You are on page 1of 5

Applied Surface Science 318 (2014) 100–104

Contents lists available at ScienceDirect

Applied Surface Science


journal homepage: www.elsevier.com/locate/apsusc

Effect of Cu concentration on the formation of Cu1−x Znx shape


memory alloy thin films
İsmail Hakkı Karahan a , Rasim Özdemir a,b,∗
a
Department of Physics, Mustafa Kemal University, Hatay 31000, Turkey
b
Kilis Vocational High School, Kilis 7 Aralık University, 79000 Kilis, Turkey

a r t i c l e i n f o a b s t r a c t

Article history: The Cux Zn1−x (x = 0.06, 0.08, 0.1) deposits were fabricated by a electrodeposition method. The structural
Received 12 November 2013 and electrical properties of the films were investigated by cyclic voltammetry (CV), X-ray diffraction
Received in revised form 17 January 2014 (XRD), Scanning electron micrograph (SEM), and DC resistivity measurements. Phase identification of
Accepted 20 January 2014
the samples was studied by the XRD patterns. XRD patterns shows the characteristics XRD peaks cor-
Available online 27 January 2014
responding to the, ˇ, and  phases. The grain sizes of the samples were decreased whereas microstrain
increased with the increase in Cu2+ substitution. The SEM study reveals the fine particle nature of the
Keywords:
samples with increasing Cu content. DC resistivity indicates the metallic nature of the prepared sam-
Electrodeposition
Cu–Zn alloy ples. It has been found that the Cu ions have a critical influence on the resultant structure and resistivity
Electrical property properties of the Cu–Zn samples.
Shape memory alloys © 2014 Elsevier B.V. All rights reserved.

1. Introduction of rubber to steel, good mechanical resistance and conformability,


hardness and decorative value [9–15].
Electrodeposited metallic layers is generally practiced to modify Commercial electrodeposition of Cu–Zn alloys in cyanide baths
the substrate surface in order to produce a wide range of use- produces high-quality deposit but environmental problems arise in
ful materials with improved mechanical, chemical and physical the use and disposal of cyanide. Thus an extensive search has been
properties. This technique has demonstrated to be very conve- made for satisfactory alternative electrolytes [12].
nient because of its simplicity and low cost in comparison with the Cu–Zn electrodeposits can be produced from electrolytes other
other method such as sputtering and vapor deposition. Thus, eco- than cyanide baths, namely, from electrolytic solutions based on
nomical materials can be used as substrates, making this process sorbitol [12], mannitol [14], sorbitol based alkaline [16], nitrilo-
economically attractive. Compared to pure metal coatings, elec- triacetic acid (NTA) [17], pyrophosphate [18–21], citrate [13,22],
trochemically deposited alloys show better properties, since their citrate based cysteine and benzotriazole [23], tartrate [24], glycerol
chemical composition can be varied to the required function. Since [25–27], and sulphate [28].
the physicochemical properties of alloys are seriously affected by The plating of Cu–Zn alloys from citrate bath has not been exten-
their compositions as well as structures [1–8], a reliable control of sively studied. However, for baths containing more than 0.02 Molar
the composition and structure for these engineering alloys is an copper ions have not been clarified.
important issue for their wide applications. In this study, we decided to study these alloy deposits pre-
Properties of an electrodeposited film can be affected by many pared from a citrate bath with various Cu/Zn ratios to explore how
parameters, such as the bath type, pH, deposition potential, and copper content would influence the deposition process and the
metal ion contents. In particular, the metal ion concentration has morphology, composition and electrical resistivity of the Cu–Zn
been observed to significantly affect the microstructure and mag- electrodeposits. The aim of this study was reduce the toxicity and
neto transport properties of electrodeposited films. the cost of the electroplating process while maintaining the quality
Cu–Zn alloys have been widely used in industry, owing to their of the coatings.
attractive features such as protection against corrosion, adhesion
2. Experimental details

∗ Corresponding author at: Mustafa Kemal University, Faculty of Art and Science,
Electrodeposition was realized in a conventional glassy cell in
Department of Physics, 31000-Hatay, Turkey. Tel.: +90 326 245 58 66;
non stirred and non aerated conditions. The bath composition
fax: +90 326 245 58 67. and deposition parameters for the Cu–Zn deposits are reported in
E-mail address: ihkarahan@gmail.com (R. Özdemir). Table 1. All the chemicals used were of analytical grade. Double

0169-4332/$ – see front matter © 2014 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.apsusc.2014.01.119
İ.H. Karahan, R. Özdemir / Applied Surface Science 318 (2014) 100–104 101

Table 1
Electrolyte composition and deposition conditions for Zn1−x Cux films electrodeposited on aluminum substrate at pH 5.8 and 20 ◦ C for the duration of 1 h.

