Anda di halaman 1dari 7

Corrosion Science 52 (2010) 118–124

Contents lists available at ScienceDirect

Corrosion Science
journal homepage: www.elsevier.com/locate/corsci

Corrosion of nickel–chromium alloys, stainless steel and niobium


at supercritical water oxidation conditions
Edouard Asselin a,*, Akram Alfantazi a, Steven Rogak b
a
Department of Materials Engineering, The University of British Columbia, Vancouver, BC, Canada
b
Department of Mechanical Engineering, The University of British Columbia, Vancouver, BC, Canada

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

Article history: Results of immersion tests of UNS N06625 (alloy 625), UNS S31609 (alloy 316 L), Ni–20Cr alloy and
Received 22 June 2009 Nb coupons exposed to oxygenated ammoniacal sulphate solution at supercritical water oxidation (SCWO)
Accepted 26 August 2009 conditions are presented. The corrosion behavior of the alloy 316 L (UNS S31603) SCWO reactor tubing is
Available online 31 August 2009
also presented under the same conditions. Immersion coupons corroded at a rate not exceeding 40 mm yr1
while the reactor tube itself corroded at a rate of between 160–1500 mm yr1, depending on which length of
Keywords: the reactor is considered to have corroded. Morphological and chemical analysis of the oxides present on the
Corrosion
coupon samples suggest that iron oxides, which had initially precipitated on the corroding alloy 316 tube
Ni–Cr alloy
Stainless steel
surface, were removed by heat flux-driven fluid mechanical action and transported through the reactor
High temperature where they deposited on the coupons. Niobium was resistant to corrosion at the tested conditions.
Supercritical water Ó 2009 Elsevier Ltd. All rights reserved.
Immersion testing

1. Introduction was supplied by two high pressure piston pumps and heated
throughout most of its length by cylindrical ceramic heaters at a
During the supercritical water oxidation (SCWO – T > 374 °C, heat flux of 95 kW m2. The system was fitted with an in-line test
P > 22 MPa) [1] of ammoniacal sulphate solution at pH298K  10, a cell which housed two weight loss coupons of each type of alloy/
failure occurred in the first pre-heater at the University of British metal. The alloy 625 and Ni–20Cr samples were commercially ob-
Columbia’s Inconel alloy 625 (22 Cr, 10 Mo, 5 Fe, 5 Nb, Bal. Ni) SCWO tained as rod stock while the alloy 316 sponge was purchased as a
reactor [2]. On a later cold static pressure test at 50 MPa the tube side disc. The alloy 316 was chosen as a sponge to attempt to differen-
of the recuperative heat exchanger burst. A subsequent failure anal- tiate between oxide deposited on the sample from the 316 reactor
ysis identified that the reactor had likely failed due to acidification of and oxide naturally grown on the coupon from exposure to solu-
the feed solution through evaporative loss of ammonia at near- tion. The Nb (>99.9% pure) samples were supplied from a specialty
supercritical temperatures [3]. This temperature/pressure regime, manufacturer. Three coupons of alloy 625 were immersed in the
commonly referred to as the ‘‘dense supercritical region”, lies reactor while two coupons each of the Nb, alloy 316 and Ni–20Cr
between approximately 380–400 °C at 22–25 MPa [4]. alloy were tested. The corrosion rate data presented below repre-
To determine the susceptibility of some selected alloy systems to sents average values. A schematic of the experimental apparatus
the above mentioned conditions, a set of immersion experiments is provided in Fig. 1.
were undertaken in a small scale 316 L SCW reactor. Coupons sam-
ples of alloys 625, 316, Ni–20Cr and Nb were immersed in the reactor. 2.2. Procedures
The following paper presents the results of these immersion tests.
The test solution used for this experiment was a 1.6 m (molal:
2. Experimental mol kgH2O1) total NH3, 0.2 m SO42, pH 10 solution (2.4 wt.%
NH3 and 2 wt.% H2SO4). This solution is consistent with the inor-
2.1. Apparatus ganic concentrations in the feed which led to the SCWO reactor
failure described in the introduction. The flow rate of the ammoni-
The test reactor consisted of a 5 m length of seamless 1=4 ” outer acal solution was 15 mL/min. Oxidant was delivered in the form of
diameter (6.35 mm) and 1.2 mm wall thickness 316 L tubing which hydrogen peroxide (H2O2). Homogeneous peroxide decomposition
occurs readily in ammoniacal alkaline solutions especially as tem-
* Corresponding author. Address: Department of Materials Engineering, The
University of British Columbia, 309-6350 Stores Road, Vancouver, BC, Canada V6T
perature is increased – the half life of hydrogen peroxide is approx-
1Z4. Tel.: +1 604 822 1918; fax: +1 604 822 3619. imately 9 min in 1.9 M NH3 at 80 °C [5]. Heterogeneous catalysis of
E-mail address: edouard@mtrl.ubc.ca (E. Asselin). hydrogen peroxide decomposition on Fe oxides is also known to

