comparison between kinetic and thermodynamic effects on grain growth

6
Comparison between kinetic and thermodynamic effects on grain growth Feng Liu a,b, * , Reiner Kirchheim b a Max-Planck-Institute for Metals Research, Heisenbergstr. 3, Stuttgart 70569, Germany b Institute for Materialphysik, University of Gottingen, Tammannstr. 1, Gottingen 37077, Germany Received 14 August 2003; received in revised form 21 January 2004; accepted 2 March 2004 Available online 27 April 2004 Abstract A detailed comparison between the kinetic and the thermodynamic effects on grain growth was given, and the corresponding models were fitted to the experimental data, respectively. It was found that both models can explain the derivation from the normal parabolic growth under ideal condition. According to the kinetic model, a single isothermal grain growth can be understood in terms of a single, thermally-activated rate process with constant Q and grain boundary (GB) energy, r b ; impurity atoms accumulated in the GBs might exert a retarding force on GB migration, but do not change the grain growth activation energy, Q. On the other hand, it is necessary to invoke variable Q and r b according to the thermodynamic model, where the probably existing impurities and/or surface oxide seem to block surface and/or GB diffusion path, thus increasing Q and reducing r b . D 2004 Elsevier B.V. All rights reserved. Keywords: Nanostructures; Grain boundary; Segregation; Diffusion 1. Introduction The stability of nanocrystalline (NC) alloy with res- pect to grain growth is discussed controversially. Some authors [1–3] claim that solute drag by the alloying or impurity atoms reduces the mobility of the grain bound- aries (GBs), whereas others attribute it to a vanishing driving force [4–6]. For NC materials, the large change in total GB area accompanying grain growth would greatly affect the grain growth kinetics. As GB area diminishes, the concentration of solute or impurity atoms segregated to the GBs is expected to increase rapidly and introduces a grain-size-dependent retarding force on GB migration. Considering this grain-size-dependent drag force, Michels et al. [2] obtained the following grain growth equation: D t ¼ D 2 max ðD 2 max D 0 Þexp 2kt D 2 max & ' 1=2 ð1Þ where D t is the final grain size, D 0 the initial grain size, D max the maximum grain size, t the annealing time and k the temperature-dependent rate constant. It is implied in the last equation that the grain growth activation energy Q and the GB energy r b are assumed to be constant. For NC alloy with a large fraction of interfacial volume, however, r b , as the driving force for grain growth, cannot be kept constant throughout the whole process. According to the Gibbs adsorption equation [7,8] and the thermodynamic treatment of Weissmu ¨ller [4] and Kirch- heim [6] based on this concept, r b reduces with solute segregation. r b ¼ r 0 C b0 tRT ln X 0 þ DH seg b ð2Þ where r 0 is the GB energy for pure solvent, X 0 the bulk content, C b0 the solute excess of GB monolayer available for segregation at saturation and DH seg the enthalpy change of segregation per mole of solute. Whenever r b is positive, grain growth will decrease the free energy of the system. Since systems with r b <0 are not thermodynamically stable, the only case where grain growth can be suppressed is where r b =0 (see Section 2). 0040-6090/$ - see front matter D 2004 Elsevier B.V. All rights reserved. doi:10.1016/j.tsf.2004.03.018 * Corresponding author. Max-Planck-Institute for Metals Research, Heisenbergstr. 3, Stuttgart 70569, Germany. E-mail address: [email protected] (F. Liu). www.elsevier.com/locate/tsf Thin Solid Films 466 (2004) 108 – 113

Upload: feng-liu

Post on 02-Jul-2016

216 views

Category:

Documents


3 download

TRANSCRIPT

Page 1: Comparison between kinetic and thermodynamic effects on grain growth

www.elsevier.com/locate/tsf

Thin Solid Films 466 (2004) 108–113

Comparison between kinetic and thermodynamic effects

on grain growth

Feng Liua,b,*, Reiner Kirchheimb

aMax-Planck-Institute for Metals Research, Heisenbergstr. 3, Stuttgart 70569, Germanyb Institute for Materialphysik, University of Gottingen, Tammannstr. 1, Gottingen 37077, Germany

Received 14 August 2003; received in revised form 21 January 2004; accepted 2 March 2004

Available online 27 April 2004

Abstract

A detailed comparison between the kinetic and the thermodynamic effects on grain growth was given, and the corresponding models were

fitted to the experimental data, respectively. It was found that both models can explain the derivation from the normal parabolic growth under

ideal condition. According to the kinetic model, a single isothermal grain growth can be understood in terms of a single, thermally-activated

rate process with constant Q and grain boundary (GB) energy, rb; impurity atoms accumulated in the GBs might exert a retarding force on

GB migration, but do not change the grain growth activation energy, Q. On the other hand, it is necessary to invoke variable Q and rbaccording to the thermodynamic model, where the probably existing impurities and/or surface oxide seem to block surface and/or GB

diffusion path, thus increasing Q and reducing rb.D 2004 Elsevier B.V. All rights reserved.

