You are on page 1of 49

Author's Accepted Manuscript

Effect of chemical composition and processing


variables on the hot flow behavior of leaded
brass alloys
A. Momeni, G.R. Ebrahimi, H.R. Faridi

www.elsevier.com/locate/msea

PII: S0921-5093(14)01503-2
DOI: http://dx.doi.org/10.1016/j.msea.2014.12.016
Reference: MSA31833

To appear in: Materials Science & Engineering A

Received date: 13 November 2014


Revised date: 29 November 2014
Accepted date: 4 December 2014

Cite this article as: A. Momeni, G.R. Ebrahimi, H.R. Faridi, Effect of chemical
composition and processing variables on the hot flow behavior of leaded brass
alloys, Materials Science & Engineering A, http://dx.doi.org/10.1016/j.
msea.2014.12.016

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 galley proof before it is published in its final citable 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.
Effect of chemical composition and processing
variables on the hot flow behavior of leaded brass
alloys

A. Momeni1*, G.R. Ebrahimi2, H.R. Faridi1

1- Department of Materials Science and Engineering, Hamedan University of Technology, Hamedan,

Iran.

2- Department of Materials and Polymer Engineering, Hakim Sabzevari University, Sabzevar, Iran.

Abstract

Hot compression tests were carried out on CuZn39Pb2 and CuZn39Pb3 leaded brasses
in temperature range of 600 °C-800 °C at strain rates of 0.001 s-1-1 s-1. The stress-
strain curves at low temperatures were characterized by a single peak. While at high
temperatures, the flow curves were characterized by a long plateau associated with
dynamic recovery. The phase diagram of Cu-Zn-Pb system drawn by the Thermocalc
software showed that increasing Pb could shift the phase border between α+β and β
towards the Zn corner, thereby increasing the volume fraction of α at the deformation
tempearutes. Hot deformation at low temperatures, e.g. 600 °C- 700 °C, changed the
strings of cornered α islands to more globularized descrete ones. However, the
conventional dynamic recrystallization could not be observed. At 800 °C, fine α
particles formed through dynamic recrystallization were coexistent with acicular ones
formed through quenching of β from high temperatures. It was found that at low
temperatures, e.g. below 700 °C- 750 °C, Pb could contribute to avoid flow
localization in both alloys. At 800 °C, more Pb could dissolve into β leading to more
tendency to flow locallization. The results showed that at low temperatures, i.e. below
700°C, both materials exhibited higher strength than that predicted by the law of
mixture. At high temperatures, particularly beyond 700°C, the predicted values of the
law of mixture laid below the experimental flow stresses. This was attributed to the
decrease in the volume fraction of α and more dissolution of Pb into β.

Key words: α-β brass; Hot compression; Mechanical properties; Dynamic recovery;
Dynamic recrystallization; Law of mixture

* Corresponding author, email: ammomeni@aut.ac.ir

1
1- Introduction

Cu-Zn brass alloys are widely used in different industries due to their corrosion

resistance and good mechanical properties. Other alloying elements may also be added to

impart special characteristics to brass. Lead, which is nearly insoluble in Cu alloys, is used to

improve machinability of leaded brass [1]. However, Pb, Bi or other elements that are used to

impart machinability often deteriorate low and high temperature ductility of brass [2, 3].

Depending upon the chemical composition of brass, a single-phase α (with fcc structure) or β

(with bcc structure) or a two-phase (duplex) α+β microstructure may form. The hot

deformation behavior of brasses with different microstructures has been the topic of some

investigations in the past decades [4-7]. During hot deformation, α , which is characterized

by low stacking fault energy (SFE), undergoes dynamic recrystallization (DRX) [7, 8];

whereas, dynamic recovery (DRV) is the restoration mechanism of β [8]. This imparts

superplastic behavior to the duplex brass alloys at high temperatures and low strain rate

deformations. The mechanisms of DRX and DRV are the same as well documented in the

literature on copper, steel and other alloys [9-13].

