Вы находитесь на странице: 1из 40

FUEL

PROCESSING
TECHNOLOGY
Fuel Processing Technology 41 (1995) 159-198

ELSEVIER

Production of hydrogen and sulfur from hydrogen sulfide


J. Z a m a n , A. C h a k m a *
Department of Chemical and Petroleum Engineering, University of Calgary,
Calgary, Alberta, Canada T2N IN4

Received 24 August 1993; accepted in revised form 1 June 1994

Abstract

Hydrogen sulfide is a fairly abundant resource whose potential is not being fully utilized by
the industry. It is in fact a mineral from which two valuable products, hydrogen and sulfur, can
be extracted. No such extraction is done today. Sulfur, however, is recovered by partial
oxidation by Claus process and low-quality steam is produced utilizing the heat of reaction.
Efforts to produce both hydrogen and sulfur from hydrogen sulfide has been made in recent
years through a diverse variety of technologies. These involve thermal, thermochemical,
electrochemical, photochemical and plasmochemical methods. An attempt has been made in
this paper to bring to focus the possibilities and limitations in each of these areas and find out
the areas of prospective research and development. A method that has the potential to develop
into a full-scale technology has been suggested.

1. Introduction

Hydrogen is currently needed in large quantities as a chemical feedstock in the


synthesis of a m m o n i a and methanol, in the desulfurization and hydrocracking in the
refineries and in the upgrading of various hydrocarbon resources such as heavy oil
and coal. Alberta is a large producer of hydrogen as well as its major consumer in the
upgrading processes [1]. Most of Alberta's hydrogen comes from the steam reforming
of natural gas, which is also the predominant method of production of hydrogen
globally [2, 3]. There are other established methods of hydrogen production such as
partial oxidation of residual oil, gasification of coal, steam-iron process and water
electrolysis [4, 5]. Extensive research is being carried out in m a n y other exotic
processes for hydrogen production such as high temperature electrolysis of steam,
thermal cracking of natural gas, thermochemical water splitting, solar photovoltaic
water electrolysis and plasma decomposition of water [6-8]. Another source of
hydrogen hitherto untapped, but being increasingly focused is the extraction of the
* Corresponding author.
0378-3820/95/$09.50 1995 Elsevier Science B.V. All rights reserved
SSDI 0 3 7 8 - 3 8 2 0 ( 9 4 ) 0 0 0 8 5 - 8

160

J. Zaman, A. Chakma/Fuel Processing Technology 41 H995) 159-198

hydrogen in hydrogen sulfide [9]. The interest in the production of hydrogen continues unabated because of the additional reason that hydrogen is perceived as the
energy of the future [10-13].
Hydrogen sulfide occurs naturally in many gas wells. It is produced in large
quantities in the desulfurization of petroleum stocks where the organic sulfur is
trapped as hydrogen sulfide. It is separated by solvent absorption methods and most
commonly sent to a Claus plant where the sulfur is recovered and a low grade steam is
obtained. Classically, hydrogen sulfide has been considered a liability which only
occasionally turned out to be an asset depending on the international sulfur price.
Only in recent times, suggestions have been made to regard hydrogen sulfide as a high
valued mineral [14]. Hydrogen sulfide has a high heating value, but its use as a fuel is
ruled out because its product of combustion is sulfur dioxide which is not environmentally acceptable. An immediately alternative route for the utilization of hydrogen
sulfide is to devise ways to break it down to its constituent elements, hydrogen and
sulfur [15, 16]. When compared to the existing method of utilization of hydrogen
sulfide for the production of sulfur in a Claus plant, the possibility of obtaining two
valuable products provides significant stimulus to the development of such a process.
On the whole, the interest in the utilization of hydrogen sulfide as a source of
hydrogen and sulfur intensified in recent years because of the following reasons:
(a) Global prospect for hydrogen energy and waste minimization. (b) The unavoidable
production of hydrogen sulfide from gas plants, refineries, upgraders and metallurgical processes. (c) In geothermal utilization, H2S is often contained in steam.
(d) Decomposition of H2S can become a part of a water splitting cycle. (e) The tail gas
clean-up from Claus plants can exceed the value of the sulfur recovered if the
environmental regulations are made more stringent.
The most direct process of obtaining hydrogen and sulfur from hydrogen sulfide is
thermal decomposition, catalytically or noncatalytically [15-49]. However, the reaction is highly endothermic and the equilibrium conversion even at high temperatures
is low (10%, 20% and 30% at 1143, 1283 and 1403K, respectively). This has
prompted research in the decomposition with equilibrium shift by several methods
such as preferential removal of products of reaction by membranes and thermal
diffusion columns. To overcome the high endothermicity and equilibrium limitation,
several two step processes such as the sulfiding of a metal or a lower sulfide to liberate
the hydrogen and then calcining the higher sulfide to decompose into sulfur and the
original metal or sulfide have also been proposed. Thermochemical cycles have been
proposed by many researchers for the splitting of a water molecule, and some of the
cycles involve a hydrogen sulfide decomposition step. This has prompted others to
devise independent thermochemical cycles for obtaining hydrogen and sulfur from
hydrogen sulfide [4]. In recent years, methods for water splitting have been applied to
H2S decomposition and we now have a whole array of methods for the decomposition
of hydrogen sulfide which closely parallels the methods of water splitting for hydrogen
generation. These methods include electrolysis, photolysis, plasmolysis and their
many variants.
As of today, there is no method of hydrogen sulfide decomposition which can be
considered commercially feasible [ 17]. Though the thermal methods with equilibrium

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

161

shift may appear to be the most promising, it would be premature to rule out the other
avenues. Particularly, the advances made in water splitting could open up an attractive
route for the splitting of the hydrogen sulfide molecule as well. The present review, therefore, deals with all the prospective methods of obtaining hydrogen and, in most cases,
hydrogen and sulfur, from hydrogen sulfide. The methods are discussed under the headings of thermal, thermochemical, electrochemical, photochemical and plasma methods.
Methods other than the thermal decomposition methods are multidisciplinary in
character and during the course of the survey of the literature, the authors were struck
by both the quality and quantity of work of researchers from various disciplines
engaged in the research on hydrogen production from HES. The present review
attempts to bring together the contribution from these various sources. Sufficient
experimental details have been provided in many cases because of their novelty and/or
the need for further research in those areas. In certain cases, work done in other
reaction systems has been included. Particularly in the areas of photodecomposition
and electrolysis, a lot can be gained from an analysis of the developments in the water
splitting methods. Similarly, in plasmolytic decomposition, examples from related
fields such as transformations of methane have been cited.

2. Thermal decomposition
2.1. Thermodynamic considerations
Hydrogen sulfide breaks down into hydrogen and sulfur on the application of heat.
At lower temperatures, molecules of vapor containing one to eight atoms of sulfur and
many sulfane species (H2S2 to H2S8) may be present in the reacting mixture. Equilibrium calculations have been made for the complex equilibrium mixture by free energy
minimization techniques and the results have been verified experimentally [18-20].
It has been observed that in the temperature range 973-1123K and pressure
101-405 kPa, the diatomic sulfur molecules constitute more than 99.8% of the total
sulfur molecules. Experimental conversions obtained at lower pressures also agreed
with these results. Overall, since the decomposition reaction requires temperatures in
excess of 1000K for significant conversions, the sulfur produced will be necessarily
diatomic. The thermal decomposition reaction can therefore be considered to follow
the stoichiometric equation below [19]:
2H2S (g) ~ 2H2 (g) + S 2 (g).

(1)

2.2. Nature of the studies by thermal methods


A study of the literature indicates that research on thermal decomposition proceeded in the following directions: (a) Develop a catalyst for the reaction. (b) Devise
a scheme for enhanced conversion by equilibrium shift. (c) Develop solar furnace for
high temperature and sub-atmospheric pressure operation so that high conversion
can be obtained in a once-through process.

. Zaman, A. Chalona/Fuel Processing Technology 41 (1995) 159-198

162

The presentation that follows discusses most of the published literature in this area
and attempts to point out the merits of the different approaches as well as their
limitations.
2.3. Studies in tubular reactors

Raymont [15, 16, 20, 21] carried out a systematic investigation of the decomposition reaction in a flow reactor at atmospheric pressure in the temperature range
750-1350K. Three catalysts were investigated: silica beads (532 CP), cobalt molybdate (HR-801) and 1% pre-sulfided platinum (182 CP), all from Air Products and
Chemicals. The results were compared with those from an empty tube reactor. Below
1250 K, the influence of catalyst was significant in enhancing the rate, the cobalt-molybdate catalyst being the most active of the three. Experimental results at
lower temperatures (< 1123 K) were explained on the basis of equilibrium model
accounting for various sulfur and sulfane species. The empty tube conversions at
higher temperatures were correlated with a rate equation of the form:
rn2 = kl CH2S-- k2 C22.

(2)

A value of 217 kJ/mol for activation energy of the forward noncatalytic reaction
was obtained. The conversion results from catalytic experiments could not be fitted
to a simple rate equation. The low temperature catalytic runs gave conversions
higher than equilibrium conversion based on reaction (1). It was shown by thermodynamic calculations that the anomaly was caused by the formation of polyatomic sulfur molecules and the sulfanes at lower temperatures. Beyond 1000K,
experimental results agreed with the equilibrium conversions predicted based on
reaction (1).
Recently, noncatalytic decomposition was studied by Kaloidas and Papayannakos
[22] in a nonisothermal flow reactor at pressures of 131-303 kPa, temperatures
873-1133K and specific flow rates of 3.4-36.0 mol/m2s. The reactor was a fused
~,-alumina tube of internal diameter 6 mm and length 403 mm. The temperature
distribution inside the reactor along its length was measured by a traveling thermocouple. The flow rates used satisfied the plug flow requirements. Filling the reactor
with crushed alumina pieces (same material as reactor) having 15-fold larger surface
area than the reactor wall did not affect the conversion, indicating that the reactor
material did not act as catalyst. The variables studied were temperature, pressure and
space velocity. An empirical rate equation describing the overall reaction as the
difference between the rates of the forward reaction and the reverse reaction was fitted
to the conversion data. Activation energies of 196 and 105 kJ/mol were obtained for
the forward and reverse reactions, respectively. The empirical rate equation was
shown to conform to the following reaction mechanism:
kl

H2S <~ HS" + H',

(3)

H" + H2S ~- H2 + HS',

(4)

k2

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159 198

163

2HS" e. H 2 S + S ' ,

(5)

S ' + S " ~ S2.

(6)

The presence of the free radicals H', S" and HS" in gaseous systems containing H2S has
been confirmed by various research workers [18, 20, 29]. Assuming reaction (3) to be
rate controlling, the above reactions give rise to the following rate equation:
1

rH2S

kl Pu2s

nl/2

KAKBKcKD vn~vs~ ,

(7)

KA, KB, Kc, and Ko are the equilibrium constants for the reactions (3)-(6). Eq. (7) is of
the same form as the empirical rate equation based on reaction (1):
1 ~

~1/2

rH2s = kl PH2s - ~ t'n2es2

(8)

The noncatalytic decomposition carried out by other workers found the homogeneous decomposition rate to be second order [23, 24]. The decomposition was also
studied catalytically using molybdenum disulfide catalyst 1-25]. Hougen-Watson type
models for the reaction suggested that the rate determining step was the cleavage of
the hydrogen-sulfur bonds of the hydrogen sulfide adsorbed on the catalyst sites. The
activation energy of this step was calculated as 217 kJ/mol. The catalytic activity of
MoS2 decreased with time, stabilizing at 65% of the initial value after about 15-25 h
of operation.
Bandermann and Harder [26] carried out the decomposition reaction in a quartz
reactor. They took advantage of the higher equilibrium conversions at lower pressures. The reactor with alumina as catalyst was operated at 1090-1230K and
13-51 kPa. Equilibrium conversions were obtained at all conditions in less than
20ms. The amount of catalyst was 2.25g and the flow rate range was
0.03-0.10Nm3/h. To hinder recombination, the reaction products were quickly
cooled down to 700 K after leaving the reactor. The sulfur was removed by condensation in a heat exchanger followed by an electro-filter. A pressure swing adsorption
system was designed to separate the gas containing 25-30% hydrogen and 75-70%
hydrogen sulfide. A commercial plant was proposed combining the low pressure
decomposition, sulfur condensation and pressure swing adsorption to separate the
hydrogen product and recycle the un-reacted H2S. Chivers et al. [27] studied the
catalytic activities of Cr2S3, WS2 and MOSE and concluded that MoS2 was a superior
catalyst at temperatures above 873 K. Table 1 gives a summary of work done on the
thermal decomposition of H2S using reactors of different types.
2.4. Studies with equilibrium shift
2.4.1. Product removal by condensation
Fukuda et al. [28] carried out the pioneering work to demonstrate by actual
experiments the level of equilibrium shift that can be obtained in the thermal
decomposition of hydrogen sulfide by the removal of one or both the products. The

164

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198


t--a
q

t.._l

"0

..~

0
e~

8
0

eq t~

"E
Y=

azr~

~'8

"r.

