Vous êtes sur la page 1sur 9

Corrosion resistance of structural materials in high-

temperature aqueous sulfuric acids in thermochemical water-


splitting iodineesulfur process
Shinji Kubo
a,
*, Masatoshi Futakawa
a
, Ikuo Ioka
a
, Kaoru Onuki
a
, Akihisa Yamaguchi
b
a
Japan Atomic Energy Agency, 4002 Naritacho, Oarai, Higashiibaraki-gun, Ibaraki 311-1393, Japan
b
Chiyoda Advanced Solutions Corporation, Technowave 100 Bldg., 1-25 Shin-Urashima-Cho 1-chome, Kanagawa-ku,
Yokohama 221-0031, Japan
a r t i c l e i n f o
Article history:
Received 21 September 2012
Received in revised form
11 January 2013
Accepted 12 January 2013
Available online 15 April 2013
Keywords:
Thermochemical process
Sulfureiodine process
Sulfuric acid
Corrosion
Silicon carbide
High silicon iron
a b s t r a c t
Very harsh environments exist in the iodineesulfur process for hydrogen production.
Structural materials for sulfuric acid vaporizers and concentrators are exposed to high-
temperature corrosive environments. Immersion tests were carried out to evaluate the
corrosion resistance of ceramics and to evaluate corrosion-resistant metals exposed to
environments of aqueous sulfuric acids at temperatures of 320, 380, and 460

C, and
pressure of 2 MPa. The aqueous sulfuric acid concentrations for the temperatures were 75,
85, and 95 wt%, respectively. Ceramic specimens of silicon carbides (SiC), silicon-
impregnated silicon carbides (SieSiC), and silicon nitrides (Si
3
N
4
) showed excellent corro-
sion resistance from weight loss measurements after exposure to 75, 85, and 95 wt% sul-
furic acid. High-silicon irons with silicon content of 20 wt% showed a fair measure of
corrosion resistance. However, evidence of crack formation was detected via microscopy.
Silicon enriched steels severely suffered from uniform corrosion with a corrosion rate in
95 wt% sulfuric acid of approximately 1 g m
2
h
1
. Among the tested materials, the ce-
ramics SiC, SieSiC, and Si
3
N
4
were found to be suitable candidates for structural materials
in direct contact with the considered environments.
Copyright 2013, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights
reserved.
1. Introduction
Hydrogen production from water using nuclear energy offers
a very large and economical hydrogen supply, superior en-
ergy security, and eco-friendly operation [1]. Thermochemical
water-splitting processes are candidates for hydrogen pro-
duction using the energy from nuclear power plants [1,2].
Furthermore, they can directly harness the heat generated
by nuclear reactors. The three-step thermochemical water-
splitting process, that is, the iodineesulfur (IS) process [3]
(or sulfureiodine (SI) process), has been investigated as
a process to use the heat generated by high-temperature gas-
cooled reactors (HTGRs), since their temperature range of
300e900

C coincides with that of the sulfuric acid decom-
position, which is a strongly endothermic chemical reaction.
This process is competitive because 1) it involves a relatively
small number of chemical reactions; 2) it could be thermally
efcient; and 3) the chemical compounds used in the process
are uids (gaseous or liquid phases). However, corrosive
chemicals must be used in the IS process because of the
* Corresponding author. Tel.: 81 29 267 1919; fax: 81 29 266 7486.
E-mail address: kubo.shinji@jaea.go.jp (S. Kubo).
Available online at www.sciencedirect.com
j ournal homepage: www. el sevi er. com/ l ocat e/ he
i nt e r na t i o na l j our na l o f hy d r og e n e ne r g y 3 8 ( 2 0 1 3 ) 6 5 7 7 e6 5 8 5
0360-3199/$ e see front matter Copyright 2013, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ijhydene.2013.01.106
environments in which the chemical reactions and unit op-
erations are performed.
Our future goal is the industrialization of a massive
hydrogen production systemthat employs the IS process with
HTGRs. To achieve this, one of the key engineering tasks is the
screening and selection of structural materials for the chem-
ical reactors, which should be resistant to the corrosive en-
vironments in which the processes take place. The chemical
reactions and temperatures involved in the process are as
follows:
SO
2
g I
2
aq 2H
2
Oaq/2HIaq H
2
SO
4
aq; ca:100

