Вы находитесь на странице: 1из 10

Corrosion Science 44 (2002) 18471856

www.elsevier.com/locate/corsci

Bulk glassy FeCrMoCB alloys


with high corrosion resistance
S.J. Pang *, T. Zhang, K. Asami, A. Inoue
Institute for Materials Research, Tohoku University, 2-1-1 Katahira, Aoba-ku, Sendai 980-8577, Japan
Received 20 July 2001; accepted 13 December 2001

Abstract

The preparation of bulk glassy alloys with high glass-forming ability and high corrosion
resistance in Fe-based system was succeeded by means of copper-mold casting. The temper-
ature interval of supercooled liquid region (DTx ) is as large as 5362 K and the reduced glass
transition temperature (Tg =Tm ) is as high as 0.620.63 for the cast Fe50x Cr16 Mo16 C18 Bx (x 4,
6, 8 at.%) glassy alloys. The corrosion rates of the Fe50x Cr16 Mo16 C18 Bx glassy alloys with a
diameter of 1.2 mm were in the range of 103 102 mm year1 in 1, 6 and 12 N HCl solutions
at 298 K. The bulk glassy alloys are spontaneously passivated in 1 and 6 N HCl solutions
with wide passive region and low passive current density. They do not suer pitting corro-
sion even when polarized anodicly in 12 N HCl solution up to 1.0 V (Ag/AgCl). The high
corrosion resistance is due to the formation of chromium-rich passive lms during immersion
in HCl solutions. In addition, the increase of boron content in alloys improves the corrosion
resistance of the bulk glassy alloys within the composition range examined. 2002 Elsevier
Science Ltd. All rights reserved.

Keywords: Bulk metallic glass; Iron-based alloy; Copper-mold casting; Glass-forming ability; Spontaneous
passivation

1. Introduction

For the last three decades, a variety of corrosion resistant amorphous alloys have
been prepared [15]. However, they exhibit low glass-forming ability (GFA) and
have been mainly fabricated as melt-spun ribbons and sputtered thin lms.

*
Corresponding author. Tel.: +81-22-215-2114; fax: +81-22-215-2111.
E-mail address: janep@imr.tohoku.ac.jp (S.J. Pang).

0010-938X/02/$ - see front matter 2002 Elsevier Science Ltd. All rights reserved.
PII: S 0 0 1 0 - 9 3 8 X ( 0 2 ) 0 0 0 0 2 - 1
1848 S.J. Pang et al. / Corrosion Science 44 (2002) 18471856

The limitations of size and shape as well as low thermal stability have prevented a
further extension of their application elds. Since the synthesis of bulk glassy alloys
in Mg- [6], Ln- [7] and Zr- [8,9] based systems which have a large supercooled liquid
region before crystallization and high resistance against crystallization, the glassy
alloys have attracted much attention as new materials in science and engineering
elds [10]. Subsequently, their alloy systems have been signicantly extended for the
last ve years, such as to Fe-, Co- and Ni-based systems [1014]. However, no de-
tailed data concerning corrosion resistance of the bulk glassy alloys in aggressive
media have been reported. As is known, Zr-based bulk glassy alloys are not cor-
rosion resistant in aggressive HCl solutions, although they exhibit good corrosion
resistance in H2 SO4 and NaCl solutions [15]. It is therefore essential to develop new
corrosion resistant bulk glassy alloys for the application of glassy alloys to engi-
neering elds with aggressive environments, such as chemical applications. The rst
Fe-based bulk glassy alloys were prepared in Fe(Al, Ga)(P, C, B) and FeM(Al,
Ga)(P, C, B) (M Cr, Mo, Nb) systems in 1995 [11], and followed by Fe(Zr, Nb)
B [10] and FeGa(P, C, B) [16] systems. All these glassy alloys have a high GFA
which is a main requisite for the formation of bulk glassy alloys. However, their
corrosion resistance was not conrmed. For Fe-based bulk glassy alloys, high me-
chanical strength, soft magnetic properties and high corrosion resistance are ex-
pected. Consequently, it is important to search for new Fe-based bulk glassy alloys
corrosion resistant in aggressive media.
In a previous work [17] the present authors have found that melt-spun FeCr
MoCB alloys have a high GFA and good corrosion resistance in aggressive HCl
solutions. The present paper aims to clarify the glass-formation composition range
and corrosion behavior of bulk glassy Fe50x Cr16 Mo16 C18 Bx alloys prepared by a
copper-mold casting technique.

