Download as pdf or txt
Download as pdf or txt
You are on page 1of 28

Accepted Manuscript

Analysis of microstructural effects on mechanical properties of copper alloys

Mitsuhiro Okayasu, Professor, Takuya Muranaga, Graduate Student, Ayana Endo,


Undergraduate Student

PII: S2468-2179(16)30191-5
DOI: 10.1016/j.jsamd.2016.12.003
Reference: JSAMD 74

To appear in: Journal of Science: Advanced Materials and Devices

Received Date: 16 November 2016


Revised Date: 17 December 2016
Accepted Date: 22 December 2016

Please cite this article as: M. Okayasu, T. Muranaga, A. Endo, Analysis of microstructural effects on
mechanical properties of copper alloys, Journal of Science: Advanced Materials and Devices (2017),
doi: 10.1016/j.jsamd.2016.12.003.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

1. Mitsuhiro Okayasu
Professor, Graduate School of Natural Science and Technology,
Okayama University
3-1-1 Tsushimanaka, Kita-ku, Okayama, 700-8530, Japan
+81-86-251-8025

PT
2. Takuya Muranaga

RI
Graduate Student, Graduate School of Natural Science and Technology,
Okayama University
3-1-1 Tsushimanaka, Kita-ku, Okayama, 700-8530, Japan

SC
3. Ayana Endo

U
Undergraduate Student, Graduate School of Science and Technology, Ehime University
3 Bunkyo-cho, Matsuyama, Ehime, 790-8577, Japan
AN
M
D
TE
C EP
AC
1
ACCEPTED MANUSCRIPT
Analysis of microstructural effects on mechanical properties of copper alloys

Abstract
With the aim of obtaining copper alloys with favorable mechanical properties (high strength

and high ductility) for various engineering applications, the microstructural characteristics of

PT
two conventional copper alloys—an aluminum bronze (AlBC; Cu–Al9.3–Fe3.8–Ni2–Mn0.8) and

a brass (HB: Cu–Al4–Zn25–Fe3–Mn3.8)—and a recently developed aluminum bronze (CADZ:

RI
Cu–Al10.5–Fe3.1–Ni3.5–Mn1.1–Sn3.7), were controlled by subjecting the alloys to two different

processes (rolling and casting) under various conditions. For the rolling process, the rolling

SC
rate and temperature were varied, whereas for the casting process, the solidification rate was

U
varied. Microstructural characteristics, as examined by electron backscatter diffraction
AN
analysis, were found to differ among the alloys. Complicated microstructures formed in

CADZ led to high hardness and high tensile strength (σUTS), but low ductility (εf). For CADZ,
M

casting at a high solidification rate allowed an increase in ductility to be obtained as a result


D

of fine-grained structure and low internal stress. In contrast, high ductility (with a fracture
TE

strain of more than 30%) was found for both cast AlBC and cast BC; moreover, both of these

alloys possessed high tensile strength when produced by warm rolling at 473 K. For CADZ,
EP

on the other hand, no clear effect of rolling on tensile strength could be found, owing to the

many microcracks caused by its brittleness. The results of this study indicate that copper
C

alloys with excellent mechanical properties can be produced. This is especially the case for
AC

the conventional alloys, with a high tensile strength σUTS = 900 MPa and a high fracture

strain εf = 10% being obtained for warm-rolled brass.

Keywords: bronze; brass; strength; ductility; microstructural characteristics


2
ACCEPTED MANUSCRIPT

1. Introduction

Copper alloys, including bronzes, are currently employed in a wide range of engineering

applications because of their high ductility, high corrosion resistance, non-magnetic properties,

excellent machinability, and high hardness [1]. Copper is used for electric wiring and in heat

PT
exchangers, pumps, tubing, and several other products, while aluminum bronze and high-

strength brass are found in marine applications, for example in propellers and propeller shafts

RI
[2]. Furthermore, shiny brass is widely employed for coins and for musical instruments.