C (1)
2HIaq/H
2
g I
2
g; ca:150 to 500

C (2)
H
2
SO
4
aq/H
2
Og SO
2
g 0:5O
2
g; ca:300 to 850

C (3)
where aq and g denote aqueous and gaseous phases, respec-
tively. In various phases and over a wide temperature range,
corrosive halogens and sulfur compounds need to be man-
aged during the process. To date, many investigations and
screening tests of structural materials for fabricating
corrosion-resistant equipment have been conducted in rep-
resentative process environments [4e18].
Candidate industrial materials have been screened from
the viewpoint of corrosion resistance to the environments in
the sulfuric acid decomposition section. Evaluation of mate-
rials for the sulfuric acid vaporization by immersion tests [4]
has revealed that silicones, silicon carbides, and silicon ni-
trides have good corrosion resistance in 98 wt% sulfuric acid
up to 452

Cand 2 MPa. Fromthe viewpoint of the construction


of equipments for the vaporization as well as acid concen-
trations, further studies expanding the range of the sulfuric
acid concentrations, a lowconcentration regionunder 98 wt%,
are indispensable. In addition, with an eye on the commercial
availability of the materials, evaluations of metallic materials
are required to broaden the variety of the structural materials.
In this study, immersion tests under conditions of high-
temperature pressurized aqueous sulfuric acids were carried
out. The corrosion resistance of ceramics and acid-resistant
metals were examined in relation to the concentration and
vaporization process using representative conditions at 2 MPa
and acid concentrations of 75, 85, and 95 wt%.
2. Corrosion environments within
iodineesulfur process
2.1. Process description
Fig. 1 shows a schematic of the IS process. The Bunsenreaction
Eq. (1) produces two acids, hydriodic acid (hydrogen iodide, HI,
in water) and sulfuric acid, fromwater, and sulfur dioxide and
iodine as rawmaterials inanaqueous solution. The mixedacid
separates into two types of acids naturally (liquideliquid
phase separation). The acid that is rich in HI is the HIx phase,
while the acid that is rich in sulfuric acid H
2
SO
4
, is the H
2
SO
4
phase. After separation, the acids are puried and con-
centrated. Then, gaseous HI is separated from the solution by
distillation and H
2
SO
4
is gasied by vaporization. In the other
two reactions, they decomposed into the gaseous phases.
The hydriodic acid decomposes into iodine and hydrogen
as in Eq. (2). The sulfuric acid decomposes into water, sulfur
dioxide, and oxygen as in Eq. (3). These three reactions as
a whole force the water to split into hydrogenand oxygen. The
products of the sulfur dioxide, iodine, and water can be used
again for the production of acids in the Bunsen reaction. The
decomposition of sulfuric acid proceeds at temperatures of
approximately 900

C and absorbs heat. The decomposition of
100
800
600
200
400
1000
HIx(aq)+H
2
SO
4
(aq)
H
2
SO
4
(aq)
HIx(aq)
HIx(aq)
Gaseous phase
Gaseous phase
Vaporization
Bunsen reaction
(exothermic)
Liquid-liqud
phase separation
Decomposition of H
2
SO
4
(endothermic)
Decomposition of HI
(endothermic)
Raw material Product Product
Purification Purification
Separation
Distillation
H
2
O(g)
Conc. T
e
m
p
e
r
a
t
u
r
e

C
Concentration
1/2O
2
(g)+SO
2
(g)+H
2
O(g)
SO
3
(g)+H
2
O(g)
SO
2
(g)
H
2
O(g)
I
2
(g)+H
2
(g) 2HI(g)
H
2
O(l) 1/2O
2
(g) H
2
(g)
Fig. 1 e Schematic of iodineesulfur process.
i nt e r na t i ona l j o ur na l o f hy d r o g e n e ne r g y 3 8 ( 2 0 1 3 ) 6 5 7 7 e6 5 8 5 6578
hydriodic acid involves a small endothermic reaction at
approximately 500

C. The Bunsen reaction occurs exo-
thermically at temperatures of approximately 100

C.
2.2. Corrosion environments of sulfuric acid
decomposition
The sulfuric acid decomposition section involves the thermal
decomposition reaction of sulfuric acid, which is the highest
temperature reaction in the IS process. In this section, the
gaseous-phase decomposition of H
2
SO
4
into SO
3
and H
2
O oc-
curs spontaneously at approximately 600

C, and then the
decomposition of SO
3
into SO
2
and O
2
occurs at approximately
850

C in the presence of a catalyst such as platinum or iron
oxide. Both reactions exhibit a strong endothermic nature and
progress with high conversion ratios. The temperature range
for the reactions is well matched with that of HTGR. Prior to
vaporization and these gaseous reactions, aqueous sulfuric
acid in a concentration of approximately 50 wt% should be
further concentrated to over 90 wt%.
Most ow sheets that were studied employed operating
conditions of high temperature under high pressure [19e21]
for the sulfuric acid concentrator and vaporizer. For a sulfu-
ric acid concentration processing [20], 57 wt% sulfuric acid is
concentrated to 88 wt% by a multiple effect evaporator with
the operating pressure ranging from 0.008 to 2 MPa, and then
vaporized at 464