Keywords: Nanostructures; Grain boundary; Segregation; Diffusion

1. Introduction

The stability of nanocrystalline (NC) alloy with res-

pect to grain growth is discussed controversially. Some

authors [1–3] claim that solute drag by the alloying or

impurity atoms reduces the mobility of the grain bound-

aries (GBs), whereas others attribute it to a vanishing

driving force [4–6]. For NC materials, the large change

in total GB area accompanying grain growth would

greatly affect the grain growth kinetics. As GB area

diminishes, the concentration of solute or impurity atoms

segregated to the GBs is expected to increase rapidly and

introduces a grain-size-dependent retarding force on GB

migration. Considering this grain-size-dependent drag

force, Michels et al. [2] obtained the following grain

growth equation:

Dt ¼ D2max � ðD2

max � D0Þexp � 2kt

D2max

� �� �1=2

ð1Þ

0040-6090/$ - see front matter D 2004 Elsevier B.V. All rights reserved.

doi:10.1016/j.tsf.2004.03.018

* Corresponding author. Max-Planck-Institute for Metals Research,

Heisenbergstr. 3, Stuttgart 70569, Germany.

E-mail address: [email protected] (F. Liu).

where Dt is the final grain size, D0 the initial grain size,

Dmax the maximum grain size, t the annealing time and k

the temperature-dependent rate constant. It is implied in

the last equation that the grain growth activation energy

Q and the GB energy rb are assumed to be constant. For

NC alloy with a large fraction of interfacial volume,

however, rb, as the driving force for grain growth, cannot

be kept constant throughout the whole process.

According to the Gibbs adsorption equation [7,8] and the

thermodynamic treatment of Weissmuller [4] and Kirch-

heim [6] based on this concept, rb reduces with solute

segregation.

rb ¼ r0 � Cb0 tRT lnX0 þ DHseg b ð2Þ

where r0 is the GB energy for pure solvent, X0 the bulk

content, Cb0 the solute excess of GB monolayer available

for segregation at saturation and DHseg the enthalpy change

of segregation per mole of solute. Whenever rb is positive,grain growth will decrease the free energy of the system.

Since systems with rb<0 are not thermodynamically stable,

the only case where grain growth can be suppressed is

where rb=0 (see Section 2).

Page 2: Comparison between kinetic and thermodynamic effects on grain growth

Fig. 1. Fits of Eq. (5) with s=2 (solid lines) and s=3 (dash-dotted lines) to

the grain sizes calculated from Eq. (1) at different temperatures.

F. Liu, R. Kirchheim / Thin Solid Films 466 (2004) 108–113 109

Grain growth in NC alloy is accompanied by a reduc-

tion in diffusion, i.e. an increase of activation enthalpy of

diffusion [9]. Upon grain growth, Q should be increased,

continuously, in contrast with a reduction of rb. Only if

grain growth stops with saturated GBs, rb could be zero

[6]. Approximately, a single process can be separated as

several domains, and Q and rb are kept constant within

each domain,

Z Ds

D0

DdD ¼Z D1

D0

DdDþ . . .þZ Ds

Ds�1

DdD

¼Z t1

t0

½A1rb1�exp � Qb1

RT

� �dt þ . . .

þZ ts

ts�1

Asrbsexp � Qbs

RT

� �� �ð3Þ

where A1,. . .,As is constant and rb1>rb2>. . .>rbs and Qb1<

Qb2<. . .<Qbs with s as the number of the domains. Inte-

grating Eq. (3) gives

D2s � D2

0 ¼ k1t1 þ k2ðt2 � t1Þ þ . . .þ ksðts � ts�1Þ ð4Þ

Table 1

Values of the rate constants and the relative errors obtained from fitting Eq. (5

temperatures

Different T Eq. (5), s=2 Eq. (5), s=3

k1 (nm2/s) k2 (nm

2/s) k1 (nm2/s) k2 (nm

2/s)

1 0.1595 0.005769 0.1442 0.01671

2 0.8861 0.04459 0.8042 0.1275

3 14.279 0.08361 13.061 0.4204

with k1>k2>. . .>ks. In general, the last equation can be

rewritten as

D2t � D2

0 ¼

k1t tVt1

k1t1 þ k2ðt � t1Þ t2 > t > t1

. . .

