You are on page 1of 12

2800 Ind. Eng. Chem. Res.

1993, 32, 2800-2811

Development of a Zero-Emissions Sulfur-Recovery Process. 1.


Thermochemistry and Reaction Kinetics of Mixtures of H2S and CO2 at
High Temperature
Gavin P. Towlert and Scott Lynn'
Department of Chemical Engineering, University of California at Berkeley, Berkeley, California 94720

When hydrogen sulfide is heated above 600 O C in the presence of carbon dioxide, the conversion
of H2S to elemental sulfur is greater than when hydrogen sulfide is heated alone. Formation of
elemental sulfur is favored by high temperature, low pressure, and low water content in the gas.
The rate-limiting step is the thermal dissociation of H2S. The hydrogen then equilibrates rapidly
with C02, forming CO and H2O via the water-gas-shift reaction. The equilibrium of H2S dissociation
is therefore shifted t o favor the formation of elemental sulfur. The main byproduct is COS, which
is formed by a reaction between C02 and H2S that is analogous to the water-gas-shift reaction. A
quench rate of 1000 "C/s or greater is sufficient to prevent loss of elemental sulfur by back-reaction
or reaction to COS during cooling. Formation of small amounts of SO2 and CS2 is thermodynamically
feasible but has not been observed. Molybdenum disulfide is the best catalyst for H2S dissociation
of those discussed in the literature. A process based on this chemistry has significant advantages
over the Claus process in that it need not produce any tail gas, it allows recovery of the chemical
(or fuel) value of the hydrogen from the H2S, and it requires much less stringent process control.

Introduction 99.8% (West, 1984). This technology allows economic


recovery of sulfur; however, it is very costly because of the
The recovery of sulfur from fossil fuels is of major large number of processing steps needed to prevent sulfur
importance in the chemical and energy industries. Sulfur emissions, and it does not recover the chemical or fuel
must be removed from fuels in order to comply with value of the hydrogen from the H2S.
legislation designed to prevent environmental damage due An alternative route for sulfur recovery suggested by
to acid precipitation. Sulfur must also be removed from Raymont (1975) was to decompose H2S either thermally
petrochemical feedstocks to prevent degradation of cat- or catalytically, and hence recover the hydrogen as well
alysts used in downstream processing. If this sulfur can as the sulfur:
be recovered in a usable form then its sale will partially
offset the cost of its removal from the fossil fuel. The H2S e H2 + (1/2)S2 (2)
amount of sulfur recovered from fossil fuels accounted for
roughly 62 % of all sulfur consumed in the United States This route received much attention in the 1970's and early
in 1991 (Chem. Eng. News, 1992). 1980's, but proved uneconomic due to the high temper-
Most industrial sulfur recovery is carried out using atures required to achieve significant conversion. At-
variants of the Claus process, which is based on the partial tempts to improve the conversion by continuous separation
oxidation of H2S by S02: of the products were attempted, but these also proved
uneconomic due to the high costs associated with sepa-
2H2S+ SO2 s 3 s + 2H20 (1) rating H2 from H2S at high temperatures (Fukuda et al.,
To provide the SO2 for this reaction, a part of the H2S is 1978). Several processes were designed using reaction 2
burned with added air, which leads to the problems (Banderman and Harder, 1982; Fukuda et al., 1978), but
described below. The gas mixture then passes through none was developed commercially.
several stages of catalytic conversion, with condensation The chemistry which led to the present work was
and removal of product sulfur between stages. Because observed experimentally while experiments on the de-
of the introduction of air to the process, there is a con- sulfurization of coal gas using limestone were carried out.
siderable amount of nitrogen flowing through the Claus These experiments involved heating limestone in an
plant (>60% of the gas stream at any point downstream atmosphere containing 96 mol% C02, 4% H2, and 1%
of the combustor). This inert material must be removed H2S. The reaction-gas mixture entered the outer tube of
as tail gas; however, after three or four conversion stages a reactor that extended into a furnace and was withdrawn
the tail gas still contains some sulfur-containing species, through a smaller, concentric inner tube. The difference
typically 2000-3000 ppm of H S plus S02. The tail gas in diameters between these tubes was such that the
must therefore be sent to a treatment plant to remove residence time in the inner tube was only one-thirtieth
these contaminants to an acceptable level. A number of the residence time in the outer tube. The gases exiting
tail-gas processes are reviewed by West (1984). Of these, the reactor were consequently quenched very rapidly by
the process most widely used industrially is the SCOT countercurrent heat exchange with the cooler incoming
process; however, a SCOT unit may cost as much as the gases. More details of the experimental design are given
Claw plant itself (West, 1984). If a sulfur-removal process in Towler (1992). When these experiments were carried
is run in conjunction witha Claus plant and tail-gas cleanup out at a temperature of 800 OC, it was observed that a
process, then overall recovery of sulfur may be as high as yellow deposit, which was found to be elemental sulfur,
formed inside the inner tube in the region where the tubes
+ Current address: Centre for ProcessIntegration, Department exited the furnace.
of Chemical Engineering, U.M.I.S.T., P.O.Box 88, Manchester The reaction by which sulfur formation occurs was
M60 lQD, United Kingdom. initially postulated as
0888-5885/93/2632-2800$04.00/0 0 1993 American Chemical Society
Ind. Eng. Chem. Res., Vol. 32, No. 11, 1993 2801
C02 + H2S CO + H20 + (1/2)S2 (3) gation of sulfur conversion in flames and found the primary
mechanism to be second order. It should be noted,
which can be seen to be the sum of reaction 2, Le., H2S however, that their experimental temperatures were
decompoeition, and the reverse of the water-gas-shift significantly higher than those of interest to this study.
reaction: Kaloidas and Papayannakos (1989) investigated the
CO + H20 F=? CO, + H, (4) noncatalytic thermal decomposition of H2S in the tem-
perature range 60+860 OC and pressure range 1.3-3.0 atm.
Reaction 3 is similar in form to the Claus reaction in that They developed a kinetic model based upon a free-radical
it achieves partial oxidation of H2S and could therefore mechanism with the splitting of H2S into free-radical
form the basis for a potential sulfur-recovery process. intermediates as the rate-limiting step. The model gave
Furthermore, the byproduct of reaction 3 is carbon good agreement with their experimental results, and they
monoxide, which can be reacted with steam to generate found the rate of decomposition to be given by
hydrogen, thereby effectively recovering the H2 from the
H2S. Initial stages in the development of such a process = kIPH2S (8)
are described below, and a full description of the process
is given in the second part of this paper (Towler and Lynn, where k l = 784.1 exp (-23600/T) (mol/(cm3.s-atm))and
1993). However, we cannot claim to be the first to have PH~S is the partial pressure of H2S in atm. They dem-
reported this reaction, as the same chemistry was patented onstrated clearly that the rate of thermal decomposition
by Bowman (1991) and used as the basis for a process was first-order in PH~S under the conditions observed and
which will be discussed below. We were unaware of cited numerous references confirming the presence of H',
Bowman's patent until the work reported here was nearly HS', and S' radicals in gaseous systems containing H2S
completed. at high temperatures (Bradley and Dobson, 1967a,b;Levy
and Merryman, 1965; Merryman and Levy, 1967, 1972;
Previous Work Norrish and Zeelenberg,1957). They also determined that
alumina (A12031 does not catalyze the decomposition
A great deal has been written on the oxidation of H2S reaction.
by SO2 as required in the Claus process. There has also Several catalysts for H2S decomposition have been
been much research into the thermal decomposition of described; in particular, molybdenum disulfide, MoS2, was
H2S as a source of hydrogen, including some efforts that identified at an early stage (Katsumoto et al., 1973).
tried to react the H2S with CO to form COS and H2 (see Fukuda et al. (1978) described the kinetics of H2S de-
below). Very little research has been reported addressing composition over molybdenum disulfide and also deter-
the partial oxidation of H2S with C02. mined that tungsten disulfide is a slightly less effective
As was noted above, North American interest in the catalyst than molybdenum disulfide, whereas NiS is much
thermal decomposition of H2S began with a paper by less effective due to the formation of NiS2. Chivers et al.
Raymont (1975), although it had received some attention (1980) performed similar testa and found that Cr2S3 gave
in Japan prior to that date (Kotera, 1976). The main focus similar performance to WS2, while FeS, COS,and a range
of Raymont's work (and much of that which followed)was of copper sulfides (CU~S, CugS5, and CuS) were not effective
on H2S decomposition as a source of hydrogen, rather than as catalysts. Chivers et al. also found that MoS2 was the
as a means of S recovery. Raymont found the decom- most effective catalyst above 600 "C, but WS2 and Cr2S3
position of H2S (reaction 2) to be thermodynamically un- were more effective below 600 "C. Possible mechanisms
favorable at temperatures below 1800 K. His kinetic for H2S decomposition over MoS2 are described by Sugioka
studies showed that the reaction proceeds rapidly to and Aomura (1984) and Katsumoto et al. (1973).
equilibrium without requiring catalyst at temperatures Chivers and Lau (1985) reported catalytic activity for
above 1250 K, though he did not state the residence time Li2S but found that sodium and potassium sulfides (Na2S,
of the reactor in which these experiments were carried K2S) and polysulfides (Na2S2, K2S2, Nan&, K2S3, Na&,
out. Below this temperature catalysis is required to achieve and K2S4) were not catalytically active; however, these
a satisfactorily rapid rate; however, Raymont did not sulfides react with the gas, forming amorphous mixed
identify a particular catalyst. Raymont also correctly polysulfideproducts. Chivers and Lau (1987)investigated
observed that the equilibrium conversion of H2S could be vanadium sulfide, v2s3, and mixed sulfide systems (v2s3/
enhanced by combining the decomposition reaction with FeS, V2S3/CugS5 and V2SdZnS) in both flow and thermal-
a more favorable reaction. Because of his emphasis on H2 diffusion-column reactors a t temperatures between 400
recovery, he chose the reaction and 800 OC. They suggested that v2s3 and V2S3/CugS6
2CO + s2 * 2cos (5) both perform better than MoS2 in flow reactors, with v2s3/
CugS5 being particularly effective at high temperatures.
giving the overall reaction Kotera (1976) investigated the decomposition route via
reaction 6, but found this pathway to be complicated by
+
H2S CO * H2 + COS (6) the reaction
The COS could then be converted to S2 via
2cos * CO, + CS, (9)
*
2 c 0 s + so2 2Co2 + (3/2)s2 (7)
He therefore switched his attention to the thermal
the SO2 being produced in a manner similar to the Claus decomposition route. Interest in the COS route has
process. He suggested that reaction 6 be carried out under recently been revived by Gangwal et al. (1991), who
conditions that allow continuous separation of Ha from proposed using coal gas to reduce SO2 to elemental sulfur
COS, i e . , in a reactor with metal-alloy-membrane walls. via
Since this technology was not available, attention shifted
to continuous removal of sulfur instead. 2co + so2F? 2CO2+ (l/n)S, (10)
Roth et al. (1982)investigated the mechanism of thermal
decomposition of H2S a t low concentrations (<200 ppm) and suggested that coal gas would also react with COS
and high temperatures (>1965 K) as part of an investi- and H2S to form elemental sulfur. The maximum tem-
2802 Ind. Eng. Chem. Res., Vol. 32, No. 11, 1993