C. For another ow sheet [21], 57 wt% sul-
furic acid is concentrated to 93 wt% by a multiple effect
evaporator with the operating pressure ranging from 0.008 to
1.2 MPa, and then vaporized at 365

C. Thus, the structural
materials of the sulfuric acid concentrator and vaporizer are
exposed to very harsh corrosive environments.
2.3. Candidate structural materials
Fig. 2(a, b) show a brief process scheme (a) and candidate
structural materials (b) corresponding to the process scheme.
The process consists of three chemical reactions; corrosive
chemicals circulate while changing chemical forms, phases,
and temperature level. The corrosive environment is roughly
classied as liquid and gaseous phases for both iodine and
sulfur circulation environments. A summary of the corrosive
environment conditions is as follows: (1) gas phase, sulfuric
acid (H
2
SO
4
), sulfur trioxide, sulfur dioxide, oxygen, water,
bubble point to 850

C; (2) gaseous phase, hydrogen iodide,
iodine, hydrogen, water, bubble point to 500

C; (3) liquid
phase, sulfuric acid, sulfuric acid with contaminated iodine
and hydrogen iodide, room temperature to bubble point; (4)
liquid phase, poly-hydriodic acid, poly-hydriodic acid with
contaminatedsulfuric acid, roomtemperature to bubble point.
In the gaseous H
2
SO
4
decomposition step, some refractory
alloys that have been used in conventional chemical plants
showed good corrosion resistance. In the gaseous HI decom-
position step, nickel-based alloys were found to show good
corrosion resistance [5,17]. There are few concerns about the
structural materials for the gaseous-phase services.
The liquidphaseenvironments are moreseverethanthegas
phase environments. For the Bunsen reaction step, Ta (tanta-
lum), Zr (zirconium), ceramics, PFA (poly-peruoroalkoxy
copolymer) resins, and glasses showed corrosion resistance,
which can be applied using lining techniques. Around the HIx
bobble point, Ta, Nb (niobium), and silicon carbides (SiC)
showed excellent corrosion resistance [10].
Around the boiling conditions of the high concentrated
sulfuric acid, ceramic materials containing silicone suchas SiC,
silicon-impregnated silicon carbides (SieSiC), silicon nitrides
(Si
3
N
4
), and high-silicon iron (FeeSi) showed excellent corro-
sionresistance[4,14,22]. Fromtheviewpoint of theconstruction
of thesulfuricacidvaporizers as well as acidconcentrators, this
study examined the silicon containing materials in a lower
range of concentrations of aqueous sulfuricacids (75wt%, 85wt
%and95wt%). SiC, SieSiC, Si
3
N
4
, FeeSi, andthesiliconenriched
alloy (Sandvik SX) could be expected to have corrosion resist-
ance in such broad concentrations.
3. Experimental
3.1. Sample materials
Sample materials, which could be expected to have good cor-
rosion resistance, were selected on the basis of the results of
earlier corrosion tests [4,9,12,14]. The selected materials were
as follows: ceramic materials; a silicon-carbide (SiC; B, 1 wt%;
C, 2 wt%), a silicon-impregnated silicon-carbide (SieSiC; SiC,
80 wt%; Si, 20 wt%), and a silicon-nitride (Si
3
N
4
; CeO
2
, 5 wt%;
MgO, 4 wt%; SrO, 1 wt%), and metallic materials; a high-silicon
iron (Fee15Si; Si, 14.8 wt%), another high-silicon iron (Fee20Si;
Si, 19.7 wt%), and a siliconenrichedalloy (SX is a brand name
from Sandvik Corp.; Fee19Nie17Cre5Sie2Cu). Quadratic-
prism shaped specimens were prepared in standard size of
4 mm 4 mm 40 mm. Surfaces of specimens, SiC, SieSiC,
Si
3
N
4
, and FeeSi, were cleaned with an ultrasonic washer
using acetone, which are so hard to be nished by polishing.
Surfaces of SX were polished with a emery paper (#600).
3.2. Conditions for immersion tests
Process designs of the IS process require to manage the sul-
furic acids in various conditions of temperatures, pressures,
and concentrations. Thus, the structural materials in direct
contact with the considered environments, in the sulfuric acid
vaporizers and acid concentrators, could be expected to be
exposed to the aqueous sulfuric acids in a broad range of
concentrations at temperatures close to the boiling points
under high pressure. A former study [4] evaluated corrosion
resistances of materials by immersion tests using just 98 wt%
sulfuric acid up to 452