Despite few investigations conducted on the hot deformation behavior of duplex leaded

brass, the hot flow curves have been hardly interpreted. This is attributed to the different

restoration mechanisms of α and β. From similar investigations into duplex stainless steels

[14, 15] and Ti alloys [16, 17] it can be inferred that different deformation mechanisms of

constituents complicate tracing the microstructural evolution and modeling the flow curves

during hot working.

Different approaches, some based on the constitutive equations [5, 18] and some others

on the self-consistent models and the law of mixture rule [19-21], have been so far adopted to

model the flow curves of duplex alloys. Despite the fact that the law of mixture rule has not

been generalized to consider the change in the strength of the constituents with deformation

2
temperature, it has been used to predict the hot flow curves of some multiphase materials [14,

21- 23]. Before applying this method on duplex leaded brass, two challenges should be dealt

with: first, the influence of Pb particles on the hot flow behavior of brass and second, the

effect of chemical compositions of α and β at different deformation temperatures on their

strength level. Hence, the present work aimed at studying the hot deformation behavior of

duplex leaded brass with regard to the effect of Pb and calculating the flow curves using the

law of mixture rule.

2- Experimental procedures

Two leaded brass with different Pb contents and chemical compositions according to

Table 1 were used in this investigation. The starting materials were supplied with cold drawn

initial condition. The initial microstructures of the materials shown in Fig. 1 indicate that both

alloys are consisted of α and β. In the micrographs, the matrix phase is β and the lighter

islands that contains most of dark particles of Pb is α.

The cylindrical specimens with 10 mm in diameter and 15 mm in height were machined

from the supplied starting materials according to ASTM E209 standard. Hot compression

tests were carried out on the specimens at temperatures of 600 ºC, 650 ºC, 700 ºC, 750 ºC,

and 1073 K (800 ºC). The isothermal compression tests were applied at strain rates of 0.001 s-
1
, 0.01 s-1, 0.1 s-1, and 1 s-1 by Zwick-Roell testing machine.

The specimens were preheated at test temperatures for 10 minutes before hot

compression. Lubrication of the flat surfaces of the specimens was done by using graphite.

The specimens were compressed to 50% reduction in height (İ = 0.7) and the load–stroke

data were recorded. The load–stroke data were converted to true stress–true strain curves

using standard equations.

3
The specimens were immediately water-quenched after hot compression to preserve the

deformation microstructure. The deformed samples were sectioned parallel to the

compression axis and prepared by standard metallographic techniques according to ASTM

E3-11. After etching by a reagent composed of 100 ml H2O, 20 ml HCL and 5 gr iron (III)

chloride, hot compressed samples were examined by optical microscopy.

3- Results and Discussion

3-1- Flow curves and microstructures

The typical true stress-strain curves of the materials calculated from the load-stroke

data at various temperatures and two strain rates of 0.001 s-1 and 1 s-1 are shown in Fig. 2.

The results indicate that at low temperatures, i.e. 600 °C and 700 °C, and all strain rates

(0.001 s-1-1 s-1) flow stress of CuZn39Pb2 is higher than those of CuZn39Pb3. It is also

notable that the difference between flow curves decreases with increasing temperature so that

at 800 °C, the flow curves are nearly overlapped.

The observations should be addressed with regard to the differences in the chemical

compositions of the alloys. It is easy to discuss the results with respect to the different

amounts of Cu and Pb. The results can be justified if we consider that replacing Cu with a

lower melting point alloying element such as Pb (Tm = 315 °C) can weaken the material

especially over the studied temperature range. Although current results imply that Pb can

decrease the flow stress of such materials, it may also degrade hot or cold ductility [1] and

cannot be therefore used as a tool to decreases hot working loads.