perature they investigated was 650"C, and they suggested conversion attained to be strongly influenced by flow rate,
a process pressure of 20 atm. Their experiments showed indicating that this reaction can be quenched.
that a high conversion of HzS to elemental sulfur could be It should be clear that for the HzS/COZsystem at high
obtained in a system containing coal gas and SOz;however, temperatures we are concerned with the interplay between
they were not aware that the H2S must be undergoing a large number of possible gas-phasereactions. For process
oxidation by COZ as well as SO2 in this system. The design purposes we wish to be able to predict the reactor
pressure selected was too high for formation of elemental size required to achieve a given conversion of H2S. We
sulfur to be strongly favored. must also understand the mechanism by which the main
Fellmuth et al. (1987) examined the reaction of H2S byproduct, COS, is formed if we are to design a satisfactory
and C02. They were, however, chiefly interested in zeolite quench. Although most of the reactions in the H2S/C02
deactivation and did not extend their studies above 300 system have been analyzed in isolation there is little
"C. The main reaction they observed was therefore information in the temperature range of interest, and no
report of the kinetics of H2S decomposition in the presence
H2S + CO, * COS + H,O (11) of C02 was found in the literature.
which they found to be catalyzed by basic zeolites. This
Thermochemical Analysis
reaction can be seen to be analogous to the water-gas-shift
reaction and is important as a mechanism for COS A series of calculations was performed to evaluate the
formation. The equilibriumconstant, KW,for this reaction equilibrium properties of mixtures of carbon dioxide and
is fitted over the temperature range 450-1100 "C by the hydrogen sulfide at high temperatures. These calculations
expression were used to confirm the experimental findings described
above and to determine suitable operating conditions for
= 0.4347 exp(-2917/T)
KsBB (12) a sulfur-recovery process based on this chemistry.
The calculations were carried out using a modified
The endothermic nature of the reaction and weak version of a computer simulation developed by Whitney
temperature dependence of the equilibrium constant give et al. (1987),whichcalculates the equilibrium composition
further evidence to support the comparison with the water- of a reaction mixture, given a feed stream cornposition,
gas-shift reaction, and we might therefore expect this together with a temperature and pressure. The calculation
reaction to have equally rapid kinetics. Unfortunately is performed by first selecting a set of independent
most of the research on this reaction has been related to chemical reactions, then determining material balances
the development of catalytic processes for COS hydrolysis for all elements present, and finally determining the
prior to gas scrubbing and was thus carried out at low composition of the mixture having the minimum Gibbs
temperatures, typically below 300 "C. Although there is free energy by using a robust, multidimensional Newton-
much discussion of the competitive absorption of different Raphson iteration procedure. For the high temperatures
species and the nature of the catalytic sites, only one paper (>600 "C) and low pressures ( 4 0 bar) of interest to this
was found to give an activation energy for the reaction. study, it was reasonableto assume that all gas-phase species
George (1974) found the activation energy to be 12 kcaU behave ideally; therefore fugacity corrections were not
mol for COS hydrolysis on cobalt molybdate at 230 "C. made. Thermochemicaldata were taken from the JANAF
The Arrhenius plot he presented showed, however, that Thermochemical Tables (Chase et al., 1985).
this energy was based on a line drawn through only four Following the findings of Kaloidas and Papayannakos
data points. The line fitted the three points at lower (19871, S2 was assumed to be the only elemental sulfur
temperatures reasonably well, but substantially under- speciespresent. In actuality, small amounts of other sulfur
predicted the point at the highest temperature. There is allotropes will also exist; therefore the calculations may
therefore good reason to believe that George may have slightly underpredict the fraction of elemental sulfur in
overestimated the activation energy, especially when we the gas at equilibrium. This is a safe-side approximation
consider that we are operating several hundred degrees that considerably shortens calculation time and obviates
above the temperatures used in his study. In any case, the selection of a thermodynamic database for the other
the value 12 kcal/mol is low for an activation energy, elemental sulfur species. The sensitivity of the results to
confirming the similarity between this and the water-gas- this assumption is discussed below.
shift reaction and giving us good reason to believe this to In analyzing the results of these calculations the most
be an important step in COS formation. important informationis the distribution of sulfur between
An alternative route by which COS may be formed in the different sulfur-containing species. The results are
the H2S/C02 system is reaction between CO and S2 via therefore presented as graphs showing the fractional
reaction 5, which is the reverse of the thermal dissociation distribution of sulfur, i.e., the fraction of the total sulfur
of COS. The kinetics of COS decomposition was studied present as each species on a molar basis. The sulfur-
by Schecker and Wagner (1969) in the temperature range containing species included in the calculations were H2S,
1500-3100 K. They found the rate to be second order and S2,SO2, CS2, COS, and SOS. Results for CS2 and SO8 are
given by not reported, however, as neither ever accounted for more
than 0.002 of the total sulfur present.
dCcos - 61000
dt --lo
14.2
(
exp -- RT )'btdccOs (13)
In discussing the results of the survey it is useful to note
that since we consider eight chemical compounds (Cop,
where CCOSis the concentration of COS (mol/cm3),Cbd CO, Ha, H20, H2S, SZ,, 9 0 2 , and COS) containing four
is the total concentration (ie., COS + inerts) (mol/cm3), elements (C, H, 0, and S), we require four reaction
and R is the ideal gas constant (cal/(mol K)). The results equations to specify the equilibrium composition of the
of Schecker and Wagner (1969) were subsequently con- gas phase. For the purposes of this discussion these will
f i i e d by Chenery et al. (1983). Dokiyaet al. (1978)looked be taken as being the water-gas-shift reaction,
at the reverse of COS decomposition, i.e., the formation CO + H20 Z= C02 + H, (4)
of COS from CO and elemental sulfur. They did not report
a kinetic expression; however, their data showed the the H2S decomposition reaction,
Ind. Eng. Chem. Res., Vol. 32, No. 11, 1993 2803
1.0, I 0.8 0.10

