Sie sind auf Seite 1von 7

Materials Science and Engineering A254 (1998) 83 89

Microstructure of hot-deformed Cu3.45 wt.% Ti alloy


A.A. Hameda a, L. Blaz b,*
b a Al-fateh Uni6ersity, PO Box 13292, Tripoli, Libya Uni6ersity of Mining and Metallurgy, Al. Mickiewicza 30, 30 -059 Krakow, Poland

Received 21 January 1998; received in revised form 18 May 1998

Abstract Hot compression tests at temperatures ranging from 673 to 1173 K were carried out on solution-treated Cu 3.45 wt.% Ti at a constant true strain rate of 1.410 4 s 1. According to literature data, an effective hardening of Cu Ti alloys results from continuous precipitation of small metastable i% particles of Cu4Ti-phase which follows the spinodal decomposition of the solid solution during static aging of a solution treated alloy. This process was also responsible for an effective hardening of the material during hot compression of the solution treated alloy in the early stages of the test. At low deformation temperatures (673773 K), shear banding and growth of intergranular cavities were found to be responsible for a ow stress decrease and for sample fracture at large strains. During hot deformation at intermediate temperatures (873 973 K), softening was also found to follow the ow stress maximum. According to the literature, the discontinuous precipitation of stable i particles of Cu3Ti-phase results in a decrease of strength during static aging, i.e. an overaging effect. However, during hot deformation of a solution treated alloy in this temperature range dynamic precipitation coarsening within shear bands is also observed to proceed in competition with cellular growth of i particles. Phase transformation of ne-scale precipitates into colonies of coarse Cu3Ti particles produces a soft structural component within the hardened matrix. This as well as intensive shear banding was responsible for the effective ow softening of the material. Moreover, discontinuous precipitation at grain boundaries is found to reduce the growth of cavities and to delay intergranular fracture. Thus, samples deformed at temperatures from 973 to 1023 K did not fracture in the deformation range used in the experiments. It is suggested that intergranular crack growth is retarded by dynamic recovery within shear bands as well as by dynamic recrystallization above the solvus temperature. 1998 Elsevier Science S.A. All rights reserved.
Keywords: Dynamic precipitation; Shear band; Dynamic coarsening; Flow softening; Copper titanium alloy

1. Introduction Addition of titanium to copper (up to 5 wt.% Ti) is used to produce age-hardenable alloys which are characterized by a very effective precipitation hardening behavior. Structural processes which are observed during static aging of Cu Ti alloys were reported to follow three distinguishable steps [1 3]: spinodal decomposition with titanium content uctuations along 100 directions; ne precipitation of coherent or semi-coherent metastable i% particles of the ordered Cu4Ti phase having the matrix/particle orientation of (001) // (001)%, [310] // [100]% and [130] // [010]% (i% is a body centered tetragonal structure (D1a) with dimensions a = 0.584 nm and c =0.362 nm);
* Corresponding author.

discontinuous precipitation of the equilibrium Cu3Ti phase. The structure of i particles of Cu3Ti-type was reported as an orthorhombic one with dimensions of a= 0.453 nm, b= 0.435 nm and c= 0.517 nm. The matrix/particle orientation relationship is [111] // [010]%and [110] // [001]%. The hardness maximum of CuTi alloys was reported to be associated with continuous nucleation and growth of very ne i% particles [4,5]. The material hardness decrease, i.e. the over aging effect, was reported to result from discontinuous precipitation of stable i particles. Cellular growth (CG) of plate-like i particles usually starts at high-angle boundaries. It was reported that some structural defects, such as subboundaries, dislocation tangles and shear bands, may produce additional sites for nucleation of discontinuous precipitation [58]. Thus, it seems reasonable that these additional nucleation sites may promote the material

0921-5093/98/ - see front matter 1998 Elsevier Science S.A. All rights reserved. PII S0921-5093(98)00753-9

84

A.A. Hameda, L. Blaz / Materials Science and Engineering A254 (1998) 8389

softening during hot deformation of the solution treated alloy as they promote the nucleation and discontinuous growth of coarse i particles. It has been observed that the ow softening of some materials during hot deformation may also result from dynamic recrystallization (DRX) and/or ow localization [8,9]. More recent results, reported for Cu Ni CrSiMg alloy have indicated a signicant interaction between dynamic precipitation and ow localization during hot deformation tests [10]. The ow stressstrain behavior for that alloy was affected and controlled by shear banding and dynamic coarsening within shear bands. Discontinuous precipitation in the alloy mentioned above was limited to a very narrow zone of the matrix nearby preferably oriented grain boundaries. It is worth mentioning that discontinuous precipitation was never observed for Cu Ni Cr Si Mg alloy in the statically aged conditions. Precipitation processes during static aging of some CuTi alloys may follow similar aging sequences to those observed in Cu Ni Cr Si Mg alloys. However, in contrast to the Cu Ni Cr Si Mg alloy, discontinuous precipitation in Cu Ti alloys with high titanium content may fully transform the material structure under both static and dynamic conditions. Thus, the purpose of the experiments described below was to study structural changes that take place during hot deformation of supersaturated Cu 3.45 wt.% Ti alloy with respect to ow stress strain behavior at various deformation temperatures. A wide range of deformation temperatures was used to investigate the strain/ precipitation interactions as well as their effect on the nal material structure with respect to the kind of precipitation mode and the morphology of growing particles.

