You are on page 1of 6

Journal of Non-Crystalline Solids 379 (2013) 48–53

Contents lists available at ScienceDirect

Journal of Non-Crystalline Solids


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

Corrosion characterization on melt spun Cu60Zr20Ti20 metallic glass:


An experimental case study
S. Vincent a, A.F. Khan a, B.S. Murty b, Jatin Bhatt a,⁎
a
Department of Metallurgical and Materials Engineering, V.N.I.T, Nagpur-440010, India
b
Department of Metallurgical and Materials Engineering, Indian Institute of Technology Madras, Chennai-600036, India

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

Article history: In this study, corrosion behavior of Cu60Zr20Ti20 (at. %) metallic glass is investigated in acidic and neutral
Received 23 April 2013 chloride solutions, at varying molar concentrations of chloride and also in alkaline solutions. Further, to in-
Received in revised form 17 June 2013 vestigate Cu60Zr20Ti20 glassy ribbon in biological conditions, it is evaluated in phosphate buffered saline
Available online 25 August 2013
(PBS) solution. Results indicate that corrosion resistance decreases with increase in chloride concentration
in case of acidic and neutral solutions. While in case of alkaline solution, corrosion resistance is found to in-
Keywords:
Metallic glass;
crease with increase in pH of solution. The glassy alloy has exhibited excellent corrosion resistance in PBS
Melt spinning; medium. In order to understand the corrosion behavior, the surfaces of the samples were studied using scan-
Corrosion resistance; ning electron microscopy (SEM). SEM examination reveals pitting type of corrosion. The possible reasons for
X-ray diffraction; pitting corrosion in metallic glasses have been addressed in the present study.
Scanning electron microscopy © 2013 Elsevier B.V. All rights reserved.

1. Introduction glasses in strong acids (i.e., H2SO4 and HNO3) and alkalies (i.e., NaOH)
[16] and relative unstable behavior in aggressive anions containing F−
Pronounced short and medium range order linked with densely and Cl− ion leads to pitting corrosion due to selective dissolution of Zr
packed liquid structure of metallic glasses [1] results in absence of has been addressed [17]. The higher corrosion resistance of Ti rich me-
crystalline defects such as grain boundaries, dislocations, stacking tallic glasses has been attributed to formation of stable protective film
faults, etc. Consequently, absence of such periodic structural features due to presence of Co, Nb and Ta but pitting corrosion still occurs at
results in unusual combination of properties such as high strength, the potential higher than 0 V due to the existence of aggressive Cl−
large elastic limit, lower modulus, high hardness, high fatigue limit, ion [15]. The importance of higher glass forming ability, good ductility
excellent formability, better wear resistance and excellent corrosion and lower cost of Cu based metallic glasses; Cu–Zr–Ti system in partic-
resistance [2–6]. Acknowledging those properties makes metallic ular, has attracted many researchers to evaluate mechanical and chem-
glasses an effective metallic material for various structural and sport- ical properties [18–21]. Corrosion behavior of binary Cu–Zr metallic
ing applications [7]. Metallic glasses also exhibits desirable properties glass has been investigated and demonstrated for pit formation in Cl−
for bio-medical applications i.e. low modulus of elasticity resulting in ion containing solutions [22]. A few reports are available on corrosion
stress shielding effect and good wear resistance to minimize wear de- behavior of Cu60Zr30Ti10 metallic glass in different electrolytes [23,24].
bris formation along with corrosion resistance to withstand harsh So far, evaluation on corrosion properties of Cu60Zr20Ti20 is not reported
in-vivo environments [8]. In addition, high strength to weight ratio to the best of the author's knowledge. Thus, the aim of the present work
along with chemical and structural homogeneity in metallic glass fa- is to examine the behavior of Cu60Zr20Ti20 under (1) different corrosive
vors biocompatibility, as a smaller implant could potentially cause environments such as acidic, neutral and alkaline media using HCl,
less damage to host tissue and thus minimize subsequent foreign NaCl and NaOH solutions; (2) physiologically relevant environment by
body response [8]. testing them in phosphate buffered saline (PBS) solution. The present
For any practical applications, it is extremely important to evaluate findings discuss the corrosion behavior of Cu60Zr20Ti20 metallic glass
chemical stability of materials under potential environments and thus using in-vitro electrochemical studies.
study of corrosion behavior becomes necessary. Corrosion behavior of
some Zr and Ti based metallic glasses in different electrolytes is well re- 2. Experimental details
ported such as Zr–Al–Co–Ag [9], Zr–Al–Co–Nb [10,11], Zr–Al–Cu–Ni–Pd
[12], Zr–Cu–Al [13], Ti–Zr–Ni–Be [14], Ti–Cu–Ni–Zr–M (M = Cu, Co, Nb Mixture of Cu, Zr and Ti pure metals (N99.9 mass %) were arc melted
or Ta) [15]. The excellent corrosion resistance of Zr based metallic in a Ti-gettered argon atmosphere to obtain alloy ingot with nominal
composition of Cu60Zr20Ti20 (at. %). Homogeneity in composition is
⁎ Corresponding author. Tel.: +91 712 2801513. achieved by remelting the alloy ingot for several times. Cu60Zr20Ti20 rib-
E-mail address: jatinbhatt@mme.vnit.ac.in (J. Bhatt). bon samples with cross section of 0.07 mm thickness and 2.6 mm width