..

~
e~

~2.~- ~
~'~
-

=.~

,~_~

"0

0
0

6
"0

"S

[..,

[-

[-

[-

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

.~

.~.8

~.

.=_

.~
.=_

~=
0

~,.

0
0

,~o

--

~=~ ~
~ ~._~
[-

=.~
"4~

=.

,,

"4

.E

[..

~.o

I ~

.~~. ~

o~

165

166

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

apparatus employed in this study was a closed circulating system made from Pyrex
glass. The dead volume of the system was 943 cm 3. Hydrogen sulfide was reserved in
a sample reservoir in advance from a liquefied hydrogen sulfide cylinder (99.99%
purity) and a known amount of hydrogen sulfide was introduced into the reaction
system when necessary. The reaction gas was circulated through the reaction vessel by
a magnetic pump. Sulfur was removed in a cold trap, causing equilibrium shift and
high conversion of H2S. It was found that molybdenum disulfide catalyst decomposes
hydrogen sulfide effectively into hydrogen and elemental sulfur over the temperature
range 773-1073 K. It was possible to convert more than 95% of the hydrogen sulfide
by continuous removal of sulfur and intermittent separation of the hydrogen when the
reaction was carried out at 1073 K and sub-atmospheric pressure (5-20 kPa).
Both molybdenum sulfide and tungsten sulfide were effective catalysts. Both the
sulfides had a preferred orientation for the (0 0 1) plane under X-ray diffraction. The
alumina supported molybdenum disulfide was prepared by the impregnation of
~-alumina (spherical 3-4 mm diameter, surface area 250 m2/g) in an aqueous solution
of ammonium thiomolybdate. The impregnation step was followed by reduction at
643 K under a hydrogen stream and calcination at 1023 or 1413 K under a nitrogen
stream for 6 h prior to use. The catalyst calcined at 1413 K was found to perform
better than the one calcined at 1023 K and the unsupported MoS2. The supported
catalyst had 8 wt% MoS2 as determined by an X-ray fluorescence method. The
surface areas were 202 and 47 m2/g for the catalysts calcined at 1023 and 1413 K,
respectively compared to 4.1 m2/g for the unsupported catalyst. The unsupported
catalyst performed better than the supported catalyst calcined at 1023 K.
Sugioka and Aomura [29-31] studied the catalytic decomposition of H2S over
MoS2 catalyst in a closed circulation system similar to that of Fukuda et al. [28]. The
dead volume of the apparatus including the reactor was 550 cm 3. Circulation of H2S
was determined by measuring the pressure of hydrogen in the system by Pirani gauge
at known times after freezing out H2S in the cell at 77 K. It was found that only
hydrogen was contained in the Pirani gauge cell by removal of sulfur in an ice trap.
Hydrogen sulfide (99.99%) was purified by passage through molecular sieve 5A and
collected in a trap cooled with liquid nitrogen. It was subjected to three such
trap-to-trap distillation before it was stored for use. Hydrogen (99.99%) was purified
by passing through molecular sieve 5A cooled with liquid nitrogen. Molybdenum
sulfide used as catalyst had a purity better than 99.2%. The impurities were silica,
below 0.05%; ferric oxide, below 0.2%; cupric oxide (below 0.05%); water soluble
substance, below 0.5%. The sample of MoS2 used was found to have a hexagonal
structure and preferred orientation parallel to the (0 0 1) plane by XRD analysis. The
decomposition of H2S with MoS2 was carried out at 773 K with 5 g of catalyst and
6 kPa of initial partial pressure of hydrogen sulfide. At this condition, direct thermal
decomposition in the absence of catalyst was shown to be negligible. It was found that
the catalytic activity of MoS2 was enhanced if the MoS2 was reduced by hydrogen at
higher temperatures. The activity of the catalyst did not increase considerably when
the catalyst was evacuated at high temperature instead of being reduced. This is an
indication that sulfur gets adsorbed during the reaction and it leaves the surface more
readily if reduced by hydrogen than if the surface is evacuated. A mechanism of the

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159 198

167

decomposition reaction was proposed where the desorption of adsorbed sulfur from
MoS2 surface was the rate determining step. The authors also studied the promoter
effects of the catalysts and found cobalt sulfide in sulfided Co-Mo/AI20 3 catalyst
acted as promoter whereas nickel sulfide in sulfided Ni-Mo/AI203 catalyst acted as
a desulfurization reagent. The acid properties of the supports are apparently important for the decomposition reaction [31].
2.4.2. Product removal by membranes
Kameyama and co-workers [33-38] carried out the decomposition of hydrogen
sulfide in glass and alumina membrane reactors. The membranes allowed preferential
permeation of hydrogen and caused equilibrium shift resulting in conversion double
the equilibrium values. The Vycor glass tubing was 600 mm in length, 15 mm in
outside diameter and 3 mm in thickness. It had a porosity of 0.79, mean pore diameter
4.5 nm and surface area 191 m2/g. The membrane tube was loaded with M o S 2 catalyst
and was surrounded by a stainless steel shell protected by a nonporous alumina tube
placed close to it. The ends of the porous glass tubing were sealed at both ends by
carbon packing. The feed gas was introduced into the membrane side at high pressure
(405 kPa) with the permeate side kept at 101 kPa. The glass membrane was tested for
durability and no degradation was observed in hydrogen and hydrogen sulfide
atmosphere between 873-1073 K for 216 h of operation. However, remarkable shrinkage occurred beyond 1073 K, setting that temperature the highest for this membrane.
The catalytic activity of the porous glass particles on the decomposition was tested
and this showed no more conversion than in an empty tube and hence it was
concluded that the Vycor glass had no catalytic activity for the reaction. The
separation factor for hydrogen was found to increase at higher temperatures, the
contribution of surface diffusion being negligible beyond 573 K. The separation factor,
however, was lower than the ideal value and this may be due to viscous flow at high
pressure (4 atm.). Permeation of sulfur was found to be low and it was concentrated
mainly in the un-diffused gases. The authors also worked on porous alumina tube
reactor with a pore diameter of 102 nm with comparable results. Alumina membrane
allowed temperatures higher than 1073 K and resulted in 30 times higher permeability
compared to Vycor glass having smaller pores. However, the separation factor was
poorer. A film coating composed of alumina sol was added to the alumina tubing to
decrease the pore diameter and hence improve separation factor, but without much
success [37].
Edlund et al. [-38] used a composite metal membrane to carry out the decomposition of hydrogen sulfide at 973 K and H2S partial pressure of 79 kPa. Vanadium was
used as the base metal which, unlike palladium, does not suffer from hydrogen
embrittlement and has good permeability for hydrogen and is less expensive. Platinum was used on the feed side to protect the vanadium from attack by hydrogen
sulfide and also to act as catalyst for the decomposition reaction. On the permeate
side, the vanadium was coated with palladium to protect it from oxidation during
start up and shut down. An inter-metallic barrier of a thin layer of S i O 2 o n each side of
the membrane was found to provide good stability for constant hydrogen flux. The
detailed procedure of preparing the composite membrane has been discussed by the

168

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

authors. The platinum coating on the feed side offered the greatest resistance to
hydrogen gas permeation. The authors found that palladium when used on the feed
side quickly gets corroded by hydrogen sulfide while platinum retained its shine after
eight hours of operation. Nearly complete conversion of H2S was achieved in the
membrane reactor, with a feed concentration of 1.5-100%.
Zaman and Chakma [39] investigated the thermal decomposition of hydrogen
sulfide in dense and porous membranes using data from the literature. The results
showed that at 1073 K high conversions in dense membranes can only be achieved at
very high sweep gas flow rate on the permeate side. The authors concluded that in
addition to improvements of membrane properties, an effective catalyst would still be
desirable in order to have enhanced conversion at temperatures below 1073 K.
Quite a number of patents have been awarded for hydrogen sulfide decomposition
in membrane reactors [33, 4042]. An early Japanese patent [40] was for a ceramic
membrane (SiO2, A1203 or Si3N4) with pore diameter 3-10 nm with the sulfide of Mo,
W or Ru as catalyst operating at temperatures greater than 873 K and pressure
202-1013 kPa. A hollow fiber construction of ceramic or glass membrane was
described in another patent [41]. A European patent [42] described membranes of
various materials such as A1203, SIO2-A1203, ZrO2, zeolite, porous glass or C supported on A1203, SIO2-A1203, mullite, cordierite, ZrO2 or C. The catalyst for
decomposition was MoS2 or precious metal such as Pt or Pd. Many of the membranes
referred to are not available commercially. However, they do point to the interest of
the industries in the processes.

2.4.3. Product removal by thermal diffusion


Thermal diffusion columns provide another means for simultaneous reaction and
separation causing equilibrium shift [-43-45]. The sulfur formed condenses on the cold
surface of the column and hydrogen moves preferentially to the top of the column.
The removal of the products shifts the equilibrium to higher conversions. It was found
that the use of two columns in series in which the lower column acts as the reactor and
the upper column as the separator gives high conversions and excellent separations.
Nishizawa et al. [43] used a thermal diffusion column having length 1000 mm, 12 mm
external diameter silica pipe for hot wall and surrounded by a Pyrex glass tube of
internal diameter 24 mm and a Pyrex glass jacket of internal diameter 32 mm. The hot
wall temperatures were taken as the mean of temperatures measured at five different
positions by means of thermocouples. The cold wall was supplied with tap water with
inlet and outlet temperatures in the ranges 288-293 and 293-298 K, respectively.
Three columns having differences in the spacing between the hot and cold surfaces,
total volume and the mode of heating were tested in batch mode. It was found that
a mixture of 35-40% hydrogen in hydrogen sulfide separated to 88-92% hydrogen at
the top of the column in 30 min. It was also shown that there was no decomposition of
H2S at this temperature in the absence of catalyst. The hot surface was coated with
catalysts and the performance of the thermal diffusion columns for the decomposition
of H2S was investigated at 773 K. The condensation of sulfur on the cold surface and
the preferential transport of hydrogen up the column caused the decomposition
reaction to continue beyond equilibrium value. The degree of decomposition after 4 h

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

169

was found to be 40% at 773 K with hydrogen composition at the top of the column
equal to 96%. It was found that the differences in spacing (6 and 10 ram) between lthe
hot and cold tubes were inconsequential when there was no reaction. But with
reaction and separation, the column with larger spacing gave substantially lower
conversion. Higher temperature of the hot wall also resulted in better separation of
hydrogen at the top of the column. Four catalysts were tested: Chromium sulfide,
cobalt sulfide, nickel sulfide and iron sulfide. At 773 K, the authors did not observe
any significant difference in the activity of any of these sulfides.
Chivers and Lau [44, 45] designed a thermal diffusion column with a central quartz
rod (12 mm diameter) equipped with an electric heating tape that could produce
temperatures up to 1073 K. The cold surface was provided by circulating an appropriate coolant through the jacket which was concentric with the quartz rod. The catalysts
used were FeTSs, Fe~S8/MoS2 and FevSs/NiSI.2 on alumina support. In all cases, the
catalysts were prepared as a paste by grinding the metal sulfides with alumina (1:1
ratio by weight) in aqueous solution. The paste was coated on to the quartz rod which
was then wrapped with quartz wool to prevent the dried catalysts from falling off the
rod. In batch operations, the concentration of H2 at the top of the reactor increased to
about 80% at 1073K for all three catalysts. It was seen that these catalysts gave
separation factor for hydrogen, 2-3 times higher than that was achieved in an
MoSz-alumina system. The separation factor was studied as a function of (a) temperature of the hot surface, (b) temperature of the cold surface, (c) composition of the gas
mixture, and (d) multiple columns in series. The separation factors for hydrogen in
H z - H 2 S mixture varied from 215 to 460 in the temperature range 673-973 K. The
distance between the hot and cold surfaces was varied between 3.7 and 12.0 mm and
an optimum distance of 7.7 mm was found for all temperatures. The use of two
columns instead of one was found to provide an improvement of separation by
a factor of 20-30. If two columns are connected in series, H2 concentration greater
than 90% was obtained in h. It has been suggested that a thermal diffusion column
reactor followed by two thermal diffusion column separators in parallel would be an
effective configuration for H2S decomposition and product separation. The increase
of cold surface temperature adversely affects the conversion as well as separation, as
has been mentioned before. However, significant reaction and separation still do take
place with the cold surface at about 373 K. The separation factor for the flow system
was found to be lower than that in a closed system; the concentration of H2 at the top
of the reactor was only half the value obtained without flow. The separation factors
were higher at low temperatures for a reacting system while it is the reverse in
a nonreacting system. The separation factors for V2S3 and V2S3/CH9S 5 catalysts were
found to be higher than MoS2 catalysts. No attempt was made by the authors to
relate these findings with any predictive theory.
2.5. Thermal methods using solar energy
The decomposition of hydrogen sulfide was investigated by several workers using
concentrated solar radiation [46--49]. Bishara et al. [46] used a quartz reactor 20 cm
long with an inside diameter of 1.2 cm and thickness 0.4 cm. Heat was provided by

