Sie sind auf Seite 1von 17

5291 r2009 American Chemical Society pubs.acs.

org/EF
Energy Fuels 2009, 23, 52915307
:
DOI:10.1021/ef900714q
Published on Web 10/16/2009
Review of Mid- to High-Temperature Sulfur Sorbents for Desulfurization of
Biomass- and Coal-derived Syngas
Singfoong Cheah,* Daniel L. Carpenter, and Kimberly A. Magrini-Bair
National Bioenergy Center, National Renewable Energy Laboratory, 1617 Cole Blvd., MS 3322 Golden, Colorado 80401
Received July 11, 2009. Revised Manuscript Received September 2, 2009
This review examines state-of-the-art mid- and high-temperature sulfur sorbents that remove hydrogen
sulfide (H
2
S) from syngas generated from coal gasification and may be applicable for use with biomass-
derived syngas. Biomass feedstocks contain low percentages of protein-derived sulfur that is converted
primarily to H
2
S, as well as small amounts of carbonyl sulfide (COS) and organosulfur compounds during
pyrolysis and gasification. These sulfur species must be removed from the raw syngas before it is used for
downstream fuel synthesis or power generation. Several types of sorbents based on zinc, copper, iron,
calcium, manganese, and ceria have been developed over the last two decades that are capable of removing
H
2
Sfromdry coal-derivedsyngas at mid- tohigh-temperature ranges. Further improvement is necessary to
develop materials more suitable for desulfurization of biomass-derived syngas because of its hydrocarbon,
tar, and potentially high steamcontent, which presents different challenges as compared to desulfurization
of coal-derived syngas.
1. Introduction
Rising world demands for oil and a finite petroleumreserve
have increased interest in alternate liquid fuel sources that are
diverse, secure, and affordable. Biomass is a renewable source
of carbon that is abundant in many regions of the world. In
the United States it has been estimated that domestic biomass
resources could be used to sustainably replace more than a
third of the U.S. demand for transportation fuel (2005 basis).
1
Worldwide, the International Energy Agency estimated that
there is sufficient nonfood biomass available to support a
substantial increase in bioenergy use.
2
With reduction of green house gas emissions to consider,
the focus of next-generationbiofuels is toproduce fuel derived
from cellulosic biomass, especially biofuels made with waste
biomass or feedstock grown on marginal lands that generate
little carbon today.
3-5
Thermochemical conversion of bio-
mass to fuels via pyrolysis or gasification is a viable route that
has the potential to be cost competitive with gasoline, as well
as biochemical biomass conversion processes.
6
During gasifi-
cationof biomass feedstocks, agas mixture comprising mainly
H
2
, CO, CO
2
, H
2
O, and CH
4
is produced along with several
unwanted byproducts, with their concentrations depending
on feedstock, gasifier design, and process conditions. These
unwanted byproducts include organic tars, sulfur and nitro-
gen heteroatom species (thiophene, pyridine), and inorganic
constituents containing sulfur (H
2
S, COS), chlorine (HCl),
nitrogen (NH
3
, HCN), and alkali metals.
6-11
Althoughthere are quite a number of studies that measured
the amount of sulfur in dry biomass feedstocks, data on the
levels and speciation of sulfur in the actual syngas that is
derived from biomass gasification are scarce. A summary of
the available data on sulfur concentrations in biomass syngas
is presented in Table 1. Even though this review focuses on
materials that remove sulfur, it also includes discussion of the
influence of other gas impurities on sorbent performance and
of the capability of the sorbents to remove multiple contami-
nants. Therefore the concentrations of other impurities such
as HCl, nitrogen compounds, and tar are also included in
Table 1.
Van der Drift et al. tested 10 different biomass feedstocks in
an atmospheric air-blown gasifier and determined H
2
S in the
biomass-derivedsyngas tobe inthe range of 50-230 ppmv for
several different types of wood(Table 1).
7
Satoet al. measured
H
2
S at the outlet of a bench-scale heated atmospheric flui-
dized bed gasifier and found less than 50 ppmv H
2
S and
*To whom correspondence should be addressed. Telephone: (303)
384-7707. Fax: (303) 384-6363. E-mail: Singfoong.Cheah@nrel.gov.
(1) Perlack, R. D.; Wright, L. L.; Turhollow, A. F.; Graham, R. L.;
Stokes, B. J.; Erbach, D. C. Biomass Feedstock for a Bioenergy and
Bioproducts Industry: The Technical Feasibility of a Billion-Ton Annual
Supply; DOE/GO-102005-2135; Oak Ridge National Laboratory: Oak
Ridge, TN, 2005.
(2) International Energy Agency. The availability of biomass resources
for energy;summary and conclusions from the IEA Bioenergy ExCo58
Workshop. http://www.ieabioenergy.com/MediaItem.aspx?id=5794 (accessed
August 27, 2009).
(3) Farrell, A. E.; Plevin, R. J.; Turner, B. T.; Jones, A. D.; O
0
Hare,
M.; Kammen, D. M. Science 2006, 311 (5760), 506508.
(4) Searchinger, T.; Heimlich, R.; Houghton, R. A.; Dong, F.;
Elobeid, A.; Fabiosa, J.; Tokgoz, S.; Hayes, D.; Yu, T.-H. Science
2008, 319 (5867), 12381240.
(5) Tilman, D.; Hill, J.; Lehman, C. Science 2006, 314 (5805), 1598
1600.
(6) Phillips, S.; Aden, A.; Jechura, J.; Dayton, D.; Eggeman, T.
Thermochemical Ethanol via Indirect Gasification and Mixed Alcohol
Synthesis of Lignocellulosic Biomass; NREL/TP-510-41168; National Re-
newable Energy Laboratory: Golden, CO., 2007.
(7) van der Drift, A.; van Doorn, J.; Vermeulen, J. W. Biomass
Bioenergy 2001, 20 (1), 4556.
(8) Sato, K.; Shinoda, T.; Fujimoto, K. J. Chem. Eng. Jpn. 2007, 40
(10), 860868.
(9) Milne, T. A.; Evans, R. J.; Abatzoglou, N. Biomass Gasifier Tars:
Their Nature, Formation, and Conversion; NREL/TP-570-25357;
National Renewable Energy Laboratory: Golden, CO, November 1998.
(10) Leppalahti, J.; Koljonen, T. Fuel Process. Technol. 1995, 43 (1),
145.
(11) Torres, W.; Pansare, S. S.; Goodwin, J. G., Jr. Cat. Rev. - Sci.
Eng. 2007, 49, 407456.
5292
Energy Fuels 2009, 23, 52915307
:
DOI:10.1021/ef900714q
30 ppmv SO
x
when wood chips were used as feedstock, and
more than 300 ppmv total sulfur when dried sewage sludge
was used as a feedstock (Table 1).
8
Kuramochi et al. modeled
H
2
S and HCl speciation during gasification of a number of
biomass feedstocks, including several that were studied by van
der Drift et al., using thermodynamic modeling.
12
There are,
however, disparities between experimental and model results,
indicating that further research is necessary. Research con-
ducted at the National Renewable Energy Laboratory
(NREL) has shown that at 850 C, the syngas H
2
S content
ranges from about 20-50 ppmv for Vermont wood to
300-600 ppmv for herbaceous feedstocks such as switchgrass
and wheat straw.
13,14
Boerrigter summarized data of typical
gas compositions obtained with three different types of gasi-
fiers and found an H
2
S content of 40-120 ppmv when wood
was used as a feedstock.
15
There are a number of reasons that sulfur needs to be
removed frombiomass derived product gas. Hydrogen sulfide
and other sulfur compounds can cause pipeline corrosion and
thus limit plant lifetime.
16
In addition, there are regulations to
limit the sulfur content in fuel such as diesel to 10-15 ppm in
United States,
17
European Union, Canada, Australia, and
New Zealand.
18
It is well-known that sulfur in fuel is a
precursor to SO
2
that will be released to the atmosphere,
resulting in downstream environmental problems. It can be
expected that fuel derived from cellulosic biomass feedstock
will have similar requirements in terms of sulfur content;
consequently it is necessary toremove sulfur compounds from
biomass-derived syngas/fuel/products at some point in the
processing.
Hydrogen sulfide is also a well-known catalyst poison.
Information on the effect of sulfur on tar reforming or steam
reforming catalysts in biomass product gas environments is
widely available and generally shows significant negative
impacts of sulfur on catalyst performance.
8,19-23
In addition
to its effect on tar reforming catalysts, sulfur also affects fuel
synthesis catalysts such as those used in mixed alcohol and in
Fischer-Tropsch hydrocarbon synthesis. For example, trace
amount of H
2
S in cedar wood derived syngas poisons ruthe-
nium catalysts used for Fischer-Tropsch synthesis.
24
The
specific activities of copper catalysts used in methanol synth-
esis
25
and cobalt catalysts used in Fischer-Tropsch fuel
synthesis decrease significantly with H
2
S/H
2
ratios as low as
tens to hundreds of ppb (Figure 5.6 and Chapter 6 of ref 26).
In contrast, a sulfided molybdenum alcohol synthesis catalyst
requires up to 100 ppm H
2
S to maintain sufficient activity.
6
Consequently, some level of sulfur removal will be necessary
to achieve biomass based catalytic fuel production.
6
A sum-
mary of syngas purity requirements for downstream catalysts
is presented in Table 2.
In a technoeconomic analysis of thermochemical ethanol
production conducted at NREL, ethanol production, tar
reforming, acid gas, and sulfur removal processes together
comprise 31%of the total fuel production cost (second largest
cost component after feedstock).
6
Thus, a more efficient
method of H
2
S removal will significantly reduce the overall
biofuel production cost.
In conventional treatment, H
2
S or other sulfur compounds
are removed via low temperature amine scrubbers. Removing
H
2
S from biomass-derived syngas using scrubbers thus re-
quires lowering the temperature of the syngas from 850 C
(temperature of gasification) to 40-50 C for cleanup
27
with
concurrent tar condensation. However, tar condensation can
cause plugging and fouling of the condenser and downstream
process piping, presenting a significant operating challenge. In
addition, the scrubbed gas must then be reheated for down-
stream fuel synthesis, which occurs from 250-400 C
(Figure 1). This cooling and reheating of the syngas is both
thermally inefficient and expensive. Theoretically, the heat
energy can be partially recovered using heat exchange equip-
ment, but tar build up on reactor walls during condensation
could render the heat recovery challenging. In addition to the
tar condenser clogging problem, tar condensation also reduces
biomass carbon utilization and produces a large waste stream
that must be treated. New mid- (400-600 C) to high-tem-
perature (600-850 C) syngas cleaning processes being devel-
oped by the coal industry, for fuel cell applications, and for
biomass syngas applications offer potential economic advan-
tages by allowing gasification, gas clean up, and downstream
processes to be operated at similar process temperatures.
Table 1. Experimentally Determined Concentrations of Impurities in Biomass-Derived Syngas
H
2
S (ppmv, dry basis) HCl (ppmv, dry basis) NH
3
(ppmv) tar (g/Nm
3
)
50-230 (wood)
a
1-200 (wood, verge grass, etc.)
a
1,000-13,000 (wood, verge grass, etc.)
a
1-150
b
<50 (wood chips)
c
<10 (wood chips)
c
<1,000 (wood chips)
c
10 (fluidized bed average)
d
20-50 (wood)
e
1,000-14,000
b
50 (updraft gasifier average)
d
300-600 (herbaceous feedstock)
e
500-1,000 (wood)
f
1 (downdraft gasifier average)
d
40-120 (wood)
f
a
Reference 7.
b
Reference 11.
c
Reference 8.
d
Reference 9.
e
References 13 and 14.
f
Reference 15.
(12) Kuramochi, H.; Wu, W.; Kawamoto, K. Fuel 2005, 84 (4), 377
387.
(13) Magrini-Bair, K. A.; Czernik, S. R.; French, R. J. Catalyst
Fundamentals: Addition of Promoters/Additives to Improve Reforming
Catalyst Sulfur Tolerance As Measured by 2x Improvement in Activity in
the Presence of 50-500 ppm H
2
S; National Renewable Energy Laboratory:
Golden, CO, March 2007.
(14) Carpenter, D. L.; Bain, R. L.; Davis, R. E.; Abhijit, D.; Feik, C.
J.; Gaston, K. R.; Jablonski, W.; Phillips, S. D.; Nimlos, M. R. Ind. Eng.
Chem. Res. 2009, Submitted.
(15) Boerrigter, H.; Rauch, R. Review of Applications of Gases from
Biomass Gasification; Vienna Institute of Technology: The Netherlands,
June 2006.
(16) Basu, P. Combustion and Gasification in Fluidized Beds; CRC
Press: Boca Raton, FL, 2006.
(17) Environmental Protection Agency. Fuels and Fuel Additives.
http://www.epa.gov/OMSWWW/regs/fuels/diesel/diesel.htm (accessed
August 27, 2009).
(18) Wikipedia. Ultra-low Sulfur Diesel. http://en.wikipedia.org/wiki/
Ultra-low_sulfur_diesel (accessed August 27, 2009).
(19) Bain, R. L.; Dayton, D. C.; Carpenter, D. L.; Czernik, S. R.;
Feik, C. J.; French, R. J.; Magrini-Bair, K. A.; Phillips, S. D. Ind. Eng.
Chem. Res. 2005, 44 (21), 79457956.
(20) Koningen, J.; Sjostrom, K. Ind. Eng. Chem. Res. 1998, 37 (2),
341346.
(21) Hepola, J.; McCarty, J.; Krishnan, G.; Wong, V. Appl. Catal., B
1999, 20 (3), 191203.
(22) Hepola, J.; Simell, P. Appl. Catal., B 1997, 14 (3-4), 287303.
(23) Hepola, J.; Simell, P. Appl. Catal., B 1997, 14 (3-4), 305321.
(24) Okabe, K.; Murata, K.; Nakanishi, M.; Ogi, T.; Nurunnabi, M.;
Liu, Y. Y. Catal. Lett. 2009, 128, 171176.
(25) Kung, H. H. Catal. Today 1992, 11 (4), 443453.
(26) Bartholomew, C. H.; Farrauto, R. J. Fundamentals of Industrial
Catalytic Processes, 2nd ed.; Wiley-Interscience: Hoboken, NJ, 2006.
(27) Vamvuka, D.; Arvanitidis, C.; Zachariadis, D. Environ. Eng. Sci.
2004, 21 (4), 525547.
5293
Energy Fuels 2009, 23, 52915307
:
DOI:10.1021/ef900714q
Even though significant progress has been achieved in the
development of midtemperature desulfurization of coal de-
rived syngas, desulfurization of biomass-derived syngas pre-
sents both unique challenges and advantages due to its
composition. Coal-derived syngas typically has 0.1-1.5%
sulfur, which is higher than that in biomass-derived syngas.
28
Coal gasification is usually a direct oxygen- or air-blown
process conducted at higher temperature than biomass gasi-
fication. Consequently, the coal syngas that is produced
usually has low water and hydrocarbon content.
Biomass gasification is accomplished using either a direct
oxygen-blown process or an indirect steam-driven process.
Both types of gasifiers are currently under development and
bothare able toproduce a syngas suitable for downstreamfuel
synthesis. For direct gasification, air is not usually considered
a gasifying agent in the context of fuel production because the
high nitrogen levels in the resulting syngas would be econom-
ically unacceptable. Therefore, oxygen is often used as a
gasifying agent in direct gasification, a major disadvantage
being the need for concurrent oxygen production. This step is
expensive andrequires large plant sizes toimprove economics.
For methanol production, a comparison of these two gasifica-
tion processes indicates that the capital intensity, that is the
capital cost per unit product, is comparable or less expensive
for indirect gasification (Figure 1 in ref 6).
Steam gasification brings its own set of challenges, such as
erosion problems inside reactor vessels due to the need to
circulate hot solids for process heat. Also, since indirect
gasifiers are usually operated at low pressure, the synthesis
gas requires downstreamcompressionfor fuel production. On
the other hand, there are several advantages to using steam.
Steam helps to produce a syngas with a H
2
/CO ratio that is
higher than one,
29
which is more suitable for mixed alcohol or
hydrocarbon fuel (e.g., via the Fischer-Tropsch route) synth-
eses. Moreover, it has been reported that the tar content of the
syngas decreases as the steam/biomass ratio increases,
30
and
more recent research suggests that less tar is produced with
steam/oxygen mixtures as a gasifying agent.
29,31
Steam in the
process gas also increases tar reforming and reduces coke
formation.
29
Using steam as a gasifying agent results in a raw
syngas that has much higher steam content (30-65%) than
that in syngas from coal conversion, with the low and high
ends from gasifiers using oxygen/steam and steam as media,
respectively.
6,9,11,15
Although biomass-derived syngas con-
tains muchless sulfur thancoal-derivedsyngas, this additional
steam content presents process challenges unique to biomass
gasification.
The other process challenge unique to biomass derived
syngas is the higher hydrocarbon and tar content compared
to coal syngas. Methane concentrations can range 10-15%
by volume, and other light hydrocarbons, such as acetylene,
ethylene, and ethane, can total a few percent.
19
Tar may be
present in concentrations of up to 5%.
9
It is well-known that
sulfur has a detrimental effect on tar reforming catalysts;
however, the potential effect of tar on high temperature
sulfur sorbents is still not clear. In addition, hydrocarbons
such as ethene have been known to form coke on steam
reforming catalysts,
32,33
and could also potentially form coke
on sulfur sorbents. Literature data on the impacts of some
these impurities on sorbent performance are summarized in
Section 2.8.
1.1. Economics of Mid- to High-Temperature Desulfuriza-
tion. Because of the lack of data about sorbent performance
in biomass-derived syngas, no economic analyses exist re-
garding the thermal efficiency and process economics of
using high temperature desulfurization versus conventional
sorbent processes. Several reports were published describing
hot gas cleanup systems applied to coal gasification. For
example, RTI International estimated that desulfurization at
370-480 C in a 600 MW integrated gasification combined
cycle (IGCC) plant could increase the overall process ther-
mal efficiency by 3.6 efficiency points over conventional
Selexol sulfur removal technology,
34
where thermal effi-
ciency is defined as the ratio of net work done to the heat
content of the fuel that is consumed. This gain would reduce
the plant cost by $269 per kilowatt, resulting in a associated
9.6% reduction of electricity cost.
34
Considering that IGCC
normally has a thermal efficiency of 40-50 points,
35,36
an
increase of 3.6 thermal efficiency points is significant.
The Netherlands Agency for Energy and the Environment
(NOVEM)
37
modeled hot gas cleanup with a 600 MW
system operated at 250, 350, and 600 C. At each tempera-
ture a different set of processes for cleaning up NH
3
, halo-
gens, sulfur, and dust was used. One exception was that the
same dehalogenation and denitrification processes were used
Table 2. Acceptable Levels of Impurities in Biomass-Derived Syngas for Downstream Fuel Synthesis Catalytic Processes
process sulfur (ppmv) halogen (ppbv) nitrogen (ppmv) tar
Fischer-Tropsch sum of sulfur compounds <1
a
HCl HBr HF < 10
a
sum of nitrogen compounds <1
a
below dew point
a
methanol synthesis <0.5, preferably <0.1
b
<1
c
a
Reference 15.
b
Reference 25.
c
Reference 179.
Figure 1. Schematic diagram of a biofuel synthesis pathway that uses
conventional amine scrubbing process to remove acid gases. Tempera-
tures for gasificationandalcohol/fuel synthesis are approximate values.
(28) Wakker, J. P.; Gerritsen, A. W.; Moulijn, J. A. Ind. Eng. Chem.
Res. 1993, 32 (1), 139149.
(29) Gil, J.; Corella, J.; Aznar, M. P.; Caballero, M. A. Biomass
Bioenergy 1999, 17 (5), 389403.
(30) Herguido, J.; Corella, J.; Gonzalez-Saiz Ind. Eng. Chem. Res.
1992, 31, 12741282.
(31) Devi, L.; Ptasinski, K. J.; Janssen, F. J. J. G. Biomass Bioenergy
2003, 24 (2), 125140.
(32) Rostrup-Nielsen, J. R.; Sehested, J.; Norskov, J. K. Hydrogen
and synthesis gas by steam- and CO
2
reforming. In Adv. Catal.,
Academic Press Inc: San Diego, 2002; Vol. 47, pp 65-139.
(33) Trimm, D. L. Catal. Today 1997, 37 (3), 233238.
(34) NETL. 2007 NETL Accomplishments; National Energy Technol-
ogy Laboratory, U.S. Department of Energy: 2007.
(35) Jiang, L.; Lin, R.; Jin, H.; Cai, R.; Liu, Z. Energy Convers.
Manage. 2002, 43 (9-12), 13391348.
(36) Ordorica-Garcia, G.; Douglas, P.; Croiset, E.; Zheng, L. Energy
Convers. Manage. 2006, 47 (15-16), 22502259.
(37) NOVEM. System Study High Temperature Gas Cleaning at
IGCCSystems; Netherlands Agency for Energy and the Environment: 1991.
5294
Energy Fuels 2009, 23, 52915307
:
DOI:10.1021/ef900714q
at 600 and 350 C, which might have contributed to the
marginal improvement in thermal efficiency at 600 C over
that at 350 C. One of the key conclusions of the report
was that the gain in thermal efficiency from using mid- to
high-temperature desulfurization is intimately tied to the
availability and choice of denitrification technology; conse-
quently, research and development of denitrification pro-
cesses is important.
Another factor to be taken into consideration is alkali
removal. Because of process equipment limitations and the
need to operate below alkali metal condensation tempera-
tures (so that the condensed metals can be collected and
filtered),
38,39
the optimum hot gas cleanup temperature
range designed for use in the coal industry is approximately
300-600 C.
38-40
Information on alkali content of biomass-
derived syngas is scarce, and its effect on the ultimate
optimum biomass syngas clean up temperature and condi-
tions will still need to be determined.
The processes downstream of syngas cleanup for IGCC
are power generation using gas and steam turbines, which
differ significantly from the downstream fuel synthesis pro-
cesses that are used for biomass-derived syngas. Therefore, it
cannot be assumed that thermal efficiency improvements in
the biomass-based process would be similar. In addition,
because of the inefficiency and cost of transporting biomass
over long distances, it is likely that most biomass gasification
plants will be significantly smaller than coal gasification
plants. This will likely entail different process economics
for heat recovery and gas cleaning in the smaller biomass
gasification plants. As a result, it is critical that thermal
efficiency and economics for biomass syngas cleanup be
analyzed as more data on sorbent performance in biomass-
derived syngas becomes available.
In this literature review, sources related to H
2
S removal in
the context of IGCC, biomass gasification, and fuel cell
applications are reviewed. The criteria for choosing an
appropriate sorbent material for use with biomass-derived
syngas are then discussed. In later sections, the basic sulfur
sorption steps and research results obtained with several
sorbent materials are discussed. The review then focuses on
sorbent regeneration pathways, which are dependent both
on the inherent thermodynamics of the material and also on
process conditions.
1.2. Criteria for Selecting Sulfur Sorption Materials. Cri-
teria that the sorbent material must satisfy, for syngas
desulfurization to be both economical and operational,
are: (1) High adsorption capacity for H
2
S. This reduces both
sorbent quantity and process equipment size. (2) Fast ad-
sorption kinetics. (3) Mechanical properties: low attrition
rate and able to tolerate high temperature. (4) Chemical
properties: stable in reducing environment containing steam
and hydrocarbon. (5) Regenerable through a suitable path-
way while maintaining efficient sulfur sorption capacity
during repeated sulfidation-regeneration cycles. (6) It is
likely that other contaminants such as HCl and NH
3
will
also require removal from raw syngas. Multifunctional mid-
to high-temperature sorbents for HCl, NH
3
, and H
2
S re-
moval could present significant advantages in terms of
operation and economics.
41
2. Sulfur Sorption Materials
2.1. Basic Sulfidation Reactions. Much of the research
conducted to date uses metal based sorbent materials that
are presumed to remove H
2
S according to the general
reaction:
M
x
O
y
s yH
2
Sg SM
x
S
y
s yH
2
Og 1
where M denotes a suitable metal. Equation 1 describes a
solid-gas reaction. This reaction is distinct from, and pre-
ceded by, a surface reaction that occurs when the time scale
for sulfidation is short.
42
Westemorland conducted a pioneering thermodynamics
study of the stable phases of various oxides andthe affinity of
these oxides for H
2
S adsorption and chose oxides or carbo-
nates of barium, calcium, cobalt, copper, iron, manganese,
molybdenum, strontium, tungsten, vanadium, and zinc as
viable candidates.
43
Slimane and Abbasian discussed in
detail how to link the thermodynamic results to practi-
cal experimental parameters such as regeneration con-
ditions.
44
Zinc, manganese, copper, iron, rare earth, and calcium
sorbents are among the most promising and most extensively
studied, and they are discussed in detail in Sections 2.2-2.7.
Selected results and experimental conditions are also
summarized in Table 3. Several other oxides, though theoreti-
cally capable of removing H
2
S, are not well suited for
desulfurization of syngas for varied reasons. For example,
molybdenum and tungsten oxides can form carbides, which
have poor desulfurization capabaility.
45
Strontium- and
barium-based carbonate materials behave similarly to cal-
cium carbonate, but calcium-based sorbents are preferable
because of their lower cost and wider operating temperature
range.
43
In addition to thermodynamic properties, kinetics of
desulfurization (or sulfidation of the solid sorbents) is critical
to establishing sorbent performance. Several studies indi-
cated the intrinsic reaction kinetics of the sulfidation of ZnO,
MnO, CaO, Fe, and Cu on ceria to have first-order depen-
dence in H
2
S.
46-48
Besides sorbent efficiency and kinetics, other side reac-
tions or catalyzing properties of the sorbents would need to
(38) Abbasian, J.; Slimane, R. B.; Lau, F. S.; Wangerow, J. R.;
Zarnegar, M. K. In Development of High Temperature Coal Gas Desul-
furization Systems ; An Overview, Fourteenth Annual International
Pittsburgh Coal Conference, Taiyuan, Shanxi, P.R. China, September
23-27, 1997; Pittsburgh Coal Conference, 1997; S4/21-S24/28.
(39) Fantom, I. R.; Cahill, P.; Sage, P. W. Hot gas cleaning ; An
overview. In Desulfurization of Hot Coal Gas, Atimtay, A. T.; Harrison, D.
P., Eds.; Springer: Berlin, 1998; Vol. 42, pp 103-116.
(40) Zhu, F.; Li, C.; Fan, H.; Li, Y. Meitan Zhuanhua (Coal Con-
version) 2000, 23 (2), 1722.
(41) Huber, G. W. Breaking the Chemical and Engineering Barriers to
Lignocellulosic Biofuels: Next Generation Hydrocarbon Biorefineries;
University of Massachusetts at Amherst. National Science Foundation,
Chemical, Bioengineering, Environmental, and Transport Systems Division:
Washington D.C., 2008.
(42) Flytzani-Stephanopoulous, M.; Sakbodin, M.; Zheng, W.
Science 2006, 312 (5779), 15081510.
(43) Westmoreland, P.; Harrison, D. P. Environ. Sci. Technol. 1976,
10 (7), 659661.
(44) Slimane, R. B.; Abbasian, J. Adv. Environ. Res. 2000, 4 (2), 147
162.
(45) Elseviers, W. F.; Verelst, H. Fuel 1999, 78, 601612.
(46) Tamhanker, S. S.; Bagajewicz, M.; Gavalas, G. R.; Sharma, P.
K.; Flytzani-Stephanopoulos, M. Ind. Eng. Chem. Proc. Des. Dev. 1986,
25, 429437.
(47) Westmoreland, P. R.; Gibson, J. B.; Harrison, D. P. Environ. Sci.
Technol. 1977, 11 (5), 488491.
(48) Flytzani-Stephanopoulous, M.; Li, Z. Kinetics of sulfidation
reactions between H
2
S and bulk oxide sorbents. In Desulfurization of
Hot Coal Gas, Atimtay, A. T.; Harrison, D. P., Eds.; Springer: Berlin, 1998;
Vol. 42, pp 179-211.
5295
Energy Fuels 2009, 23, 52915307
:
DOI:10.1021/ef900714q
T
a
b
l
e
3
.
M
a
t
e
r
i
a
l
s
,
S
u
l
f
i
d
a
t
i
o
n
,
a
n
d
R
e
g
e
n
e
r
a
t
i
o
n
C
o
n
d
i
t
i
o
n
s
F
r
o
m
S
e
l
e
c
t
e
d
P
u
b
l
i
c
a
t
i
o
n
s
r
e
f
s
m
a
t
e
r
i
a
l
s
s
u
l
f
i
d
a
t
i
o
n
c
o
n
d
i
t
i
o
n
s
o
u
t
l
e
t
[
H
2
S
]
r
e
g
e
n
e
r
a
t
i
o
n
c
o
n
d
i
t
i
o
n
s
G
i
b
s
o
n
,
I
I
I
,
a
n
d
H
a
r
r
i
s
o
n
(
1
9
8
0
)
a
Z
n
O
3
7
5
-
8
0
0