0.08

B
VI
0.06
5 0.6 - In
e

1
Q
VI I.
c 2
004 d"
v)
0.4 e
Q

002 89

". , 0.00
0 2 4 6 8 10 12

nn I
"." I
900 1000 1100 1200 1300 1400
Temperature (K)
Figure 1. Equilibrium distribution of sulfur-containing species aa
a function of temperature for H2S only (1atm). The equilibrium yield of S2 increases with increasing
temperature because of the endothermic nature of HzS
decomposition (reaction 2). Increasing H2S decomposition
causes an increase in the amount of HzO present in the
gas, which helps drive the reverse of the Claus reaction
(1);therefore the equilibrium fraction of SO2also increases
with temperature. This imposes an important limit on
the operating conditions for the sulfur-recovery process
as formation of SO2 is detrimental to process operation
and should be avoided whenever possible. The formation
of COS is favored by low temperatures, since reaction 5
is exothermic as written. An important consequence of
this thermal behavior is that if a mixture of C02 and H2S
is brought to equilibrium at high temperature and then
cooled slowly, the elemental sulfur is able to back-react
to H2S and COS, thereby losing part or all of the reaction
- /
yield. This can be prevented by quenching, Le., cooling
0.0 0 00 the gas rapidly to a temperature a t which the sulfur-
900 1000 1100 1200 1300 1400
Temperature (K)
consuming reactions are very slow compared to processing
timescales. Quenchingto about 600"C a t a rate of roughly
Figure 2. Effect of temperature on sulfur distribution for a feed lo00 K/s was found experimentally to be sufficient.
containing 50% HzS, 50% COZ at 1 atm.
Effect of Pressure. The water-gas-shift reaction (4)
is equimolar, and therefore its equilibrium composition is
H2S ~i H2 + (1/2)S, (2) independent of pressure. All of the elemental-sulfur-
the Claus reaction, forming reactions lead to a net increase in the number of
moles; consequently the amount of S present as elemental
2H2S + SO, e (3/2)S2+ 2H20 (1) sulfur decreases with increasing pressure as can be seen
and the formation of COS from CO and S2, in Figure 3. An interesting, and perhaps unexpected, result
is that the equilibrium fraction of SO2 also decreases. If
2co + s, Fi 2 c o s (5) we consider the ratio (number of moles of products)/
It should be noted that these reactions merely form an (number of moles of reagents) for the S2-formingreactions,
independent set for equilibrium calculations and are not we find that for reactions 2 and 5 the ratio is 1.5, whereas
necessarilyindicative of the reaction mechanism. Reaction for the Claus reaction it is only 1.17. COS formation and
mechanism and kinetics are discussed below. H2S decomposition are therefore much more sensitive to
Effect of Temperature. Figure 1shows the equilib- pressure than the Claus reaction, and consequently when
rium sulfur distribution given by the decomposition of the pressure is increased S2 is more likely to be converted
H2S as a function of temperature at 1atm in the absence to H2S or COS via these routes than to SO2 via the Claw
of C02. Figure 2 shows the distribution for an initial feed reaction. Obviously this is somewhat dependent on the
of 50% H2S, 50% C02. Comparing the figures, it can be initial composition of the gas (particularly the amount of
seen that the fraction of S present as elemental sulfur at HzO present) since the water-gas-shift equilibrium is
equilibrium has been enhanced by a factor of roughly 2 pressure-independent.
a t all temperatures. This significant increase in conversion Effect of the COdH2S Ratio in the Feed. Figure 4
suggests that a process based on this chemistry would be shows the effect of increasing the initial mole fraction of
more successfulthan a process based on the decomposition H B in C02 on the equilibrium distribution of sulfur
of H2S in the absence of COa and demonstrates the compounds. At very low H2S/C02 ratios, SO2 is the most
effectiveness of the water-gas-shift reaction for shifting favored product; however, the fraction of the S present as
the equilibrium of H2S decomposition. SO2 falls rapidly as the initial HzS concentration is
2804 Ind. Eng. Chem. Res., Vol. 32, No. 11, 1993
, 0.20

Feed sat w i t h H,O


a t 4OoC

i\\
0.15

Yield of S2 per mole of H,S


0 5 i i in Reactor Feed
0.10 04

0.3

0.05

0.2

0.0 n ""
"."I 0.1
0 10 20 30 40 50
Inltid X H,S In COz

Figure 4. Effect of the initial H&C02 ratio on the equilibrium 0.0 1


I , , 1 I / , I l
o io 20 30 40 so 60 70 a0 90 loo
distribution of sulfur-containing species at 1 atm, 900 OC.
Initial Z H,S in CO, (900°C, 1 bar)

Figure6. Yield of elemental sulfur as a function of feed concentration


,E
0
E
c 0.6
I
~ 10.06
(900 OC,1 bar).