0010-938X/$ - see front matter Ó 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.corsci.2009.08.053
E. Asselin et al. / Corrosion Science 52 (2010) 118–124 119

Peroxide Pulsation
pump damper Ceramic cylindrical heaters 1,2,3,4

1 2
Pressure
Test solution P
relief valve
pump

Test cell 4 3

Gas/Liquid
Separator

Double pipe heat exchanger P


Back
pressure
regulator

Fig. 1. Schematic of bench-top SCWO apparatus.

inder of the high pressure piston pump. In fact, the homogeneous


decomposition reaction was so rapid in the alkaline solution that,
even without the catalytic metallic surface contact, the thermo-
Glass sample
holder plastic pump feed lines were also clearly full of gas. For this reason,
a second pump was required to inject the 19 wt.% peroxide solu-
Sample Sample tion at a rate of 5 mL min1. The total combined solution flowrate
Solution flow
through the reactor was therefore 20 mL min1 containing 4.8 wt.%
Cell wall H2O2. This yielded an estimated oxygen concentration of 2–3 m
(depending on density) in the supercritical solution running
through the test cell. This concentration is similar to the one em-
ployed in the larger SCWO reactor and which led to failure (4 m).
Fig. 2. Sectional schematic view of test cell coupon assembly. The test cell (Figs. 1 and 2), which was not heated, consisted of an
alloy 316 L cylindrical vessel mounted in-line with the direction of
occur at 30 °C in neutral solutions [6]. Similarly most transition solution flow. The solution was made to pass through tubular
metals and chromium are known to increase the rate of both boro-silicate glass sample holders that were designed to stack hori-
homogeneous and heterogeneous peroxide decomposition [7,8]. zontally in the test cell. The sample holders had an inlet diameter
Clearly, since the pump internals were made of stainless steel which was larger than their outlet diameter. Test coupons were rect-
and since the reactor tubing was stainless steel we expected a high angular and approximately 1 cm  0.5 cm  1.5 mm so as to fit in-
rate of peroxide decomposition. This reaction occurs through dis- side the larger diameter section of the glass holders, but also so as
proportionation according to Eq. (1) [5]. not to fit through the smaller diameter section. A schematic cross-
sectional view of the test cell is given in Fig. 2. The glass holders were
2H2 O2 ¼ O2 þ 2H2 O ð1Þ
necessary to remove any possibility of galvanic interaction. The
Initially experiments were run with only one pump such that tested coupons were ground to 600 SiC, ultrasonically cleaned in eth-
the ammoniacal solution and peroxide were mixed on the low anol, dried and weighed prior to being put in the test cell.
pressure side of the feed pump. However, complete loss of head oc- The bench-top system was brought to temperature with dis-
curred within 5 min due to compressible oxygen gas filling the cyl- tilled water over a period of approximately 1 h such that the tem-

550
Heater Bank 1 Heater Bank 2 Heater Bank 3 Heater Bank 4
500 Un-heated
Bend
450 Test
Cell
400 Dense supercritical region
Temperature, °C

350
300

250
200

150
100
50
0
0 50 100 150 200 250 300 350 400 450 500
Distance from inlet to first heater, cm

Fig. 3. Temperature profile in SCWO reactor and test cell.


