Sie sind auf Seite 1von 4

Materials Chemistry and Physics 128 (2011) 539542

Contents lists available at ScienceDirect


Materials Chemistry and Physics
j our nal homepage: www. el sevi er . com/ l ocat e/ mat chemphys
Simple thermodynamic model of the extension of solid solution of CuMo alloys
processed by mechanical alloying
C. Aguilar
a,
, D. Guzman
b
, P.A. Rojas
c
, Stella Ordo nez
d
, R. Rios
e
a
Departamento de Ingeniera Metalrgica y de Materiales, Universidad Tcnica Federico Santa Mara, Avenida Espana 1680, Valparaso, Chile
b
Departamento de Metalurgia, Facultad de Ingeniera, Universidad de Atacama, Av. Copayapu 485, Copiap, Chile
c
Escuela de Ingeniera Mecnica, Facultad de Ingeniera, Ponticia Universidad Catlica de Valparaso, Av. Los Carrera 01567, Quilpu, Chile
d
Departamento de Ingeniera Metalrgica, Facultad de Ingeniera, Universidad de Santiago de Chile, Av. L. Bernardo OHiggins 3363, Santiago, Chile
e
Instituto de Materiales y Procesos Termomecnicos, Facultad de Ciencias de la Ingeniera, Universidad Austral de Chile, General Lagos 2086, Valdivia, Chile
a r t i c l e i n f o
Article history:
Received 21 December 2010
Received in revised form 25 February 2011
Accepted 21 March 2011
Keywords:
Thermodynamic properties
Powder metallurgy
Metals
Coherent X-ray scattering
a b s t r a c t
The objective of this work is proposing a simple thermodynamic model to explain the increase in the
solubility limit of the powders of the CuMo systems or other binary systems processed by mechanical
alloying. In the regular solution model, the effects of crystalline defects, such as; dislocations and grain
boundary produced during milling were introduced. The model gives results that are consistent with the
solubility limit extension reported in other works for the CuCr, CuNb and CuFe systems processed by
mechanical alloying.
2011 Elsevier B.V. All rights reserved.
1. Introduction
Copper is widely used in engineering because of their prop-
erties, such as, resistance to corrosion, electrical and/or thermal
conductivity, good fatigue resistance and easy fabrication [1,2].
Moreover, copper has the disadvantage of lowmechanical strength.
But its strength can be increased by adding solute atoms (e.g. Mo,
Cr, Nb, Fe) for form a solid solution. Interesting alloys are CuMo
alloys, which are used in applications of heat sink material, vac-
uum technology and electronic packing devices due their physical
and electronic properties [1,3]. In most of the applications, high-
density CuMo materials with homogeneous microstructure are
required for high performance [3]. The disadvantage of CuMo sys-
tems is that present a positive mixing enthalpy at solid state [4],
thus the solubility of Mo in Cu at room temperature is negligi-
ble [4]. But it is known that by using a non-equilibrium process
the range of solubility of many systems can be extended [57].
Mechanical alloying (MA) is a non-equilibrium process used for
extension of solubility [8]. During MA, due to the impact of the
ball, the energy is transferred to metallic powders, and so increases
the density of crystalline defects, such as grain boundary, disloca-
tions, vacancies, stacking fault, etc. [810]. It has been estimated
that the maximum departure from the equilibrium for materials

Corresponding author. Tel.: +56 32 2654228; fax: +56 32 2654228.