SC
However, in spite of their excellent material characteristics, there is still scope for technical

improvements to increase the strength and ductility of these alloys. To achieve improvements

U
in mechanical strength, several copper alloys with high dislocation density and fine
AN
microstructure, containing solid solutions, have been proposed. The mechanical strength of

ultrafine-grained or nanocrystalline Cu–Al alloys, prepared by equal-channel angular pressing


M

(ECAP), has been investigated, and the strength and uniform elongation of these alloys have
D

been simultaneously improved by lowering the stacking fault energy [3]. The hardness of
TE

even nanocrystalline copper with grain size as small as 10 nm still follows the Hall–Petch

relation [4].
EP

A variety of methods have been used to make high-strength copper alloys. Maki et al. [5]

attempted to create a higher-strength Cu–Mg alloy through a solid-solution hardening effect,


C

in which supersaturation with Mg increases the strength compared with that of a


AC

representative solid-solution Cu–Sn alloy [5]. A high tensile strength of 600 MPa was

reported by Sarma et al. [6], who produced a Cu–Al alloy with ultrafine-grained

microstructure and very fine annealing twins by cryorolling and annealing at 523 K for

15 min. The higher strength of this Cu–Al alloy was interpreted in terms of the enhanced

solid-solution strengthening effect of Al, which is about 1.7 times higher than the
3
ACCEPTED MANUSCRIPT

corresponding effect in Cu–Zn alloys [6]. In recent years, Cu–Zn30–Al0.8 alloys exhibiting

nanostructure have been fabricated by cryomilling of brass powders and subsequent spark

plasma sintering [7]. Such alloys have a high compressive yield strength of 950 MPa, which

is much higher than the values of 200–400 MPa found in commercially available alloys. This

PT
increase in mechanical strength has been attributed to precipitation hardening and grain

boundary strengthening [7]. The effect of grain size on yield stress was examined in

RI
polycrystalline copper and Cu–Al alloys at 77 and 293 K, and the yield stress was found to

SC
satisfy the Hall–Petch relation in both materials [8]. The influence of hydrogen on the

mechanical properties of aluminum bronze was investigated, and it was found that neither

U
tensile nor fatigue properties were affected [9]. After low-temperature thermal treatment,
AN
strained Cu–Al alloys exhibited high mechanical strength, which is caused by increases both

in the degree of order and in the electron-to-atom (e/a) ratio [10]. The effects of
M

microstructural characteristics on the mechanical strength of Cu–Ni26–Zn17 alloy were


D

investigated, and it was found that solid-solution strengthening of the alloy was affected by
TE

the interaction of Ni and Zn atoms with screw dislocations and by the effective interaction

caused by the modulus mismatch [11]. In order to understand the material properties of
EP

copper alloys, it is important to investigate their microstructural characteristics, including

texture. The textures of copper alloys after rolling and recrystallization were analyzed by
C

electron backscatter diffraction analysis (EBSD) [12]. The evaluation of grain boundaries in
AC

copper bicrystals during one-pass ECAP was systematically investigated by several methods,

including EBSD [13].

The above literature survey shows that there are various approaches that can be adopted

to improve the mechanical properties of copper alloys, including grain refinement, solid

solutions, and high dislocation density. In many practical applications, it is desirable to reduce
4
ACCEPTED MANUSCRIPT

the weight of components and structures made from such alloys by enhancing their

mechanical properties. Thus, in the present work, an attempt is made to create copper alloys

with favorable tensile properties (high strength and ductility) via microstructural modification

using forging and casting processes under various conditions. To analyze the mechanical

PT
strength and ductility of these alloys, their microstructural characteristics are investigated by

EBSD.