170

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

a solar furnace consisting of a heliostat (3 3 m), which tracked the sun automatically,
and a parabolic dish concentrator. The characteristics of the concentrator were as
follows: 1.5 m diameter, 0.66 focal distance, 61 mm diameter of focal image and 1.5 kW
power. The reactor position was adjusted near or away from the focal point to obtain
the reaction temperature. The variables studied were temperature and residence time
using three different catalysts: alumina (pore volume 0.75-0.95 cm3/g, surface area
220m2/g, particle diameter 1.6mm), cobalt-molybdenum oxide (CoO 4.5 wt%,
MoO 3 16.2 wt%, A120 3 79.3 wt%, surface area 264.5 m2/g, pore volume 0.52 cm 3,
pore radium 3.93 nm) and nickel-molybdenum oxide (NiO 3.33 wt%, MoO3
20.25 wt%, AI20 3 76.42 wt%, surface area 160.34 m2/g, pore volume 0.46 cm a, pore
radius 5.7 nm). The oxides of Ni, Co and Mo were sulfided by passing a stream of H2S
at high temperatures. Steady state conversions were obtained in all cases at a short
residence time (0.25 s) in the temperature range 893-1073 K. Conversions higher than
equilibrium values were obtained. This happened because the upper part of the
reactor was exposed to the atmosphere which caused removal of sulfur by quenching
and consequent equilibrium shift. A maximum conversion of 15% was achieved at
1043 K. The data were analyzed successfully to fit a second-order rate equation for all
the three catalysts. The rate equations and parameters from various sources are
compared in Table 2. The authors [45] also investigated three other catalysts in this
reaction system: nickel-tungsten (Harshaw Ni-4301), nickel-molybdenum (Harshaw
HT-500E) and cobalt-molybdenum (Harshaw HT-400). Comparable results were
obtained for these three and the previous three catalysts, with HT-500E catalysts
giving a maximum conversion of 14.5% at 973 K at a residence time of 0.3 s. The
activation energy values for these six catalysts were in the range of 60-75 kJ/mol.
A conceptual design of a plant was proposed with separation of sulfur by condensation and of hydrogen by pressure swing adsorption. Type Y (sodium ion) molecular

Table 2
Rate equation and activation energy for the thermal decomposition of hydrogen sulfide
Rate equation

Activation energy
(forward reaction)
(kJ/mol)

Catalyst

Temperature
range (K)

Reference

r = klpH2s - - k 2 P-H 2 P s,1/2


2
Initial rate
Initial rate
Initial rate
r = kI-H2S] 2
r = k1,H2S] 2
r = k[H2S] 2
r = k1,H2S] 2
r = k1,H2S] 2
r = k1,H2S] 2
r = k[H2S] 2
r = kI[H2S] - k21,H212[S] 1/2

196
175.8
112.2
112.2
74.9
69.1
59.9
75.7
90.0
209
49.8
217

None
None
MoS2
WS2
(Ni-W) sulfide
(Ni Mo) sulfide
(Co-Mo) sulfide
Alumina
Alumina
None
Alumina
None

873-1133
773 1073
773-1073
773-1073
933 1073
933-1073
933 1073
933-1073
933 1073
-773-873
1061-1201

[22]
[28]
1,28]
1,28]
[46, 47]
[46, 47]
[46, 47]
1,46, 47]
[26]
1,24]
[63]
1,23]

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

171

sieve was suggested as the adsorbent. Kappauf et al. [49] claimed that ceramic
materials can be used in solar furnace and suggested that the reaction be carried out at
temperature 1873 K and pressure 10-50 kPa, achieving high conversions in a single
pass.
2.6. Assessment of the thermal methods

An examination of the research work conducted on thermal decomposition


shows that only a limited amount of work has been carried out on the kinetics,
mechanism and catalyst development [20-26, 46-47]. The kinetic studies have
been mostly empirical and as shown in Table 2, there is a large amount of disagreement in the parameter values, with the order of reaction cited as 1 or 2 and
the activation energies for the catalytic reaction varying from 50 to 112 kJ/mol.
Among the catalysts, only MoS2 has been investigated to any detail. Even in
this case, information on improving the catalyst by additional components is
scarce [31].
As regards to the processes, none of the thermal methods stand out as being the
most promising. In fact, none of the processes has been investigated to great
depths. The severe reaction environment (thermal as well as chemical) possibly
acted as an inhibiting factor. Processes based on membrane reactors are not
technologically available today, in spite of a number of patents in this area
[33, 40-42]. Developments in ceramic membranes and supported ceramic membranes
may offer possibilities in the future 1-50-54]. High temperature solar furnace in a
large scale has not been a reality yet; the problem of material of construction is
not going to be trivial. The thermal diffusion column reactors have not developed
as practical devices and the system has been found to be energetically inefficient
[26]. The product removal scheme of Fukuda et al. [28] is unworkable on a large
scale because of the large heating/cooling load. The scheme suggested by Bandermann
and Harder [26] is based on today's technology and has been claimed to be
commercially feasible.

3. Thermochemical methods

Thermochemical approach attempts to devise a number of easier reactions to arrive


at a difficult overall reaction. A large number of thermochemical cycles has been
proposed to carry out the highly energetic water splitting reaction [55-58]. Hydrogen
sulfide decomposition is not energetically as intense as water splitting, but the low
equilibrium yield at fairly high temperature of 1000 K has prompted many researchers
to investigate a workable thermochemical cycle for the decomposition reaction. The
incorporation of this reaction in some of the water splitting cycles also put added
impetus to the study of the reaction [56, 58]. The cycles are discussed in appropriate
headings. A comprehensive description of the various cycles are provided by Clark
and Wassink [17].

172

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

3.1. Two-step closed cycles with metals or metal sulfides


The two-step decomposition of H2S in presence of metal or lower metal sulfides
may be represented by the following equations [59-66]:
(M or MxSr) + zH2S ~(MSz or MxSy+z) + zH2,

(9)

(MSz or MxSr+z ) ~ (M or MxSr) + zS',

(10)

Overall: zH2S ~ zH2 + zS'.

(11)

In this process, the metal or metal sulfide is sulfided with H2S while hydrogen is
liberated. The metal sulfide subsequently decomposes into sulfur and the original
metal or metal sulfide. Kiuchi et al. 1-59-62] carried out the cyclic experiments by
sulfurizing a lower metal sulfide powder by passing H2S for 1 h. The powder was then
heated at reduced pressure to the thermal decomposition temperature of the higher
sulfide. Argon was used as the carrier gas to remove the sulfur gas. The decomposition
was carried out for I h and the cycle repeated. Silver metal and the following sulfides
were subjected to this cycle: FeS, Co9S8, NiaS2, NiaS2-MoS2, natural chalcopyrite
(CuFeS2_ x) and synthetic pentlandite (Ni4.5 Fe4.5 $8). Table 3 summarizes the results.
The iron sulfide was not an effective agent, as the actual amount of H2 evolved was no
more than 30% of theoretical amount. The long term use of nickel sulfide was not
promising, because the generation of hydrogen decreases possibly because of sintering
or partial fusion of the sulfide. Nickel sulfide mixed with an equal amount of MoS2
was found to give excellent results every time. The cycle with cobalt sulfide is
characterized by a small difference in temperature between sulfurization and decomposition. Cycles with chalcopyrite and synthetic pentlandite were unsuccessful,
like iron sulfide cycle. A cycle with metallic silver was found to be promising with
sulfurization at 873 K and decomposition at higher temperatures and reduced pressure.
Chivers and Lau [44, 45] conducted HzS decomposition experiments using MoS2
catalyst as well as by a two-step process using the following sulfides: Fe7S8, FevS8/
MoS2 and FeTSs/NiSI./ (all supported on alumina), V2S3, V2S3/FeS, W2S3/fu9S 5
and V2S3/ZnS. Experiments were carried out in thermal diffusion reactors operated
continuously or batch-wise as well as in a flow reactor continuously or in a closed
loop. The two-step processes were found to be nearly as effective as MoS2 catalyst for
the conversion of HES to hydrogen and sulfur. A1-Shamma and Naman [63, 64]
studied the two-step decomposition with different mixtures of vanadium oxide and
sulfide in alumina in a Pyrex tube reactor. Kinetic studies were made based on the
following reaction sequence:
V2S 2 "~ 5H2S ~ 2VS 4 + 5H2,

(12)

2VS4 ~ VzSa + 5S,

(13)

overall: 5H2S ~ 5H2 + 5S.

(14)

The Arrhenius activation energies were between 31-53 kJ/mol and the orders of the
reactions were between zero and one, depending on catalyst and temperature. Chen

J. Zaman, /1. Chakma/Fuel Processing Technology 41 (1995) 159-198

~D

~= .~ .r.

~'~

~-=~=

,~ o ~

~
~ - .

:.=

.~ ~.~
~
~.~.~

:.=

~
~

L~

e~

cO
0
e~

=~
e~

--t

=0
oo

oo

:L

.0'2_
Z
e~
a)

r~
,,2
.2

Z
0
e~

0
~3

173

174

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

et al. [65] recently presented results of the kinetics of desulfurization using metallic
copper as the desulfurizer. The reaction of hydrogen sulfide with copper was found to
become diffusion limited as the desulfurization progresses across the metal. Yumura
[66] observed that sulfur from hydrogen sulfide is adsorbed by Mn nodules and
hydrogen is liberated at 873-1073 K. The adsorption reaction was half order while the
decomposition reaction liberating hydrogen was found to be first order. Table 4 gives
a s u m m a r y of the results for the systems where the sulfiding and desulfiding have been
carried out at the same temperature.
3.2. Two-step closed cycle with carbon monoxide
This cycle is represented by the following two reactions [55, 67]:
H2S + C O ~ COS + H2,

(15)

COS ~ C O + S.

(16)

Reaction (15) takes place at 373-473 K and 101 k P a with either cobalt sulfide or
nickel sulfide as catalyst. However, reaction (16) is accompanied by the following side
reaction:
(17)

2COS ~ CO2 + CS2.


The inability to control reaction (17) makes the process unattractive.
3.3. Two-step closed cycle with iodine
This cycle can be described by the following two reactions 1-56, 68, 69-1:
H2S + I2 ~ 2HI + S,

(18)

2HI ~ H2 + I2.

(19)

Table 4
Thermochemical two-step process for hydrogen generation and sulfur recovery
Sulfide

Reaction
temperature (K)

Main findings

References

V2S3

803 1023

V2S3 + FeS (1 : 1)

673-1073

V2S3 + CugSs (1 : 1)

673-1073

V2S3 + ZnS (1 : 1)

673-1023

Hydrogen yield comparable to that obtained [45]


with Mo,S2
Hydrogen evolution superior to MoS2 below 1-45]
923 K, but worse at higher temperatures
Higher H2 yield than MoS2 at temperature
[45]
above 873 K. The mixed sulfides form Cu3VS4.
This acts as a better catalyst than MoS2
Better than MoS2 below 923, but similar to
[45]
MoS2 at higher temperature

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

175

The first reaction takes place at moderate conditions, but has difficulties regarding
isolation of pure sulfur. The second is a well studied reaction as a component of
water-splitting cycles and hydrogen can be recovered in widely different ways such as
electrolysis, thermal or photochemical methods.

3.4. Two-step open cycle with metals


In this process, a metal is sulfided liberating hydrogen and the sulfide is then
oxidized to regenerate the metal and produce SO2 [59 62, 70, 71].
M + zH2S-~ MS~ + zH2,

(20)

MSz + zO2 ~ M + zS02,

(21)

overall: zHzS + zO2 ~ zH2 + zSOz.

(22)

The SOz formed can be recovered as sulfuric acid or as elemental sulfur by the Claus
process. The metals have been used in powder form or as liquid, in order to avoid the
reaction hindrance by the sulfide film on the surface of the solid metal. The metals
used for the purpose have been lead, silver, copper and nickel. In the liquid metal
process, H2S was bubbled into molten lead or soft blown at the surface of the molten
lead. The addition of nickel for the bubbling and copper for soft blowing was found to
have catalytic effect. The regeneration of the metal took place by oxidation on low
oxygen pressure. At temperatures above 1023 K, only metallic Pb is formed, while
PbO and PbSO4 were also formed at lower temperatures. Similar results are obtained
with Cu, Ag and Ni. But in these latter cases, the reaction of sulfide and sulfate are
known to give rise to the metal in metallurgical processes. Hence, the recovery of
metals in these cases can be considered complete. A brief summary of the work done
using various metals is presented in Table 5.

