You are on page 1of 6

Materials Science and Engineering A 527 (2010) 28752880

Contents lists available at ScienceDirect

Materials Science and Engineering A


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

A simple technique of nanocrystallizing metallic surfaces for enhanced


resistances to mechanical and electrochemical attacks
X.Y. Mao a,b , D.Y. Li b, , F. Fang a , R.S. Tan a , J.Q. Jiang a
a
b

School of Materials Science and Engineering, Southeast University, Nanjing 211189, Jiangsu, China
Department of Chemical and Materials Engineering, University of Alberta, Edmonton, Alberta T6G 2V4, Canada

a r t i c l e

i n f o

Article history:
Received 30 October 2009
Received in revised form
31 December 2009
Accepted 6 January 2010

Keywords:
Surface nanocrystallization
Electron work function
Mechanical behavior
Corrosion resistance

a b s t r a c t
Surface nanocrystallization is an effective approach that can be used to enhance many metallic materials for increased resistances to electrochemical and mechanical attacks. Efforts have been continuously
made to improve existing surface nanocrystallization processes for feasible industrial applications. In this
article, we demonstrate a simple process for surface nanocrystallization achieved by repeated punching
with subsuent recovery treatment. As a sample study, we applied this process to a Cu30Ni alloy and
showed that the nanocrystallized CuNi surfaces exhibited increased hardness, corrosion resistance and
electron stability; the properties were further improved by increasing the punching time, which led to
smaller nano-sized grains and thicker nanocrystalline layers. This study demonstrates that the repeated
punching plus recovery treatment is a simple and effective process for surface nanocrystallization.
2010 Elsevier B.V. All rights reserved.

1. Introduction
Different from their coarse-grained polycrystalline counterparts, nanostructured materials possess improved or modied
mechanical, physical and chemical properties, such as strength,
resistances to wear and corrosion, and workability, etc. [14].
Recent studies show that nanocrystallization also helps to reduce
the afnity of stainless steel for bacterial biolms [5], attributed
to the formation of an enhanced passive lm that blocks the surface electronic interaction with the surrounding medium more
effectively. This makes nanocrystallization also promising for suppressing bio-corrosion, i.e. the deterioration of materials as a result
of the metabolic activity of microorganisms, which may produce
enzymes, organic and inorganic acids, and volatile compounds such
as hydrogen sulde, and thus inuences surface electrochemical
processes [6]. Nanocrystallization provides many new opportunities for materials engineers to mitigate problems with wear and
corrosion including bio-corrosion. Various processing techniques
have been developed to fabricate nanocrystalline materials, such
as the consolidation of ultrane power [2], ball-milling [7], severe
plastic deformation (SPD) [8], crystallization of amorphous solids
[4] and electrodepositing [9]. Among these processes, SPD is a simple method to generate nanocrystalline metallic materials but it
is less easy for treating hard bulk metals or alloys [10,11]. However, failure of a work piece often initiates at surface caused by,

Corresponding author.
E-mail address: dongyang.li@ualberta.ca (D.Y. Li).
0921-5093/$ see front matter 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2010.01.024

e.g., fatigue, fretting, wear or corrosion. Many mechanical components, e.g., gears, are primarily required to have superior surface
properties, since the surface damage changes the dimensions of
the components thus making them no longer functional. In this
case, surface nanocrystallization could be sufcient, which is much
easier to be achieved by applying SPD to a target surface. Surface
nanocrystallization produced by SPD has attracted many interests [1214]. Processes for surface nanocrystallization through SPD
include surface mechanical attrition [13,14], shot peening [1518],
sliding wear [19], surface rolling [20], wire-brushing [21], and
sandblasting [2224]. Researchers have been continuously making
efforts to improve these techniques in order to produce thicker and
more effective nanocrystalline surface layers for industrial applications.
In this article, we demonstrate a simple surface nanocrystallization process, which combines repeated punching and subsequent
recovery treatment. This process has the capability for generating
relatively thicker nanocrystalline layers with smaller nano-grains
by repeated punching due to the facts that there is no material
removal and the punching force and duration can be varied to
achieve larger depths of the affected layers with smaller dislocation
cells.
The main objective of this work is to demonstrate the effectiveness of this approach with emphasis on effects of the punching
duration on surface grain size, thickness of the nanocrystalline
layer, and corresponding surface mechanical and electrochemical
properties. We use Cu30Ni alloy as a sample material for the study.
Choosing Cu30Ni alloy as the sample material is attributed to its
resistance to corrosion for many applications in the marine envi-