RI
SC
2. Experimental procedures

2.1. Sample preparations

U
Two commercial copper alloys, namely, an aluminum bronze (AlBC: Cu–Al9.3–Fe3.8–
AN
Ni2–Mn0.8) and a brass (HB: Cu–Al4–Zn25–Fe3–Mn3.8), were studied, as well as a newly

developed aluminum bronze (CADZ: Cu–Al10.5–Fe3.1–Ni3.2–Mn1.1–Sn3.7). It should be


M

pointed out that CADZ was developed on the basis of a Cu–Al10.5 alloy in Dozen-Kogyo Co.
D

Ltd. The material characteristics of CADZ were originally developed by described in detail
TE

elsewhere [14].

The test samples of the alloys were produced by casting and forging (rolling). In the
EP

casting process, two different cooling rates, and thus solidification speeds, were adopted. At

the low cooling rate (slow cooling, SC: 20 K/s), the melts were solidified slowly in a furnace.
C

In this case, the solidification process was carried out under an argon gas atmosphere to
AC

prevent oxidation. At the high cooling rate (rapid cooling, RC: 150 K/s), the melts were

solidified rapidly in a copper mold. The solidification speeds for both the rapid and slow

cooling processes were measured directly using a thermocouple. In the rolling process, the

alloys were forged at different deformation rates, using a 10-ton twin-rolling machine

(Yoshida Kinen Co., Ltd.) with high-strength rollers made of hot rolled steel (SKD11:
5
ACCEPTED MANUSCRIPT

150 mm diameter × 200 mm). Samples of thickness 10 mm were forged under severe

deformation at different temperatures: 293 K (cold rolling, CF), 493 K (warm rolling, WF),

and 1073 K (hot rolling, HF).

PT
2.2. Material properties

Tensile tests were conducted at room temperature using a hydraulic servo-controlled

RI
testing machine with 50 kN capacity. Rectangular dumbbell-shaped specimens were

SC
employed with dimensions 3 mm × 20 mm × 2 mm. The loading speed was set at 1 mm/min

until final failure. The tensile properties (ultimate tensile strength σUTS and fracture strain εf)

U
were evaluated via tensile stress versus tensile strain curves, which were monitored by a data
AN
acquisition system in conjunction with a computer through a standard load cell and strain

gauge. Hardness measurements were made using a micro-Vickers tester at 2.94 N for 15 s. In
M

this test, a diamond indenter was loaded manually at about 0.3 N/s to the sample surface,
D

which had been polished to a mirror finish.


TE

The microstructural and lattice characteristics of the alloys were investigated by EBSD

using a field emission scanning electron microscope (SEM; JEOL JSM-7000F), with an
EP

acceleration voltage of 15 kV, a beam current of 5 nA, and a step size of 20 µm. The samples

were sectioned to less than 10 mm thick, and the sample faces for the observation were
C

polished to a mirror finish in a vibropolisher, using colloidal silica for no longer than 2 h.
AC

3. Results and discussion

3.1. Microstructural characteristics

Fig. 1 shows the maximum possible rolling rates for the alloys. With the cold-rolling

process, the maximum rolling rate of CADZ is about 9%, which is about 50% and 33% lower
6
ACCEPTED MANUSCRIPT

than those for HB and AlBC, respectively. With the warm-rolling process, the rolling rate is

still as low as 12% for CADZ, although severe deformation of more than 75% is obtained for

HB and AlBC after warm rolling. With the hot-rolling process, a high rolling rate of more

than 75% is obtained for the three alloys.

PT
Fig. 2 shows optical micrographs of the three alloys, made using both the casting and the

rolling processes. Essentially, the three alloys consist of matrix and eutectic structures. The

RI
main eutectic phases of the CADZ sample are found to be (Fe, Ni)3Cu, Cu–Ni–Sn and Cu–Al

SC
[14], as indicated by the arrows in cast CADZ. The AlBC sample is essentially formed from

eutectic Fe-, Cu–Al–Ni-, and Cu–Al-based phases, while for the HB sample, eutectic Fe-, Cu–

U
Zn–Al-, and Cu–Zn-based phases are observed. The grain size clearly varies for all the cast
AN
alloys, where the higher the cooling rate, the smaller the grain size. For the rolled samples, no

clear changes in grain size can be detected, especially for CADZ. This could be due to the low
M

rolling rates for CF- and WF-CADZ. On the other hand, grain growth (or recrystallization)
D

occurs for HF-CADZ and for HF-HB. For CF- and WF-AlBC, slightly strained
TE

microstructural formations can be seen. To understand these microstructural characteristics in

detail, an EBSD analysis was carried out.