120 E. Asselin et al. / Corrosion Science 52 (2010) 118–124

perature at the exit of heater bank 4 was  390 °C. The tempera- Table 1
ture profile was also arranged so that high-density supercritical Concentrations of dissolved metal after SCWO experiment in 316 L tube.

fluid was obtained within the 316 L reactor tubing in banks 2 Mo Ni Fe Cr Nb


and 4 as well as in the test cell. The obtained temperature profile mg L 1
1.28 ± 0.01 1.18 ± 0.03 0.50 ± 0.11 0.17 ± 0.05 <0.05
is shown in Fig. 3.
Temperature was measured by thermocouples on the outside of
the tubing and the whole system was heavily insulated with a
is well known that the corrosion of stainless steels and nickel–
thermal ceramic. The temperature inside the tube was roughly
chromium alloys is significantly increased at sub- and supercritical
the same as that of the thermocouples. As the hot fluid exited hea-
temperatures by both the presence of heteroatoms such as sulphur
ter bank 4 it went directly into the test cell. Because the test cell
and chlorine and the presence of oxygen in large amounts [4,12]. In
was insulated but not heated, the incoming fluid was cooled as it
particular, the latter is linked to transpassive corrosion [13–15].
passed the samples. The test cell was thermocoupled in six loca-
For this reason a small increase in the amount of oxidant available
tions and temperatures measured on the surface of the cell and
can significantly impact the corrosion rate and it seems reasonable
at the solution exit did not go below 380 °C. Thus the fluid temper-
to conclude that the corrosion rates observed in this work are
ature in the test cell was between 390 and 380 °C which is exactly
linked to transpassive corrosion as demonstrated in the case of al-
the temperature range of interest and that which led to the failure
loy 625 in the same solution as used here [16].
in the larger reactor.
The dissolved metal concentrations from the filtrate of the
After the 1 h heat-up period, the influent feed stream was
experiment effluent are presented in Table 1. The order of solubil-
switched from distilled water to the ammoniacal sulphate solution
ity was Mo > Ni > Fe > Cr > Nb. This is similar to the order for the
and hydrogen peroxide mixture. Within approximately 10 min,
effluent concentrations collected from the alloy 625 tubing failure
oxygen bubbles were visible in the gas/liquid separator (made
which were: Mo > Ni > Cr [2]. The high rate of Ni and Mo dissolu-
from transparent acrylic). Approximately 35 min later the reactor
tion (mostly as compared to Fe) is surprising given that the alloy
effluent became dark brown in colour and contained suspended
concentrations are so minor: for 316 L Ni  12 wt.%, Mo  2.5 wt.%.
solids. The experiment was terminated and the system was flushed
As discussed earlier, the effluent solution composition is necessar-
with distilled water and allowed to cool for 40 min. The effluent
ily a reflection of the room temperature at which it was collected
was collected and filtered to remove suspended solids. The filtrate
and analyzed and it should be emphasized that precipitation of
was analyzed with inductively coupled plasma atomic emission
spectrometry to determine dissolved metal concentrations. The fil-
ter cake was dried and weighed.

3. Results and discussion

3.1. Reactor tubing

Based on the dried filter cake material it was found that at least
11.1 g of solids had been corroded and removed from the reactor.
The solids were analyzed by energy dispersive X-ray spectroscopy
(EDX) and found to contain approximately 33 wt.% O2, 26 wt.% Fe,
22 wt.% Cr, 8 wt.% Ni, 5 wt.% S and 4.3 wt.% Mo. By removing the
mass associated with oxygen and sulphur it can be roughly deter-
mined that 6.9 g of the reactor’s alloy 316 L had been corroded in
the 45 min time period for which it was exposed to the ammonia-
cal solution. Assuming an even corrosion rate distribution across
the entire internal surface of the 5 m tube this weight loss would
imply a corrosion rate of approximately 160 mm yr1 or complete
tube dissolution in roughly 65 h. If one considers only the 0.55 m
length of tube which was exposed to the high-density supercritical
solution at steady state then the corrosion rate becomes
1500 mm yr1, or wall thickness penetration in 8 h. These calcula-
tions demonstrate that the observed corrosion rate is obviously
completely unacceptable. Hayward et al. have presented results
on the corrosion of alloy 316 L SCWO reactor tubing in water con-
taining 3 wt.% hydrogen peroxide at 375 °C. Based on solution
analysis alone these authors obtained a maximum corrosion rate
of 3 lm yr1: several orders of magnitude less than observed here
[9]. However, it is acknowledged in the work by Hayward et al. that
the corrosion rates they report are significantly lower than those
observed by others. It should also be noted that solubility clearly
changes with temperature such that measurements of dissolved
metal concentrations from room temperature effluents may tend
to give an erroneous corrosion rate. For example, Boukis et al. have
reported corrosion rates of 24 mm yr1 for alloy 316 in oxygen-free
Fig. 4. (a) Morphology of porous scale before cleaning and (b) high magnification
0.05 m HCl [10]. Son et al. have reported a corrosion rate of approx- morphology of deposits on the surface of coupons exposed to 380–390 °C SCWO
imately 20–30 mm yr1 for alloy 316 in 5 wt.% NaCl and 1 wt.% environment. White square represents location of EDX spectra reveal that the
H2O2 at 350–400 °C based on mass change measurements [11]. It porous scale is mostly Fe oxide and Cr oxide removed from the 316 L reactor tube.
E. Asselin et al. / Corrosion Science 52 (2010) 118–124 121