Table I. Effect of Including Sulfur Allotropes


v) fraction of fraction of
P 8 Sulfur-Containing Sulfur-Containing
Species Species
species S2only S2, ...,Se species S2only S2, ...,SS
S2 0.324 0.318 Se 0 O.ooOo24
ss 0 0.00649 H2S 0.614 0.613
s
4 0 0.000303 SO2 0.000811 0.000798
s5 0 O.ooOo5 COS 0.0607 0.0604
S6 0 0.000123 CS2 0.0000087 0.0000087

-
SI 0 0.000107

In particular, the equilibrium fraction of S2 decreases only


from 0.32 to 0.29, a change of under 10%.
Increasing the water content helps drive the reverse
Clam reaction; thus the SO2 fraction increases, though
not strongly. The effect of increasing the feed H20 content
on the water-gas-shift reaction is to drive the reaction
toward C02 and H2, thereby tending to reduce H2S
increased and the mixture becomes more reducing in decomposition. This also reduces the CO concentration
nature. and hence favors the reverse of reaction 5, leading to COS
The fraction of sulfur present as S2 passes through a decomposition and causing the COS fraction to decrease
maximum at around 6% H2S in C02, since increasing the as the H2O fraction increases. The increased COS de-
amount of H2S present not only makes the mixture less composition to some extent counteracts the reduced H2S
oxidizing in nature, but also lowers the S2 fraction by decomposition, giving rise to the observed weak depen-
dilution with H2S. A similar effect is observed for COS. dence of S2 fraction on H20 content.
It should be noted that the results for COS are plotted on Allowance for Sulfur Allotropes. As was stated
an expanded scale (4X magnification) and that the COS above, elemental sulfur can exist as a number of allotropes
fraction is therefore much less sensitive to the feed S2, ...,SSunder the conditions of this study. To confirm
composition than the S2 fraction. the results of Kaloidas and Papayannakos (1987) cited
Effect of H20. For processing reasons, it is likely that above, the base calculation (1 atm, 900 "C, 5050 H2S:
any gas stream containing H2S and CO2 would also contain CO2) was repeated with the sulfur speciesS3,...,Sa included.
a small amount of water vapor. (This arises from the use The results are shown in Table I. The fraction of S present
of absorber/stripper loops for acid-gas cleaning-the acid as elemental sulfur was 0.3240 in the case where only S2
gas stream usually leaves the stripper saturated with water was considered and 0.3256 in the case where all S allotropes
at condenser temperature.) It is therefore important to were allowed. This confirms that the effect of including
understand the effect of the initial HzO concentration on the compounds S3,..., Se is negligible.
the equilibrium distribution of sulfur-containing species. Reactor Yield. In considering the design of a sulfur-
Figure 5 shows this variation at 900 OC and 1atm, for a recovery process the optimum reactor feed is determined
feed containing 1:l C02 and HzS, as the water fraction is by several factors, which are somewhat dependent upon
increased from 0 to 109%. At l-atm pressure, 10% initial the process configuration, and will therefore be discussed
water concentration corresponds to the feed gas being in greater detail in the second part of this paper (Towler
saturated with water at 46 OC. Perhaps the most salient and Lynn, 1993). To a first approximation, however, we
feature of Figure 5 is that none of the species is strongly can consider the best feed composition to be that which
affected by increasing the H2O fraction over this range. maximizes the number of moles of elemental sulfur
Ind. Eng. Chem. Res., Vol. 32,No. 11,1993 2806
Vent

U U I
T
hookout
TUlk

-H N.08AbiorpUon
1 ScNbbclr

I I I I I I L

Pipellno
I v
I
11 u u u u u
C02 NP E2S E2 CO
SRI 8610 Gal Chromatoirsph
8
Sulfur
Condemer
Figure 7. Experimental apparatus.
Inner tube o.d, 0 3 m m
td 3.9 mm QwtrT b r m o w l l
C u Inlet
II II h
I
I
Outer Tuba 1.d. 21.8 mm
a.d 7 mm

T
. I

cu Cu Sampling/
outlet
I . .
smep G u Port

kction
Figure 8. Reactor dimensions.

produced per mole fed to the reactor, i.e., the yield per of C02. As noted above, the optimum reactor feed com-
reactor pass. We can determine the yield at chemical position may depend upon other factors as well as reactor
equilibrium quite easily, by multiplying the fraction of S yield, and this will be discussed in the second part of this
present as elemental sulfur by the fraction of HzS in the paper (Towler and Lynn, 1993). It is, however, important
feed. to note that the maximum in the yield curve is in fact
Figure 6 shows the reactor yield at 900 "C, 1 bar, as a rather broad and that high yield can therefore be achieved
function of the concentration of HzS in the feed (drybasis). over a wide range of feed compositions. This has some
The solid lines in Figure 6 are for a dry feed, in which case important implications (discussed below) when a sulfur-
the maximum yield is 0.167 at an initial concentration of recovery process based on this chemistryis compared with
67% HzS in C02. This represents an improvement of 36% a process based on the Claus reaction.
over the yield in the absence of C02. This could raise the
question of whether the presence of C02 is really necessary Experimental Section
to the process; however, it must be remembered that the
water-gas-shift reaction (41, by significantly lowering the Experiments were carried out in the continuous-flow
hydrogen concentration, also minimizes back-reaction apparatus shown in Figure 7. The C02 and H2S (industrial
during the quench. It is therefore much easier to maintain grade, supplied by Mattheson Gas Products, Newark, CA)
a high yield in the presence of C02 than in its absence. were mixed at the desired flowrates using high-precision,
As noted above, the feed will normally contain some low-flow-rate rotameters (Omega Engineering Inc., Stam-
H2O; therefore the calculations were repeated for a feed ford, CT), before being sent to the reactor. The reactor
saturated with water at 40 "C, 1 atm. The results are consisted of two concentric quartz tubes mounted in a
shown as the dashed lines in Figure 6. In this case the high-temperaturefurnace. The gas entered the outer tube
maximum yield of elemental sulfur per mole of feed gas of the reactor, where it was preheated by contact with the
is 0.1452 a t a feed fraction 71% H2S in C02. This is exiting gas and by heating tape around the outside of the
somewhat lower than the dry case, mainly due to dilution: reactor. The gas was then further heated in the furnace
if we correct the results to a dry basis, the yield is 0.157, section of the reactor before exiting through the inner
Le., an improvementof 27.7% over the yield in the absence tube. The difference in the tube diameters caused the
2806 Ind. Eng. Chem. Res., Vol. 32, No. 11, 1993
900 1000

800 900

h
700 800
U
0- 0-
u

3e
0
2
600 e
D
700

c
i Inner tube entrance
located here
F
i
500 600

0 Thermowell temperature
400 V Inner tube temperature 500
0 Thermowell temperature
(repeat) /
300 , I f I , I , 400
-5 0 5 10 15 20 25 30 35 0 1 2 3 4 5 6 7 0
Distanoe Along Reactor Time (8)

Figure 9. Reactor temperature profile (T-t = 800 "C, flow rate = Figure 10. Typical Reactor Temperature-Time Ronie
2.935 mL/s). (T- = 900 T,F = 2.935 mlis)