C and 2 MPa. This is the highest con-
centration of the aqueous sulfuric acid because it is about
azeotropic composition. This study investigated the corrosion
resistance of the sample materials in a lower region of sulfuric
acid concentrations to reach such high concentration so that
the concentrations of sulfuric acids used in this study were
75 wt%, 85 wt% and 95 wt%.
3.3. Experimental apparatus and procedure
Immersion tests were carried out using an apparatus sche-
matically illustrated in Fig. 3. The apparatus provided corro-
sive environments of high-temperature aqueous-phase
i nt e r na t i o na l j our na l o f hy d r og e n e ne r g y 3 8 ( 2 0 1 3 ) 6 5 7 7 e6 5 8 5 6579
sulfuric acid by pressurization. It consisted of an autoclave
made of a stainless steel (JIS-SUS316), quartz ampoules con-
taining the specimens and the sulfuric acid, electric heaters,
and a helium gas cylinder.
The specimens with 5-mL sulfuric acid were encapsulated
in the quartz ampoules; the ampoules were cooled by liquid
nitrogen, evacuated, and then sealed under vacuum condi-
tions. The sealed ampoules were placed in the autoclave and
were covered with aluminum foils to prevent quartz frag-
ments scattering.
The atmosphere in the autoclave was replaced with helium
gas before heating. Then, the autoclave was heated up to
a specied temperature that was maintained at a constant
level for a specied time; the heating rate was 100

C h
1
.
While heating up, the pressure inside the autoclave was
adjusted so as to be the same or somewhat higher than the
vapor pressure [23] on the sulfuric acid, which is shown in
Fig. 4. This was done to keep the pressure difference of the
ampoules small.
Test conditions simulated the stages of the sulfuric acid
concentrator and the vaporizer, as they appear in earlier ow
sheets [20,21]. Three types of sulfuric acid were used as test
solutions with concentrations of 75, 85, and 95 wt%. Test
temperatures were 320

C (sulfuric acid, 75 wt%), 380

C (85 wt
%), and 460

C (95 wt%), which are close to the boiling points
[23] under high pressure (2 MPa).
The exposure times were 100 h and 1000 h in the case of
95 wt% sulfuric acid. In the cases of 75 and 85 wt% sulfuric
acid, the exposure times were 100 h.
Some ampoules were brokenduring the immersiontests as
detected by internal pressure monitors. When this occurred,
the broken ampoules were removed from the autoclave, and
the immersion tests were continued with the remaining
sound ampoules.
Fig. 2 e Candidate materials for prototypical environments of iodineesulfur process. FeSi indicates high silicon iron.
i nt e r na t i ona l j o ur na l o f hy d r o g e n e ne r g y 3 8 ( 2 0 1 3 ) 6 5 7 7 e6 5 8 5 6580
After exposure, the specimens were taken out from the
chopped ampoules, and the specimens were weighed after
cleaning and drying. Corrosion rates were determined by
weight changes. Microscopic observations and elemental
analysis of cross sections using electron probe micro analysis
(EPMA) were conducted.
4. Results and discussions
4.1. External view
Figs. 5 and 6 show visual appearances of the quartz ampoules
and the specimens before and after exposure to 95 wt% sul-
furic acid for 100 h. No scales were observed on the ceramic
specimens (SiC, SieSiC, and Si
3
N
4
) after the exposure. In the
1000 h exposure to 95 wt%sulfuric acid case, a fewscales were
observed on the SiC and SieSiC specimens; the scales were
easily removed by the cleaning process. No scales were
observed on the specimen of Si
3
N
4
after the exposure. There
was a lm-like scale on Fee15Si, but the scale was also easily
removed by the cleaning process. For SX, greenish-yellow
scales were observed. In the 1000 h exposure to 95 wt% sul-
furic acid case, the thickness of the SX specimens underwent
an reduction of 0.24 mm.
4.2. Corrosion rate, microscopy, and electron probe
micro analysis in cross section
Table 1 shows the corrosion rates measured by the weight
changes in the experiments, and Table 2 shows a general
criterion of the corrosion rate for chemical plant materials
[24]. Fig. 7 shows the microscopic observations of specimens
cross sections.
Concerning the ceramics in the experiments using 75, 85,
and 95 wt% sulfuric acid, little change was detected in weight,
Fig. 3 e Immersion test apparatus for material evaluation
in high-temperature aqueous-phase sulfuric acid at 2 MPa.
Fig. 4 e Vapor pressure on sulfuric acid in high
concentration region.
Fig. 5 e Visual appearances of quartz ampoules containing
specimens and sulfuric acid, before and after exposure to
95 wt% sulfuric acid for 100 h (#1, SiC; #2, SieSiC; #3, Si
3
N
4
;
#4, SX; 5, Fee15Si; #6, Fee20Si).
i nt e r na t i o na l j our na l o f hy d r og e n e ne r g y 3 8 ( 2 0 1 3 ) 6 5 7 7 e6 5 8 5 6581
indicating their superior corrosion resistance. The corrosion
rates of SiC, SieSiC, and Si
3
N
4
were excellent
(<0.021 g m
2
h
1
) according to the ratings in Table 2. As
shown in Fig. 7(a, b, c), microscopic observations reveal no
evidence of cracks in the ceramics.
For SiC, the oxide layer was recognized on the surface of
the specimens through EPMA. The thickness of the layer was
a few micrometers after its exposure to 95 wt% sulfuric acid
for 1000 h. For SieSiC and Si
3
N
4
, the oxide layer was also
observed after exposure to 95 wt% sulfuric acid for 1000 h; the
thicknesses were several micrometers. A previous study [13]
reported that amorphous silicon oxide lms of 1-mm thick-
ness containing sulfur species were formed on surfaces of the
SieSiC after exposure to 95 wt%sulfuric acid at approximately
320