2. Experimental procedures

The alloys with nominal compositions of Fe50x Cr16 Mo16 C18 Bx (x 3, 4, 6, 8, 10


at.%) were investigated. The ingots were prepared from a mixture of pure Fe, Cr,
Mo, C and B by high frequency induction melting in an argon atmosphere. From the
ingots, bulk alloy samples in a rod form with a diameter of 1.2 mm were prepared by
a copper-mold casting method. The structure of specimens thus prepared was ex-
amined by X-ray diraction. The thermal stability associated with glass transition,
supercooled liquid and crystallization was investigated by dierential scanning cal-
orimetry (DSC) at a heating rate of 0.67 K s1 . The melting temperature (Tm ) was
measured by dierential thermal analysis.
Corrosion behavior of the glassy alloys was evaluated by weight loss and elect-
rochemical measurements. Prior to immersion test and electrochemical measure-
ments, the specimens were degreased in acetone, washed in distilled water and dried
in air. Electrolytes used were 1, 6 and 12 N HCl solutions, which were prepared from
reagent grade chemical and distilled water. The corrosion rates were estimated from
the weight loss after immersion in HCl solutions open to air at 298 K for one week.
S.J. Pang et al. / Corrosion Science 44 (2002) 18471856 1849

After the immersion test, the specimens were washed in distilled water, dried, and
subjected to weight loss measurement and observation by scanning electron micro-
scope (SEM). Electrochemical measurements were conducted in a three-electrode cell
using a platinum counterelectrode and a Ag/AgCl reference electrode. Potentiody-
namic polarization curves were measured with a potential sweep rate of 50 mV min1
in HCl solutions open to air at 298 K after immersing the specimens for 20 min,
when the open-circuit potentials became almost steady.
For surface analysis of the bulk metallic glasses before and after immersion in HCl
solutions at room temperature controlled at 298 K, X-ray photoelectron spectro-
scopy (XPS) measurements were performed by using a SSI SSX-100 photoelectron
spectrometer with Al-Ka excitation (hm 1486:6 eV). The composition and thick-
ness of the surface lm and the composition of the underlying alloy surface were
determined quantitatively by a previously proposed method using the integrated
intensities of photoelectrons under the assumption of a three-layer model [18,19].
The binding energies of electrons were calibrated by using the method described
elsewhere [20,21].

3. Results and discussion

The cast rod samples of Fe50x Cr16 Mo16 C18 Bx (x 4, 6, 8 at.%) with a diameter of
1.2 mm consisted of a single glassy phase as was evidenced from a main halo peak
without crystalline peaks in their X-ray diraction patterns. They had a smooth
surface and metallic luster, and no contrast of a crystalline phase was seen over the
outer surface. On the other hand, the diraction patterns of the cast Fe50x Cr16 -
Mo16 C18 Bx (x 3, 10 at.%) alloys exhibited crystalline peaks besides a broad peak,
indicating that they were composed of glassy phase plus crystalline phases. There-
fore, the compositional range of the bulk glass-forming alloys in this system is
Fe50x Cr16 Mo16 C18 Bx (x 4, 6, 8 at.%) under this casting condition.
Fig. 1 shows the DSC curves of the bulk glassy Fe50x Cr16 Mo16 C18 Bx (x 4, 6, 8
at.%) rods with a diameter of 1.2 mm, where Tg and Tx correspond to glass transition
temperature and onset temperature of crystallization, respectively. These glassy alloys
exhibit the sequential transition of the glass transition, supercooled liquid region and
crystallization. The Tg increases with increasing boron content. As shown in Table 1,
the supercooled liquid region, DTx , dened by the dierence between Tg and Tx is as
large as 5362 K, implying a high thermal stability of the supercooled liquid. The
reduced glass transition temperature dened as Tg =Tm [22], which is also a GFA
index, is determined to be 0.620.63, indicating a high GFA of the present alloys.
The origin of the high GFA for the FeCrMoCB alloys is discussed in the
framework of the three empirical rules for the achievement of high GFA [10,23], i.e.,
(1) multicomponent alloy systems consisting of more than three elements, (2) sig-
nicant atomic size mismatch above 12% among main constituent elements, and (3)
suitable negative heats of mixing among their main constituent elements. When the
three empirical rules are applied to the present alloy system, the alloys are also
composed of three types of elements of a late transition metal (LTM Fe), early
1850 S.J. Pang et al. / Corrosion Science 44 (2002) 18471856

