2021-Development of Catalysts For Sulfuric Acid Decomposition in The Sulfur Iodine Cycle A Review

Download as pdf or txt
Download as pdf or txt
You are on page 1of 37

Catalysis Reviews

Science and Engineering

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/lctr20

Development of catalysts for sulfuric acid


decomposition in the sulfur–iodine cycle: a review

Hassnain Abbas Khan, Ahsan Jaleel, Eyas Mahmoud, Shoaib Ahmed, Umair
Hassan Bhatti, Muhammad Bilal & Hussain

To cite this article: Hassnain Abbas Khan, Ahsan Jaleel, Eyas Mahmoud, Shoaib Ahmed,
Umair Hassan Bhatti, Muhammad Bilal & Hussain (2021): Development of catalysts for
sulfuric acid decomposition in the sulfur–iodine cycle: a review, Catalysis Reviews, DOI:
10.1080/01614940.2021.1882048

To link to this article: https://doi.org/10.1080/01614940.2021.1882048

Published online: 04 Apr 2021.

Submit your article to this journal

Article views: 185

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=lctr20
CATALYSIS REVIEWS
https://doi.org/10.1080/01614940.2021.1882048

Development of catalysts for sulfuric acid decomposition in


the sulfur–iodine cycle: a review
Hassnain Abbas Khan a, Ahsan Jaleel b, Eyas Mahmoudc, Shoaib Ahmedd,
Umair Hassan Bhatti e, Muhammad Bilalf, and Hussaing
a
Clean Combustion Research Centre, Division of Physical Sciences and Engineering, King Abdullah
University of Science and Technology (KAUST), Thuwal, Saudi Arabia; bClean Energy Research Centre,
Korea Institute of Science and Technology, Seongbuk-gu, Republic of Korea; cDepartment of Chemical
and Petroleum Engineering, College of Engineering, United Arab Emirates University, Al Ain, United Arab
Emirates; dChemical Engineering Department, Dawood University of Engineering and Technology,
Karachi, Pakistan; eKorea Institute of Energy Research, University of Science and Technology, Daejeon,
Republic of Korea; fSchool of Life Science and Food Engineering, Huaiyin Institute of Technology, Huaian,
China; gDepartment of Metallurgy and Materials Engineering, University of the Punjab, Lahore Pakistan

ABSTRACT KEYWORDS
To achieve carbon-neutral energy vectors, researchers have inves­ SI cycle; sulfuric acid
tigated various sulfur-based thermochemical cycles. The sulfur– decomposition; hydrogen
iodine cycle has emerged as a cost-effective global process with production cycle; water
splitting; catalyst stability
massive hydrogen production potentials. However, all sulfur-based
thermochemical cycles involve sulfuric acid decomposition reac­
tion, which is highly corrosive and energy intensive. The activation
energy of this reaction can be reduced using catalysts that decrease
the onset temperature of the reaction. Renewable heat sources
such as solar and waste nuclear heat demand high stability to
operate within a wide temperature window (650°C–900°C).
Several metal/metal oxide systems based on noble and transition
metals have been investigated over the last twenty years. In the
literature, supported Pt-based catalysts are regarded as the prime
choice for stable operations. However, during catalytic operations,
noble metals are degraded owing to sintering, oxidation, leaching,
and other processes. Transition metal oxides such as Fe, Cu, Cr, and
Ni exhibit promising catalytic activity at high temperatures; how­
ever, at low temperatures (>600°C), their activation is reduced
owing to poisoning and the formation of stable sulfate species.
The catalytic activity of transition metal oxides is determined by the
decomposition temperature of its corresponding metal sulfate;
thus, the metal sulfate formation is considered as the rate-limiting
step. Herein, the catalytic systems studied over the last decade are
summarized, and recommendations for designing robust catalysts
for commercial applications are presented.

1. Introduction
Hydrogen is considered as the cleanest fuel among all-natural and synthetic
fuels.[1] Hydrogen has been extracted from natural sources using steam

CONTACT Hassnain Abbas Khan [email protected] Clean Combustion Research Centre, Division
of Physical Sciences and Engineering, King Abdullah University of Science and Technology (KAUST), Thuwal 23955-
6900, Saudi Arabia; Eyas Mahmoud [email protected] Department of Chemical and Petroleum
Engineering, College of Engineering, United Arab Emirates University, Al Ain 15551, United Arab Emirates
© 2021 Taylor & Francis
2 H. A. KHAN ET AL.

reforming, hydrogen sulfide decomposition, ammonia decomposition, and


water splitting. As a raw material, water has garnered considerable interest;
however, hydrogen production from water requires a high temperature to
decompose stable water molecules.[2] Cyclic routes that couple low-
temperature reactions into a net chemical reaction are more feasible routes to
realize water decomposition.[2,3] Potential pathways for sulfuric acid (SA)
decomposition reaction include hybrid sulfur-based thermochemical cycles,
[4,5]
such as sulfur–bromine and sulfur–iodine (SI) cycles. [6,7] SI cycles are
particularly promising for generating massive amounts of hydrogen with high
efficiency (30%–47%) using nuclear or solar heat sources. In particular, a high-
temperature gas-cooled reactor (HTGR) supplies heat at approximately 1000°
C. [2,8,9]
The three-step reaction SI cycle is listed.
Overall reaction
( H2 SO4 ðlÞ þ 2HI ðlÞ ð80� C
SO2 ð g Þ þ I2 ðlÞ þ 2H2 O ðlÞ + 120� CÞ (1)
Bunsen Reaction

2HI ð g Þ (
+ I2 ð g Þ þ H2 ð g Þ ð,450 CÞ (2)
SA thermal dissociation

H2 SO4 ð g Þ (
+ H2 O ð g Þ þ SO3 ð g Þ ð400 C 450� CÞ (3)
Catalytic sulfur trioxide decomposition
( SO2 ð g Þ þ 1=2 O2 ð g Þ
SO3 ð g Þ + (4)
Equations (3) and (4) are the two steps of the SA decomposition reaction. The
first step is the thermal dissociation reaction yielding SO3 and H2O at tem­
peratures above 350°C, and the second is the catalytic decomposition of SO3 to
afford SO2 and O2.[9]
Several catalysts for SA decomposition have been explored and
reported.[10,11] However, the stability of various metal oxide catalysts has
been examined only at elevated temperatures (>800°C). Kim et al.[12] proposed
a mechanism based on temperature-programmed desorption (TPD) and
infrared (IR) spectroscopy data. They reported that O2 and SO2 are formed
via the thermal decomposition of metal sulfates, which is generated as an
intermediate product during the reaction.[12] Based on the TPD patterns of the
metal sulfates of spent catalysts, the activity of metal oxide catalysts has been
correlated with the decomposition temperature of metal sulfates.[13,14] In other
words, the reaction relies on the decomposition temperature of reaction-
formed species such as Al2(SO4)3. If the reaction temperature within the
recommended range (550°C–850°C) is below the decomposition temperature
of the metal sulfate, SO3 cannot be catalyzed. At temperatures below the
CATALYSIS REVIEWS 3

deposition temperature, metal sulfate is inevitably observed on metal oxide


catalysts.[12,13] Transition metals such as iron, copper, and chromium and
bimetallic catalysts formed from these metals have also been widely considered
for SA decomposition because their corresponding metal sulfate decomposi­
tion temperatures are lower.[15,16] Soft platinum nanoparticles exhibit the
highest catalytic activity.[17] Rashkeev et al. matched the experimental results
with the results of density functional theory (DFT) based on first principles.[17]
However, similar to unsupported metal oxides, metal oxides (including Al2O3,
ZrO3, CeO2, TiO2, BaSO4, and Ta2O5) supported by noble metal (Pt/C)
systems are applicable to SA decomposition only at temperatures above
~800°C,[18–21] because their activity is limited owing to metal sulfate formation
on the support. For example, a Pt/Al2O3 catalyst is rigorously deactivated at
reaction temperatures below 800°C owing to stable Al2SO4 formation
(poisoning).[22] The sulfate can be retained on the alumina support by adding
silicon carbide coatings.[22]The silicon carbide coating (Pt/SiC–Al2O3) is stable
in the temperature range of 650°C–850°C but is not immune to aluminum
sulfate formation during the SA reaction, and some evidence of Al(SO4)2 is
observed. Pure nanosized SiC-supported Pt catalysts (Pt/n-SiC) have also been
reported.[23] Silicon carbide resists acid corrosion and remains sulfate-free in
the presence of SA. Although these catalysts are stable and active in a wide
temperature range, Pt is weakly bound to the SiC support, thus causing active
Pt metal loss. Pt-impregnated anatasis TiO2(Pt/TiO2–A) is highly active and
stable for SO3 decomposition at lower temperatures (~600°C), which is pro­
mising for the low-temperature SI hydrogen cycle. Moreover, the catalytic
activity of Pt/TiO2–A is significantly superior to that of a rutile TiO-supported
Pt catalyst (Pt/TiO2–R) at a high reaction temperature (800°C).[24,25] A major
challenge in SA decomposition is the strong corrosive effect of gaseous species
such as H2SO4, SO3, SO2, O2, and H2O. Nevertheless, Pt-based catalysts have
proven their practicality.[16,26–32]
Iron oxide (Fe2O3) has been extensively investigated as an alternative to
Pt-based catalysts; however, its vulnerability to corrosive environments at
low temperatures has remained a challenging issue. Cr-doped Fe2O3 exhibits
similar behavior to Pt/Al2O3, and is more stable.[16,31–35] In extended time
stability tests conducted at ~800°C, Cr doping stabilized Fe2O3 and con­
trolled the agglomeration.[35] SiO2 has also gained attention as an inert
support in high-temperature corrosive reactions. To prevent poisoning, an
Fe2O3 catalyst was supported on a SiO2 support, thus stabilizing Fe2O3
nanoparticles and minimizing deactivation at low temperatures.[33] The
catalytic performance of Cu was significantly improved in the Cu@SiO2
composite.
In this short but comprehensive review, the catalytic systems and materials
used in the SA decomposition step of the SI cycle are summarized. Previously
published papers related to this topic are evaluated to discuss the advantages
4 H. A. KHAN ET AL.

and disadvantages of different catalysts for commercial applications.


Moreover, a computational approach is suggested for investigating the deac­
tivation mechanism and intermediate species, which can be the cause of
deactivation. Finally, potential catalytic materials are highlighted for SA
decomposition reactions in the SI cycle for massive hydrogen production.

1.1. Background of thermochemical water-splitting cycles for hydrogen


production

High-temperature thermal decomposition of water via cyclic reactions com­


bined with a sustainable heat source is considered as the most promising
hydrogen production strategy.[36] The high-temperature source (which can
be solar or nuclear) drives the chemical reactions in the hydrogen production
cycle.[6,37,38] The primary energy source of thermochemical hydrogen produc­
tion is the medium-temperature (600°C–850°C) solar or thermal output of
helium from nuclear power plants.[39] When the oil crisis occurred in the late
1970s, water splitting for alternate energy carriers gained enormous
popularity.[6,37,38,40] Nearly 280 hydrogen production cycles with different
sets of chemical reactions and temperatures from alternate sources (solar
and nuclear) were explored by the PROMES-CNRS research group.[41]
However, some of these cycles are thermodynamically and capitally infeasible.
The most thermodynamically favorable cycles are sulfur-based thermochemi­
cal cycles such as hybrid sulfur-based cycles,[37,39] the IS cycle of general
atomics (GA),[37,39,42] and UT-3 sulfur–bromine hybrid cycle.[43–47] Owing
to many thermodynamic, economic, solar/nuclear, environmental, and safety
factors, these cycles are more promising for further development than the
other 280 cycles. Moreover, all sulfur-based cycles operate at high tempera­
tures; thus, they are heated using waste nuclear materials to increase the
efficiency of the reaction. The reaction is performed in HTGR operated at
730–950 K.[48]
The SI cycle is the most efficient thermochemical cycle, with a theoretical
efficiency of 52%.[5,49] This efficiency can be further improved and the capital
cost can be lowered. In ongoing research efforts, the energy input of the SI
cycle is increased using the waste heat from advanced nuclear power reactors,
which also afford clean hydrogen from the SI cycle.

1.2. Emerging hydrogen economy concept


Energy is the driving force behind every growing economy and modernization.[50]
Power and energy are essential to the modern human lifestyle, driving technolo­
gical growth, societal development, and economies.[51] The hydrogen economy
concept was formulated by John O’Mara Bockris in 1970, who was a professor of
chemistry at the University of Pennsylvania. At that time, when energy crises
CATALYSIS REVIEWS 5

became a global concern, Bockris proposed several thermochemical cycles for


hydrogen production. Several of these cycles showed promising capabilities when
coupled with an elevated-temperature nuclear reactor. In 1970, an American
scientist working for GA introduced the first SI cycle. The SI cycle was first
described and explained by Russell in 1976.[52] Since then, it has gained interest
owing to its higher thermodynamic efficiency and lower operational temperature
than other cycles.[53] Researchers have concluded that GA’s SI cycle is the most
thermodynamically and economically feasible option for massive thermochemical
hydrogen production.[53] The research group at the France atomic energy institute
almost reached the same conclusion.[54] The United States, Korea, Japan, and
many other countries have developed commercial-level thermochemical water-
splitting facilities for hydrogen production using the SI cycle.[55] By the end of
2026, a high-temperature reactor with an efficiency of 30 MW (t) is expected.[55]

1.3. Gaseous SA decomposition-based studies and their economic evaluation

Economic evaluation is a prerequisite for commercial testing of any process.