As mentioned, ∆ı (the difference between the flow stresses of the materials) decreases

as temperature rises and nearly vanishes at 800 °C. This is an important finding that needs to

be elucidated with considering the phase equilibrium diagrams of the studied materials. The

equilibrium phase diagrams of Cu-Zn, Cu-Zn-2 wt. % Pb and Cu-Zn-3 wt. % Pb system has

4
been presented in Fig. 3. The phase diagrams of Fig. 3(b) and 3(c) confirm that Pb can push

the phase border between α+β and β towards the Zn corner [24]. This effect clearly changes

the volume fractions and chemical compositions of α and β that coexist at any deformation

temperature in the range of 873 K (600 ºC) to 1073 K (800 ºC). Previous investigations on

Cu-Zn-Pb system reported that the solubility of Pb in α is limited to 0.08 %; Whereas, Pb can

dissolve in β up to 0.35 % [24]. Therefore, the presence of Pb particles almost in α in Fig. 1

can be justified accordingly.

The weight percents of α calculated based on Fig. 3(b) are represented in Fig. 4. It is

seen that CuZn39Pb3 has greater amount of α than CuZn39Pb2 at any deformation

temperature. This notifies that Pb resides in α and stabilizes it versus β. At high temperatures,

i.e. beyond 700°C, both Figs. 3(b) and 4 indicate that the weight percent of α considerably

decreases. With decreasing α more Pb can dissolve into β and this in turn decreases the

difference between flow curves. It is worthy of attention that unlike the Pb particles in α, the

dissolved Pb atoms can participate in the strengthening of β and the similar flow curves of

materials at 800°C corroborates this idea.

The flow curves presented in Fig. 2 confirm that the balance of α and β at different

deformation temperatures controls the behavior of two sorts of samples. At low temperatures,

e.g. 600 ºC, where α mostly constitutes the microstructure of both materials, the flow curves

are characteristic of a material undergoing DRX alone. However, as the contribution of β to

the microstructure increases at high temperature, the materials exhibit less discernible peaks

and behave as though they were wholly β. Some typical micrographs of the materials at three

different temperatures of 600 ºC, 700 ºC and 800 ºC presented in Fig. 5 corroborate the

mechanical results. The change in the volume fraction of α and β with increasing temperature

is clearly observed for both alloys. It is manifesting that increasing temperature from 600 ºC

5
to 700 ºC leads to a remarkable decrease in the volume fraction of α (the light phase). This is

consistent with the results of Figs. 3 and 4 and the changes in the flow behavior of the

materials in Fig. 2. The micrographs of 600 ºC and 700 ºC (Fig. 5 (a)-(d)) also suggest a

change in the morphology of α with respect to initial morphology shown in Fig. 1. It is

thought that the initial string-like morphology of α with cornered faces turns into more

discrete globular α due to hot deformation at 600 ºC and 700 ºC. This observation suggests a

kind of rotation and globularization that may be partly responsible for the flow softening

during hot deformation. Similar reorientation in α was also reported by Roberts in some other

leaded brass alloys deformed at high temperatures [25]. The mechanism of rotation and

globularization may be an alternative to the conventional DRX that is apparently absent in the

α islands at 600 ºC and 700 ºC. In a similar research on duplex stainless steels, it has been

reported that conventional DRX in austenite which exists as the second-phase islands may be

remarkably postponed or inhibited due to the extensive DRV in the ferritic matrix [26]. At

800 ºC, according to Figs. 5 (e) and 5 (f), the morphology of α has changed to the acicular or

widmanstatten appearance. It seems that the acicular α grains have been formed through a

displacive transformation of quenched β. However, some fine-grained α patches which are

observed near or at the β grain boundaries suggest that DRX could occur in the α grains

remained at 800 ºC. According to the micrographs of Figs. 5 (e) and 5 (f) at 800 ºC the

volume fraction of β increases considerably and therefore the flow curves of both materials

behave as if they were wholly β.

6
3-2- Constitutive analysis

The constitutive description of flow stress provides a way to generalize the laboratory results

to industrial hot working operations. Here, it can also shed light on the influence of chemical

composition on the deformation resistance of the studied materials. The temperature and

strain rate dependence of flow stress is generally expressed in terms of an Arrhenius type

equation. The hyperbolic sine equation has been more widely used than other constitutive

relations to relate the hot flow stress of a material to deformation condition [27-29]. This

equation can be used at all ranges of low and high stresses, as follows [28]:

Q
) = A{sinh(Į ı)}
n
Z = İ exp( (1)
RT
where, Q is the apparent activation energy, n is the stress exponent, A and α are the material

constants and R is the universal gas constant.

