You are on page 1of 8

Applied Energy 92 (2012) 57–64

Contents lists available at SciVerse ScienceDirect

Applied Energy
journal homepage: www.elsevier.com/locate/apenergy

Fate of sulfur with H2S injection in methane/air flames


H. Selim, A. Al Shoaibi 1, A.K. Gupta ⇑
Department of Mechanical Engineering, University of Maryland, College Park, MD 20742, United States

a r t i c l e i n f o a b s t r a c t

Article history: Investigation of sulfur chemistry of hydrogen sulfide in methane/air flame under different equivalence
Received 7 July 2011 ratios is conducted. A range of equivalence ratios extending from fuel-lean (U = 0.5), stoichiometric
Received in revised form 13 October 2011 (U = 1.0), to fuel-rich (Claus condition, U = 3.0) are examined here. Hydrogen sulfide, methane, and air
Accepted 1 November 2011
were premixed prior to the injection of mixture into the combustor. Spatial gas sampling was carried
Available online 29 November 2011
out both axially and radially along the reactor at various locations downstream of the reactor. Oxidation
competition between H2S and CH4 was found to be significant due to the premixed conditions. Subse-
Keywords:
quently, reaction of H2S formed SO2 rather than sulfur (S2). In addition, sulfur dioxide showed a consis-
Sulfur chemistry
Hydrogen sulfide combustion
tent (almost-constant) trend under all equivalence ratios examined here. On the other hand, results
Carbon disulfide formation showed the presence of carbon sulfide (CS2) under both lean and stoichiometric conditions. The forma-
Enhanced sulfur recovery tion of carbon disulfide was found to be due to the reaction of methane with sulfur compounds to form
Higher hydrocarbon formation CS2. Moreover, intermediate species such as, CH react with H2S to form CS2. Higher hydrocarbons (ethane
and ethylene) were observed under Claus conditions. This was due to the coupling catalytic effect of SO2
which enhances the dimerization of CH3 to form ethane. Dehydrogenation of ethane forms ethylene.
Under stoichiometric and lean conditions higher hydrocarbons were not formed due to the lack of high
SO2 concentration in the reaction pool.
Published by Elsevier Ltd.

1. Introduction environment and industry. Claus process [2–6] is commonly used


for hydrogen sulfide treatment wherein reaction between H2S
Hydrogen sulfide is known to be present in different quantities and O2 occurs under rich conditions (U = 3) to form elemental
in crude natural gas extracted from the gas/oil wells. The use of sulfur (S2). During this reaction one third of H2S is burned to form
crude natural gas in any chemical to thermal energy transforma- SO2 (reaction (1)). Then the reaction continues between the so
tion process, e.g. combustion, results in the formation of acid gases, formed SO2 and unburned H2S to form sulfur (reaction (2)) which
such as, SO2 and SO3 which are then transformed to sulfurous and is then captured in liquid or solid form. Practically, the Claus pro-
sulfuric acid. If these gases are released to the atmosphere in trace cess is divided into two main stages of thermal stage and catalytic
amounts, they would lead to the problem of acid rain. In addition, stage. Both stages have same chemical reactions, but catalysts are
the presence of these gases lowers the dew point of the combus- used in the later stages that possess considerably low concentra-
tion gases so that exhaust gases must be released at higher temper- tions of H2S.
atures to prevent condensation of gases and to mitigate corrosion
3H2 S þ 1:5O2 ! H2 O þ SO2 þ 2H2 S DHr ¼ 518 kJ=mol ð1Þ
problems. The release of gases at higher temperatures to the atmo-
sphere reduces the plant efficiency significantly. Therefore, hydro-
2H2 S þ SO2 ! 1:5S2 þ H2 O DHr ¼ 47 kJ=mol ð2Þ
gen sulfide must be removed from natural gas prior to its
utilization. Amine extraction process [1] is commonly used for Furthermore, there is increased need for the hydrolysis and
the removal of acidic gases, mainly H2S, from crude natural gas hydrogenation of other sulfur compounds present in fossil fuel
wherein alkaline-based organic compounds are used to absorb wells, such as, carbonyl sulfide and carbon disulfide [7] that are
H2S from the fuel stream. Although the concentration of the ab- precursors to the H2S. These compounds can also be formed in
sorbed H2S is fairly low, it is crucial that hydrogen sulfide under- Claus process in presence of methane or higher hydrocarbons [8].
goes treatment process to hinder its harmful effects on both the With the shrinking reserves of fossil fuels we must place increased
emphasis on extracting energy from wells that contain higher
amounts of sulfur compounds. Hydrogenation and hydrolysis of
⇑ Corresponding author. Tel.: +1 301 405 5276. sulfurous compounds transform them into hydrogen sulfide [5,9].
E-mail address: akgupta@umd.edu (A.K. Gupta).
1
Address: The Petroleum Institute, Abu Dhabi, United Arab Emirates.
COS þ H2 O () H2 S þ CO2 ð3Þ