. . . ts�1 > t > ts�2

k1t1 þ k2ðt2 � t1Þ þ . . .þ ksðt � ts�1Þ ts > t > ts�1

8>>>>>>>>>>>>>>>><>>>>>>>>>>>>>>>>:

9>>>>>>>>>>>>>>>>=>>>>>>>>>>>>>>>>;

:

ð5Þ

Apparently higher value of s gives better fits to the

experimental data (see Section 3). Generally, s can be

chosen as 2 or 3.

By fitting both models (Eqs. (1), (3), (4) and (5)) to the

experimental data obtained in crystalline growth in nano-

nickel samples with different oxygen amounts at 673 K

[10] and in normal grain growth in silver thin film at

different temperatures [11], a detailed comparison between

the kinetic and the thermodynamic effects on grain growth

inhibition has been given in the present paper.

2. Comparison between kinetic effect and

thermodynamic effect

According to Eq. (1), Dt, at constant annealing temper-

ature T, is determined by D0, Dmax and k, and can be

described as a function of t (see Fig. 1). Eq. (1) is derived

by assuming a linear velocity/driving force (V/rb) relation

where the pinning force, assumed to be dependent on grain

size, is subtracted directly from the capillary force term

[1], whereas it is necessary to invoke variable Q and rbaccording to Eq. (5). Fitting Eq. (5) to Dt calculated from

Eq. (1) gives the values of k1 and k2 with s=2 or k1, k2 and

k3 with s=3, as well as the relative error of the fit, as

gathered in Table 1, respectively. Obviously, good fits are

obtained, and fit corresponding to s=3 is better than that

corresponding to s=2. This further indicates that it is

reasonable to assume variable Q and rb.Furthermore, a relation kik1

s=2ik1s=3>k2

s=2ik2s=3>k3

s=3,

deduced from Table 1, implies that higher T strengthens the

) with s=2 and s=3 to the grain sizes calculated from Eq. (1) at different

Eq. (1) Error of fit (%)

k3 (nm2/s) Dmax (nm) k (nm2/s) s=2 s=3

0.001384 15 0.1 2.9 2.1

0.002173 35 0.5 4.0 1.6

0.01173 120 10 3.1 1.5

Page 3: Comparison between kinetic and thermodynamic effects on grain growth

Table 2

Values of the rate constants and the relative errors obtained from fitting Eqs.

(1) and (5) (with s=2) to the grain sizes measured in nickel thin film with

oxygen contents between 956 and 6039 wt. ppm at 673 K [10]

T=673 K Eq. (5), s=2 Eq. (1) Error of fit (%)

ppm O2 k1(nm2/s)

k2(nm2/s)

Dmax

(nm)

k

(nm2/s)

Eq.

(5)

Eq.

(1)

956 53.561 0.1370 142 68.843 2.5 0.8

1805 9.1921 0.04928 62.2 9.6801 3 3.2

6309 1.7336 0.03181 33.6 1.8827 0.4 1.8

F. Liu, R. Kirchheim / Thin Solid Films 466 (2004) 108–113110

effect of solute segregation, and thus increases Q, e.g.

activation enthalpy for GB diffusion. Regarding both kinetic

and thermodynamic effects, which one controls the grain

growth inhibition is really a difficult problem.

Another mechanism for grain growth inhibition was

proposed considering the correlation between GB struc-

ture and grain growth [12]. The GBs at low temperature

tend to be ordered (facet or straight) with mobility much

lower than that of disordered (defacet or curved) GBs at

high temperature. When grain growth occurs at relatively

low temperature, the fraction of high energy GBs

decreases while that of low energy GBs increases. When

most of GBs are facet or straight, the mobility becomes

extreme low and the grain growth can practically stop.

Therefore, this interpretation indicates that, although rb

decreases with segregation to a degree, it is difficult to

accept that rb decreases eventually to zero, where grain

growth stops. In reality, ‘‘zero’’ value according to the

thermodynamic interpretation should be considered as the

ideally reachable limit for rb. This can be supported by

the fact that the {111}, {110} and {211} coherent twist

boundary in Cu possesses rb=0 if the inclination angles

/111, /110 and /211 are close to zero [13]. As for NC

alloy, however, experiment has shown that the effect of

solute atoms on rb can be substantial [14], and theoretical

considerations predict that driving rb to zero is possible

in alloy with high segregation energy [6,14].