2876

X.Y. Mao et al. / Materials Science and Engineering A 527 (2010) 28752880

ronment [25,26]. Besides, CuNi is a FCC single-phase alloy, which


helps to avoid complication in the study.
2. Experimental details
2.1. Sample preparation
Samples were cut from a commercial Cu30Ni alloy tube, consisting of (in wt.%): Ni 30.57, Mn 0.87, Fe 0.98, C 0.04, Zn 0.10, P
0.005 and Cu balance. The nal dimensions of samples were 30 mm
in diameter and 3 mm in thickness. The samples were ground with
silicon carbide paper up to 600-grit, rinsed with acetone, ethanol
and deionized water successively, and dried in an air ow. The samples were then punched repeatedly for a certain period of time
using a roto-hammer (Robert Bosch Tool Corporation, USA) at a frequency of 50 Hz. The head of the punching hammer, made of tool
steel, has a semi-spherical shape of 5 mm in diameter. The impact
energy of this punching machine was 1.6 ft-lbs (2.207 Nm) and the
punching head scanning rate was about 5 mm/s. Fig. 1 schematically illustrates a punching process, during which the punching
head scans the target surface to punch it at different locations.
The repeated punching results in accumulated plastic deformation.
The process was well controlled in order to have the entire sample
surface punched homogeneously. The punched surfaces had their
roughness in the range of 10 m. Samples punched for 30 min were
annealed (recovery treatment) at 275, 350 and 425 C, respectively,
in an argon atmosphere for 1 h in order to identify the optimal
recovery temperature. For determining the effect of punching duration on the thickness of nanocrystallized layer, a set of samples
were punched for 60, 90, 120 min, respectively, and then annealed
at 350 C in argon for 1 h. The annealing treatment was necessary
to turn dislocation cells with diffuse boundaries into nano-sized
grains. These treated samples were polished using SiC paper of
1200-grit and then etched using a 4 wt.%HNO3 solution. For electrochemical testing, an electrical wire was connected to the back face
of a sample prior to its encapsulation in insulating cement. The face
exposed to a 3.5 wt.%NaCl solution was a square of 5 mm 5 mm.

CA, USA). Hardness and  = We/Wt that is the ratio of the elastic deformation energy (We) to the total deformation energy
(Wt) were measured using a triboscope (nano-indenter made by
Hysitron, Minneapolis, USA) that was attached to the AFM. During the nanoindentation tests, indentation loaddepth curves were
recorded under maximum forces of 50, 100, 200 N, respectively.
Microhardness of the samples was also measured using a microindenter probe (Fisher Technology Ltd., Winsor, CT, USA) with a
diamond tip under a larger load of 50 mN. The hardness of each
tested location was measured at least three times.
Dynamic polarization measurements were carried out using
a commercial electrochemical system (model Pc4-750, Gamry
Instruments Inc., Warminster, PA, USA) by sweeping the potential
at a scan rate of 3 mv/s from 0.4 to 0.1 V vs. the open corrosion
potential (OCP) in a 3.5 wt.%NaCl solution. Prior to the electrochemical tests, mounted samples with electrical connecting wire were
stabilized in the solution for 50 min. The exposed square area of
5 mm 5 mm acted as the working electrode. A saturated calomel
electrode (SCE) was used as the reference electrode and a platinum
plate of 20 mm 20 mm was used as the counter electrode. During
the polarization tests, corrosion current and potential curves were
recorded.
Electron work functions (EWF) of treated and untreated sample surfaces in the ambient environment were determined using a
scanning Kelvin probe (SKP, KP Technology Ltd., Wick, UK). Surfaces
of the samples were lightly polished and etched using a 4% NHO3
solution to remove the deformed top layer (caused by polishing).
A gold tip of 1 mm in diameter was used to scan a sample surface
within an area of 2 mm 2 mm covering 10 10 = 100 measurement points. The EWF value was obtained by averaging EWF values
collected from 100 points over the scanned area. The oscillation
frequency of the SKP tip was 173.5 Hz. The principle and experimental procedure of the SKP system for measuring EWF have been
described in detail in the literature [27,28].
3. Results and discussion

2.2. Characterization

3.1. Formation of a nanocrystalline surface by punching and


recovery treatment

The morphology and grain size of sample surfaces were analyzed using an atomic fore microscope (AFM, Digital Instruments,

Severe plastic deformation of metals or alloys with multiple


slip systems can generate nano-sized dislocation cells, forming a

Fig. 1. Schematic illustration of a punching process: the hammer was moved up and down at a frequency of 50 Hz to impact a target surface and introduced severe plastic
deformation in the surface layer.