C
,
1
-
6
%
H
2
S
L
e
w
e
t
a
l
.
(
1
9
8
9
)
b
Z
n
O
-
T
i
O
2
6
0
0
-
6
5
0

C
,
1
%
H
2
S
i
n
1
3
%
H
2
,
1
9
%
H
2
O
,
6
7
%
N
2
<
1
0
p
p
m
i
n
a
l
l
c
a
s
e
s
7
0
0

C
,
1
0
%
a
i
r
-
9
0
%
N
2
W
o
o
d
s
e
t
a
l
.
(
1
9
9
0
,
9
1
)
c
Z
n
O
-
T
i
O
2
,
z
i
n
c
f
e
r
r
i
t
e
2
6
0
-
8
7
0

C
,
0
.
2
5
-
2
.
5
%
H
2
S
i
n
m
i
x
t
u
r
e
o
f
C
O
,
C
O
2
,
H
2
,
N
2
,
a
n
d
1
5
%
H
2
O
5
c
y
c
l
e
s
,
7
2
0
-
7
6
0

C
,
2
-
4
%
O
2
S
a
s
a
o
k
a
e
t
a
l
.
(
1
9
9
5
)
d
Z
n
O
5
0
0

C
,
5
2
0
p
p
m
C
O
S
,
5
0
0
a
n
d
1
1
0
0
p
p
m
H
2
S
J
u
n
e
t
a
l
.
(
2
0
0
4
)
e
Z
i
n
c
t
i
t
a
n
a
t
e
(
Z
T
)
p
r
o
m
o
t
e
d
b
y
C
o
a
n
d
N
i
o
x
i
d
e
s
4
8
0
/
6
5
0