EP

Fig. 3 displays the inverse pole figure (IPF) and misorientation (MO) angle maps of the

cast alloys, obtained by EBSD. As can be seen, complicated microstructures with high MO
C

angles are formed almost throughout both the cast and rolled CADZ samples. In contrast, high
AC

MO angles are found mainly in the eutectic phases of AlBC, whereas high MO angles are

widely distributed in RC- and CF-AlBC. Similar trends are observed in the corresponding HB

samples. The MO angles for the CADZ samples are overall higher than those for AlBC and

HB. The higher MO angles for the CF samples are considered to be due to increased

dislocation density, while the low MO angles for the HF samples result from a reduction in
7
ACCEPTED MANUSCRIPT

internal stress due to the high-temperature processing. In addition, it is notable that

deformation twins can be clearly detected in the rolled AIBC samples, but not in the others.

This can be attributed to the different extents of stacking fault energy (SFE): the lower the

SFE, the weaker the deformation twins. In previous work, it has been reported that the SFE

PT
decreases with an increasing proportion of Al in the alloy composition: for example, the SFE

of Cu–Al2.3 alloy is about 6 times higher than that of Cu–Al11.6 alloy [15]. Since it has been

RI
reported that the SFE of Cu–Zn24 alloy is about 4 times higher than that of Cu–Al8 alloy [16],

SC
our AIBC should have a much lower SFE compared with HB, leading to deformation

twinning in the former but not the latter.

U
The grain size of the cast alloys was measured directly, and the results are summarized in
AN
Table 1. It should be pointed out first that for the Cu-based phases, measurements were made

of straight diagonal lines on each grain, and the grain size was determined as the mean value
M

of more than 50 measurement data. Since grain formation is not clearly seen for CADZ,
D

image analysis was conducted on the cast CADZ, with the Fe element being removed from
TE

the IPF maps; see Fig. 3(d). From Table 1, it can be seen that the grain size varies, depending

on the sample and the casting speed. The differences in microstructural characteristics lead to
EP

differences in mechanical properties. The average grain size of the alloys made by rapid

cooling is less than 38 µm. The grain size increases with decreasing cooling rate: for example,
C

for HB, a large grain size of 680 µm is obtained, which is more than 10 times greater than that
AC

for AIBC.

3.2. Mechanical properties

Figs. 4 and 5 show Vickers hardness data for the three alloys made by rolling and casting

processes under different conditions. For the cast samples shown in Fig. 4(a), a high hardness
8
ACCEPTED MANUSCRIPT

is obtained overall for CADZ: for example, the value of about 2.5 GPa for AS-CADZ is about

15% and 70% higher than those for AS-HB and AS-AlBC, respectively. An improvement in

hardness is obvious for all the alloys with a higher solidification rate (the RC samples): for

example, for RC-CADZ, the hardness is as high as 3.3 GPa, which is more than 1.4 times that

PT
for SC-CADZ. On the other hand, the lowest hardness of about 1.4 GPa is obtained for the

SC-AlBC samples. These differences in hardness are due to a number of reasons, including

RI
the fact that different grain sizes lead to different grain boundary strengths. Fig. 4(b) shows

SC
the relationship between grain size and hardness for the alloys. Although there are only a few

data points, clear correlations can be seen, and the Hall–Petch relation appears to be satisfied.