metal oxides would have occurred throughout the reactor, partic- Table 2
ularly in those areas subjected to low density supercritical fluid Coupon corrosion rates after exposure to ammoniacal sulphate solution at 380–
390 °C.
and in the heat exchanger where the fluid was cooled [4]. The
11.1 g of metal oxide filtered from the effluent is clear evidence Mass change before Mass change after Corrosion rate
of this occurrence. Sue et al. have reported a slightly different order de-scaling (%) de-scaling (%) (mm.yr1)

of solubility based on both theoretical calculations and experimen- Alloy 625 0.10 0.13 18
tal measurement: Fe > Ni > Cr [17]. The increased solubility of Ni Ni/20Cr 0.07 0.27 29
Alloy 316 1.40 0.78 39
(with respect to Fe) in the case of our experiments is almost cer- Nb 0.20 0.12
tainly due to the stability of ammine complexes in the effluent. It
is well known that ammonia is relatively stable under supercritical
water conditions particularly at temperatures less than 500 °C
sample holders had expanded within the cell and fused themselves
[18,19]. However, the effluent total ammonia and total sulphur
to its wall. Also, it was observed that the glass had either frag-
concentrations were also measured and found to be approximately
mented under differential expansion/contraction pressure or re-
one quarter of the influent concentrations. Sulfur was obviously
acted with the solution to form a compact and sandy powder. To
adsorbed or co-deposited with the metal oxides as indicated by
extract the coupons, the test cell was carefully face milled until
the filter cake composition which would account for its decrease
the exposed glass holders were visible. The sand-like product
in concentration. The loss of ammonia throughout the reactor
was then removed with an acid brush and the samples were
likely occurred by degassing in the gas liquid separator and by
collected.
partial conversion to molecular nitrogen. Nevertheless, the remain-
The exposed coupons were all covered with a black scale
ing ammonia concentration (4200 mg L1) was more than enough
(Fig. 4). The scale morphology appears to be the result of an
to stabilize nickel ammine complexes [3,20,21]. This order of
agglomeration of sub-micron sized spherical deposits. These
solubility and propensity for corrosion has also been measured
deposits were analyzed with EDX and found to be of the same com-
electrochemically in a previous investigation on the corrosion of
position as the dried solids filtered out from the effluent i.e. origi-
Ni–Cr alloys in ammoniacal solutions [21].
nating from the reactor. There are two possible mechanisms for the
precipitation of such particulates: (1) the metals were corroded
3.2. Immersion tests from the reactor and precipitated from the bulk both homoge-
neously and heterogeneously or (2) the particulate was formed di-
The alloy 625, Ni–20Cr, 316 and Nb samples were removed rectly at the corroding tube/solution interface and was carried
from the test cell with great difficulty. It was found that the glass through the reactor by fluid motion. In the latter proposition the

Fig. 5. Surface of corrosion coupons after de-scaling (a) Alloy 625 (b) Nb (c) Alloy 316 sponge and (d) Ni–20Cr alloy.
122 E. Asselin et al. / Corrosion Science 52 (2010) 118–124

Fig. 6. Cross sections of corrosion coupons after de-scaling (a) Alloy 625 (b) Nb (c) Alloy 316 sponge and (d) Ni–20Cr alloy.