0306-2619/$ - see front matter Published by Elsevier Ltd.


doi:10.1016/j.apenergy.2011.11.005
58 H. Selim et al. / Applied Energy 92 (2012) 57–64

CS2 þ 2H2 O () 2H2 S þ CO2 ð4Þ the above quoted values. One of the main reasons of this difference
in the efficiency is the formation of undesired sulfur byproducts
SO2 þ 2H2 () H2 S þ 2H2 O ð5Þ such as CS2 and COS that are formed at other reactor conditions.
As discussed earlier the Claus process is followed by hydrolysis
Sn þ nH2 () nH2 S ð6Þ
and hydrogenation stages to prevent the escape of these products.
Reactions (3) and (4) represent the hydrolysis of carbonyl sul- Formation of CS2 and COS could be due to the reaction of hydrocar-
fide and carbon disulfide, while reactions (5) and (6) describe the bons that exist in the acid gas stream with hydrogen sulfide. On the
hydrogenation of sulfur dioxide and elemental sulfur. The formed other hand, the non-uniformity of reactants causes non-uniform
hydrogen sulfide is absorbed from pure fuel stream via efficient local equivalence ratio distribution in the reactor to cause non-uni-
absorbents such as zinc oxide (ZnO) or alkanolamines [10]. form thermal field distribution in the reactor. In this paper we
Several researchers have investigated the fate of hydrogen sul- investigate the combustion of hydrogen sulfide in methane/air
fide flame with focus on structure, products yield and evolution of flame at difference equivalence ratios under perfectly mixed reac-
various species under various reactor and operational conditions tants conditions. Three specific cases have been examined in detail.
[11–20]. In these investigations hydrogen sulfide has been used The first represents fuel-lean conditions (U = 0.5), the second rep-
either as the primary fuel or a secondary fuel introduced into an- resent stoichiometric conditions (U = 1.0), and the third represent
other flame. Bernez-Cambot et al. [11] investigated experimentally Claus condition having fuel-rich conditions (U = 3.0). Gas sampling
the flame structure of H2S/air diffusion flame under Claus condi- has been used to examine the spatial distribution of species in the
tions. In their results, they divided the flame into three distinct reactor in order to provide better insight on the reaction chemistry.
zones. First zone is the thermal and chemical decomposition of
H2S wherein hydrogen is the major product. Second zone contains
2. Experimental
the oxidation of both H2S and H2 formed in the first zone. Third
zone incorporated partial consumption of hydrogen and to a lesser
A schematic diagram of the experimental setup is shown in
extent sulfur diffusing from the flame. Hydrogen sulfide decompo-
Fig. 1. The facility consists of a quartz tube reactor of 19 cm length
sition was shown to be significant at high temperatures [12,13],
and 4 cm diameter. A double concentric tubular burner was used
where considerable amounts of hydrogen are formed. Bowman
for all the experiments wherein air is premixed with methane
and Dodge [12] examined the decomposition of H2S in a mixture
and hydrogen sulfide and injected into the outer annulus of the
of hydrogen sulfide and argon. Different hydrogen sulfide concen-
burner. The central tube was not used in this study. Bluff body
trations (mole fraction) were examined (0.025, 0.014, 0.007, and
was used to stabilize the flame immediately downstream from
0.014). So there was no chemical reaction in their studies. Note
the burner exit. The gases from the burner were allowed to flow
that under Claus process conditions, wherein the chemical reac-
into the quartz tube reactor. Sonic throat quartz sampling probe
tions occur under fuel-rich conditions, H2S decomposition is ex-
was used for gas sampling. Throat diameter of the sampling probe
pected to be substantial. Hawboldt et al. [14] studied the kinetics
was of the order of few microns so that the flow is chocked at its
of H2S pyrolysis under Claus process conditions. They determined
throat. The rapid expansion of the gases after the throat section