C
,
1
.
5
%
H
2
S
i
n
m
i
x
,
a
m
b
i
e
n
t
p
r
e
s
s
u
r
e
1
0
-
1
5
c
y
c
l
e
s
,
3
-
5
%
O
2
/
N
2
a
t
5
8
0
-
8
0
0

C
F
e
n
o
u
i
l
a
n
d
L
y
n
n
(
1
9
9
5
)
f
u
n
c
a
l
c
i
n
e
d
l
i
m
e
s
t
o
n
e
(
9
7
.
8
%
C
a
C
O
3
)
5
7
0
-
8
6
0

C
,
0
.
5
-
1
.
8
5
%
H
2
S
i
n
m
i
x
,
a
m
b
i
e
n
t
P
F
e
n
o
u
i
l
a
n
d
L
y
n
n
(
1
9
9
5
)
g
c
a
l
c
i
n
e
d
l
i
m
e
s
t
o
n
e
(
9
7
.
8
%
C
a
C
O
3
)
5
6
0
-
1
1
0
0

C
,
0
.
0
5
-
1
.
8
%
H
2
S
i
n
m
i
x
,
a
m
b
i
e
n
t
P
Y
r
j
a
s
e
t
a
l
.
(
1
9
9
6
)
h
l
i
m
e
s
t
o
n
e
,
d
o
l
o
m
i
t
e
7
5
0
,
9
5
0

C
,
0
.
2
%
H
2
S
i
n
m
i
x
,
2
0
b
a
r
A
k
i
t
i
,
e
t
a
l
.
(
2
0
0
2
)
i
l
i
m
e
s
t
o
n
e
a
n
d
c
a
l
c
i
u
m
s
u
l
f
a
t
e
h
e
m
i
h
y
d
r
a
t
e
-
b
a
s
e
d
c
o
r
e
-
i
n
-
s
h
e
l
l
8
4
0
-
9
2
0

C
,
1
.
1
%
H
2
S
i
n
N
2
1
0
c
y
c
l
e
s
,
a
i
r
a
t
1
0
5
0

r
e
d
u
c
t
i
o
n
,
3
0
%
C
O
/
N
2
A
b
a
d
,
e
t
a
l
.
(
2
0
0
4
)
j
l
i
m
e
s
t
o
n
e
,
d
o
l
o
m
i
t
e
8
0
0
-
1
0
0
0

C
,
0
.
2
5
-
1
.
0
%
H
2
S
i
n
m
i
x
,
1
0
b
a
r
<
2
5
0
p
p
m
P
a
t
r
i
c
k
e
t
a
l
.
(
1
9
8
9
)
k
C
u
O
-
A
l
2
O
3
5
5
0
-
8
0
0

C
,
0
.
2
-
1
%
H
2
S
7
0
0
-
8
0
0

C
,
9
0
%
N
2
/
1
0
%
a
i
r
o
r
1
0
0
%
a
i
r
L
i
a
n
d
F
l
y
t
z
a
n
i
-
S
t
e
p
h
a
n
o
p
o
u
l
o
s
(
1
9
9
7
)
l
C
u
-
C
r
-
O
a
n
d
C
u
-
C
e
-
O
7
5
0
-
8
5
0

C
,
0
.
5
-
2
%
H
2
S
,
1
0
-
2
0
%
H
2
,
0
-
1
0
%
H
2
O
,
b
a
l
a
n
c
e
N
2
<
5
p
p
m
6
%
O
2
,
b
a
l
a
n
c
e
N
2
A
b
b
a
s
i
a
n
a
n
d
S
l
i
m
a
n
e
(
1
9
9
8
)
m
C
u
O

C
r
O
3
,
A
l
2
O
3
5
5
0
-
6
5
0

C
,
2
%
H
2
S
i
n
m
i
x
5
-
1
0
p
p
m
1
4
c
y
c
l
e
s
,
a
i
r
/
N
2
a
t
7
5
0

C
W
a
k
k
e
r
e
t
a
l
.
(
1
9
9
3
)
n
M
n
O
o
r
F
e
O
o
n

-
A
l
2
O
3
4
0
0
-
8
0
0

C
,
0
-
1
%
H
2
S
,
0
-
1
%
C
O
S
0
-
1
0
0
p
p
m
1
0
t
o
>
3
0
0
c
y
c
l
e
s
,
g
a
s
c
o
n
t
a
i
n
i
n
g
s
t
e
a
m
B
e
n
-
S
l
i
m
a
n
e
a
n
d
H
e
p
w
o
r
t
h
(
1
9
9
5
)
o
M
n
O
-

-
A
l
2
O
3
/
T
i
O
2
7
0
0
-
1
0
0
0

C
,
3
%
H
2
S
i
n
m
i
x

2
0
0
p
p
m
2
5
c
y
c
l
e
s
,
a
i
r
a
t
9
0
0

C
G
a
s
p
e
r
-
G
a
l
v
i
n
,
e
t
a
l
.
(
1
9
9
8
)
p
C
u
/
M
o
/
M
n
o
n
z
e
o
l
i
t
e
8
7
1

C
,
0
.
2
%
H
2
S
i
n
m
i
x
,
w
h
i
c
h
c
o
n
t
a
i
n
e
d
1
%
C
H
4
a
n
d
1
9
%
H
2
O
<
1
0
t
o
1
0
0
p
p
m
5
c
y
c
l
e
s
,
5
0
/
5
0
a
i
r
/
s
t
e
a
m
a
t
8
7
1

C
B
a
k
k
e
r
,
e
t
a
l
.
(
2
0
0
3
)
q

-
A
l
2
O
3

M
n
O
4
0
0
-
1
0
0
0

C
,
1
.
0
%
H
2
S
,
i
n
H
2
/
A
r
<
5
p
p
m
1
1
0
c
y
c
l
e
s
,
3
0
%
S
O
2
/
5
0
%
H
2
O
/
2
0
%
A
r
a
t
8
5
0

C
Y
o
o
n
e
t
a
l
.
(
2
0
0
4
)
r
M
n
o
r
e
(

-
M
n
O
2
)
5
5
0
-
8
5
0

C
,
1
.
0
%
H
2
S
i
n
m
i
x
<
1
p
p
m
2
0
c
y
c
l
e
s
,
4
%
O
2
/
N
2
a
t
5
5
0
-
8
5
0

C
o
r
4
%
O
2
/
N
2

0
.
5
%
N
H
3

5
-
2
0
%
H
2
O
;
f
o
l
l
o
w
e
d
b
y
C
O
/
H
2
/
C
O
2
/
N
2
r
e
d
u
c
t
i
o
n
A
l
o
n
s
o
a
n
d
P
a
l
a
c
i
o
s
(
2
0
0
2
)
s
Z
n
d
o
p
e
d
m
a
n
g
a
n
e
s
e
o
x
i
d
e
7
0
0

C
,
1
.
0
%
H
2
S
i
n
m
i
x
<
1
5
p
p
m
7
0
c
y
c
l
e
s
,
a
i
r
a
t
8
0
0

C
W
a
n
g
a
n
d
F
l
y
t
z
a
n
i
-
S
t
e
p
h
a
n
o
p
o
u
l
o
s
(
2
0
0
5
)
t
C
e
O
2