According to Ref. [12], grain growth will re-start

because the GBs undergo a transition from ordered

to disordered structure with increasing T. This kind

of strong temperature-dependence can be also deduced

from the kinetic effect (Eq. (1)). Without considering this

correlation between GB structure and grain growth,

potentially more effective would be the thermodynamic

Fig. 2. Fits of Eq. (1) (dash-dotted lines) and Eq. (5) with s=2 (solid lines)

to the grain sizes measured in nickel thin film with oxygen contents

between 956 and 6039 wt. ppm at 673 K (data points were taken from [10];

note that not all the data were taken).

effect, which manifests only weak temperature-depen-

dence [15].

3. Application of the kinetic and the thermodynamic

models

3.1. Grain growth in nanocrystalline nickel doped with

nickeloxide

According to Ref. [10], a pulse reverse technique was

used to deposit nanocrystalline nickel (19 nm) doped with

nickel oxide. The effect of GB nickel oxide on the

thermal stability of NC Nickel was studied on samples

with oxygen contents between 956 and 6039 wt. ppm at

673 K. Fig. 2 shows different fits to the measured data

points using Eqs. (1) and (5), respectively. The corres-

ponding values for the fitting parameters are gathered in

Table 2.

From Fig. 2 and Table 2, both models give good fits to

the experimental data. According to the kinetic mechanism,

constant Q and linear V/rb relation prevail throughout a

single process with definite oxygen content, and increasing

Fig. 3. Fits of Eq. (1) (dashed lines) and Eq. (5) with s=2 (dash-doted lines)

and s=3 (solid lines) to the grain sizes measured in silver thin film at

different annealing temperatures [11].

Page 4: Comparison between kinetic and thermodynamic effects on grain growth

Table 3

Values of the rate constants and the relative errors obtained by fitting Eqs. (1) and (5) (with s=2 and s=3) to the grain sizes measured in silver thin film at

different annealing temperatures [11]

T Eq. (5), s=2 Eq. (5), s=3 Eq. (1) Error of fit (%)

(K)k1 (nm

2/s) k2 (nm2/s) k1 (nm

2/s) k2 (nm2/s) k3 (nm

2/s) Dmax (nm) k (nm2/s) Eq. (1) Eq. (5)

s=3 s=2

523 15.136 1.3132 14.060 7.0260 5.1427 94.9 13.180 2.4 2.9 4

473 5.4695 0.7227 6.4965 1.9641 0.4576 69.3 5.6039 3.1 2.8 3.5

448 1.3528 0.1944 2.6034 0.7351 0.1088 53.1 4.9807 0.4 0.33 2.8

F. Liu, R. Kirchheim / Thin Solid Films 466 (2004) 108–113 111

the oxygen content enhances Q (see Table 2). So the

stagnation of grain growth after longer annealing is due to

the segregated impurities and/or second phase precipitates

to stabilize GB migration. According to Eq. (5), however, a

single process consists of several domains—Q for the later

domain is much higher than that for the earlier domain and

a linear V/rb relation does not hold within the whole

process. Furthermore, k1 and k2 are reduced with increasing

the oxygen content, but k2 differs in smaller order of

magnitude than k1 does, inferring that Q in the second

domain does not change a lot with the oxygen content.

Therefore, it can be concluded that if grain growth has

(almost) stopped with saturated GBs, variations of Q and rb

Fig. 4. Arrhenius plot of the parameters (a) k (in Eq. (1)) and (b) k1, k2 and

k3 (in Eq. (5) with s=3) against the reciprocal annealing temperature.

should be negligible, independent of the overall solute

content [10].

3.2. Normal grain growth in silver thin film

Normal grain growth in 80-nm-thick sputter-deposited

Ag films was studied via in situ heating stage transmission

electron microscopy [11]. A grain growth exponent n=1/3

from the law D1/n�D01/n=k(T)t was calculated by minimiz-

ing the deviation in the fitting function to the experimental

data, and Q=0.53 eV was found, which is close to surface

diffusion [16]. According to Ref. [11], n=1/3 indicates that

some impurities and/or surface oxide might play a role in

inhibiting GB migration by segregation in normal grain

growth in silver thin film. Here, re-fitting Eqs. (1) and (5)

that assume n=0.5 to the experimental data [11] gives

good results, as shown in Fig. 3 and Table 3. Comparing

Table 3 with Table 1 demonstrates that the relation,

kik1s=2ik1

s=3>k2s=2ik2

s=3>k3s=3 is also deduced, and fit

by Eq. (5) with s=3 is better than that with s=2.