For economic evaluations of nuclear hydrogen production projects in devel­
opment, the hydrogen economic assessment program (HEEP) has been estab­
lished. In this subsection, detailed investigations and production routes of
hydrogen production were analyzed.
Figure 1 shows the different routes for developing hydrogen-based energy
systems. Waste heat from nuclear plants is the cheapest solution for producing
net hydrogen from plant to consumers. Although the overall efficiency of an
electrolysis plant from the primary source to electricity is ~40%, the opera­
tional efficiency is 90%. The overall efficiency of transforming the energy from
one form to another is 36%. However, the initial estimated theoretical effi­
ciency of the investigated thermochemical cycles is nearly 50%. The economic
evaluation results presented in Table 1 have been reproduced several times and
were consistent with the HEEP results.[56,57]
Table 1 presents the results of five generic and detailed case studies using
different technological and engineering facilities. The first three cases involved
conventional water electrolysis and an advanced pressurized water reactor.
Cases 1, 2 and 3 were performed in small-scale reactors at conventional
electrolysis plants, which output small production rates (4 and 5 kg/s, respec­
tively). Cases 4 and 5 were performed using next-generation nuclear reactor
technologies. The reactor in Case 4 generated heat at 900°C and was used to
perform high-temperature steam electrolysis. The reactor in Case 5 had
a maximum operating temperature of approximately 950°C and powered an
SI thermochemical plant. The storage and transportation capital costs were
included in the economic evaluations of all generic cases.
6 H. A. KHAN ET AL.

Figure 1. Various scenarios for producing renewable hydrogen and electricity: sustainable hydro­
gen economy concept.

Table 1. Benchmark results of HEEP with G4-ECONS for five generic cases to produce hydrogen
from water.
Generic Case 1 Generic Case 2 Generic Case 3 Generic Case 4 Generic Case 5
G4-ECONS LUHC $5.41/kgH2 $4.17/kgH2 $3.39/kgH2 $2.98 /kgH2 $2.46/kgH2
HEEP LUHC $5.61/kgH2 $4.25/kgH2 $3.70/kgH2 $3.34/kgH2 $2.95/kgH2
Prodution rate APWR + CE APWR + CE APWR + CE HTGR+HTSE HTGR+SI
H2 : 4 kg/s H2 : 8 kg/s H2 : 12 kg/s H2 : 4 kg/s H2 : 4 kg/s
LUHC Levelised unit hydrogen cost
HEEP Hydrogen Economic Evaluation Programmed
HTRGR High-temperature REVERSE
HTSE High-temperature steam electrolysis
SI; Sulfur Iodine
APWR Advanced pressurized water reactor
CE Conventional electrolysis

2. Review of catalytic systems


2.1. Transition metal-based supported catalysts
Catalytic decomposition of SO3 is the O2 evolution step in the SI cycle (see
Eq. 3). In the literature, SO3 has been decomposed over Cu, Fe, and Cu–Fe
supported on Al2O3.[29] In all catalyst preparations, the cost and stability of the
catalyst in strong acidic environments must be considered. Catalysts are
synthesized using the Yoldas and oil-drop methods. The selected synthesis
method strongly affects the physiochemical properties of the resulting materi­
als. The temperature range and gas hourly space velocity (GHSV) of the
reaction are 750°C–950°C and 72,000 mL gcat−1 h−1, respectively. High activity
has been reported for a Cu/Fe/Al2O3 catalyst at the optimum [Cu]:[Fe] ratio of
CATALYSIS REVIEWS 7

Figure 2. Data from Abimanyu et al., [29] Copyright© (2008) with permission from Elsevier.

0.5 and a [Cu] + [Fe] concentration of 0.125 M. The highest reported SO3
conversion was ~88.8% over a CuFeAl−4 catalyst at 950°C. Figure 2 shows the
temperature dependences of SO3 conversion for 13 samples. The catalyst
activity remained constant over time, and no deactivation was observed. The
CuAl catalysts outperformed the FeAl catalysts in SO3 decomposition. Metal
sulfate was formed when SO3 contacted the metal. The decomposition tem­
perature of the corresponding metal sulfate depends on individual metal.
CuSO4 and FeSO4 decompose at 650°C and 780°C, respectively. The CuAl
systems are more feasible for SO3 decomposition. Moreover, the metal pre­
cursor and support play a significant role in SO3 decomposition. Al2O3
catalysts protecting Cu, Fe, and TiO2 were prepared using a similar
method.[29]
When TiO2 was added to Cu–Fe/Al2O3, the specific surface area of the
catalyst (which was determined using the Brunauer–Emmett–Teller (BET)
method) decreased from 270.22 to 233 m2 g−1 as the titanium content
increased from CuFe/Ti:Al of 1:1 to CuFe/Ti:Al of 2:1. The surface area also
decreased when the Al2O3 content increased. The change in the surface area is
likely caused by pore filling using amorphous or crystalline metal oxides. The
results of a quantitative analysis conducted using atomic absorption spectro­
scopy and inductively coupled plasma optical emission spectroscopy (ICP-
OES) were consistent with the theoretical values, confirming the good
8 H. A. KHAN ET AL.

optimization of the synthesis method. No prominent Fe or Cu peaks appeared


in the X-ray diffraction (XRD) spectrum of the prepared catalyst, indicating
that the metals were well dispersed. UV–visible spectroscopy confirmed 69%
conversion of TiO2.CuFe/Ti:Al of 1:2 at 800°C, with a GHSV of 72,000 mL
gcat−1 h−1. The SO3 conversion over the investigated catalysts increased as the
temperature increased from 700°C to 850°C.
To validate the experimental results, the H2SO4 decomposition kinetics
were investigated over an Al2O3-supported Fe (III) oxide catalyst.[58] The
experimental kinetic parameters included temperature (750°C–900°C) and
space velocity (0.146–0.731 kmol kg−1 h−1).
In HYSYS simulations, the conversion of thermal SA decomposition to SO3
and H2O increased from 92% to ~99% as the temperature increased from 750
to 950 K. Moreover, the equilibrium conversion of SO3 decreased with an
increase in pressure (from 10 to 150 kPa). Within this pressure range, the
equilibrium conversion of SO3 to SO2 and O2 was below 10% at 700 K,
increasing to approximately 50% and more than 90% at 900 and 1200 K,
respectively. The space velocity significantly affected the SO3 conversion. The
activation energy (Ea) of SA decomposition over Fe2O3/Al2O3 was experi­
mentally determined as 138.6 kJ mol−1. The SA decomposition to SO3 was
abrupt and reached nearly 100%. The kinetically limiting reaction was SO3
decomposition to SO2 and O2, which must overcome a substantial kinetic
barrier under these SI cycle conditions. Metal sulfate decomposition is
a crucial step in cyclic thermochemical water splitting.[59] Another study on
metal sulfate decomposition was performed using ZnSO4, which was selected
for its high activity, low toxicity, and low cost. Ag–Pd and Fe–Pd alloy catalysts
supported on γ-Al2O3 and SiC were prepared using microwave-assisted synth­
eses. In this approach, a glycerol solvent was used as the reducing and
stabilizing agent. Bimetallic alloying was confirmed using XRD analysis.
A morphological scanning electron microscopy (SEM) analysis showed
a uniform dispersion of the nanoparticle catalysts. The zinc sulfate decom­
position temperature was 85°C lower in Fe–Pd/Al2O3 than in previously
reported systems.[60]
A granular Pt/Al2O3 catalyst for SA decomposition in a dual-tube reactor
with no carrier or diluent gas was investigated.[31] The two zones in the reactor
served as a boiler and preheater. Increasing the temperature from 700°C to
825°C increased the SO2 yield from approximately 30% to 80%, and increasing
the SA flux from 0 to 2.5 ml/min decreased the SO2 yield. The periodic
reduction of the SO2 product indicated that the catalyst had started to deac­
tivate. The yield decreased from 80% to 73% at 800°C. The SA acid flux was
0.63 ml min−1. The deactivation was associated with physicochemical changes
in the alumina support. Characterization using ICP-OES, N2 adsorption,
XRD, Fourier transform infrared spectroscopy, and SEM demonstrated that
in the deactivated system, the porosity, surface area, stable sulfation, and
CATALYSIS REVIEWS 9

Figure 3. Comparison of the DTG profiles of the spent Pt/Al2O3 catalysts (a) with a 1:4 molar
mixture of Al2(SO4)3 with Al2O3 (b) and Al2(SO4)3 (c). Data from Banerjee AM et al.[31] copyright©
(2015) with permission from Elsevier.

partial phase transformations decreased in the alumina support. Transmission


electron microscopy with energy dispersive X-ray (TEM-EDX) spectroscopy
confirmed that the active metal Pt was thermally sintered, forming large,
poorly dispersed agglomerates. Differential thermogravimetric (DTG) profiles
of the used Pt/Al2O3 catalyst is shown in Figure 3. The studies based on TG-
DTG evolved gas analyses confirmed a synergistic effect between the Pt
particles and the Al2O3–Al2(SO4)3 support.
In the presence of Pt, Al2(SO4)3 decomposes into Al2O3 using aluminum
oxy-sulfate species, thus reducing the decomposition temperature of corre­
sponding Al2(SO4)2. The two-step reaction is expressed in Equations (5)
and (6).

Al2 ðSO4 Þ3 Al2 OðSO4 Þ2 þ SO2 þ 1=2 O2 (5)

Al2 OðSO4 Þ2 Al2 O3 þ 2SO2 þ O2 (6)


The low- and high-temperature (>500°C) peaks in Figure 3 are attributed to
water loss and sulfate decomposition, respectively. The one-step decomposi­
tion of bulk aluminum sulfate occurred at 650°C, and aluminum sulfate in the
mechanical Al2O3 mixture formed at 665°C. In the used Pt/Al2O3 catalyst,
sulfate underwent a two-step decomposition starting at a low temperature of
510°C. The area ratio of the first and second peaks (with maxima at 630°C and
751°C, respectively) was 1:2. Results suggested that in addition to the adsorp­
tion–decomposition mechanism, the reaction can proceed through aluminum
sulfate formation and decomposition (Equations 5 and 6). The interaction
10 H. A. KHAN ET AL.

between Pt particles and the modified support (sulfated alumina) can play
a crucial role in preserving the performance of the Pt/Al2O3 catalyst during the
extended SA decomposition reaction.

2.2. Introduction of vanadium (oxides) to copper catalyst for deactivation


resistance at low temperatures
Copper (bulk Cu, CuO, or CuFe2O4) has long been studied for the SA
decomposition reaction; however, owing to certain limitations, it is
a subideal solution for less expensive catalysts. Stable CuSO4 decomposes
only at high temperatures; thus, it is not operational over a wide temperature
range (550°C–800°C). Cu–V oxide has emerged as a competitive alternative to
platinum catalysts for the SA decomposition reaction.[61] The synthesis
method strongly influences the applicability of a specific reaction system.
Machida et al.[61] achieved a highly efficient substitute of Pt catalyst for SA
decomposition using a novel dissolution reprecipitation route. Transition
metal oxides and non-noble metals are of limited use in SA decomposition
because they are severely poisoned at low temperatures. A modified catalyst
was synthesized that operates at low temperatures (550°C–600°C) under SA
feed conditions. Briefly, Cu(NO3)2 and NH4VO3 were impregnated onto
three-dimensional ordered mesoporous SiO2 and calcined at 650°C, forming
copper pyrovanadate (Cu2V2O7, with a melting point of 780°C) both intern­
ally and externally. After high-temperature annealing at 800°C, Cu2V2O7
melted and mildly penetrated the mesoporous and homogeneous covering of
the cavity walls. After dissolution and reprecipitation, the mesoporous silica
was substantially transformed into a macroporous silica matrix. The thin Cu2
V2O7 coating on the inner walls stabilized the matrix at low temperatures,
preventing its deactivation in the presence of high water percentage.[61]
A morphological TEM analysis revealed uniform mesostructured channel-
like pores. As shown in Figure 4(c), Swiss cheese-like macropores were formed
throughout the sample. Despite the loss of small mesopores and the evolution
of the macroporous structure in Cu–V/SiO2 after heat treatment at 800°C, the
SO3 decomposition was well accomplished, as shown in Figure 4(d).
The performance of Cu–V was compared with that of a commercially avail­
able model catalyst, Pt/Al2O3 (Figure 4(d)). Upon initial exposure to SO3, P/Al2
O3 was strongly deactivated by Al2O3 sulfate formation and the deactivation
mechanism had been demonstrated in a previous study.[62] Masato et al. impreg­
nated Cu–V species in mesoporous silica walls using different heat treatments,
thus obtaining different pore structures. The calcination temperature of the
Cu–V/SiO2 catalyst, which was crucial for impregnation, significantly affected
the catalytic performance. Moreover, the aging temperature considerably
affected the pore size, i.e., mesopores to macropores. The catalyst calcined at
a low temperature (650°C) was deactivated shortly after SO3 contact. After
CATALYSIS REVIEWS 11

Figure 4. FE-TEM photographs of (a) as synthesized 3-D mesoporous SiO2, (b) Cu–V/SiO2@650, and
(c) Cu–V/SiO2@800. Arrows in panel b show the deposition of Cu–V oxide.(d) Catalytic SO3
conversion versus the reaction time from the beginning of H2SO4 supply. Cu–V/SiO2@650 (open),
Cu–V/SiO2@800 (closed). Reaction temperature: 600°C (○, ●) and 650°C (Δ, ▲). SO3 conversions for
0.5 wt % Pt/γ-Al2O3 measured at 600°C (···) and 650°C (–) are also shown. 14% SO3, 17% H2O, and
N2 balance, WHSV = 55.2 g-H2SO4 g-cat–1 h–1.Data reprinted from[61] copyright©(2012) with
permission from ACS publishers.