The logarithmic form of Eq. (1) that is usually used to determine the materials constants is

written as follows:

Q
lnZ = ln İ + = ln A + n ln{sinh(Į ı)} (2)
RT
According to Eq. (2), the values of n and Q/R can be determined as ∂lnİ / ∂ln{sinh(Į ı)}and

∂ln{sinh(Į ı)}/∂ (1/T) from the plots of İ − ı and ı − (1/T) , respectively. α is basically an

adjustable constant and is determined in a way to bring the curves to the most linear and

parallel condition. The approach to determine α has been described elsewhere in detail [14].

The influence of strain rate on the hyperbolic sine function of flow stress at typical strain of

0.4 is shown in Fig. 6. The data points in Fig. 6 have been fitted in the straight lines from

which the average slopes (navg) determine the stress exponents (n) of the alloys. The results of

n summarized in Fig. 7 show that the value of n in both materials increases as temperature

rises. The sharp variation of n with deformation temperature beyond the range of 700-750°C

7
can be accounted for by the results of Figs. 4 and 5. As indicated, the flow behaviour also

changes when temperature passes the same range and β becomes the major microstructural

constituent. In general, for most industrial alloys, n tends to decrease with increasing

temperature which results in more resistance to flow localization [30].

1/n
According to Eq. (1), sinh(αı) is proportional to İ . In a material with low n (high 1/n),

when plastic deformation is localized, the local strain rate significantly increases. Then, the

flow stress of such region considerably increases due to the direct relationship between

1/n
sinh(αı) and İ , thereby leading to the transfer of deformation to other regions with lower

flow stress. As discussed, Fig. 6 indicates that both materials resist flow localization at lower

temperatures where possess lower n values. The Pb particles seem to be responsible for the

resistance to flow localization. As they are liquid at the deformation temperatures, they can

play the role of an internal lubricant and help α (the most constituent phase at low

temperatures) to reorient itself and accommodate more strain. By increasing temperature,

more Pb is dissolved into β and less contributes to postpone flow localization. The resistance

of two sorts of samples versus flow localization does not necessarily imply that Pb leads to

better hot ductility, because other mechanisms such as cavitation, wedge cracking and hot

shortness may still jeopardize hot workability.

The influence of temperature on flow stress is shown in the frame of the hyperbolic sine

constitutive equation in Fig. 8. According to Eq. (2), the slopes of curves equals to Q/nR,

from which the value of deformation resistance (Q) can be measured. The average of slopes

in Fig. 8 shows that the deformation resistance of alloys with 3 and 2 wt.% Pb can be

calculated as 267.4 and 234.1 kJ/mol, respectively. The values of Q confirm the previous

observation that Pb decreases the strength of leaded brass. However, according to Figs. 3 and

4, Pb also tends to increase the volume fraction of α that is the harder constituent.

8
3-3- Law of mixture analysis

As discussed before, the individual flow behaviors of α and β have been investigated

and modeled using experimental or phenomenological methods. However, in case of the

coexistence of these phases, in a duplex brass, the contribution of each phase should be

considered carefully. The contribution of α and β to the total strength of a duplex brass

depends on the chemical composition of each phase, which actually varies with deformation

temperature. The chemical compositions (Zn content) of α and β at various deformation

temperatures, calculated from the phase diagrams of Fig. 3(b), are shown in Fig. 9. In order to

model the flow curves of the studied materials the specific strength of each phase should be

determined as a function of temperature and strain rate. Then, the law of mixture role could

be applied to bring out the strength of the duplex alloys using the specific strengths and the

volume fractions of the constituents, as follows:

ı = f Į ı Į + fȕ ıȕ (3)

where, f stands for the volume fraction and ı is the strength of leaded brass at a given

deformation temperature and strain rate.