E-mail address: claudio.aguilar@usm.cl (C. Aguilar).
with a dislocation density of 10
16
mm
3
and a crystallite size of
1nm is 1kJ mol
1
and 10kJ mol
1
, respectively [11]. These values
showthe large amount of stored energy in nanocrystalline materi-
als, whichcanbe compared withthe standard heat value of melting
copper; 12.9kJ mol
1
[12]. In this work the effects of reduction of
crystallite size and increment of strain due to the presence of crys-
talline defects are considered in a regular solution model to explain
the increase in the solubility limit of Mo in Cu during MA.
2. Experimental procedure
The powder mixtures used during this work were obtained by blending 99wt.%
pure copper containing less than 1000ppm of oxygen, 99.9wt.%, and particle sizes
ranged between 170 and 400mesh; and 325 and 400mesh for Mo powders. In a
typical run, the copper powder was blended with 8wt.% Mo and 1wt.% of stearic
acid as the process control agent. The mixture was placed in a 25ml container and
milled in a SPEX 8000Dmill under an argon atmosphere for 0.5, 4, 8 and 50h, main-
taining a ball/powder load ratio of 10/1. All experiments were run with same milling
speed of 875 cycles min
1
(clamp speed), which was measured with a stroboscopic
light. X-ray diffraction was recorded with a Siemens D5000 diffractometer of Cu
K
1
radiation (0.154056nm). The whole patterns were measured over an angu-
lar range of 20 =38122

. The X-ray prole analysis was made using the modied


Warren-Averbach method [13], using annealed pure copper as a standard.
2.1. Determination of crystallite size and dislocation density
The basic equation of the modied Warren-Averbach method is the expression
1, where A(L) is the real Fourier coefcient, A
S
(L) is the size contribution, L =na
3
,
a
3
=z/2(sin0
2
sin0
1
), (0
2
0
1
) the angular range of the measured diffraction pro-
le, zis thewavelengthof X-rays, B=b
2
/2, bis theBurger vector, is thedislocation
density, Re is the outer cutoff radius of dislocations, R
1
andR
2
are auxiliary constants
0254-0584/$ see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.matchemphys.2011.03.047
540 C. Aguilar et al. / Materials Chemistry and Physics 128 (2011) 539542
Fig. 1. Effect of the milling time on the crystallite size and the dislocation density
of Cu8wt.% Mo.
not interpreted physically, Q is the correlation factor and stands for higher order
terms in K
2
C, K=2sin0/z and C is the average dislocation contrast factor for a
particular reection.
ln/(L)

= ln/
S
(L) 8L
2
ln(
Rc
L
)(K
2
C) +U8
2
L
4
ln(
R
1
L
) ln(
R
1
L
)(K
2
C)
2
(1)
The real part of the Fourier coefcients, A(L) are plotted against K
2
C and are
tting to a parabolic equation. The intersections of the quadratic curves at K=0
provide a crystallite sizes (the values of A
S
(L) are plotted against L and then the
crystallite size can be determined from the intercept of the initial slope on the L-
axis). The dislocation density, , can be determined from slopes of the quadratics
curves [13].
3. Results and discussion
3.1. Milling
Fig. 1 shows the evolutionof crystallite size anddislocationden-
sity as function of milling time of Cu8wt.% Mo. This gure also
shows that crystallite size decreases to values of around 10nm
at 50h of milling which is in agreement with previous results
[14]. Additionally, a smaller crystallite size was obtained with a
Cu8wt.% Mo than pure copper suggesting that crystallite size
is affected by chemical composition due to solid solutions being
harder and stronger than pure metal. Thus, when solute atoms are
dissolved into the solid solution, they increase the hardness of the
solution so that its fragmentation tendency is increased. This fact
is in agreement with the results reported by Eckert et al. [15]. Also,
it can be observed that when milling time increases, the values of
dislocation density increase due to the severe plastic deformation
of the CuMo powders produced by the shock of balls. At 50h of
milling time, the value of dislocation density is slightly lower than
the value obtained at 8h of milling time suggesting that the mini-
mal crystallite size was reached in both alloys and further milling
would not produce more dislocations, which is in agreement with
a previous work [16]. This is explained by the difculty of generat-
ing dislocations at very small crystallite size, whereby the existing
dislocations are rearranged and some will be annihilated, causing
the strain and the dislocation density to show a decrease.
Fig. 2 shows the X-ray proles of Cuwt.% Mo alloys. This g-
ure shows the evolution of the Cu and Mo peaks as a function of
milling time. In gure the strongest peaks (110) for Mo remain
up to 8h of milling then disappear at 50h of milling. Thus, when
the Mo peaks are not observed in the X-ray pattern, it can be con-
sidered that atoms of Mo are dissolved in copper. On other hand,
the strongest peak (111) for Cu shows a broadening due to high
degree of cold deformation (increasing of crystalline defects) and
microstructural renement (decrease of crystallite size) produced
Fig. 2. XRD pattern for various milling times of Cu8wt.% Mo.
during milling [8]. Finally, the lattice parameter, which was cal-
culated by Cohens method [17], increased when milling time is
increased, until 0.3618nm at 50h of milling. This is because the
atomic radius of Mo is greater than the atomic radius of Cu [18].
Therefore Cu8wt.% Mo solid solution is formed at a milling time
of over 8h.
3.2. Thermodynamic model
When crystallite size decreases the free energy of the system
increases, which can be determined by using Eq. (2), where; ) is
the surface energy, M the mean curvature of the surface at point
p, and M1/2(1/r
1
+1/r
2
). The radii r
1
and r
2
are the principal radii
of curvature. The total strain energy due to the presence of dislo-
cations in materials can be estimated using Eq. (3), where o = ,
is the dislocation elastic energy per unit length of dislocation lines
and is the dislocation density. The term oV
m
considers the effect
of the hydrostatic component of the stress tensor due to the pres-
ence of dislocations. In MA, the dislocation density has a high value
due to several plastic deformations produced by the ball impact
and the grain sizes are of nano order [19,20].
^C
cs
= 2) Mv
m
(2)
^C
s
= ov
m
(3)
The regular solution model will be applied to several systems.
The advantage of this thermodynamic model is its simplicity of
mathematic formalism. Therefore, the free energy can be written
thus [21]:

C
j
i
= C
j
i
+(1 x
i
)
2
+R1 lnx
i
(4)
where i is A or B component, j is or phase,

C
j
i
is the partial molar
Gibbs free energy of i component in j phase, C
j
i
is the molar Gibbs
freeenergyof i component inj phase(or themolar Gibbs freeenergy
standard), =^H
M
/(x
A
x
B
), ^H
M
the enthalpy of mixing, x
A
and x
B
are mole fractions of A and B components, respectively, R is the gas
constant and T the temperature. The Eq. (4) can be expressed for
both phases, and , in a binary system, but it is only necessary to
knowx
B
in the phase (considering A=Cu, B=Mo, phase rich in A
andphase richinB). Thus

C

8
is the partial molar Gibbs free energy
of B component in the phase, G
B

is the molar Gibbs free energy


C. Aguilar et al. / Materials Chemistry and Physics 128 (2011) 539542 541
Fig. 3. Schematic gure shows the change of xB as a function of temperature for
different Q values, considering Q=K/.
of pure B in the phase. Thus Eq. (4) can be expressed as:

C

8
=
C

8
+(1 x
8
)
2
+R1 lnx
8
. It is necessary to know the free energy
change when Mo atoms (or other solute atom) change from bcc to
fcc solutions (B

) which can be expressed as ^C


8
= C

8
C

8
,
which is the difference in free molar energy between B atoms in the
and phase. The maximum concentration of B in (x
B
) is given
by the condition (a)

C

8
=

C

8

= C

8
, thus Eq. (4) yields:
^C
8
= C

8
C

8
= (1 x
8
)
2
R1 lnx
8
(5)
After milling the free curves of both phases, and , are
shifted by the amounts E

and E

, respectively. Thus, ^G
B
(after
milling) =(G
B

+E

) (G
B

+E

) = ^G
B
(before milling) +E

.
Assuming that E

is only due to the effects of grain boundaries


and stress and E

is assumed to be zero due to the fact that Mo


atoms enter in solid solution in Cu, then; ^G
B
(after milling) =
^G
B
(before milling) +E