2. Experimental The experiments were carried out on CuTi alloy with chemical composition given in Table 1. Vacuum cast ingots were homogenized at 1173 K and then cold-rolled, with intermediate annealing at 1173 K. Cylindrical samples, 9.4 mm in diameter and 12 mm in height, were machined from cold rolled rods. Convex depressions 0.2 mm deep were machined on both ends of the sample in order to maintain the graphite lubricant during compression tests. Then, the specimens were solution treated for 1 h at 1173 K and water quenched. A set of the samples was also over aged by furnace cooling in order to test a reference material having a stable structure during testing at intermediate deformation temperatures. Compression tests at temperatures from 673 to 1173 K were carried out on an MTS machine equipped with a furnace and fast cooling system. The samples were water quenched within less than 3 s after deformation had stopped. The samples were subsequently sectioned along the compression axis using a spark erosion machine and prepared for structural observations by optical and transmission electron microscopy. Transmission electron microscopy observations were carried out on a JEM 2010 ARP 200 kV microscope equipped with a scanning transmission electron microscopy device and a LINK-PENTAFET X-ray analysis system.

3. Results and discussion Typical ow strain-stress curves obtained during hot compression of solution treated (ST) Cu3.45 wt.% Ti alloy at a strain rate of 1.4 10 4 s 1 are shown in Fig. 1 (solid lines). The total aging time, which includes 3 min of temperature stabilization and : 7 min of compression test, was not long enough to detect any precipitation hardening during static aging experiments at 673 K [11]. Thus, during the compression test at 673 K, continuous strain hardening was observed over the entire range of deformation. The structure of a sample deformed at 673 K is shown in Fig. 2(a). Coarse slip bands were observed in the grain interiors at the beginning of deformation with shear band (SB) development at larger strains. The stressstrain curves obtained at 773 K, exhibit a peak stress in the initial stage of deformation, after which the ow stress decreased monotonically until the sample fractured at the strain of mt = 0.20.3. Microstructural observations revealed SB development and growth of intergranular cavities which were responsible for the sample fracture at larger strains (Fig. 2(b)). Occurrence of cracking became more restricted with increased deformation temperature. Neither intergranular cracks nor fractures of the specimens were observed

Table 1 Chemical composition of the alloy used in the experiments Element P Ag Sb Zn Fe As S Cd Pb Ni Sn Ti Cu ppm 83 B1 18 B1 37 12 56 5 B1 68 1 3.45 In balance wt.%

A.A. Hameda, L. Blaz / Materials Science and Engineering A254 (1998) 8389

85

pearlitic-like structure due to discontinuous growth of i particles. The nal structure of hot deformed samples was transformed to 8595% by the discontinuous precipitation process (Fig. 4(a)). Pronounced deformation of lamellar particles during testing resulted in their fragmentation and spheroidization. Thus, the wavy shaped lamellae and quasi spherical i-particles were observed in hot deformed samples whereas broad diffuse shear bands were hardly distinguishable in the material structure (Fig. 4(b)). Such structural processes are considered to control the ow softening of OA samples during hot deformation at the same temperature (Fig. 1). Discontinuous growth of i particles was found to be a very efcient structural softening process during compression testing at 1023 K. Three structural components could be observed in these hot deformed samples (Fig. 5(a)): coarsened particles (B), spherical in shape, plate-like particles (C) in the grain interiors resulting from cellular growth and discontinuously grown of aligned single particles (A). Chemical compositions of coarse particles by the X-ray analysis revealed titanium contents of : 2333 at.%, thus conrming the literature data which suggest the development of the Cu3Ti-

Fig. 1. True stress strain curves for ST Cu3.45 wt.% Ti alloy deformed at strain rate of 1.4 10 4 s 1 (deformation temperatures are marked in the gure). Some stressstrain curves for OA material are included for comparison.