The flow stress of brass with various weight percents of Zn were obtained from the published

literature on the hot deformation of single-phase and duplex alloys [5, 7, 8, 30]. The results of

flow stress at a typical strain of 0.4 for different conditions of temperature and strain rate

have been summarized in Fig. 10 [5, 7, 8, 30]. The dashed lines in Fig. 10 show the α/α+β

and α+β/β borders calculated from the equilibrium phase diagram in Fig. 3(a). According to

the equilibrium phase diagram (Fig. 3(a)), β appears in the single-phase α microstructure

when Zn exceeds 35-38 wt.% depending on temperature. Similarly, when Zn exceeds 40-43

wt.%, β is substituted for the duplex microstructure.

Irrespective of deformation temperature or strain rate, the results imply that Zn tends to

decrease the strength of brass, especially in the α and α+β phase regions. Unlikely, for alloys
9
with Zn below 10-15 wt. percent or in a wholly β alloy the strength remains nearly

unchanged. This was also reported by Spigarelli et al. who clearly indicated the higher

strength of α with respect to β and highlighted the negative effect of Zn on the hot strength

of brass alloys in range of 30-44 wt.% [31]. The influence of Zn on the weakening of brass

should be associated with high temperatures because, at low temperatures, Zn behaves as a

strong solid solution hardener for copper. The results lead us to the idea that the effect of Zn

on the weakening of brass at high temperatures would be in connection with the mutual

diffusion coefficient in Cu-Zn system. The diffusivities of Zn and Cu are often formulated

using the following Arrhenius-type by equations [32, 33]:

189924
D Zn = 0.28exp(− ) cm 2 /s
RT (4)
212100
D Cu = 0.78exp(− ) cm 2 /s
RT
Although the frequency factor (pre exponential coefficient) and the activation energy for

diffusion in Eq. (4) are dependent on the chemical composition [31], in particular on the

amount of Zn, the above equations can be qualitatively used to account for the effect of Zn on

the hot strength of copper.

It is clear that under hot working temperatures the diffusion coefficient of Zn in Cu is higher

than that of self-diffusion in copper and therefore the mutual diffusion coefficient increases

with increasing Zn in a Cu-Zn system. Higher diffusivity leads to more expeditious DRV in a

high-Zn brass and the strength level and work hardening rate descends thereby. Zn also tends

to stabilize β versus α and microstructure gradually turns from single-phase α for low-Zn

alloys to α+β and single-phase β for high-Zn alloys. The microstructural change with the

variation of Zn also keeps the decreasing trend of strength, because β is a high SFE material

which undergoes DRV at high temperatures.

10
Fig. 10 can be used to determine the specific strength of α and β at any deformation

temperature and strain rate. For this purpose, at a given temperature, the Zn content of each

phase is simply determined from Fig. 9 and then the specific strengths are estimated from

Fig. 10. It is to be noted that as Zn contents of α and β in a duplex alloy are very close

(according to Fig. 9), the specific strengths does not differ a lot. This, in turn, eliminates

considerable strain partitioning between α and β, unlike to other two-phase systems such as

duplex stainless steels [21]. Therefore, the law of mixture rule (Eq. (3)) can be used in

combination with Figs. 9 and 10 to predict the flow stress of the duplex brass at any

deformation condition. It is to be noted that in Eq. (3) the effect of Pb has been ignored and

therefore the discrepancy between predicted and flow stresses reflects the influence of

phosphor. The experimental results and law of mixture predictions are compared in Figs. 11

and 12, respectively for CuZn39Pb2 and CuZn39Pb3.

The results show that at low temperatures (e.g. 650°C), both materials exhibit higher strength

than that is predicted by the law of mixture. This can be associated with the effect of Pb on

expanding the region of α+β (Fig. 3) that results in increasing the weight percent of α. At

high temperatures, particularly beyond 700°C, α decreases and Pb tends to more dissolve into

β. The dissolved Pb atoms effectively decrease the strength of Pb-bearing alloys with respect

to the strength of corresponding CuZn39, as predicted by the law of mixture. A simple

statistical analysis shows that the average difference between "experimental" and "law of

mixture" curves for CuZn39Pb2 and CuZn39Pb3 are respectively as 11.5% and 13.2%, that

emphasize on the effect of Pb.