0022-3093/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jnoncrysol.2013.07.007
S. Vincent et al. / Journal of Non-Crystalline Solids 379 (2013) 48–53 49

a Melt Spun Cu60Zr20Ti20


b

Intesity (a.u)

20 30 40 50 60 70 80 90
2θ (Degree)

Fig. 1. (a) XRD pattern and (b) SEM micrograph (before corrosion) of melt spun Cu60Zr20Ti20 metallic glass.

were produced by melt spinning with a wheel velocity of 20 m/s in an electrical isolation. The WE was allowed to attain a stable open circuit
argon atmosphere. The structure of melt spun ribbon was examined potential (OCP) in the test solution. All the samples attained stable
by X-ray diffraction (XRD) using a PANalytical X'pert-Pro diffractometer OCP in 30 min. 400 ml of fresh test solution was used for every polari-
with Cu–Kα (λ = 1.54 Ǻ) radiation on solid flat melt spun ribbons sur- zation test. The polarization curves were measured at least three
face. The XRD is operated at reflection mode at 40 kV and 40 mA with a times to confirm reproducibility. The surface morphologies of the rib-
beam size of 12 mm wide × 5 mm long. bon before and after corrosion were examined using scanning electron
The corrosion behavior of Cu60Zr20Ti20 ribbon samples was eval- microscopy (SEM) in back scattered (BS) mode coupled with energy
uated by electrochemical polarization experiments using VersaSTAT dispersive spectroscopy (EDS) in a JEOL JSM-820.
3 potentiostat. Test was performed at room temperature in a three-
electrode cell using saturated calomel reference electrode (SCE)
(USCE = 241 mV) and platinum gauge counter electrode. Prior to 3. Results
electrochemical measurements, the specimens were ultrasonically
degreased in acetone, washed in distilled water and dried with pulsed Fig. 1(a) shows XRD analysis of Cu60Zr20Ti20 melt spun ribbon. The
air. For testing in acidic and neutral media, electrolytes used are presence of broad diffused hump and absence of any sharp crystalline
aqueous solutions of HCl and NaCl, respectively, with concentrations of peak in the diffraction pattern confirm the amorphous nature of melt
0.01, 0.1 and 1 M. Tests were conducted in alkaline media at two differ- spun ribbon. Fig. 1(b) shows surface microstructure of melt spun
ent pH levels, 8 and 10. Alkaline solutions were prepared by adjusting the Cu60Zr20Ti20 before corrosion.
pH of 0.1 M NaCl with 0.1 M NaOH. Test was also conducted in PBS solu- The potentiodynamic polarization measurements were performed
tion having pH of 7.2. All electrolytes were prepared from reagent grade on Cu60Zr20Ti20 melt spun ribbon to evaluate its corrosion behavior. Po-
chemicals and distilled water. Potentiodynamic polarization experi- larization results indicate that the corrosion process in the present sys-
ments were performed at a scan rate of 0.1666 mV/s. The working elec- tem is under cathodic control as reflected by the Tafel slopes. Fig. 2
trode (WE) was exposed only to an area of 0.8–1.2 cm2 while the rest of shows polarization curves for Cu60Zr20Ti20 melt spun in neutral NaCl so-
the specimen was embedded in a thermoplastic resin to provide lution (pH = 7) at three different molar concentrations (0.01, 0.1 and
1 M). Corrosion current density (icorr) and corrosion potential (Ecorr)
values were determined by applying Tafel extrapolation of polarization
curves. The current density increases to a very high value after the Ecorr
-0.1
is reached indicating high rate of metal dissolution and no sign of pas-
sivity was observed. The values of icorr and Ecorr for melt spun ribbon
-0.2
0.01M NaCl in NaCl solution are shown in Table 1. From Fig. 2 and Table 1, it is evi-
0.1M NaCl dent that the higher the concentration of Chloride (Cl−), the higher is
-0.3
1M NaCl the icorr value which in turn indicates that the rate of corrosion increases
with increase in concentration of Cl− in NaCl solution. Corrosion in gen-
Potential, E (V)