3.5. Assessment of the thermochemical methods


The closed cycle thermochemical processes present interesting possibilities to
obtain hydrogen and sulfur at less severe conditions. One of the drawbacks of the
two-step sulfide processes (both open and closed cycles)is the handling of large mass
of sulfides and metals and the need to cool and re-heat them because of the differences
in operating temperatures of the two steps. The cycle with carbon monoxide appears
promising, but a highly selective catalyst is essential for the commercialization of the
process. The iodine cycle may become promising if photochemical method of decomposing HI becomes successful. The work carried out in single reactors with V2S3 and
mixed sulfides appear promising, but the results obtained are not adequate to make an
assessment of the prospects. A number of patents have been awarded on cyclic
processes of various kinds [72-79]. One patent considers a variation of the iodine
cycle where the HI formed is reacted with C6H 6 in the presence of Ru hydroxide at
423 K to give cyclohexane [75]. Catalytic dehydrogenation of cyclohexane at 573 K
with Pt supported by alumina gave H2. Behie et al. [76] suggested the use of mixed
metals in a fluidized bed at 623-823 K, followed by occasional regeneration of the

176

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

.~..~ ~

.~.

~,

._

t~

~
.~

O~

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159 198

177

metals at 823-1223 K. Two processes were described where organic metal complexes
were used in solvents. In one [78], the organic complex takes up the sulfur liberating
hydrogen, while the complex is regenerated with a mild oxidant, giving sulfur as the
other product. In the other process [79], the organic complex takes up the hydrogen,
releasing the sulfur. The complex is regenerated by dehydrogenation at 473 573 K
and hydrogen is obtained as a product. Both these processes require milder operating
conditions, but sufficient information is not available to evaluate them. None of the
thermochemical processes described in the open literature or in the patents have been
known to be commercialized or even piloted.

4. Electrochemical methods

Production of hydrogen by electrolysis of alkaline water is a well-known technology [4, 5]. The developments in water electrolysis such as advanced alkaline
water electrolysis (AWE), inorganic membrane alkaline water electrolysis (IME),
solid polymer electrolysis (SPE), high temperature electrolysis (HTE) and intermediate temperature electrolysis (ITE) [80-86] can be adapted to the electrolysis
of hydrogen sulfide. The passivation of the electrodes by sulfur, however, poses
a problem not encountered in water electrolysis. Three distinct routes for the
electrolysis of hydrogen sulfide emerged over the years and they are discussed
under the headings direct electrolysis, indirect electrolysis and high temperature
electrolysis.
4.1. Direct electrolysis
The electrolysis of hydrogen sulfide presents difficulties because of its low solubility
and low electrical conductance as a liquid. The electrolysis is carried out in acid or
basic solution, preferably in basic solution with H2S. Passivation of electrodes by
sulfur, difficulties in handling and removal of sulfur and the occurrence of undesirable
secondary electrochemical reactions make the system challenging, in spite of the
considerably lower energy need for this decomposition compared to the decomposition of water. Various remedies for the mitigation or elimination of the deposition of
sulfur on the anode has been attempted, such as the use of organic vapor and solvent,
porous electrodes and higher temperature and higher concentration of alkaline
solution [87-93].
Shih and Lee [91] fed a mixture of toluene and basic sulfide solution in a continuous stirred tank electrochemical reactor (CSTER). It was found that sulfur yield
was 40-80%, depending on the volume flow rate (10-66 ml/h), temperature
(293-333 K), applied voltage (1.8-4.8 V), sulfide concentration and electrolyte concentration. Dandapani et al. [89] achieved the electrolytic conversion without passivation using high concentrations of HS- and O H - and high temperature. In most
experiments, carbon electrodes were used while Ni and Pt cathodes were used in some
experiments.

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

178

In general, when hydrogen sulfide is dissolved in alkaline solution, sulfide ion is


formed:
H2S + O H - ~ H 2 0 + HS-.

(23)

Sulfur is formed at the anode and hydrogen is liberated at the cathode according to
the following electrochemical reactions:
Anode:

HS- + O H - ~ S + H20 + 2e-.

(24)

Cathode:

2H20 + 2e- ~ H2 + 2OH-.

(25)

Sulfur dissolves in the alkaline media forming polysulfides (only disulfide is shown in
the equations):
S + H S - + O H - -o S~- + H20.

(26)

Polysulfides may also form at the anode:


2HS- + 2OH- ~ Sz2- + 2H20 + 2e-.

(27)

Polysulfides are highly soluble in water, but when the solubility is exceeded, sulfur
may be precipitated:
Sz2- + H20 --, S + HS- + O H - .

(28)

However, with the accumulation of polysulfides, a cathodic reaction competing with


hydrogen evolution may take place:
Sz2- + 2e- + H 2 0 ~ HS- + 82- + O H - .

(29)

This decreases the formation of Hz with time. To overcome this problem, the cathode
was separated from the anode with a cation conducting membrane, preventing the
polysulfides from contacting the cathode [89]. This improved the hydrogen production efficiency, but the membrane lost its effectiveness within several days.
Anani et al. [92, 93] optimized the operating conditions of a double compartment
electrochemical cell employing Nation membranes as separator, producing crystalline
elemental sulfur and high purity hydrogen at high current efficiencies. Anode materials employed included graphite, nickel and porous nickel-chromium alloy as well as
titanium. Cathode materials were either nickel or graphite. In the process, the HzS
was absorbed in a NaOH scrubber and an equimolar concentrations of NaOH and
NariS were used at 80C. The equimolar concentration fixes the pH of the solution at
an optimum value, so that electrode passivation was delayed by sulfur dissolution.
This pH value also prevents the oxidation of polysulfide species to sulfur and
oxyanions.

4.2. lndirect electrolysis


Kalina and Maas [94, 95] describes processes for the indirect conversion of
hydrogen sulfide to hydrogen and sulfur. An acidic [94] electrochemical process was
based on the electrochemical oxidation of iodide in aqueous hydriodic acid at high

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

179

current densities and current efficiencies. Hydrogen gas was produced concurrently
with soluble triiodide to yield an impure sulfur product recovered in an unusually
sticky plastic form. Re-crystallization of this product from toluene yielded sulfur of
comparable purity to that of commercially produced sulfur.
Reactions in the electrolysis cell:
Anode:

31- ~ I3 + 2e-.

(30)

Cathode:

2H + + 2e-

(31)

Overall:

3I

-+ H 2.

+ 2H + --+ I3 + H2.

(32)

Reaction in chemical reactor/absorber:


I2 (as I3) + HeS ~ 2H + + 3I- + S.

(33)

In a typical experiment, 2.68 mol of HzS were passed into the reaction vessel over
a 5 h period. Afterward, 92.47 g of acid sulfur were recovered from the system. After
treatment with 1.21 hot toluene, this acid sulfur yielded 85.35 g of sulfur for a 99.3%
sulfur recovery based on the HzS added.
An indirect electrolytic basic process was also developed based on iodine cycle [95].
In a strongly basic solution (pH 13-14), the following overall reaction takes place in
the electrochemical cell:
I- + 3H20--+ 103 + 3H2.

(34)

The reaction of hydrogen sulfide with the iodate takes place in a separate reactor:
3HaS + IO3 ~ 3S + 3H20 + I - .

(35)

The basic process has advantages in material of construction, the quality of sulfur and
negligible iodine loss. However, it requires higher voltage and gives lower recovery of
elemental sulfur than the acid process.
Indirect processes based on iron complexes and iron chloride have also been
developed [96-101]. Olson [96, 97] describes a process involving a chemical step in
which an iron chelate reacts with H2S to produce sulfur, followed by an electrochemical step to recover the chelate and generate hydrogen. The chemical step is complicated by possible reaction of the elemental sulfur with the iron chelate. A hybrid
process based on iron chloride has been described by Mizuta and co-workers
[98-101]. The process consists of absorption of H/S in FeC13 aqueous solution and
subsequent electrolysis of aqueous FeCI/ solution. The following reactions are
involved:
H2S (g) + 2FeC13 (aq) ~ 2FeCI2 (aq) + 2HC1 (aq) + S (c),

(36)

2FeCI2 (aq) + 2HCI (aq) - [electrolysis] ~ H2 (g) + 2FeC13 (aq),

(37)

Overall:

(38)

HzS (g) ~ H2 (g) + S (c).

H2S absorption with almost 100% yield at 343 K and a satisfactorily low electrolysis
voltage (0.7 V) at a current density of 100 mA/cm 2 at 70C were obtained. The whole
process was carried out in a large excess of HC1. A bench scale unit consisting of HzS

180

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

absorption and electrolysis with a production capacity of 2.1 Nm 3 of H 2 per day and
3 kg sulfur per day was constructed and intermittent operation of the plant for 1000 h
was performed satisfactorily. The feed used was 30% H2S and 70% Ar. In an
industrial environment, the process gas can be directly fed to the system, thus
eliminating the need for a waste gas treatment plant such as an amine plant for the
separation of H2S.

4.3. High temperature electrolysis


A high temperature electrochemical process for the removal of H2S from coal gas or
natural gas has been developed [102-104]. The process has the advantage of converting H2S directly to hydrogen and sulfur; hydrogen goes with the process stream while
sulfur is condensed and collected as a product. Hence, the need for waste gas
treatment plant to separate the H2S is eliminated. The hot process gas flows past the
cathode of carbon, cobalt sulfide or nickel sulfide and the following electrochemical
reaction takes place:
H2S + 2e- ~ S2- + H2.

(39)

The hydrogen is evolved at the cathode and the sulfide ion moves along the electrical
potential gradient through a molten salt electrolyte suspended in an inert ceramic
matrix, to the opposite electrode with the following electrochemical reaction at the
anode:
S 2- --15 2 -~- 2 e - .

(40)

The process can also remove CO2 from natural gas and up to 80.7% removal of H2S
from a simulated process gas containing 2000 ppm HE S could be achieved. The sulfur
liberated at the anode is swept away for condensation with an inert gas stream.

4.4. Assessment of the electrochemical methods


The electrochemical approach provides a number of promising ways to generate
hydrogen and sulfur from hydrogen sulfide. Mao and co-workers [93] proposed
a process flow diagram for continuous electrolysis based on their results from a batch
process in the laboratory. They claim to have resolved the problem of sulfur deposition on the anode by optimizing the electrolyte concentration. The high temperature
electrolysis method developed by Winnick and co-workers [102, 103] may prove to be
successful where hot separation of H2S is preferable. The Fe-C1 process of Mizuta and
co-workers [98-100] combining chemical and electrochemical steps and producing
crystalline sulfur needs further evaluation. Fletcher and co-workers [ 105, 106] suggest
a process in which H2 and S are produced from H2S in an electrolysis cell, the electric
power for which is provided by an H2-O2 fuel cell which uses part of the H2 from the
electrolysis cell and atmospheric 02 as its reactants. The other avenue for the success
of the electrochemical methods may lie in the integration of the electrochemical
technology with the photovoltaic technology utilizing solar energy. In the mean time,
parallel to research on improving the performance of an individual cell, the practical

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

181

aspects of scale-up and design of cell modules should also be emphasized. In evaluating the performance of a single cell, both kinetic and transport rates should be
investigated in order to obtain meaningful design information.

5. Photochemical methods
Production of hydrogen by photolytic decomposition of water aroused intense
interest in recent years E8, 107-111]. Photodecomposition is perceived as a way to
harness solar energy, with hydrogen acting as the energy carrier. However, simple
molecules like water cannot absorb solar radiation without the aid of a photosensitizer or photocatalyst. The role of a photocatalyst is the same as the role
chlorophyll plays in the photosynthetic system in living cells. The catalysts can be
semiconductor electrodes, semiconductor particles, colored redox species such as dyes
and metal complexes, or colored redox species adsorbed on semiconductor electrodes.
When semiconductor electrodes are used, the process is commonly referred to as
photoelectrochemical process. It is called photochemical or photocatalytic process if
powder semiconductors are used. The methods and techniques developed for the
photolytic decomposition of water has been successfully applied for the decomposition of hydrogen sulfide to hydrogen and sulfur. The availability of large amounts of
sulfide and sulfite in chemical processing industries has also added to the impetus for
developing photolytic techniques for hydrogen sulfide decomposition [112, 113].

5.1. General principles


It was found that when RuO2-1oaded CdS particles were dispersed in aqueous
solutions of hydrogen sulfide and illuminated with visible light, hydrogen was generated at an astonishingly high rate [114, 115]. The CdS particles having band gap
2.4 eV (517 nm) absorbs visible light of wave length less than the band gap and this
generates electrons in the conduction band and holes in the valence band. The
electrons migrate to the surface where reduction of water to hydrogen takes place
while the holes react with sulfide to form sulfur. The different oxides such as TiO2,
RuO2 etc. are added to act as catalyst for the redox reaction and prevent or minimize
electron-hole recombination. TiO2, for example, can accept electrons from the conduction band of CdS while R u O 2 catalyzes the hole transfer from the valence band of
CdS to H2S or sulfide in solution. The different steps in the process can be represented
by the following equations [116, 117]:
H2S dissociation:

H2S + OH

~ HS- + H20.