Table 3 Alloy 625 was the most resistant alloy (18 mm yr1) while alloy
Composition of oxides (EDS) on de-scaled coupons exposed to SCWO conditions. 316 had the highest corrosion rate (39 mm yr1). The corrosion
Composition Alloy Nb Alloy 316 Alloy 316 Ni– rate generally decreased with increasing chromium content but
(wt.%) 625 surface inside 20Cr the difference between alloy compositions is significant enough
Fe 47 30 46 41 4 that other factors such as minor element contributions cannot be
Cr 5 27 24 20 overlooked. Nevertheless, Shroer et al. have also concluded that
Ni 10 6 6 69 binary Ni–Cr alloy corrosion decreases with increasing Cr content
Mo 4 4 4 at 390 °C in SCWO systems [23]. The corrosion rates observed by
Nb 1 50
Shroer et al. were between 20 and 40 mm yr1 for binary Ni–Cr al-
O 33 20 17 25 7
loys in 0.12 m HCl/0.06 m O2 at 400 bar, 390 °C. The coupon corro-
sion rates observed in our work are therefore consistent with those
presented in three previous studies on Ni–Cr alloys and alloy 316
implication is that oxides which form at the corroding tube/solu- [10,11,23]. However, a recent study by Sun et al. found only mass
tion interface are not adherent or compact as would be expected gain for 316 coupons immersed in 2.0 wt.% H2O2 supercritical
of oxides which are formed at lower temperatures. This view is water at 350 °C [24]. This difference in results is likely due to the
consistent with the interpretation put forward in our earlier work absence of any heteroatoms in their work. One of the most inter-
which demonstrated that the inner chromium oxide on the surface esting findings of this work is the large disagreement between
of alloy 625 (assuming a bi-layered oxide structure [22]) was the corrosion rates observed in the alloy 316 L reactor tubing
transpassively dissolved at temperatures exceeding 100 °C and (160–1500 mm yr1) and those observed on the alloy 316 cou-
that the outer layer which remained was porous and highly defec- pons (40 mm yr1). There are several possible reasons for this
tive [16]. discrepancy, all of which are likely responsible to some extent
The immersion coupons were first weighed after the experi- for our findings. Firstly, the corrosion products from the reactor
ment without further treatment. They were then subsequently tubing may have protected the underlying coupons. Deposition of
de-scaled ultrasonically for 15 min in a solution containing 8.8 g iron oxides is well known for protecting underlying steel in high
of 2-butyne-1,4-diol, 30 mL of reagent HCl and 100 mL of distilled temperature water where it has been demonstrated that inward
water and then re-weighed to determine the mass gain/loss with- (towards metal surface) diffusion of oxygen or water through the
out the deposited scales from the reactor. The results of the mass outer oxide layer is rate limiting [16,25]. Secondly, the high rate
measurements are presented in Table 2. All three alloys showed of reactor tubing corrosion may have substantially reduced the
a mass loss after de-scaling indicating a positive corrosion rate. available oxygen for coupon corrosion downstream (Fig. 1). Finally,
E. Asselin et al. / Corrosion Science 52 (2010) 118–124 123