4- Conclusions

11
Hot compression tests were conducted on two leaded brass of CuZn39Pb2 and
CuZn39Pb3 in temperature range of 600 °C-800 °C at strain rates of 0.001 s-1-1 s-1.
The major results of this research can be summarized as follows:
1) The stress-strain curves at low temperatures were characterized by a single peak. At
high temperatures, the flow curves were characterized by a long plateau associated
with dynamic recovery.
2) The phase diagram of Cu-Zn-Pb system drawn by the Thermocalc software showed
that increasing Pb could shift the phase border between α+β and β towards the Zn
corner, thereby increasing the volume fraction of α at the deformation tempearutes.
3) Hot deformation at low temperatures, e.g. 600 °C- 700 °C, changed the strings of
cornered α islands to more globularized descrete ones. The conventional dynamic
recrystallization could not be observed. At 800 °C, fine α particles formed through
dynamic recrystallization were coexistent with acicular ones formed through
quenching of β from high temperatures.
4) It was found that at low temperatures, e.g. below 700 °C- 750 °C, Pb can contribute to
avoid flow localization in both alloys. At 800 °C, more Pb could dissolve into β
leading to more tendency to flow locallization.
5) The results showed that at low temperatures, i.e. below 650°C -700°C, both materials
exhibit higher strength than that predicted by the law of mixture. At high temperatures,
particularly beyond 650°C -700°C, α decreases and Pb tends to more dissolve into β.
Therefore, the predicted values of the law of mixture lie below the experimental flow
stresses.

References

[1] L.V. Zhuravel, V.A. Yatsenko, A.G. Ivashkevich, Effect of lead on plastic deformation of

brass, Met. Sci. Heat Treat., 18 (1976) 247-248.

[2] L. Blaz, Z. Konior, T. Majda, Structural aspects of α/β transformation in hot deformed

CuZn-39Pb3 alloy, J. Mater. Sci., 36 (2001) 3629-3635.

[3] Y. Jang, S. Kim, S. Han, Effect of misch metal on elevated temperature tensile ductility of

the Cu-Zn-Bi alloy, Met. Mater. Trans., 36A (2005) 1060-1065.

12
[4] M. Sharififar, S.A.A. Akbari Mousavi, Tensile deformation and fracture behavior of

CuZn5 brass alloy at high temperature, Mater. Sci. Eng., A594 (2014) 118-124.

[5] Y.-H. Xiao, C. Guo, X.-Y. Guo, Constitutive modeling of hot deformation behavior of

H62 brass,Mater. Sci. Eng., A528 (2011) 6510-6518.

[6] M. Hatherly, A.S. Malin, C.M. Carmichael, F.J. Humphreys, J. Hirsch, Deformation

processes in hot worked copper and α brass, Acta Metall., 34 (1986) 2247-2257.

[7] D. Padmavardhani, Y.V.R.K Prasad, Characterization of hot deformation behavior of

brasses using processing maps: Part I β Brass and α-β Brass Metall. Trans., 22A (1991)

2993-3001.

[8] D. Padmavardhani, Y.V.R.K Prasad, Characterization of hot deformation behavior of

brasses using processing maps: Part I. Į Brass, Metall. Trans., 22A (1991) 2985-2992.

[9] A. Momeni, G.R. Ebrahimi, M. Jahazi, P. Bocher, Microstructure evolution at the onset of

discontinuous dynamic recrystallization: A physics-based model of subgrain critical size, J.

Alloys Comp., 587 (2014) 199-210.

[10] X.-M. Chen, Y.C. Lin, D.-X. Wen, J.-L. Zhang, M. He, Dynamic recrystallization

behavior of a typical nickel-based superalloy during hot deformation, Mater. Design, 57

(2014) 568-577.