(41)

Light absorption:

CdS --* ecB(CdS) - h +(CdS).

(42)

Oxidation:

2HS- + 2h+(CdS)~S~- + 2H + .

(43)

Reduction:

2HzO + 2ecB(CdS) ~ H2 + 2 O H - .

(44)

182

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

Electron transfer:

ecB(CdS)+ TiO2 ~ eca(TiO2) + CdS.

(45)

Reduction:

2H20 + 2ecB(TiO2) ~ H2 + 2 O H - .

(46)

The overall rate of the photochemical process also includes a mass transfer step
involving diffusion of the generated hydrogen from the catalyst surface to the bulk of
the liquid phase and further transport to the gas stream bubbling through the
suspension.

5.2. Effects of variables


Effects of different variables such as pH and concentrations of different semiconductors, their combinations and preparation techniques were studied I-115-129]. The pH
had a pronounced effect on the yield of H2, increasing from 1.5 to 7 cm 3 H2/h for pH
from 0 to 14 [114, 118]. The production of hydrogen goes through a maximum with
both CdS and RuO2 concentrations. Table 6 gives the initial rates of photodecomposition with semiconductor dispersions of different compositions. The irradiation was
carried out with 25 cm 3 samples in Pyrex glass vials closed with a septum using
a 450 nm cut-off filter to remove i.r. and u.v. radiation, respectively. The solution
contained 0.1 M Na2S, 1 M NaOH and 50 mg CdS. It was found that naked CdS
particles produced H2 at a rate of 0.41 cm3/h. Addition of A1203 leads to a small
increase while TiO2 decreases it by more than a factor of 2. CdS loaded with Pt
increases the rate somewhat but CdS loaded with RuO2 is much better. The H2
generation decreases if TiO2 is added to RuO2 loaded CdS, while it increases if RuO2
loaded TiO2 is added to CdS. The rate is very small if Pt loaded TiO2 is added to CdS
while the rate drops to zero for RuO2 loaded TiO2. The hydrogen evolution-phenomena observed and presented in Table 6 can generally be explained by mechanisms
discussed in the previous section. The optimum choice of variables become very
difficult because of the possibilities of a wide range of choices for the semiconductors

Table 6
Initial rates of hydrogen production with semiconductor dispersion of different compositions
Dispersion

Hydrogen production (cm3/h)

CdS
CdS + TiO2
CdS + A1203
CdS/0.5 wt% Pt
CdS/0.5 wt% RuO 2
CdS/1.0 wt% RuO2
CdS/1.0 wt% RuO2 + TiOz
CdS + TIO2/0.5 wt% RuO2
CdS + TIO2/0.5 wt% Pt
TIO2/0.5 wt% RuO2

0.41
0.19
0.45
0.60
1.57
1.86
1.66
2.23
0.05
0.00

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159 198

183

and their respective proportions and their methods of preparation. A significant effort
has therefore been directed towards the formulation of the photocatalysts. However,
the design aspects of a practical photoreactor has not received much attention.
The addition of sulfite ions to the solution was helpful for the production of
hydrogen because of the conversion of disulfide ions ($22-) to thiosulfate ions ($20 2 -)

[1183.
S~- + S O 2--oS203z- + S 2-.

(47)

The disulfide ion acts as an optical filter, lowering H2 production by decreasing light
absorption. Borgarello and Serpone [118] studied the photochemical cleavage of
thiosulfate according to the reaction:

1.5H20+ 1.5S20~- hv 2SO~- +3H+.

(48)

They propose a two-compartment photocleavage system for H2S in the presence of


sulfite ions where hydrogen is produced in the CdS/RuO2 containing half cell and S2
is concurrently oxidized to $203z-. The thiosulfate is transferred to a second half cell
containing TiO2 particles where light driven dismutation of $20~- takes place. The
process leads to the production of 3 mol of H2 from 1 tool of H2S according to the
overall reaction:
3 H 2 0 + S 2-

h v S O 2 - ..~ 3H2.

(49)

The absence of accumulation of either sulfur or thiosulfate makes the photoreaction


very efficient.
5.3. Kinetic studies

Most studies on photoconversion of water and hydrogen sulfide have dealt with the
development and evaluation of different photocatalysts. Elaborate procedures of
preparation and treatment of these photocatalysts have been presented. However,
studies of the rate processes to develop data for the design ofa photoreactor have been
extremely limited [116, 117, 130-132]. An attempt has recently been made to describe
the light absorption, chemical and mass transfer steps in the production of hydrogen
from sulfide and sulfite solutions. Rate expressions have been developed to test how
well they fit the experimental results. Sabate et al. [117] irradiated a Pyrex quasispherical photoreactor with a 150 W Xe lamp as the light source through a round flat
window of area 7.5 cm 2. CdS catalyst was suspended in 70 cm 3 of an appropriate
substrate solution by stirring with a magnetic stirrer. The temperature of the reactor
was regulated using an air stream directed towards the reactor. The lamp was
equipped with an elliptic reflector which concentrated the radiation on the window of
the photoreactor. Oxalic-uranyl actinometry and a laser power meter indicated that
the radiation entering the reactor with wave'lengths between 300 and 520 nm was
8.9 x 10- 7 einstein/s cm 2 (equivalent to 260 mW/cmZ). Hydrogen evaluation was used
as a measure of the reaction rate. Hougen-Watson type models were developed by
considering the following steps in the overall reaction: (1) Absorption of a photon.
(2) Absorption of reactants on the catalyst surface. (3) Reaction on the catalyst surface.

184

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

(4) Desorption of intermediates and/or products. (5) Subsequent reactions of intermediates in the liquid phase.
Based on different rate limiting steps, 10 equations were developed out of which three
equations were found to fit the experimental data well. These equations could explain
the influence of sulfide, sulfate and thiosulfate concentrations, but were not adequate
to quantitatively predict the hydrogen evolution rate.
Escudero et al. [ 132] studied water photolysis using a three-phase solid-liquid-gas
photoreactor, continuous with respect to the gas phase. The continuous system
provides a greater insight into the behavior of the system while the batch system
primarily used in photodevices allows only a global view, integrating almost all effects
present in the process. The authors used two catalysts: Pt-TiO2 and Pt, RuO2-TiO2.
TiO2 was used from three different sources and the impregnation of TiO2 or
RuO2-TiO2 on Pt was done in different ways. The different catalysts thus obtained
were tested for water cleavage in two annular photoreactors of equal height and
double volume (150 and 300 cm 3 suspension). The lamp centered annular geometry
allowed better mathematical representation of the radiation flux. The energy source
was a mercury vapor UV lamp with emission spectrum 254-590 nm, the usable range
in the study being 300-416 nm, considering the band gap of titania and the optical
characteristics of the Pyrex glass used in the reactor. The radiation flux reaching the
reaction chamber of both reactors was 18 x 10-6 einstein/scm 2. The catalytic powders were kept suspended in water by means of gas bubbling through the reactor. The
effects of different variables such as catalyst composition, size, surface area and
preparation methods as well as operating variables such as catalyst concentration and
bubbling gas flow rate were studied and the results analyzed in terms of light
absorption, chemical reaction and mass transfer effects. It was concluded that the
overall rate of the photolytic process is basically determined by the radiation absorption in the photoreactor, the recombination of photoproducts on the catalyst surface
and the desorption of products from the liquid to gas. The importance of mass transfer
in limiting the process yield was shown and it was suggested that this step be
considered in the development of photocatalysts at early stages.

5.4. Developments in photocatalysts


Investigations have been made to develop photocatalysts which do not require
noble metals such as Pt, Ru, Rh or their oxides for electron transfer [119-129,
133-136]. Kakuta et al. [133] developed a ZnS-CdS catalyst supported on silica
Nation films for H2 production in aqueous sulfide solution. The activity of ZnS.CdS
system was found to be comparable to Pt/CdS system. Similar activity was observed
for the silica-supported catalysts. Kiwi and Gratzel [134] used ~-Fe203 powders and
colloids while Kudo et al. [135] used NiO. K4Nb6017 powder for the photodecomposition of water. Lu and Li I-119] developed a new photocatalyst, ZnFe204 spinel
powder without any noble metal loading for photodecomposition of H2S. The spinel
powder was prepared by adding NH3/H20 into a solution of the required amount of
Fe(NO3),9H20 and Zn(NO3)2,6H20. The precipitate was washed with water, dried
and calcined in air for 10 h, then powdered. The photoactivity was determined from

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

185

the measurement of hydrogen volume (25C, 760mm Hg) produced during the
irradiation with a 750 W xenon lamp of a suspension of ZnFe204 spinel powder in
0.1 M NazS solution. The photoreactor was a quartz glass filled with 25 ml solution at
pH 12 and room temperature. The photocatalytic activity of ZnFe204 was much
higher than that of ZnS/CdS and was comparable to Pt/CdS.
The rates of hydrogen production from the photodecomposition of H2S using
VO/VS mixed semiconductor was not successful because of low efficiency [ 120, 121 ].
Mixed sulfides of Cd with Zn allowed more efficient hydrogen generation than CdS
alone [121]. Use of CdS particles loaded with 0.5% RuS2 gave a seven-fold improvement of hydrogen evolution from alkaline S2 /SO32 solutions compared to
CdS/0.5 wt% RuO2 [123]. Active catalysts have been prepared by developing suspensions of semiconductors with micro-hetero-junctions formed by CdS or solid solutions of ZnyCda_yS and Cu~S. The optimization of the conditions leading to the
production of the junctions is still an unresolved matter [124]. The use of nonaqueous
solution of H2S to avoid the difficulties with thiosulfate ions formed in the presence of
sulfite ions was explored. Suspension of CdS/RuO2 in acetonitrile saturated with H2S
in presence of an alkene generates hydrogen on irradiation. However, there was
evidence that hydrogen production was reduced because the alkene intercepts the
hydrogen as well as sulfur ions [125].
Polymer immobilized semiconductor catalysts have recently been used for photodecomposition of water and hydrogen sulfide [-126, 127] (see Table 7). The polymersupported dispersed semiconductors are specially appealing because it can
considerably ease problems related to the removal of reaction products from the
reactor. Besides, the polymeric matrix can participate in the photocatalyst's action
and change the physicochemical properties of the semiconductor. The cation or anion
exchange properties of polymers also open up wide possibilities for modifying photocatalyst content and structure. Gruzdkov et al. [126] used cadmium sulfide supported
on cation-exchange polymer for decomposing water-soluted hydrogen sulfide. Finely
dispersed group VIII metals co-precipitated with cadmium sulfide on the polymer
considerably enhanced the reaction rate. A special treatment of cadmium sulfide with
copper or silver was found to result in a sufficient expanding of the spectral range of
the photocatalyst as well as in a substantial rise of its activity.

5.5. High temperature photolysis


Irradiation at high temperatures (623-973 K) in the presence of vanadium sulfide
on A1203, TiO2, ZnO and ZnS leads to an increase of 14-40% in the decomposition
of HzS, compared to the reaction carried out in absence of light [128]. The high
temperature photoreaction was extended to a HzS/CO 2 mixture and the results
indicated that 9-13% of the H2 produced by the splitting of HzS is converted to CH4
or C2H6 and 50% of the CO2 converts to CH4, C2H6, CO and COS. It was suggested
that the consumption of hydrogen in the hydrocarbon forming reactions results in
equilibrium shift and hence a further rise in the conversion of H2S. The authors could
not compare their results because no data could be found in the literature on high
temperature photolytic system.

186

J. Zaman, ,4. C h a k m a / F u e l Processing Technology 41 (1995) 159-198

.,=

~.~~~

~o '~
,- ~

~.~
0

;~

~''~

0.~

ID~

r . . ~

.-~

~-~
~5

.8

's

,S

.o.

o,=

i=

_9o

-8
~

.o
~s
0

c~
"r'

0
f~

;>

"~
&

0
0
e~
o

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

r,j ~
.o

~ ~

o
0

~,

,,~

o~

~
~

..~ .~ ~,
~a ~) r..- . <

o ~

~.o_~ ~,~

.-~.

=.=_ ~ o

o .o ~ ~ ~ ~ ~
m

r ,

.~ ~ ~~

".~

:~

~o "s
.~.. .~. ~ .o. . ._. ~ = ~0 ,

~ ' ~ - ~

"0

o~.r.

"d
"~

.o

. m

o
o
r.~

-~-,~, r' um

,4

+~
"uO0

~o~

r..)

0 ~

~o~
~o

r,.)