90 min, the stability and conversion (11%) of the catalyst calcined at a high
temperature (800°C) exceeded those of Pt/γ-Al2O3 (8% conversion). As Al2O3-
based catalysts are inactive at low temperatures, the catalytic performance was
evaluated after slightly increasing the reaction temperature (up to 50°C). The
Cu–V/SiO2 catalyst aged at 850°C converted 25% of the SO3, while the catalyst
calcined at 600°C and the Pt/γ-Al2O3 catalyst converted 7.0% and 18%,
12 H. A. KHAN ET AL.

Figure 5. Temperature programmed desorption of O2 from Cu2V2O7, CuO, and V2O5 measured in
a stream of He. Heating rate: 10°C min–1. Reprinted from.[63], copyright© (2013) with permission
from ACS publishers.

respectively.[63] This result confirmed the higher activity of the Cu–V/SiO2


catalyst calcined at 800°C than that of the Pt/γ-Al2O3 catalyst. The catalytic
transformation of SO3 to SO2 requires energy and (most importantly) lattice
oxygen vacancies in the catalyst material. SO3 decomposition has also been
evaluated on Cu2V2O7 (α, β) polymorphs. The α phase (blossite type) was stable
during the tests. In this polymorph, oxygen vacancies were prominent and
oxygen desorption was accompanied by charge compensation through the
reduction of Cu2+ to Cu+. Consequently, the number of oxygen deficiencies
was comparable to that of Cu16V16O55 at 600°C. The formation of oxygen
vacancies was confirmed using DFT. Figure 5 shows the O2-TPDs of Cu2V2
O7, CuO, and V2O5. V2O5 showed no noticeable O2 desorption up to its melting
point (675°C). However, at approximately 600°C, CuO exhibited O2 desorption
owing to Cu2O decomposition, although the O2 loss was negligible below 700°C.
In contrast, the O2 desorption from Cu2V2O7 was significant over 600°C.
Oxygen vacancies were favorably formed at the Cu–O–V bonding sites than
at the V–O–V sites in the pyrovanadate unit. The formation energy of oxygen
vacancies was considerably lower at the (100) surface than in bulk Cu16V16O56.
Takahiro et al. found that molten CuV2O6 efficiently catalyzes SO3 decom­
position at ~600°C. The SO3 decomposition over CuV2O6 was significantly
increased near the melting temperature of CuV2O6 (~630°C), surpassing those
of solid-phase Cu2V2O7 and other materials with slightly higher melting
temperatures (780°C), such as CuO–V2O5. CuSO4 is formed when SO3
CATALYSIS REVIEWS 13

Figure 6. Catalytic SO3 conversion over a) CuV2/SiO2 and b) CuV/SiO2 versus time-on-stream
during stepwise changes in reaction temperature. Equilibrium conversions are shown as dashed
lines. Reprinted from[65] copyright© (2011) with permission from RSC.

adsorbs onto the Cu oxide surface, and it is rapidly decomposed into SO2 and
O2 during contact with the molten phase. The molten catalyst also contained
a significant fraction of monovalent Cu, formed by the impulsive desorption of
O2 species.[64] Other Cu–V oxides with high melting points (Cu3V2O8 and Cu5
V2O10) were found to be less active than Cu2V2O7.[65]
Figure 6 shows the performances of different catalysts as functions of
temperature. The compared catalysts exhibited different behaviors at 650°C.
Low but sustainable SO3 conversion (<1.5%) was recorded for CuV/SiO2. The
SO3 conversion of CuV2/SiO2 increased over time, reaching 20 folds higher
than that of Cu2V2O7. Similarly, the conversion was remarkably enhanced
near the melting point temperature of Cu2V2O7. The catalyst, which remained
as solid-state CuV/SiO2 owing to its high melting point, delivered poor
performance, as also reported elsewhere.[63] Cu–V-based catalysts are promis­
ing for low-temperature applications.
Table 2 summarizes the published works on supported and unsupported Cu–
V catalyst systems. Highly surface-amorphous SiO2 is the most promising
support.[70] nCuO–V2O5 supported on SiO2 catalysts was prepared using dif­
ferent molar ratios of CuO (0–5) for SO3 decomposition. Depending on n, the
binary compounds, CuV2O6, Cu2V2O7, Cu3V2O8, and Cu5V2O10, were formed
on the three-dimensional mesoporous SiO2 surface. During thermal aging, these
compounds were incongruently melted and Cu2V2O7 was precipitated. The
pore structures were significantly changed due to high-temperature annealing;
14 H. A. KHAN ET AL.

Table 2. SO3 decomposition on based Cu-V supported and unsupported catalysts for low-
temperature application.
Onset
Catalyst(s) Reaction Conversion
researched Main conclusions Temperature Space Velocity (%) References
[65]
Cu2V2O7 Highest activity is achieved at (650 600 °C WHSV = 55.2 g-H2 18
°C). It is reported that SO4 (g-cat)−1 h−1
pyrovanadate framework is
resistant to sulfate formation
which makes it stable for SO3
decomposition.
[64]
CuV2O6 Molten phase Cu2V2O7 is more 630 °C WHSV = 55.2 g-H2 ~12
active as compared to CuV2O6 SO4 (g-cat)−1 h−1
[66]
CeV4 Bimodal mesoporous SiO2 was 600 °C WHSV = 3.6–110 g- 44
supported utilized to deposit Ce-V active H2SO4 g−1 h−1
on SiO2 component. 50 fold higher TOF
was achieved using said
strategy compared to
unsupported CeVO4 system.
[61]
Cu-V-O/SiO2 The macroporous catalyst showed 650 °C WHSV = 55.2 g-H2 25
promising activity at moderate SO4 (g-cat)−1 h−1
temperatures. It remained
stable for 100 h at a space
velocity of 55.2 H2SO4 g.cat−1.
h−1
[63]
Cu2V2O7 (α, β, The phase depend activity is - -
δ) investigated. The blossite type
(α-phase) is stable phase for SO3
decomposition. It is proposed
that oxygen desorption
accompanied by charge
compensation through the
reduction of Cu2+ to Cu+
produced an oxygen deficiency
corresponding to Cu16V16O55 at
600°C.
[67]
CuV2O6 Hydrothermally stabilized Cu and 650 °C WHSV = 55.2 g-H2 27
supported V over SiO2 shows promising SO4 (g-cat)−1 h−1
on 3D SiO2 activity when it is melted by
high temperature in the SiO2
porous network .
nCuO-V2O5 (n = 0,1,2,3,4 and 5) SO3 decomposition is 600 °C
/SiO2 structure
sensitive and
phase-sensitive,
catalyst
treatment at
different
temperature has
a strong
influence on SO3
decomposition.
Moles of CuO-V2
O5 effect was
[68] also studies.
WHSV = 55.2 g-H2SO4 15
(g-cat)−1 h−1
[69]
CuO-CeO2 The catalysts CeCu–Al2O3, 650 °C WHSV = 10.0 g-H2 40
CeCu–SiC–Al2O3 CeCu–SiC, and SO4 (g-cat)−1 h−1
CeCu–SiC–Al2O3 was prepared
to check the effect of supports.
CATALYSIS REVIEWS 15

Figure 7. SO3 conversion measured at 600°C of nCuO–V2O5/SiO2 (n = 0, 1, 2, 3 and 5) before and


after thermal aging at 700°C (n = 0), 750°C (n = 1), 800°C (n = 2, 3) and 850°C (n = 5). Data from
Kawada T et al.[68] Copyright© (2015) with permission from Elsevier.

however, a modified surface with a uniform coating of copper vanadate achieved


improved performance at low temperatures (~600°C). The catalyst with the
lowest melting point (n = 1) exhibited the highest activity (see Figure 7).[67,68]

2.3. Introduction of ceria (oxides) to copper catalysts for low-temperature


applications
Inexpensive alternatives to Pt-based catalysts are greatly desired. Zhang et al.
prepared complex metal oxides CexCu1_xO2_δ (x = 0.2–0.8) using the sol–gel
method performed the SA decomposition reaction at a space velocity of
5000 mL gcat−1 h−1 within a temperature window of 727°C–877°C).[71] At
temperatures above 800°C, the catalyst Ce0.8Cu0.2O2 calcined at 900°C showed
higher activity than the Pt catalyst. In a stability test, this catalyst remained
active for up to 60 h at 850°C. This facile synthesized CeO2 support and CuO
clusters showed synergistic effects. The reactive oxygen provided by the CeO2
support accelerated the rate-limiting step of SO3 decomposition, namely, the
decomposition of SO3 to SO2.[71] Figure 8 shows that the Ce0.8Cu0.2 catalyst
with a Ce:Cu ratio of 4:1 calcined at 900°C exhibited the highest catalytic
activity. The SO3 conversion values of the CuO/CeO2 catalysts increased with
an increase in the aging temperature, particularly in the reaction temperature
16 H. A. KHAN ET AL.

Figure 8. SO3 conversion ratio with complex oxides catalysts of CexCu1−x-900, CexCu1−x-700 or
pure oxide CuO-900, and CeO2-900, where x = 0.2, 0.5, and 0.8, −900 and −700 represents the
calcination temperatures of 700 and 900°C. Data from Zhang Y et al.[71] Copyright© (2015) with
permission from Elsevier.

range of 727°C–777°C. The SO3 conversion significantly increased with the


Ce:Cu ratios. Increasing the temperature of SO3 decomposition reduced the
effect of catalyst aging temperature on the sample activities, thus improving
their consistency. The SO3 conversion ratio of pure CeO2 rapidly increased at
temperatures higher than 777°C. The increased activity was attributed to the
high oxygen storage capacity of CeO2.
Su et al.[72] investigated SA decomposition on a CuCr–O2 catalyst prepared
using the coprecipitation method. They found that SA decomposition over the
CuCrO2 catalyst is structure sensitive. To optimize the structure of CuCrO4,
they carefully monitored the pH and aging temperature. Catalytic stability is
always a priority in applications requiring highly corrosive reactions. The
CuCr2O4 catalyst was relatively stable for up to 70 h. The activity decreased
during the first few hours, possibly because the reaction conditions were
unstable; thereafter, no severe deactivation was observed. The size of the
CuCrO4 crystals largely depended on pH and remained within a few nan­
ometers. The catalytic activity of CuCrO4 reasonably approached that of the
commercially available Pt/Al2O3 catalyst. Therefore, CuCrO4 is as a suitable
candidate for SA decomposition in the SI cycle.[72]
Wang et al.[69] evaluated the effectiveness of various supports for noble
metal-based catalysts. However, the support effects on transition metal-based
catalysts have not been extensively investigated. As the candidate supports for
CATALYSIS REVIEWS 17

Figure 9. SO3 conversion ratio with complex oxide catalysts supported by different carriers
Reprinted from[69] copyright© (2018) with permission from Elsevier.

SA decomposition, they selected SiC, Al2O3, SiC–Al2O3, and SiC–Al2O3 balls


with high surface areas. The authors reported a significant influence of support
material on the catalytic activity. The recorded activities are plotted in Figure
9. In the temperature range 600°C–700°C, the CeCu activity was lowest on the
SiC support, with a conversion of only 10% at 650°C. The activity was
improved on the Al2O3 support and a mixture of Al2O3 with SiC (Al2O3–
SiC). The net conversion of CeCu/Al2O3 was 20% at 650°C, higher than that of
CeCu on pure SiC. On the Al2O3–SiC support, CeCu conversion was highly
stable and reached nearly 60% at 650°C, close to the equilibrium conversion
value. Next, the catalysts were characterized using different techniques. The
catalyst performance was found to depend on the oxidation state of the
catalyst. X-ray photoelectron spectroscopy (XPS) data showed that SiC was
partially oxidized to SiO/SiOCx, consistent with common knowledge that
CeO2 is a strong oxygen carrier. In a temperature-programmed reduction
analysis, CeO2/SiC–Al2O3 showed the highest redox potential and CuO pre­
sented the lowest reduction temperature.[69] SiC and SiO2 supports are known
to resist acid corrosion. The results demonstrated the high activity and low
deactivation of CeCu supported on SiC–Al2O3. The reaction rate significantly
depends on the space velocity, and increasing the GHSV decreases the activity
and hastens the deactivation.
Cu- and Fe-based catalysts are considered to activate the SA decomposition
reaction. Banerjee et al.[34] attempted to optimize the catalytic activity of Fe2O4
catalysts. In previous literature, Fe2O4 has been reported as strongly unstable
18 H. A. KHAN ET AL.