an expression for H2S pyrolysis over a temperature range of 850–
rapidly quenches the gases to freeze the gas compositions. Two
1150 °C. They found that the rate of dissociation of H2S is minimal
computer controlled traversing mechanism were used to move
below 1000 °C which interprets the need for high temperatures in
the sampling probe axially and radially along the reactor. The inner
Claus process. Muller et al. [15] studied sulfur chemistry in fuel-
and outer diameters of the sampling probe were 3 and 4 mm,
rich H2/O2/N2 flames with low concentrations (0.25%, 0.5% and
respectively. A suction pump was connected to the sampling line
1% of H2S) in the mixture. They measured concentrations of SH,
to introduce the sampled gas to gas chromatograph (GC) for gas
S2, SO, SO2, and OH using quantitative laser fluorescence measure-
analysis. The sampled gas is split inside the GC into two streams.
ments. With the help of the aforementioned radical measurements
First stream is injected into thermal conductivity detector which
they were able to provide kinetics parameters for various possible
is responsible for the analysis of carbon monoxide and hydrogen.
intermediate chemical reactions of sulfur compounds. Selim et al.
Second stream is injected into flame photometric detector which
[16] examined the effect of hydrogen sulfide in methane/air flames
is responsible for gas sampling of stable sulfur compounds (hydro-
on the products speciation. They introduced hydrogen sulfide at
gen sulfide and sulfur dioxide). Mean temperatures were measured
different concentrations into slightly-lean methane/air flame. Their
using a K-type thermocouple that was also connected to a traverse
goal was to examine the effect of H2S/O2 equivalence ratio on the
mechanism. This allowed for the measurements of axial and radial
products evolution rate and behavior. The results revealed that
temperature distribution from within the reactor. The whole
most of the hydrogen sulfide is transformed into sulfur dioxide un-
experimental setup was placed inside a fume hood. The fume hood
der stoichiometric or lean conditions. However, under Claus condi-
was connected to an exhaust duct wherein a fan is used to induce
tions considerable sulfur depositions were observed on the reactor
air into the fume hood for safety purposes.
walls and sulfur dioxide concentration was dropped significantly.
Azatyan et al. [17] examined the behavior of hydrogen sulfide, car-
bon disulfide, and carbonyl sulfide combustion at low-pressure 3. Experimental conditions
using electron spin resonance technique along with gas chroma-
tography. Their results revealed that first stage of H2S formation in- Experiments were conducted to investigate the behavior of
cludes the formation of H2, SO2 and SO and they suggested that the hydrogen sulfide combustion in methane/air flame under perfectly
presence of H2S is considered inhibitor of H2 oxidation. The second premixed conditions at difference equivalence ratios. Methane/air
stage includes hydrogen oxidation coupled with the formation of mixture was combusted under slightly fuel-lean conditions with
hydroxyl group radical. These results agree with the findings of H2S injected at different flow rates to vary the equivalence ratio
Bernez-Cambot et al. [11], but the major contradiction was the lack with control of H2S/O2 ratio in the mixture. Spatial distribution
of OH presence in first stage of the reaction. Theoretical results of gas species was obtained by gas sampling of the local species
showed that up to 74% of sulfur can be captured from within the along longitudinal axis of the reactor.
thermal stage of Claus process [18]. However, in practical applica- Experiments procedure included the following steps. First
tions the average efficiency of thermal Claus process is only 65% methane/air equivalence ratio was adjusted to achieve slightly
[19] at best. Mostly the thermal stage efficiency is much lower than fuel-lean conditions. Table 1 shows the flow rates of air, methane,
H. Selim et al. / Applied Energy 92 (2012) 57–64 59

