaps of the branches and aids in disintegrating the oxide lm; this process may be regarded as an eective method for improving bond quality.

In general, shear banding in rolled sheet is enhanced by lamellar structures of hard and soft alternating layers [11]. The micrograph developed in Figures 5a and b comprised dierent substructures near the bonding

962

M. Z. Quadir et al. / Scripta Materialia 58 (2008) 959962

region and away from the bonding line. Adjacent to the bounding line, the material comprised ne equiaxed blocks of sharp greyscale contrast, which is normally harder than the far region materials having a smeared appearance, which is an indicator of coarser substructures. This suggests that the interior substructure was produced from rolling and the surface substructure of each sheet was produced from brushing [1217] plus rolling. Since shear banding increases with the degree of rolling reduction [18], this phenomenon will also improve the bond toughness as indicated in Figure 3. For multiple cycles of ARB, the oxide lm is expected to substantially fragment into ne oxide particles, and their dispersion in the matrix may aid in material (i.e. bond) strengthening [19,20]. The strengthening phenomenon by incoherent Al2O3 has been recently overviewed by Karnesky et al. [21]. Figure 2 indicates that, once the oxide lm is eectively broken with rolling reduction, the bond toughness increases with increasing temperature due to an enhanced rate of diusion at the metal metal interface [5,9]. It has been shown recently by a molecular dynamics (MD) simulation of Chen et al. [22] that the eect of surface roughness is inuenced by temperature: higher temperatures enhances material ow and promotes the lling of voids created between any two rough surfaces in contact, thereby improving bonding. It is worth noting that the length scale of the surface roughness in the MD simulation is 25 nm, which is much smaller than the scratches generated in the present samples by wire brushing. Nevertheless some other experimental data of larger length scale produced outcomes consistent with the MD simulation of Chen et al. [22]. For example, Derby and Wallach [23] and Elzey and Wadley [24] found that microscale surface roughness improved bonding more eectively at high rolling temperatures. This implies that their MD simulation results are applicable for wider length scales. Finally, Figure 4 conrms that bond toughness is improved by recrystallization. The present work highlights an additional mechanism for enhancing bond toughness by recrystallizing the materials trapped in between the branches of shear bands in Figure 6: growth of those materials across the oxide boundaries creates a kind of composite-like strengthening eect from the unidirectional aligned structures of both oxides and aluminium generated after rolling. However, extensive grain growth is expected to have a minor negative eect on bond toughness.

The authors gratefully acknowledge the Australian Research Council (ARC) for supporting this work through the ARC Centre of Excellence for Design in Light Metals (CEO561574), and the Electron Microscope Unit of Analytical centre of University of New South Wales.
[1] R.Z. Valiev, T.G. Langdon, Prog. Mater. Sci. 51 (2006) 881. [2] Y. Saito, H. Utsunomiya, N. Tsuji, T. Sakai, Acta Mater. 47 (1999) 579. [3] M.G. Nicholas, D.R. Milner, Br. Weld. J. 8 (1961) 375. [4] N. Tsuji, Progress in Severe Plastic Deformation, in: B.S. Altan (Ed.), NOVA Publishers, New York, 2006, p. 545. [5] S. Ohsaki, S. Kato, N. Tsuji, T. Ohkubo, K. Hono, Acta Mater. 55 (2007) 2885. rez-Prado, O.A. Ruano, Mater. [6] J.A. del Valle, M.T. Pe Sci. Eng. A 410411 (2005) 353. [7] M.Z. Quadir, O. Al-Buhamad, L. Bassman, M. Ferry, Acta Mater. 55 (2007) 5438. [8] A.G. Atkins, Y.-W. Mai, Elastic and Plastic Fracture, John Wiley & Sons, Chichester, 2005. [9] W.A. Owczarski, D.F. Paulonis, Weld. J. 62 (1981) 22. [10] C.S. Lee, W.T. Hui, B.J. Duggan, Scripta Metall. Mater. 24 (1990) 752. [11] M.Z. Quadir, B.J. Duggan, ISIJ Int. 46 (2006) 1495. [12] M. Sato, N. Tsuji, Y. Minamino, Y. Koizumi, Mater. Sci. Forum. 426432 (2003) 753. [13] M. Sato, N. Tsuji, Y. Minamino, Y. Koizumi, Sci. Tech. Adv. Mater. 5 (2004) 45. [14] N. Takata, K. Yamada, K. Ikeda, F. Yoshida, H. Nakashima, N. Tsuji, Mater. Sci. Forum. 503504 (2006) 919. [15] B.L. Li, N. Tsuji, N. Kamikawa, Mater. Sci. Eng. A 423 (2006) 331. [16] M. Slamova, P. Homola, P. Slama, M. Karlik, M. Cieslar, Y. Ohara, N. Tsuji, Mater. Sci. Forum. 519521 (2006) 1227. [17] N. Takata, K. Yamada, K. Ikeda, F. Yoshida, H. Nakashima, N. Tsuji, Mater. Trans. 48 (2007) 2043. [18] M.R. Barnett, J.J. Jonas, ISIJ Int. 37 (1997) 697. [19] M.Y. Drits, Phys. Met. Metall. 57 (1984) 118. [20] M. Ferry, N. Burhan, Acta Mater. 55 (2007) 3479. [21] R.A. Karnesky, L. Meng, D.C. Dunand, Acta Mater. 55 (2007) 1299. [22] S. Chen, F. Ke, M. Zhou, Y. Bai, Acta Mater. 55 (2007) 3169. [23] B. Derby, E.R. Wallach, Metal Sci. 16 (1984) 49. [24] D.M. Elzey, H.N.G. Wadley, Acta Metall. Mater. 41 (1993) 2297.

Das könnte Ihnen auch gefallen