Figure 10. Temperature-dependent catalytic activity profiles of the three ferrospinel catalysts for
H2SO4 decomposition reaction. Reprinted from[34] copyright© (2011) with permission from
Elsevier.

in acidic environments. To minimize the effect of acid poisoning and digestion


of the catalyst at high temperatures, Fe2O3 was doped with Co, Ni, or Cu.
CuFe2O4 demonstrated high activity (Figure 10), with a conversion of 78% at
800°C. Chemical states and redox properties are highly desirable for SO3
decomposition. Banerjee et al. probed the properties of their synthesized
catalysts using Mossbauer spectroscopy and other advanced techniques. The
proposed reaction mechanism over the copper ferrite catalyst was similar to
that described using Equations (5) and (6), namely, metal sulfate decomposi­
tion followed by oxygen release. Again, the metal sulfate decomposition was
the rate-determining step. The other catalysts (NiFe2O4 and CoFe2O4) were
moderately active. The activity of the doped catalysts decreased in the follow­
ing order: CuFe2O4 > NiFe2O4 > CoFe2O4. The authors claimed that the high
electronegativity of Cu in the mixed metal sulfates contributes to SO bond
dissociation. Cu-based ferrospines are characterized by the decomposition
temperature and high reducibility of low metal sulfates; consequently, they
are stably active in SA decomposition.[,34,73] The same authors compared the
catalytic performances of Fe1.8Cr0.2O3 and a model Fe2O3 catalyst. Fe1.8Cr0.2
O3 was highly stable for up to 100 h, achieving an SA flow of 0.63 mL/min in
CATALYSIS REVIEWS 19

the test reactions. Moreover, the stable activity of Fe1.8Cr0.2O3 exceeded that of
pure Fe2O3.
The activities of different catalysts are compared in Table 3. Here, the
literature data on different noble metals are tabulated for a convenient com­
parison of various results.

2.4. Significance of silicon carbide in resisting support-induced deactivation

Silicon carbide (SiC) is an inexpensive material with excellent corrosion and


thermal shock resistance, remarkable thermal conductivity, and high radiation
absorbance. Accordingly, it is a promising candidate for H2SO4 decomposi­
tion. Alberto et al.[81] evaluated the performance of an iron(III) oxide catalyst
supported by siliconized silicon carbide (SiSiC) for the thermochemical split­
ting of 96 wt% SA. They monitored the influence of weight hourly space
velocity (WHSV = 12.3–17.9 h−1) and reaction temperature (800°C–875°C)
on the catalytic activity during a prolonged reaction (100 h). The catalyst
presented acceptable activity (nearly 80%) with negligible deactivation during
the long run. The spent catalyst was characterized, and slight sulfate species
were observed on the surface of the spent catalyst, confirming that intermedi­
ate species were formed and the reaction mechanism was governed by decom­
position sulfates.[81]
Karagiannakis et al.[13,78] developed a series of catalysts supported on
monolithic SiSiC honeycomb structures. The catalysts were Fe2O3 and CuO
as well as mixed oxides of Cu–Fe, Fe–Cr, Cu–Al, and Cu–Fe/Al. The catalyst
performances were assessed in a 100-h SA decomposition reaction. Under the
reaction conditions (850°C, 5–35-h−1 WHSV), all catalysts exhibited high
conversion (≥70%) with only slight activity loss (5%–10%) during the 100-h
reaction. Post-reaction characterization of the catalysts showed extensive
phase separation and particle sintering in the Cu-contained mixed oxides.
During the 100-h reaction, the SiC-based Fe–Cr-mixed oxide catalyst achieved
significant activity (close to that of Pt/Al2O3) and remarkable stability. As
confirmed by XRD results, the chemical composition of this catalyst was
unchanged after the reaction. TGA analysis recorded a maximum weight
loss of 3.5% in the samples, attributed to the in situ formation and decom­
position of iron and copper sulfates. Sulfates were detected in the XRD
analysis, strongly evidencing the involvement of sulfate species in the reaction
mechanism. However, whether sulfate formation was linked to SiC or Fe/Cr
species was not clarified.
Owing to its high thermal conductivity and poisoning resistance, SiC has
garnered significant interest. Lee et al.[22] prepared a Pt–SiC–Al2O3 catalyst
by coating an alumina support with SiC via chemical vapor deposition,
followed by Pt impregnation. When reacted at 650°C–850°C for 6 h, the
fabricated catalyst achieved superior stability and activity than the Pt–Al2O3
20

Table 3. Summary of sulfuric acid decomposition catalyst and its catalytic performance at various conditions.
Stability test Space velocity
Catalyst(s) Main conclusions Duration (if given) References
[74]
G-5 and P-25 TiO2 Pt supported It was studied that the anatase phase stabilizes the Pt nanoparticles and anatase phases were resistant 100 h 46000–144000 ml
1 wt% to metal sulfate formation. TiO2 anatase keeps the stable phase of Pt, PtO at low and high g−1h−1
temperature (600–850) which makes this catalysts best choice for wide temperature range.
[29,27]
Pt supported on SBA-15, SBA-SiC Pt was supported on SBA-15, and SBA-15 template was used to make SiC with high surface area and 100 h 46000–144000 ml
-15 well-defined pore. Pore disintegration and phase change during the reaction, which stabilized the g−1h−1
H. A. KHAN ET AL.

Pt nanoparticle from sintering, and lead, which resulted in high stability. SBA-15 Functionalizing
approach was also introduced, make made SBA-15, higher stable and remained stable for 100 h.
[26]
Pt supported on SiO2, hollow SiO2 The new strategy implemented to synthesis controlled shell thickness approach to make Pt hollow 100 h 46000–144000 ml
linings of SiO2. Catalyst remained active for a long duration. g−1h−1
[20]
Pt supported on Ta2O3 Pt supported on Ta2O3 shows promising stability as compated to othe SO3 resistant supports. The 2000 h 22 − 880 g of H2
catalyst remained active for 1800 hour atr low temperature of 600 °C (activity loss ≤ 1.5% per 1,000 SO4 g−1 h−1.
h).
[27,,22,,75]
Pt supported on SiC, Pt-SiC polyol SiC was pre-treated with strong acids (HNO3&H2SO4). Pt was impregnated on modified surface charge 10–30 h 46000–144000 ml
and impregnation supports. Wet impregnation and EG polyol method were used to impregnate. EG synthesis method g−1h−1
stabilized the catalyst and higher activity at 650 and 850 C)
[30]
SiC phase change and Pt Initially, Hollow SiC was prepared, and Pt was impregnated using Polyol method. Pt size was 50 h 46000–144000 ml
supported inside hollow sphere controlled to ultra-small size (1–3 nm). hSiC was oxide to bulk SiO2 phase, and it helped to offer Pt g−1h−1
oxidation states PtOx which was considered as active phase.
[34]
Fe2O3 supported on SiO2 It was found that supported Fe2O3 was more active and stable as compared to unsupported at mash 100 h 27 g acid g−1
size effect was very prominent for SO2 decomposition. The stability test was performed at 800 °C catalyst h−1
over 20–45 mesh size
[31]
Cu-, Fe- & CuFe-Al2O3 composites Cu showed better activities as compared to Fe. It is proposed that, Cu (650 °C) has lower sulfate 10 h
decomposition temperature as compared to Fe (780 °C) which is intermediate step.
[36]
Fe2O3, Fe1.8Cr0.2O3, Fe1.6Cr0.4O3 In repeated redox cycle, Fe1.8Cr0.2O3 showed less deactivation and increased reproducibility . – 1200 ml g−1h−1
Regeneration was enhanced.
[35]
CoFe2O4, NiFe2O4, CuFe2O4 CuFe2O4 concluded as most promising catalyst, depending on lowest sulfate decomposition – 0.06 ml.min−1
temperature.
[18]
Ag-Pd alloy & Fe2O3 on SiO2 Alloy and Fe2O3 showed higher activity with overall decrease in initial decomposition temperature 18 h 25–100sccm.
support with 300 °C. Fe2O3 was stable up to 16 h. min−1
[76]
FeTiO3, MnTiO3, NiFe2O4, CuFe2 Complex metal oxides recorded activity was in order: 2CuOCr2O3 > CuFe2O4 > NiCr2O4 = NiFe2O4 160 h 50 g acid g−1
O4, NiCr2O4, CuOCr2O3 > MnTiO3 = FeTiO3. Both 2CuOCr2O3 and NiCr2O4 leached Cr into the sulfur dioxide. FeTiO3 was catalyst h−1
unstable at higher temperature (>850°C).
[77]
Pt, Fe2O3, V2O5, CuO, MnO2, Cr2 It was noted that Pt was most active, and order of activity for tested catalyst was: Pt > Fe2O3 > V2O5 10000 ml g−1h−1
O3, CeO2, CoO, ZnO, Al2O3 > CuO > MnO2 > Cr2O3 > CeO2 > CoO > ZnO > Al2O3.
(Continued)
Table 3. (Continued).
Stability test Space velocity
Catalyst(s) Main conclusions Duration (if given) References
[,13,78]
Iron oxides, Al2O3, CuO, Cr2O3, Cu (650 °C) has lower sulfate decomposition temperature as compared to Fe (780 °C). It is noted that – –
Pt/Al2O3 addition of Al, Cu & Cr to Fe structure enhances conversion.
[79]
Pt, V2O4, Cr2O3, Fe2O3, NiO It is recorded that, Pt > V2O4 > Cr2O3 > Fe2O3 > NiO. Ru, V2O4 & Cr2O3 as order of activity. Among all, – –
Cr2O3 showed poor reproducibility and decreased activity.
[80]
Fe2O3 supported by Al2O3 Catalyst deactivated over period of 20 h. The deactivation was proposed as result of sulfation of the 25 h 0.08 ml.min−1
Al2O3 support material.
[10]
NiO, CoO, SiO2, Al2O3, ZnO, Cr2O3, Catalytic activity per unit weight of catalyst is in order: Fe2O3 > V2O5 > CuO > Cr2O3 > Co3O4 > TiO2 150 h 1.8 ml.min−1
V2O5, CuO, Fe2O3, MnO, TiO2 > ZnO > MnO2 > NiO > SiO2 > Al2O3.
[81]
Fe2O3 (pellets) or on SiSiC support SiSiC is sulfate resistant and good thermal conductor, Fe2O3 supported on SiSiC showed good catalytic 100 h 6.0–7.9 h−1
activity (approximately 80% of equilibrium conversion) with negligible deactivation after 100 hrs.
[66]
Pt on Al2O3, TiO2 & ZrO2 supports TiO2 is also resistant to sulfate formation. It is noticed that Pt on TiO2 was stable, however lower Pt and 250 h for Pt/TiO2 52 g acid .g−1
TiO2 interaction resulted in Pt loss. Al2O3 & ZrO2 supports showed good activity at 850°C, but catalyst .h−1
activity loss at lower temperature (800°C).
[82]
Fe2O3, V2O5 and Cr2O3 on supports High temperature volatility of V2O5 andCr2O3 makes them feasible for cyclic oxidation and reduction – –
of SO2. Fe2O3 most active and stable as compared to other candidates.
[24]
Pt on TiO2 support Platinum loss and support interaction were major factor for Pt loss and activity decrease during long 548 49.5 g acid g−1
time testing. catalyst h−1
[17]
Pt, Pd, Rh, Ir, Ru supported on TiO2 Platinum is most stable and active for SO3 decomposition. 230 g acid g−1
catalyst h−1
[83]
Proprietary catalysts – –
[12]
Fe supported by Al and Ti. Fe: Fe loading as significant role and it increased catalyst performance. No evidence of Fe sulfate detected – 72000 ml gcat−1.
support = 25%, 50%, 75% & on the 850°C, but was detected on the 550°C sample. Fe-Ti based system could work at low h−1
100% temperature of 700 °C, while Fe-Al are active above 800 °C.
[59]
Fe2O3 on Al2O3 support Fe2O3 demonstrated good activity at high temperatures (900°C) and low temperature deactivation is 0.146–0.731
because of aluminum sulfate formation. kmol/kg.h
[21]
Pt supported by BaSO4 High dispersion of Pt on BaSO4 by modified synthesis methods, improved catalytic performance. 72000 ml gcat−1.
h−1
[84]
Pt on SiSiC, Fe2O3 on SiSiC, blank When Fe2O3 is competed with Pt, It is found that Pt showed high catalytic activity even at low – 0.8–6 ml g−1.
SiSiC residence times (0.2 s). h−1h−1
[85]
Supports: Al2O3, SiO2, CeO2, TiO2, SiO2 and TiO2 is resistant to metal sulfate formation and it showed that Pt supported on Al2O3 – –
ZrO2, BaSO4 supports failed due to sulfate formation (Al2(SO4)3 poisoning). ZrO2 and BaSO4 supports
deactivated because of sulfate formation.
CATALYSIS REVIEWS

[86]
Fe2O3, CuFe2O4 Study focused more on solar reactor design and CuFe2O4 is more active in this study. – 2–6 ml gcat−1.h−1
[73]
ZnFe2O4, NiFe2O4 No significant data for stability studies. Synthesis methods were studied. – 1200 ml. gcat−1.h−1
[15]
Fe2O3, Cr2O3, Al2O3, CeO2, NiO, Pt, Fe & Cr activities were similar > 850°C. CeO2 did not show good activity. CuO2 is active above 700 ° – 4300 h−1
CuO C.
21
22 H. A. KHAN ET AL.

catalyst. The Pt/SiC–Al2O3 catalyst converted 71% of SO3 at 850°C.