C and under atmospheric pressure. In this study, similar
thin oxide layers were formed on the surfaces of silicon con-
taining ceramics (SiC, SieSiC, and Si
3
N
4
).
Weight losses of SiC, 0.0004e0.14 g m
2
h
1
, were reported
in high-temperature water (290

C) [25]. A hypothesized re-
action model, in the case that a nonprotective layer was pro-
duced, was proposed as follows:
SiC 4H
2
O SiOH
4
CH
4
; (4)
so that the hydro-silica sol dissolved inthe water. On the other
hand, a reactionmodel, in the case of a lowwater content, was
shown [26] as follows:
SiC 2H
2
O SiO
2
C 2H
2
; (5)
so that a protective silica-layer is produced. Since high con-
centrated sulfuric acid contains a low level of water, these
silicon containing ceramics such as SiC showed excellent
corrosion resistance by the corrosion-resistant lm formation
in spite of the sulfuric acid concentration of this level and the
high temperature and pressure.
The as-prepared sample of Fee15Si corroded after expo-
sure to 95 wt% sulfuric acid for 100 h, and was rated as a fail-
ure (>0.42 g m
2
h
1
), although a past study [27] reported that
high silicon irons containing Si higher than 12 wt% showed
excellent corrosion resistance in 95 wt% sulfuric acid at at-
mospheric pressure. This aberration is probably causes by the
corrosive conditions at high temperature and pressure.
The Fee20Si annealed samples were tougher than Fee15Si.
The corrosionrate of Fee20Si after exposure to 95 wt%sulfuric
acid for 1000 h was rated fair (0.042e0.21 g m
2
h
1
) and those
after exposure to 75 and 85 wt% sulfuric acid for 100 h were
rated excellent. Using EPMA, an oxide layer containing sulfur
species was observed on the specimen surface; the thickness
of the layer was approximately 20 mmafter exposure to 95 wt%
sulfuric acid for 1000 h. An earlier study [27] reported that
Fig. 6 e Visual appearances of specimens before and after
exposure to 95 wt% sulfuric acid for 100 h.
Table 1 e Corrosion rates of specimens.
Sample Corrosion rate
a
(g m
2
h
1
)
Aqueous H
2
SO
4
,
75 wt%
b
Aqueous
H
2
SO
4
, 85 wt%
c
Aqueous H
2
SO
4
,
95 wt%
d
100 h exposure 100 h exposure 100 h exposure 1000 h exposure
SiC 0.0 (0.0)
f
0.10 0.002
SieSiC (SiC, 80 wt%; Si, 20 wt%) e 0.0 0.0 0.006
Si
3
N
4
0.0 0.0 0.0 0.007
Fee15Si (Si, 14.8 wt%; as-prepared) e e 1.1 e
Fee20Si (Si, 19.7 wt%; as-prepared) e e 0.12 0.13
Fee20Si (Si, 19.7 wt%; annealed
h
) 0.0 0.0 e 0.065
Sandvik SX

(Fee19Nie17Cre5Sie2Cu, pre-oxidized
i
) e (4.5)
f
0.28 (0.96)
g
Negative values indicate weight gain.
Conversion formula of corrosion rates c
2
(mm y
1
) c
1
(g m
2
h
1
) 24 0.365 r.
(r are densities (g cm
3
) of specimens, SiC, 2.9; SieSiC, 2.9; Si
3
N
4
, 3.2; Fee15Si, 7.2; Fee20Si, 6.8; SX, 7.6).
a Values were calculated from weight losses.
b Test conditions: 320