Figure 10. %icd reactor temperaturetime profiie (T,t = 900 "c,


residence time in the inner tube to be roughly 4% of that flow rate = 2.935 mL/s).
in the outer tube, which, together with the temperature
profile obtained in the reactor and judicious selection of Table 11. Exwrimental Conditions and Results
flow rates, ensured that the gas spent long enough at high outlet
temperature to guarantee significant conversion of H2S, temp flowrate feedconcn conv COSmole
before being cooled rapidly as it left the furnace. The (OC) (d/s) ( % HzS in COz) (%) fraction x 103
physical dimensions of the reactor are given in Figure 8. 950 2.935 5 34.4 4.05
These dimensions are of great importance in the modeling 950 2.935 5 35.2 4.43
of the reactor; each dimension is the average of several 950 2.935 10 34.0 2.82
950 2.935 10 34.4 3.99
measurements (taken at different points of the reactor 950 1.468 5 54.3 3.57
where possible). 950 1.468 5 53.1 4.15
The gas leaving the reactor passed through a condenser 950 1.468 10 43.8 4.79
that collected more than 99% of the elemental sulfur 950 1.468 10 44.3 5.16
900 2.935 5 21.5 3.86
formed during the reaction. The gas was then sampled 900 2.935 10 20.0 3.36
using an SRI 8610 Gas Chromatograph (SRI Instruments, 900 2.935 10 22.2 2.99
Torrance, CA), which was used to detect for C02, CO, 900 1.468 5 36.1 3.32
COS, H2S, and S02. This, together with knowledge of the
feed rates, allowed the composition of the gas leaving the Rasults
reactor to be determined.
Because we are concerned with t L e interp.dy between
The temperature profile of the reactor was measured several reactions over a wide range of temperature and
for different heater set-points and gas flow rates. The composition,where substantial changes in the equilibrium
profile was found to vary slightly with flow rate; therefore behavior of the system can occur, the results are best
three standard flow rates were selected and the temper- presented in tabular, rather than graphical, form. The
ature profile was measured for each. A typical temperature conversion of HzS achieved for various experimental
profile for the reactor is shown in Figure 9. The tem- conditions is given in Table 11.
perature profile in the inner tube was found by removing Comparingthe experiments at high flow rate (where we
the condenser and inserting a thermocoupleinto the tube. are furthest from equilibrium conversion),we can see that
The temperature in the outer tube was measured by a the fractional conversion obtained with a feed of 10% H2S
thermocouple in the reactor thermowell and a slight in C02 is roughly the same as that obtained with 5% H2S
correction was made to allow for radiative effects. The in C02. This behavior would be expected if the rate-
uncertainty in the temperatures measured was roughly 3 limiting step were the first-order thermal decomposition
OC. of HzS. At low H2S feed concentrations the water-gas-
Knowledge of the temperature-distance profile and the shift reaction lowers the hydrogen concentration to such
gas flow rate allows calculation of the reactor temperature an extent that back-reactionby recombination of hydrogen
time profile, i.e., the temperature history experienced by and elemental sulfur is negligible. The overall reaction
the gas as it passed through the reador. We can also would thus behave as a first-order irreversible reaction,
calculate the quench rate from the highest temperature for which the conversion in a plug-flow reactor is inde-
attained to any lower temperature. Such calculations were pendent of the feed concentration. We see that this limit
incorporated into the kinetic models described below. A is obtained at both set-point temperatures (but note, how-
typical temperature-time profile is shown in Figure 10. ever, that the gas does not experience a single temperature,
To investigate the effect of quench rate, some experi- but a range or profile, as indicated in Figure 9). Obviously
ments were run with a 3.06-mm quartz rod inserted into as the conversion approaches its equilibrium value this
the inner tube. This reduced the flow area and increased approximation can no longer be true, and we see that the
the gas velocity by a factor of roughly 2.6, allowing the experimentsrun at high temperature and low flow rate do
quench rate to be increased by roughly the same factor. not satisfy this condition as equilibrium effects are more
Ind. Eng. Chem. Res., Vol. 32,No. 11,1993 2807
Table 111. Outlet Compositions Measured for Quenched If this is the mechanism of H2S conversion, then the
and Nonquenched Experiments. rate of change of the H2S partial pressure is given by
feed outlet composition
H2S quench Hz d@H2S)/dt = -klPH# + k&-@S:'2 (14)
fraction rate and
(%) (K/d C02 H2S CO H2O COS S2
where kl should be the first-order rate constant as found
by Kaloidas and Papayannakos (1989) and other symbols
5 990 0.928 0.0218 0.00973 0.0260-000357 0.0112 are as defined in the Nomenclature section.
5 2560 0.930 0.0218 0.00884 0.0248 0.00415 0.0103
10 1040 0.873 0.0524 0.0108 0.0408 0.00479 0.0180 If we use suitable average values for the quantities in
10 2700 0.872 0.0519 0.0113 0.0412 0.00516 0.0180 the second term on the right-hand side of eq 14 and
integrate, we obtain a "partially integrated" algorithm:
In all cases the flow rate was 1.468 mL/s and the maximum
(I

temperature in the profile was 950 "C.


PH2S,n
-- PH,S,n-l exp(-kl,nAtn) +
important for these runs. The error in the reported
conversions can be assessed from the difference between
repeated runs and is typically of the order of 3% ,though
it may be as high as 11%for the experiments with the which may be used to update the H2S partial pressure
lowest conversion. from one interval to the next. The first term in this
The outlet compositions found for the high-quench equation allows for decomposition with no reverse reaction,
experiments (in which a quartz rod was inserted into the and the second term is a correction to allow for the reverse
reactor exit tube) and experiments run under the same reaction. It can be seen that in the limit as n and At,, tend
conditions with lower quench rates are shown in Table to infinity at a constant temperature, the first term goes
111. Table I11 shows that the outlet compositions for the to zero while the second term approaches the equilibrium
high-quench experiments are practically identical to those concentration.
for the corresponding low-quench experiments; i.e., the Having found the number of moles of H2 formed from
increase in quench rate gives no added conversion. (The the H2S decomposition and knowing the number of moles
apparent slight increase in COS formation for both of CO2 fed, the concentrations of Con, CO, H2, and H2O
quenched experiments is negligible compared with the are then found by solving the water-gas-shift reaction
error in the COS measurements, but see below.) This equilibrium equation and three mass balances (C, H, and
confirms that the quench rates used in the bulk of the 01,and the concentration of elemental sulfur is found
experiments (>lo00 "C/s) were high enough to quench from a sulfur mass balance.
completely both the H2S decomposition back-reaction and This basic model gave good agreement with the exper-
the COS-forming reactions. These results are also useful imental data using the published kinetic parameters of
in analyzing the formation of COS, as discussed below. Kaloidas and Papayannakos (1989);however, it is unsat-
isfactory in that it does not describe the formation of the
byproducts COS and SO2. Sulfur dioxide formation was
Kinetic Model accounted for by assuming that the Claus reaction was
always at equilibrium. This is a reasonable approximation
In developing a model for the kinetic behavior of the above 700 "C, but in practice this did not significantly
H2S/C02 system at high temperature, our main goal is to alter the predictions of the model as SO2 was formed at
be able to predict the conversion of H2S to elemental sulfur, very low levels under the conditions studied. Sulfur
i.e.,to identify the rate-limiting step(s) in H2S conversion. dioxide was not detected in any of the noncatalytic
A secondary goal is to predict the formation of carbonyl experiments; therefore the validity of this assumption
sulfide, so that reactor conditions can be arranged to could not be tested and remains a subject for future
minimize formation of this byproduct. Several models investigation.
were tried; however, in the interest of brevity we shall The formation of COS is not so simple to describe. There
concentrate upon the model that gave the best agreement are two likely mechanisms for COS formation, namely
with the data. reactions 5 and 11. The kinetics of these reactions was
The reactor is assumed to be in plug flow and is divided described above, though it should be noted that the work
into sections corresponding to the interkls at which the of George (1974) was carried out under conditions some-
temperature was measured. For each interval the average what removed from those of this study.
temperature can be found, and hence, knowing the feed Since most of the formation of CO and S2 occurs a t the
flow rate and pressure, the time the gas spends in the highest temperatures in the profile, we would expect
interval may be calculated (the flow rate can also be reaction 5 to be significant as a COS-forming mechanism
corrected to allow for the increase in molar flow rate that only at the highest temperatures or during cooling, whereas
occurs on reaction). Given the residence time and tem- reaction 11 would be a reasonable mechanism at lower
perature for each interval we can use a suitable algorithm temperatures when substantial amounts of CO and S2 have
to find the concentration of each species a t the end of the not yet formed. (Indeed, reaction 11plus the reverse of
interval. By treating the intervals as a set of plug-flow reaction 5 would be a second-order alternative mechanism
reactors in series, we eventually obtain a prediction of the for H2S conversion.) In practice, both reactions are
final outlet concentration. important, as will be shown by considering the experi-
The pseudo-first-order-irreversible behavior observed mental results.
a t high flow rates suggested that the rate-limiting step of The COS concentrations detected were greater than the
H2S conversion is the thermal decomposition of H2S, with concentrations that would have been found if COS had
the hydrogen thereby produced reacting with C02 via the been in equilibrium with the other species at the maximum
water-gas-shift reaction. The water-gas-shift reaction is temperature in the profile. As was shown above, COS
known to equilibrate very rapidly at temperatures above formation is favored by lower temperatures. These high
600 "C and may therefore be assumed to be always a t COS concentrations must therefore have been formed
equilibrium. either as the gas entered the reactor or else as it cooled
2808 Ind. Eng. Chem. Res., Vol. 32, No. 11, 1993