Whether both the kinetic and the thermodynamic models

can explain the deviation from the ideal parabolic manner

for normal grain growth needs further consideration of

the revolution of Q with time or temperature. When n

deviates from the ideal value of 0.5, Q is generally diffi-

cult to assess particularly. Since n=0.5 can be guaranteed

when modelling by using Eqs. (1) and (5), Q can be

subsequently obtained from the slope of a plot of ln k against

1/T. It is assumed that the rate constant k has an Arrhenius

relation with temperature [1]:

k ¼ k0 exp � Q

RT

� �ð6Þ

where R is the gas constant and k0 a constant. The results are

shown in Fig. 4 and Table 4.

Table 4

Grain growth activation energies obtained from Arrhenius plot of the

parameters (a) k (in Eq. (1)) and (b) k1, k2 and k3 (in Eq. (5) with s=3)

against the reciprocal annealing temperature

Temperature Eq. (1) Eq. (5), s=3

range (K)Q (eV) Q1 (eV) Q2 (eV) Q3 (eV)

448–523 0.31 0.4 0.6 1

Page 5: Comparison between kinetic and thermodynamic effects on grain growth

F. Liu, R. Kirchheim / Thin Solid Films 466 (2004) 108–113112

4. Discussion

The present experimental data strongly suggests that the

probably existing impurities and/or surface oxide segregated

to GBs are the cause of the kinetic or thermodynamic

mechanism that suppresses grain growth in silver thin film

[11]. The difficult question is how or what is the mechanism

by which the impurities and/or surface oxide acts.

In most cases, the activation energy of grain growth is

close to that of GB diffusion. Due to the limited temper-

ature range for grain growth in silver thin film, constant

value for Q could be obtained by fitting the data points

derived from k (Eqs. (1) and (6)) with a straight line (see

Fig. 4a). On this basis, Q=0.31 eV is obtained (Table 4),

which is significantly less than that of GB diffusion in

silver, 0.95 eV, but close to that measured for the

abnormal grain growth, 0.274 eV [17]. This value is

consistent with a surface diffusion process, reported to

be 0.3 eV in electro-migration pore formation experiments

[17]. So the enhanced thermal stability due to the pres-

ence of impurities can be probably explained on the basis

of GB pinning by an oxide phase that seems to be

present in the silver sample. These impurities accumulated

in GBs, as an external agent, might exert a retarding force

on GB migration, but do not change the activation energy

for grain growth, i.e. surface diffusion (see Fig. 4a and

Table 4).

As solute segregation proceeds upon grain growth, GB

concentration reaches the saturated value [18,19], thus

eliminating the GB vacancies, altering the GB dislocation

structure, and reducing the free volume by simply attach-

ing itself to the areas of poor fits [20]. This will increase

the activation energy for GB diffusion, DGb, as grain

growth relies upon diffusion while diffusion is concerned

with the number and type of neighbouring atoms. As

shown in Refs. [18,19], grain growth in nano-grained

FexY1�x (0<x<0.3) produced by gas condensation was

apparently influenced by solute segregation of Fe to the

Y GBs. When GBs are saturated with Fe atoms, the

grains and GBs reach a metastable state and grain growth

stops; grain size keeps stable and grain growth will not

occur until precipitation happens. This kind of phenome-

non was also observed in GB self-diffusion in a Cu

polycrystal of different purity showing that increasing

the GB impurity concentration leads to a decrease in

GB energy, while enhancing the activation enthalpy for

GB self-diffusion [21].

By fitting the data points derived from k1, k2 and k3 (Eq.

(5) with s=3 and Eq. (6)), Q, kept constant within each

domain, is declared to increase dramatically with annealing

time (Fig. 4b and Table 4). This shows that isothermal

annealing results in equilibrium grain size with saturated

GBs, while Q is increased with solute segregation and grain

growth. As given in Table 4, Q1, Q2 and Q3 for the first, the

second and the third domains are obtained as 0.4, 0.6 and 1

eV, respectively. Therefore, it should be due to the reduced

GB energy resulting from solute segregation that suppresses

grain growth.