Fig. 1. DSC curves of the cast glassy Fe50x Cr16 Mo16 C18 Bx rods with a diameter of 1.2 mm.

Table 1
Thermal stability of the bulk glassy FeCrMoCB alloys with a diameter of 1.2 mm
Alloy Tg (K) Tx (K) DTx (K) Tm (K) Tg =Tm
Fe46 Cr16 Mo16 C18 B4 862 915 53 1389 0.62
Fe44 Cr16 Mo16 C18 B6 870 932 62 1414 0.62
Fe42 Cr16 Mo16 C18 B8 887 947 60 1405 0.63

transition metals (ETM Cr, Mo) and metalloids (M C, B) and the atomic size
of the constituent elements decreases in the order Mo > Cr > Fe > B > C [24].
The heats of mixing are negative for the atomic pairs of ETMLTM, ETMB and
LTMB and the negative values are in the range of 119 kJ mol1 [25]. It is regarded
that the present alloys satisfy the three empirical rules for the stabilization of the
supercooled liquid. It has previously been reported that the alloy liquid satisfying the
three empirical rules can have a new liquid structure with highly dense random
packed atomic congurations, new local atomic congurations and long-range ho-
mogeneous atomic congurations with strong interaction [10,23], leading to the
suppression of nucleation and growth reactions of a crystalline phase, a higher
viscosity and a lower atomic diusivity. Therefore, these FeCrMoCB alloys with
the low nucleation and growth reactions and low atomic diusivity can have a low
melting temperature leading to high Tg =Tm values above 0.62 and the large DTx above
53 K, which enable us to form the bulk glassy alloys.
Fig. 2 shows the corrosion rates of Fe50x Cr16 Mo16 C18 Bx (x 4, 6, 8 at.%) bulk
metallic glasses estimated by immersion tests in 1, 6 and 12 N HCl solutions for one
week at 298 K open to air. It is indicated that these bulk glassy alloys had signi-
cantly low corrosion rate of 103 102 mm year1 in 1, 6 and 12 N HCl solutions at
S.J. Pang et al. / Corrosion Science 44 (2002) 18471856 1851

Fig. 2. Corrosion rates of the cast glassy Fe50x Cr16 Mo16 C18 Bx rods with a diameter of 1.2 mm in HCl
solutions open to air at 298 K.

room temperature, and their corrosion rates exhibited a decreasing tendency with a
increase of boron content in the alloys. In addition, these bulk glassy alloys main-
tained metallic luster after the immersion tests, indicating their high corrosion re-
sistance.
Fig. 3 shows the SEM micrographs of the surfaces of the glassy Fe50x Cr16 -
Mo16 C18 Bx (x 4, 6, 8 at.%) rods of /1.2 mm after the immersion test in 12 N HCl
solution. The bulk glassy rods containing 6 and 8 at.% boron do not suer pitting