Table IV. Comparison of Experiments and Model


Predictions of the Conversion of HsS CO and S2 in Equilibrium
-
Dredictions of model with various assumptions
E,& from 5% variation 5% variation
-2
4
exptl published of E& for of E.& for -4 -
conv data reaction 1 reactions 1 and 8 Scheckcr and Wagner
34.4 35.1 34.6 34.2 -6 -
35.2 35.4 34.8 34.5
34.0 34.3 33.8 33.5
34.4 34.3 33.8 33.5
54.3 58.6 57.9 57.5 Entrance t o Quench
53.1 58.8 58.1 57.7 -a
-10 -- Tube
43.8 55.2 54.6 54.3
44.3 55.2 54.6 54.3
21.5 19.6 19.8 20.0 -12 -
20.0 19.0 19.2 19.2 0 Measured
22.2 19.0 19.3 19.3 -14 - Concentration
36.1 36.9 37.3 37.4
-16 1
upon leaving. If COS were formed during cooling (via 0 10 20 30 40 50
reaction 5), then the outlet concentration would show a Distance From Reactor Entry (cm)
dependence upon the quench rate, i.e.,the residence time Figure 11. Variation of COS concentration withd i c e into reactor
for cooling. It can be seen from Table I11that this is clearly (maximum temperature 950 "C, flow rate 1.468 mL/s, 5% H2S in
not the case; in fact, the COS outlet concentrations may Cod.
even be higher at higher quench rates. Alternatively, COS
may be formed via reaction (11)when the gas first enters of Schecker and Wagner (1969). Clearly neither of these
the furnace. As the gas flows to regions of higher gives an adequate description of COS formation.
temperature, a point would be reached a t which the COS We can improve the model predictions somewhat if we
concentration is greater than its equilibrium value, at which allow constrained variation of the activation energies for
point COS would begin to dissociate via the reverse of reactions 2 and 11. For HzS decomposition this is not
reaction 5 (thoughnot necessarilyvia the reverse of reaction unreasonable, since the experiments of Kaloidas and
11 since COZ and HzS could still be present at greater- Papayannakos (1989) were performed outside our tem-
than-equilibrium concentrations). By this hypothesis the perature range and must necessarily contain some exper-
outlet COS fraction would not be greatly affected by the imental error; furthermore, Fukuda et al. (1978) found an
quench rate (as long as this is rapid enough to prevent activation energy of 42 kcal/mol, which is somewhat lower
significant reaction when the gas reaches lower temper- than the 43.7 kcal/mol found by Kaloidas and Papayan-
atures) and might even increase at higher quench rates as nakos. If we use an activation energy 5 % lower than that
the time for COS decomposition a t high temperature is of Kaloidas and Papayannakos (41.5 kcal/mol), we obtain
reduced. the third column of Table IV. The preexponential constant
This mechanism not only gives a better explanation of in this case is 2.4 X 10 s-l. The root-mean-square error
the data, but is also more physically reasonable. We know is reduced to 11.3%,which is of the order of experimental
that reaction 11 is somewhat similar to the water-gas- error. The COS concentration predictions are not affected
shift reaction, (4), (see above), and, given that it has such by this change, as COS formation occurs in the low-
alow activation energy,it would be unreasonable to neglect temperature region of the entry tube, Le., before the HzS
its effect at temperatures above 800 "C. Both reactions conversion is significant. We can, however, improve the
5 and 11were therefore incorporated into the model, again prediction of COS concentration somewhat by allowing
using published kinetic parameters, and using an algorithm similar variation of the activation energy of reaction 11.
similar to eq 15 to upgrade the COS concentration from Again this is not unreasonable, due to the rather dubious
one interval to the next. Equation 15 was also modified nature of the graph presented by George (1974). We should
to account for the presence of reaction 11. Mathematical also note that although George was studying the reaction
details of the algorithm used are given in Towler (1992). in the presence of catalyst, his experiments were carried
Table IV shows the experimental results and model out 600 "C below ours, so extrapolation may not be
predictions obtained using the published values of the justified. If the activation energy for COS hydrolysis is
activation energies (Kaloidas and Papayannakos (1989) taken as 8 kcal/mol, we obtain the results of the fourth
for HzS decomposition, and George (1974) for reaction column of Table IV, which predict both conversion and
11). The preexponential constants used were adjusted to COS formation to within experimental error (for details
fit the data over the temperature range observed and were of the COS predictions see Towler (1992)). Obviously the
6.0 X lo6 s-1 for HzS decomposition and 2.4 X lo6 cm3 model predictions could be further improved by including
mole-1 s-l for reaction 11. The root-mean-square error more reactions in the model; however, this merely increases
between the experimental conversions and those predicted the number of adjustable parameters available and is
by the model is about 12%, and the error between the therefore rejected. The model thus obtained gives an
predicted and observed COS concentrations is roughly adequate description of the experimental data over the
the same as the error in the COS measurements. range of conditions studied and can be used for reactor
The predicted COS concentration for a typical exper- design. Some limitations of this model will be discussed
iment is plotted against temperature in Figure 11, from below.
which we see that the model gives the right qualitative
behavior. Figure 11also shows the concentration of COS Effect of Catalysis
that would be found if COS and SZwere in equilibrium
and the concentration that would be present if COS were Having identified the rate-limiting reaction for the HzSl
only formed from CO and SZas predicted by the kinetics COZsystem to be the thermal dissociationof HzS, it follows
Ind. Eng. Chem. Res., Vol. 32, No. 11,1993 2809
that one would improve the overall reactor performance H2S conversion and the formation of the main reaction
by supplying a catalyst for this reaction. This is of great byproduct, COS. On a more fundamental level, however,
importance in the development of a process based on this the model is somewhat less satisfactory.
chemistry, as use of catalysis considerably reduces the Firstly, undoubtedly the greatest source of error in the
reactor volume required. Even with catalysis the reactor model is also, unfortunately, the hardest to quantify. The
is likely to be the most expensivecapital item in the process model assumed the reactor to consist of two plug-flow
designed (Towler, 1992). Catalysts for the thermal de- reactors in series, each with a different residence time,