U
A similar Hall–Petch relation is also obtained for nanocrystalline copper (10 nm) [4].
AN
For the rolled samples shown in Fig. 5, the hardness value increases with increasing

rolling rate and decreasing rolling temperature. These trends are presumably due to the
M

differences in dislocation density, deformation twinning, and internal stress arising during the
D

rolling process, as indicated by the distributions of MO angles seen in Fig. 3. In particular, a


TE

high hardness is obtained for the cold-rolling process, owing to dislocation tangling, despite

the low rolling rate. On the other hand, the low hardness of the samples made by hot rolling is
EP

a consequence of their recrystallization and grain growth, as described previously. It should

also be pointed out that the deformation characteristics of AIBC and HB can vary depending
C

on the SFE, as mentioned above. In general, it appears that deformation twining occurs for the
AC

alloys with lower SFE, namely, the AIBC samples. This deformation occurs when dislocation

is dominated by the rolling process, i.e., work hardening occurs [17].

Fig. 6 shows representative tensile stress versus tensile strain curves for the three alloys

made by rolling and by casting, while Fig. 7 summarizes their tensile properties in terms of

ultimate tensile strength versus fracture strain. It should be pointed out that more than three
9
ACCEPTED MANUSCRIPT

specimens were employed here to obtain the tensile properties. From the stress–strain curves,

it can be seen that high ductility is obtained for the cast samples, with the fracture strain for

AIBC being higher than that for HB and CADZ. The reason for this is the presence of

deformation twinning in AIBC, as mentioned above. Huang et al. [18] reported that the

PT
deformation twins in coarse-grained Cu occurred mainly in shear bands and at their

intersections, as a result of the very high local stress caused by severe plastic deformation. On

RI
the other hand, a high tensile strength is obtained overall for the rolled samples compared

with the cast ones. In particular, higher tensile strengths σUTS are obtained overall for AlBC

SC
and HB made at a high rolling rate and a rapid cooling rate. The highest σUTS values

U
(>900 MPa) are obtained for WF-AlBC, WF-HB, and RC-CADZ. On the other hand, low
AN
σUTS values are found for HF-HB and HF-AlBC, even when high rolling rates were applied.

The data plots of tensile properties are relatively scattered for CADZ, which may be due to
M

the low sample quality. Fig. 8 shows an SEM image of the HF-CADZ sample after rolling but
D

before the tensile test. As can be seen, several microcracks have been generated along the
TE

grain boundaries, as indicated by the dashed lines. Such microcracks could lead to a

deterioration in mechanical properties. It should be noted that no clear microcracks were


EP

detected in the other rolled alloys, because of their high ductility.

For the cast samples in Fig. 7, higher tensile strengths are obtained for the alloys made at
C

a high solidification rate (the RC alloys). For cast CADZ, the highest σUTS value is obtained
AC

for RC-CADZ, and is higher than the value for the corresponding rolled alloy. This may be a

consequence of the fine-grained structure as well as the high sample quality (with no

microcracks). The tensile strength of the cast samples decreases with decreasing solidification

rate. Unlike the tensile strength of CADZ, high tensile strengths of both HB and AlBC result

from cold and warm rolling at a high rolling rate. In addition, the rolled AlBC and HB alloys
10
ACCEPTED MANUSCRIPT

(e.g., the CF and WF samples) show a raised ductility εf of more than 15%, although this

strain value is lower than those for the RC-AlBC and RC-HB samples. From this result, it can

be considered that the cast and rolled samples are overall located on the right- and left-hand

sides, respectively. On the other hand, no clear trend in tensile properties is seen for CADZ.

PT
This may be due to its low deformability and the microcracks generated by the rolling process,

as mentioned above.

RI
SC
4. Conclusions

The mechanical properties of copper alloys made by different processes have been

U
investigated. The results can be summarized as follows:
AN
1) The mechanical properties of the alloys depend on the production process: rolling or

casting. For the CADZ alloy, high mechanical strength was obtained for the rapidly cooled
M

cast sample, although low ductility was found. High ductility (>30% in some cases) was
D

obtained for cast AlBC and HB alloys. High tensile strength with high ductility was
TE

obtained by warm rolling at a high rolling rate, especially for HB and AlBC.