it is expected that the locations of high corrosion in the reactor roded at a rate of between 160 and 1500 mm yr1. The high rate
tubing (380–400 °C) were exposed to some degree of fluid of 316 L tubing corrosion (as compared to coupons) is thought to
mechanical effects due to heat flux-driven density differences be- be due to fluid mechanical effects associated with buoyancy
tween the solution at the tube wall (low density) and the solution changes in the dense supercritical fluid region. It is hypothesized
at the centre of the tube (high-density). Indeed, supercritical water that enough turbulence was caused in this region to dislodge any
at 25 MPa changes from a density of 600 kg m3 at 370 °C to loosely adherent oxide particles from the tube surface. Much of
200 kg m3 at 390 °C [26]. This could have induced enough fluid the oxide present on the surfaces of the alloy coupons located
motion to dislodge oxide particulate as hypothesized above. downstream of the reactor could be attributed, through composi-
The rate of corrosion of the niobium coupons is low – these cou- tion, to the particulate released from the corroding tube. Niobium
pons showed a net increase in weight. Similar results have been coupons were not attacked under the conditions tested here.
observed by Kritzer et al. for niobium exposed to various solutions
at 350 °C, 24 MPa [27]. We can conclude, as did Kritzer et al., that Acknowledgements
niobium may be an adequate material for the severe service condi-
tions encountered in SCWO reactors in the subcritical and dense The authors acknowledge the Natural Sciences and Engineering
supercritical temperature regime. Research Council of Canada, the Canadian Foundation for Innova-
The surfaces of the corrosion coupons after scale removal are tion and the University of British Columbia for financial support.
shown in Fig. 5. These SEM micrographs clearly demonstrate that Mr. Ross Mcleod is acknowledged for his considerable assistance
not all of the particulate was removed through ultrasonication in related to experimental apparatus design.
the inhibited HCl solution. The spherical nature of the deposited
particulate (originating from the tube corrosion) is shown on the References
surface of the Nb sample in Fig. 5(b). Larger agglomerates appear
to have formed on the Ni–20Cr alloy shown in Fig. 5(d). [1] H. Schmieder, J. Abeln, Supercritical water oxidation: state of the art, Chemical
Engineering and Technology 22 (1999) 903–908.
The chemical composition of the oxides/scales as obtained [2] I. Vera-Perez, Supercritical water oxidation of phenol and 2,4 dinitrophenol,
through EDX analyses of cross sections (Fig. 6) of the coupons are Masters Thesis, The University of British Columbia, 2002.
presented in Table 3. Generally the oxide/scales were between 2 [3] E. Asselin, A. Alfantazi, S. Rogak, Thermodynamics of the corrosion of alloy 625
supercritical water oxidation reactor tubing in ammoniacal sulphate solution,
and 4 lm thick. It is interesting to note that the oxygen concentra- Corrosion 64 (2008) 301–314.
tion at the metal/alloy surfaces decreased with stacking order. Al- [4] P. Kritzer, N. Boukis, E. Dinjus, Factors controlling corrosion in high-temperature
loy 625 coupons were located at the front of the immersion test aqueous solutions: a contribution to the dissociation and solubility data
influencing corrosion processes, The Journal of Supercritical Fluids 15 (1999)
cell, Nb coupons were second and so on. Similarly, the oxides are
205–227.
increasingly stoichiometrically deficient in oxygen. Assuming [5] J.E.A.M. van den Meerakker, M.H.M. van der Straaten, A mechanistic study of
Cr2O3, Fe2O3, NiO and Nb2O5, we get an excess of approximately silicon etching in NH3/H2O2 cleaning solutions, Journal of the Electrochemical
13 atoms of oxygen for the oxide formed on alloy 625 while there Society 137 (1990) 1239–1243.
[6] P. Baldrian, V. Merhautova, J. Gabriel, F. Nerud, P. Stopka, M. Hruby, M.J. Benes,
is a 36, 46 and 70 oxygen atom deficiency in the case of the Nb, al- Decolorization of synthetic dyes by hydrogen peroxide with heterogeneous
loy 316 and Ni–20Cr alloys, respectively. This indicates a lack of catalysis by mixed iron oxides, Applied Catalysis B: Environmental 66 (2006)
available oxygen near the end of the test cell which supports the 258–264.
[7] G.C. Hobbs, J. Abbot, Two-stage peroxide bleaching of eucalypt-SGW with
hypothesis that the rate of corrosion of the coupons was limited chromium catalysts, Journal of Wood Chemistry and Technology 11 (1991)
by a lack of oxygen (it having reacted up-stream). A similar conclu- 329–347.
sion may be drawn as regards the rate of iron oxide deposition on [8] J.T. Burton, L.L. Campbell, Hydrogen peroxide decomposition by metal
catalysts: bad actors in a bleaching stage, in: Proceedings of the 1985
the coupons. For example there is a large amount of iron oxide International Symposium on Wood and Pulping Chemistry, American
deposited on the surface of alloy 625 and Nb (wt.% Fe in Table 3) Chemical Society, Vancouver, BC, 1985, pp. 255–259.
while the amount of iron oxide deposited on the last coupons of [9] T.M. Hayward, I.M. Svishchev, R.C. Makhija, Stainless steel flow reactor for
supercritical water oxidation: corrosion tests, Journal of Supercritical Fluids 27
the Ni–20Cr alloy is significantly less. Of course, since the coupons (2003) 275–281.
were de-scaled prior to this EDX analysis one may also infer that [10] N. Boukis, C. Friedrich, W. Habicht, M. Schacht, E. Dinjus, Corrosion screening
the deposits could have been less adherent at the furthest locations tests in supercritical water containing hydrochloric acid and oxygen, in:
Proceedings of Eurocorr 1997, Event No. 208, European Federation of
due to temperature difference. Further work would be required to
Corrosion, Trondheim, Norway, vol. 1, 1997, pp. 617–622.
clarify this point. The relevance of this analysis is simply to indi- [11] M. Son, Y. Kurata, Y. Ikushima, Corrosion behavior of metals in SCW
cate that coupon testing is open to interpretation in SCWO reactors environments containing salts and oxygen, Corrosion, 2002, Paper No.
because the relative position of such coupons can dictate the mor- 02357, NACE International.
[12] P. Kritzer, E. Dinjus, An assessment of supercritical water oxidation (SCWO)
phology and chemical composition of the oxides which are formed existing problems, possible solutions and new reactor concepts, Chemical
or deposited on the coupon surface. The oxide located inside the Engineering Journal 83 (2001) 207–214.
pores of the alloy 316 sponge contains significantly more oxygen [13] P. Kritzer, N. Boukis, E. Dinjus, Transpassive dissolution of alloy 625,
chromium, nickel and molybdenum in high temperature solutions
than that measured on the surface (24 vs. 17 wt.%, Table 3). More- containing hydrochloric acid and oxygen, Corrosion 56 (2000) 265–272.
over, the oxide was clearly deposited within the pores of the [14] P. Kritzer, N. Boukis, E. Dinjus, Review of the corrosion of nickel-based alloys and
sponge microstructure (Fig. 6c) indicating that it came from the stainless steels in strongly oxidizing pressurized high-temperature solutions at
subcritical and supercritical temperatures, Corrosion 56 (2000) 1093–1104.
corroding reactor tube. During the first few minutes of the test [15] P. Kritzer, N. Boukis, E. Dinjus, Corrosion of alloy 625 in high-temperature, high
the alloy 316 L tube surface would have been exposed to a large pressure sulphate solutions, Corrosion 54 (1998) 689–699.
amount of unreacted oxygen and the deposits found downstream [16] E. Asselin, A. Alfantazi, S. Rogak, Effect of oxygen on the corrosion behavior of
alloy 625 from 25 to 200 degrees C, Journal of the Electrochemical Society 154
in the pores of the 316 sponge may be a reflection of this high ini-
(2007) C215–C229.
tial rate of oxidation. [17] K. Sue, N. Tsujinaka, T. Adschiri, K. Arai, Y. Watanabe, Relationship between
corrosion rate and metal oxide solubility in supercritical water, Corrosion,
2002, Paper No. 02354, NACE International.
4. Conclusions
[18] R.K. Helling, J.W. Tester, Oxidation of simple compounds and mixtures in
supercritical water: carbon monoxide, ammonia, and ethanol, Environmental
Alloys 625, 316 and Ni–20Cr experience unacceptably high cor- Science and Technology 22 (1988) 1319–1324.
rosion rates (20–40 mm yr1) when exposed to heavily oxygen- [19] P.A. Webley, J.W. Tester, H.R. Holgate, Oxidation kinetics of ammonia and
ammonia–methanol mixtures in supercritical water in the temperature range
ated ammoniacal sulphate solutions in the high-density 530–700 °C at 246 bar, Industrial and Engineering Chemistry Research 30
supercritical region (380–390 °C). Alloy 316 L reactor tubing cor- (1991) 1745–1754.
124 E. Asselin et al. / Corrosion Science 52 (2010) 118–124