L
a
,
C
u
6
0
0
-
8
5
0

C
,
1
0
0
0
p
p
m
H
2
S
i
n
5
0
%
H
2
/
1
0
%
H
2
O
/
b
a
l
a
n
c
e
h
e
l
i
u
m
,
a
m
b
i
e
n
t
p
r
e
s
s
u
r
e
.
<
1
p
p
m
5
c
y
c
l
e
s
,
3
%
O
2
/
H
e
f
o
l
l
o
w
e
d
b
y
5
0
%
H
2
/
1
0
%
H
2
O
/
H
e
(
f
o
r
r
e
d
u
c
t
i
o
n
)
a
t
t
h
e
s
a
m
e
t
e
m
p
e
r
a
t
u
r
e
a
s
s
u
l
f
i
d
a
t
i
o
n
.
a
R
e
f
e
r
e
n
c
e
5
1
.
b
R
e
f
e
r
e
n
c
e
6
0
.
c
R
e
f
e
r
e
n
c
e
s
6
5
,
8
2
.
d
R
e
f
e
r
e
n
c
e
s
5
4
,
5
5
.
e
R
e
f
e
r
e
n
c
e
1
6
2
.
f
R
e
f
e
r
e
n
c
e
1
1
2
.
g
R
e
f
e
r
e
n
c
e
1
0
9
.
h
R
e
f
e
r
e
n
c
e
1
1
4
.
i
R
e
f
e
r
e
n
c
e
1
1
6
.
j
R
e
f
e
r
e
n
c
e
1
1
8
.
k
R
e
f
e
r
e
n
c
e
9
5
.
l
R
e
f
e
r
e
n
c
e
9
8
.
m
R
e
f
e
r
e
n
c
e
9
9
.
n
R
e
f
e
r
e
n
c
e
2
8
.
o
R
e
f
e
r
e
n
c
e
1
2
1
.
p
R
e
f
e
r
e
n
c
e
1
0
8
.
q
R
e
f
e
r
e
n
c
e
1
2
6
.
r
R
e
f
e
r
e
n
c
e
1
3
0
.
s
R
e
f
e
r
e
n
c
e
1
2
8
.
t
R
e
f
e
r
e
n
c
e
1
5
2
.
5296
Energy Fuels 2009, 23, 52915307
:
DOI:10.1021/ef900714q
be considered. Iron-, cobalt-, and copper-oxide based mate-
rials readily reduce in reducing environments at high tem-
perature.
49
Iron oxide and copper metal are both water gas
shift catalysts that promote conversion of CO and H
2
O to
CO
2
andH
2
, at 350-500 Candbelow200 C, respectively.
26
Consequently the potential effect of iron (and possibly
copper) sorbent components on the resulting CO/H
2
ratio
in the desulfurized syngas should be considered as they may
offer additional compositional tuning of the syngas for
downstream fuel synthesis.
2.2. Zinc-Based Materials. Zinc-based sorbents have
been used in desulfurization of natural gas feedstock,
typically at 370 C.
50
Consequently, it was one of the first
materials to be studied extensively for coal syngas desul-
furization. Early work with zinc oxide investigated the
effects of product layer diffusion, pore diffusion, and gas-
film diffusion on the kinetics of sulfidation. For example,
Gibson and Harrison (Table 3) attributed an apparent
20% utilization of the ZnO sorbent mass to be due to slow
diffusion from pore closure and formation of a dense
sulfide layer.
51
Sulfate formation is a factor that needs to be carefully
addressed when zinc-based sorbent is used.
50
Siriwardane
and Woodruff characterized reactions of water vapor and
oxygen with zinc sulfide at 550-650 C using Fourier
transform infrared spectroscopy (FTIR).
52,53
Oxygen pro-
motes the formation of SO
2
and sulfate on the surface, but
the extent of sulfate formation is less at higher tempera-
tures.
52
The presence of water vapor promotes the forma-
tion of sulfite (SO
3
2-
), at the expense of sulfate (SO
4
2-
). The
authors suggested that it may be advantageous to have
steam present during regeneration of ZnS since water
promotes formation of the sulfite species, which are easier
to decompose or remove from the surface than sulfate
species.
Sasaoka et al. studied the reaction of sulfided ZnO
sorbent with COS (Table 3).
54,55
They found that ZnS
catalyzes hydrolysis of COS to H
2
S, and the H
2
S formed
can be removed by ZnO downstream in a packed bed
reactor.
Though zinc oxide has a high H
2
S sorption capacity, at
temperatures above 600 C, early investigation by Gibson
and Harrison (Table 3) has clearly demonstrated that
reduction of ZnO in the highly reducing atmosphere of
syngas followed by vaporization of elemental zinc can
present a significant problem.
51
Consequently, a significant
part of the research on zinc-based material involved mod-
ifying it to improve its stability in the temperature range of
interest.
Sasaoka et al. examined the stability of zinc oxide with
ZrO
2
, TiO
2
, and Al
2
O
3
addition and concluded that even
though these oxides improve the performance, they do
not completely eliminate reduction and vaporization of
zinc.
56
Many other research groups, however, found
TiO
2
to be a very useful stabilizer of ZnO. Sections
2.2.1 and 2.2.2 summarize some of the key findings in
the development of zinc ferrite and zinc titanate sulfur
sorbents.
Baird et al. found that transition metals coprecipitated
with zinc oxide tend to produce mixed oxides of higher
surface areas.
57
Quantification of the improvements in per-
formance due to surface area increases versus other mechan-
isms will be tremendously useful.
Zinc-based materials remain promising, particularly in
the low- (<400 C) to midtemperature ranges. Turton et al.
conducted experiments of ZnO-based sorbents using ther-
mogravimetric analysis (TGA) and a small pilot-scale
transport reactor at 480-590 C.
58
The reaction of H
2
S
with the sorbent was modeled using a grainy-pellet model.
The transport reactor data were described with a plug-flow
model. The authors correlated laboratory and pilot scale
data and successfully predicted sorbent performance in the
transport reactor using their TGA data. More recently, a
zinc oxide-based sorbent developed by RTI International
was field-tested for 3000 h on a slipstream of a 1200 t/d
Eastman Chemical Company quench gasifier plant in Ten-
nessee.
34
2.2.1. Zinc Ferrite. Modifying zinc with iron oxide has
been found to produce a material that has high desulfuriza-
tion efficiency and capacity at approximately 500 C.
59-63
At higher temperature, disintegration of the sorbent can
reduce its performance. Focht et al. found that in the
temperature range 500-700 C a ZnFe
2
O
4
sorbent broke
down into Fe
2
O
3
and ZnO, and the produced Fe
2
O
3
further
reduced to Fe
3
O
4
or FeO, depending on the reduction
potential of the atmosphere.
64
Woods et al. reached similar
conclusions regarding reduction of zinc ferrite.
65
They also
conducted studies of the sulfidation and regeneration ki-
netics of zinc ferrite materials. Gupta et al. evaluated
various methods to synthesize zinc ferrite sorbents for
fluid-bed reactor applications and found that a granulation
technique produced an excellent sorbent that performed
well below 550 C.
66
Above that temperature, disintegra-
tion of the ferrite material into its metal oxide components
reduced its performance.
66
(49) Swisher, J. H.; Schwerdtfeger, K. J. Mater. Eng. Perform. 1992, 1
(3), 399408.
(50) Harrison, D. P. Performance analysis of ZnO-based sorbents in
removal of H
2
S from fuel gas. In Desulfurization of Hot Coal Gas,
Atimtay, A. T.; Harrison, D. P., Eds.; Springer: Berlin, 1998; pp 213-242.
(51) Gibson, J. B., III; Harrison, D. P. Ind. Eng. Chem. Proc. Des.
Dev. 1980, 19 (2), 231237.
(52) Siriwardane, R. V.; Woodruff, S. Ind. Eng. Chem. Res. 1995, 34
(2), 699702.
(53) Siriwardane, R. V.; Woodruff, S. Ind. Eng. Chem. Res. 1997, 36
(12), 52775281.
(54) Sasaoka, E.; Taniguchi, K.; Hirano, S.; Uddin, M. A.; Kasaoka,
S.; Sakata, Y. Ind. Eng. Chem. Res. 1995, 34 (4), 11021106.
(55) Sasaoka, E.; Taniguchi, K.; Uddin, A.; Hirano, S.; Kasaoka, S.;
Sakata, Y. Ind. Eng. Chem. Res. 1996, 35 (7), 23892394.
(56) Sasaoka, E.; Hirano, S.; Kasaoka, S.; Sakata, Y. Energy Fuels
1994, 8 (3), 763769.
(57) Baird, T.; Denny, P. J.; Hoyle, R.; McMonagle, F.; Stirling, D.;
Tweedy, J. J. Chem. Soc., Faraday Trans. 1992, 88 (22), 33753382.
(58) Turton, R.; Berry, D. A.; Gardner, T. H.; Miltz, A. Ind. Eng.
Chem. Res. 2004, 43 (5), 12351243.
(59) Gangwal, S. K.; Harkins, S. M.; Woods, M. C.; Jain, S. C.;
Bossart, S. J. Environ. Prog. 1989, 8 (4), 265269.
(60) Lew, S.; Jothimurugesan, K.; Flytzani-Stephanopoulos, M. Ind.
Eng. Chem. Res. 1989, 28, 535541.
(61) Ayala, R. E.; Marsh, D. W. Ind. Eng. Chem. Res. 1991, 30 (1),
5560.
(62) Grindley, T.; Steinfeld, G. Development and Testing of Regener-
able Hot-Coal-Gas Desulfurization Sorbents; DOE/MC/16545-1125;
CONF-820610-3; 1981.
(63) Jha, M. C.; Blandon, A. E.; Hepworth, M. T. Durable Zinc
Ferrite Sorbent Pellets for Hot Coal Gas Desulfurization. U.S. Patent
4,732,888, Mar. 22, 1988.
(64) Focht, G. D.; Ranade, P. V.; Harrison, D. P. Chem. Eng. Sci.
1988, 43 (11), 30053013.
(65) Woods, M. C.; Gangwal, S. K.; Harrison, D. P.; Jothimurugesan,
K. Ind. Eng. Chem. Res. 1991, 30 (1), 100107.
(66) Gupta, R.; Gangwal, S. K.; Jain, S. C. Energy Fuels 1992, 6 (1),
2127.
5297
Energy Fuels 2009, 23, 52915307
:
DOI:10.1021/ef900714q
Kobayashi et al. studied zinc ferrite and zinc ferrite-silica
composite sorbents.
67-71
Using in situ X-ray diffraction,
Mossbauer spectroscopy, and fixed bed reactor tests, they
showed that when the H
2
S concentration was less than
80 ppmv, zinc was the reactive sorbent and a mixture of zinc
sulfide and iron oxide (Fe
3
O
4
and FeO) formed. At higher
H
2
S concentrations, both zinc and iron were reactive and
formed zinc and iron sulfides. The authors conducted sulfi-
dation and regeneration for 20 cycles and used that data to
extrapolate the loss in activity to 500 cycles of desulfuriza-
tion. They concluded that verification of the long-term
durability and activity of the sorbent would require further
testing of up to 40-50 total cycles.
71
In a separate study, Ayala and Marsh characterized and
conducted long-range testing (50 cycles of sulfidation and
regeneration) of zinc ferrite sorbent at approximately 550 C
and detected no undesired metal carbide, sulfate, and ele-
mental iron based on XRD.
61
Gangwal et al. evaluated zinc ferrite based sorbents
59,72
and
found that in as little as 5%steam, soot formation, which they
hypothesized to be catalyzed by iron, was a process chal-
lenge.
59
Sasaoka et al. also investigated the zinc ferrite system
and found that soot formed on zinc ferrite sorbent at 500 C
under atmospheric pressure.
73
Soot formation decreased as
reaction temperature (450-600 C) increased, and soot for-
mation was accelerated by the presence of H
2
and CO, but
suppressed by H
2
Oand CO
2
. On the basis of scanning electron
microscopy and XRD data, they hypothesized that soot and
iron carbide (Fe
3
C and Fe
x
C) formation are correlated.
Besides the binary zinc ferrite system, a number of ternary
modified zinc ferrite systems were also studied. Siriwardane
and Poston investigated the zinc copper ferrite system.
74
The
starting material was 28.5 mol % ZnO, 65% Fe
2
O
3
, 4.6%
CuO, and 2% bentonite, but XRD showed that the synthe-
sized material did not significantly incorporate copper and
had a primary composition of ZnFe
2
O
4.
This XRD result
suggests that copper may exist in an amorphous phase or is
bound to zinc and iron to form dispersed species that cannot
be detected by XRD. X-ray photoelectron spectroscopy
(XPS), however, showed that when this material was heated
to 550 C, most of the copper was on the surface in the 1
oxidation state. Enrichment of copper on the surface was
attributed by Siriwardane and Poston to the high perfor-
mance of zinc copper ferrite as sulfur sorbent. In a reducing
environment, iron in zinc copper ferrite exists in both Fe
3
and Fe
2
oxidation states.
Pineda et al. characterized zinc oxide and zinc ferrite
sorbents doped with titanium or copper using XRD and
Raman spectroscopy.
75
For the study on zinc copper ferrite,
they fixed the iron atomic content at 50% and varied the
copper content from7 to 40%.
75
They concluded that copper
does not affect the stability of zinc ferrite. X-ray photoelec-
tron spectroscopy measurements indicate that the ratioof Cu
to Fe on the surface is always much higher than the bulk,
indicating copper migration to the surface. The authors
hypothesized that copper migrates to the surface during
calcinations and regeneration steps. They attributed the
improved sorbent performance (lower H
2
S concentration
after sulfidation) to the enhanced level of copper on the
surface, in agreement with results from Siriwardane and
Poston.
74
The finding of high H
2
S removal with copper
addition into zinc ferrite is in agreement with results reported
by Gangwal et al.,
59
who found that addition of copper into
zinc ferrite improves its efficiency, and the copper modified
zinc ferrite sorbent removes H
2
S to less than 1 ppmv at
600 C even with 20% steam in the atmosphere. Pineda et al.
also found that during sulfidation Cu-Zn-Fe mixed oxides
are completely converted into low oxidation sulfides, that is,
the sulfide species on the surface are Cu
2
S, FeS, and ZnS and
that in the case of FeS there were a lot of Fe vacancies.
75
Multicycle tests of these sorbents show that the decay in
performance is not related to structural changes but to
decreased porosity.
76
Pineda et al. also investigated the effect of incorporating
titaniuminto zinc ferrite, with zinc content fixed at 50 atom%,
and titanium content varied from 7 to 40 atom %.
75
At
650 C, the inclusion of titanium actually impeded the
formation of zinc ferrite. However, with calcination con-
ducted at 1100 C, the addition of titanium to zinc ferrite
prevented decomposition into its component oxides, that is,
titanium addition stabilized the ferrite lattice. The stabiliza-
tion of titanium is particularly effective in noniron-contain-
ing sorbents.
77
In summary, zinc ferrite is an efficient sorbent at approxi-
mately 500 C. Efforts to produce ternary materials included
preparation of zinc copper ferrite, which can reduce H
2
S to
lower levels than zinc ferrite, and manufacture of zinc ferrite
doped with titanium, which shows higher stability under
certain preparation conditions. The interest in zinc ferrite
materials for moderate temperature desulfurization con-
tinues, and more recently the sol-gel method has been
explored as a way to synthesize a high surface area zinc
ferrite sorbent.
78
2.2.2. Zinc Titanate. Combining zinc and titaniumdioxide
is one of the most successful approaches to stabilize ZnO
against reduction at mid- to high-temperature ranges. Lewet
al. conducted extensive research on the kinetics and mechan-
isms of reduction and sulfidation of a series of Zn-Ti-O
materials with different zinc to titanium ratios (Table 3).
79-81
Lew et al. found that between 400 and 700 C, the activation
energies for sulfidation of Zn-Ti-Oand ZnOsorbents were
similar (9-10 kcal/mol), indicating that the sulfidation
(67) Kobayashi, M.; Shirai, H.; Nunokawa, M. Energy Fuels 1997, 11
(4), 887896.
(68) Kobayashi, M.; Shirai, H.; Nunokawa, M. Ind. Eng. Chem. Res.
2000, 39 (6), 19341943.
(69) Kobayashi, M.; Shirai, H.; Nunokawa, M. Energy Fuels 2002, 16
(3), 601607.
(70) Kobayashi, M.; Shirai, H.; Nunokawa, M. Ind. Eng. Chem. Res.
2002, 41 (12), 29032909.
(71) Kobayashi, M.; Shirai, H.; Nunokawa, M. Energy Fuels 2002, 16
(6), 13781386.
(72) Gangwal, S. K.; Stogner, J. M.; Harkins, S. M.; Bossart, S. J.
Environ. Prog. 1989, 8 (1), 2634.
(73) Sasaoka, E.; Iwamoto, Y.; Hirano, S.; Uddin, M. A.; Sakata, Y.
Energy Fuels 1995, 9 (2), 344353.
(74) Siriwardane, R. V.; Poston, J. A. Appl. Surf. Sci. 1993, 68 (1), 65
80.
(75) Pineda, M.; Fierro, J. L. G.; Palacios, J. M.; Cilleruelo, C.;
Garcia, E.; Ibarra, J. V. Appl. Surf. Sci. 1997, 119 (1-2), 110.
(76) Pineda, M.; Palacios, J. M.; Alonso, L.; Garcia, E.; Moliner, R.
Fuel 2000, 79 (8), 885895.
(77) Garcia, E.; Cilleruelo, C.; Ibarra, J. V.; Pineda, M.; Palacios, J.
M. Ind. Eng. Chem. Res. 1997, 36 (3), 846853.
(78) Zhang, R.; Huang, J.; Zhao, J.; Sun, Z.; Wang, Y. Energy Fuels
2007, 21 (5), 26822687.
(79) Lew, S.; Sarofim, A. F.; Flytzani-Stephanopoulos, M. Chem.
Eng. Sci. 1992, 47 (6), 14211431.
(80) Lew, S.; Sarofim, A. F.; Flytzani-Stephanopoulos, M. Ind. Eng.
Chem. Res. 1992, 31 (8), 18901899.
(81) Lew, S.; Sarofim, A. F.; Flytzani-Stephanopoulos, M. AlChE J.
1992, 38 (8), 11611169.
5298
Energy Fuels 2009, 23, 52915307
:
DOI:10.1021/ef900714q
mechanism on these two types of solids is likely the same.
80
However, for sorbents containing more than 25 mol % Ti,
the initial sulfidation rate of Zn-Ti-O sorbents was 1.5-2