The atomic mechanism involved in the motion of the

GBs consists of jumps across the interface from one

crystallite to another. This is expected to occur with Q

between 0.75 and 1.25 eV [11,17], more consistent with

GB diffusion in silver. For a thin film specimen, GB

motion and grain growth can be inhibited by the formation

of thermal groove, which pins the GBs at the film surface

[17]. The thermal groove forms by surface diffusion, and

surface diffusion should have activation energy, 0.3 eV

[17]. As given in Table 4, Q=0.4 eV is initially close to

that of surface diffusion, but upon grain growth, the

probably existing impurities and/or surface oxide seem to

block surface and/or GB diffusion path, thus increasing Q

(=1 eV) and reducing rb.

5. Conclusion

A detailed comparison between the kinetic and the

thermodynamic effects on grain growth was given. Both

mechanisms provide good interpretation for derivation

from the normal parabolic growth under ideal conditions.

It was found that, both models originate from segregation

of solute and/or impurity atoms at GBs, present different

explanations for grain growth inhibition, but give analo-

gous results. According to the thermodynamic effect, the

probably existing impurities and/or surface oxide seems to

block surface and/or GB diffusion path, thus enhances the

activation energy, and in turn, reduces GB energy. Accord-

ing to the kinetic effect, however, impurity atoms accu-

mulated in GBs, as an external agent, might exert a

retarding force on GB migration, but does not change

the activation energy for grain growth.

Acknowledgements

This research is supported by the Alexander von

Humboldt Foundation.

References

[1] C.H. Moelle, H.J. Fecht, Nanostruct. Mater. 6 (1995) 421.

[2] A. Michels, C.E. Krill, H. Ehrhardt, R. Birringer, D.T. Wu, Acta

Mater. 47 (1999) 2143.

[3] K.W. Liu, F. Mucklich, Acta Mater. 49 (2001) 395.

[4] J. Weissmuller, Nanostruct. Mater. 3 (1993) 261.

[5] Z.Q. Gao, B. Fultz, Nanostruct. Mater. 4 (1994) 839.

[6] R. Kirchheim, Acta Mater. 50 (2002) 413.

[7] J.W. Gibbs, Trans. Conn. Acad., vol. III, 1876, pp. 108–248; (1878)

343–524, Also in: The Collected Works of J.W. Gibbs, vol. 1, Long-

mans, Green and Co., New York 1928, p. 55.

[8] J.W. Cahn, in: W.C. Johnson, J.M. Blakely (Eds.), Surface Segrega-

tion in Metals and Alloys, ASM, Metals Park, 1979, p. 279.

Page 6: Comparison between kinetic and thermodynamic effects on grain growth

F. Liu, R. Kirchheim / Thin Solid Films 466 (2004) 108–113 113

[9] H.J. Hofler, R. Tao, L. Kim, R.S. Averback, C.J. Altstetter, Nano-

struct. Mater. 6 (1995) 901.

[10] H. Natter, M.S. Loffler, C.E. Krill, R. Hempelmann, Scr. Mater. 44

(2001) 2321.

[11] R. Dannenberg, E.A. Stach, J.R. Groza, B.J. Dresser, Thin Solid Films

379 (2001) 133.

[12] S.B. Lee, N.M. Hwang, D.Y. Yoon, M.F. Henry, Metall. Mater.

Trans., A Phys. Metall. Mater. Sci. 31A (2000) 985.

[13] F. Ernst, M.W. Finnis, A. Koch, C. Schmidt, B. Straumal, W. Gust, Z.

Met.Kd. 87 (1996) 911.

[14] C.E. Krill, H. Ehrhardt, R. Birringer, in: E. Ma, B. Fultz, R. Shull, J.

Morral, P. Nash (Eds.), Chemistry and Physics of Nanostructures and

Related Non-Equilibrium Materials, The Minerals, Metals & Mate-

rials Society, Warrendale, PA, 1997, p. 115.

[15] M. Hillert, Acta Metall. 13 (1965) 227.

[16] H.V. Atkinson, Acta. Metall. 36 (1988) 469.

[17] R. Dannenberg, E.A. Stach, J.R. Groza, B.J. Dresser, Thin Solid Films

370 (2000) 54.

[18] J. Weissmuller, W. Krauss, T. Haubold, R. Birringer, H. Gleiter,

Nanostruct. Mater. 1 (1992) 439.

[19] F. Liu, F. Liu, R. Kirchheim, J. Cryst. Growth 264 (2004) 385.

[20] H. Gleiter, B. Chalmers, Prog. Mater. Sci. 16 (1972) 77.

[21] T. Surholt, C.H.R. Herzig, Acta Mater. 45 (1997) 3817.