[11] A. Momeni, K. Dehghani, G.R. Ebrahimi, Modelling the initiation of dynamic

recrystallization using a dynamic recovery model, J. Alloys Comp., 509 (2011) 9387-9393.

13
[12] T.I. Chashchukhina, L.M. Voronova, M.V. Degtyarev, D.K. Pokryshkina, Deformation

and dynamic recrystallization in copper at different deformation rates in Bridgman anvils,

Phys. Met. Metallogr., 111 (2011) 304-313.

[13] A. Rohatgi, K.S. Vecchio, G.T. Gray III, The influence of stacking fault energy on the

mechanical behavior of Cu and Cu-Al alloys: Deformation twinning, work hardening, and

dynamic recovery, Met. Mater. Trans., 32A (2001) 135-145.

[14] H. Farnoush, A. Momeni, K. Dehghani, J.A. Mohandesi, H. Keshmiri, Hot deformation

characteristics of 2205 duplex stainless steel based on the behavior of constituent phases,

Mater. Design, 31 (2010) 220-226.

[15] A. Momeni, S. Kazemi, A. Bahrani, Hot deformation behavior of microstructural

constituents in a duplex stainless steel during high-temperature straining, Int. J. Minerals.

Metall. Mater., 20 (2013) 953-960.

[16] S.S. Babu, J. Livingston, J.C. Lippold, Physical Simulation of Deformation and

Microstructure Evolution During Friction Stir Processing of Ti-6Al-4V Alloy, Met. Mater.

Trans., 44A (2013) 3577-3591.

[17] L. Sahebdel, S.M. Abbasi, A. Momeni, Microstructural evolution through hot working

of the single-phase and two-phase Ti-6Al-4V alloy, Int. J. Mater. Res., 102 (2011) 41-47.

[18] A. Momeni, K. Dehghani, Hot working behavior of 2205 austenite-ferrite duplex

stainless steel characterized by constitutiv equations and processing maps, Mater. Sci. Eng.,

A528 (2011) 1448-1454.

[19] J.H. Kim, S.L. Semiatin, Y.H. Lee, C.S. Lee, A self-consistent approach for modeling

the flow behavior of the alpha and beta phases in Ti-6Al-4V, Met. Mater. Trans., 42A (2010)

1805-1814.

14
[20] S.L. Semiatin, F. Montheillet, G. Shen, J.J. Jonas, Self-consistent modeling of the flow

behavior of wrought alpha/beta titanium alloys under isothermal and nonisothermal hot-

working conditions, Met. Mater. Trans., 33A (2001) 2719-2727.

[21] A. Momeni, K. Dehghani, M.C. Poletti, Law of mixture used to model the flow behavior

of a duplex stainless steel at high temperatures, Mater. Chem. Phys., 139 (2013) 747-755.

[22] S. Spigarelli, M. El Mehtedi, P. Ricci, C. Mapelli, Constitutive equations for prediction

of the flow behavior of duplex stainless steels, Mater. Sci. Eng., A527 (2010) 4218-4228.

[23] A. Momeni, K. Dehghani, M. Heidari, M. Vaseghi, Modeling the flow curve of AISI

410 martensitic stainless steel, J. Mater. Eng. Perform., 21 (2012) 2238-2243.

[24] P. Perrot, Non-Ferrous Metal Systems. Part 2, Landolt-Börnstein - Group IV Physical

Chemistry, 2007, 11C2, 408-419.

[25] W. Roberts, Dynamic changes that occur during hot working and their significance

regarding microstructural development and hot workability, ASM materials science seminar

on Deformation, Processing, and Structure, G. Krauss (Ed.), 23-24 Oct. 1982, Missouri,

USA, 109-184.

[26] A. Momeni, K. Dehghani, X.X. Zhang, Mechanical and microstructural analysis of 2205

duplex stainless steel under hot working condition, Journal of Materials Science 47 (2012),

2966-2974.

[27] Y.C. Lin, X.M. Chen, A critical review of experimental results and constitutive

descriptions for metals and alloys in hot working, Mater. Design, 32 (2011) 1733-1759.