Moreover, although aluminum sulfate was formed at low reaction tempera­
tures, the formation of sulfate species was considerably lower on Pt/SiC–Al2
O3 than on bare Pt–Al2O3 because it was largely suppressed by SiC coating
on the alumina of Pt/SiC–Al2O3.
The effectiveness of SiC against bulk sulfating is unarguable; however, its
corrosion resistance must be improved. Lee et al.[22] reported that SiC is an
active support material for SA decomposition; however, SiC–Al did not
eliminate that formation of sulfate species. Noh et al.[23] synthesized Pt on
a nanoSiC (Pt/n-SiC) support and operated the catalyst at 72 L/g for 6 h in the
temperature range 650°C–850°C (Figure 11) Error! Reference source not
found. The highest conversion (79%) was recorded at 850°C. A Pt catalyst
prepared on a commercial SiC support (acquired from SICAT Catalyst Inc.)
converted 64% of the SA under similar reaction conditions. XPS results of used
samples suggested that Si oxides were observed on both catalysts, indicating
that both surfaces (n-SiC and commercial SiC) were partially oxidized during
the reaction. However, the effect of Pt was less severe on the commercial SiC
than on the nanosilica, and the Si oxides were weaker on the commercial
support.
SiC has been considered as an alternative to other supports, such as Al2O3,
CeO2, ZrO2, and CuO, in a broader temperature range (600°C–900°C). In
previous studies, SiC was oxidized to SiO2; however, the resistance to metal

Figure 11. Sulfuric acid decomposition with respect to reaction time at 650°C (Δ, ▲), 750°C (○, ●),
850°C (□, ■) and a GHSV of 76,000 mL/gcat/h: open symbols for Pt/SiC catalysts and filled symbols
for Pt/n-SiC. Data from Seonchel Noh et al.[23] (2014)© with permission from Elsevier.
CATALYSIS REVIEWS 23

sulfate formation makes this support superior to others. After further


detailed investigations, the material properties were tuned to enhance the
surface area with well-defined morphology and the stability issues were
elucidated over extended times. Khan et al.[27] supported a Pt catalyst on
SBA-15-templated mSiC-15 for SA decomposition. The mSiC-15 support
was derived by carbonizing the SBA-15 template. Their catalyst exhibited
excellent stability at 850°C during a 50-h reaction, and no deactivation was
observed at low temperatures (~650°C). To validate the support efficiency of
SiC and the SBA-15 template, the authors compared the conversion profiles
of their catalyst with that of Pt on SBA-15. Pt/SBA-15 (reference catalyst)
was severely deactivated after only 6 h of reaction. Notably, SBA-15 (SiO2)
resists sulfate formation; however, the plausible deactivation cause was oxide
formation and Pt sintering at high temperatures, which reduce the Pt con­
tent. The used Pt/mSiC-15 catalyst was characterized using XRD and XPS.
Results revealed that bulk mSiC was fully oxidized to m-SiO2 after a 12-h
reaction. The BET surface area also deteriorated during the phase change,
and the pore structure was severely disintegrated. Nevertheless, the Pt
particles were presumably stabilized owing to strong encapsulation in the
SiOC surface; consequently, the Pt loss was reduced and the Pt/mSiC-15
catalyst was highly stable.
The SO3 dissociation mechanism on the Pt surface was suggested as
follows[87 88}:
First, SO3 decomposes into SO2 and oxygen atoms.
SO3 SO2 þ O (7)
In the intermediate step, oxygen generated using Equation (7) engages in
oxygen evolution from the SO3 molecules.
SO3 þ O SO2 þ O2 (8)
The reaction scheme of SO3 decomposition on Pt can be suggested as follows:
SO3 þ Pt ! SO2 þ Pt O (9)

Pt O þ SO3 þ Pt O ! SO2 þ O2 (10)

Pt O þ Pt O O2 (11)
The S–O bonds become weak when SO3 adsorbs on the Pt surface (Equation
9). The SO3 molecules may then dissociate through an unstable intermediate
state of adsorbed SO2 and O2. The reaction proceeds similarly in Equations (7)
and (8), which together constitute the rate-determining step.[88]
The desirable characteristics of the SiC support were further exploited by
depositing ultrasmall Pt particles on the inner walls of hollow SiC spheres with
thick-core walls. During SA decomposition at 850°C, the initial conversion of
24 H. A. KHAN ET AL.

Pt deposited on the hollow SiC catalyst was approximately 80%. The Pt/hSiC
catalyst was highly stable throughout the 100-h reaction. Moreover, the cata­
lyst delivered stable performance at high space velocity (GHSV = 120,000 mL
gcat−1 h−1, twice that of the standard GHSV, i.e., 48,000 mL gcat−1 h−1). In the
characterization analysis, the Pt/hSiC in the spent catalyst was entirely con­
verted to Pt/hSiO2 after 6 h, which was consistent with previous
reports.[22,23,29,70] The conversion was attributed to oxidation during the
reaction, the breakdown of the shell wall, and the oxidation of Pt particles
on the outer walls of the hollow sphere. In contrast, the Pt particles on the
inner walls of the hollow sphere were quite resilient to the harsh reaction
conditions and exhibited no particle loss or sintering.
Figure 12 presents the TEM-EDX results of the Pt nanoparticles dispersed
in the inner and outer spaces of the thick-cored hollow SiC sphere. Clearly, the
particles were pushed owing to the pressure exerted by the gas molecules and
the high-temperature mobility of PtO to the inner core. Therefore, fabricating
active metal particles on the inner walls of hollow spheres is a promising
strategy for SA decomposition without sintering and leaching. The hollow
sphere behaves as nanoreactors, and the support morphologies and catalytic
activities were retained at high space velocities. The reaction conditions
strongly affected the BET pore size and structures as compared to pristine
catalysts. Pore structure modification due to high temperature and oxidative
environment resulted from sintering of SiC shells, since it is known that

Figure 12. TEM particle size distribution and TEM elemental mapping and EDX intensity line graph
of (a–a2) pristine Pt/hSiC, (b–b2) spent Pt/hSiC after 6 h reaction, and (c–c2) spent Pt/hSiC after
30 h reaction at GHSV of 76,000 mL/gcat/h. Data from Hassnain et al.[30] copyright© (2018) with
permission from Elsevier.
CATALYSIS REVIEWS 25

sintering temperature is 0.8 to melting temperature. SiC partially oxidizes to


SiO2/SiOC at 850°C, and initiates sintering as well. This phase transition and
morphological changes have an effect on particles, impregnated inside the
pore of hollow SiC. It is reported that sintering was prevented by trapping
ultrasmall particles into the shell regions of hollow SiC and newly converted
SiO2 phase. As shown in the EDX line graph analysis, the Pt particles were
densely accumulated in the shell regions after the reaction (Figure 12).
SiC-based support materials weakly interact with noble metals (low metal–
support interactions). The metal–support interaction plays a vital role in
catalyst stability. PtOx is volatile, implying that the metal is lost because of
evaporation, sintering, and leaching at high temperatures. To improve this
situation, the metal–support interaction was modified by applying opposite
charges to the support surface. When metal ions react with an oppositely
charged support surface, the interaction is strengthened. Khan and Jung
extensively treated SiC with concentrated acids (HNO3 and H2SO4), achieving
Pt/SiC catalyst with improved catalytic performance supported by commercial
SiC.[27] The strong acid treatment modified the surface change, which was
experimentally confirmed through zeta potential analysis. The impregnated Pt
ions were generated using an oppositely charged platinum precursor. The
acid-treated supported catalyst was highly active and stable over a prolonged
duration than the untreated catalyst, and a severe metal loss was observed.

2.5. Role of support materials against sintering- and leaching-based


deactivation
The catalyst performance depends on its morphology, pore structure, and
composition.
Fe2O3 and Fe doped with Cr, Cu, and Ni are less expensive than noble
metals such as Pt and Pd. However, transition metal alternatives are chal­
lenged by the formation of intermediate metal sulfates. The advantages of Fe-
and Co-based supports have been emphasized in several studies. The SO3
decomposition over these alternative catalysts proceeds by representative
sulfate formation and dissociation of sulfate to SO2 and O2. Among many
support materials explored in the literature, inert and highly stable SiO2
modifies the fate of Fe2O3 in the SA decomposition reaction. Accordingly,
an Fe catalyst on an SiO2 support (5 wt% Fe/SiO2) has been prepared and
tested. The Fe phase exhibited an unusual crystallinity loss in the Fe–O–Si
arrangement. Under aggressive gaseous SA at 800°C, Fe2O3 sintering was
inhibited by the SiO2 support. Surprisingly, the activity of Fe2O3 nanoparticles
remained stable for up to 100 h.
Incorporating Fe2O3 in the SiO2 matrix significantly improved the catalytic
stability.[34] Unsupported Fe2O3 performed poorly, confirming the stabilizing
role of the support.[66] In the absence of the support matrix, Fe2O3 was severely
26 H. A. KHAN ET AL.

deactivated and sulfate species were observed in the TGA and TPD
experiments.[89] However, on an Al2O3 support, nobel metals, iron oxide,
copper oxide, and copper ferrite were severely deactivated at low temperatures
due to stable sulfate formation. The silica morphology and surface charge are
crucial for stabilizing the nanoparticles to extend their reaction times and
improve their stabilities. Recently Pt@mSiO2 has been reported by Khan et al.,
as post-stable catalyst for SA decomposition. core shell structures play sig­
nificant role in mini minimizing surface migration and sintering which makes
catalytic materials suitable for high-temperature applications.[90]
CeO2-based catalysts are affected by the reaction conditions. As stable
sulfate species cover the active sites, the catalyst eventually deactivates to
noticeable levels over time. Such catalysts are inapplicable at low
temperatures.[91] ZrO2-based materials show higher affinity and dispersal
ability of noble metals than CeO2-based catalysts. However, ZrO2-supported
catalysts are severely damaged and quickly deactivated in harsh reaction
environments.[66]
Recently, Khan et al.[74] reported a strong metal–support interaction of Pt
on anatase and rutile TiO2 supports. Anatase-TiO2 supports are competitive
candidates for SA decomposition. TiO2 resists the metal sulfate formation at
low temperatures and minimizes the metal loss due to Pt sintering, the
leaching of volatile material, and high feed flow rates at high temperatures
(850°C–900°C). Similar results for anatase TiO2 at low temperatures were
reported by Nur et al. They reported higher activity and stability of the anatase
phase than of the rutile phase.[25] Tiny Pt particles on anatase TiO2 were
observed in TEM images. The Pt particles were incorporated into the TiO2
lattice, as observed by the overlapped Pt and TiO2 in the EDX mapping. The
particles were likely retained by the bulk phase of TiO2, and were apparently
difficult to isolate.

2.6. Effect of preparation methods on the activity and stability of Pt-based


catalysts

The synthesis method significantly affects the performance of catalysts.