2) The high hardness of the CADZ alloy was attributed to severe lattice strains almost
EP

throughout the material. Vickers hardness was clearly related to grain size for all three

alloys, with larger grains leading to lower hardness, i.e., the Hall–Petch relationship.
C

3) The CADZ alloy could not be subjected to intense rolling owing to its brittleness, arising
AC

from its complicated microstructure. A large number of microcracks were created in rolled

CADZ, resulting in reduced tensile strength. On the other hand, intense rolling was possible

for the HB and AlBC alloys, allowing samples to be produced with high strength and high

ductility.
11
ACCEPTED MANUSCRIPT

Acknowledgements

The authors appreciate financial support from the Japan Copper and Brass Association,

and the Cu alloys used in the present work were provided by Dozen-Kogyo Co. Ltd.

PT
References
1 H. Imai, Y. Kosaka, A. Kojima, S. Li, K. Kondoh, J. Umeda, H. Atsumi, Characteristics
and machinability of lead-free P/M Cu60-Zn40 brass alloys dispersed with graphite,

RI
Powder Tech. 198(2010)417-421.

2 A.H. Tuthill, Guidelines for the use of copper alloys in seawater, Mater. Perform.

SC
Sept.(1987)12-22.

3 S. Qu, X.H. An, H.J. Yang, C.X. Huang, G. Yang, Q.S. Zang, Z.G. Wang, S.D. Wu, Z.F.
Zhang, Microstructural evolution and mechanical properties of Cu-Al alloys subjected to

U
equal channel angular pressing, Acta Mater. 27(2009)1586-1601.
AN
4 J. Chen, L. Lu, K. Lu, Hardness and stain rate sensitivity of nanocrystalline Cu, Scripta
Mater. 54(2006)1913-1918.
M

5 K. Maki, Y. Ito, H. Matsunaga, H. Mori, Solid-solution copper alloys with high strength
and high electrical conductivity, Scripta Mater. 68(2013)777-780.
D

6 V.S. Sarma, K. Sivaprasad, D. Sturm, M. Heilmaier, Microstructure and mechanical


properties of ultrafine grained Cu-Zn and Cu-Al alloys produced by cryorolling and
TE

annealing, Mater. Sci. Eng. A 489(2008)253-258.

7 H. Wen, T.D. Topping, D. Isheim, D.N. Seidman, E.J. Lavernia, Strengthening mechanisms
in a high-strength bulk nanostructured Cu-Zn-Al alloy processed via cryomilling and spark
EP

plasma sintering, Acta Mater. 61(2013)2769-2782.

8 T. Tabata, K. Takagi, H. Fujita, The effect of grain size and deformation sub-structure on
C

mechanical properties of polycrystalline copper and Cu-Al alloys, Trans. JIM


16(1975)569-579.
AC

9 T. Michler, J. Naumann, Influence of high pressure hydrogen on the tensile and fatigue
properties of a high strength Cu-Al-Ni-Fe alloy, Int. J. Hydrogen Energy, 35(2010)11373-
11377.

10 J.M. Popplewell, J. Crane, Order-strengthening in Cu-Al alloys, Metal. Trans.


2(1971)3411-3420.

11 S. Nagarjuna, M. Srinivas, K.K. Sharma, The grain size dependence of flow stress in a Cu-
26Ni-17Zn alloy, Acta Mater. 48(2000)1807-1813.

12 S.S. Vadlamani, J. Eickemeyer, L. Schultz, B. Holzapfel, Rolling and recrystallization


12
ACCEPTED MANUSCRIPT
textures in Cu-Al, Cu-Mn and Cu-Ni alloys, J. Mater. Sci. 42(2007)7586-7591.

13 W.Z. Han, H.J. Yang, X.H. An, R.Q. Yang, S.X. Li, S.D. Wu, Z.F. Zhang, Evolution of
initial grainboundaries and shear bands in Cu bicrystals during one-pass equal-channel
angular pressing, Acta Mater. 57(2009)1132-1146.