C, 2 MPa.
c Test conditions: 380

C, 2 MPa.
d Test conditions: 460

C, 2 MPa.
e Quartz ampoule broken due to trouble in temperature control (1 h, 390

C).
f Quartz ampoule containing specimen broke after 17 h.
g Quartz ampoule containing specimen broke after 800 h.
h Annealed at 1100

C under vacuum for 100 h.
i Oxidized in air at 800

C for 90 h.
i nt e r na t i ona l j o ur na l o f hy d r o g e n e ne r g y 3 8 ( 2 0 1 3 ) 6 5 7 7 e6 5 8 5 6582
silicon dioxide lms containing sulfur species were formed on
the surfaces of FeeSi (silicon content of 13 wt%) after exposure
to 95 wt% sulfuric acid at atmospheric pressure. In this study,
the oxide layers showed corrosion resistance in spite of the
range of sulfuric acid concentrations and the harsh conditions
(the high temperatures and pressures).
Microscopic observation of FeeSi revealed the formation of
a number of cracks, as shown in Fig. 7(d, e); the annealing
process was found to be effective in reducing the formation of
cracks. The annealing process also improved the corrosion
resistance (Table 1). In the case of the experiments using 75
and 85 wt% sulfuric acid, microcracks were also observed.
This fact suggests that the stress imposed in the sample
preparation was responsible for the observed corrosion.
Because it was impossible to suppress the crack formation
completely, FeeSi materials might not be suitable as struc-
tural materials in the considered environments.
SX suffered from uniform corrosion; the corrosion rate of
SX, after exposure to 95 wt%sulfuric acid for 1000 h, was rated
as a failure (>0.42 g m
2
h
1
). Note that the corrosion rate
could be decreased by a pre-oxidation treatment. Oxide layers
(with thickness of approximately 40 mm) containing sulfur
species were formed after the exposure. As shown in Fig. 7(f,
g), the specimen of SX in 85 wt% sulfuric acid suffered more
severe corrosion than that in the 95 wt% sulfuric acid. This is
ascribed to the stronger chemical reactivity of sulfuric acids of
lower concentrations.
4.3. Material selection for sulfuric acid concentration
and vaporization
Fig. 8 provides a summary of corrosion rates of the silicon
containing materials to the aqueous sulfuric acids in various
concentrations. Items appear on the vertical axis are material
names divided into groups of the materials; the each group is
Table 2 e General criteria of corrosion rate for chemical
plant materials.
Rating of corrosion
resistance
Criteria of rates
Weight
loss
a
/g m
2
h
1
Penetration
depth
b
/mm y
1
Excellent 0e0.021 0e0.05
Good 0.021e0.042 0.05e0.1
Fair 0.042e0.21 0.1e0.5
Poor 0.21e0.42 0.5e1.0
Failure >0.42 >1.0
a General corrosion.
b Maximum value.
Fig. 7 e Microscopic observations of specimen cross
section.
Fig. 8 e Corrosion rates of silicon containing materials to
aqueous sulfuric acids in various concentrations; (a),
present; (b), cited from [4]; (c), cited from [9]; (d), cited from
[11]; (e) cited from [27]. SX is a brand name from Sandvik
Corp.; Fee19Nie17Cre5Sie2Cu. Duriron is a brand name
of a high silicon iron containing 14.5 wt% silicon. Bold
letters denote results of this study.
i nt e r na t i o na l j our na l o f hy d r og e n e ne r g y 3 8 ( 2 0 1 3 ) 6 5 7 7 e6 5 8 5 6583
arranged so that the highest concentration of the sulfuric acid
is placed at the bottom.
To combine our results with the results of [4,11], we can
state that SiC and Si
3
N
4
show excellent corrosion resistance
to a broad range of sulfuric acid concentrations of 75e98 wt%
(the azeotropic point) at the high pressure of 2 MPa.
In addition, they resist to the acid in low concentration of
50 wt%.
FeeSi, silicon content of 20 wt%, showed some measure of
corrosion resistances to the sulfuric acid concentrations in
50 wt% to 95 wt%, which rated not fewer than fair; however,
we observed the evidences of crack formation. FeeSi, silicon
content of 15 wt%, showed the corrosion resistances in the
sulfuric acid concentrations in 95 and 98 wt%; however, we
observed the evidences of crack formation and the higher
corrosion rate than other studies [27]. For lower concentration
of 40 wt%, the corrosion rate remarkably intensies by
including the iodine species, through Duriron exhibits very
good resistance, 0.10e0.41 g m
2
h
1
, to the entire range of
sulfuric acid concentrations at temperatures including the
boiling points [28]. Note that it rapidly attacks in service con-
taining free sulfur trioxide [28]. Because of the crack formation
and the iodine contamination effect, considerable care should
be exercised in FeeSis adoption to the considered sulfuric
acid processing.
The corrosion rates of SX were rated as failure; the sulfuric
acid at lower concentration (85 wt%) provide severer corrosive
environment than the 95 wt% sulfuric acid. SX is not suitable
for use as structural materials inthe considered environments.
We revealed that the silicon-containing ceramics are the
candidate structural materials in direct contact with the
aqueous sulfuric acids in broad range of concentrations (50 wt
% to 98 wt%) at near bubbling temperature and high pressure
of 2 MPa.
5. Conclusions
Immersion tests were carried out to evaluate corrosion re-
sistances of ceramics and acid resistant metals for the fab-
rication of sulfuric acid processing equipment in the
sulfureiodine water splitting hydrogen production process.
Silicon-carbide (SiC), silicon-impregnated silicon-carbide
(SieSiC), silicon-nitride (Si
3
N
4
), high silicon iron (FeeSi), and
a silicon enriched alloy (Sandvik SX), which could be expected
to have a high corrosion resistance, were examined from the
viewpoint of constructing the sulfuric acid vaporizers and
concentrators. Corrosive environments using sulfuric acid
with concentrations of 75, 85, and 95 wt%, were provided by
the autoclave containing the specimens and the sulfuric acid
encapsulated with quartz ampoules. The immersion tests
were conducted at temperatures of 320, 380, and 460