times slower than that of ZnO, suggesting the frequency
factor in the rate expression for Zn-Ti-O is smaller, that is,
there are fewer reaction sites on Zn-Ti-O than ZnO. The
smaller number of reaction sites on Zn-Ti-Owas presumed
to be due to nonreactive titanium on the surface.
80
Lew et al.
conducted studies on the kinetics of Zn-Ti-O reduction
and found that in H
2
-N
2
gas mixtures, Zn-Ti-O solids
have a lower reduction rate than ZnO in the temperature
range 550-1050 Candthe associatedactivationenergies for
Zn-Ti-O and ZnO reduction were calculated at 37 and
24 kcal mol
-1
, respectively.
79
Modeling of the results with
the overlapping grain model suggests that the primary
limitation to reaction rate on both Zn-Ti-O and ZnO is
diffusion through a ZnS product layer.
81
Gangwal et al. studied various binary oxides of zinc and
found that zinc titanate had excellent mechanical strength
and capacity retention at 700 C.
59,72
The reduction,
sulfidation, and regeneration kinetics of zinc oxide-tita-
nium oxide single pellet sorbents were studied and mod-
eled by Woods et al.
82
(Table 3) and Jothimurugesan and
Harrison.
83
In agreement with Lewet al.,
79
they found that
the addition of titanium oxide increased the maximum
sorbent operating temperature range by stabilizing zinc
oxide against reduction and subsequent volatilization.
Mass transfer and product layer diffusion resistance were
the main parameters that controlled the global reaction
rate.
Pineda et al.
75
found spectroscopic evidence that the
addition of titanium increases the stability of ZnO against
reduction through the formation of Zn
2
TiO
4
, in agreement
with Lewet al. and Woods et al.
79,82
Because of the interest in
ZnO-TiO
2
materials, Yang and Swisher studiedthe stability
of Zn
2
Ti
3
O
8
in detail and provided additional information
on the ZnO-TiO
2
phase diagram.
84
Siriwardane and Poston studied the sulfidation of zinc
titanate in the presence of H
2
and CO.
85
Using XPS to study
the surface species after H
2
S, CO, and H
2
exposures at
670-800 C, they deduced that when Zn
2
TiO
4
reacts with
H
2
S, zinc is sulfided while titanium is not, with oxygen
release according to the reaction:
Zn
2
TiO
4
2H
2
SS2ZnSTiO
2
2H
2
O
2
2
The authors also concluded that some of the O
2
can
oxidize H
2
S to SO
2
and also ZnS to ZnSO
4
. The reaction
of SO
2
withZn
2
TiO
4
canalso result in ZnSO
4
formation. Ina
reducing syngas, both H
2
and CO can reduce ZnSO
4
to
ZnO.
85
Hatori et al. studied the role of TiO
2
in oxidative regen-
eration of a ZnO-TiO
2
sorbent.
86
They found that TiO
2
increased the oxidative regeneration rate by accelerating the
reactionbetweenZnSandO
2
as well as that betweenZnSand
H
2
O. The rate increase could be explained by activation of
H
2
O over TiO
2
and subsequent spillover of the activated
H
2
O species to react with ZnS. The overall reaction is
ZnS3H
2
OSZnOSO
2
3H
2
3
3H
2
3=2O
2
S3H
2
O 4
A few research groups reported on modifying or doping
Zn-Ti based materials further with other oxides, and this
approach yielded some promising results. Sasaoka et al.
investigated modification of a 50% ZnO-TiO
2
mixture, a
mixed oxide that primarily has the Zn
2
Ti
3
O
8
structure.
87
Addition of 5-10% ZrO
2
improved its reactivity for H
2
S
removal and regenerability. They authors hypothesized that
addition of ZrO
2
improved the pore structure and helped to
maintain a large surface area after regeneration.
87
Siriwardane et al. studied molybdenum-containing zinc
titanate sorbents.
88
They found evidence of sulfate forma-
tion, and that changes in the structure due to the large
volume difference between sulfate and oxide during sulfida-
tion and regeneration may have contributed to sorbent
spalling. In addition, they also found that when regeneration
was conducted at 649-760 C, there was incomplete refor-
mation of the titanate structure, and that the degree of
incomplete reformation increased (amount of pure TiO
2
increased) with an increasing number of sulfidation and
regeneration cycles.
Liu et al. investigated the effect of V
2
O
5
, B
2
O
3
, and
tungsten doping on sintering and phase transition of zinc
titanate ceramics.
89-91
Jothimurugesan and Gangwal eval-
uated the coprecipitation of a series of transition metal
oxides with zinc-titanum oxides as a means of lowering
the regeneration temperature of zinc titanate from 650 C.
92
They found that a combination of a small weight percent of
nickel and cobalt added to zinc titanate effectively reduced
the regeneration temperature by at least 100 C without
impacting sorbent performance in subsequent cycles.
92
Jun et al.
93
investigated cobalt oxide-doped zinc titanate
further and found it to have higher sulfur sorption capacity
and better regenerability (particularly at 480 C) than un-
doped zinc titanate (Table 3). They hypothesized that the
modified materials have better regenerability partly because
addition of cobalt into the structure minimized volume
expansion and contraction during sulfidation and regenera-
tion.
93
A separate report by the same research group found
that the doped material was capable of decomposing NH
3
.
94
2.3. Copper-Based Materials. Copper-based sorbents have
been widely investigated because of the favorable equilibri-
um between copper oxides and H
2
S. Copper oxide is able to
achieve low levels of H
2
S in the clean fuel gas provided the
sorbent is not reduced to elemental copper. This is because
copper oxide readily reduces in high temperature reducing
(82) Woods, M. C.; Gangwal, S. K.; Jothimurugesan, K.; Harrison,
D. P. Ind. Eng. Chem. Res. 1990, 29 (7), 11601167.
(83) Jothimurugesan, K.; Harrison, D. P. Ind. Eng. Chem. Res. 1990,
29 (7), 11671172.
(84) Yang, J.; Swisher, J. H. Mater. Charact. 1996, 37 (2-3), 153159.
(85) Siriwardane, R. V.; Poston, J. A. Appl. Surf. Sci. 1990, 45 (2),
131139.
(86) Hatori, M.; Sasaoka, E.; Uddin, M. A. Ind. Eng. Chem. Res.
2001, 40 (8), 18841890.
(87) Sasaoka, E.; Sada, N.; Manabe, A.; Uddin, M. A.; Sakata, Y.
Ind. Eng. Chem. Res. 1999, 38 (3), 958963.
(88) Siriwardane, R. V.; Poston, J. A.; Evans, G. Ind. Eng. Chem. Res.
1994, 33 (11), 28102818.
(89) Liu, X. C.; Gao, F.; Zhao, L. L.; Tian, C. S. J. Alloys Compd.
2007, 436 (1-2), 285289.
(90) Liu, X. C.; Gao, F.; Zhao, L. L.; Tian, C. S. J. Mater. Sci. -
Mater. Electron. 2007, 18 (8), 863868.
(91) Liu, X. C.; Gao, F.; Zhao, L. L.; Zhao, M.; Tian, C. S. J.
Electroceram. 2007, 18 (1-2), 103109.
(92) Jothimurugesan, K.; Gangwal, S. K. Ind. Eng. Chem. Res. 1998,
37 (5), 19291933.
(93) Jun, H. K.; Lee, T. J.; Ryu, S. O.; Kim, J. C. Ind. Eng. Chem. Res.
2001, 40 (16), 35473556.
(94) Jun, H. K.; Jung, S. Y.; Lee, T. J.; Ryu, C. K.; Kim, J. C. Catal.
Today 2003, 87, 310.
5299
Energy Fuels 2009, 23, 52915307
:
DOI:10.1021/ef900714q
atmospheres, and elemental copper is an order of magnitude
less active for sulfidation than Cu
2
O and CuO. Stabilizing
copper oxide with a variety of metal oxides has been quite
successful.
Tamhankar et al. studied several pure and mixed oxide
sorbents (ZnO, CuO, ZnO-Fe
2
O
3
, CuO-Fe
2
O
3
, CuO-
Al
2
O
3
, and CuO-Fe
2
O
3
-Al
2
O
3
) prepared as porous solids
with very high pore volume.
46
Both Fe
2
O
3
and Al
2
O
3
play a
role in stabilizing Cu in the oxidized state in the temperature
range 538-600 C. However, even in the mixed oxides,
copper is reduced stepwise from oxidation state 2 to
1 to 0, particularly at and above 650 C. The sorbent
CuO-Fe
2
O
3
-Al
2
O
3
performs better than the binary mixed
oxides.
Patrick et al. found sulfidation of CuO-Al
2
O
3
sorbent
produced digenite (Cu
9x
S
5
) and that alumina stabilizes
CuO against complete reduction to Cu at 550-800 C,
contributing to sulfidation of copper at the more desirable
oxidation states of 1 or 2 (Table 3).
95
They found sulfate
formation during regeneration step, which they attributed to
either copper sulfate or surface aluminum sulfate. Sick and
Schwerdtfeger studied desulfurization with copper particles
embedded in alumina pellets and were able to lower H
2
S
levels from 3000 to 300 and 150 ppmv at 750 and 650 C,
respectively.
96
Yoo et al. found that in an oxidizing environ-
ment, sulfation of -Al
2
O
3
support to form Al
2
(SO
4
)
3
oc-
curred,
97
and the extent of aluminum sulfation (surface
or bulk sulfate formation) depends on temperature and
copper loadings. Their results confirmthe findings of Patrick
et al.,
95
and both copper and aluminum sulfation need to be
taken into account in regeneration of copper-on-alumina
sorbents.
Two other metals that have been investigated exten-
sively for stabilization of copper are chromium and man-
ganese. Li and Flytzani-Stephanopoulos studied the
binary Cu-Cr-O and Cu-Ce-O oxides (Table 3) for
desulfurization, including their kinetics of reduction, sul-
fidation, and regeneration under a variety of conditions at
650-850 C.
98
They found that in the CuO-Cr
2
O
3
mix-
ture, the stable compound copper chromite (CuCr
2
O
4
) was
formed, which has the lowest reducibility of all copper
oxide-containing compounds in the literature. Thus,
this compound, which contains copper in the 1 or 2
oxidation states, has very high H
2
S removal efficiency,
resulting in H
2
S levels of less than 5 ppmv prior to break-
through.
Abbasian and Slimane researched copper chromium
mixed oxide and copper supported on alumina
(Table 3).
99
They used thermodynamic analysis to ratio-
nalize the selection of chromia (Cr
2
O
3
), for the stabiliza-
tion of copper oxide (Cu
2
O) against complete reduction to
elemental copper. The Cu-Cr-O mixed sorbent, most
efficient at 600 C, was able to reduce H
2
S levels to
less than 5 ppmv. Regeneration with a dilute O
2
-N
2
mixture at 750 C proved to be an efficient protocol
without sulfate formation or reactivity deterioration over
15 cycles.
99
Slimane and Abbasian developed formulations of copper
supported on alumina and manganese oxide matrix
(MnAl
2
O
4
was alsodetected inthe XRD), whichwas suitable
for use over the temperature range 350-600 C.
100
These
highly attrition resistant sorbents were able to remove H
2
S
concentrations to <1 ppmv and were regenerable at
650-725 C using a 6% O
2
-N
2
gas mixture.
Alonso et al. and Garcia et al. studied manganese
copper mixed sorbents with the manganese to copper ratio
ranging from 0.13 to 1.6.
101,102
They used XPS, XRD,
scanning electron microscopy (SEM), FTIR, and tempera-
ture programmed reduction (TPR) to characterize freshly
synthesized sorbents, sorbents exposed to simulated syn-
gas with and without H
2
S, and regenerated samples.
101
They found that their freshly synthesized sorbents con-
tained two copper phases, one consisting mainly of CuO
with inclusions of manganese ions, whereas the other was
the spinel structure CuMn
2
O
4
. Upon exposure to simu-
lated coal gas (10% H
2
, 15% H
2
O, 5% CO
2
, and 15% CO,
balance N
2
) for 2 h, the sorbents reduced to metallic Cu
and MnO, indicating manganese oxide does not stabilize
copper against reduction.
101
Nevertheless, the use of cop-
per was necessary to achieve low(sub-ppm) levels of H
2
S in
the outlet.
102
Sulfidation at a minimum of 700 C and
regeneration at 800 C with diluted air was found to be the
optimal conditions. This is because 700 and 800 C are the
minimum temperature where MnSO
4
is unstable under
reducing and oxidizing conditions, respectively. It was
found that at these temperatures in simulated syngas there
was no performance deterioration after five sulfidation
and regeneration cycles.
102
Karayilan et al. also studied Mn-Cu and Mn-Cu-V
mixed oxide sorbents prepared with a complexation meth-
od.
103
The major crystalline phases in the Mn-Cu mixed
oxide were Cu
1.5
Mn
1.5
O
4
and CuMn
2
O
4
.
103
They found that
regeneration at temperatures lower than 700 C resulted in
significant reduction in activity, which they attributed to
sulfate formation. When conducting sulfidation at 627 C
and regeneration at 700 Cwith a gas mixture containing 6%
O
2
in nitrogen, they found less than 10% degradation in
activity after 5 cycles.
Yasyerli et al. studied the reaction of copper oxide and two
mixed oxides, Cu-V and Cu-Mo, with H
2
S at 300 and
700 C.
104,105
They found that a significant amount of SO
2
was produced with a CuO sorbent in the absence of hydro-
gen. X-ray diffractionstudies of the sulfidedsamples indicate
the major phase of solid product formedis Cu
1.8
S, suggesting
that some of the H
2
S reduces copper in the CuO oxide, with
the concurrent formation of SO
2
. When Cu-V and Cu-Mo
sorbents are used, SO
2
formation was detected even in the
presence of 10% H
2
.
(95) Patrick, V.; Gavalas, G. R.; Flytzani-Stephanopoulos, M.; Jothi-
murugesan, K. Ind. Eng. Chem. Res. 1989, 28 (7), 931940.
(96) Sick, G.; Schwerdtfeger, K. Metall. Trans. B 1987, 18 (3), 603
609.
(97) Yoo, K. S.; Kim, S. D.; Park, S. B. Ind. Eng. Chem. Res. 1994, 33
(7), 17861791.
(98) Li, Z. J.; Flytzani-Stephanopoulos, M. Ind. Eng. Chem. Res.
1997, 36 (1), 187196.
(99) Abbasian, J.; Slimane, R. B. Ind. Eng. Chem. Res. 1998, 37 (7),
27752782.
(100) Slimane, R. B.; Abbasian, J. Ind. Eng. Chem. Res. 2000, 39 (5),
13381344.
(101) Alonso, L.; Palacios, J. M.; Garcia, E.; Moliner, R. Fuel
Process. Technol. 2000, 62 (1), 3144.
(102) Garcia, E.; Palacios, J. M.; Alonso, L.; Moliner, R. Energy Fuels
2000, 14 (6), 12961303.
(103) Karayilan, D.; Dogu, T.; Yasyerli, S.; Dogu, G. Ind. Eng. Chem.
Res. 2005, 44, 52215226.
(104) Yasyerli, S.; Dogu, G.; Ar, I.; Dogu, T. Ind. Eng. Chem. Res.
2001, 40, 52065214.
(105) Yasyerli, S.; Dogu, G.; Ar, I.; Dogu, T. Chem. Eng. Commun.
2003, 190 (5-8), 10551072.
5300
Energy Fuels 2009, 23, 52915307
:
DOI:10.1021/ef900714q
Atimtay et al. and Gasper-Galvin et al. explored incorpor-
ating copper, manganese, and molybdenum into silica-rich
zeolite to stabilize these metal sorbents (Table 3).
106-108
Those formulations with Cu had the highest sulfur capacity
and copper was found to be the primary desulfurization
agent with manganese and molybdenum acting as promo-
ters.
2.4. Calcium-Based Materials. Calcium-based sorbents
such as limestone (CaCO
3
) or dolomite (CaCO
3
3
MgCO
3
)
are relatively low cost and can be used in both reducing and
oxidizing conditions. Equilibrium calculations and experi-
ments by several research groups have shown that calcium is
most active for H
2
S removal near 880 C, the temperature at
which CaCO
3
decomposes to CaO(lime) and CO
2
.
43,109
This
property makes it suitable for systems operating at higher
temperatures and thus attractive from a thermal efficiency
standpoint. However, other research suggested that H
2
S
removal by limestone was restricted to conditions far away
from the calcination equilibrium and that the simultaneous
presence of CO
2
and H
2
S would inhibit both calcination and
sulfidation reactions.
110
More importantly, due to thermo-
dynamic constraints, H
2
S removal efficiencies with calcium-
containing materials are only on the order of 90% under
typical gasification conditions, resulting in residual H
2
S
levels of 100 ppm or greater. This may limit the sorbents
usefulness strictly to bulk H
2
S removal, requiring an external
bed to further polish the gas if an application requires more
stringent sulfur cleanup.
Fenouil and Lynn studied the kinetics of H
2
S sorption by
uncalcined and calcined limestone.
109,111,112
They found that
for the uncalcined material, the kinetics are reaction limited
below 660 C, and diffusion limited above 660 C. The
conversion of uncalcined CaCO
3
to CaS was limited to
10% (Table 3).
112