14 M. Okayasu, D. Izuka, Y. Ninomiya, Y. Manabe, T. Shiraishi, Mechanical and wear


properties of Cu-Al-Ni-Fe-Sn-based alloy, Adv. Mater. Res. 2(2013)221-235.

PT
15 C.X. Huang, W. Hu, G. Yang, ZF. Zhang, S.D. Wu, Q.Y. Wang, G. Gottstein, The effect of
stacking fault energy on equilibrium grain size and tensile properties of nanostructured
copper and copper-aluminum alloys produced by equal channel angular pressing, Mater.

RI
Sci. Eng. A 556(2012)638-647.

16 A. Rohatgi, K.S. Vecchio, G.T. Gray III, The influence fo stacking fault energy on the

SC
mechanical behavior of Cu and Cu-Al alloys: deformation twinning work hardening, and
dynamic recovery, Metall. Mater. Trans. A 32A(2001)135-145.

17 X.X. Wu, X.Y. San, X.G. Liang, Y.L. Gong, X.K. Zhu, Effect of stacking fault energy on

U
mechanical behavior of cold-forging Cu and Cu alloys, Mater. Des. 47(2013)372-376.
AN
18 C.X. Huang, K. Wang, S.D. Wu, Z.F. Zhang, G.Y. Li, S.X. Li, Deforamtion twinning in
polycrystalline copper at room temperature and low strain rate, Acta Mater. 54(2006)655-
665.
M

Captions
D

Table 1 Grain sizes of the copper alloys CADZ, AlBC, and HB (SD: standard deviation)
TE

Fig. 1 Maximum rolling rates for copper alloys at different temperatures.

Fig. 2 Optical micrographs of copper alloys made by casting and by rolling: (a) CADZ; (b)
EP

AlBC; (c) HB.

Fig. 3 (a–c) Inverse pole figure (IPF) and misorientation angle maps of (a) CADZ, (b) AlBC,
and (c) HB. (d) IPF map for CADZ with and without Fe element.
C

Fig. 4 (a) Vickers hardness of copper alloys made by casting. (b) Relationship between
AC

Vickers hardness and grain size.

Fig. 5 Vickers hardness of copper alloys made by rolling: (a) CADZ; (b) AlBC; (c) HB.

Fig. 6 Stress–strain curves for copper alloys made by rolling and by casting: (a) CADZ; (b)
AlBC; (c) HB.

Fig. 7 Relationship between ultimate tensile strength and tensile strain for copper alloys made
by rolling and by casting: (a) CADZ; (b) AlBC; (c) HB.

Fig. 8 SEM image of the hot-rolled CADZ sample, showing some microcracks.
ACCEPTED MANUSCRIPT

PT
RI
SC
Table 1 Grain sizes of the copper alloys CADZ, AlBC, and HB (SD: standard deviation)
Cu alloys
U CADZ AlBC HB
AN
Rapid cooling (µm) 38(SD:8.6µm) 15(SD:10.9µm) 38(SD:4.6µm)
Casting
As-cast (µm) 132(SD:16.3µm) 45(SD:14.0µm) 81(SD:19.7µm)
process
Slow cooling (µm) 220(SD:34.6µm) 60(SD:52.1µm) 680(SD:176.1µm)
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP

Fig. 1 Maximum rolling rates for copper alloys at different temperatures.