Fig. 3. SEM micrographs of the surfaces of the cast glassy Fe50x Cr16 Mo16 C18 Bx rods with a diameter of
1.2 mm immersed in 12 N HCl solution for 168 h at 298 K.
1852 S.J. Pang et al. / Corrosion Science 44 (2002) 18471856

corrosion because no pit is observed on the surfaces of the samples after the im-
mersion test. The pit-like contrasts on the alloy with 8 at.% B are due to cavities
formed during casting. The surface of the alloy with 6 at.% B is roughened by general
corrosion. On the other hand, on the surface of the Fe46 Cr16 Mo16 C18 B4 glassy rod,
many pits are observed in places after the long period immersion as shown in Fig.
3(a), indicating its lower corrosion resistance compared to that of the alloys con-
taining more boron, which is in agreement with the corrosion rates of these bulk
metallic glasses.
The electrochemical measurements in HCl solutions also indicate that bulk glassy
Fe50x Cr16 Mo16 C18 Bx (x 4, 6, 8 at.%) alloys with a diameter of 1.2 mm exhibit high
corrosion resistance as shown in Fig. 4, where anodic and cathodic polarization
curves measured in 1 N HCl solution at room temperature for these bulk glassy
alloys are plotted. It can be seen that they exhibit similar polarization behavior to
each other and are spontaneously passivated with low passive current density of the
order 102 A m2 . Wide passive regions are recognized until the transpassive dis-
solution of chromium occurs. Moreover, the increase of boron content in the glassy
alloy results in a decrease in the passive current density, leading to an improvement
of corrosion resistance. The bulk glassy alloys are spontaneously passivated also in 6
N HCl solution as shown in Fig. 5, though the passive current density is higher and
the passive region is smaller than that in 1 N HCl solution. In addition, it is also seen
that the alloys containing a larger amount of boron show lower current density and
hence higher corrosion resistance in 6 N solution. In 12 N HCl solution, the surface
lms of the bulk glassy alloys are not stable by anodic polarization, as can be seen in
Fig. 6, the anodic current density increases rapidly with the increase of potential.
However these bulk glassy alloys do not suer pitting corrosion in 12 N HCl solu-
tion.

Fig. 4. Potentiodynamic polarization curves of the cast glassy Fe50x Cr16 Mo16 C18 Bx rods with a diameter
of 1.2 mm in 1 N HCl solution open to air at 298 K.
S.J. Pang et al. / Corrosion Science 44 (2002) 18471856 1853

Fig. 5. Potentiodynamic polarization curves of the cast glassy Fe50x Cr16 Mo16 C18 Bx rods with a diameter
of 1.2 mm in 6 N HCl solution open to air at 298 K.

Fig. 6. Potentiodynamic polarization curves of the cast glassy Fe50x Cr16 Mo16 C18 Bx rods with a diameter
of 1.2 mm in 12 N HCl solution open to air at 298 K.

Since the high corrosion resistance is based on passive lms formed in aggressive
hydrochloric acids, it is necessary to clarify the origin of passivity by analyzing the
passive lms formed in HCl solutions and the surface lms before immersion. The
XPS spectra over a wide binding energy region apparently exhibited peaks of carbon,
oxygen, iron, chromium and molybdenum. All the spectra from the alloy constitu-
ents consisted of peaks corresponding to the oxidized state (ox) in surface lm and
1854 S.J. Pang et al. / Corrosion Science 44 (2002) 18471856