For precalcined limestone, the conversion
from CaO to CaS was 100% in an hour, and the sulfidation
rate for calcined material was largely insensitive to tempera-
ture (Table 3).
109
Yrjas et al. and Zevenhoven et al. studied reactions of
limestone (CaCO
3
) and dolomite (CaMg(CO
3
)
2
) under ele-
vated pressure conditions.
113-115
They found that calcined
limestone, half calcined dolomite (CaCO
3
MgO), and fully
calcined dolomite (CaO MgO) all have faster sulfidation
kinetics and higher percentages of sorbent conversion than
uncalcined limestone (Table 3).
114
The unreacted shrinking
core model with variable effective diffusivity was used to
model the data.
115
The results of Fenouil and Lynn as well as
those of Yrjas et al. and Zevenhoven et al. indicate that
precalcination of limestone and dolomite leads to faster
sulfidation kinetics and higher sorbent conversion.
Very few studies considered regenerating spent calcium-
based sorbents largely due to the fact that limestone and
dolomite are relatively soft and easily broken up,
116
leading
to high attrition rates in fluidized applications. In addition,
they tend to form a stable sulfate layer during regeneration,
leading to a loss in active material. As a result, calcium-based
sorbents are usually considered for once-through applica-
tions, requiring stabilization and disposal of large amounts
of solid CaS, although this can be partially offset by conver-
sion to salable gypsum (CaSO
4
3
2H
2
O). Some attempts have
been made to improve the attrition resistance and reactivity
of limestone and dolomite by forming agglomerates of the
pulverized material with various binders.
117
Akiti and co-workers reported promising results using
pelletized calcium sulfate hemihydrate (plaster of Paris)
coated onto a porous shell of powdered alumina and lime-
stone.
116
When heat treated, the resulting core-in-shell
materials were reported to be mechanically strong, active
for H
2
S removal, and regenerable for 10 cycles with no loss in
activity (Table 3). However, several oxidation/reduction
cycles were required to regenerate these sorbents.
Abad et al. investigated desulfurization of coal gas using
limestone and dolomite in a moving bed reactor.
118
They
found both countercurrent and concurrent configurations to
have high desulfurization levels but the countercurrent con-
figuration was more effective (Table 3).
Bjorkman and Sjostrom found that dolomite can also
catalyze ammonia decomposition but the catalytic activity
was inhibited in the presence of hydrocarbons and steam.
119
They hypothesized that the hydrocarbons formed carbonac-
eous materials on dolomite that likely inhibited ammonia
decomposition.
2.5. Manganese-Based Materials. In a reducing gas envir-
onment, manganese oxides of higher oxidation states are
likely reduced to MnO.
120
The thermodynamics of MnO
sulfidation is not as favorable as some other metal oxides
such as zinc oxide and copper oxide. However, MnO does
not decompose to elemental manganese readily in reducing
gas environment. Thus, manganese-based materials offer the
advantage of stability at high temperatures, which better
matches biomass gasification and tar reforming process
temperatures. The potential disadvantage is that manga-
nese-based sorbents are prone to sulfate formation and have
to be regenerated at very high temperature.
101,121
Ben-Slimane and Hepworth studied manganese-based
sorbents for desulfurization and modeled the kinetics of
the sulfidation, obtained through TGA studies, using the
shrinking core model.
120
In this model the reaction proceeds
at a narrow front which moves into the solid particle. Three
different diffusion controlled rate models, gas film diffusion
control, product layer diffusion control, and surface reaction
control, were used to fit the data. Product layer (MnS)
diffusion control was found to best fit the experimental data.
(106) Atimtay, A. T. Development of supported sorbents for hydro-
gen sulfide removal from fuel gas. In Desulfurization of Hot Coal Gas,
Atimtay, A. T.; Harrison, D. P., Eds.; Springer: Berlin, 1998; Vol. 42, pp
315-329.
(107) Atimtay, A. T.; Gasper-Galvin, L. D.; Poston, J. A. Environ.
Sci. Technol. 1993, 27 (7), 12951303.
(108) Gasper-Galvin, L. D.; Atimtay, A. T.; Gupta, R. P. Ind. Eng.
Chem. Res. 1998, 37, 41574166.
(109) Fenouil, L. A.; Lynn, S. Ind. Eng. Chem. Res. 1995, 34 (7), 2334
2342.
(110) de Diego, L. F.; Abad, A.; Garcia-Labiano, F.; Adanez, J.;
Gayan, P. Ind. Eng. Chem. Res. 2004, 43 (13), 32613269.
(111) Fenouil, L. A.; Lynn, S. Ind. Eng. Chem. Res. 1995, 34 (7), 2343
2348.
(112) Fenouil, L. A.; Lynn, S. Ind. Eng. Chem. Res. 1995, 34 (7), 2324
2333.
(113) Yrjas, K. P.; Zevenhoven, C. A. P.; Hupa, M. M. Ind. Eng.
Chem. Res. 1996, 35 (1), 176183.
(114) Yrjas, P.; Iisa, K.; Hupa, M. Fuel 1996, 75 (1), 8995.
(115) Zevenhoven, C. A. P.; Yrjas, K. P.; Hupa, M. M. Ind. Eng.
Chem. Res. 1996, 35 (3), 943949.
(116) Akiti, T. T.; Constant, K. P.; Doraiswamy, L. K.; Wheelock, T.
D. Ind. Eng. Chem. Res. 2002, 41 (3), 587597.
(117) Voss, K. E. Limestone-Based Sorbent Agglomerates for Removal
of Sulfur Compounds in Hot Gases and Method of Making. U.S. Patent
4,316,813, Feb 23, 1982.
(118) Abad, A.; Adanez, J.; Garcia-Labiano, F.; de Diego, L. F.;
Gayan, P. Energy Fuels 2004, 18 (5), 15431554.
(119) Bjorkman, E.; Sjostrom, K. Energy Fuels 1991, 5 (5), 753760.
(120) Ben-Slimane, R.; Hepworth, M. T. Energy Fuels 1994, 8, 1175
1183.
5301
Energy Fuels 2009, 23, 52915307
:
DOI:10.1021/ef900714q
The thermodynamic data for the Mn-S-O system have
considerable variability. Turddogan as well as Ben-Slimane
and Hepworth evaluated the data to determine which ther-
modynamic information should be given more credence.
121,122
Ben-Slimane and Hepworth then used the information to
help design regeneration conditions for a manganese sorbent
they had synthesized. They concluded that oxidative regen-
eration is more favorable and that a temperature as high as
900 C is necessary to avoid sulfate formation.
123
The
sorbent was tested for 25 cycles and showed no degradation
in performance.
121
Atakul et al. studied the sulfidation and regeneration of
MnO/-Al
2
O
3
sorbent.
124,125
Both sulfidation and regenera-
tion were conducted at 600 C. Steamwas more efficient than
hydrogen for regeneration, and the rate of sulfur removal
was found to be proportional to steam content. Steam
consumption for regeneration increased slowly during the
initial 60-70%of regeneration and rose rapidly afterward to
reach complete regeneration, that is, regeneration was sig-
nificantly slower after about 60-70% regeneration.
Bakker et al. developed both regenerable monolith- and
particle-based sorbents consisting mostly of crystalline
Mn
3
O
4
.
126
The material is able to remove H
2
S from gases
with inlet concentrations of 6400 ppm down to 5-50 ppm.
After sulfidation, XRD and TEM detected mostly MnA-
l
2
O
4
, MnS, and some amorphous phases. After regeneration,
there was still considerable MnAl
2
O
4
, together with MnO
and amorphous phases. The researchers were able to use SO
2
to directly regenerate elemental sulfur from the sulfided
material and demonstrated more than 100 regeneration
cycles with no significant loss of sorption capacity. Accord-
ing to the theoretical estimates of the authors, MnAl
2
O
4
can
also be used to remove HF and HCl.
126
Wakker et al. studied manganese and iron supported on
-Al
2
O
3
at 400-800 C (Table 3).
28
It is one of the few
laboratory-scale studies where the sorbents have gone
through several hundred sulfidation and regeneration cycles
(100 cycles for iron-, and more than 300 cycles for manga-
nese-based sorbents, respectively). The deactivation profile
indicates that most of the deactivation occurred in the first
10 cycles.
Yasyerli prepared a series of sorbents (equimolar ratios of
vanadium-manganese, zinc-manganese, iron-manganese,
and several different ratios of cerium-manganese) and
foundthat a cerium-manganese mixedoxide hadthe highest
H
2
S sorption rate constant. Regeneration of the sulfided
sorbent converted 90% of the sorbed sulfur to elemental
sulfur.
127
Alonso and Palacios studied Zn- and Cu-doped manga-
nese oxide as a high temperature, regenerable sorbent
(Table 3).
128,129
These sorbents were prepared by calcining
the extrudate formed from mixed powders of pure oxides
with a Mn/dopant ratio of 10:1. For zinc-doped materials,
XRD and transmission electron microscopy (TEM) studies
indicated that in the freshly prepared sorbent Zn was in-
corporated into the Mn
3
O
4
lattice, and the resultant mixed
oxide was the dominant phase (77%) of the prepared sor-
bent. After sulfidation, most of the sulfide species was MnS
(89%), with mixed sulfide (Mn,Zn)S comprising the remain-
ing 10% of the material. Copper doping, on the other hand,
mainly impacted the degree of dispersion of the fresh sorbent
with only a small amount of Mn-Cu mixed oxide form-
ing.
129
After sulfidation the dominant phases detected by
XRD were MnS and MnO. Any copper oxide or copper
sulfide phase, if present, was not detectable by XRD, poten-
tially due to poor crystallinity. After regeneration, only one
phase was detected in each of the doped materials, with zinc
or copper incorporated into the Mn
3
O
4
tetragonal struc-
ture.
129
In a long-term, 70-cycle test, with regeneration
conducted at 800 C in air, MnSO
4
was detected using FTIR
spectroscopy.
128
Alonso and Palacios hypothesized that the
formation and decomposition of some MnSO
4
during re-
generation could be advantageous as it may help maintain
sorbent pore structure.
128
Yoon et al. investigated desulfurization by and regenera-
tion of manganese ore (Table 3).
130,131
The manganese
ore used consisted primarily of -MnO
2
(pyrolusite). They
found sulfate formation could cause a decline in break-
through time in subsequent cycles. Adding 0.5% ammonia
gas (NH
3
) in the regeneration gas stream prevented sulfate
formation, and (NH
4
)
2
SO
4
was detected as one of the
products.
One recent study investigated the performance of MnO
2
supported on a monolith as a catalyst for decomposing NH
3
and found it to be very effective.
132
However, it is not clear
whether this material can be used for both NH
3
and H
2
S
removal because the experiment was not conducted in the
presence of H
2
S or steam.
2.6. Iron-Based Materials. Iron-based and iron-modified
materials (e.g., zinc ferrite that was discussed in Section
2.2.1) have received considerable attention. They are rela-
tively low cost and have considerable sulfidation efficiency
even when reduced.
Tamhankar et al. studied the kinetics of desulfurization
using an iron oxide sorbent (45 wt. % Fe
2
O
3
, 55 wt. %
SiO
2
).
133
In syngas (CO and H
2
), the iron oxide rapidly
reduced to metallic iron, which then reacted with H
2
S.
X-ray diffraction, Mossbauer spectroscopy, and TGA
weight change analysis revealed that the product is FeS
1.1
,
a form of pyrrhotite that has the general formula Fe
1-x
S.
The kinetic studies also showed that hydrogen concentration
had no effect on the sulfidation rate over a wide range of
concentrations, and that the sulfidation reaction was first
order with respect to H
2
S concentration. The activation
energy for sulfidation, 3.3 kcal/mol, was low, which the
authors attributed to the adsorption mechanism, with H
2
S
decomposition as the rate limiting step. For small and
(121) Ben-Slimane, R.; Hepworth, M. T. Energy Fuels 1995, 9 (2),
372378.
(122) Turkdogan, E. T. Ironmaking Steelmaking 1993, 20 (6), 469
475.
(123) Ben-Slimane, R.; Hepworth, M. T. Energy Fuels 1994, 8, 1184
1191.
(124) Atakul, H.; Wakker, J. P.; Gerritsen, A. W.; van den Berg, P. J.
Fuel 1996, 75 (3), 373378.
(125) Atakul, H.; Wakker, J. P.; Gerritsen, A. W.; Vandenberg, P. J.
Fuel 1995, 74 (2), 187191.
(126) Bakker, W. J. W.; Kapteijn, F.; Moulijn, J. A. Chem. Eng. J.
2003, 96 (1-3), 223235.
(127) Yasyerli, S. Chem. Eng. Process. 2008, 47 (4), 577584.
(128) Alonso, L.; Palacios, J. M. Energy Fuels 2002, 16 (6), 15501556.
(129) Alonso, L.; Palacios, J. M. Chem. Mater. 2002, 14 (1), 225231.
(130) Yoon, Y. I.; Chun, B. H.; Yun, Y.; Kim, S. H. J. Chem. Eng.
Jpn. 2004, 37 (7), 835841.
(131) Yoon, Y. I.; Kim, M. W.; Yoon, Y. S.; Kim, S. H. Chem. Eng.
Sci. 2003, 58 (10), 20792087.
(132) Ismagilov, Z. R.; Shkrabina, R. A.; Yashnik, S. A.; Shikina, N.
V.; Andrievskaya, I. P.; Khairulin, S. R.; Ushakov, V. A.; Moulijn, J. A.;
Babich, I. V. Catal. Today 2001, 69 (1-4), 351356.
(133) Tamhankar, S. S.; Hasatani, M.; Wen, C. Y. Chem. Eng. Sci.
1981, 36 (7), 11811191.
5302
Energy Fuels 2009, 23, 52915307
:
DOI:10.1021/ef900714q
large particles, the reaction rate was controlled by chemical
reaction and pore diffusion, respectively.
Tseng et al. studied the kinetics of regeneration of sulfided
iron oxide sorbent, using O
2
and SO
2
separately, in the
temperature range 600-900 C.
134
At equilibrium, oxygen
oxidizes the FeS
1.1
that was formed during sulfidation to
Fe
2
O
3
, with the simultaneous liberation of sulfur as SO
2
(reaction 5). The activation energy for this reaction was
determined to be 15.63 kcal/mol.
2FeS
1:1
3:7O
2
SFe
2
O
3
2:2SO
2
5
With SO
2
, the oxidation of FeS
1.1
proceeds to magnetite,
with the formation of elemental sulfur; this reaction can be
written as
3FeS
1:1
2SO
2
SFe
3
O
4
2:65S
2
6
The activation energy for this reaction was determined to be
17.5 kcal/mol.
In a separate publication, Tamhankar et al. studied the
oxidation of FeS with steam and steam-air mixture using
TGA and Mossbauer spectroscopy.
135
They determined that
oxygen and steamoxidized FeS in parallel to Fe
2
O
3
and Fe
3
O
4
,
respectively. Further oxidation of Fe
3
O
4
by O
2
also occurred in
the reactor, but for as yet unknown reasons it is likely to have
negligible impact since Fe
3
O
4
is detected even in samples that
have almost completely reacted. They also deduced that ele-
mental sulfur is formed from H
2
S decomposition.
White, et al. further studied the regeneration of FeS in
steam-O
2
mixtures with the H
2
O/O
2
ratio varied from 7.7 to
200 at 550-700 C.
136
Building on the reactions proposed by
Tamhankar et al.,
135
they interpreted their experimental
results on the basis of four simultaneous reactions: (1)
oxygen reacts rapidly with FeS to produce SO
2
and Fe
2
O
3
,
(2) steamreacts relatively slowly with FeS to formFe
3
O
4
and
H
2
S, (3) further oxidation of Fe
3
O
4
formed in the second
reaction by O
2
, and (4) SO
2
and H
2
S react to form elemental
sulfur and water. The route for elemental sulfur formation is
the main difference between Tamhankar et al. and White
et al. in their interpretation of reactions that occur between
iron sulfide and steam-air mixture. On the basis of their data,
White et al. concluded that to achieve high yields of ele-
mental sulfur, the experimental conditions need to be ad-
justed to have a large fraction of the FeS convert to H
2
S (the
relatively slow second reaction) and that both SO
2
and H
2
S
be formed at positions within the bed with sufficient time to
react. Among the experimental parameters investigated by
the authors, the maximum yield of elemental sulfur was
approximately 75% of the theoretical, which was achieved
using a H
2
O/O
2
ratio of 200 and a temperature of 600 C.
2.7. Rare Earth-Based Materials. Rare earth-based sor-
bents show very good promise. There are several advantages
of rare earth-based sorbents including their use at high
temperature (>700 C), their favorable sulfidation equilib-
rium in reducing gas environments, and their potential to be