-0.4
eral is enhanced by the presence of anions like S2− and Cl− and in
metallic glasses, preferential adsorption of Cl− ion is responsible for re-
-0.5
placement of metal cations. It can also be observed from Table 1 that
Ecorr becomes more negative with the increase in concentration of Cl−
-0.6
ion. This further indicates that stability of melt spun ribbon decreases
with increase in Cl− ion concentration.
-0.7
Furthermore, surface morphology examination using SEM on these
ribbon samples gives us more information on their corrosion behavior.
-0.8
Fig. 3 shows SEM images of sample subjected to potentiodynamic polar-
1E-11 1E-10 1E-9 1E-8 1E-7 1E-6 1E-5 1E-4 1E-3 ization measurements in NaCl neutral solution. SEM micrographs indi-
Current Density, I (A/cm2) cate pitting type of corrosion. From Fig. 3(a) it is evident that at
0.01 M Cl− concentration, pits are less penetrative in terms of depth of
Fig. 2. The polarization curves at different Cl− concentrations in neutral NaCl solution. area under attack. For 0.1 M Cl− concentration as shown in Fig. 3(b),
50 S. Vincent et al. / Journal of Non-Crystalline Solids 379 (2013) 48–53

Table 1 -0.1
0.01M HCl
The corrosion potential (Ecorr) and corrosion current density (Icorr) for Cu60Zr20Ti20 0.1M HCl
glassy alloy in different aqueous solutions. -0.2
1M HCl
Type of solution Concentration, M icorr Ecorr
-0.3
μA/cm2 V vs. SCE

Potential, E (V)
-0.4
NaCl (neutral) 0.01 0.0259 −0.1589
0.1 0.16612 −0.2228
1 5.1525 −0.3019 -0.5
HCl (acidic) 0.01 0.4621 −0.1769
0.1 1.1832 −0.2466 -0.6
1 6.2896 −0.4138
NaOH (alkaline) pH 8 0.4976 −0.2342 -0.7
pH 10 0.2099 −0.2118
PBS pH 7.2 0.0216 −0.1243
-0.8