Sample CuSO4 .5H2 O (mol L−1 ) ZnSO4 .7H2 O (mol L−1 ) Na3 C6 H5 O7 (mol L−1 ) Current (mA)

Zn74 Cu26 0.06 0.2 0.5 60


Zn63 Cu37 0.08 0.2 0.5 60
Zn48 Cu52 0.10 0.2 0.5 60

distilled water was used for preparation of the electrolyte solu-


tion. Pure aluminum foil with a surface area 3.8 cm2 was used
as substrate and before each experiment; the samples were pol-
ished, degreased, and activated by dipping in a 1 M NaOH with
surfactant at 70 ◦ C during 5 min and finally rinsed with the twice
distilled water (18 M cm) and dried with fresh air. The CV exper-
iments were conducted in a freshly prepared electrolyte given as
Table 1, without purging. A scan rate of 10 mV s−1 was selected
based on preliminary experiments. The scan began from 0.5 V in
the positive direction, up to −1.6 V, and finally reversed to 0.5 V.
The potentials were measured versus a saturated calomel refer-
ence electrode (SCE). Characterization of the coatings was done
by a variety of analytical facilities. The surface morphology of
the deposits was observed by means of a JEOL scanning electron
microscope (SEM). The attached energy dispersive spectroscopy Fig. 1. Influence of the copper concentration on the cyclic voltammograms for the
(EDS) detector was used to determine atomic composition of the electrodeposition from zinc solution (0.2 M ZnSO4 .7H2 O) in 0.5 M Na3 C6 H5 O7 , pH
alloy. Each sample was analyzed at three locations, to confirm 5.8, v = 10 mV s−1 .
uniformity.
A Rigaku diffractometer was used to analyze the crystallo- in the temperature range of 100–350 K were carried out by using
graphic structure of alloys. The X-ray diffractometer was operated a Keithley 2400 source measure unit. The samples were placed
at 30 kV and 30 mA with CuK˛ radiation. The film composition was onto cold finger of a Janis liquid nitrogen cryostat and the tempera-
determined via an Energy Dispersive X-ray Spectrometer (EDX). ture was accurately monitored with a Lake–Shore 320 temperature
The grain size (D) was calculated using Scherer’s formula [29]. controller.
0.9
D= (1) 3. Results and discussion
ˇ × cos 

The microstrain was obtained by using the relations [30] 3.1. Electrodeposition and properties of Zn, Cu, and Zn1−x Cux thin
films
ˇ × cos q
ε= (2)
4
To define the effect of electrolyte concentration on the Cu–Zn
where ˇ is the full-width at half maximum,  is the wavelength of alloy deposition process cyclic voltammetry technique was used.
X-ray used (1.5400 Å), and  is the diffraction angle. Fig. 1 shows voltammograms for the aluminum electrode in the
The resistivity measurements were carried out using the four Cu–Zn baths. From this figure, the influence of copper content
point probe method. For this purpose; the electrodes were attached on the potentiodynamic cathodic and anodic polarization parts
to the contact regions with silver paint. The ohmic behavior of of cyclic voltammograms curves can be seen. The potential was
the contacts in the studied temperature region was confirmed by scanned from 0.5 V to the cathodic direction up to −1.6 V and
the linear variations of the I–V characteristics which is indepen- ended at 0.5 V. It is apparent from Fig. 1 that the anodic part
dent from the reversal of the applied currents. The thermal voltage in the cyclic voltammograms consisted of one anodic current
effect was eliminated by taking the average of voltage readings peak for all the films. This shows that the alloy dissolution takes
with two reverse currents at each temperature. Each sample was place under one potential values. The anodic process, in general,
measured several times to make sure that the obtained data was shows an anodic dissolution peak A (A1 for 0.10 M, A2 for 0.08 M
reliable. The temperature-dependent conductivity measurements and A3 for 0.06 M CuSO4 solutions) at 0.1 V vs SCE. In solutions,

Fig. 2. XRD analysis of Cu–Zn thin films.