187

188

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

5.6. Photoelectrochemical methods

The advantage with photoelectrochemical method is that one can not only produce
hydrogen but also electricity. Kainthla and Bockris [ 129] used a photoelectrolytic cell
with chemically deposited thin CdSe film as photoanode and Pt as cathode, the anode
and cathode compartments being separated by nation membrane. The electrolyte
used in the cathode compartment was 1 M NaOH. For use in the anode compartment,
H2S was bubbled through 1 M NaOH to make the solution saturated with HzS.
Appropriate amounts of sulfur and NaOH were then added to make the solution 1 M
in each. Irradiation was done through a quartz window using a Xe arc lamp adjusted
to 100 mW/cm 2 at the electrode surface. The light intensity was measured using
a standardized Eppley precision radiometer. The photogenerated holes on the surface
of the photoanode oxidize the sulfide ions to polysulfide:
S2 - + 2h --* S,

(50)

S 2 - + S --, S ~ - .

(51)

The electrons flow through the back of the semiconductor and then through the
external resistor to the cathode and reduce H20 to H2:
2H20 + 2e- ~ H 2 + 2 O H - .

(52)

The hydrogen evolved during the cell operation can be collected. The polysulfide
formed dissolves in the anolyte till the solubility limit is reached after which sulfur
starts to precipitate out. The maximum light to stored chemical energy conversion
efficiency occurs for zero cell voltage when the amount of photocurrent flowing
through the cell is maximum. The efficiency of this process was found to be 1.5%. At
any other cell voltage, both electrical and stored chemical energies could be
withdrawn simultaneously. The maximum light to electrical energy conversion in this
case occurred at a cell voltage of 300 mV and had a value of 1.8%. The maximum light
to total energy conversion (electricity and chemical energy in the form of H2) occurred
at a cell voltage of 275 mV and had a value of 2.85%. The cell showed 97% current
conversion efficiency to hydrogen. If the hydrogen produced in the cell were used in
a H2/air fuel cell for electricity generation, practical efficiency of 10.6% could be
obtained. If the cathode is placed in the same compartment as the anode, the following
cathodic reaction takes place and there is no production of hydrogen:
$2z- + 2e- --* S2-.

(53)

The efficiency of conversion of light to electricity increases from 1.8% to 2.4%. The
increase in efficiency is due to the absence of the Nation membrane (to prevent the
sulfide and polysulfides from coming into contact with the cathode) which increases
the cell internal resistance and reduces the short circuit current and fill-factor, thus
i'educing the efficiency. This points to the possibility of improving the conversion
efficiencies by reducing the membrane resistance and optimizing the cell geometry to
reduce cell internal resistance.

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

189

5. 7. Assessment of photolytic methods


Significant advances have been made in the development of semiconductors
and photocatalysts over the past few years. Compared to that, work on the
development of an effective photoreactor has been very limited. There is a great
need for research in kinetics and transport processes during photoreaction and
the application of chemical engineering approach to the design of photoreactors
will likely be worthwhile.

6. Plasma methods

Plasma is an ionized gas which may be produced by a number of methods:


combustion, flames and explosions, nuclear reactions, electrically heated furnaces,
electrical discharges (coronas, sparks, glows, arcs, microwave discharges, plasma jets),
shocks (electrical, magnetic and chemically driven), adiabatic compression technique
and bombardment of quanta of electromagnetic fields or high energy particles [137,
138]. Plasmas are variously classified depending on their energy level, temperature
and ionic density. In chemical literature, we come across terms like high temperature
plasma, hot plasma, low temperature plasma or cold plasma. High temperature
plasmas usually have applications in thermonuclear reactions. Most chemical and
metallurgical applications deal with low temperature plasma, often called thermal or
hot plasma. They also go by the term low pressure plasma or high pressure plasma,
depending on the pressure level of operation. These plasmas have energies less than
10eV per molecule, temperature below 105K and ionic density greater than
102 perm 3. Gauvin [139] records a number of applications of these plasmas for
metallurgical and chemical processing. The cold plasmas, usually generated by glow
discharges have slightly higher energy density but much lower ionic density compared
to hot plasmas. These plasmas do not reach thermal equilibrium and possess high
electronic excitation without gas heating. The cold plasmas operate at low pressures
and can be referred to as low pressure plasmas. These plasmas have found applications in chemical processing as a source of active species and have been used by
chemists for many exotic reactions. The cold plasmas are at nonequilibrium condition
and can produce product beyond thermal equilibrium [140].
Chemical processes can be operated at temperatures up to 20 000 K and pressures
from 10 to 109 Pa. using low temperature olasma. The high temperature generated by
the plasmas allow very high conversions for thermodynamically limited endothermic
reactions. The ability to attain high conversions in practice depends on the effectiveness of the quenching system. Not only is a high quenching rate needed, in many
reactions such as the cracking of hydrocarbons to produce acetylene, the moment
when quenching is applied is also important, because of the possibility of undesirable
reactions [141-143]. The quenching has most often been done by surface heat
exchange, direct injection of water, expansion through nozzle or by means of
a fluidized bed. In all cases, cooling rate of 10 6 deg/s or higher has been obtained
[142, 143].

190

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

The hot plasmas are generated using a variety of high intensity arcs, sometimes
added with jet devices (plasma jets) and using induction heating in radio frequency
field (electrodeless) [144-148]. Information on the application of either hot or cold
plasmas for the thermal decomposition is scarce. The material that follows includes
application to other gas phase reactions and have been included because similar
techniques could be extended to hydrogen sulfide decomposition [-149-163]. A brief
description of the different applications of plasma reactors is provided in Table 8.
A large amount of work on plasma decomposition of hydrogen sulfide has been
carried out in former Soviet Union [149-153]. Bagautdinov et al. [149-151] report
the construction of a pilot plant for the decomposition of hydrogen sulfide based on
experiments in the laboratory. Laboratory investigations at low power MCW installations showed that plasma dissociation allows to obtain 1 m 3 of hydrogen and
1.4kg of sulfur from 1 m 3 of H2S with energy cost of about 0.5-1.0eV/mol
(0.6-1.2 kWh/m3). Based on this, a powerful plasmochemical plant was developed and
built in order to explore this method at MCW power up to 1 MW and productivity
about 1000 Nma/h. Both the laboratory and pilot units were used successfully for the
decomposition of H2S alone or mixed with CO2.
Traus and co-workers 1-154, 155] used an ozonizer and an atmospheric pressure
rotating glow discharge to decompose H2S. An ozonizer discharge is characterized by
a dielectric barrier between the electrodes. The discharges are weak and the system is
usually operated at low temperatures. The authors built a double quartz wall ozonizer
which will withstand the chemical and electrical environments at higher temperatures.
The decomposition was studied with electrode temperature 443-833 K, gas flow 50
and 100 sccm, initial concentration of H2S 20-100%, input voltage 0-15 kV and
added gases Ar, H2 and N2. Silver electrodes were used and the product could be
removed as a liquid or vapor without problem.
The essential feature of atmospheric rotating glow discharge was the use of two
concentric electrodes mounted in an axial electric field [155]. Both electrodes were
made of nonmagnetic stainless steel. The outer electrode was built with a double wall
with circulating heating oil to heat the electrode up to 443 K. The operating temperature was about 449 K where liquid sulfur has a pronounced low viscosity. The initial
concentration of H2S was in the range 10-100/% using hydrogen or argon as the
diluent gas. The total gas flow rates were 50-250 sccm, the electrical power 45-75 W
at electrode gap of 3.5 mm and magnetic coil current 5-10 A. Table 9 compares the
performance of the ozonizer and the atmospheric glow discharge. The conversion in
the ozonizer is considerably lower, because the micro-discharges at the low pressure of
the ozonizer are too short to achieve higher conversions. It is seen that the efficiency of
the rotating glow unit was 1-2 and of the ozonizer 0.3-0.8, while the value on
a theoretical basis is 75. The authors expect that the efficiency of the glow unit could
be improved by an order of magnitude and then the decomposition by the plasma
process can become economically attractive. A mechanism of the reaction was also
proposed by the authors. The following reaction path postulated based on study with
short-lived radio frequency pulse was verified by computer simulation [157]:
H2S --* 2H + S,

(54)

.I. Zaman, .4. Chakma/Fuel Processing Technology 41 (1995) 159 198


o

.~ .~

= ~

~?

e
tt~

"T~
.=

~ ~~

-=

,6

.=_
e!
o

c5

"6
o

ea

+
+0

~=

"~

?4

d~

d
"6

S
t

r~

"r

b-

,.r-

191

192

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

Table 9
Comparison of H2S dissociation in the ozonizer and the glow discharge
Parameter

Ozonizer

Glow discharge

Conversion y%
Energy efficiency y+ (mol/kWh)

0.5-12
0.34).8

17-40
1-2

Maximum y%
Corresponding y+ (tool/kWh)
Conditions

12
0.3
50 sccm, 50 W

40
1
50 sccm, 45-70 W

Maximum Ye (mol/kWh)
Corresponding y%
Conditions

0+7
2
100 sccm, 7 W

2
20
250 sccm, 45-75 W

H + H 2 S --~ SH + H 2 ,

(55)

H + SH --, H2 + S,

(56)

2SH ~ S + H2S,

(57)

2S + M ~ $ 2 + M,

(58)

252

+ M + S+ + M.

(59)

A patent was awarded to a Japanese company for the recovery of hydrogen and
sulfur from H2S through decomposition by electric discharge [158]. 99.5% H2S was
passed through a cylindrical electrode (27.85 mm dia, stainless steel), where glow
discharge was carried out with a current of 0.264 A at 825 V and 218 mm pressure.
After a discharge for 1 min with 1.361 H2S, the exhaust gas was cooled, filtered from
S and the unconverted H2S removed. The remaining 0.381 gas was pure H2. The
recovered sulfur was 0.53 g with a purity of 99.6%.
Givotov et al. [159] examined the hydrogen production in plasma both in a onestage process of water vapor decomposition and in a two-stage process with preliminary reduction of CO2 to CO followed by water gas shift reaction of CO and H20,
forming hydrogen. Carbon dioxide decomposition was studied in U.H.F. discharge,
H.F. discharge and electron beam discharge. U.H.F. radiation of 1.7 kW and
2400 MHz was used with a rectangular wave guide. The reactor was a quartz tube
crossing the wide wall of the wave guide. The CO2 pressure in the reactor ranged from
6.6 to 26.6 kPa and the flow rate was adjusted for the specific energy input to the
discharge equal to 0.7-1.4 eV/mol CO2, this energy range being optimal for nonequilibrium discharge. It was found out that the double stage was energetically more
efficient than the single step process of water dissociation.
Rubtsova and co-workers [161] discussed the role of catalysts in plasma reactions.
Suib and Zerger [162] presented some interesting results on the conversion of
methane to ethylene, ethane and acetylene in a catalytic plasma reactor. The low
temperature plasma was generated by a microwave discharge. The catalyst was placed
downstream of the reactor at the edge of the plasma zone. The plasma provided the

,1. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

193

radicals subsequently used by the catalyst to produce the products. Nickel powder
was found to be an effective catalyst as opposed to Li/MgO, Sm203 and other
materials used for the oxidative coupling of methane to produce similar products. The
conversion of methane was as high as 52% with selectivities to ethane 50%, ethylene
25% and acetylene 25%. This is in sharp contrast to other oxidative coupling systems
where high selectivities are obtained only at very low conversion of methane. The
thermal temperature of the plasma reactor was less than 473 K and the electronic
temperature in the order of 2000 K and gas pressure of 1.3 to 6.6 kPa.
Laflamme et al. [163] designed a thermal plasma reactor to study the direct
conversion of natural gas to liquid hydrocarbons, namely, benzene, toluene and xylene
(BTX). Equilibrium calculations showed that the aromatic species were stable at
temperatures below 2000 K and the maximum BTX concentrations could be obtained
at 1450 K and atmospheric pressure. Kinetic calculations showed that a residence time
of at least 1 s at 1300-1500 K was needed for maximum BTX conversion, which appear
as intermediate compounds in the synthesis route between acetylene and soot. As the
long residence time was incompatible with standard plasma reactor technology, it was
achieved by the combined use ofa DC plasma torch for the rapid heating of the reaction
mixture to the required temperature and of resistive heating elements to extend the high
temperature zone throughout the one-meter long tubular reactor (4.2 cm i.d.). Experiments were conducted using the plasma power level of 7.0 kW d.c. plasma torch of
standard design. The plasma gas (62% mixture of Ar/N2) flow rate was adjusted in the
range 8-101/min to give a residence time of 0.63-0.94 s with the ratio of methane to
plasma gas maintained at 0.185 by volume. Experiments conducted with or without
plasma, but at the same thermal conditions showed that there was no reaction of
methane in the absence of plasma. In the presence of plasma, conversion of methane
rises from 34.4% to 39.5% by weight with the BTX content from 2.3% to 3.2% by
weight as the residence time increased from 0.62 to 0.93 s. The plasma torch was
provided with a buffer section so that high temperature plasma gas (5000-8000 K) is
cooled to 2000-2500 K before its contact with the methane gas. The product gas from
the reactor was quenched by direct quench with water or by a cooling surface. The
direct quenching was more effective leading to the complete elimination of soot and
a slight increase in acetylene and BTX formation.