C
AC
ACCEPTED MANUSCRIPT

(a)CADZ
Cast sample

Rapid cooling (RC) As-cast (AS) Slow cooling (SC)

PT
Cu-Ni-Sn

RI
Cu-Al

SC
Fe-Ni-Cu

Rolling sample

Cold rolling (CF)


U
Warm rolling (WF) Hot rolling (HF)
AN
9% 12% 25%
M
D
TE

50%
EP

100µm
C
AC
ACCEPTED MANUSCRIPT

(b)AlBC
Cast sample

Rapid cooling (RC) As-cast (AS) Slow cooling (SC)

PT
Cu-A

RI
Cu-Al-N

SC
Fe

Rolling sample

Cold rolling (CF)


U
Warm rolling (WF) Hot rolling (HF)
AN
25% 25% 25%
M

Strained grain
D
TE

50% 50%
EP

100µm
C
AC
ACCEPTED MANUSCRIPT

(c)HB
Cast sample

Rapid cooling (RC) As-cast (AS) Slow cooling (SC)

PT
RI
Fe

Cu-Zn

U SC
Rolling sample
AN
Cold rolling (CF) Warm rolling (WF) Hot rolling (HF)

19% 25% 25%


M

Grain growth
D
TE

50% 50%
EP

100µm
C
AC

Fig. 2 Optical micrographs of copper alloys made by casting and by rolling:


(a) CADZ; (b) AlBC; (c) HB.
ACCEPTED MANUSCRIPT

(a) CADZ
Cast sample
Rapid cooling (RC) As-cast (AS) Slow cooling (SC)
IPF MO IPF MO IPF MO

PT
RI
U SC
AN
Rolling sample
M

Cold rolling (CF) Warm rolling (WF) Hot rolling (HF)


D

IPF MO IPF MO IPF MO


TE

9% 12% 50%
C EP
AC

70µm

111
2°~ 5°
5°~ 15°
15°~180°
001 101
ACCEPTED MANUSCRIPT

(b) AlBC
Cast sample
Rapid cooling (RC) As-cast (AS) Slow cooling (SC)
IPF MO IPF MO IPF MO

PT
RI
U SC
AN
Rolling sample
M

Cold rolling (CF) Warm rolling (WF) Hot rolling (HF)


D

IPF MO IPF MO IPF MO


TE

25% 50% 50%


C EP
AC

30µm
ACCEPTED MANUSCRIPT

(c) HB
Cast sample
Rapid cooling (RC) As-cast (AS) Slow cooling (SC)
IPF MO IPF MO IPF MO

PT
RI
U SC
AN
M

Rolling sample
Cold rolling (CF) Warm rolling (WF) Hot rolling (HF)
D

IPF MO IPF MO IPF MO


TE

19% 50% 50%


C EP
AC

100µm
ACCEPTED MANUSCRIPT

(d) IPF map for CADZ with and without Fe element

PT
Cast sample
Rapid cooling (RC) As-cast (AS) Slow cooling (SC)

RI
With Without With Without With Without

U SC
AN
M
D
TE

70µm
EP

Fig. 3 (a–c) Inverse pole figure (IPF) and misorientation angle maps of (a) CADZ, (b) AlBC,
and (c) HB. (d) IPF map for CADZ with and without Fe element.
C
AC
ACCEPTED MANUSCRIPT

PT
RI
(a) (b)

U SC
AN
M
D
TE

Fig. 4 (a) Vickers hardness of copper alloys made by casting. (b) Relationship
EP

between Vickers hardness and grain size.


C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
C EP
AC

Fig. 5 Vickers hardness of copper alloys made by rolling: (a) CADZ; (b) AlBC; (c) HB.
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
C EP
AC

Fig. 6 Stress–strain curves for copper alloys made by rolling and by casting: (a) CADZ; (b) AlBC; (c) HB.
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
C EP
AC

Fig. 7 Relationship between ultimate tensile strength and tensile strain for copper
alloys made by rolling and by casting: (a) CADZ; (b) AlBC; (c) HB.
ACCEPTED MANUSCRIPT

PT
RI
CADZ (HF sample)

U SC
AN
Micro-crack
s
M
D

1mm
TE

Fig. 8 SEM image of the hot-rolled CADZ sample, showing some microcracks.
C EP
AC

You might also like