102 İ.H. Karahan, R. Özdemir / Applied Surface Science 318 (2014) 100–104

Fig. 3. (a) SEM images of Zn74 Cu26 alloy (0.06 M CuSO4 .7H2 O). (b) SEM images of Zn63 Cu37 alloy (0.08 M CuSO4 .7H2 O). (c) SEM images of Zn48 Cu52 alloy (0.10 M CuSO4 .7H2 O).

during the forward scan towards the negative direction, the peak was observed at 0.15 V. When the copper concentration was
cathodic current increased sharply when the deposition begins. increased to 0.08 M, the electroreduction peak shifted to −1.2 V
In addition, the increase in the current density at potentials more and the dissolution peak of the deposit showed an increase. When
negative than ∼−1.20 V can be attributed to a significant hydrogen the copper concentration was increased to 0.1 M, the reduction
evolution reaction (HER) in parallel to Cu–Zn electrodeposition. It peak was practically unaffected, but the dissolution peak increased
was observed that in presence of more copper ion the cathodic again.
current was increased. Cu–Zn electrodeposition in the presence It may be noted, in Fig. 1, that the deposition rate is affected in
of 0.06 M copper presents a reduction peak at −1.05 V versus cathodic region by the concentration of copper, since the cathodic
SCE. Upon reversal of the potential-sweep direction, a dissolution current density in this region increased when the concentration
İ.H. Karahan, R. Özdemir / Applied Surface Science 318 (2014) 100–104 103

of this complex in the bath increased (20 mA cm−2 , 0.06 M CuSO4 105
in the bath; 30 mA cm−2 , 0.08 M CuSO4 in the bath; 34 mA cm−2 , 100
0.08 M CuSO4 in the bath). In this region a primary bulk nucleation
95
and growth of copper occurred.

Grainsize (nm)
The cathodic current starts to increase and forms the cathodic 90
peak at a potential that depends on bath condition. 85
The reaction responsible for the cathodic peak is reaction (1): 80
2+ − 0 75
(1)Cu + 2e = Cu
70
Upon the sweep reversal, the cathodic current density gradually 65
decreases, until it crosses 0 and turns into anodic current. Further 60
sweep in the positive direction results in the anodic peak, which 20 30 40 50 60
corresponds to the reaction (1) in the reverse direction. The current Copper in Deposited CuZn (%)
density past the peak is 0, indicating the completion of oxidative
dissolution of metallic copper at the electrode surface. In the anodic Fig. 4. Grain size (D) with Cu content.
direction, however, the oxidation of copper starts from the surface
that already has copper deposited, resulting in a potential close to
the Cu/Cu2 equilibrium potential. Due to the difference in depo- Fig. 4 shows that the crystallite diameter decreased from
sition and dissolution potentials, a crossover occurs between the 100 to 66 nm with increase in Cu2+ substitutions, which is in
cathodic and anodic current traces [31] at the crossover poten- agreement with the results observed in the literature [35,36]. Fur-
tial Ex. The presence of the crossover is diagnostic for the nuclei ther, the substitution of Cu2+ in place of Zn2+ ion increased the
formation on the electrode [32]. crystalline size, due to the large size mismatch between Cu2+
The preferential crystal orientation of zinc electrodeposits and Zn2+ ions. Hence, in order to relax the volume strain the
depends on the experimental conditions [33,34], such as current crystallite size decreases with the increase in Cu2+ concentra-
density, bath composition, pH and temperature. Fig. 2 shows the tion. Other possible explanation to decrease of crystallite size is
X-ray diffraction patterns obtained for several Cu–Zn coatings elec- that the ionic radius of Cu is smaller than the ionic radius of
trodeposited at gradually increasing copper concentration. From Zn.
the XRD measurements in the electrodeposited samples ˛, ˇ, and The temperature dependent electrical resistivity measurements
 phases were detected. When the deposit was formed in the pres- for all samples were carried out in the temperature range of
ence of 0.06 M Cu (Zn74 Cu26 ), a considerable decrease of the (1 1 0) 100–350 K and the variations of normalized resistivity of five sam-
peak was observed, as well as of the peaks associated with planes ples were plotted as a function of temperature in Fig. 5. There is a
(2 1 1), due to the change in orientation of the Cu–Zn crystallo- gradual linear drop in resistivity as the temperature is lowered from
graphic planes by the incorporation of copper. When the copper 320 K. The resistivity increased with the increase in Cu substitu-
content in solution was increased to 0.08 M (Zn63 Cu37 ) and 0.1 M, tion. The resistivity increased with the Cu substitution as it should
the peaks associated with plane (1 1 0) and (2 1 1) diminished fur- since increasing disorder, brought about by rising Zn concentra-
ther and (1 1 1) peak was appeared. When the copper content in tion, contributes to the increasing electron scattering and hence to
solution was increased to 0.1 M (Zn48 Cu52 ), a peak (3 1 1) then increasing resistivity.
appeared at a value of 2 = 88◦ . Electrical resistivity increases with increasing Cu content and
When Zn electrodeposited with other metals the morphol- more increase in these elements in the sample decreases the resis-
ogy of the deposits modifies because of their influence on the tivity [37,38]. Fig. 6 illustrates increase in electrical resistivity
growth of the initial nuclei [33]. Fig. 3 shows SEM micrographs of values for the alloys with increasing Cu concentrations in film. Fig. 6
deposits formed from the Zn74 Cu26 (Fig. 3(a) 0.06 M CuSO4 .7H2 O); shows the strong dependence of the resistivity on the composition
Zn63 Cu37 (Fig. 3(b) 0.08 M CuSO4 .7H2 O) and Zn48 Cu52 (Fig. 3(c) of thin films. This strong dependence on the composition can be
0.10 M CuSO4 .7H2 O) the acidic citrate–sulfate Zn–Cu solutions at understood qualitively by considering the densities of states. The
various copper concentrations. It is observed that the deposits density of states splits up into contributions from the components
are generally composed of fine grains. The metallic luster and the (Zn, Cu) of the alloy. The two contributions are well separated in
brightness of the deposit increased with increasing of the copper. energy.
In the presence of 0.06 M copper in the bath, the deposit being
compact and formed by small cubical crystallites. With a copper Zn63Cu37 Zn48Cu52 Zn74Cu26
concentration of 0.08 M, a deposit with even smaller grain size was
observed with slim rods (Fig. 3b), and at the 0.1 M copper grain size
1.00
was began to bigger non-coalesced globular crystallites and the film
Normalized Resisvity