the metallic state (m) in the underlying alloy surface. The Fe spectrum was composed
of peaks of iron species in metallic, Fe2 and Fe3 states. The Cr spectrum consisted
of peaks of chromium species in metallic and Cr3 states. The Mo spectrum was
deconvoluted into peaks corresponding to metallic, Mo4 , Mo5 and Mo6 states.
The C 1s peaks were those from carbon in alloy, and so-called contaminant carbon
on the top surface of the specimen. The O 1s spectrum consisted of peaks originating
from oxygen in metalOmetal bond, metalOH bond and/or bound water.
The composition of passive lm gives important information about the stability
and protectiveness of passive lm. Therefore, change in the surface composition of
the FeCrMoCB metallic glasses by immersion in a solution was analyzed to
clarify the origin of their high corrosion resistance. The integrated peak intensities of
the C 1s from contaminant and alloy, O 1s, Fe 2p3=2 in oxidic and metallic states,
Cr 2p3=2 in oxidic and metallic states, Mo 3d in oxidic and metallic states, B 1s in
oxidic and metallic states and Cl 2p peaks were estimated from the measured spectra
by deconvolution and curve tting. The metallic state peaks were assumed to come
from the underlying alloy surface, the oxidized state peaks, from the surface lms.
The contaminant carbon is assumed to cover the top surface of the specimen.
Fig. 7 shows the atomic fractions in the underlying substrate alloy surface and the
cationic fractions in the surface lm as a function of alloying boron content for the
cast glassy Fe50x Cr16 Mo16 C18 Bx alloys exposed to air after mechanical polishing.
Iron is rich in the air-formed lm with respect to alloy composition, while the reverse
is seen for the underlying alloy surface. Accordingly, air exposure of the metallic
glasses results in preferential oxidation of iron, which is often observed for initial
oxidation of Fe-based alloys. On the other hand, chromium concentration in surface
lms increases with the increase of boron content in the alloys, suggesting that the
existence of boron promotes diusion of Cr or depresses that of Fe, and gives eect
on improvement of corrosion resistance with increasing boron content.
The thickness of the surface lms formed on the bulk FeCrMoCB metallic
glasses before and after immersion in 1 N HCl solution at 298 K was also analyzed.

Fig. 7. The change in the atomic fractions in the underlying substrate alloy surface and the cationic
fractions in the surface lm as a function of alloying boron content for the cast glassy Fe50x Cr16 Mo16 -
C18 Bx alloys exposed to air after mechanical polishing.
S.J. Pang et al. / Corrosion Science 44 (2002) 18471856 1855

Fig. 8. The change in the atomic fractions in the underlying substrate alloy surface and the cationic
fractions in the surface lm as a function of alloying boron content for the cast glassy Fe50x Cr16 Mo16 -
C18 Bx alloys immersed in 1 N HCl solution at 298 K for 168 h.

As a result, the thickness of the passive lms formed on the metallic glasses after
immersion in the acid ranged from 20 to 25 nm, which was almost the same as those
of the air-formed lms of the alloys. However, it does not simply mean that the
surface lm formed during air-exposure is responsible for the high corrosion resis-
tance. Actually, the compositions of the surface lms on the specimens are largely
changed as shown in Fig. 8.
Fig. 8 shows the atomic fractions in the underlying substrate alloy surface and the
cationic fractions in the surface lm as a function of alloying boron content for the
cast glassy Fe50x Cr16 Mo16 C18 Bx alloys immersed in 1 N HCl solution at 298 K for
168 h. Comparing these results shown in Fig. 8 to those of the as-prepared specimens
in Fig. 7, it can be seen that chromium concentration in the surface lms formed
on the FeCrMoCB metallic glasses after immersion in 1 N HCl solution is
apparently higher than that of the air-formed lms. Chromium-enriched passive
lms should be formed on the FeCrMoCB metallic glasses during immersion in
1 N HCl solution due to the reforming of air-formed lms by preferential dissolution
of iron. Accordingly, the corrosion resistance of these metallic glasses is essentially
originated from the same mechanism as that reported for melt-spun amorphous Fe
Cr-based alloys with similar composition [26]. In addition, chromium concentration
in both surface lms and underlying alloy surfaces for the FeCrMoCB metallic
glasses after immersion in 1 N HCl solution shows a signicant tendency to increase
with the increase of boron content in the alloys, which promotes improvement of
corrosion resistance of these metallic glasses.