[28] Y.C. Lin, M.-S. Chen, J. Zhong, Effect of temperature and strain rate on the compressive

deformation behavior of 42CrMo steel, J. Mater. Process. Technol., 205 (2008) 308-315.

[29] A. Momeni, K. Dehghani, G.R. Ebrahimi, Sh. Kazemi, Developing the processing maps

using the hyperbolic sine constitutive equation, Met. Mater. Trans., 44A (2013) 5567-5576.

15
[30] Hot working guide, a compendium of processing maps, Ed. Y.V.R.K. Prasad and S.

Sasidhara, ASM, Materials Park, Ohio, USA, 1997, 203-233.

[31] S. Spigarelli, M.El Mehtedi, M.Cabibbo, F.Gabrielli, D.Ciccarelli, High temperature

processing of brass: Constitutive analysis of hot working of Cu–Zn alloys, Mat. Sci. Eng.,

A615 (2014) 331-339.

[32] D.V. Butrymowicz, J.R. Manning, M.E. Read, Diffusion in copper and copper alloys,

Part I, volume and surface self-diffusion in copper, J. Phys. Chem. Ref. Data, 2 (1973) 643-

655.

[33] M.B. Dutt, S.K. Sen, Diffusion of zinc in copper and silver, Japan. J. Appl. Phys., 18

(1979) 1025-1029.

Figure Captions
Fig. 1. Microstructure of the as-received materials used in this investigation: (a) CuZn39Pb2,
(b) CuZn39Pb3.
Fig. 2. Representative true stress–strain curves of the studied alloys at various temperatures

and strain rates of (a) 0.001 s-1, (b) 1 s-1.

Fig. 3. Phase diagram of Cu-Zn with two different percentages of Pb drawn using the

Thermocalc software.

Fig. 4. Weight percent of α as a function of temperature in the studied alloys calculated from

the phase diagram of Fig. 3.

Fig. 5. Micrographs of the hot deformed samples from CuZn39Pb2 (a, c, e) and CuZn39Pb3

(b, d, f) at different deformation temperatures of (a, b) 600 °C, (c, d) 700 °C, and (e, f) 800

°C.

Fig. 6. Constitutive dependence of flow stress on strain rate according to Eq. (2).

Fig.7. Variation of stress exponent (defined in Eq. (1)) with temperature in the studied alloys.

16
Fig. 8. Constitutive dependence of flow stress on the reciprocal of temperature according to

Eq. (2).

Fig. 9. Variation of chemical composition (weight percentage of Zn) of α and β with

temperature in the studied alloys.

Fig. 10. Flow stress of a pure brass (Cu-Zn) alloy at a typical strain of 0.4 as a function of the

percentage of Zn at different temperatures and strain rates of (a) 0.001 s-1, (b) 0.01 s-1, (c) 0. 1

s-1, (d) 1 s-1.

Fig. 11. Comparison of the flow stress determined experimentally at typical strain of 0.4 and

the law of mixture predictions for CuZn39Pb2.

Fig. 12. Comparison of the flow stress determined experimentally at typical strain of 0.4 and

the law of mixture predictions for CuZn39Pb3.

Table captions

Table1. Chemical composition of brass alloys used in this investigation.

17
Alloy Zn Pb Sn Fe Ni Cu

CuZn 39. 2.2 0.0 0.0 <0. Re

39Pb2 4 9 36 48 005 m.

CuZn 39. 2.8 0.0 0.0 <0. Re

39Pb3 8 7 94 43 005 m.

18
Figure1a
Figure1b
Figure2a
Figure2b
Figure3
Figure4
Figure5a
Figure5b
Figure5c
Figure5d
Figure5e
Figure5f
Figure6a
Figure6b
Figure7
Figure8a
Figure8b
Figure9
Figure10a
Figure10b
Figure10c
Figure10d
Figure11a
Figure11b
Figure11c
Figure11d
Figure12a
Figure12b
Figure12c
Figure12d

You might also like