was compact. Naturally, morphologies of Cu–Zn deposits were con-


0.90
nected with the discharge mechanism of the complex agent adding
to the bath. 0.80
The average crystallite size of the samples was calculated
from the peak width at the half maximum of the peak (ˇ), 0.70
using Eq. (1).The microstrain (ε) for preferential orientation were
calculated using the Eq. (2). Table 2 shows the calculated aver- 0.60
age crystallite size of the samples obtained at different bath
compositions. The grain sizes were found to be within the 0.50
range of approximately 66–100 nm as seen from Fig. 4. The 50 100 150 200 250 300 350
average crystallite size of the Cu–Zn alloys decreased and micros- Temperature (K)
train increased when the increase in Cu concentration in the
bath. Fig. 5. Normalized electrical resistivity against temperature for the Cu–Zn alloys.
104 İ.H. Karahan, R. Özdemir / Applied Surface Science 318 (2014) 100–104

Table 2
Grain size, micro strain values for Zn74 Cu26 , Zn63 Cu37 and Zn48 Cu52 alloys.

Samples ␪ (◦ ) d(Å) FWHM Grain size (nm) Micro strain

Zn74 Cu26 43.201 2.0924 0.382 100.1 0.1183


Zn63 Cu37 43.240 2.0906 0.579 66.1 0.1862
Zn48 Cu52 42.306 2.1346 0.541 69.7 0.1622

0.25 [7] E. Beltowska-Lenman, A. Riesenkampf, An investigation of the electrodepo-


sition kinetics of permalloy thin films using a rotating disc electrode, Surf.
Technol. 11 (5) (1980) 349–355.
Electrical resisvity (μΩcm)