4. Conclusions

Fe-based bulk glassy alloys with high GFA and good corrosion resistance were
found in the Fe50x Cr16 Mo16 C18 Bx (x 4, 6, 8 at.%) system, which satises the three
1856 S.J. Pang et al. / Corrosion Science 44 (2002) 18471856

empirical rules for the achievement of high GFA. The DTx is as large as 5362 K and
the Tg =Tm is as high as 0.620.63. The corrosion rates of the bulk glassy Fe50x Cr16 -
Mo16 C18 Bx alloys with a diameter of 1.2 mm are in the range of 103 102 mm year1
in 1, 6 and 12 N HCl solutions at room temperature. The bulk glassy alloys are
spontaneously passivated in 1 and 6 N HCl solutions with wide passive region and
low passive current density. They do not suer pitting corrosion even by anodic
polarization in 12 N HCl solution. The high corrosion resistance is due to the for-
mation of chromium-rich passive lms during immersion in HCl solutions. In ad-
dition, the increase of boron content in alloys improves the corrosion resistance of
the bulk glassy alloys within the composition range examined.

References

[1] M. Naka, K. Hashimoto, T. Masumoto, J. Jpn. Inst. Metals 38 (1974) 835.


[2] H. Yoshioka, K. Asami, A. Kawashima, K. Hashimoto, Corros. Sci. 27 (1987) 981.
[3] K. Asami, K. Kawashima, K. Hashimoto, Mater. Sci. Eng. 99 (1988) 475.
[4] D.B. Lee, H. Mitsui, H. Habazaki, A. Kawashima, K. Hashimoto, Corros. Sci. 38 (1996) 2031.
[5] K. Hashimoto, H. Habazaki, E. Akiyama, H. Yoshioka, J.H. Kim, P.Y. Park, A. Kawashima, K.
Asami, Sci. Rep. RITU A42 (1996) 99.
[6] A. Inoue, K. Ohtera, K. Kita, T. Masumoto, Jpn. J. Appl. Phys. 27 (1988) L2248.
[7] A. Inoue, T. Zhang, T. Masumoto, Mater. Trans. JIM 30 (1989) 965.
[8] T. Zhang, A. Inoue, T. Masumoto, Mater. Trans. JIM 32 (1991) 1005.
[9] A. Peker, W.L. Johnson, Appl. Phys. Lett. 63 (1993) 2342.
[10] A. Inoue, Bulk Amorphous Alloys, Trans. Technol. Pub., Zurich, 1998, pp. 136.
[11] A. Inoue, Y. Shinohara, J.S. Gook, Mater. Trans. JIM 36 (1995) 1427.
[12] A. Inoue, T. Zhang, A. Takeuchi, Appl. Phys. Lett. 71 (1997) 464.
[13] A. Inoue, T. Zhang, H. Koshiba, T. Itoi, Mat. Res. Soc. Symp. Proc. 554 (1999) 251.
[14] X. Wang, I. Yoshii, A. Inoue, Mater. Trans. JIM 41 (2000) 539.
[15] S.J. Pang, T. Zhang, H.M. Kimura, K. Asami, A. Inoue, Mater. Trans. JIM 41 (2000) 1490.
[16] A. Inoue, J.S. Gook, Mater. Trans. JIM 37 (1996) 32.
[17] S.J. Pang, T. Zhang, K. Asami, A. Inoue, Mater. Trans. JIM 42 (2001) 376.
[18] K. Asami, K. Hashimoto, S. Shimodaira, Corros. Sci. 17 (1977) 713.
[19] K. Hashimoto, K. Asami, Corros. Eng. (Boshoku Gijutsu) 26 (1977) 375.
[20] K. Asami, J. Electron Spectrosc. 9 (1976) 469.
[21] K. Asami, K. Hashimoto, Corros. Sci. 17 (1977) 559.
[22] D. Turnbull, J.C. Fisher, J. Chem. Phys. 17 (1949) 71.
[23] A. Inoue, Mater. Trans. JIM 36 (1995) 866.
[24] Metals Databook, edited by Japan Inst. Metals, Maruzen, Tokyo, 1993, p. 8.
[25] F.R. Boer, R. Boom, W.C.M. Mattens, A.R. Miedema, A.K. Niessen, Cohesion in Metals, North-
Holland, Amsterdam, 1989, pp. 103637.
[26] K. Asami, K. Hashimoto, T. Masumoto, S. Shimodaira, Corros. Sci. 16 (1976) 387.

Вам также может понравиться