A catalyst for heterogeneous catalysis is usually synthesized using the wet
impregnation method. Catalysts prepared using the polyol method have
been least studied for high-temperature applications but are widely used in
electrochemical and photochemical reactions. For catalysts in the SA decom­
position reaction, the charge on the support surface is usually determined
using the zeta potential. Khan and Jung[28] introduced surface charge by
treating the commercial SiC-extruded support material with concentrated
nitric and SA mixtures at 80°C. After surface modification, the dispersion
and metal–support interaction of the impregnated Pt particles significantly
improved and the Pt particle sizes remained smaller (3–7 nm). The positively
CATALYSIS REVIEWS 27

charged Pt particles interacted with the negatively charged support molecules,


and the interaction was strengthened by thermal treatment. After the partial
oxidation of the SiC support during an intense acid treatment and further
calcination, a stable catalyst was formed. The polyol metal deposited on
a modified SiC support stabilized the Pt particles in the support material,
realizing tiny Pt nanoparticles. The two catalysts synthesized on the surface-
modified SiC supports using different methods (wet impregnation and polyol
deposition) achieved different conversions in SA decomposition. In particular,
the catalyst synthesized using the polyol deposition method was highly active
and stable in prolonged reactions.
Similar behaviors were reported for Pt/hSiC catalysts. Here, a hollow SiC
support was used with no surface modification. The polyol method yielded
ultrasmall nanoparticles in the hollow SiC sphere. When the catalyst was
activated at high temperatures in N2 atmosphere, the particles stabilized in
hollow SiC and did not leach during the reaction. Thermal treatment and
exposure to SA modified the pore structure of the support material but
retained the metal particles in the hollow region. The role of hollow-sphere
supports as nanoreactors are widely accepted. The hollow spheres retain the
reactants and release the products. During this sequential mass transfer, the
metal is not directly affected by the intense pressure on the reaction and gas-
product particles.
Pt was deposited on polystyrene using the modified impregnation method.
The charge on the polystyrene spheres was modified using silver nitrate. The
modified charges on the solid surface strongly attract the oppositely charged Pt
ions and distribute them uniformly across the surface. The Pt-loaded poly­
styrene sphere was then coated with a SiO2 precursor (TEOS). This synthesis
method was designed to capture Pt inside the silica sphere. Under high-
temperature treatment, the polystyrene sphere is decomposed and its depos­
ited Pt particles are trapped inside the hollow SiO2. In SA decomposition, SO3
can decompose on the active Pt in the inner walls of the hollow spheres.
Because the walls of the hollow spheres suppressed Pt sintering and the thick-
silica matrix did not disintegrate, this synthesis strategy minimized Pt losses.
The catalytic activities of silica spheres with internally and externally attached
Pt particles (HMSS-Pt and Pt-HMSS, respectively) were then compared. The
catalyst with Pt nanoparticles encapsulated in the hollow silica sphere was
highly active and stable over extended times.[26]
Khan et al.[28] also synthesized a Pt-SBA-15-based catalyst for high-
temperature SA decomposition reactions. Pt/SBA-15 prepared using the
impregnation method achieved high activity in both low and high-
temperature ranges (650°C–850°C). Although the Pt particles were well
dispersed in Pt/SBA-15, the low affinity of the tiny nanoparticles caused
fast degradation at a high temperature (850°C). The authors concluded that
the synthesis method used to fabricate the Pt/SBA-15-impregnated catalyst
28 H. A. KHAN ET AL.

did not adequately utilize the hexagonal pore symmetry of SBA-15. The high
surface area ensured good particle dispersion; however but a strong metal–
support interaction was lacking. This limitation was addressed by modifying
the synthesis method. One-pot methods are considered as the most efficient
approaches for entrapping silica particles. The chemical nature of silica can
be modified by adding MPTMS at the initial stage, which facilitates the
chemical interaction between Pt nanoparticles and their supporting
materials.[28]

3. Remarks and proposed research directions


The SA decomposition reaction faces several challenges. After reviewing the
available literature, we attributed the low performance and severe deactivation
of catalysts to three main factors:
(i)The high-energy barrier required for detaching the oxygen atoms from
SOn (where n = 1, 2, or 3)
(ii)The removal of decomposed or intermediate products of SA, such as
atomic O, S, and SOn components, from the catalyst surface
(iii)Most importantly, the loss of active metal due to the digestion of metals
in acids and leaching owing to the high flow rates of feed gases.

Nanoscale phenomena (the detachment of SOn, atomic oxygen, and sulfur) collectively
influence the reaction behavior; accordingly, an atomistic study of the detachment
process is necessary for understanding and preventing deactivation.

Pt and Pt-group metals have shown desirable catalytic activities and stabilities
under various conditions at 600°C–900°C. The major challenges of Pt-based
catalytic systems are a loss of active components and poisoning due to
unwanted byproducts. For example, Pt on Al2O3 forms sulfate species that
affect the neighboring active metal particles. Moreover, sintering and surface-
mediated ripening in the gas phase have been reported over the surfaces of
SiO2 and TiO2.
The sintering phenomena are explained by the transport of Pt atoms and
small clusters (Ostwald and Smoluchowski mechanisms). Sintering occurs
when the particle perimeter contacts the support (forming a triple-phase
boundary). In contrast, the gas phase involves the entire particle surface,
particularly in PtOx formation. The encapsulation of Pt in SiO2 and TiO2
suppresses the sintering of Pt particles but does not prevent Pt loss (the
measured losses were approximately 16%). The modifications of nanoparticle
structures, such as core–shell synthesis and alloying, can hinder the sintering,
ripening, and leaching of active key components.
CATALYSIS REVIEWS 29

3.1. Preliminary investigations of catalyst design

Phase transitions significantly affect the decomposition efficiency of catalysts.


In the literature, the decomposition process is usually simulated using empiri­
cal formulas and single-step transitions; however, such models typically
neglect the H2SO4 phase transitions during the boiling step. For instance,
during the thermal splitting process, boiling SA separates into multiple
intermediates.
In the thermal dissociation of SA, H2SO4(l) dissociates into H2O(g) and SO3
(g) at approximately 345°C. The gas-phase species include the well-known
species and several others, i.e., H2SO4(g), SO3(g), SO2(g), O2(g), H2O(g),
SO(g), and H2S(g). Other gas phase species, such as OH, H, HSO, HO2, HS,
H2, H2SO, HSOH, H2O2, and H2S2, are possible. Gaseous coordination com­
plexes, such as SO3–H2O, SO2–2H2O, SO3–3H2O, H2SO4–H2O, and H2SO4
–2H2O, should also be considered in thermodynamic studies. These gaseous
complexes are closely linked adducts among monomeric molecules.

3.2. Catalyst resilience

The biggest technological barrier in sulfur iodine cycle commercialization is


catalyst durability in the operable temperature window of 500°C–900°C. To
investigate this problem, researchers must use the latest simulation tool, which
considers two key catalyst components (active metal and support materials).
Parametric investigations must be performed on the interactions among metal
particles (the active component) and its support, particle wettability on the
support (miscibility of phases), phase stability (oxidation), atomic and cluster
mobilities across the surface, and catalyst dissolution (leaching).
To improve the economic viability of SO2 decomposition reaction,
researchers must investigate the use of inexpensive and easily available mate­
rials. The SI cycle can dually meet current energy needs and harvest environ­
mental SO2 (a hazardous gas), thus protecting our ecosystems.
In the present review, the key findings of catalytic systems for the SA
decomposition reaction are summarized. Systems with various preparation
parameters, such as support materials, reaction times and temperatures,
and active metals, were compared. The post-reaction characterizations of
catalysts prepared using various methods were also discussed. Based on
the previous results and their shortcomings, several suggestions for the
design of stable catalysts and the commercial feasibility of catalytic sulfu­
ric decomposition are provided.
30 H. A. KHAN ET AL.

Acknowledgments
The authors would like to acknowledge the research support provided by Prof Kwang Deog Jung
at Clean Energy Research Center, Korea Institute of Science and Technology, South Korea.

Disclosure statement
All the figures and tables are reproduce with permission from original source.

ORCID
Hassnain Abbas Khan http://orcid.org/0000-0003-2779-638X
Ahsan Jaleel http://orcid.org/0000-0002-3750-8356
Umair Hassan Bhatti http://orcid.org/0000-0003-3565-7582

References
[1] Rosen, M. A.; Koohi-Fayegh, S. The Prospects for Hydrogen as an Energy Carrier: An
Overview of Hydrogen Energy and Hydrogen Energy Systems. Energy, Ecol. Environ.
2016, 1(1), 10–29. DOI:10.1007/s40974-016-0005-z.
[2] Onuki, K.; Kubo, S.; Terada, A.; Sakaba, N.; Hino, R. Thermochemical Water-Splitting
Cycle Using Iodine and Sulfur. Energy Environ. Sci. 2009, 2(5), 491–497. DOI: 10.1039/
b821113m.
[3] Brecher, L. E.; Spewock, S.; Warde, C. J. The Westinghouse Sulfur Cycle for the
Thermochemical Decomposition of Water. Int. J. Hydrogen Energy. 1977, 2(1), 7–15.
DOI: 10.1016/0360-3199(77)90061-1.
[4] Jung, Y. H.; Jeong, Y. H. Development of the Once-through Hybrid Sulfur Process for
Nuclear Hydrogen Production. Int. J. Hydrogen Energy. 2010, 35(22), 12255–12267.
DOI: 10.1016/j.ijhydene.2010.07.168.
[5] O’Keefe, D.; Allen, C.; Besenbruch, G.; Brown, L.; Norman, J.; Sharp, R.; McCorkle, K.
Preliminary Results from Bench-Scale Testing of a Sulfur-Iodine Thermochemical
Water-Splitting Cycle. Int. J. Hydrogen Energy. 1982, 7(5), 381–392. DOI: 10.1016/
0360-3199(82)90048-9.
[6] Beghi, G. E. A.;. Decade of Research on Thermochemical Hydrogen at the Joint Research
Centre, Ispra. Int. J. Hydrogen Energy. 1986, 11(12), 761–771. DOI: 10.1016/0360-
3199(86)90172-2.
[7] Pen˜a, M. A.; Gómez, J. P.; Fierro, J. L. G. New Catalytic Routes for Syngas and
Hydrogen Production. Appl. Catal. A Gen. 1996, 144(1), 7–57. DOI: 10.1016/0926-
860X(96)00108-1.
[8] Favuzza, P.; Felici, C.; Nardi, L.; Tarquini, P.; Tito, A. Kinetics of Hydrogen Iodide
Decomposition over Activated Carbon Catalysts in Pellets. Appl. Catal. B Environ. 2011,
105(1–2), 30–40. DOI: 10.1016/j.apcatb.2011.03.032.
[9] Terada, A.; Iwatsuki, J.; Ishikura, S.; Noguchi, H.; Kubo, S.; Okuda, H.; Kasahara, S.;
Tanaka, N.; Ota, H.; Onuki, K.;, et al. Development of Hydrogen Production Technology
by Thermochemical Water Splitting IS Process Pilot Test Plan. J. Nucl. Sci. Technol.
2007, 44(3), 477–482. DOI: 10.1080/18811248.2007.9711311.
CATALYSIS REVIEWS 31

[10] Dokiya, M.; Kameyama, T.; Fukuda, K.; Kotera, Y. The Study of Thermochemical
Hydrogen Preparation. III. An Oxygen-Evolving Step through the Thermal Splitting of
Sulfuric Acid. Bull. Chem. Soc. Jpn. 1977, 50(10), 2657–2660. DOI: 10.1246/bcsj.50.2657.
[11] van der Merwe, A. F.;. Reaction Kinetics of the Iron-Catalysed Decomposition of SO3.
2014, April, 89. Dissertation submitted in fulfilment of the requirements for the degree
Magister in Chemical Engineering at the Potchefstroom Campus of the North-West
University.
[12] Kim, T. H.; Gong, G. T.; Lee, B. G.; Lee, K. Y.; Jeon, H. Y.; Shin, C. H.; Kim, H.;
Jung, K. D. Catalytic Decomposition of Sulfur Trioxide on the Binary Metal Oxide
Catalysts of Fe/Al and Fe/Ti. Appl. Catal. A Gen. 2006, 305(1), 39–45. DOI: 10.1016/j.
apcata.2006.02.052.
[13] Karagiannakis, G.; Agrafiotis, C. C.; Zygogianni, A.; Pagkoura, C.;
Konstandopoulos, A. G. Hydrogen Production via Sulfur-Based Thermochemical
Cycles: Part 1: Synthesis and Evaluation of Metal Oxide-Based Candidate Catalyst
Powders for the Sulfuric Acid Decomposition Step. Int. J. Hydrogen Energy. 2011, 36
(4), 2831–2844. DOI: 10.1016/j.ijhydene.2010.11.083.
[14] Machida, M.; Miyazaki, Y.; Matsunaga, Y.; Ikeue, K. Efficient Catalytic Decomposition of
Sulfuric Acid with Copper Vanadates as an Oxygen-Generating Reaction for Solar
Thermochemical Water Splitting Cycles. Chem. Commun. 2011, 47(34), 9591–9593.
DOI: 10.1039/c1cc12382c.
[15] Tagawa, H.; Endo, T. Catalytic Decomposition of Sulfuric Acid Using Metal Oxides as
the Oxygen Generating Reaction in Thermochemical Water Splitting Process. Int.
J. Hydrogen Energy. 1989, 14(1), 11–17. DOI: 10.1016/0360-3199(89)90151-1.
[16] Banerjee, A. M.; Shirole, A. R.; Pai, M. R.; Tripathi, A. K.; Bharadwaj, S. R.; Das, D.;
Sinha, P. K. Catalytic Activities of Fe 2O 3 and Chromium Doped Fe 2O 3 for Sulfuric
Acid Decomposition Reaction in an Integrated Boiler, Preheater, and Catalytic
Decomposer. Appl. Catal. B Environ. 2012, 127, 36–46. DOI: 10.1016/j.
apcatb.2012.07.030.
[17] Rashkeev, S. N.; Ginosar, D. M.; Petkovic, L. M.; Farrell, H. H. Catalytic Activity of
Supported Metal Particles for Sulfuric Acid Decomposition Reaction. Catal. Today.
2009, 139(4), 291–298. DOI: 10.1016/j.cattod.2008.03.029.
[18] Barbarossa, V.; Brutti, S.; Diamanti, M.; Sau, S.; De Maria, G. Catalytic Thermal
Decomposition of Sulphuric Acid in Sulphur-Iodine Cycle for Hydrogen
Production. Int. J. Hydrogen Energy. 2006, 31(7), 883–890. DOI: 10.1016/j.
ijhydene.2005.08.003.
[19] Banerjee, A. M.; Pai, M. R.; Tewari, R.; Raje, N.; Tripathi, A. K.; Bharadwaj, S. R.; Das, D.
A Comprehensive Study on Pt/Al2O3 Granular Catalyst Used for Sulfuric Acid
Decomposition Step in Sulfur-Iodine Thermochemical Cycle: Changes in Catalyst
Structure, Morphology and Metal-Support Interaction. Appl. Catal. B Environ. 2015,
162, 327–337. DOI: 10.1016/j.apcatb.2014.07.008.
[20] Nur, A. S. M.; Matsukawa, T.; Funada, E.; Hinokuma, S.; Machida, M. Platinum
Supported on Ta2O5 as a Stable SO3 Decomposition Catalyst for Solar
Thermochemical Water Splitting Cycles. ACS Appl. Energy Mater. 2018, 1(2),
744–750. DOI: 10.1021/acsaem.7b00195.
[21] Nagaraja, B. M.; Jung, K. D.; Ahn, B. S.; Abimanyu, H.; Yoo, K. S. Catalytic
Decomposition of So3 over Pt/BaSO4 Materials in Sulfur-Iodine Cycle for Hydrogen
Production. Ind. Eng. Chem. Res. 2009, 48(3), 1451–1457. DOI: 10.1021/ie801328w.
[22] Lee, S. Y.; Jung, H.; Kim, W. J.; Shul, Y. G.; Jung, K. D. Sulfuric Acid Decomposition on
Pt/SiC-Coated-Alumina Catalysts for SI Cycle Hydrogen Production. Int. J. Hydrogen
Energy. 2013, 38(14), 6205–6209. DOI: 10.1016/j.ijhydene.2013.01.107.
32 H. A. KHAN ET AL.