Computer controller
traverse mechanisms

Gas sample to gas chromatograph


Sonic-throat
8.94
sampling probe
3.58

9.94 Quartz tube


reactor

Burner
Dimensions
in
millimeters

Air+CH4+H2S

Fig. 1. A schematic diagram of the experimental setup.

bustion of H2S in methane/air flame. Dimensionless axial (W) and


Table 1 radial (R) distances are used to show the trends for temperature
Reactants flow rate at each equivalence ratio.
and the evolution of combustion products. Jet diameter was used
Equivalence Air flow CH4 flow Excess oxygen H2S flow rate to transform the linear distances to dimensionless parameters,
ratio (U) rate (l/min) rate (l/min) (cm3/min) (cm3/min) i.e., W = axial distance/Djet, and R = radial distance/Djet.
U = 3.0 9.7 0.99 55 110
U = 1.0 9.7 0.97 90 60
U = 0.5 9.7 0.95 144 48 4.1. Reactor temperature distribution

Reactor mean temperatures were measured using a K-type


thermocouple (thermocouple bead diameter was 0.8 mm). Flame
and hydrogen sulfide at each condition. Gas sampling was carried photos showed acceptable flame symmetry, hence mean tempera-
out at different axial and radial locations and the gas samples ob- ture measurements were carried out on only one half section of the
tained were injected directly into the GC. Due to the reactor sym- reactor in the longitudinal axis of the reactor. Mean temperature
metry only half of the reactor was examined. Gas samples were distribution in the reactor was measured radially and axially using
obtained along three radial locations of 0, 0.635, and 1.27 cm from computer controlled traverse mechanisms. Fig. 2 shows the tem-
the burner centerline. Samples were taken at axial distances of 0, perature contours of CH4/air flame for the left half of the reactor
0.635, 1.27, 1.9, 2.54, 3.81, 5.08, 6.35, 7.62, 10.16, 12.7, 15.24, under Claus conditions. Under fuel-lean and stoichiometric condi-
and 17.78 cm) along the longitudinal axis of the reactor. Data tions wherein H2S flow rate is minimal, temperature distribution
was repeated at least three times at each location for data repeat- did not show considerable change as compared to CH4/air flame.
ability and reliability. The flow rate and conditions are given in A direct comparison between CH4/air flame temperatures with-
Table 1. out/with H2S revealed that temperatures did not change consider-
ably away from the flame zone. However, H2S addition increased
the temperature around the flame zone with maximum increase
4. Results and discussion
in temperature of 50 °C at W = 2.83 and R = 0.0.

Temperature scan of only one half side of the reactor was per-
formed and this will first be presented. The gas sampling was car- 4.2. Hydrogen sulfide combustion analysis
ried out to determine the spatial distribution of species on half of
the longitudinal cross section of the reactor. Nitrogen and carbon The gas analysis was carried out using gas chromatograph (GC).
dioxide were excluded from the focus. This is because the mole The results obtained were estimated for accuracy of the data. The
fraction of both N2 and CO2 did not change significantly under inaccuracy arising from gas chromatograph (GC), temperature,
any of the conditions examined. Results only show behavior of flow rates, and traverse mechanism in the experimental results
H2S combustion with specific focus on the evolution of various presented here was considered. The GC used is known to have
product species in methane/air flame. The comparison between ±0.1% accuracy. Temperature effects of the reactants were consid-
CH4/air flame without and with H2S addition is presented in one ered to be negligible (<0.1 °C). The flow rate accuracy is estimated
of our previous studies [16]. In this paper we only discuss the com- to be 1.5% of the full scale. The traversing mechanism used had an
60 H. Selim et al. / Applied Energy 92 (2012) 57–64

CH 4/Air, with H2S, =1.0


1.2

H2 Mole Fraction (%)


R=0.0
0.8
R=1.77
0.6 R=3.54

0.4

0.2

0
0 10 20 30 40 50
Dimensionless Axial Distance (W)

Fig. 4. Hydrogen mole fraction. Flame conditions: methane/air with H2S, U = 1.0,
W = axial distance/Djet, R = radial distance/Djet.

CH 4/Air, with H2S, =3.0


1.8
1.6

H2 Mole Fraction (%)


1.4
1.2
1
0.8
R=0.0
0.6
R=1.77
0.4
R=3.54
0.2

Fig. 2. Spatial temperature distribution of the reactor under Claus conditions, 0


W = axial distance/Djet, R = radial distance/Djet.
0 10 20 30 40 50
Dimensionless Axial Distance (W)

accuracy of 0.01%. These errors were found to be within the exper- Fig. 5. Hydrogen mole fraction. Flame conditions: methane/air with H2S, U = 3.0,
W = axial distance/Djet, R = radial distance/Djet.
imental data point symbols presented in the figures.
Figs. 3–5 describe the distribution of hydrogen mole fraction at
different equivalence ratios along the reactor at three radial loca-
tions. Along the reactor centerline a constant decrease in H2 mole spread radially outward. Moreover, the presence of hydrogen sul-
fraction is observed at any equivalence ratio. Presence of hydrogen fide is considered an inhibitor to hydrogen oxidation primarily
at the reactor centerline is attributed to the flow recirculation as a [16]. Further downstream hydrogen starts to be more competitive
direct result of the bluff body presence in the flow path. This sug- with hydrogen sulfide reaction. Therefore, hydrogen sulfide inhibi-
gests that the constant decrease in H2 mole fraction to be due to tion to hydrogen oxidation justifies the increase of H2 mole fraction
either oxidation or reaction with sulfur intermediate species, such upstream. Further downstream, hydrogen decreases due to the
as, S and SO [16]. However, radially outwards, hydrogen mole frac- oxidation competition between H2 and H2S. On the other hand,
tion tends to peak to a maximum then a monotonic decrease is ob- at U = 1.0 and U = 0.5, one can notice that H2 peak at R = 1.77 ex-
served. The peak of H2 mole fraction is attributed to the ceeds the corresponding value of H2 mole fraction at R = 0.0. This
configuration of the burner which dictates the reactants flow to is attributed to the availability of higher amounts of oxygen at
R = 1.77. This is translated to higher oxidation rates of hydrogen.
Subsequently, this leads to lower amounts of hydrogen present
in the flow recirculation zone downstream of the bluff body. In
CH 4/Air, with H2S, =0.5 addition, the lower amount of H2S at lower equivalence ratios
0.5
(U = 1.0 and U = 0.5) lessens the oxidation inhibition effect of
hydrogen in the reaction pool.
H2 Mole Fraction (%)