, where E

=^G
cs
+^G
s
. Therefore, grain
boundaries and stresses, produced by decreasing the grain bound-
ary and increasing the dislocation densities, respectively, change
the free energy and lead to an enhanced solubility. Therefore, the
Eq. (5) can be modied by introducing these effects given by Eqs.
(2) and (3), as shown in Eq. (6) (where B can be any solute atom).
^C
8
= (1 x
8
)
2
R1 lnx
8
+ov
m
+2)Mv
m
(6)
Thus, the solubility of B in A (e.g. Mo in Cu), considering the new
effects is obtained fromEq. (6), and can be expressed as the Eq. (7).
x
8
= exp

^C
8
+(1 x
8
) (o +2)M)v
m
R1

(7a)
x
8
= exp

K /
R1

(7b)
If the solubility of solute atoms is smaller, it can be considered
as x
B
<1, thus (1x
B
) 1, and the Eq. (7a) can be solved easily.
But if x
B
is not smaller, the Eq. (7) must be solved with numer-
ical methods. The Eq. (7a) can be expressed as Eq. (7b), where
K=^G
B
+(1x
B
) and /=(o +2)M)V
m
. If there are no effects of
grain boundaries and stresses, the term/will be zero, and only the
term K will remain. Term / denotes external energy contribution
produced during MA and thus this contribution shifts the system
away fromthe equilibrium. In conclusion, /value is the necessary
energy for the departure from the equilibrium. The K value is posi-
tive for immiscible systems suchas CuMo. Greater /value implies
smaller K/ values and greater solubility of solute atoms in sol-
vent atoms as shown in Fig. 3 (considering Q=K/). If / value
Fig. 4. Comparison of the extended solid solubility limits of several solute elements
in copper achieved by MA with values obtained applied Eq. (6).
increases enough on the energy value to form BA bonds, enough
driving force exists for solute atoms to enter into a solid solution.
Thus, / xes the departure from equilibrium and the new solu-
bility of B in A. Therefore, it is proposed to consider / as the term
departure from the equilibrium.
Eq. (7) was applied to four systems, CuMo, CuCr, CuNb and
CuFe using the data listed in Table 1. Data of the CuMo system
was determined from X-ray diffraction peak prole analysis made
by our group (Fig. 1). Data of CuNb [2225], CuCr [26] and CuFe
[27,28] were taken from works reported in the literature. The
values were determined using ^H
M
=X
Cu
X
B
(where B is solute
atoms and can be Mo, Nb, Cr or Fe), the crystallite size was taken
as 7nm due to results obtained from Fig. 1 for CuMo and results
reported by other works [24,28]. The surface energy used was the
value of surface energy grain boundary/grain boundary for pure
copper, ) =625mJ m
2
[29]. The temperature was taken as 550K
as some previous studies have reported an increase in the internal
temperature of up to approximately this value [30,31]. The stress
acting on a solute atom located at a distance r from a dislocation
core is given by Eq. (8) [27], where; is the shear modulus, b the
Burgers vector, D the Poissons ratio of the metal, and 0 the angle
about the glide plane. The considered for Cu was 48.3GPa [32]
and r is smallest crystallite size determined from X-ray analysis,
Fig. 1, which was taken as 10nm. Thus, a value of o of 0.53MPa
was calculated. An approximation for the stress, can be considered
as /30 [33].
o =
b(1 +D) sin0
3(1 D) r
(8)
The Eq. (7) was solvednumerically using the bisectionandNew-
ton methods. The x
B
(solid solubility) values for the CuMo, CuCr,
CuNb and CuFe systems are given by Fig. 4 and are compared
with the extended solid solubility limits achieved by MA reported
by Suryanarayana [8]. It may be noted that the solid solubility val-
ues of Cr, Nb and Fe in Cu determined by Eq. (7) are close to the
values reported, and for Mo there are no values reported. Thus, it
can be said that in general the Eq. (7) gives reasonable results for
the four systems.
For the Cu8wt.% Mo alloy studied in this work, the / term is
large enough to allow solid solutions to form from stored mechan-
ical energy at 50h of milling time. This result is in agreement with
the fact that neither Mo peak is observed in the X-ray pattern at
50h of milling time, and due that the lattice parameter increased
whenmilling time is increased. Also, it canbe saidthat these results
are in agreement with those reported by Echert et al. [15,27] and
542 C. Aguilar et al. / Materials Chemistry and Physics 128 (2011) 539542
Table 1
Data of the systems for solve Eq. (6).
CuMo CuCr CuFe CuNb
^H
M
(kJ mol
1
) 4.3 [4]