0.22 [8] M. Dogan, E. tirasoglu, İ.H. Karahan, N. Kup Aylikci, V. Aylikci, A. Khoul, H.A.
Cetinkara, O. Serifoglu, Alloying effect on K X-ray intensity ratio and production
100 K cross section values of Zn and Zn–Cr alloys, Radiat. Phys. Chem. 87 (2013) 6–15.
[9] F.A. Lowenheim, Modern Electroplating, Wiley, New York, 1974.
0.19 149 K
[10] A. Brenner, Electrodeposition of Alloys. Principles and Practice, vol. 1, Academic
198 K Press, New York, 1963.
248 K [11] H. Strow, Metal Finishing Guidebook 99 (1A) (2001) 206.
0.16 [12] I.A. Carlos, M. R.H. de Almeida, Study of the influence of the polyalcohol sor-
297 K
bitol on the electrodeposition of copper–zinc films from a non-cyanide bath,
346 K J. Electroanal. Chem. 562 (2004) 153–159.
0.13 [13] F.B.A. Ferreira, F.L.G. Silva, A.S. Luna, et al., Response surface modeling and
optimization to study the influence of deposition parameters on the electrode-
position of Cu–Zn alloys in citrate medium, J. Appl. Electrochem. 37 (2007)
0.1 473–481.
[14] R. Juskenas, V. Karpaviciene, V. Pakstas, A. Selskis, V. Kapocius, Electrochemical
20 30 40 50 60
and XRD studies of Cu–Zn coatings electrodeposited in solution with<** D-
Copper in thin film (%) mannitol, J. Electroanal. Chem. 602 (2007) 237–244.
[15] K.M. Ismail, R.M. Elsherif, W.A. Badawy, Effect of Zn and Pb contents on the elec-
Fig. 6. Effect of Cu ratio in the film and electrolyte in different temperature on the trochemical behaviour of brass alloys in chloride-free neutral sulfate solutions,
electrical resistivity. Electrochim. Acta 49 (2004) 5151–5160.
[16] M.R.H. de Almeida, E.P. Barbano, M.F. de Carvalho, I.A. Carlos, J.L.P. Siqueira,
L.L. Barbosa, Electrodeposition of copper–zinc from an alkaline bath based on
4. Conclusions EDTA, Surf. Coat. Tech. 206 (2011) 95–102.
[17] R.M. Krishnan, V.S. Muralidharan, S.R. Natarajan, A non-cyanide brass plating
Zn74 Cu26 , Zn63 Cu37 and Zn48 Cu52 alloy films were synthesized bath, Bull. Electrochem. 12 (1996) 274–277.
[18] K. Johannsen, Plat. Effect of temperature & bulk stirring on electroplating of
via electrodeposition method. X-ray diffractograms assured the brass from pyrophosphate electrolyte, Surf. Finish. 88 (2001) 104–108.
cubic spinel structure for all the investigated samples with the [19] Y. Fujiwara, H. Enomoto, Composition, structure and morphology of Cu–Zn
appearance of small peaks represented a secondary phase due to alloy deposits from pyrophosphate baths, Plat. Surf. Finish. 80 (1993) 52–56.
[20] D. Page, S. Roy, Electrodeposition of thin film Cu–Zn shape memory alloys,
the presence of rare earth La3+ ions. XRD patterns confirmed that J. Phys. IV (1997) 269–274.
the electrodeposited samples have ˛, ˇ, and  phases. The average [21] L.F. Senna, S.L. Díaz, L. Sathler, Electrodeposition of copper–zinc alloys in
crystallite size of the samples varied from 67 to 100 nm. Micros- pyrophosphate-based electrolytes, J. Appl. Electrochem. 33 (2003) 1155–1161.
[22] F.H. Assaf, S.S.A. Elrehim, A.S. Mohamed, A.M. Zaky, Electroplating of brass from
train have been calculated Scherrer’s formula The strain of the films citrate-based alloy baths, Indian J. Chem. Technol. 2 (1995) 147–152.
increased firstly linearly with the increase of Cu substitution. But, [23] F.L.G. Silva, D.C.B. do Lago, L.F. Senna, Electrodeposition of Cu–Zn alloy coatings
the crystallite size was reduced with the increase of Cu2+ ions. The from citrate baths containing benzotriazole and cysteine as additives, J. Appl.
Electrochem. 40 (2010) 2013–2022.
structural characterization of the powders using XRD confirmed the
[24] D. Defilippo, A. Rossi, D. Atzei, A tartrate-based alloy bath for brass-plated steel