-0.9
1E-7 1E-6 1E-5 1E-4 1E-3 0.01
pits are more penetrative indicating decrease in corrosion resistance of Current Density, I (A/cm2)
melt spun ribbon with increase in Cl− ion concentration.
It is well known that Cl− ion concentration in normal sea water is Fig. 4. The polarization curves at different Cl− concentrations in acidic HCl solution.
around 2.5 to 3.5 wt.%. In 1 M NaCl solution, concentration of Cl− ion
is around 6 wt.% which is two times stronger corrosive environment Large numbers of pits were observed for sample tested at 0.1 M
than normal sea water. Conventional metals such as 304 L stainless (Fig. 5(b)) in comparison with 0.01 M Cl− concentration (Fig. 5(a))
steel (SS) exhibits Ecorr of − 0.69 V (vs. SCE) in 1 M NaCl solution which again indicates that corrosion resistance of melt spun sample
[25] whereas Cu60Zr20Ti20 metallic glass exhibits Ecorr of − 0.30 V decreases with increase in Cl− concentration.
(vs. SCE) (Table 1). This suggests that Cu60Zr20Ti20 metallic glass is Fig. 6 shows polarization curves for Cu60Zr20Ti20 melt spun ribbon
nobler than 304 L SS. Furthermore 304 L SS is reported to exhibit a in NaOH alkaline solution. The tests were carried out in two different
very high icorr (~ 105 μA/cm2) [25] in comparison to melt spun ribbon pH solutions i.e. pH 8 and 10, respectively. The icorr and Ecorr were de-
(~ 5 μA/cm2) which indicates that corrosion rate in the former is termined from Fig. 6 and shown in Table 1. The results exhibit differ-
much higher than the latter. The tests performed in present investiga- ent trend in comparison with neutral and acidic Cl− solutions. From
tion are from very low (0.01 M) to a very high (1 M) concentration of Fig. 6 and Table 1, it can be observed that increase in pH resulted in
Cl− ions and results indicates higher corrosion resistance of metallic decrease in icorr. Ecorr gets nobler with increase in pH of solution.
glasses in comparison to stainless steel. This can be attributed to the This behavior indicates that the rate of corrosion decreases with in-
presence of chemical and structural homogeneity in metallic glasses, crease in pH. This further signifies that Cu60Zr20Ti20 melt spun ribbon
thus inhibiting the absorption of Cl− ions on glassy surface. exhibits better corrosion resistance in alkaline conditions than acidic
Potentiodynamic polarization curves for Cu60Zr20Ti20 melt spun in and neutral pH. The above results are evident from SEM examination
HCl acidic solution is shown in Fig. 4. The tests were carried out in shown in Fig. 7. SEM for melt spun sample subjected to corrosion test
three different molar concentrations of HCl i.e. 0.01, 0.1 and 1 M, re- in pH 8 exhibited deep pits (Fig. 7(a)) whereas for pH 10, pits were
spectively. The values of icorr and Ecorr were determined from Fig. 4 observed only on the surface, which further confirms that corrosion
and shown in Table 1. The results exhibited by acidic Cl− solution fol- resistance of melt spun sample increases in alkaline conditions.
low same trend as that of neutral Cl− solutions i.e. increase in Cl− Fig. 8 shows behavior of Cu60Zr20Ti20 melt spun in terms anodic and
concentration resulted in increase in icorr. Though the trend is the cathodic polarization curves in biological conditions using PBS solution.
same, values exhibited by icorr in acidic Cl− solution is very high in A pH of 7.2 is maintained for PBS solution which is similar to body fluid
comparison with neutral Cl− solution. The above results indicate conditions. As evident from Fig. 8, the polarization curve exhibits signif-
that Cl− is more aggressive in acidic conditions. This may be due to icant passive region which implies better corrosion resistance of melt
dissolution of the protective film formed on the surface of melt spun ribbon in biological conditions. The icorr and Ecorr were determined
spun ribbon. Also Ecorr is more negative as compared to neutral solu- from Fig. 8 and shown in Table 1. It can be seen from Table 1 that lowest
tion of Cl−, which again signifies aggressive nature of Cl− in acidic icorr value is observed for PBS solution. Further, melt spun ribbon
conditions. SEM examinations on these samples are shown in Fig. 5. also exhibits lowest Ecorr in this solution which indicates the passive

a b

Fig. 3. SEM micrographs of corroded surface in NaCl solution at (a) 0.01 M and (b) 0.1 M concentration.
S. Vincent et al. / Journal of Non-Crystalline Solids 379 (2013) 48–53 51

a b

Fig. 5. SEM micrographs of corroded surface in HCl solution at (a) 0.01 M and (b) 0.1 M concentration.