X.Y. Mao et al. / Materials Science and Engineering A 527 (2010) 28752880

so-called cellular structure. The boundaries or walls between


adjacent dislocation cells are diffuse in nature [30] and the cell
interior may still have dislocations. As computational studies
demonstrate [31,40], the cell interior dislocation density could
reach 1015 m2 , which is about 1/5 of that in the cell walls
in nanocrystallined metals achieved by SPD. Thus, the punched
samples were annealed (recovery treatment) in order to turn the
nano-sized dislocation cells in the surface layer into nano-sized
grains with sharp grain boundaries and fewer interior dislocations.
To identify a suitable annealing temperature, we annealed samples
that were punched for 30 min respectively at 275, 350 and 425 C
for 1 h. It appeared that 350 C was the most suitable recovery
temperature (such treated sample showed the best performance
during corrosion test). A group of samples were then punched for
different periods, respectively, and annealed at 350 C for investigating how the duration of punching affected the thickness of
nanocrystallized surface layer and corresponding properties. In the
present study, the size of grains in the nanocrystallized surface layer
was determined using an atomic force microscope. Previous studies [5,22,24] on nanocrystalline materials using AFM, transmission
electron microscopy and X-ray diffraction techniques gave consistent results; this indicates that AFM is acceptable for determination
of grain size in nanocrystalline materials.
Representative AFM images of treated samples are illustrated in
Fig. 2(a) and (b). As shown, the grain size of samples punched for
30 min is about 85 nm, while samples punched for 120 min shows
their grain size in the range of 30 nm. The decrease in the grain
size with the punching duration is understandable. Based on the
dislocation theory [2931], repeated punching should introduce
high-density dislocations with a dense dislocation network consisting of nano-sized dislocation cells, which can be turned into
nano-grains by subsequent recovery treatment. Longer duration of
punching helps to reduce the grain size, since repeated impact leads
to more accumulated dislocations or higher dislocation density (t ),
causing a decrease in the average cell diameter (D) as expressed by
the following equation [32]:
K1
D=
t

(1)

where K1 is a material constant.


The accumulated plastic deformation introduced by repeated
punching should more or less increase the sub-grain misorientation [33], which raises the energy barrier to possible annihilation
of dislocation walls or sub-boundaries at elevated temperatures.
This also facilitates the formation of nanocrystalline structures
with smaller and more stable nano-sized grains, corresponding to
increased hardness or a larger barrier to dislocation movement.

2877

Fig. 3. Loaddepth curves of the original sample and a nanocrystallized one


(punched for 30 min and annealed at 350 C for 1 h) under indentation loads of 50,
100 and 200 N, respectively.

3.2. Effects of the punching duration on thickness and mechanical


behavior of the nano-layer
Hardness and  ratio of treated and untreated sample surfaces
were evaluated using a nano-mechanical probe (nano-indenter).
Fig. 3 illustrates loaddepth curves of an original sample and a
nanocrystallized one (Punched for 30 min and annealed at 350 C
for 1 h) measured under various loads of 50, 100 and 200 N,
respectively, using a nano-mechanical probe. From the loaddepth
curves, the hardness (Hv) and  values of the samples were determined. In general, under a xed load, the smaller the indentation
depth, the harder is the sample. As shown in Fig. 3, the nanocrystallized surface was considerable harder than the original one, or it
exhibited much higher resistance to penetration.
The ratio () of the elastic deformation energy (We) to the total
deformation energy (Wt) is a useful parameter reecting the elastic behavior of a material [34]. The area enclosed by the unloading
curve and the maximum penetration depth represents the elastic deformation energy, We, while the total deformation energy,
Wt, is represented by the area enclosed by the loading curve and
the maximum penetration depth. The ratio,  = We/Wt, represents
the contribution of elastic deformation to the total deformation,
which reects the elastic behavior of a material or its capability
to absorb deformation energy without damage. As shown in Fig. 3,
the nanocrystalline surface has a much higher  value. For instance,

Fig. 2. Representative AFM image of samples punched for 30 min (a) and 120 min (b), respectively, and annealed at 350 C for 1 h.