regenerated via a route that directly produce elemental
sulfur, which reduces the complexity of the regeneration step
(Section 3.3).
137
The thermodynamic, oxidation-reduction, and electronic
properties of ceriumoxides, ceria-zirconia mixed oxides, and
lanthanum oxides have been widely investigated.
138-143
These studies yield detailed information on the reduction
of CeO
2
to CeO
2-x
, which further reduces to Ce
2
O
3
. This
thermodynamic information laid a foundation for using rare
earth oxides as sulfur sorbents. In addition, sulfidation
thermodynamics of ceria and reduced ceria, as illustrated
in (reaction 7 and 8) have also been well studied.
144,145
2CeO
2
s H
2
Sg H
2
g SCe
2
O
2
Ss 2H
2
Og 7
Ce
2
O
3
s H
2
Sg SCe
2
O
2
Ss H
2
Og 8
In contrast to copper, where the reduced form is not very
efficient in H
2
S removal, the reduced forms of ceria and
lanthanum are more active. This makes rare earth-based
materials suitable for desulfurization in highly reducing
syngas.
Wheelock et al. used rare earth metals, with particular
focus on lanthanum-containing sorbents, to remove H
2
S and
COS.
146
Kay et al. recognized that the thermodynamic of the
reactions between rare earth oxides and sulfur is particularly
favorable in a reducing environment. They developed meth-
ods of using rare earth oxides, in particular cerium oxide, to
desulfurize gases as well as molten iron and steel that have
high sulfur content.
144,147-150
Use of the sorbents and meth-
ods for regenerating the sorbents and for SO
x
and NO
x
removal were also developed.
Li and Flytzani-Stephanopoulos studied binary
Cu-Cr-O and Cu-Ce-O oxides (Table 3) for desulfuriza-
tion.
98
For the CuO-CeO
2
system, they concluded that
copper remains in a dispersed state in the sorbent and that
the reduced ceria explains the desulfurization efficiency.
Kobayashi and Flytzani-Stephanopoulous studied the sulfi-
dation kinetics of cerium oxide and Cu-modified cerium
oxide.
151
By obtaining experimental data of the initial re-
duction rate of CeO
2
and Cu-CeO
x
(where Cu content
ranged from 5 to 40 atom %) as a function of temperature,
the authors clearly demonstrate that it is easier to reduce the
(134) Tseng, S. C.; Tamhankar, S. S.; Wen, C. Y. Chem. Eng. Sci.
1981, 36 (8), 12871294.
(135) Tamhankar, S. S.; Garimella, S.; Wen, C. Y. Chem. Eng. Sci.
1985, 40 (6), 10191025.
(136) White, J. D.; Groves, F. R., Jr; Harrison, D. P. Catal. Today
1998, 40 (1), 4757.
(137) Zeng, Y.; Zhang, S.; Groves, F. R.; Harrison, D. P. Chem. Eng.
Sci. 1999, 54 (15-16), 30073017.
(138) Zhou, G.; Shah, P. R.; Montini, T.; Fornasiero, P.; Gorte, R. J.
Surf. Sci. 2007, 601 (12), 25122519.
(139) Zinkevich, M.; Djurovic, D.; Aldinger, F. Solid State Ionics
2006, 177, 9891001.
(140) Kim, T.; Vohs, J. M.; Gorte, R. J. Ind. Eng. Chem. Res. 2006, 45
(16), 55615565.
(141) Xiao, W.; Guo, Q.; Wang, E. G. Chem. Phys. Lett. 2003, 368
(5-6), 527531.
(142) Huang, S.; Li, L.; Van der Biest, O.; Vleugels, J. Solid State Sci.
2005, 7 (5), 539544.
(143) Andersson, D. A.; Simak, S. I.; Johansson, B.; Abrikosov, I. A.;
Skorodumova, N. V. Phys. Rev., B, Condens, Matter Mater. Phys. 2007,
75, (3), 35109-35101 to 35109-35106.
(144) Kay, D. A. R.; Wilson, W. G.; Jalan, V. J. Alloys Compd. 1993,
193 (1-2), 1116.
(145) Ferrizz, R. M.; Gorte, R. J.; Vohs, J. M. Appl. Catal., B2003, 43
(3), 273280.
(146) Wheelock, K. S.; Aldridge, C. L. Sulfide Removal Process. U.S.
Patent 3,974,256, Aug. 10, 1976.
(147) Kay, A. R.; Wilson, W. G. Methods of Desulfurizing Gases. U.S.
Patent 4,604,268, Aug. 5, 1986.
(148) Kay, D. A. R.; Wilson, W. G. Methods of Desulphurizing Iron
and Steel and Gases, Such As Stack Gases and the Like. U.S. Patent
4,084,960, Apr 18, 1978.
(149) Kay, D. A. R.; Wilson, W. G. J. Metals 1988, 40 (11), 57.
(150) Kay, D. A. R.; Wilson, W. G. Method for the Regeneration of
Sulfided Cerium Oxide Back to a Form That Is Again Capable of
Removing Sulfur from Fluid Materials. U.S. Patent 4,857,280, Aug.15,
1989.
(151) Kobayashi, M.; Flytzani-Stephanopoulos, M. Ind. Eng. Chem.
Res. 2002, 41 (13), 31153123.
5303
Energy Fuels 2009, 23, 52915307
:
DOI:10.1021/ef900714q
composite Cu-CeO
x
structure than the nonmodified ceria.
In addition, the authors also determined that the sulfidation
rate of all Cu-containing ceria samples is much higher than
that of pure ceria. This canbe explained by the higher affinity
of reduced Ce
2
O
3
for sulfur and the increased number of
reduced Ce
2
O
3
sites in copper doped ceria.
Wang and Flytzani-Stephanopoulos studied the activity
and stability of lanthanum- and copper-containing cerium
oxide sorbents (Table 3).
152
Sorbents were tested using
simulated hot reformate gas at 650 and 800 C and regener-
ated at the same temperature (details of gas composition are
in Table 3). Lanthanum doping (up to 50 atom %) was
effective against crystal growth and sintering, thus moderat-
ing the surface area loss of ceria in H
2
S-free reformate gas at
800 C.
152
On the basis of XRD, copper oxide exists in highly
dispersed form in ceria and is not effective in moderating
sintering. However, copper-doped ceria had the best sulfida-
tion kinetics among the ceria-based sorbents tested.
152
In
samples containing lanthanum, La
2
O
2
S was detected after
sulfidation. However, in samples containing ceria, Ce
2
O
2
S
was not detected via XRD, which the authors interpreted as
due to the oxidation of Ce
2
O
2
S in air to Ce
2
O
2.5
S.
152,153
In
their study, Wang and Flytzani-Stephanopoulos determined
that, by using very high space velocities, sulfidation could be
limited to the sorbent surface only.
Using presulfided CeO
2
, La
2
O
3
, and lanthanum-substi-
tuted ceria, Flytzani-Stephanopoulos et al. were able reduce
H
2
S concentrations to sub-ppm levels and to limit surface
area loss to just 4% after 15 regeneration cycles, indicating
that fast sulfidation-regeneration may reduce surface area
loss and prolong the overall sorbent lifetime.
42
The authors
suggested that by using two reactors that each rapidly
alternates its function between a sorber and a regenerator,
and conducting the sulfidation and regeneration both under
high space velocities, the time required for sulfidation and
regeneration would be similar and the reactor design could
be compact.
152
Zeng et al. investigated the effect of H
2
O and H
2
/H
2
O on
ceria sorbent efficiency in the temperature range 700-850 C
and interpreted the results based on equilibrium O
2
pressure
exerted by the product gas in the gas composition.
154
No
carbon monoxide or carbon dioxide was used in the experi-
mental gas. They also investigated the use of SO
2
to regen-
erate sulfided ceria as described in Section 3.3.
137
Other studies found that addition of ZrO
2
increases the
oxygen mobility within the CeO
2
crystal lattice
155
and the
resulting CeO
2
-ZrO
2
solid solution has 3-5 times the
reversibly stored oxygen than pure CeO
2
.
156
Assuming that
these properties are related to enhanced performance as a
sulfur sorbent, Yi et al. studied the desulfurization properties
of CeO
2
-ZrO
2
.
157
They found that ZrO
2
doping increased
the reducibility of ceria, increased the specific surface area of
the sorbent produced, and improved its sintering resistance.
Addition of small amounts of CO
2
in the feed gas signi-
ficantly reduced the sorbent capacity, resulting in much
shorter prebreakthrough time.
2.8. The Effect of Hydrocarbon, Tar, and Steam on Sulfur
Sorption. Most sorbent tests using simulated coal-derived
syngas were conducted with relatively dry syngas and in the
absence of tar and other contaminants. Although the effect
of some of these contaminants on sorbent performance or
stability can be modeled using thermodynamics (e.g., effect
of steam), other effects (e.g., soot formation) cannot be
modeled easily. Experimental verification of the effect of
these other syngas components would be tremendously
important for efficient pilot and demonstration plant testing.
2.8.1. Steam. Thermodynamics predicts that water vapor
has a negative effect on the equilibrium between H
2
S and
metal oxide sorbents. Novochinskii et al.
158
hypothesized
that in addition to shifting the thermodynamic equilibrium
shown in reaction 1, the competitive adsorption of H
2
O on
surfaces also plays a role in water vapors effect on desulfur-
ization.
Actual experimental results indicate varying levels of
water vapor impact. This may be because reported data were
obtained in different conditions with different factors con-
trolling the reaction rates. In general, the effect of steam to
sulfur sorbent performance is expected to be more severe at
higher temperatures, although there are a lot fewer studies of
the effect of steam at higher temperatures. According to
thermodynamics, zinc-based sorbents should perform effec-
tively even with high steam content below 500 C. Experi-
mental results generally confirmthat. For example, Gupta et
al. found that a zinc titanate sorbent efficiently removed H
2
S
from simulated syngas containing 50% steam (other gas
components comprising CO
2
, CO, H
2
, no hydrocarbons) at
493 C.
159
Sanchez et al. used Z-Sorb III, a sorbent contain-
ing less than 50% ZnO and 10% nickel oxide, and found no
detrimental effects on sorbent performance from H
2
, CO, or
10-15% H
2
O.
160
Kim at al. conducted a study of the effect of steam (0-45
vol %, in nitrogen) on H
2
S removal and found that the
presence of 45% steam reduced the H
2
S breakthrough time
by almost half (from approximately 460 to 280 min, at
363 C).
161
Before breakthrough though, the ZnO sorbent
was able to reduce an inlet H
2
S concentration of 2000 ppmv
to 1 ppmv even in 45% steam.
161
Jun et al. compared the performance of Zn-Ti-based
sorbents doped by cobalt and nickel oxide as a function of
water vapor content (5-20%by volume inCO, CO
2
, H
2
, and
N
2
). They found that at 480 C the reaction rate of zinc
titanate with H
2
S is decreased in water vapor whereas those
of the doped materials are not affected.
162
2.8.2. Gas Components Such As CO and CO
2
. In the highly
reducing environment of coal- and biomass-derived syngas,
H
2
S should be the dominant sulfur species. Nevertheless,
here are a number of studies that indicate COS can be a non-
negligible species. Yang et al. studied the effect of individual (152) Wang, Z.; Flytzani-Stephanopoulos, M. Energy Fuels 2005, 19
(5), 20892097.
(153) Sourisseau, C.; Cavagnat, R.; Mauricot, R.; Boucher, F.; Evain,
M. J. Raman Spectrosc. 1997, 28 (12), 965971.
(154) Zeng, Y.; Kaytakoglu, S.; Harrison, D. P. Chem. Eng. Sci. 2000,
55 (21), 48934900.
(155) Colon, G.; Pijolat, M.; Valdivieso, F.; Vidal, H.; Kaspar, J.;
Finocchio, E.; Daturi, M.; Binet, C.; Lavalley, J. C.; Baker, R. T.;
Bernal, S. J. Chem. Soc., Faraday Trans. 1998, 94 (24), 37173726.
(156) Hori, C. E.; Permana, H.; Ng, K. Y. S.; Brenner, A.; More, K.;
Rahmoeller, K. M.; Belton, D. Appl. Catal., B 1998, 16 (2), 105117.
(157) Yi, K. B.; Podlaha, E. J.; Harrison, D. P. Ind. Eng. Chem. Res.
2005, 44 (18), 70867091.
(158) Novochinskii, I. I.; Song, C.; Ma, X.; Liu, X.; Shore, L.;
Lampert, J.; Farrauto, R. J. Energy Fuels 2004, 18, 576583.
(159) Gupta, R. P.; Turk, B. S.; Portzer, J. W.; Cicero, D. C. Environ.
Prog. 2001, 20 (3), 187195.
(160) Sanchez, J. M.; Ruiz, E.; Otero, J. Ind. Eng. Chem. Res. 2005, 44
(2), 241249.
(161) Kim, K.; Jeon, S. K.; Vo, C.; Park, C. S.; Norbeck, J. M. Ind.
Eng. Chem. Res. 2007, 46, 58485854.
(162) Jun, H. K.; Jung, S. Y.; Lee, T. J.; Kim, J. C. Korean J. Chem.
Eng. 2004, 21 (2), 425429.
5304
Energy Fuels 2009, 23, 52915307
:
DOI:10.1021/ef900714q
components of CO, CO
2
, and water content on H
2
S break-
through and COS formation.
163
They reported that CO and
CO
2
do not significantly affect H
2
S removal by ZnO/SiO
2
,
but do react with H
2
S to formCOS (up to 1000 ppm, froman
input H
2
S of 4000 ppm), which was not removed by ZnO.
This result differs fromthat of Sasaoka et al., who found that
COS removal by ZnO was possible.
54,55
Xie at al. used iron oxide and cerium oxide supported on
fine coal ash for desulfurization in the temperature range of
400-650 C.
164
The inlet simulated coal gas (mixture of H
2
,
CO, CO
2
, and N
2
) contained 0.47% H
2
S. The sorbent was
efficient at removing the H
2
S, with the breakthrough time
being longer at higher temperatures (520 and 620 C). Both
H
2
S and COS were detected in the exit gas after break-
through had started, with a H
2
S to COS ratio of approxi-
mately 13, indicating that some H
2
S was converted to COSin
the simulated coal gas environment.
2.8.3. Hydrocarbon. Wakker et al. investigated the effects
of water, CO, and hydrocarbons separately on MnO/-
Al
2
O
3
and FeO/-Al
2
O
3
sorbents.
28
They found that lower
alkanes such as CH
4
and C
2
H
6
do not affect desulfurization
efficiency. At 400 C, alkenes also do not affect desulfuriza-
tion activity. However, at 600 C, when 10% propene was
used, their measurements and material balance calculations
indicated some sulfur did not exit the reactor in gaseous
form. In the propene experiments, Wakker et al. also ob-
served brown deposit in their reactor system at the higher
temperature, which led them to hypothesize that polymeri-
zation or vulcanization of H
2
S with propene had occurred.
28
Gupta et al. studieddesulfurization withsimulated natural
gas containing 58% methane (other components were N
2
and CO, no steam), the H
2
S content was reduced from 3 vol
% to less than 20 ppmv when sulfidation temperatures were
greater than 427 C.
159
However, the total sulfur content in
the simulated gas, consisting of COS, SO
2
, and CS
2
, re-
mained greater than 100 ppmv at 427 C.
159
These results
show that the efficiency of the sorbent for total sulfur
removal can be affected when H
2
S is transformed to other
sulfur species. They also illustrate the importance of sulfida-
tion conditions as sorbent performance and sulfur speciation
are bothfunctions of temperature andother gas constituents.
2.8.4. Other Impurities Such As Chlorine. There are reports
of both negative and positive effects of chlorine on the
performance of sulfur sorbents. Wakker et al. found that
0.5-1% HCl in the feed gas reduced H
2
S breakthrough
capacity of their manganese on -Al
2
O
3
sorbent by
30-40%.
28
The deactivation by HCl was found to be
reversible, and adsorbed HCl was assumed to be removed
from the sorbent during regeneration.
Jun et al. reported on the sorption capacity of cobalt- and
nickel-doped zinc titanate in the presence of 0.2%HCl.
162
At
650 C, the doped materials were not affected by HCl over
4-5 cycles of sulfidation and regeneration. At that tempera-
ture, the undoped zinc titanate initially had good sorption
capacity (0.2 g sulfur/g sorbent) but lost its capacity gradu-
ally with increasing cycles, most likely due to the formation
of ZnCl
2
.
162
At a lower temperature, 480 C, cobalt- and
nickel-doped zinc titanate were not affected by HCl, but
the sulfur sorption capacity of undoped zinc titanate was
significantly reduced, even for a fresh sorbent (0.05 g sulfur/g
sorbent).
A different effect was observed by Gupta and OBrien,
who studied the desulfurization of zinc titanate sorbents in
simulated syngas containing 0-1500 ppmv HCl at 538-
750 C.
165
The zinc titanate sorbent used had a ZnO/TiO
2
molar ratio of 1.5 with inorganic and organic binders.
165,166
Gupta and OBrien found no deleterious effects of HCl at
538 and 650 C, and in most cases the desulfurization rates
increased with HCl present. For the simulated lowBTUcoal
gas, the prebreakthrough time increased significantly with