[23] Noh, S. C.; Lee, S. Y.; Shul, Y. G.; Jung, K. D. Sulfuric Acid Decomposition on the Pt/
n-SiC Catalyst for SI Cycle to Produce Hydrogen. Int. J. Hydrogen Energy. 2014, 39(9),
4181–4188. DOI: 10.1016/j.ijhydene.2014.01.021.
[24] Petkovic, L. M.; Ginosar, D. M.; Rollins, H. W.; Burch, K. C.; Pinhero, P. J.; Farrell, H. H.
Pt/TiO2 (Rutile) Catalysts for Sulfuric Acid Decomposition in Sulfur-Based
Thermochemical Water-Splitting Cycles. Appl. Catal. A Gen. 2008, 338(1–2), 27–36.
DOI: 10.1016/j.apcata.2007.12.016.
[25] Nur, A. S. M.; Matsukawa, T.; Hinokuma, S.; Machida, M. Catalytic SO 3 Decomposition
Activity and Stability of Pt Supported on Anatase TiO 2 for Solar Thermochemical
Water-Splitting Cycles. ACS Omega. 2017, 2(10), 7057–7065. DOI: 10.1021/
acsomega.7b00955.
[26] Khan, H. A.; Iqbal, M. I.; Jaleel, A.; Abbas, I.; Abbas, S. A.; Deog-Jung, K. Pt Encapsulated
Hollow Mesoporous SiO2 Sphere Catalyst for Sulfuric Acid Decomposition Reaction in
SI Cycle. Int. J. Hydrogen Energy. 2019, 44(4), 2312–2322. DOI: 10.1016/j.
ijhydene.2018.07.161.
[27] Khan, H. A.; Jung, K.-D. Preparation Scheme of Active Pt/SiC Catalyst and Its Phase
Changes during Sulfuric Acid Decomposition to Produce Hydrogen in the SI Cycle.
Catal. Letters. 2017, 147(8), 1931–1940. DOI: 10.1007/s10562-017-2105-6.
[28] Khan, H. A.; Natarajan, P.; Jung, K.-D. Stabilization of Pt at the Inner Wall of Hollow
Spherical SiO2 Generated from Pt/Hollow Spherical SiC for Sulfuric Acid
Decomposition. Appl. Catal. B Environ. 2018, 231, 151–160. DOI: 10.1016/j.
apcatb.2018.03.013.
[29] Abimanyu, H.; Jung, K. D.; Jun, K. W.; Kim, J.; Yoo, K. S. Preparation and
Characterization of Fe/Cu/Al2O3-Composite Granules for SO3 Decomposition to
Assist Hydrogen Production. Appl. Catal. A Gen. 2008, 343(1–2), 134–141. DOI:
10.1016/j.apcata.2008.03.033.
[30] Nagaraja, B. M.; Jung, K. D.; Yoo, K. S. Synthesis of Cu/Fe/Ti/Al2O3 Composite
Granules for SO3 Decomposition in SI Cycle. Catal. Letters. 2009, 128(1–2), 248–252.
DOI: 10.1007/s10562-008-9747-3.
[31] Khan, H. A.; Jaleel, A.; Natarajan, P.; Yoon, S.; Jung, K.-D. Pt Stabilization on Pt/SBA-15
through Surface Modification Using MPTMS for Sulfuric Acid Decomposition in SI
Cycle to Produce Hydrogen. Int. J. Hydrogen Energy. 2019. DOI: 10.1016/j.
ijhydene.2019.05.139.
[32] Banerjee, A. M.; Shirole, A. R.; Pai, M. R.; Tripathi, A. K.; Bharadwaj, S. R.; Das, D.;
Sinha, P. K. Catalytic Activities of Fe 2 O 3 and Chromium Doped Fe 2 O 3 for Sulfuric
Acid Decomposition Reaction in an Integrated Boiler, Preheater, and Catalytic
Decomposer. Appl. Catal. B Environ. 2012, 127, 36–46. DOI: 10.1016/j.
apcatb.2012.07.030.
[33] Nadar, A.; Banerjee, A. M.; Pai, M. R.; Meena, S. S.; Pai, R. V.; Tewari, R.; Yusuf, S. M.;
Tripathi, A. K.; Bharadwaj, S. R. Nanostructured Fe2O3 Dispersed on SiO2 as Catalyst
for High Temperature Sulfuric Acid Decomposition—Structural and Morphological
Modifications on Catalytic Use and Relevance of Fe2O3-SiO2 Interactions. Appl.
Catal. B Environ. 2017, 217, 154–168. DOI: 10.1016/j.apcatb.2017.05.045.
[34] Banerjee, A. M.; Pai, M. R.; Meena, S. S.; Tripathi, A. K.; Bharadwaj, S. R. Catalytic
Activities of Cobalt, Nickel and Copper Ferrospinels for Sulfuric Acid Decomposition:
The High Temperature Step in the Sulfur Based Thermochemical Water Splitting Cycles.
Int. J. Hydrogen Energy. 2011, 36(8), 4768–4780. DOI: 10.1016/j.ijhydene.2011.01.073.
[35] Banerjee, A. M.; Pai, M. R.; Bhattacharya, K.; Tripathi, A. K.; Kamble, V. S.;
Bharadwaj, S. R.; Kulshreshtha, S. K. Catalytic Decomposition of Sulfuric Acid on
Mixed Cr/Fe Oxide Samples and Its Application in Sulfur-Iodine Cycle for Hydrogen
CATALYSIS REVIEWS 33

Production. Int. J. Hydrogen Energy. 2008, 33(1), 319–326. DOI: 10.1016/j.


ijhydene.2007.07.017.
[36] Sano, H.;. Energy Carriers And Conversion Systems With Emphasis On Hydrogen:
Volume 2. Energy Carriers Convers. Syst. with Emphas. Hydrog. 2009, 1, 170–178.
[37] Funk, J. E.;. Thermochemical Hydrogen Production: Past and Present. Int. J. Hydrogen
Energy. 2001, 26(3), 185–190. DOI: 10.1016/S0360-3199(00)00062-8.
[38] Yalçin, S. A.;. Review of Nuclear Hydrogen Production. Int. J. Hydrogen Energy. 1989, 14
(8), 551–561. DOI: 10.1016/0360-3199(89)90113-4.
[39] Brown, L.; Besenbruch, R.; Schultz, K.; Funk, J.; Pickard, P.; Marshall, A.; Showalter, S.
High Efficiency Generation of Hydrogen Fuels Using Nuclear Power. Gen. At. U.S. Dep.
Energy. 2003 December.
[40] Brown, L.; Funk, J.; Showalter, S. Initial Screening of Thermochemical Water-Splitting
Cycles for High Efficiency Generation of Hydrogen Fuels Using Nuclear Power. GA
Proj.. 2000 April.
[41] Abanades, S.; Charvin, P.; Flamant, G.; Neveu, P. Screening of Water-Splitting
Thermochemical Cycles Potentially Attractive for Hydrogen Production by
Concentrated Solar Energy. Energy. 2006, 31(14), 2469–2486. DOI: 10.1016/j.
energy.2005.11.002.
[42] Perret, R.;. Solar Thermochemical Hydrogen Production Research (STCH)
Thermochemical Cycle Selection and Investment Priority. Sandia Rep. 2011 May,
1–117.
[43] Sakurai, M.; Bilgen, E.; Tsutsumi, A.; Yoshida, K. Adiabatic UT-3 Thermochemical
Process for Hydrogen Production. Int. J. Hydrogen Energy. 1996, 21(10), 865–870.
DOI: 10.1016/0360-3199(96)00024-9.
[44] Sakurai, M.; Bilgen, E.; Tsutsumi, A.; Solar, Y. K. UT-3 Thermochemical Cycle for
Hydrogen Production. Sol. Energy. 1996, 57(1), 51–58. DOI: 10.1016/0038-092X(96)
00034-5.
[45] Sakurai, M.; Miyake, N.; Tsutsumi, A.; Yoshida, K. Analysis of a Reaction Mechanism in
the UT-3 Thermochemical Hydrogen Production Cycle. Int. J. Hydrogen Energy. 1996,
21(10), 871–875. DOI: 10.1016/0360-3199(96)00029-8.
[46] Bilgen, C.;. The Solar Cristina Process for Hydrogen Production. 1986, 36(3), 267–280.
[47] Brown, L. C.; Besenbruch, G. E.; Schultz, K. R.; Showalter, S. K.; Marshall, A. C.;
Pickard, P. S.; Funk, J. F. High Efficiency Generation of Hydrogen Fuels Using
Thermochemical Cycles and Nuclear Power. 2002, March, 1–31.
[48] Kodama, T.; Gokon, N. Thermochemical Cycles for High-Temperature Solar
Hydrogen Production. Chem. Rev. 2007, 107(10), 4048–4077. DOI: 10.1021/
cr050188a.
[49] Ozturk, I. T.; Hammache, A.; Bilgen, E. An Improved Process for H2SO4 Decomposition
Step of the Sulfur-Iodine Cycle. Energy Convers. Manag. 1995, 36(1), 11–21. DOI:
10.1016/0196-8904(94)00036-Y.
[50] Jorgenson, D. W.;. Productivity and Postwar U.S. Economic Growth. J. Econ. Perspect..
1988, 2(4), 23–41. DOI: 10.1257/jep.2.4.23.
[51] Oyediran, O. B.; Centre, N. M.; Equations, O. D. Energy the Backbone of Economic and
Technological Development A Support for President Alhaji Umaru Y ’ Adua, GCFR
Human and Financial Development Programme. 2016, April. 10.13140/
RG.2.1.1946.4085.
[52] Russell, J.-L. J.; McCorkle, K.-H.; Norman, J.-H.; Porter, J.-T. I. I.; Roemer, T.-S.;
Schuster, J.-R.; Sharp, R.-S. Water Splitting - A Progress Report. In 1st World
Hydrogen Energy Conference, Volume 1, Vol. 1, 1976, 1A_105-1A_124.
34 H. A. KHAN ET AL.