2878

X.Y. Mao et al. / Materials Science and Engineering A 527 (2010) 28752880

the  values of the original sample and a nanocrystalline one are


approximately 15% and 60%, respectively, under a load of 200 N.
The larger  value is related to a greater capability to reduce the
mechanical damage.
One of main objectives of this study is to increase the thickness of the nanocrystallized surface by continuous punching with
subsequent recovery treatment. In order to determine the effect of
punching duration on the thickness of the nanocrystallized layer,
cross-sectional hardness proles of punched samples were measured using a micromechanical probe (micro-indenter) under a
maximum load of 50 mN. We also measured top surfaces hardness
values of nanocrystallized samples having different grain sizes to
establish the hardness nano-grain size relationship so that indirectly we could approximately determine variations in grain size
versus the distance from the top surface of nanocrystallized samples. In this way, we may obtain the information on the inuence of
the punching duration on the thickness of nanocrystallized surface
layer. Fig. 4 illustrates the cross-sectional hardness proles of samples punched for 30, 60, 90, 120 min, respectively, and recovered at
350 C for 1 h. As shown, increasing the punching duration resulted
in higher hardness and larger thickness of the nanocrystallized
layer. For instance, the sample punched for 30 min had its hardness equal to 360 Hv at the top surface (corresponding to an average
grain size of 85 nm) and the hardness decreased to a stable value of
90 Hv in the unaffected zone after the depth reached 70 m. While
when the punching duration was increased to 120 min, both the
hardness of top surface and the thickness of nanocrystallized layer
increased. As shown in Fig. 4, the top surface of the sample punched
for 120 min had a higher hardness of 540 Hv, corresponding to an
average grain size of 30 nm, and the lower and stable hardness of
unaffected zone was reached at a distance about 200 m from the
top surface. Clearly, prolonging the punching duration increased
the thickness of nanocrystallized layer and reduced the grain size.

Fig. 5. Average grain size with the punching time (annealed at 350 C for 1 h).

For more quantitative values, Fig. 5 illustrates how the average


grain size at top surface decreases with an increase in the punching
duration. If we dene 100 nm as a cut-off grain size for nanocrystallized surface layer, then the thickness of surface layer having grains
smaller than 100 nm in the sample punched for 30 min is approximately 30 m, while that in the sample punched for 120 min is
around 120 m.
As has been demonstrated, a longer punching duration generates more dislocations, forming smaller dislocation cells as Eq. (1)
indicates. The thickness of the nanocrystallized layer also increases
with the punching duration. However, under a xed punching force
or impact energy, the increase in thickness of affected surface layer

Fig. 4. Hardness versus the depth for samples punched for 30, 60, 90, 120 min, respectively. Annealing condition: 350 C, 1 h.

X.Y. Mao et al. / Materials Science and Engineering A 527 (2010) 28752880

2879

Fig. 8. EWFs of samples having different grain sizes.

Fig. 6. Potentiodynamic polarization curves of samples with different grain sizes


measured in a 3.5 wt.%NaCl solution.

may become saturated as Fig. 4 illustrates, since the plastic ow


stress waves cannot penetrate too deep as they are blocked by
the plastically deformed surface layer. However, further increase in
thickness of the nanocrystallized layer is possible if larger punching
forces or impact energies are used.
3.3. Corrosion resistance and the electron work function
Potentiodynamic polarization tests were performed to evaluate
the corrosion resistance of nanocrystallized surfaces having various nano-grain sizes in a 3.5 wt.%NaCl solution. Fig. 6 illustrates
obtained polarization curves. From the polarization tests, corrosion
potentials and passivation currents of the samples were determined (see Fig. 7). As demonstrated, the nanocrystallized surfaces
showed lower passivation currents and higher corrosion potentials,
in comparison with those of the microcrystalline sample. The corrosion resistance increases with a decrease in the grain size, which
is attributed to the fact that as the grain size is reduced to the
nano-scale, the high-density grain boundaries make the surface
more active with enhanced local atomic migration, thus accelerating passivation and forming a more protective or stable lm with

Fig. 7. Effects of the grain size on corrosion potential and passivation current.