0.4
R=0.0 Figs. 6–8 depict carbon monoxide mole fraction at different
equivalence ratios along the reactor at three radial locations. Sim-
0.3 R=1.77
ilar to the behavior of hydrogen, carbon monoxide decreases
R=3.54 monotonically along the centerline of the reactor. However, radi-
0.2
ally outward, carbon monoxide peaks up to a maximum and then
it decreases. All three CO mole fraction values are almost equal
0.1
at the reactor exit. In the reaction zone CO mole fraction at
R = 1.77 is higher as compared to the values at R = 0.0 and
0
0 10 20 30 40 50 R = 3.54. This is attributed to the presence of bluff body which im-
poses the flow to spread radially outwards close to R = 1.77, where
Dimensionless Axial Distance (W)
higher reactions rates are observed. Furthermore, CO oxidation oc-
Fig. 3. Hydrogen mole fraction. Flame conditions: methane/air with H2S, U = 0.5, curs so that small percentage of CO is presented at R = 0.0 and
W = axial distance/Djet, R = radial distance/Djet. R = 3.54.
H. Selim et al. / Applied Energy 92 (2012) 57–64 61

CH 4/Air, with H2S, =0.5 CH 4/Air, with H2S, =0.5


0.9
0.004
0.8
CO Mole Fraction (%)

0.0035
0.7 R=0.0

H2S Mole Fraction (%)


R=0.0
0.6 0.003 R=1.77
R=1.77
0.5 0.0025 R=3.54
0.4 R=3.54
0.002
0.3
0.2 0.0015

0.1 0.001
0 0.0005
0 10 20 30 40 50
Dimensionless Axial Distance (W) 0
0 10 20 30 40 50
Fig. 6. Carbon monoxide mole fraction. Flame conditions: methane/air with H2S, Dimensionless Axial Distance (W)
U = 0.5, W = axial distance/Djet, R = radial distance/Djet.
Fig. 9. Hydrogen sulfide mole fraction. Flame conditions: methane/air with H2S,
U = 0.5, W = axial distance/Djet, R = radial distance/Djet.
CH 4/Air, with H2S, =1.0
1.8
1.6 CH 4/Air, with H2S, =1.0
CO Mole Fraction (%)

0.035
1.4
R=0.0
1.2 0.03
R=1.77

H2S Mole Fraction (%)


1 R=0.0
R=3.54 0.025
0.8 R=1.77
0.6 0.02
R=3.54
0.4
0.015
0.2
0 0.01
0 10 20 30 40 50
0.005
Dimensionless Axial Distance (W)
0
Fig. 7. Carbon monoxide mole fraction. Flame conditions: methane/air with H2S,
0 10 20 30 40 50
U = 1.0, W = axial distance/Djet, R = radial distance/Djet.
Dimensionless Axial Distance (W)

Fig. 10. Hydrogen sulfide mole fraction. Flame conditions: methane/air with H2S,
CH 4/Air, with H2S, =3.0
U = 1.0, W = axial distance/Djet, R = radial distance/Djet.
3

2.5 R=0.0
CO Mole Fraction (%)

R=1.77
2 CH 4/Air, with H2S, =3.0
R=3.54 0.12
1.5
0.1
H2S Mole Fraction (%)