behavior of alloy in biological conditions. SEM examination on sample possible mechanism for Cl− induced pitting is explained as follows
subjected to PBS media (Fig. 9) indicates formation of small amounts [30]. At the interface between these small crystallites and glassy matrix,
of pits on the surfaces. Cl− ions can be preferentially absorbed. Pits can be easily initiated at
these regions by destroying the passive film even at lower concentra-
4. Discussion tion of Cl− ions. The glassy matrix gets pitted at glass/crystal interface
creating weak interfacial bonding resulting in detachment of crystalline
Pitting corrosion in metallic glasses is well reported in previous in- particles from the surface, leaving pits behind. Further, Cl− ions accu-
vestigations [22–24]. Though the single phase amorphous nature in me- mulate in these pits and the process gets repeated making the pits
tallic glasses is considered to be resistant to localized pitting corrosion, deeper. With increase in Cl− concentration in both NaCl and HCl solu-
the possible presence of defects in cast sample, chemical heterogene- tion, pit initiation and growth process occurs at a rapid rate thereby
ities and composition fluctuation acts as initiation points for pitting cor- resulting in deeper and large number of pits. In case of alkaline solution,
rosion. It is also well known that pitting corrosion is highly accelerated if with increase in pH, pits density is found to decrease which can be at-
aggressive anions such Cl−, Br− or S− 2 ions are present in environment. tributed to decrease in Cl− concentration in solution.
In the present study, SEM examination on the Cu60Zr20Ti20 metallic The following series of reactions occur in the presence of Cl− ions;
glass has shown pitting corrosion in all electrolyte solution. It has

been reported in earlier studies that presence of nanocrystalline parti- M þ Cl →MClads þ e−
cles in glassy matrix serve as an active sites for Cl− induced pitting cor-
rosion [26–28]. Transmission electron microscopy (TEM) study on melt
spun Cu60Zr20Ti20 metallic glass by the present authors [29] reported þ
MClads →MCl þ e−
the presence of ~ 2–5 nm sized spherical to irregular cluster. Though
the melt spun Cu60Zr20Ti20 ribbon is X-ray amorphous, the presence of
compositional heterogeneity is reported earlier [29]. The presence of
such compositional heterogeneity and small crystallites could be prob- þ 2þ −
MCl →M þ Cl
able reasons for pitting corrosion.
The results in all electrolyte solutions indicate that corrosion occur- Pitting corrosion in Cu60Zr20Ti20 melt spun ribbon could also be due to
ring in melt spun alloy is due to Cl− induced pitting corrosion. The local selective dissolution of galvanically active metals. The standard equi-
librium electrode potentials with respect to normal hydrogen electrode
for the Zr/Zr4+, Ti/Ti2+ and Cu/Cu2+ couples are −1.529 V, −1.628 V
and +0.337 V, respectively, which clearly indicates that Cu is nobler
-0.2 than Zr and Ti. The large electrochemical potential difference between
8 pH Cu and Zr (Ti) acts as a driving force for selective dissolution of Zr /Ti in
-0.3 10 pH Cu–Zr–Ti alloy. Due to an auto-catalytic reaction of Cl− ions at some
weak regions such as compositional heterogeneities and glass/crystal
Potential, E (V)