6.1. Assessment o)Cplasma methods


Plasmas have not so far been used for any large scale chemical processing. The biggest
obstacle possibly is the use of electrical energy. However, the plasma process is environmentally friendly and operationally simple with minimum waste generation. None of the
work reported in the literature, however, has reached a mature level of development.

7. Conclusions
The diverse attack on hydrogen sulfide to obtain two salable products is very
striking. Every year a large amount of potential resource is being wasted and there is

194

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

no doubt it should be stopped. The success in the development of a suitable technology for the production of hydrogen and sulfur will signify the attainment of the triple
objectives of waste minimization, resource utilization and environmental pollution
reduction. An assessment of the current level of progress in different technologies
- thermal, thermochemical, electrochemical, photolytic and plasmolytic, has been
made at the end of discussions in the respective sections. The purpose here is to draw
the line between what is achievable today and what is waiting on the horizon in order
to assess whether the full utilization of hydrogen sulfide would be feasible in the near
future.
As of today, there is no commercial technology for the production of hydrogen and
sulfur from hydrogen sulfide. Any technology proposed for the purpose has to
overcome the built-in inertia of the existing Claus technology and the tail gas
technologies. At the same time, it must be realized that, in many environments,
hydrogen is not a scarce commodity and it can be manufactured quite cheaply using
standard technologies. Keeping these factors in view, there seems to be no doubt that
a venture technology will not be acceptable in the market today and the processing
technology has to follow the pattern of established technologies.
Photochemical and plasmochemical technologies are still in the developmental
stages and are not mature enough to be applied to large scale chemical processing.
Electrochemical technology is established in certain areas, but its application to
hydrogen sulfide requires further development. In addition, it is unlikely that electrochemical process can be competitive in today's electricity costs. This leaves us with the
thermal and thermochemical methods, both of which fall in the realm of established
technologies.
Of the thermal methods, the membrane, thermal diffusion and solar technologies
have not yet developed very well. In fact, the membrane technology which appears
very attractive is essentially a technology of the future. Particularly for system as
difficult as hydrogen sulfide, the application has to wait till the technology matures
with less demanding systems. Based on these considerations, the authors came to the
view that the following two processes merit further analysis to act as the basis for
a prospective technology for commercialization: (1) Thermal decomposition at
around 1073 K at reduced pressure in presence of catalysts in a fixed bed reactor. (2)
Two-step sulfide process in the temperature range 773-923 K in fluidized bed reactors
as reactor-regenerator system.
The products from both the systems may pass through heat exchange, sulfur
removal by condensation and hydrogen removal by pressure swing adsorption systems followed by the recycle of the unconverted H2S. In order to establish the
technical and economic feasibilities of such a process, the following set of actions are
needed: (1) Preliminary estimates of process plants based on the above technologies.
(2) Research on the development of catalysts and choice of sulfides and obtain kinetic
information. (3) Set up pilot plants and develop scale up procedures to procure design
information to build a commercial plant.
The above suggestions are in no way intended to discourage research effort in other
approaches for treating hydrogen sulfide. There may, in fact, not be a universal
approach for the desulfurization process. The economics of a process may be

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

195

influenced by diverse factors to make different processes desirable depending on the


size of the plant, the source of H2S, its temperature and concentration, the local energy
situation or the environmental regulations. While research in other areas proceed, it is
recommended that the proper utilization of H2S should start immediately using
processes based on conventional technologies. It is important that the resource
utilization and waste minimization aspects of the proposal be appreciated and both
the government and the industry be persuaded to put their efforts on this and change
the status of hydrogen sulfide from an unwelcome material to a profitable resource.

References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]

Bowman, C.W. and Du Plessis, M.P., 1986. Int. J. Hydrogen Energy, 11(1): 43.
Gregory, D.P., Tsaros, C.L, Arora, J.L. and Nevrekar, P., 1980. ACS Syrup. Ser., 116: 3.
Minet, R.G. and Desai, K., 1983. Int. J. Hydrogen Energy, 8: 285.
Schuetz, G.H., 1985. Int. J. Hydrogen Energy, 10(7/8): 439.
Steinberg, M. and Cheng, H.C., 1989. Int. J. Hydrogen Energy, 14: 797.
LeRoy, R.L., 1983. Int. J. Hydrogen Energy, 8(6): 401.
Wendt, H. and Imarisio, G., 1988. J. Appl. Electrochem., 18: 1.
Somorjai, G.A., Hendewerk, M. and Turner, J.E., 1984. Catal. Rev. Sci. Eng., 26(3/4): 683.
Fletcher, E.A., Noring, J.E. and Murray, J.P., 1984. Int. J. Hydrogen Energy, 9: 587.
Bockris, J. O'M, Dandapani, B., Cocke, D. and Ghoroghochin, 1985. Int. J. Hydrogen Energy, 10(3):
179.
Vezirglu, T.N. and Barbir, F., 1992. Int. J. Hydrogen Energy, 17(6): 391.
Nitsch, J., Klaiss, H. and Meyer, J., 1992. Int. J. Hydrogen Energy, 17: 651.
Scott, D.S,, 1993. Int. J. Hydrogen Energy, 18(3): 197.
Fletcher, E.A., 1983. Int. J. Hydrogen Energy, 8: 835.
Raymont, M.ED., 1975. Hydrocarb. Proc., 54(7): 139.
Raymont, M.ED., 1980. ACS Syrup. Ser., 116: 333.
Clark, P.D. and Wassink, B., 1990. Alberta Sulfur Res. Quart. Bull., 26(2/3/4): 1.
Raymont, M.E.D. and Hyne, J.B., 1975. Alberta Sulfur Res. Quart. Bull., 12(3): 1.
Kaloidas, V.E. and Papayannakos, N.G., 1987. Int. J. Hydrogen Energy, 12: 403.
Raymont, M.E.D., 1974. The thermal decomposition of hydrogen sulfide, PhD Thesis, University of
Calgary, Canada.
Raymont, M.ED. and Hyne, J.B., 1975. Alberta Sulfur Res. Quart., 12(2): 3.
Kaloidas, V.E. and Papayannakos, N.G., 1989. Chem. Eng. Sci., 44: 2493.
Tesner, P.A., Nemirovsky, M. and Motyl, D.N., 1991. Kinetics and Catalysis, 31: 1081.
Darwent; R. de B. and Roberts, R., 1953. Proc. Roy. Soc. London, A216: 344.
Kaloidas, V.E. and Papayannakos, N.G., 1991. Ind. Eng. Chem. Res., 30: 345.
Bandermann, F. and Harder, K.B., 1982. Int. J. Hydrogen Energy, 7: 471.
Cbivers, T., Hyne, J.B. and Lau, C., 1980. Int. J. Hydrogen Energy, 5: 499.
Fukuda, K., Dokiya, M., Kameyama, T. and Kotera, Y., 1978. ind. Eng. Chem. Fund., 17: 243.
Sugioka, M. and Aomura, K., 1984. Int. J. Hydrogen Energy, 9: 891.
Sugioka, M. and Aomura, K., 1984. Adv. Hydrogen Energy (Hydrogen Energy Prog. 5, Vol. 2), 4:
477.
Sugioka, M. and Kanazuka, T., 1988 Nippon Kagaku Kaishi, 8: 1294.
Dokiya, M., Kameyama, T. and Fukuda, K., 1977. Denki Kagaku, 45: 701.
Kameyama, T., Fukuda, K., Dokiya, M. and Kotera, Y., 1979. Jpn patent 78,99,078.
Kameyama, T., Dokiya, M., Fukuda, K. and Kotera, Y., 1979. Sep. Sci. Technol., 14: 953.
Kameyama, T., Fukuda, K., Fujishige, M., Yokokawa, H. and Dokiya, M., 1981. Adv. Hydrogen
Energy (Hydrogen Energy Prog. 2), 2: 569.

196

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

[36] Kameyama, T., Dokiya, M., Fujishige, M., Yokokawa, H. and Fukuda, K., 1981. Ind. Eng. Chem.
Fund., 20: 97.
1-37] Kameyama, T., Dokiya, M., Fujishige, M., Yokokawa, H. and Fukuda, K., 1983. Int. J. Hydrogen
Energy, 8: 5.
[38] Edlund, D.J. and Pledger, W.A., 1993. J. Membr. Sci., 77(213): 255.
[39] Zaman, J. and Chakma, A., 1993. In: R. Clift and J.P.K. Seville (Eds.), Gas Cleaning at High
Temperatures Blackie Academic and Professional, London, p. 671.
[40] Dokiya, M., Kameyama, T. and Fukuda, K., 1978. Jpn Patent 78,130,291.
[41] Toyobo Co. Ltd., 1980. Jpn Patent 80,119,439.
[-42] Abe, F., 1987. Eur. Pat. Appl 228, 885.
[43] Nishijawa, T., Tanaka, Y. and Hirota, K., 1979. Int. Chem. Eng., 19: 517.
1-44] Chivers, T. and Lau, C., 1987. Int. J. Hydrogen Energy, 12: 561.
1-45] Chivers, T. and Lau, C., 1987. Int. J. Hydrogen Energy, 12: 235.
[46] Bishara, A., Salman, O.A., Kharishi, N. and Marafi, A., 1987. Int. J. Hydrogen Energy, 12: 679.
[47] Salman, O.A., Bishara, A. and Marafi, A., 1987. Energy, 12: 1227.
[48] Diver, R.B., Pederson, S., Kappauf, T. and Fletcher, E.A., 1983. Energy 8: 947.
[49] Kappauf, T., Murray, J.P., Palumbo, R., Diver, R.B. and Fletcher, E.A., 1985. Energy, 10:1119.
[50] Hsieh, H.P., Bhabe, R.R. and Fleming, H.L., 1988. Microporous alumina membranes, J. Membr. Sci.,
39: 221.
[51] Hsieh, H.P., 1991. Catal. Rev. Sci. Eng., 33(1 & 2): 1.
[52] Uzio, D., Peureux, J., Fendler, A.G., Torres, M., Ramsay, J. and Dalmon, J.A., 1993. Appl. Catal., 96:
83.
[53] Wu, J.C.S., Flowers, D.F. and Liu, P.K.T., 1993. J. Memb. Sci., 77: 85.
[54] Shu, J., Grandjean, B.P.A., Seste, A. van and Kaliaguine, S., 1991. Can. J. Chem. Eng., 69: 1036.
[55] Kotera, Y., 1976. Int. J. Hydrogen Energy, 1: 219.
[56] O'Keefe, D.R. and Norman, J.H., 1980. Catal. Rev. Sci. Eng., 22(2): 325.
[57] Yoshida, K., Kameyama, H., Auchi, T., Nobue, M., Aihira, M., Amir, R., Kondo, H., Sato, T.,
Tadokora, Y., Yamaguchi, T. and Sakai, N., 1990. Int. J. Hydrogen Energy, 5(3): 171.
1-58] Nishimoto, Y., Mizumoto, Y., Hasegawa, S. and Misuoka, S., 1976. Jpn. patent 76,02,696.
[59] Kiuchi, H., Iwasaki, T., Nakamura, I. and Tanaka, T., 1980. ACS Symp Ser. 116: 349.
[60] Kiuchi, H., Nakamura, T., Funaki, T. and Tanaka, T., 1982. Int. J. Hydrogen Energy, 7: 477.
[61] Kiuchi, H., Funaki, K., Nakai, Y. and Tanaka, T., 1982. Adv. Hydrogen Energy (Hydrogen Energy
Prog. 4), 2: 543.
[62] Kiuchi, H., Funaki, K. and Tanaka, T., 1983. Metall. Trans. B., 14: 347.
[63] AI-Shamma, L.M. and Naman, S.A., 1989. Int. J. Hydrogen Energy, 14(3): 173.
[64] A1-Shamma, L.M. and Naman, S.A., 1990. Int. J. Hydrogen Energy, 15(1): 1.
[65] Chen, Q., Li, L. and Hepler, L.G., 1991. Can. J. Chem. Eng., 69: 1160.
[66] Yumura, M. and Fwimsky, E., 1985. Appl. Catal., 16(2): 157.
[67] Dokiya, M., Fukuda, K., Yokokawa, H. and Karneyama, T., 1978. Bull. Chem. Soc. Japan, 51: 150.
[68] Mehra, A. and Sharma, M.M., 1988. Chem. Eng. Sci., 43: 1071.
[69] Dokiya, M., Kameyama, T. and Fukuda, K., Thermochemical hydrogen preparation. V., 1979.
A feasibility study of the sulfur-iodine cycle, Int. J. Hydrogen Energy, 4(4): 267.
1-70] Shell International Research, 1965. Neth. patent 302 075.
[71] Soliman, M.A., Calty, R.H., Couger, W.L. and Funk, J.E., 1975. Can. J. Chem. Eng., 53: 164.
[-72] Odera, Y., Todo, N. and Fukuda, K., 1973. Jpn. patent 73,28,393.
[73] Fukuda, K., Dokiya, M. and Kameyama, T., 1978. Jpn. patent 78,134,791.
[-74] Kameyama, T., Dokiya, M. and Fukuda, K., 1979. Jpn. patent 79,16,395.
[75] Cocuzza, G. and Musso, G., 1978. U.S. patent 4,094,962.
1-76] Behie, L.A., Berk, D., Bishnoi, P.R. and Svrcek, W.Y., 1981. Can. patent, 1,134,596.
[77] Herrington, D.R., and Kuch, P.L., 1984. U.S. patent 4,432,960.
[78] Plummer, M.A., 1986. U.S. patent 4,592,905.
[79] James, B.R., Li-Lee, C., Lilga, M.A. and Nelson, D.A., 1987. U.S. patent 4,693,875.
[-80] Doenitz, W., Schmidberger, R. and Steinhill, E., 1980. Int. J. Hydrogen Energy, 5: 55.