during hot deformation at 873 K up to strains of 0.3 (Fig. 3(a)) despite the observation of intensive ow softening during hot deformation (Fig. 1). Cavity growth within shear bands and at grain boundaries appears to have been limited due to cellular growth and intensive particle coarsening within SBs. Highly recovered dislocation structures and dynamically coarsened particles within SBs were revealed by TEM and STEM structure analysis (Fig. 3(b)). Thus, it seems reasonable to conclude that the interaction of the non-uniform deformation and particle coarsening within shear bands, accompanied by cellular growth of i particles at grain boundaries, were responsible for the material ow softening during hot compression tests. Flow stress decrease during compression tests at the temperature of 973 K is ascribed to intensive cellular growth of i particles rather than non-homogeneous deformation processes. The time for temperature stabilization before compression testing was long enough to transform : 30% of the material volume and produce a

Fig. 2. Microstructure of ST Cu 3.45 wt.% Ti alloy deformed at: (a) 673 K and strain rate 1.410 4 s 1, mt =0.34; (b) 773 K and strain rate 1.410 4 s 1, mt =0.17. Shear band position is marked in the gure (SB).

86

A.A. Hameda, L. Blaz / Materials Science and Engineering A254 (1998) 8389

Fig. 3. Microstructure of ST Cu3.45 wt.% Ti alloy deformed at the temperature of 873 K and strain rate of 1.4 10 4 s 1, mt =0.31: (a) optical microscopy; (b) TEM microstructure and titanium distribution revealed by STEM and Ti (Kh ) X-ray mapping within the area marked in the gure. Continuous precipitation on the left, and discontinuous precipitation on the right side of shear band are shown (b); CG, cellular growth, SB, shear bands.

type phase. The precipitation processes mentioned above had been nished practically at the beginning of compression test and did not result in the material hardening during hot deformation tests. Thus, in general, the ow stress value, was controlled only by dynamic recovery processes. Above the solvus temperature, i.e. 1020 K, the shape of stressstrain curve indicated that the restoration process was controlled by intensive dynamic recovery and dynamic recrystallization. The strain value at the ow stress peak decreased with increasing deformation temperature. A steady state ow stress range, following the ow stress maximum, was observed during hot

deformation at 1173 K. Dynamically recrystallized structures of single phase alloy were observed for samples deformed at the temperature of 1123 K and above (Fig. 5(b)). Dynamic recrystallization (DRX) was not detected in the presence of any type of the precipitates below the solvus temperature.. The ow softening during hot compression tests as well as the results of hardness measurements carried out at room temperature, can be related to the specic volume fraction of the material transformed by discontinuous precipitation process (Fig. 6). For comparison purposes, the results for over aged material (OA) are also displayed in the gure. Dynamic precipitation in

A.A. Hameda, L. Blaz / Materials Science and Engineering A254 (1998) 8389

87

Fig. 4. TEM microstructure of ST Cu3.45 wt.% Ti alloy deformed at the temperature of 973 K and strain rate of 1.4 10 4 s 1, mt = 0.11. Lamellae of i particles and their local spheroidization are shown.

solution treated material deformed at low temperature resulted in an increase of maximum ow stress value and was found to be much more effective hardening process than the strain hardening for OA material. It

Fig. 6. The effect of deformation temperature on the maximum ow stress value for ST Cu 3.45 wt.% Ti alloy deformed at strain rate of 1.4 10 4 s 1. Room temperature hardness of hot deformed ST samples and the material volume fraction (X) transformed by discontinuous precipitation (cellular growth) are indicated in the gure. The temperature range for aligned needle-like particle growth (NP) and dynamic recrystallization development (DRX) are marked in the gure. Data for OA material and the initial hardness of solution treated and undeformed material are also displayed for comparison proposes.

Fig. 5. Optical microstructure of ST Cu3.45 wt.% Ti alloy deformed at: (a) 1023 K and strain rate of 1.4 10 4 s 1, mt = 0.38: A, aligned discontinuous precipitates; C, cellular growth of plate-like particles; B, spheroidal particles within the matrix (continuous precipitation); (b)1123 K and strain rate of 1.410 4 s 1, mt = 0.39 (dynamically recrystallized structure).

could also be concluded from Fig. 6 that the ow stress decrease during hot compression tests at the temperature of 873 K and above was, in general, dependent on discontinuous (cellular) growth of i particles which and could be considered as the result of the increase of the soft structural component volume. However, in the light of previous results, it should be stressed that the interaction of shear banding and dynamic coarsening of