formation of nanosized films. Bright, compact and good adherent wire production, J. Appl. Electrochem. 22 (1992) 64–72.
alloy coatings were obtained with galvanostatic conditions. Elec- [25] T. Vagramyan, J.S.L. Leach, J.R. Moon, A method for pH measurement in
trical resistivity measurements showed the strong dependence of the immediate vicinity of the electrode surface, Electrochim. Acta 24 (1979)
231–236.
the resistivity on the Cu content in the electrolyte and in the Cu–Zn [26] I.A. Carlos, Ph.D. Thesis, Universidade Federal de São Carlos, São Paulo, 1990.
films. This strong dependence on the composition can be under- [27] A. Brenner, Electrodeposition of Alloys, vol. 1, Academic Press, New York, 1963,
stood qualitively by considering the densities of states. pp. 411.
[28] S.M. Rashwan, Trans. Inst. Metal Finish. 85 (2007) 217.
[29] B.D. Cullity, S.R. Stock, Elements of X-ray diffraction, vol.170, 3rd ed., Prentice
Acknowledgement Hall, Upper Saddle River NJ, 2001.
[30] K.L. Chopra, Thin Film Phenomena, Robert E. Krieger publishing Company,
Huntingten,Newyork, 1979.
Financial support of this research by the Mustafa Kemal Uni- [31] Southampton Electrochemistry Group, in: T.J. Kemp (Ed.), Instrumental Meth-
versity Scientific Research Projects is gratefully acknowledged ods in Electrochemistry, Ellis Horwood Ltd., Chichester, UK, 1985.
(MKU-BAP- 1005M 0118 and 1204 D 0110). [32] D. Grujicic, B. Pesic, Electrodeposition of copper: the nucleation mechanisms,
Electrochim. Acta 47 (2002) 2901–2912.
[33] J.B. Bajat, V.B. Miskovic-Stankovoc, M.D. Maksimovic, D.M. Drazic, S. Zec, Elec-
References trochemical deposition and characterization of Zn–Co alloys and corrosion
protection by electrodeposited epoxy coating on Zn–Co alloy, Electrochim. Acta
[1] W. Schwarzacher, D.S. Lashmore, Giant magnetoresistance in electrodeposited 47 (2002) 4101–4112.
films, IEEE Trans. Hagn. MAG-32 (1996) 3133–3153. [34] İ.H. Karahan, A study on electrodeposited ZnFe alloys, J. Mater. Sci. 42 (24)
[2] S.S. Djokic, M.D. Maksimovic, in: J.O’M. Bockris (Ed.), Modern Aspects of Elec- (2007) 10160–10163.
trochemistry, No. 22, Plenum Press, New York, 1992, p. 417. [35] L. Zhao, H. Yang, L. Yu, Y. Cui, X. Zhao, Y. Yan, S. Feng, The studies of nanocrys-
[3] T. Osaka, Acta 44, recent development of magnetic recording head core mate- talline Ni0.7 Mn0.3 Ndx Fe2–x O4 (x = 0–0.1x = 0–0.1) ferrites, Phys. Lett. A. 332
rials by plating method, Electrochem. 21–22 (1999) 3885–3890. (2004) 268–274.
[4] T. Osaka, M. Takai, K. Ohashi, M. Saito, K. Yamada, A soft magnetic CoNiFe [36] M.A. Ahmed, N. Okasha, M.M. El-Sayed, Enhancement of the physical properties
film with high saturation magnetic flux density and low coercivity, Nature 392 of rare-earth-substituted Mn–Zn ferrites prepared by flash method, Ceram. Int.
(1998) 796–798. 33 (2007) 49–58.
[5] E.M. Kakuno, D.H. Mosca, I. Mazzaro, N. Mattoso, W.H. Schreiner, M.A.B. [37] D.C. Leitao, C.T. Sousa, J. Ventura, J.S. Amaral, F. Carpinteiro, K.R. Pirota, M.
Gomes, M.P. Cantao, Structure, composition, and morphology of electrode- Vazquez, J.B. Sousa, J.P. Araujo, Characterization of electrodeposited Ni and
posited Cox Fe1-x alloys, J. Electrochem. Soc. 144 (1997) 3222–3226. Ni80 Fe20 nanowires, J. Non-Cryst. Solids 354 (2008) 5241–5243.
[6] K.-M. Yin, Otentiostatic deposition model of iron–nickel alloys on the rotating [38] A. Thakur, P.K. Ahluwalia, Electrical resistivity of NaSn compound forming liq-
disk electrode in the presence of organic additive, J. Electrochem. Soc. 144 (5) uid alloy using ab initio pseudopotentials, Physica B-Condens. Matter 373 (1)
(1997) 1560–1566. (2006) 163–168.

You might also like