enhanced passive lms interfacial adherence through, e.g., possible pegging of the passive lm into the substrate with high-density
grain boundaries [35].
During corrosion, material loss occurs through electrochemical
reactions. The exchange of electrons between a metal and surrounding medium constitutes an electric circuit at the interface.
Corrosion is related to the surface electron behavior or electron
activity [35,36], which may be characterized by the electron work
function (EWF) that is the minimum energy required to move an
electron from inside a metal to its surface [37]. EWF is an informative parameter for further understanding the corrosion behavior or
electrochemical stability of a material. In this work, EWFs of sample surfaces experienced punching and recovery treatment were
measured after light polishing and etching.
Fig. 8 illustrates representative EWFs of nanocrystallized samples having different grain sizes. As shown, the EWF increases as
the grain size decreases, corresponding to an increase in the surface inertness to electrochemical attack. The rise of EWF resulting
from surface nanocrystallization is largely attributed to the formation of an enhanced passive lm. The activity or inertness of a
metal can be characterized by its absolute electrode potential (Em ),
which is dependent on the electron work function (EWF: m ) of
the metal and the contact potential difference (m
) at the metal
s
(M)solution (S) interface [38]
E m = m + m
s

(2)

Thus, the surface reactivity or corrosion behavior of a metal


is intrinsically determined by the electron work function and the
surrounding medium or environment that inuences the surface
electrochemical reactions as well, i.e. affecting the term of m
.
s
If a more protective passive lm forms on the metal, escape of
electrons from the metal will be more difcult, corresponding to
a higher apparent or measured EWF.
The observed increase in EWF with a decrease in the grain size
is a good indication that the passive lm on a nanocrystalline surface is more effective to reduce the electron activity and enhances
the resistance of the surface to corrosion. It needs to be indicated
that Cu30Ni alloys is passivable in NaCl solutions, in which a
CuO/Cu(OH)2 /Cu2 O passive layer forms on surface of the alloys
[39]. It has been well claried that, unlike non-passive materials,
nanocrystallization of passive alloys results in enhanced resistance
to corrosion [35]. Since the thickness of nanocrystallized surface
layer and grain size vary with the punching time, higher corrosion
resistance and longer service life can be achieved by prolonging
the punching process. Such a treated surface is more durable when
used in a corrosive environment. If mechanical attacks are involved
simultaneously, e.g., corrosive wear, a thicker nanocrystalline layer
would be more protective than thin nanocrystalline layers.

2880

X.Y. Mao et al. / Materials Science and Engineering A 527 (2010) 28752880

4. Conclusions
This study demonstrates a simple surface nanocrystallization
process combining repeated punching and subsequent recovery
treatment to produce nanocrystalline surface layers, using Cu30Ni
alloy as an example. It was demonstrated that the surface grain
size decreased while the thickness of the nanocrystallized layer
increased with respect to the punching period or duration, resulting
in increases in both hardness and corrosion resistance of Cu30Ni
alloy (in a NaCl solution) with a longer expected service life. The
surface EWF also increased with surface nanocrystallization, corresponding to higher surface stability and enhanced corrosion
resistance.
Acknowledgements
This work is nancially supported by Natural Science and Engineering Research Council of Canada, Major State Basic Research
Development Program of China (Grant No.2007CB616903), and
China Scholar Council (CSC). The authors would like to thank Dr.
Xinhu Tang at University of Alberta for his assistance in using
various instruments for this study and valuable suggestions. The
authors are also grateful to express their thanks to Gao Xin Zhangtong Limited Company (China) for the provision of experimental
materials of Cu30Ni alloy tube.
References
[1] K.S. Kumar, H. Van Swygenhoven, S. Suresh, Acta Materialia 51 (2003)
57435774.
[2] H. Gleiter, Progress in Materials Science 33 (4) (1989) 223315.
[3] C. Suryanarayana, International Materials Reviews 40 (1995) 4164.
[4] K. Lu, Materials Science and Engineering R 16 (4) (1996) 161221.
[5] B. Yu, R.S. Hodges, E. Davis, R.T. Irvin, D.Y. Li, Nanotechnology 19 (2008) 335101.
[6] I.B. Beech, J. Sunner, Current Opinion in Biotechnology 15 (2004) 181186.
[7] C.C. Koch, Nanostructure Materials 2 (2) (1993) 109129.
[8] R.Z. Valiev, A.V. Korznikov, R.R. Mulyukov, Materials Science and Engineering
A 168 (2) (1993) 141148.
[9] U. Erb, A.M. El-Sherik, G. Palumbo, K.T. Aust, Nanostructure Materials 2 (4)
(1993) 383390.