=6.5 [39] 11.9 [25] 6.0 [22]
(kJ mol
1
)

=83

=79

=50

=32
^GB (kJ mol
1
)

=18 [40,41]

=7.5 [39]

=4.8 [25]

=10 [23]
Mo Cr Fe Nb
Vm (m
3
mol
1
) [42] 9.4010
6
7.23910
6
7.10510
6
1.08610
5
Schwarz [34]. They suggested a mechanism for the formation of
solid solution based on interaction between dislocation eld and
solutes. There is a preferential segregationof solute atoms into core
regions of dislocations, which are abundant in powder processing
by MA [8]. Nanomaterials obtained by MA have higher stress and
interfacevalues duetohighdislocationdensityandsmall crystallite
size. At 50hof milling a highdislocationdensity anda small crystal-
litesizecanbeobserved, whichincreasethefreeenergyof powders.
Another interesting aspect is that nanomaterials have log-normal
crystallite size [35,36], and particles distributions [37], thus there
is a fraction of crystallites with sizes smaller than 10nm as had
been reported by other authors [20,35], which increase the solid
solubility limit. Yavari et al. [38] proposed that capillary forces in
fragments with a small neck or a tip with a radius of 2nminuence
increase in solubility in immiscible systems and a fraction of grains
have grain sizes of around 1 and 2nm.
4. Conclusions
The thermodynamic model proposed gives results of solubility
change for CuMo systems that are in agreement with experimen-
tal data obtained fromX-ray diffraction patterns. Also gives results
that are consistent with the solubility limit extension reported in
other works for the CuCr, CuNb and CuFe systems processed by
MA.
In the application of the model to the CuCr, CuNb and CuFe
systems, some assumptions have to be made (e.g. surface energy
and shear modules were assumed as pure Cu due to the low per-
centage of Mo used). Therefore the introduction of more accuracy
values in the model should give better results.
Acknowledgements
The authors acknowledge the support given by CONICYT, Fondo
de Desarrollo Cientco y Tecnolgico de Chile, FONDECYT, project
No 11070052.
References
[1] J.L. Johnson, Int. J. Powder Metall. 35 (1999) 3948.
[2] Z.A. Ali, O.B. Drury, M.F. Cunningham, IEEE Trans. Appl. Supercond. 15 (2005)
5269.
[3] A. Sun, D. Wang, Z. Wu, X. Zan, J. Alloys Compd. 505 (2010) 588591.
[4] P.R. Subramanian, DE Laughlin, Bull. Alloy Phase Diagrams 11 (1990) 169.
[5] E. Gaffet, C. Louison, M. Harmelin, F. Faudet, Mater. Sci. Eng. A 134 (1991)
13801384.
[6] J. Kuyama, K.N. Ishihara, P.H. Shingu, Mater. Sci. Forum 88/90 (1992) 521528.
[7] C. Suryanarayana, Non-Equilibrium Processing of Materials, 1 ed., Pergamon,
1999.
[8] C. Suryanarayana, Mechanical Alloying and Milling, rst ed., Marcel Dekker,
New York, 2004.
[9] P.S. Soni, Mechanical Alloying, rst ed., Cambridge International Science Pub-
lishing, U.K., 2000.
[10] L. Lu, M.O. Lai, Mechanical Alloying, Kluwer Academic Publishers, USA, 1998.