[20] J. Bjerrum, Metal Ammine Formation in Aqueous Solution, P. Haase and Son, [24] M. Sun, X. Wu, Z. Zhang, E.H. Han, Oxidation of 316 stainless steel in
Copenhagen, 1957. supercritical water, Corrosion Science 51 (2009) 1069–1072.
[21] E. Asselin, A. Alfantazi, S. Rogak, Polarization Study of Alloy 625, Ni, Cr [25] L. Tomlinson, C.B. Ashmore, P.J.B. Silver, A.P. Mead, The high temperature
and Mo in Ammoniated Sulphate Solutions, Corrosion 61 (2005) 579– aqueous corrosion of ferritic steels: effect of heat flux on corrosion and
586. magnetite deposition, Corrosion Science 30 (1990) 511–535.
[22] D.D. Macdonald, The point defect model for the passive state, Journal of the [26] A.H. Harvey, A.P. Peskin, S.A. Klein, NIST Standard Reference Database 10, NIST/
Electrochemical Society 139 (1992) 3434–3449. ASME Steam Properties Database.
[23] C. Shroer, J. Konys, J. Novotny, J. Hausselt, Corrosion investigations on binary [27] P. Kritzer, N. Boukis, G. Franz, E. Dinjus, The corrosion of niobium in oxidizing
and ternary Ni–Cr alloys in SCWO systems, Corrosion, 2002, Paper No. 02359, sub- and supercritical aqueous solutions of HCl and H2SO4, Journal of Materials
NACE International. Science Letters 18 (1999) 25–27.

Anda mungkin juga menyukai