1
0.08 R=0.0
0.5
0.06 R=1.77
0
R=3.54
0 10 20 30 40 50 0.04
Dimensionless Axial Distance (W)
0.02
Fig. 8. Carbon monoxide mole fraction. Flame conditions: methane/air with H2S,
U = 3.0, W = axial distance/Djet, R = radial distance/Djet.
0
0 10 20 30 40 50
Dimensionless Axial Distance (W)
The mole fraction distribution of hydrogen sulfide at different
equivalence ratios at three different radial locations in the reactor Fig. 11. Hydrogen sulfide mole fraction. Flame conditions: methane/air with H2S,
is shown in Figs. 9–11. Under any conditions, H2S peaks to a max- U = 3.0, W = axial distance/Djet, R = radial distance/Djet.
imum at R = 1.77 and then decreases in a consistent fashion. This is
attributed to the fact that at R = 1.77 the sampling probe is in the
vicinity of the path lines of injected reactants. Under Claus condi- Figs. 12–14 present the behavior of sulfur dioxide mole fraction
tions, considerable amount of H2S is observed at R = 3.54. This is at different equivalence ratios along the reactor at three radial
attributed to the lack of oxygen (mixture is slightly rich) which locations. Under Claus conditions, SO2 mole increases somewhat
prevents the rapid combustion of H2S. Hydrogen sulfide around until it reach to a maximum and decreases slightly further down-
the centerline is always negligible since most of the H2S is com- stream of the reactor. The decrease in SO2 is attributed to its reac-
busted in the recirculation region downstream of the bluff body. tion with H2S to form sulfur (reaction (2)). In previous studies
62 H. Selim et al. / Applied Energy 92 (2012) 57–64

CH 4/Air, with H2S, =0.5 mole fraction increases along the reactor. This is attributed to the
0.25 unlikelihood of oxygen depletion under lean or stoichiometric
conditions. Under all conditions SO2 mole fraction does not change
SO2 Mole Fraction (%)

0.2 significantly in the radial direction at the reactor exit.


Figs. 15 and 16 show the distribution of carbon disulfide mole
0.15 fraction at U = 3.0 and U = 1.0. Note that CS2 was not observed un-
R=0.0 der lean conditions (U = 0.5). The formation of carbon disulfide is
0.1 R=1.77 attributed primarily to the reaction of methane and sulfur com-
R=3.54 pounds. Reactions (7)–(9) describe the possible channels for CS2
0.05 formation [22,23]:

CH4 þ 2S2 () CS2 þ 2H2 S ð7Þ


0
0 10 20 30 40 50
CH4 þ 2S () CS2 þ 2H2 ð8Þ
Dimensionless Axial Distance (W)

Fig. 12. Sulfur dioxide mole fraction. Flame conditions: methane/air with H2S, CH þ 2H2 S () CS2 þ 4H2 ð9Þ
U = 0.5, W = axial distance/Djet, R = radial distance/Djet.
Under Claus conditions and stoichiometric conditions it is more
common for these reactions to take place. However, under lean
conditions, oxidation of H2S and CH4 to form SO2 and CO2, respec-
CH 4/Air, with H2S, =1.0 tively, is more dominant. Under Claus conditions, CS2 mole fraction
0.3 is almost one order of magnitude higher than CS2 value under stoi-
chiometric conditions. This is mainly emanated from the rarity of
0.25
SO2 Mole Fraction (%)

CH and S radicals where the availability of more oxygen will trans-


form these radicals to SO and OH. The latter radicals will dictate
0.2
R=0.0 the reaction to form other end-products, such as, SO2, H2O, and
0.15 R=1.77 CO2.
R=3.54 Figs. 17 and 18 depict the formation of higher hydrocarbons
0.1 (ethane and ethylene) in the reaction pool at R = 1.77 and
R = 3.54 under Claus conditions. The formation of higher hydrocar-
0.05 bons is attributed to CH3 dimerization to form C2H6. On the other
hand, Ethane dehydrogenation forms ethylene. The dimerization
0
0 10 20 30 40 50 of alkyl groups emanates from the presence of SO2 which acts as
a coupling catalyst. In one of previous study [16] we discuss the
Dimensionless Axial Distance (W)
formation of higher hydrocarbons around the centerline of the
Fig. 13. Sulfur dioxide mole fraction. Flame conditions: methane/air with H2S, reaction zone wherein H2S was injected in the middle tube. The
U = 1.0, W = axial distance/Djet, R = radial distance/Djet. difference here is that H2S is injected in the annular part in the
present study. The absence of higher hydrocarbons around the cen-
terline, in this study, justifies and validates the hypothesis of SO2
effect as a coupling catalyst for CH3. Higher hydrocarbons were
not observed under stoichiometric and lean conditions. This is
CH 4/Air, with H2S, =3.0 attributed to the relatively low SO2 concentration which reduces
0.4
the chances of CH3 dimerization. This justification supports with
0.35 the findings of Sofranko et al. [24]. Moreover, the higher oxygen
SO2 Mole Fraction (%)

0.3 availability enhances the chances of oxidation of the higher hydro-


carbons. On the other hand, C2H4 and C2H6 abruptly decompose
0.25 R=0.0
0.2 R=1.77
0.15 R=3.54
CH 4/Air, with H2S, =3.0
0.1 0.018
0.05 0.016
CS2 Mole Fraction (%)

