Skip to main content

Homogeneous oxidation of SO2 in the tail gas incinerator of sulfur recovery unit

Abstract

The formation and emission of sulfur trioxide (SO3) in sulfur recovery unit has received increasing attention due to its adverse effects on the operation of plant and environment. Due to the excess oxygen, high concentration of SO2 and high temperature, SO3 formation in the sulfur recovery unit tail gas incinerator may significantly increase. A small horizontal tube reactor was employed to simulate the homogeneous oxidation of SO2 in the tail gas incinerator. The SO3 concentration was measured with a controlled condensation method at the outlet of the reactor. The present work focuses on the gas-phase chemistry and examines the impact of different combustion parameters and atmospheres on the formation of SO3 in the tail gas incinerator. Experiment results show that the increased O2 and SO2 concentrations along with increasing temperature are favorable for enhancing SO3 formation over the range of tested parameters. The presence of water vapor has an enhancing effect of SO2 oxidation in the experiments conducted. No significant effect of CO2 was found to the oxidation of SO2.

Peer Review reports

Introduction

The oxidation of SO2 relate to the sulfur compounds combustion becomes an increasing concern. The high concentration of oxidation products such as sulfur trioxide (SO3), which could react with water molecule form H2SO4 or sulfuric acid mist, is not only a corrosion risk to equipment but also an environmental concern. In sulfur recovery unit, the tail gas stream from Claus tail gas treatment plant is predominantly N2, H2O vapor and CO2 depending on the original acid gas. This stream will also contain a number of residual sulfur compounds such as H2S, SO2, COS, CS2 and S vapor together with other components, such as CO and H2, at low concentrations. Compliance with regulatory guidelines of pollutant emission, requires the destruction of all reduced sulfur compounds in the tail gas down to low levels before allowing the stream to enter [1] the atmosphere [1, 2]. This is typically done by oxidizing the reduced sulfur compounds in the tail gas to SO2 within the plant incinerator. One of the preferred methods to incinerate the Claus tail gas stream is thermal incineration, where high temperature homogeneous gas phase reactions take place for oxidation of the reduced sulfur compounds and then all the sulphur compounds are converted to SO2. When excess oxygen is present and gas temperature is sufficiently high, a small amount of SO2 could be converted to SO3 in the flue gas [3,4,5]. SO3 is a great concern in anti-corrosion design, due to the presence of SO3 and vapor in the flue gas. Sulfuric acid can condense at very high temperature in flue gas, wherein the dew point of the sulfuric acid in the flue gas from the sulphur tail gas incinerator is typically 120–170 °C [6]. The Claus tail gas and sulphur degasser off-gases are both routed to the incinerator, resulting in relatively high concentration of SO2 in the flue gas. Due to the high concentration of O2, SO2 and water vapor in the incinerator, more SO3 tends to be formed. Higher SO3 concentration together with a mass of water vapor in the flue gas will obviously increase the acid dew point of the flue gas, increasing the risk of corrosion on the equipment. Care must be taken to prevent the steel shell of the equipment from getting below the dew point of the acid gas [7]. Beyond that, sulfuric acid in flue gas is an environmental concern, and a small amount of SO3 emission may cause serious environmental problems. With the decrease in temperature, the H2SO4 vapor condenses to form acid mist, leading to acid deposition [8], and furthermore, sulfuric acid mist is often the cause of the blue haze that often appears as the flue gas plume dissipates.

Many studies have been done to explore the formation of SO3 in combustion process by both experiment method and theoretical modeling [9,10,11,12,13,14]. In principle, SO3 formation from SO2 oxidation during the combustion process results from either homogeneous gas reactions or heterogeneous catalytic reactions. Sulfur trioxide is thermodynamically favored at lower temperatures, but kinetic limitations and short reaction time often prevent an SO3/SO2 partial equilibrium from being attained [10]. From the research work of Hindiyarti et al. [15] two reaction routes are recognized as the dominant pathways in the oxidation of SO2 in gas phase reaction.

The primary reaction is the direct reaction between SO2 and oxygen radical at temperature higher than 900 ℃:

$${\text{SO}}_{2} + {\text{O}}( + {\text{M}}) \Leftrightarrow {\text{SO}}_{3} ( + {M})$$
(1)

The other reaction path occurs under moist atmosphere, where oxidation of SO2 is strengthened in the presence of water vapor which increases the O/H radical concentration:

$${\text{SO}}_{2} + {\text{OH}}( + {M}) \Leftrightarrow {\text{HOSO}}_{2} ( + {M})$$
(2)
$${\text{HOSO}}_{2} + {\text{O}}_{2} \Leftrightarrow {\text{SO}}_{3} + {\text{HO}}_{2}$$
(3)

The first reaction path is the main source of SO3 formation at high temperature, and the second reaction path contributes mostly to the generation of SO3 under moist atmosphere at low temperature. HOSO2 is unstable at high temperature, and production of SO3 via HOSO2 is insignificant at temperatures of 1000 K and above [13, 16, 17].