interface, an enhanced local selective dissolution of less noble metals


-0.4 (Zr and Ti) is possible.
Table 2 shows composition of passive film obtained by EDS after
-0.5 polarization in 0.01 M HCl, 0.01 M NaCl, pH 8 and 10 NaOH, and
PBS solution. The compositions of passive film consists of Cu, Zr, Ti
and O which indicates that passive film formed is a mixture of simple
-0.6
oxides or complex oxidic compounds of all alloying elements. The
passive film in all cases consists of more Cu than Zr and Ti, indicating
-0.7 that the less noble Zr and Ti are leached out of the sample due to cor-
rosion. From Table 2, it can be observed that the content of Cu, Zr and
1E-7 1E-6 1E-5 1E-4 1E-3 Ti is more in 0.01 M HCl in comparison with 0.01 M NaCl. This indi-
Current Density, I (A/cm2) cates that passive film formation is easier and faster in HCl than in
NaCl. It is also evident from SEM images of 0.01 M NaCl (Fig. 3) and
Fig. 6. The polarization curves at different pH values in alkaline NaOH solution. 0.01 M HCl (Fig. 5). Pit morphology in Fig. 3 shows deeper and
52 S. Vincent et al. / Journal of Non-Crystalline Solids 379 (2013) 48–53

a b

Fig. 7. SEM micrographs of corroded surface in NaOH solution at (a) pH 8 and (b) pH 10.

more penetrative pits in comparison with Fig. 5. This can be attribut-


ed to fact that easier and faster passive film formation in HCl solution
offers better resistance to pitting corrosion in comparison with NaCl
solution. Similar interpretation was observed for alkaline solutions
as well in which the content of Cu, Zr and Ti is more for pH 10 than
in pH 8 (Table 2). This could be due to more dissolution of Cu, Zr
and Ti in pH 8 than pH 10 by oxygen. SEM image shows deeper pen-
etrative pits for pH 8 (Fig. 7(a)) than that in pH 10 (Fig. 7(b)) which
clearly suggests that passive film formation is faster and easier at
higher pH and thus better resistance to pitting.

5. Conclusions

The corrosion behavior of Cu60Zr20Ti20 metallic glass is evaluated in


different service environments such as acidic, alkaline, neutral media
and also in physiologically relevant environment. The corrosion rate of
Cu60Zr20Ti20 metallic glass is directly proportional to the Cl− concentra-
tion in acidic and neutral media and varies inversely with pH in alkaline Fig. 9. SEM micrograph of corroded surface in PBS solution.
solution. Cl− induced pitting corrosion is observed in all electrolyte so-
lutions. Small nanocrystallites present in the glassy matrix along with
selective dissolution of less nobler elements are expected to control Table 2
pitting corrosion in the electrolyte solutions. The glassy alloy shows ex- Composition of passive film on the alloy at different aqueous solutions.
cellent corrosion resistance in PBS solution. This hints the possibility of Solution Cu (at.%) Zr (at.%) Ti (at.%) O (at.%)
using Cu60Zr20Ti20 metallic glass as potential biomaterials. However,
Base alloy 59.4 21.5 19.1 –
further efforts are necessary to check cytotoxicity to achieve thorough
0.01 M HCl 46.4 18.4 14.9 20.3
understanding of biocompatibility of Cu60Zr20Ti20 metallic glass. 0.01 M NaCl 39.1 15.2 11.9 33.8
NaOH pH 8 34.6 12.5 11.2 41.7
NaOH pH 10 41.9 18.6 14.7 24.9
0.3 PBS (pH 7.2) 43.4 18.5 14.7 23.4
Phosphate Buffered
Saline Solution
0.2
References
0.1
[1] A.L. Greer, E. Ma, MRS Bull. 32 (2007) 611–619.
[2] W.H. Wang, C. Dong, C.H. Shek, Mater. Sci. Eng., R 44 (2004) 45–89.
Potential, E (V)