[10] Y. Iwnhashi, Z. Horita, M. Nemoto, T.G. Langdon, Acta Materialia 46 (9) (1998)
33173331.
[11] F. Wetscher, A. Vorhauer, R. Pippan, Materials Science Engineering A 410/411
(25) (2005) 213216.
[12] K. Lu, J. Lu, Journal of Science and Technology 15 (1999) 193197.
[13] J.W. Ren, A.D. Shan, J.B. Zhang, H.W. Song, J.L. liu, Materials Letters 60 (2006)
20762079.
[14] W.L. Li, N.R. Tao, K. Lu, Scripta Materialia 59 (2008) 546549.
[15] T. Wang, J. Yu, B. Dong, Surface and Coatings Technology 200 (2006) 47774781.
[16] X. Wu, N.R. Tao, Y. Hong, B. Xu, J. Lu, K. Lu, Acta Materialia 50 (8) (2002)
20752084.
[17] Y. Todaka, M. Umemoto, J. Li, K. Tsuchiya, Reviews on Advanced Materials
Sciences 10 (2005) 409416.
[18] M. Umemoto, K. Todaka, K. Tsuchiya, Materials Science Engineering A 375377
(2004) 899904.
[19] D.A. Hughes, D.B. Dawson, J.S. Korellls, L.I. Weingarten, Journal of Materials
Engineering and Performance 3 (4) (1994) 459475.
[20] L. Waltz, D. Retraint, A. Roos, P. Olier, Scripta Materialia 60 (1) (2009) 2124.
[21] M. Sato, N. Tsuji, Y. Minamino, Y. and, Koizumi, Materials Science Forum
426432 (2003).
[22] D.Y. Li, L. Wang, W. Li, Materials Science and Engineering A 384 (1/2) (2004)
355360.
[23] X.S. Guan, Z.F. Dong, D.Y. Li, Nanotechnology 16 (2005) 29632971.
[24] X.Y. Wang, D.Y. Li, Electrochemistry Acta 47 (2002) 39393947.
[25] X.Y. Mao, F. Fang, F. Yang, J.Q. Jiang, R.S. Tan, Journal of Materials Processing
Technology 209 (2009) 21452151.
[26] J.K.W. Roberta, Wear 261 (2006) 10121023.
[27] W. Li, D.Y. Li, Acta Materialia 53 (14) (2005) 38713878.
[28] I.D. Baikie, P.J.S. Smish, D.M. Portereld, P.J. Estrup, Review of Scientic Instruments 70 (3) (1999) 18421850.
[29] J.P. Hirth, J. Lothe, Theory of Dislocations, 2nd ed., John Wiley & Sons, Inc., New
York, 1982.
[30] T.H. Courtney, Mechanical Behavior of Materials, 2nd ed., McGraw Hill, Boston,
2000.
[31] P.W.J. Mckenzie, R. Lapovok, Y. Estrin, Acta Materialia 55 (2007) 29852993.
[32] Y. Estrin, L.S. Toth, A. Molinari, Y.B. Chet, Acta Materialia 46 (1998) 55095522.
[33] M. Ohnami, Plasticity and High Temperature Strength of Materials, Elsevier
Applied Science, London, 1988.
[34] R. Liu, D.Y. Li, Y.S. Xie, R. Liewellyn, H.M. Hawthorne, Scripta Materialia 41
(1999) 691.
[35] S. Tao, D.Y. Li, Philosophical Magazine Letters 88 (2008) 137144.
[36] W. Li, D.Y. Li, Applied Surface Science 240 (2005) 388395.
[37] N.W. Ashcroft, N.D. Mermin, Solid State Physics, Saunders College Publishing,
New York, 1976.
[38] J.O. Bockris, S.U.M. Khan, Surface Electrochemistry: A Molecular Level
Approach, Springer, New York, 1993.
[39] I. Milosev, M. Metikos-Hukovis, Electrochimica Acta 42 (10) (1997) 15371548.
[40] E. Hosseini, M. Kazeminezhad, Computational Materials Science 44 (2009)
11071115.

You might also like