C and
a pressure of 2 MPa. The temperature conditions were set
close to the boiling points. As a result, excellent corrosion
resistances and corrosion rates were measured for all SiC,
SieSiC, and Si
3
N
4
specimens, and they showed no weight
changes. FeeSi with silicon content of 20 wt% showed some
measure of corrosion resistance that rated fair. However, ev-
idence of crack formationwas found. The annealing operation
could not prevent the crack formation completely as far as we
could determine. Sandvik SXsuffered fromuniformcorrosion.
The corrosion rate in 95 wt% sulfuric acid was approximately
1 g m
2
h
1
, which was rated as a failure. Within the scope of
the test conditions, the FeeSi and Sandvik SX metallic mate-
rials are not suitable for use as structural materials in the
considered environments. We conclude that the silicon con-
taining ceramics are excellent candidates for structural ma-
terials in direct contact with aqueous sulfuric acids in a broad
range of concentrations (75 wt% to 98 wt%) at the high pres-
sure of 2 MPa.
Acknowledgments
We thank NGK Insulators, Ltd. and Mitsui Engineering &
Shipbuilding Co., Ltd. for providing sample materials.
r e f e r e n c e s
[1] Onuki K, Kubo S, Terada A, Sakaba N, Hino R.
Thermochemical water-splitting cycle using iodine and
sulfur. Energy Environ Sci 2009;2:491e7.
[2] Winter CJ. Hydrogen energy e abundant, efcient, clean:
a debate over the energy-system-of-change. Int J Hydrog
Energy 2009;34(14, Supplement 1):S1e52.
[3] Norman JH, Besenbruch G, Keefe DO. Thermochemical
water-splitting for hydrogen production. Final report
(January 1975-December 1980), GRI-80/0105; March 1981.
[4] Ammon R. Status of materials evaluation for sulfuric acid
vaporization and decomposition application. Proc. 4th world
hydrogen energy Conf.. California, USA: June 1982;2:623e44.
[5] Imai Y, Kanda Y, Sasaki H, Togano H. Corrosion resistance of
materials in high temperature gases composed of iodine,
hydrogen iodide and water. Boshoku Gijutsu 1982;31:714e21.
[6] Kondo W, Kaneko M, Takemori Y, Sasaki H, Fujii K. Corrosion
resistance of ceramic materials in high temperature gases
composed of iodine, hydrogen iodide and water
(environment of the 3rd and 4th stage reactions). Bousyoku
Gijutsu 1982;31:722e7.
[7] Imai Y, Mizuta S, Nakauchi H. Material problems associated
with hydrogen production for waterewith special interest
in thermochemical method. Bousyoku Gijutsu 1986;35:
230e40.
[8] Porisini FC. Selection and evaluation of materials for the
construction of a pre-pilot plant for thermal decomposition
of sulfuric acid. Int J Hydrog Energy 1989;14:267e74.
[9] Onuki K, Nakajima H, Shimizu S, Sato S, Tayama I. Materials
of construction for the thermochemical is process, (i). J
Hydrog Energy Syst Soc Japan 1993;18:49e56.
[10] Onuki K, Ioka I, Futakawa M, Nakajima H, Shimizu S, Sato S,
et al. Materials of construction for the thermochemical is
process, (ii). J Hydrog Energy Syst Soc Japan 1994;19:10e6.
[11] Onuki K, Ioka I, Futakawa M, Nakajima H, Shimizu S,
Tayama I. Screening tests on materials of construction for
the thermochemical is process. Corros Eng 1997;46:141e9.