[11] F.H. Froes, C. Suryanarayana, K.C. Russell, M. Ward, Proceeding of Symposium
held at Materials Week, The Minerals, Metals & Materials Society (TMS) and
the Materials Information Society, ASM International, 1994, pp. 121.
[12] O. Kubaschewski, C.B. Alcock, Metallurgical Thermochemistry, fth ed., Perga-
mon Press, Oxford, 1979.
[13] T. Ungr, A. Borbly, Appl. Phys. Lett. 69 (1996) 31733175.
[14] A. Al-Hajry, M. Al-Assiri, S. Al-Heniti, S. Enzo, J. Efne, A. Shahrani, A.E. Salami,
Mater. Sci. Forum 386/388 (2002) 205210.
[15] J. Eckert, J.C. Holzer, C.E. Krill III, W.L. Johnson, J. Mater. Res. 7(1992) 19801983.
[16] I. Lucks, P. Lamaparter, E.J. Mittemeijer, Acta Mater. 49 (2001) 24192428.
[17] B.D. Cullity, S.R. Stock, Elements of X-ray Diffraction, 3 ed., Addison-Wesley,
2001.
[18] W.D. Callister, Materials Science andEngineering: AnIntroduction, 7ed., Wiley,
2006.
[19] T. Ungr, J. Gubicza, P. Hank, I. Alexandrov, Mater. Sci. Eng. A 319/321 (2001)
274278.
[20] T. Ungr, J. Gubicza, G. Ribrik, A. Borbly, J. Appl. Crystallogr. 34 (2001)
298310.
[21] D.R. Gaskell, Introduction to the Thermodynamics of Materials, 5 ed., Taylor &
Francis, USA, 2008.
[22] E. Ma, Scripta Mater. 49 (2003) 941.
[23] A. Puthucode, M.J. Kaufman, R. Banerjee, Metall. Mater. Trans. A 39 (2008)
15781584.
[24] R. Tomasi, E.M.J.A. Pallome, F.W. Botta, Mater. Sci. Forum 312314 (1999)
333338.
[25] E. Botcharova, J. Freudenberger, L. Schultz, J. Alloys Compd. 365(2004) 157163.
[26] C. Aguilar, J. Marn, S. Ord nez, D. Celentano, F. Castro, V. Martnez, Rev. Metal.
42 (2006) 334344.
[27] J. Eckert, J.C. Holzer, C.E. Krill III, W.L. Jonson, J. Appl. Phys. 73 (1993) 131141.
[28] Z. Caama no, G. Prez, L.E. Zamora, S. Suri nach, J.S. Mu noz, M.D. Bar, J. Non-
Cryst. Solids 287 (2001) 1519.
[29] R. Dehoff, Thermodynamics in Materials Science, McGraw-Hill Inc., Singapore,
1993.
[30] R.M. Davis, B. McDermontt, C.C. Koch, Metall. Trans. A 19 (1988) 28742876.
[31] R. Schulz, M. Trudeau, J. Huot, A. Van Neste, Phys. Rev. Lett. 62 (1989)
28492852.
[32] R. Hertzberg, Deformation and Fracture Mechanics of Engineering Materials,
John Wiley & Sons, USA, 1988.
[33] D. Hull, D.J. Bacon, Introduction to Dislocations, Pergamon, Oxford, 1984.
[34] R.B. Schwarz, Mater. Sci. Forum 269/272 (1998) 665674.
[35] C.E. Krill, R. Birringer, Philos. Mag. A 77 (1998) 621640.
[36] Y. Zhong, D. Ping, X. Song, F. Yin, J. Alloys Compd. 476 (2009) 113117.
[37] Z. Hong, T. Wen-Ming, X. Guang-Qing, W. Yu-Cheng, Z. Zhi-Xiang, Mater. Chem.
Phys. 122 (2010) 64.
[38] A. Yavari, P.J. Desr, T. Benamuer, Phys. Rev. Lett. 68 (1992) 22352238.
[39] C. Michaelsen, C. Gente, R. Bormann, J. Mater. Res. 12 (1997) 14631467.
[40] B. Zhao, D.M. Li, F. Zeng, F. Pan, Appl. Surf. Sci. 207 (2003) 334340.
[41] Y.G. Chen, B.X. Liu, J. Alloys Compd. 261 (1997) 217224.
[42] HSC 6.0, Outokumpu Research Oy, 19742006.

Das könnte Ihnen auch gefallen