composition of H2S were discussed above; in particular, corresponding to the inner and outer flow volumes. In
molybdenum disulfide (MoS2) is well-known as a catalyst practice the Reynolds numbers used experimentally were
for H2S decompositionat high temperatures. A short series 2 or 4 for the outer tube and 48 or 96 for the inner tube,
of experiments was therefore performed to demonstrate so the flow was laminar in all cases. For a first-order
that MoSz would act as a catalyst for the H2S/C02 system. isothermal reaction the conversion in a laminar-flow
It should be stressed that this was only a preliminary tubular reactor is somewhat less than that in a plug-flow
investigation to demonstrate feasibility and that a full
analysis of the catalysis of this system would be a major reactor, mainly due to bypassing by material that flows
piece of work in its own right and was beyond the scope through the reactor at velocities faster than the average
of this study. velocity. If the velocity profile is known (e.g., for a simple
The apparatus described above was modified by packing tube), the conversion can be found mathematically (this
the reactor with MoS2 catalyst supported on quartz wool. involves use of the exponential integral function for a first-
Initially, experiments were performed with the reactor order isothermal reaction). In our case, however, the outer
packed with quartz wool in the absence of catalyst; the tube contains two nonconcentric,nonparallelsmaller tubes
quartz wool alone was determined to have a negligible (the inner tube and thermowell) and has a marked
effect on reaction kinetics. The experiments were then longitudinal (and possibly also a slight radial) temperature
repeated with the same mass of quartz wool onto which gradient,all of which make evaluation of the velocity profile
MoS2 powder had been loaded. The conversion increased a complex computational problem in fluid dynamics.
significantly at all temperatures studied (up to 900 "C), Furthermore, the high molecular diffusivities at the high
and no evidence of catalyst deactivation was observed over temperatures obtaining in the reactor cause some radial
52 h of operation. This does not eliminate the possibility mixing and tend to offset the dispersion of residence times
that catalyst deactivation is occurring;however, it suggests caused by the velocity profile. Exact characterization of
that if it occurs at all then it is a rather slow process. In the laminar-flowreactor is therefore difficult and the plug-
several cases the distribution of sulfur-containing species flow assumption was thus necessary, but this may impose
was close to that expected for chemical equilibrium at the a large error on the model predictions. Experiments that
highest temperature in the profile, with the exception that were run with the outer tube entirely packed with quartz
the fraction of SO2 was lower than expected. It is not wool (as preparation for the catalysis experiments) gave
clear whether SO2 was not formed at high temperature or somewhat higher conversionsthan experiments performed
whether it was lost by back-reaction during cooling. in the absence of this packing. It was not clear, however,
The experimental conditions were not sufficiently well whether the quartz wool was promoting radial mixing and
defied to allow detailed modeling of the experiments with thus bringing the reactor closer to plug flow or whether
catalyst (Towler, 1992). It would be possible to extend channelingeffects (the quartz wool could not be distributed
the model described above by using the kinetic expression evenly) were causing changes in the temperature profile
developed for H2S thermal decomposition over MoS2 by experienced by the gas.
Fukuda et al. (19781,but this would not account for the Secondly, at the high temperatures in question it is not
effect of catalysison reaction 11 and hence COS formation. reasonable to suppose that the net rate of reaction is
Since the catalytic mechanism is known to involve affected only by the collisions of molecular species. In
formation of H*, HS*, and S* free radicals (Katsumoto practice there are likely to be many short-lived free radical
et al., 1973),it is unreasonable to expect the catalyst to species present at low concentrations (for example, H*,
affect only the H2S-decomposition reaction. We may HS*, S*,OH*, SO*, etc.) which probably play an integral
however, use the kinetics of Fukuda et al. (1978)as a first- part in the reaction mechanism(s)via a series of initiation,
order approximation for process design if we assume the chain propagation, and termination reactions, much the
gas reaches the equilibrium concentration (a reasonable
assumption in the light of the experimental results) and same as is found in the analysis of combustion systems.
the catalyst loading is the same as that used experimentally The presence of these species was referred to by Kaloidas
by Fukuda et al. For a fundamental understanding of the and Papayannakos (1989)and earlier workers, as noted
catalytic behavior a more complete analysis is necessary, above. The success of the experiments with a catalyst
which should include not only the effect of the catalyst on that is known to promote the formation of HS* radicals
the other reactions in the system, but also the preparation is also good evidence of free-radical initiation as the rate-
of catalysts and optimization of properties such as surface limiting step; however, HO* radicals are also known to be
loading, etc. This was beyond the scope of this work; of great importance in the water-gas-shiftreaction kinetics
however, the beneficial effects of catalysis on the H2S/ and we did not feel confident enough in the experimental
COz system were clearly demonstrated. More details of resulta (for reasons discussed above) to extend our analysis
this work are given in Towler (1992). to proposing a free-radical mechanism. Our analysis was
thus limited to molecular species. A more complete
Discussion analysis of the kinetics would take account of free-radical
effects, and also of other minor reactions and byproducts.
The goal of this study was to develop an understanding This would require rather more accurate data than are
of the kinetics that could be used to design a reactor for presented here, in particular, the flow pattern in the reactor
the Zero-Emissions Sulfur Process discussed in part 2 of would have to be more accurately described. Such an
this series (Towler and Lynn, 1993). For this purpose the analysis must therefore be the subject of future work, and
model developed above is adequate, since it describes both a suitable concluding note is that the engineering model
2810 Ind. Eng. Chem. Res., Vol. 32, No. 11, 1993