0 0.014
0 10 20 30 40 50
0.012
Dimensionless Axial Distance (W)
0.01
Fig. 14. Sulfur dioxide mole fraction. Flame conditions: methane/air with H2S, 0.008 R=0.0
U = 3.0, W = axial distance/Djet, R = radial distance/Djet.
0.006 R=1.77
0.004
R=3.54
0.002
[20,21] it has been shown that H2S reaction tends to form SO2
0
rather than S2 in presence of oxygen. Since all reactants are pre- 0 10 20 30 40 50
mixed, probability of H2S oxidation is high as it competes with
methane. This justifies the strong presence of SO2 just downstream Dimensionless Axial Distance (W)
of the burner tip where most of H2S has swiftly transformed to SO2. Fig. 15. Carbon disulfide mole fraction. Flame conditions: methane/air with H2S,
On the other hand, under stoichiometric and lean conditions SO2 U = 3.0, W = axial distance/Djet, R = radial distance/Djet.
H. Selim et al. / Applied Energy 92 (2012) 57–64 63

CH 4/Air, with H2S, =1.0 that H2 and CO mole fractions exist with higher concentrations
0.005 away from the reactor centerline. This is emanated from the reac-
0.0045 tants tendency to spread radially outward. The presence of H2S and
CS2 Mole Fraction (%)

0.004 CO near the centerline is attributed to flow recirculation


0.0035 downstream of the bluff body. Therefore, hydrogen and carbon
0.003 monoxide are oxidized before reaching the centerline. Hydrogen
R=0.0 sulfide mole fraction showed its maximum values at R = 1.77 at
0.0025
any equivalence ratio. This is because that at R = 1.77 the sampling
0.002 R=1.77
probe is in the vicinity of the injected reactants pathlines. Sulfur
0.0015 R=3.54 dioxide showed a consistent (almost-constant) trend for all the
0.001 equivalence ratios examined. This is attributed to reactants being
0.0005 premixed prior to combustion. This enhances the oxidation compe-
0 tition between H2S and CH4 which leads the reaction of H2S to form
0 10 20 30 40 50 SO2 rather than S2. Carbon sulfide was formed under lean and stoi-
Dimensionless Axial Distance (W) chiometric conditions only. It was found that carbon disulfide is
formed due to the presence of methane which reacts with sulfur
Fig. 16. Carbon disulfide mole fraction. Flame conditions: methane/air with H2S,
U = 1.0, W = axial distance/Djet, R = radial distance/Djet.
compounds to form CS2. On the other hand, other intermediate
species, such as, CH reacts with H2S to form CS2 as well. Finally,
higher hydrocarbons (ethane and ethylene) were observed under
CH 4/Air, with H2S, =3.0 Claus conditions. This is attributed to the presence of SO2 which
0.04 acts as a coupling catalyst to the alkyl group (CH3) to form ethane.
0.035
Afterwards, the dehydrogenation of ethane forms ethylene. Under
C2H6 Mole Fraction (%)

stoichiometric and lean conditions higher hydrocarbons were not


0.03
formed due to the lack of high SO2 concentration in the reaction
0.025 pool.
R=1.77
0.02
0.015 R=3.54
Acknowledgments
0.01
The authors gratefully acknowledge the research support pro-
0.005
vided by The Petroleum Institute and ADNOC, Abu Dhabi, UAE.
0
0 10 20 30 40 50
Dimensionless Axial Distance (W) References

Fig. 17. Ethane mole fraction. Flame conditions: methane/air with H2S, U = 3.0, [1] Jensen AB, Webb C. Treatment of H2S-containing gases: a review of
W = axial distance/Djet, R = radial distance/Djet. microbiological alternatives. Enzyme Microb Technol 1995;17(1):2–10.
[2] Zagoruiko AN, Matros YS. Mathematical modeling of Claus reactors undergoing
sulfur condensation and evaporation. Chem Eng J 2002;87(1):73–88.
[3] Larraz R. Influence of fractal pore structure in Claus catalyst performance.
CH 4/Air, with H2S, =3.0 Chem Eng J 2002;86(3):309–17.
0.07 [4] Mora RL. Sulfur condensation influence in Claus catalyst performance. J Hazard
Mater 2000;79(1–2):103–15.
C2H4 Mole Fraction (%)