increasing HCl concentrations. They hypothesized that this
could be due to a two step reaction:
ZnO2HCl ZnCl
2
H
2
O 9
ZnCl
2
H
2
S ZnS2HCl 10
According to the hypothesis of Gupta and OBrien, the
increased efficiency is attributed to the presence of liquid
ZnCl
2
at 500-750 C and that the liquid form may be more
accessible for reactions with H
2
S than ZnO solid.
165
Opera-
tion at 650 C or above is undesirable when chlorine content
reached 800 ppmv due to vaporization of ZnCl
2
. In addition
to the regenerable adsorbed chlorine on zinc titanate, a
portion of the HCl cannot be regenerated. It is hypothesized
that non-regenerable NaCl was formed through the reaction
of HCl with Na
2
O in the binder. Even though zinc titanate
itself is rated to perform up to 700 C, the optimum tem-
perature of operation for zinc titanate that is prepared with
alkali based binder is below 550 C because of the alkali
vaporization issue.
165
In summary, even though there are some reported results,
a detailed understanding of the performance of sulfur sor-
bents in process gases with high steam concentrations,
hydrocarbons, and halogens, as is found in biomass-derived
syngas, is still lacking. As illustrated in the fewinvestigations
in the presence of HCl, the effects of these other contami-
nants on sulfur sorbents is likely to be dependent on tem-
perature, sorbent, contaminant concentrations, and binder/
support. Whether these impacts can be entirely predicted by
thermodynamics remain unclear, and further research in this
area is necessary.
3. Regeneration
During regeneration, metal sulfide in the sorbent material is
regenerated back to metal oxide. Depending on the process
conditions (total pressure, temperature, and oxygen partial
Figure 2. Sulfur sorbent regeneration process using steam. Hydro-
gen sulfide gas is produced which may need to be scrubbed prior to
its release to the atmosphere.
(163) Yang, H. Y.; Sothen, R.; Cahela, D. R.; Tatarchuk, B. J. Ind.
Eng. Chem. Res. 2008, 47 (24), 1006410070.
(164) Xie, W.; Chang, L.; Wang, D.; Xie, K.; Wall, T.; Yu, J. Fuel
2009, In Press.
(165) Gupta, R. P.; OBrien, W. S. Ind. Eng. Chem. Res. 2000, 39 (3),
610619.
(166) Gupta, R. P.; Gangwal, S. K.; Jain, S. C. Fluidizable Zinc
Titanate Materials with High Chemical Reactivity and Attrition Resis-
tance. U.S. Patent 5,254,516, October 19, 1993.
5305
Energy Fuels 2009, 23, 52915307
:
DOI:10.1021/ef900714q
pressure P
O
2
) and the inherent thermodynamics of the sorbent
material, SO
2
, H
2
S, elemental sulfur (S
2
), or a combination of
these gases is also produced. Another important aspect to
consider is the regenerationscheme, whichdefines the number
of reactors or vessels needed, and whether any byproducts of
economic value are generated.
3.1. Regeneration with Production of H
2
S. One potential
regeneration scheme is the reverse of reaction 1, that is,
M
x
S
y
s yH
2
Og SM
x
O
y
s yH
2
Sg 11
Froma process engineering perspective, assuming station-
ary packed beds are used, a dual bed reactor system is
required for this regeneration process. One of the reactors
contains active sorbent used during the sulfidation step,
while the other contains spent material being regenerated
(Figure 2). Regenerated process gas would contain H
2
S, a
precursor to SO
2
and a regulated emission. In addition, H
2
S
is itself an odor nuisance, with an odor threshold of 0.5-
2 ppb by volume.
167
Depending on process conditions, the
outlet gas may exceed the emission limit for H
2
S and/or SO
2
,
requiring further scrubbing before it is discharged to the
atmosphere. Another potential approach that could be used
in a biomass-derived syngas application, if a sulfided mo-
lybdenum catalyst was chosen for alcohol synthesis, would
be to sendsome of the H
2
S to the mixed alcohol fuel synthesis
unit as the sulfided molybdenum catalyst requires up to
100 ppm H
2
S to maintain catalyst activity.
6
However, this
might entail sulfur removal from the synthesized fuel.
Even though it is normally assumed that the role of H
2
Ois
to regenerate the metal oxide with release of H
2
S, several
studies have shown that H
2
O can also cause SO
2
release.
Sasaoka et al. studied the role of H
2
O in the oxidation of
sulfided ZnO, Fe
2
O
3
and CuO sorbents using TPR and
isotopic H
2
18
O.
168,169
They concluded that H
2
O directly
contributes to the oxidation of ZnS, FeS, and Cu
2
S, resulting
in the release of SO
2
.
3.2. Regeneration with Production of SO
2
. This regenera-
tion scheme works for several different oxides. A zinc oxide-
zinc sulfide system will be used as an example in the discus-
sion of this regeneration scheme. The regeneration of the
sulfided material, ZnS, in an atmosphere containing O
2
proceeds as
2ZnSs 3O
2
g S2ZnOs 2SO
2
g H
600C
-779kJ=mol 12
Because this reaction is highly exothermic, a low O
2
content of the regeneration gas must be maintained to
control the regeneration temperature and to avoid sintering
and the corresponding loss of sorbent effectiveness.
137
The
other reason for controlling O
2
content is to avoid formation
of sulfate, an undesirable side reaction. Again, a zinc based
system is used to illustrate the reaction:
ZnSs 2O
2
g SZnSO
4
s 13
2ZnOs 2SO
2
g O
2
g S2ZnSO
4
s 14
The formation of sulfate leads to potential physical da-
mage to the sorbents as there is a large molar volume
difference between ZnO or ZnS (15 and 24 cm
3
3
mol
-1
respectively) and ZnSO
4
(46 cm
3
3
mol
-1
).
170
The presence
of some oxygen during regeneration promotes formation of
SO
2
, but an oxygen concentration that is too high forms
undesirable ZnSO
4
, with the suitable level of oxygen being
pressure dependent. To prevent formation of ZnSO
4
, a suf-
ficiently high temperature and adjustments to pressure and
oxygen concentration in the process gas are necessary. (See
Elseviers and Verelst
45
for modeling results, and Gupta
et al.
159
for regeneration conditions used in their experiments).
The disadvantage of a regenerationscheme that produces SO
2
(g) is that SO
2
is aregulatedpollutant. The presence of SO
2
inthe
atmosphere can (1) cause health problems, (2) reduce visibility,
and (3) form acid rain.
171,172
As in the case of a regeneration
scheme that produces H
2
S, it would be necessary to determine
whether the outlet gas meets the requiredemissionstandards and
whether further processing of the outlet gas is necessary.
The conventional Claus process
171,173
can be used to
recover elemental sulfur from the SO
2
gas that is produced.
Another method to recover pure sulfur is the direct sulfur
recovery process (DSRP) patented by a research group at
RTI International.
174,175
This process uses a slipstream of
Figure 3. Simplified schematic diagram of regeneration with pro-
duction of SO
2
followed by DSRP to produce elemental sulfur.
Figure 4. Regeneration with direct production of elemental sulfur
(adapted from Figure 10 in Zeng et al.
137
).
(167) Gabriel, D.; Deshusses, M. A. Proc. Nat. Acad. Sci. U.S.A.
2003, 100 (11), 63086312.
(168) Sasaoka, E.; Hatori, M.; Yoshimura, H.; Su, C.; Uddin, M. A.
Ind. Eng. Chem. Res. 2001, 40 (11), 25122517.
(169) Sasaoka, E.; Hatori, M.; Sada, N.; Uddin, A. Ind. Eng. Chem.
Res. 2000, 39 (10), 38443848.
(170) Armstrong, T. R.; Carneim, R. D.; Berry, D. A. In A Review of
Current State-of-the-Art Materials for Hot Gas Desulfurization, Nine-
teenth Annual International Pittsburgh Coal Conference, Pittsburgh, PA,
September 24-26, 2002; Pittsburgh Coal Conference, 2002; pp 1238-1254.
(171) Stirling, D. The Sulfur Problem: Cleaning up Industrial Feed-
stocks; Royal Society of Chemistry: Cambridge, UK, 2000.
(172) Environmental Protection Agency. Sulfur Dioxide;Health and
Environmental Impacts of SO
2
. http://www.epa.gov/air/urbanair/so2/hlth1.
html (accessed August 27, 2009).
(173) Kohl, A. L.; Nielsen, R. B. Gas Purification, 5th ed.; Gulf
Publishing Company: Houston, TX, 1997.
(174) Gangwal, S. K.; McMichael, W. J.; Dorchak, T. P. Environ.
Prog. 1991, 10 (3), 186191.
(175) Dorchak, T. P.; Gangwal, S. K.; Harkins, S. M. Method for
producing elemental sulfur from sulfur-containing gases. US 5,798,088,
Aug 25, 1998.
5306
Energy Fuels 2009, 23, 52915307
:
DOI:10.1021/ef900714q
coal-derived syngas containing CO and H
2
to produce ele-
mental sulfur from the SO
2
released during regeneration
(Figure 3). A skid-mounted, 6.0 in. diameter reactor was
designed and fabricated for conducting pilot tests of this
process.
176
According to RTI International, DSRP is more
economical than the Claus process.
3.3. Regeneration with Direct Production of Elemental
Sulfur. Regeneration using O
2
or SO
2
with the direct produc-
tion of elemental sulfur was reported in a number of studies
of sulfided iron sorbents (Figure 4).
134,136,177
More recently,
this approach was studied for manganese- and ceria-based
sorbents. The economic advantage of directly producing
elemental sulfur is that pure sulfur can be condensed and
sold. The challenging issue for this regeneration scheme is to
eliminate elemental sulfur condensation in other parts of the
reaction system.
Zeng at al.
137
conducted research with ceria and modified
ceria and their ceria regeneration scheme consisted of:
2Ce
2
O
2
Ss 2O
2
g S4CeO
2
s S
2
g 15
Ce
2
O
2
Ss SO
2
g S2CeO
2
s S
2
g 16
Bakker et al. regenerated MnS with an SO
2
/O
2
mixture
and with SO
2
alone to produce elemental sulfur (Table 3).
126
Regeneration of this sorbent with SO
2
needs to be conducted
at over 600 C to avoid sulfate formation. At 850 C, the
kinetics of regeneration using SO
2
is 2 times faster than
regeneration with steam. During regeneration, a mixture of
MnO and MnAl
2
O
4
was formed, and greater than 90% of
the adsorbed sulfur was recovered as elemental sulfur.
4MnSs 2SO
2
g S4MnOs 3S
2
g 17
4MnSs 4Al
2
O
3
s 2SO
2
g S4MnAl
2
O
4
s 3S
2
g
18
3.4. Factors that Affect Regenerability. The factors that
affect whether a material can be regenerated over many
cycles are not well understood. Several research groups
believe that reactivity is directly related to available sorbent
surface area or pore volume and that sintering during high
temperature regeneration causes loss of surface area and
subsequent loss of reactivity. However, Jun et al. noted that
their regenerated zinc titanate had a higher surface area than
fresh zinc titanate though the material steadily lost perfor-
mance over ten cycles.
93
Ryu et al. conducted a 100-cycle test
of a ZnO-TiO
2
based sorbent modified with cobalt oxide,
nickel oxide, molybdenumoxide, and iron hydroxide.
178
The
total pore area of the sorbent increased, while the median
pore diameter decreased after 100 cycles. These results
demonstrate that higher surface area does not always corre-
spond to improved performance. The authors of both studies
hypothesized that repeated expansion and contraction dur-
ing sulfidation and regeneration caused spalling or cracking
of the sorbent.
93,178
The volume expansion and contraction
problem, specifically when sulfate species formed, was also
discussed by Armstrong et al. as a mechanism for failure of
ZnO-based sorbent after regeneration.
170
5. Conclusions
This review has identified several different sorbent systems
that can potentially be used to effectively remove H
2
S from
coal- or biomass-derived syngas at mid- to high-temperature
ranges. For zinc, copper, and iron, reduction to metallic
elements will occur at high temperature in a reducing envir-
onment. Much of the research in the past two decades has
focused on stabilization of these materials through formation
of mixed oxides. The optimum sorbent depends on process
conditions and target sulfur levels. Rare earth-based sorbents
show significant promise for high temperature applica-
tions (700-850 C). Manganese- and zinc titanate-based
sorbents also have potential for applications in the 700 C
range if care is taken to limit sulfate formation. Both manga-
nese- and ceria-based materials may be regenerated with the
direct production of elemental sulfur. Copper-based and zinc
ferrite materials are well suited for approximately 500 C.
Unmodified zinc oxide-based materials are generally more
applicable to the low temperature ranges. Kinetic studies in
general indicate product layer diffusion as the most useful
model todescribe sulfidation kinetics. Some of these materials
additionally may have the ability to remove NH
3
and/or HCl,
contaminants also found in biomass-derived syngas.
For high-temperature desulfurization of biomass-derived
syngas, further research is necessary to determine the perfor-
mance of sorbents in a high steam environment containing
light hydrocarbons and tar, which is quite different from
sulfur removal in coal-derived syngas. An understanding of
the extent of removal of other sulfur species such as COS is
also essential.
There is increasingly more research on binary oxides and
promoted binary oxides to determine whether these mod-
ified materials provide better attrition resistance, higher sulfi-
dation equilibrium constants, and the ability to remove
multiple gas contaminants. Molecular understanding and
modeling, which has been very effective in aiding catalyst
design, could be a powerful tool for the design of next
generation sorbents or combined sorbent/catalyst systems.
Spectroscopic and imaging characterization tools could pro-
vide some of the mechanistic information that is needed for
rational design and iterative testing and modification of such
materials.
Key remaining questions include whether multiple con-
taminants in syngas can be removed simultaneously and
whether one contaminant might interfere with the removal
of another. Aneeded and ancillary requirement is an accurate
determination and measurement of the concentration and
speciation of biomass-derived syngas contaminants (chlorine,
sulfur, and ammonia) as a function of feedstock (corn stover,
switch grass, hard and soft wood, etc.) in varied gasification
environments, ideally at the operating temperature of the
reactor, sothat insituspeciationinformation canbe obtained.
With information on syngas contaminant concentrations and
speciation, the appropriate sorbents and multifunctional
materials can be better designed and tested in the labora-
tory. At a minimum, the effect of other gas components
(e.g., hydrocarbons) and impurities (e.g., chlorine, ammonia)
(176) Portzer, J. W.; Turk, B. S.; Gangwal, S. K. In Durability Testing
of the Direct Sulfur Recovery Process, Advanced Coal-Fired Power
Systems Review Meeting, Morgantown, WV, July, 1996.
(177) Patrick, V.; Gavalas, G. R.; Sharma, P. K. Ind. Eng. Chem. Res.
1993, 32 (3), 519532.
(178) Ryu, S. O.; Park, N. K.; Chang, C. H.; Kim, J. C.; Lee, T. J. Ind.
Eng. Chem. Res. 2004, 43 (6), 14661471.
(179) Twigg, M. V.; Spencer, M. S. Appl. Catal., A 2001, 212 (1-2),
161174.
5307
Energy Fuels 2009, 23, 52915307
:
DOI:10.1021/ef900714q
on the performance of sulfur sorbents needs to be modeled
and experimentally verified.
Pilot scale testing of such sorbents will provide results and
experience for eventual commercialization. As more scientific
data on mid- to high-temperature sulfur sorbent becomes
available, technoeconomic analysis and systems modeling to
determine the potential improvements in thermal efficiency,
environmental benefits, andprocess economics throughuse of
such sorbent systems will be critical.
Acknowledgment. Funding for this research was provided by
the Office of the Biomass Program, U.S. Department of Energy,
under contract DE-AC36-99GO10337 with the National Renew-
able Energy Laboratory.

Das könnte Ihnen auch gefallen