[12] Ioka I, Onuki K, Futakawa M, Kuriki Y, Nagoshi M,
Nakajima H, et al. Corrosion resistance of feesi alloys in
boiling sulfuric acid. Zairyo (J Soc Mat Sci, Japan) 1997a;46:
1041e5.
[13] Futakawa M, Onuki K, Steinbrech RW. Corrosion resistance
of oxide scale formed on sisic in boiling sulfuric acid. J Surf
Finish Soc Japan 1997;48(6):l662e3.
[14] Nishiyama N, Futakawa M, Ioka I, Onuki K, Shimizu S, Eto M,
et al. Corrosion resistance evaluation of brittle materials in
i nt e r na t i ona l j o ur na l o f hy d r o g e n e ne r g y 3 8 ( 2 0 1 3 ) 6 5 7 7 e6 5 8 5 6584
boiling sulfuric acid. Zairyo (J Soc Mat Sci, Japan) 1999;48:
746e52.
[15] Ioka I, Mori J, Kato C, Futakawa M, Onuki K. The
characterization of passive lms on fe-si alloy in boiling
sulfuric acid. J Mater Sci Lett 1999;18:1497e9.
[16] Kurata Y, Tachibana K, Suzuki T. High temperature tensile
properties of metallic materials exposed to a sulfuric acid
decomposition gas environment. J Japan Inst Met 2001;65:
262e5.
[17] Futakawa M, Kubo S, Wakui T, Onuki K, Shimizu S,
Yamaguchi A. Mechanical property evaluation of surface
layer corroded in thermochemical-hydrogen-production
process condition. J Japanese Soc Exp Mech 2003;3(2):109e14.
[18] Kubo S, Futakawa M, Ioka I, Onuki K, Shimizu S, Ohsaka K,
et al. Corrosion test on structural materials for iodineesulfur
thermochemical water splitting cycle. In: Proc 2nd topical
Conf on fuel cell technology, AIChE 2003 spring national
meeting held at New Orleans. NY: AIChE; 2003.
[19] Normanl J, Besenbruch G, Brown L, Oeefe D, Allen C.
Thermochemical water-splitting cycle, bench-scale
investigations and process engineering; 1982. GA-A 16713.
[20] Knoche K, Schepers H, Hesselmann K. Second law and cost
analysis of the oxygen generation step of the general atomic
sulfureiodine cycle. Proc 5th world hydrogen energy Conf..
Toronto, Canada: July 1984;2:487e502.
[21] Ozturk IT, Hammache A, Bilgen E. An improved process for
h2so4 decomposition step of the sulfureiodine cycle. Energy
Convers Mgmt 1995;36(1):11e21.
[22] Roeb M, Thomey D, Graf D, de Oliveira L, Sattler C, Poitou S,
et al. Hycycles project on solar and nuclear hydrogen
production by sulphur-based thermochemical cycles. In:
Proceedings of 18th world hydrogen energy conference 2010
May 16-21; 2010. HP.2 thermochemical cycles. Essen,
Germany.
[23] Helming J. Beitrag zur korrelation von
phasengleichgenwichtsdaten. Technischen hochschule
Aachen, thesis, 1983.
[24] Saichi H. Kagaku sochi taisyoku-hyo. Kagaku kougyo-sya;
1998.
[25] Hirayama H, Kawakubo T, Goto A. Corrosion behavior of
silicon carbide in 290c water. J Am Ceram Soc 1989;72(11):
2049e53.
[26] Yoshio T, Oda K. Corrosion resistance of structural ceramics
in aqueous environment. Zairyo-to-Kankyo 1995;44(7):
405e15.
[27] Ioka I, Onuki K, Futakawa M, Kuriki Y, Nagoshi M,
Nakajima H, et al. Corrosion resistance of feesi alloys in
boiling sulfuric acid. J Soc Mater Sci Japan 1997b;46(9):
1041e5.
[28] Ryusan-kyokai, editor. Sulfuric acid handbook; 1977.
i nt e r na t i o na l j our na l o f hy d r og e n e ne r g y 3 8 ( 2 0 1 3 ) 6 5 7 7 e6 5 8 5 6585

Vous aimerez peut-être aussi