0.06 [5] El-Bishtawi R, Haimour N. Claus recycle with double combustion process. Fuel
Process Technol 2004;86(3):245–60.
0.05 R=1.77
[6] Monnery WD, Hawboldt KA, Pollock A, Svrcek WY. New experimental data and
kinetic rate expression for Claus reaction. Chem Eng Sci 2000;55(21):5141–8.
0.04 R=3.54 [7] Stumpf A, Tolvaj K, Juhasz M. Detailed analysis of sulfur compounds in gasoline
range petroleum products with high-resolution gas chromatography-atomic
0.03 emission detection using group-selective chemical treatment. J Chromatogr
1998;819(1–2):67–74.
0.02 [8] Mcintyre G, Lyddon L. Claus sulfur recovery options. Petroleum Technology
Quarterly, Spring; 1997. pp. 57–61.
0.01
[9] Rhodes C, Riddel SA, West J, Peter Williams B, Hutchings GJ. The low-
0 temperature hydrolosis of carbonyl sulfide and carbon disulfide: a review.
0 10 20 30 40 50 Catal Today 2000;59(3–4):443–64.
[10] Rhodes C, Riddel SA, West J, Peter Williams B, Hutchings GJ. The low-
Dimensionless Axial Distance (W) temperature hydrolysis of carbonyl sulfide and carbon disulfide: a review.
Catal Today 2000;59(3–4):443–64.
Fig. 18. Ethylene mole fraction. Flame conditions: methane/air with H2S, U = 3.0, [11] Bernez-Cambot J, Vovelle C, Delbourgo R. Flame structure of H2S–air diffusion
W = axial distance/Djet, R = radial distance/Djet. flame. In: 18th Symposium (international) on combustion. The Combustion
Institute; 1981. p. 777–83.
[12] Bowman CT, Dodge LG. Kinetics of the thermal decomposition of hydrogen
downstream to form lower hydrocarbons that contribute in CO, sulfide behind shock waves. In: 16th Symposium (international) on
CO2, and CS2 formation. combustion. The Combustion Institute; 1977. p. 971–82.
[13] Cullis CF, Mulcahy MFR. The kinetics of combustion of gaseous sulphur
compounds. Combust Flame 1972;18:225–92.
[14] Hawboldt KA, Monnery WD, Svrcek WY. New experimental data and kinetic
5. Conclusions rate expression for H2S pyrolysis and re-association. Chem Energy Sci
2000;55(5):957–66.
[15] Muller III CH, Schofield K, Steinberg M, Brodia HP. Sulfur chemistry in flames.
Hydrogen sulfide combustion in methane/air mixtures has been In: 17th Symposium (international) on combustion. The Combustion Institute;
investigated at different equivalence ratios ranging from lean, stoi- 1979. p. 867–79.
chiometric to rich conditions. Gas sampling was carried out axially [16] Selim H, Al Shoaibi A, Gupta AK. Effect of H2S in methane/air flames on sulfur
chemistry and products speciation. J Appl Energy 2001;88(8):2593–600.
at three different radial locations (R = 0.0, R = 1.77, and R = 3.54, [17] Azatyan VV, Gershenson UM, Sarkissyan EN, Sachyan GA, Nalbandyan AB.
where R is a dimensionless radial distance). The results showed Investigation of low-pressure flames of a number of compounds containing
64 H. Selim et al. / Applied Energy 92 (2012) 57–64

sulfur by the ESR method. In: 12th Symposium (international) on combustion. [21] Leeds University. Sulfur mechanism extension to the Leeds methane
The Combustion Institute; 1969. p. 989–94. mechanism; May 2002. <http://www.chem.leeds.ac.uk/combustion/sox.htm>.
[18] Selim H, Gupta AK, Sassi M. Acid gas composition effects on the reactor [22] Thacker CM, Miller E. Carbon disulfide production. Ind Eng Chem
temperature in Claus reactor. In: 6th AIAA international energy conversion 1944;36(2):182–4.
engineering conference (IECEC), Cleveland, OH; July 28–30, 2008 [AIAA 2008- [23] Zhu T, Dreher A, Flytzani-Stephanpoulos M. Direct reduction of SO2 to
5797]. elemental sulfur by methane over ceria-based catalysts. Appl Catal., B:
[19] Khudenko BM, Gitman GM, Wechsler EP. Oxygen based Claus process for Environ 1999;21(2):103–20.
recovery of sulfur from H2S gases. J Environ Eng 1993;119(6):1233–51. [24] Sofranko JA, Leonard JJ, Jones CA. The oxidative conversion of methane to
[20] Selim H, Gupta AK, Sassi M. Novel error propagation approach for reducing higher hydrocarbons. J Catal 1987;103(February):301–10.
H2S/O2 reaction mechanism. Appl Energy 2011. doi:10.1016/
j.apenergy.2011.01.04.

You might also like