88

A.A. Hameda, L. Blaz / Materials Science and Engineering A254 (1998) 8389

Ti-rich particles within shear bands is the most efcient structural processes giving rise to the pronounced load decrease (ow softening) at intermediate deformation temperatures. The Arrhenius plot of ow stress versus reciprocal temperature of deformation was used to estimate the activation energy for a hot deformation process according to the formula: m; = A| n exp(Q/RT) where: A, n are empirical constants; R is the gas constant; | is the maximum ow stress value. The Q-value of 389 kJ mol 1 (105 kcal mol 1) was calculated for experimental data of the single phase material (above Tsolvus). The slope of ln(|) versus T 1 line and calculated Q-value do not differ signicantly while passing thorough the solvus temperature within the deformation temperature range 9731173 K (Fig. 7). The balance of dynamic recovery and dynamic recrystallization within the deformation temperature range of : 9 100 K below and above Tsolvus have practically no effect on the Q-value. This may suggest that the Qvalue depends upon DR processes rather than DRX. The data for OA material were also displayed for comparison purposes [11]. In practice, the maximum ow stress value for both ST and OA samples at deformation temperatures above and very close to the Tsolvus, was not dependent on the samples heat treatment before the hot compression test. At low deformation temperatures, a nonlinear relation of ln(| n) versus T 1 has resulted from the temperature dependent

strain/precipitation interaction and the change of deformation mechanisms into intensely non-homogeneous processes.

4. Conclusions Hot deformation of solution treated Cu3.45 wt.% Ti alloy in the temperature range 673773 K was found to result in intensive strain/precipitation hardening of the material. Continuous precipitation of very ne i% particles at 773 K was accompanied by a nonuniform deformation and intergranular cavity growth which resulted in sample fracture at larger strains. The most effective interaction of non-homogeneous deformation and discontinuous precipitation of i-particles was observed in the deformation temperature range 873973 K. The ow stress maximum was followed by a ow softening range which resulted from non-homogeneous deformation of the material. Dynamic coarsening of precipitates and intense DR of the h-matrix within shear bands intensied the localized plastic ow and was found to be responsible for the marked ow stress decrease at larger strains. Due to intensive dynamic recovery within the localized ow areas, sample fracture was effectively retarded. Cellular growth of i particles in the vicinity of grain boundaries was also considered to produce a soft structural component which retards the intergranular fracture of the material. Discontinuous precipitation was practically complete during the temperature stabilization and early stages of deformation at 1073 K. In consequence, the interaction of phase transformation and non-uniform deformation was limited, and the bending of elongated particles as well as their fracture and spheroidization within the heavy deformed matrix were responsible for a slight decrease in ow stress at larger strains. Non-homogeneous deformation was not observed for samples deformed at 1023 K and above. During hot deformation of the CuTi alloy at 1073 and 1123 K, i.e. above the solvus temperature, dynamic recrystallization of the single phase material was observed. Arrhenius analysis in the vicinity of the solvus temperature (range : 9 100 K) suggested that the occurrence of dynamic recrystallization did not distinctly change the activation energy of the hot deformation process. A tentative conclusion is that dynamic recovery becomes the dominant structural process controlling the ow stress value during hot deformation within this temperature range.

Fig. 7. Arrhenius plot of maximum ow stress value versus reciprocal temperature of deformation for ST Cu3.45 wt.% Ti alloy. The Q-value used to t the experimental points for ST samples above the solvus temperature is marked in the gure. Data for OA material are included for comparison.

Acknowledgements The authors are grateful to Professor Bevis Hutchinson for his help and proving the text and to Professor

A.A. Hameda, L. Blaz / Materials Science and Engineering A254 (1998) 8389

89

Andrzej Korbel for his valuable advice and helpful discussion. Special thanks go to Marian Orkisz for his assistance. One of the authors gratefully acknowledges the nancial support of Al-Fateh University at TripoliLibya. The work was supported by the State Committee for Science Research, Grant No: PB501/T08/97/12. The Foundation for Polish Science is acknowledged for funding the electron microscope.

References
[1] J.A. Cornie, A. Datta, W.A. Soffa, Metall. Trans. 4 (1973) 727.

[2] H.T. Michel, I.B. Cadoff, E. Levine, Metall. Trans. 3 (1972) 667. [3] D.E. Laughlin, J.W. Cahn, Acta Metall. 23 (1975) 329. [4] A.W. Thomson, J.C. Williams, Metall. Trans. A 15A (1984) 931. [5] J. Dutkiewicz, Metall. Trans. A 8A (1977) 751. [6] L. Blaz, T. Sakai, J.J. Jonas, Mater. Sci. 17 (1983) 609. [7] A. Korbel, W. Bochniak, L. Blaz, J.D. Embury, Metall. Sci. 18 (1984) 216. [8] S. Wierzbinski, A. Korbel, J.J. Jonas, Mater. Sci. Technol. 8 (1992) 153. [9] T. Sakai, J.J. Jonas, Acta Metall. 32 (1984) 189. [10] L. Blaz, E. Evangelista, M. Niewczas, Metall. Mater. Trans. 25A (1994) 257. [11] A.A. Hameda, PhD. Thesis, University of Mining and Metallurgy, Cracow, Poland.

Das könnte Ihnen auch gefallen