developed above will be useful not only in process design, does not significantly affect the Zero-Emissions Sulfur
but also in designingan apparatus for more detailed kinetic Process (except to slightly improve the quality of the
studies. product gas). Other advantages of the new technology
compared to the Claus process will be discussed in part
Process Synthesis 2 (Towler and Lynn, 1993).
One can easily envisiona processbased on this chemistry, Acknowledgment
in which a stream containing H2S and C02 in the correct
proportions is produced by absorbing these species from This research was funded by the Morgantown Energy
a suitable sour gas. This mixture is sent to a furnace, TechnologyCenter through the U.S.Department of Energy
where it reacts forming S2, CO, HzO, H2, and small amounts under Contract DE-AC03-76SF00098.
of COS, SOZ,and CS2. The sulfur is condensed and the
remaining acid gases are absorbed in a second absorber, Nomenclature
leaving a product gas consisting chiefly of Hz and CO. The
acid gases can than be stripped from the solvent in a Ci = concentration of species i (mol/cm3)
stripper and recycled to the reactor. Cbd = total concentration of the gas phase (mol/cm3)
A detailed description of the process, operating con- Eact= activation energy (cal/(mol-K))
ditions, and flowsheet will be presented in part 2 (Towler kl = first-order, noncatalytic, H2S thermal decomposition
and Lynn, 1993). There are, however, some advantages rate constant (mol/(cm3.s.atm))
to a process based on this chemistry that were apparent kz = rate constant for noncatalytic thermal-recombination
from an early stage, including some that arise from the reaction between Hz and S2 (moF5 ~ m - ls-1 . ~atm-1)
thermochemical considerations discussedabove. The most KWs = equilibrium constant for the "sulfur-gas-shift"reaction,
important advantage is that by using CO2 instead of air reaction 8
as oxidant the tail gas problem is eliminated. The n = (subscript) property evaluated under the conditions of
concentration of sulfur-containing species in the gas the nth interval
streams leaving the process (the sweetened gas and the P = total pressure (atm)
product gas) is therefore determined by the operating Pi = partial pressure of species i (atm)
conditions of the two absorber/stripper loops, and may be r = rate of H2S thermal decomposition (mol/(cm3-s))
controlled to a very low value (a few ppm) depending on R = ideal gas constant (cal/(mol.K))
how lean the solvent is stripped in each stripper. Note t = time (a)
that the process is strictly "zero-emissions" in the sense At = residence time (s)
that there is no tail gas. Effluent streams of course have T = temperature (K)
some sulfur content, but this may be controlled to as low Tset= furnace set-point temperature ("C)
a level as can be achieved using absorber/stripper tech- AT = temperature difference (K)
nology. * = denotes that the species is a free radical
Secondly, the product gas leavingthe process is a mixture
of CO and HZsaturated with water vapor a t the absorber Literature Cited
temperature. This gas may be used as fuel or synthesis
Bandermann, F.; Harder, K. B. Production of Hz via Thermal
gas; in the latter case it may be mixed with steam and sent Decomposition of H2S and Separationof Hz and H2S by Pressure-
to a shift reactor for conversion to hydrogen. The choice Swing Adsorption.Znt. J. Hydrogen Energy 1982,7 (6), 471-475.
between these uses depends on economic considerations Bowman, M. G. Thermochemical Cycle for Splitting Hydrogen
that may be expected to vary from site to site; however, Sulfide. U.S. Patent 4,999,178, March 12, 1991.
the value of the hydrogen from the H2S is recovered in all Bradley,J. N.;Dobson, D. C. Oxidationof Hydrogen Sulfidein Shock
cases. Waves, I. Absorption Studies of OH and SO2 in H&S-02-Ar
Further advantages arise from the breadth of the Mixtures. J . Chem. Phys. 1967a, 46 (8), 2865.
Bradley, J. N.;Dobson, D. C. Oxidationof Hydrogen Sulfidein Shock
maximum in the yield curve (Figure 6). The Claus process Waves, 11. The Effect of Added Hydrogen on the Absorption of
requiresvery tight control of the air-to-HzS ratio to achieve OH and SOz. J. Chem. Phys. 1967b, 46 (8), 2872.
an exact stoichiometric ratio of 2 H2S to 1 SO2 in the Chem. Eng. News 1992, June 29, 37.
reactor sequence. If this ratio is upset slightly, then the Chase, M. W., Jr.; Davies, C. A.; Downey, J. R., Jr.; Frurip, D. J.;
unconverted HzS or SO2 is passed on to the tail-gas unit, McDonald,R. A,;Syvemd,A. N. JAIVAF Thermochemical Tables,
creating considerable difficulties. For a process based on 3rd Ed.;National Bureau of Standards: Washington, DC, 1985;
the interaction between COZ and H2S, the COz-to-HaS Vol. 14.
Chenery,J. A.;Fakhr, A.; Wood,M. I.; Simpson,C. J. S. M. Vibrational
ratio is not critical owing to the rather broad maximum Analysisof the Products fromthe ThermalDecompositionof OCS
in the reactor yield. The Zero-Emissions Sulfur Process and COz. Chem. Phys. Lett. 1983, 96(2), 143-147.
is therefore much more robust and requires much less Chivers,T.;Lau, C. TheThermalDecompositionof HydrogenSulfide
sophisticated control. A similar problem is caused for the Over Alkali Metal Sulfides and Polysulfides. Znt. J . Hydrogen
Claus process if the incoming H2S stream is contaminated Energy 1986, 10 (l),21-25.
by hydrocarbon material. This is often the case if low- Chivers,T.;Lau, C. The Thermal Decompositionof HydrogenSulfide
temperature, high-pressure absorption is used, and is Over Vanadium and Molybdenum Sulfides and Mixed-Sulfide
Catalyst in Quartz and Thermal-DiffusionColumn Reactors.Znt.
caused by condensation of organicsin the absorber solution J. Hydrogen Energy 1987,12 (4), 235-243.
and subsequent vaporization in the stripper. If hydro- Chivers, T.; Hyne, J. B.; Lau, C. The Thermal Decomposition of
carbon material is present in the Claus process feed, it HydrogenSulfide Over TransitionMetal Sulfides.Int. J. Hydrogen
consumes some of the oxygen from the air feed, making Energy 1980,5 (5), 499-506.
control of the H2S-to402 ratio more difficult. Fluctua- Dokiya, M.; Fukuda, K.; Yokokawa, H.; Kameyama, T. The Study
tions in the amount of hydrocarbon material present of Thermochemical Hydrogen Preparation. VI. A Hydrogen-
evolving Step through the H2S-CO Cycle. Bull. Chem. SOC.Jpn.
obviously exacerbate this problem. For the Zero-Emissions 1978,51 (l), 150-153.
Sulfur Process any organic material sent to the furnace Fellmuth,P.;Lutz, W.;Biilow, M. Influenceof Weakly Co-ordinated
will react with COz, forming carbon monoxide and Cations and Basic Sites Upon the Reaction of H2S and COZon
hydrogen, and will end up as part of the product gas. This Zeolites. Zeofites 1987, 7 (4), 367-371.
Ind. Eng. Chem. Res., Vol. 32, No.11,1993 2811
Fukuda, K.; Dokiya, M.; Kameyama, T.; Kotera, Y. Catalytic Raymont, M. E. D. Make Hydrogen from Hydrogen Sulfide.
Decomposition of Hydrogen Sulfide. Znd. Eng. Chem. Fundam. Hydrocarbon Process. 1975,54 (7), 139-142.
1978,17(4),243-248. Roth,P.; L&r, R.; Barner, V. Thermal Decomposition of Hydrogen
Gangwal, S.K.;McMichael, W. J.;Dorchak, T. P. The Direct Sulfur- Sulfide at Low Concentrations. Combust. Flame 1982,45(31,273-
Recovery Process. Environ. h o g . 1991,lO (3,186-191. 286.
George, Z.M. Kinetics of Cobalt-Molybdate-Catalyzed Reactions of Schecker, H. G.; Wagner, H. G. On the Thermal Decomposition of
SO2 with H2S and COS and the Hydrolysis of COS. J . Catal. 1974, COS. Int. J. Chem. Kinet. 1969,1, 541-549.
32,261-271. Sugioka, M.; Aomura, K. A Possible Mechanism for Catalytic
Kaloidas, V.; Papayannakos, N. Hydrogen Production from the Decomposition of Hydrogen Sulfide over Molybdenum DisuKde.
Decomposition of Hydrogen Sulfide. Equilibrium Studies on the Znt. J . Hydrogen Energy 1984,9(111, 891-894.
System HzS/H2/Si (i = 1,...,8)in the Gas Phase. Znt. J. Hydrogen Towler, G. P. Synthesis and Development of Processes for the
Energy 1987,12(6),403-409. Recoveryof Sulfur From Acid Gases.Ph.D. Dissertation, University
Kaloidas, V.;Papayannakos, N. Kinetics of Thermal, Non-catalytic of California, Berkeley, 1992.
Decomposition of Hydrogen Sulfide. Chem. Eng. Sci. 1989, 44 Towler, G. P.; Lynn, S. Development of a Zero-Emissions Sulfur-
(ll), 2493-2500. Recovery Process. 2. A Sulfur-Recovery Process Based on the
Katsumoto, M.; Fueki, K.; Mukaibo, T. An Investigation of the Gas- Reactions of HzS and COz at High Temperature. Znd. Eng. Chem.
Solid Interface Reaction. Bull. Chem. SOC.Jpn. 1973,46,3641- Res. 1993,following paper in this issue.
3644. West, J. R. Sulfur Recovery. In Kirk-Othmer Encyclopedia of
Kotera, Y. The Thermochemical Hydrogen Program at N.C.L.I. Znt. Chemical Technology; Grayson, M., Ed.; Wiley: New York, 19W,
J. Hydrogen Energy 1976,1,219-220. Vol. 22,pp 267-297.
Levy, A.; Merryman, E. L. The Microstructure of Hydrogen Sulfide Whitney, G. M.; Bang, Y.; Denn, M. M.; Petersen, E. E. Sulfur
Flames. Combust. Flame 1965,9,229. Capture During Partial Coal Combustion. Chem. Eng. Commun.
Merryman, E. L.; Levy, A. Kinetics of Sulfur Oxide Formation in 1987,55(1-6),83-93.
Flames: I1Low Pressure HzS Flames. J. Air Pollut. Control 1967,
17,800. Received for review January 11, 1993
Merryman, E. L.; Levy, A. Disulfur and the Lower Oxides of Sulfur Revised manuscript received June 8, 1993
in Hydrogen Sulfide Flames. J. Phys. Chem. 1972,76(14),1925. Accepted June 24, 1993.
Norrish, R. G. W.; Zeelenberg, A. P. The Combustion of Hydrogen
Sulfide Studied by Flash Photolysis and Kinetic Spectroscopy. Abstract published in Advance ACS Abstracts, September
R o c . R. SOC.Ser. A 1957,240,293. 1, 1993.

You might also like