[53] Brown, L. C.; Besenbruch, G. E.; Lentsch, R. D.; Schultz, K. R.; Funk, J. F.; Pickard, P. S.;
Marshall, A. C.; Showalter, S. K. High Efficiency Generation of Hydrogen Fuels Using
Nuclear Power Final Technical Report for the Period. 1999, June.
[54] Vitart, X.; Le Duigou, A.; Carles, P. Hydrogen Production Using the Sulfur-Iodine Cycle
Coupled to a VHTR: An Overview. Energy Convers. Manag. 2006, 47(17), 2740–2747.
DOI: 10.1016/j.enconman.2006.02.010.
[55] Russ, B.; Sulfur Iodine Process Summary for the Hydrogen Technology Down-Selection:
Process Performance Package. 2009, June, INL/EXT-12-25773.
[56] El-Emam, R. S.; Khamis, I. International Collaboration in the IAEA Nuclear Hydrogen
Production Program for Benchmarking of HEEP. Int. J. Hydrogen Energy. 2017, 42(6),
3566–3571. DOI: 10.1016/j.ijhydene.2016.07.256.
[57] Cerri, G.; Salvini, C.; Corgnale, C.; Giovannelli, A.; De Lorenzo Manzano, D.;
Martinez, A. O.; Le Duigou, A.; Borgard, J. M.; Mansilla, C. Sulfur-Iodine Plant for
Large Scale Hydrogen Production by Nuclear Power. Int. J. Hydrogen Energy. 2010, 35
(9), 4002–4014. DOI: 10.1016/j.ijhydene.2010.01.066.
[58] Kondamudi, K.; Upadhyayula, S. Kinetic Studies of Sulfuric Acid Decomposition over
Al-Fe 2O 3 Catalyst in the Sulfur-Iodine Cycle for Hydrogen Production. Int.
J. Hydrogen Energy. 2012, 37(4), 3586–3594. DOI: 10.1016/j.ijhydene.2011.05.026.
[59] Orozco-mena, R. E.; Rom, M.; Soto-dı, O.; Romero-paredes, H.; Camacho-d, A. A.
ScienceDirect Metal Sulfate Decomposition Using Green Pd-Based Catalysts
Supported on G Al 2 O 3 and SiC : A Common Step in Sulfur-Family
Thermochemical Cycles. 2018, 4, 2–7. doi:10.1016/j.ijhydene.2018.08.045
[60] Soto-Díaz, O.; Orozco-Mena, R. E.; Román-Aguirre, M.; Romero-Paredes, H.;
Camacho-Dávila, A. A.; Ramos-Sánchez, V. H. Metal Sulfate Decomposition Using
Green Pd-Based Catalysts Supported on Γal2o3 and SiC: A Common Step in Sulfur-
Family Thermochemical Cycles. Int. J. Hydrogen Energy. 2019, 4, 12309–12314. DOI:
10.1016/j.ijhydene.2018.08.045.
[61] Machida, M.; Kawada, T.; Hebishima, S.; Hinokuma, S. Macroporous Supported
Cu-V Oxide as A Promising Substitute of Pt Catalyst for Sulfuric Acid
Decomposition in Solar Thermochemical Hydrogen Production. 2–7. doi:10.1186/
1756-6606-4-2.
[62] Machida, M.; Kawada, T.; Yamashita, H.; Tajiri, T. Role of Oxygen Vacancies in Catalytic
SO3 Decomposition over Cu2V2O7 in Solar Thermochemical Water Splitting Cycles.
J. Phys. Chem. C. 2013, 117(50), 26710–26715. DOI: 10.1021/jp410431a.
[63] Kawada, T.; Tajiri, T.; Yamashita, H.; Machida, M. Molten Copper Hexaoxodivanadate:
An Efficient Catalyst for SO3 Decomposition in Solar Thermochemical Water Splitting
Cycles. Catal. Sci. Technol. 2014, 4(3), 780. DOI: 10.1039/c3cy00880k.
[64] MacHida, M.; Miyazaki, Y.; Matsunaga, Y.; Ikeue, K. Efficient Catalytic Decomposition
of Sulfuric Acid with Copper Vanadates as an Oxygen-Generating Reaction for Solar
Thermochemical Water Splitting Cycles. Chem. Commun. 2011, 47(34), 9591–9593.
DOI: 10.1039/c1cc12382c.
[65] Kawada, T.; Ikematsu, A.; Tajiri, T.; Takeshima, S.; Machida, M. Structure and SO3
Decomposition Activity of CeVO4/SiO2 Catalysts for Solar Thermochemical Water
Splitting Cycles. Int. J. Hydrogen Energy. 2015, 40(34), 10726–10733. DOI: 10.1016/j.
ijhydene.2015.07.007.
[66] Kawada, T.; Yamashita, H.; Zheng, Q.; Machida, M. Hydrothermal Synthesis of CuV2O6
Supported on Mesoporous SiO2 as SO3 Decomposition Catalysts for Solar
Thermochemical Hydrogen Production. Int. J. Hydrogen Energy. 2014, 39(35),
20646–20651. DOI: 10.1016/j.ijhydene.2014.06.162.
CATALYSIS REVIEWS 35

[67] Kawada, T.; Hinokuma, S.; Machida, M. Structure and SO3 Decomposition Activity of
NCuO-V2O5/SiO2 (N = 0, 1, 2, 3 and 5) Catalysts for Solar Thermochemical Water
Splitting Cycles. Catal. Today. 2015, 242(PB), 268–273. DOI: 10.1016/j.
cattod.2014.05.023.
[68] Wang, L.; Zhu, Y.; Yang, H.; He, Y.; Xia, J.; Zhang, Y.; Wang, Z. SO3 Decomposition
over CuO–CeO2 Based Catalysts in the Sulfur–Iodine Cycle for Hydrogen Production.
Int. J. Hydrogen Energy. 2018, 43(32), 14876–14884. DOI: 10.1016/j.
ijhydene.2018.06.056.
[69] Ginosar, D. M.; Petkovic, L. M.; Glenn, A. W.; Burch, K. C. Stability of Supported
Platinum Sulfuric Acid Decomposition Catalysts for Use in Thermochemical Water
Splitting Cycles. Int. J. Hydrogen Energy. 2007, 32(4), 482–488. DOI: 10.1016/j.
ijhydene.2006.06.053.
[70] Zhang, Y.; Yang, H.; Zhou, J.; Wang, Z.; Liu, J.; Cen, K. Catalytic Decomposition of
Sulfuric Acid over CuO/CeO2in the Sulfur-Iodine Cycle for Hydrogen Production. Int.
J. Hydrogen Energy. 2015, 40(5), 2099–2106. DOI: 10.1016/j.ijhydene.2014.12.048.
[71] Su, T.; Zhang, P.; Wang, L.; Chen, S. Preparation of Copper Chromite by Vacuum
Freezing Drying Method and Its Catalytic Activity for Sulfuric Acid Decomposition.
Chinese J. Chem. 2010, 28(12), 2339–2344. DOI: 10.1002/cjoc.201190001.
[72] Yannopoulos, L. N.; Pierre, J. F. Hydrogen Production Process: High
Temperature-Stable Catalysts for the Conversion of SO3 to SO2. Int. J. Hydrogen
Energy. 1984, 9(5), 383–390. DOI: 10.1016/0360-3199(84)90058-2.
[73] Khan, H. A.; Kim, S.; Jung, K.-D. Origin of High Stability of Pt/Anatase-TiO2 Catalyst in
Sulfuric Acid Decomposition for SI Cycle to Produce Hydrogen. Catal. Today. 2019.
DOI: 10.1016/j.cattod.2019.10.037.
[74] Khan, H. A.; Natarajan, P.; Jung, K.-D. Synthesis of Pt/Mesoporous SiC-15 and Its
Catalytic Performance for Sulfuric Acid Decomposition. Catal. Today. 2018, 303,
25–32. DOI: 10.1016/j.cattod.2017.09.026.
[75] Zhang, P.; Su, T.; Chen, Q. H.; Wang, L. J.; Chen, S. Z.; Xu, J. M. Catalytic
Decomposition of Sulfuric Acid on Composite Oxides and Pt/SiC. Int. J. Hydrogen
Energy. 2012, 37(1), 760–764. DOI: 10.1016/j.ijhydene.2011.04.064.
[76] Ginosar, D. M.; Rollins, H. W.; Petkovic, L. M.; Burch, K. C.; Rush, M. J. High-
Temperature Sulfuric Acid Decomposition over Complex Metal Oxide Catalysts. Int.
J. Hydrogen Energy. 2009, 34(9), 4065–4073. DOI: 10.1016/j.ijhydene.2008.09.064.
[77] Ishikawa, H.; Ishii, E.; Uehara, I.; Nakane, M. Catalyzed Thermal Decomposition of
H2SO4 and Production of HBr by the Reaction of SO2 with Br2 and H2O. Int.
J. Hydrogen Energy. 1982, 7(3), 237–246. DOI: 10.1016/0360-3199(82)90087-8.
[78] Karagiannakis, G.; Agrafiotis, C. C.; Pagkoura, C.; Konstandopoulos, A. G.; Thomey, D.;
De Oliveira, L.; Roeb, M.; Sattler, C. Hydrogen Production via Sulfur-Based
Thermochemical Cycles: Part 3: Durability and Post-Characterization of Silicon
Carbide Honeycomb Substrates Coated with Metal Oxide-Based Candidate Catalysts
for the Sulfuric Acid Decomposition Step. Int. J. Hydrogen Energy. 2012, 37(10),
8190–8203. DOI: 10.1016/j.ijhydene.2012.02.090.
[79] Brittain, R. D.; Hiidenbrand, D. L. Catalytic Decomposition of Gaseous SO3. J. Phys.
Chem. 1983, 87, 3713–3717. DOI: 10.1021/j100242a027.
[80] Brutti, S.; De Maria, G.; Cerri, G.; Giovannelli, A.; Brunetti, B.; Cafarelli, P.; Semprin, E.;
Barbarossa, V.; Ceroli, A. Decomposition of H 2so 4 by Direct Solar Radiation. Ind. Eng.
Chem. Res. 2007, 46(20), 6393–6400. DOI: 10.1021/ie070245l.
[81] Giaconia, A.; Sau, S.; Felici, C.; Tarquini, P.; Karagiannakis, G.; Pagkoura, C.;
Agrafiotis, C.; Konstandopoulos, A. G.; Thomey, D.; De Oliveira, L.;, et al. Hydrogen
Production via Sulfur-Based Thermochemical Cycles: Part 2: Performance Evaluation of
36 H. A. KHAN ET AL.

Fe2O3-Based Catalysts for the Sulfuric Acid Decomposition Step. Int. J. Hydrogen
Energy.2011, 36(11), 6496–6509. DOI: 10.1016/j.ijhydene.2011.02.137.
[82] O’keefe, D. R.; Norman, J. H.; Williamson, D. G. Catalysis Research in Thermochemical
Water-Splitting Processes. Catal. Rev. 1980, 22(3), 325–369. DOI: 10.1080/
03602458008067537.
[83] Spewock, S.; Brecher, L. E.; Talko, F. Thermal Catalytic Decomposition of Sulfur
Trioxide to Sulfur Dioxide and Oxygen. In Miami Univ. First World Hydrogen Energy
Conf. Proc; 1976; Vol. 1 Sess 9A, pp 53–68.
[84] Noglik, A.; Roeb, M.; Sattler, C.; Pitz-Paal, R. Experimental Study on Sulfur Trioxide
Decomposition in a Volumetric Solar Receiver-Reactor. In 2008 Proceedings of the 2nd
International Conference on Energy Sustainability, ES 2008; 2009. 10.1115/es2008-54171.
[85] Myers, B. F.; Francisco, S.; Office, O. General Atomic Company. 1979, September.
[86] Thomey, D.; De Oliveira, L.; Säck, J. P.; Roeb, M.; Sattler, C. Development and Test of
a Solar Reactor for Decomposition of Sulphuric Acid in Thermochemical Hydrogen
Production. Int. J. Hydrogen Energy. 2012, 37(21), 16615–16622. DOI: 10.1016/j.
ijhydene.2012.02.136.
[87] Astholz, D. C.; Glänzer, K.; Troe, J. The Spin-forbidden Dissociation–Recombination
Reaction SO 3 ?SO 2 +O. J. Chem. Phys. 1979, 70(5), 2409–2413. DOI: 10.1063/1.437751.
[88] Zhang, Y.; Yang, H.; Zhou, J.; Wang, Z.; Liu, J.; Cen, K. Detailed Kinetic Modeling of
Homogeneous H2SO4 Decomposition in the Sulfur-Iodine Cycle for Hydrogen
Production. Appl. Energy. 2014, 130, 396–402. DOI: 10.1016/j.apenergy.2014.05.017.
[89] Pelovski, Y.; Petkova, V. Mechanism and Kinetics of Inorganic Sulphates
Decomposition. J. Therm. Anal. 1997, 49(3), 1227–1241. DOI: 10.1007/bf01983679.
[90] Khan, H. A.; Jung, K.; Ahamad, T.; Ubaidullah, M.; Imran, M.; Alshehri, S. M. Pt-Core
Silica Shell Nanostructure: A Robust Catalyst for the Highly Corrosive Sulfuric Acid
Decomposition Reaction in Sulfur Iodine Cycle to Produce Hydrogen †. 2020. DOI:
10.1039/d0nj04830e.
[91] Banerjee, A. M.; Pai, M. R.; Tripathi, A. K.; Bharadwaj, S. R.; Das, D. Development of
Catalyst for Decomposition of Sulfuric Acid: The Energy Intensive Step in Sulfur-Iodine
Thermochemical Cycle for Hydrogen Generation Using Nuclear Heat. BARC Newsl.
2013, 332, 7–12.

You might also like