0.0 [3] M.Z. Ma, R.P. Liu, Y. Xiao, D.C. Lou, L. Liu, Q. Wang, W.K. Wang, Mater. Sci. Eng., A
386 (2004) 326–330.
-0.1 [4] T. Wada, T. Zhang, A. Inoue, Mater. Trans. JIM 43 (2002) 2843–2846.
[5] T. Zhang, A. Inoue, Mater. Trans. JIM 43 (2002) 267–270.
-0.2 [6] T. Zhang, A. Inoue, Mater. Sci. Eng., A 375 (2004) 432–435.
[7] M. Telford, Mater. Today 7 (2004) 36–43.
[8] L. Huang, Z. Cao, H.M. Meyer, P.K. Liaw, E. Garlea, J.R. Dunlap, T. Zhang, W. He,
-0.3
Acta Biomater. 7 (2011) 395–405.
[9] N. Hua, L. Huang, J. Wang, Y. Cao, W. He, S. Pang, T. Zhang, J. Non-Cryst. Solids 358
-0.4 (2012) 1599–1604.
[10] X.Y. Lu, L. Huang, S.J. Pang, T. Zhang, Chin. Sci. Bull. (2012) 1–5.
-0.5 [11] S.J. Pang, T. Zhang, K. Asami, A. Inoue, J. Mater. Res. 18 (2003) 1652–1658.
[12] F.X. Qin, H.F. Zhang, P. Chen, F.F. Chen, D.C. Qiao, Z.Q. Hu, Mater. Lett. 58 (2004)
1E-11 1E-10 1E-9 1E-8 1E-7 1E-6 1E-5 1246–1250.
[13] H.-H. Huang, Y.-S. Sun, C.-P. Wu, C.-F. Liu, P.K. Liaw, W. Kai, Intermetallics 30
Current Density, I (A/cm2) (2012) 139–143.
[14] M.L. Morrison, R.A. Buchanan, A. Peker, P.K. Liaw, J.A. Horton, J. Non-Cryst. Solids
Fig. 8. The polarization curve for PBS solution. 353 (2007) 2115–2124.
S. Vincent et al. / Journal of Non-Crystalline Solids 379 (2013) 48–53 53

[15] F. Qin, X. Wang, A. Kawashima, S. Zhu, H. Kimura, A. Inoue, Mater. Trans. JIM 47 [23] K. Asami, C.-L. Qin, T. Zhang, A. Inoue, Mater. Sci. Eng., A 375–377 (2004) 235–239.
(2006) 1934–1937. [24] W.-K. An, A.-h. Cai, X. Xiong, Y. Liu, Y. Luo, T.-l. Li, X.-S. Li, Mater. Sci. Appl. 2
[16] A. Gebert, K. Buchholz, A. Leonhard, K. Mummert, J. Eckert, L. Schultz, Mater. Sci. (2011) 546–554.
Eng., A 267 (1999) 294–300. [25] F. Arjmand, A. Adriaens, Int. J. Electrochem. Sci. 7 (2012) 8007–8019.
[17] U.K. Mudali, S. Baunack, J. Eckert, L. Schultz, A. Gebert, J. Alloys Compd. 377 [26] F.X. Qin, H.F. Zhang, Y.F. Deng, B.Z. Ding, Z.Q. Hu, J. Alloys Compd. 375 (2004)
(2004) 290–297. 318–323.
[18] M.K. Tam, S.J. Pang, C.H. Shek, J. Non-Cryst. Solids 353 (2007) 3596–3599. [27] K. Mondal, B.S. Murty, U.K. Chatterjee, Corros. Sci. 47 (2005) 2619–2635.
[19] D. Zander, B. Heisterkamp, I. Gallino, J. Alloys Compd. 434–435 (2007) 234–236. [28] K. Mondal, B.S. Murty, U.K. Chatterjee, Corros. Sci. 48 (2006) 2212–2225.
[20] A. Inoue, W. Zhang, T. Zhang, K. Kurosaka, Acta Mater. 49 (2001) 2645–2652. [29] S. Vincent, Joysurya Basu, B.S. Murty, Jatin Bhatt, Mater. Sci. Eng., A 550 (2012)
[21] A. Inoue, W. Zhang, T. Zhang, K. Kurosaka, Mater. Trans. JIM 42 (2001) 1149–1151. 160–166.
[22] H.B. Lu, L.C. Zhang, A. Gebert, L. Schultz, J. Alloys Compd. 462 (2008) 60–67. [30] C. Suryanarayana, A. Inoue, Bulk Metallic Glasses, CRC Press, Boca Raton, 2010.

You might also like