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159 198

[81]
[82]
[83]
[84]
[85]
[86]
[87]
[88]
[89]
[90]
[91]
[92]

[93]
[94]
[95]
[96]
[97]
[98]
[99]
[100]
[101]
[102]
[103]
[104]
[105]
[106]
[107]
[108]
[109]
1-110]
[111]
[112]
[113]
[114]
[115]
[116]
[117]
[118]
[119]
[120]
[121]
[122]
[123]
[124]

197

Divisek, J., Malinowsky, P., Mergel, J. and Schmitz, H., 1985. int. J. Hydrogen Energy, 10(6): 383.
Janjua, M.B. and LeRoy, R.L., 1985. Int. J. Hydrogen Energy, 10(1): 11.
Balej, J., 1985. Int. J. Hydrogen Energy, 10(2): 89.
Millet, P., Durand, R. and Piner, M., 1990. Int. J. Hydrogen Energy, 15(4): 245.
Dutta, S., 1990. Int. J. Hydrogen Energy, 15(6): 379.
Schreiber, M., Lucier, G., Ferrante, J.A. and Huggins, R.A., 1991. Int. J. Hydrogen Energy, 16(6): 373.
Bolmer, P.W., 1966. U.S. patent 3,249,522.
Bolmer, P.W., 1968. U.S. patent 3,409,520.
Dandapani, B., Scharifker, B.R, and Bockris, J.O'M., 1986. Proc. Intersoc. Energy Convers. Eng.
Conf., 1: 228.
Venkateson, B., Pinsky, N. and Sapru, K., 1985. U.S. patent 4,433,461.
Shih, Y.S. and Lee, J.L., 1986. Ind. Eng. Chem. Proc. Des. Dev., 25: 834.
Anani A.A., Mao, Z., White, R.E., Srinivason, S. and Appleby, A.J., 1990. Electrochemical production
of hydrogen and sulfur by low temperature decomposition of hydrogen sulfide in an aqueous
alkaline solution, J. Electrochem. Soc., 137: 2703.
Mao, Z., Anani, A., White, R.E., Srinivason, S. and Appleby, A.J., 1991. J. Electrochem. Soc., 138:
1299.
Kalina, D.W. and Maas, E.T., Jr., 1985. Int. J. Hydrogen Energy, 10(3): 157.
Kalina, D.W. and Maas, E.T., Jr., 1985. Int. J. Hydrogen Energy, 10(3): 163.
Olson, D.C., 1984. U.S. patent 4,443,423.
Olson, D.C., 1985. U.S. patent 4,540,561.
Mizuta, S., Kondo, W. and Fujii, K., 1984. Denki Kagaku, 52: 688.
Kondo, W., Mizuta, S. and Fujii, K., 1984. Denki Kagaku, 52: 693.
Mizuta, S., Kondo, W., Fujii, K., Iida, H., Isahiki, S., Noguchi, A., Kikuchi, T., Sue, H. and Sakai, K..
1991. Ind. Eng. Chem. Res., 30: 1601.
Fujii, K., Kondo, W., Mizuta, S., Oosawa, Y. and Kumagai, T., 1989. Jpn. patent 1-53201.
Lira, H.S. and Winnick, J., 1984. J. Electrochem. Soc. 131: 562.
Weaver, D. and Winnick, J., 1987. J. Electrochem. Soc., 134: 2451.
Winnick, J., 1990. Sep. Sci. Techno., 25: 2057.
Fletcher, E.A., 1983. Energy, 8: 835.
Nygren, K., Atansoski, R., Smyrl, W.H. and Fletcher, E.A., 1989. Energy, 14(6): 323.
Gratzel, M. (Ed.), 1993. Energy Resources Through Photochemistry and Catalysis, Academic Press,
New York.
Getoff, N., 1990. Int. J. Hydrogen Energy, 15: 407.
Wald, F.V., 1984. Catal. Rev. Sci. Eng., 26: 709.
Kudo, A., Tanaka, A., Domen, K., Maruya, K., Aicha, K. and Onishi, T., 1988. J Catal., 111: 67.
Zamarev, K.I. and Parmon, V.N., 1983. In: M. Gratzel (Ed.), Energy Resources Through Photochemistry and Catalysis, Academic Press, New York, p. 123.
Sabate, J., Cervera-March, S., Simarro, R. and Gimenez, J., 1990~ Int. J. Hydrogen Energy, 15(2): 115.
Cervera-March, S., Borrell, L., Gimenez, J., Simarro, R. and Andujar, J.M., 1992. Int. J. Hydrogen
Energy, 17: 683.
Borgarello, E., Kalyanasundaram, K. and Gratzel, M., 1982. Helvetica Chimica Acta, 65: 243.
Serpone, N., Borgarello, E. and Gratzel, M., 1984. J. Chem. Soc. Chem. Commun., 2: 342.
Grzyll, L.R., Thomas, J.J. and Bardie, R.G., 1989. Int. J. Hydrogen Energy, 14: 647.
Sabate, J., Cervera-March, S., Simarro, R. and Gimnez, J., 1990. Chem. Eng. Sci., 45: 3089.
Borgarello, E. and Serpone, N., 1985. Int. J. Hydrogen Energy, 10: 737.
Lu, G. and Li, S,, 1992. Int. J. Hydrogen Energy, 17: 767.
Naman, S.A., Aliwi, S.M. and A1-Emara, K., 1986. Int. J. Hydrogen Energy, I1(1): 33.
Naman, S.A. and AI-Emara, K., 1987. Int. J. Hydrogen Energy, 12: 629.
Khan, M.M.T., Bhardwaz, R.C. and Bhardwaz, C., 1988. Int. J. Hydrogen Energy, 13(2): 7.
Thewissen, D.H.M.W., Zonwen-Assink, E.A., Timmer, K., Tinnemans, A.H.A. and Mackor, A., 1984.
J. Chem. Soc. Chem. Commun., 941.
Savinov, E.N., Gruzdkov, Yu A. and Parmon, V.N., 1989. Int. J. Hydrogen Energy, 14(1): 1.

198

J. Zaman, A. Chakma/Fuel Processing Technology 41 (1995) 159-198

[125] Green, M. and Elofson, R.M., 1985. J. Chem. Soc. Chem. Commun. 830.
[126] Gruzdkov, Yu. A., Savinov, E.N. and Parmon, V.N., 1987. Int. J. Hydrogen Energy, 12(6): 393.
[127] Mau, A.W.H. Huang, C.B., Kakuta, N., Bard, A.J., Campion, A., Fox, M.A., White, J.M. and Weber,
S.E., 1984. J. Am. Chem. Soc., 106: 6537.
[128] Naman, S.A., 1992. Int. J. Hydrogen Energy, 17: 499.
[129] Kainthla, R.C. and Bockris, J, O'M., 1987. Int. J. Hydrogen, 12(1): 23.
[130] Domen, K., Kudo, A. and Onishi, T., 1986. J. Catal., 102: 92.
[131] He, Z.X. and Pong, W., 1990. Int. J. Hydrogen Energy, 15(2): 99.
[132] Escudero, J.C., Simarro, R., Cervera-March, S. and Gimenez, J., 1989. Chem. Eng. Sci., 44: 583.
[133] Kakuta, N., Park, K.H., Finlayson, M.F., Ueno, A., Bard, A.J., Campion, A., Fox, M.F., Webes, S.E.
and White, J.M., 1985. J. Phys. Chem., 89: 732.
[134] Kiwi, J. and Gratzel, M., 1987. J. Chem. Soc. Farad. Trans., 83: 1101.
[135] Kudo, A., Tanaka, A., Domen, K., Maruya, K., Aika, K. and Onishi, T., 1988. J. Catal., 111: 67.
[136] Kutty, T.R.N. and Avudaithai, X., 1992. Catal. Rev. Sci. Eng., 34(4): 373.
[137] Venugopalan, M. and Veprek, S., 1983. In Topics in Current Chemistry: Plasma Chemistry IV, Vol.
107. Springer, Berlin, p. 1.
[138] Venugopalan, M., Roychowdhury, U.K., Chan, K. and Pool, M.L., 1980. In Current Chemistry:
Plasma Chemisrty II, Vol. 90, Springer, Berlin, p. 1.
[139] Gauvin, W.H., 1990. Novel reactors for plasma applications, Chem. Eng Sci., 45: 2453-2460.
[140] Polak, L., 1971. In: M. Venugopalan (Ed.) Reactions Under Plasma Conditions, Vol. 2. Wiley
Interscience, New York, p. 141.
[-141] Vursel, F. and Plak, L., 1971. In: M. Venugopalan (Ed.) Reactions Under Plasma Conditions, Vol. 2,
Wiley Interscience, New York, p. 299.
[142] Goldberger, W.M. and Oxley, J.H., 1963. A.I.Ch.E.J., 9: 778.
[143] Skrivan, J.F. and Jaskowsky, W.V., 1965. Ind. Eng. Chem. Proc. Des. Dev., 4(4): 371.
[144] Baddour, R.F. and Timmins, R.S. (Eds). The application of Plasmas to Chemical Processing, MIT
press, Cambridge, MA.
1-145] Freeman, M.P., and Skrivan, J.F., 1962. A.I.Ch.E.J., 8: 450.
1-146] Baddour, R.F., and Bronfini, B.R., 1965. Ind. Eng. Proc. Des. Dev., 4(2): 162.
1-147] Bell, A.T. and Kwong, K., 1972. A.I.Ch.E.J., 18: 990.
[148] Brown, L.C. and Bell, A.T., 1974. Ind. Eng. Chem. Fund., 13(3): 203.
[149] Bagautdinov, A.Z. Zhivotov, V.K., Kalachev, I.A., et al., 1991. Soy. Phys. Tech. Phys. 36(4): 488.
[150] Bagautdinov, A.Z., Zhivotov, V.K., Kalachev, I.A., et al., 1989. Topics in Atomic Science and
technology, Atomic Hydrogen Power Series, No. 3, Moscow, p. 56.
[151] Azizov, R.I., Fridman, A.A. and Zhivotov, V.K. et al., 1981. In: B. Waldie and G.A. Farel (Eds.),
Proceedings of fifth International Symposium on Plasma Chemistry, No. 2, Herriot-Watt University, Edinburgh, UK, p. 774.
[152] Rusanov, V.D., Fridman, A.A. and Macheret, S.O., 1985. Dokl. Akad. Nauk SSSR, 283(3): 590.
[153] Azizov, R.I., Vakar, A.K., Zhivotov, V.K. et al., 1985. Sov. Phys. Tech. Phys., 30: 44.
1-154] Traus, I. and Suhr, H., 1992. Plasma Chem. and Plasma Proc., 12(3): 275.
[155] Traus, I., Suhr, H. and Harry, J.E., 1993. Plasma Chem. and Plasma Proc., 13(1): 77.
[156] Helfritch, D.J., 1992. Proceedings of IEEE Industry Applications Society Annual Meeting, 28
Sept.-4 Oct. 1991, 756.
[157] Nicholas, J.E., Amodio, C.A. and Baker, M.J., 1979. J. Chem. Soc. Faraday Trans., I75: 1868.
1-158] Maruzen Oil Company Ltd., 1965. Jpn patent 14,413.
[159] Givotov, V.K., Fridman, A.A., Krotov, M.F., Kreshenninikor, E.G., Patrushev, B.I., Rusanov, V.D.
and Sholin, G.V., 1981. Int. J. Hydrogen Energy, 6: 441.
[160] Belousov, I.G., Legasov, V.A. and Rusanov, V.D., 1980. Int. J. Hydrogen Energy 5: 1.
[161] Rubtsova, E.A., Eremin, E.N. and Maltsev, A.N., 1966. Russ. J. Phys. Chem. 40: 1671.
1-162] Suib, L.S. and Zerger, R.P., 1993. J. Catal, 139: 383.
[163] Laflamme, C.B., Jurewicz, J.W., Gravelle, D.V. and Boulos, M.I., 1990. Chem. Eng. Sci. 45: 2483.

Вам также может понравиться