Due to the significantly increasing concentration of SO2 and oxygen compared to air firing, lots of research works focus on the SO3 formation in oxy-fuel circulating fluidized bed [4, 14, 18,19,20,21]. SO3 concentration measurement under oxy-fuel condition also indicates an increase of SO3 concentration in flue gas [5]. Compared with the conditions in oxy-fuel combustion, because the Claus tail gas and sulphur degasser off-gases are both routed to the incinerator, the concentration of SO2 in tail gas incinerator are around 1% vol, higher than that in oxy-fuel combustion [14, 22]. Since limited work has been published focusing on the generation of SO3 in incinerators of the sulfur recovery unit, it is necessary to have a deep look at the formation of SO3 in the tail gas incinerator of the sulfur recovery unit.

In the present work, the focus is to investigate the homogeneous gas phase SO3 generation in the tail gas incinerator of the sulfur recovery unit under different combustion conditions. The effects of concentration of reactant gas, reaction temperature and the presence of impurities in flue gas were investigated on the formation of SO3.

Experimental section

The experimental setup consists of a gas feeding system, a mixing chamber, a reactor system, an acid mist condenser and a tail gas treatment unit, as shown in Fig. 1. The gas feeding system consists of gas cylinders that are connected to the reactor system with mass flow controllers from the Beijing Sevenstar Electronics Co., Ltd. to the reactor system. High purity air, CO2, SO2, N2 (21% Oxygen in N2, 99.99% CO2, 10% SO2, 99.99%N2,) are used to simulate the reaction gas, N2 is the balance gas. The concentration of O2, SO2, CO2 in the reaction gas are adjusted by controlling the flowrate of corresponding mass flow controller. The controllers are adjusted before the experiment using a bubble flow meter. All the reaction gases including water vapor are mixed in the mixing chamber. A water saturator maintained at a controlled temperature was used to introduce water vapor into the N2 carrier gas. The amount of water vapor introduced to the reaction system is measured by the loss of weight of evaporated water. The concentration of water vapor is controlled by the flowrate of carrier gas. According to the actual condition in the tail gas incinerator of the sulfur recovery unit, the concentration of SO2 and oxygen in the tube reactor were set to be around 1% vol and 4% vol respectively. The reactor system consists of a quartz glass tube flow reactor that is inserted for the oxidation of SO2 [11]. The tube reactor characterized with 16 mm inner diameter and 60 cm total length, is located in a one zone electrically heated horizontal furnace with an isothermal length of 18 cm. According to the residence time of flue gas in tail gas incinerator, the residence time of reaction gas in the isothermal zone was set to be 2 s at the actual reaction temperature, and the total gas flow rate into the reactor was adjusted to meet the residence time, typically 263 ml/min (@1 atm, 273 K) for the reaction temperature at 1123 K. The temperatures in the tube reactor were measured under inert atmosphere with a type K thermoelectric couple (± 5 K). The gas line behind the outlet of reactor was electrically heated with a heating tape to a temperature around 250 ℃ for the purpose of avoiding condensation [10].

Fig. 1
figure 1

Experimental setup for homogeneous SO2 oxidation experiments

To determine the oxidation of SO2 in the reactor, the controlled condensation method (CCM) was employed to measure the concentration of SO3 at the outlet of tube reactor. SO3 was extremely reactive in flue gas, with the decrease of flue gas temperature, and water vapor and SO3 beginning to form vapor sulfuric acid. When the temperature of flue gas below is 200 ℃, almost all the SO3 is converted to H2SO4 in the humid condition of typical flue gas [23]. With the continuous decrease of flue gas temperature, once the temperature was below acid dew point, the gaseous H2SO4 began to condense to sulfuric acid. The temperature of gas line between the outlet of the tube reactor and the inlet of the condenser was controlled around 250 ℃ to avoid condensation. The temperature of flue gas in the condenser was set around 85 ℃ which was above the dew point of water vapor to prevent the condensation of water and associated SO2 capture by condense water, but also kept below the acid dew point which allow the condensation of sulfuric acid. With this approach, SO3 in the flue gas was selectively condensed and the interference of high concentration of SO2 in the flue gas is avoided. Moreover, about 0.5 g quartz wool is placed inside the condenser to improve the recovery rate of acid mist. When exploring the reaction between SO2 and O2, water vapor is needed to condense the SO3 in the flue gas, as shown in Fig. 1. A three-way valve was used at the outlet of the tube reactor for introducing water vapor into the flue gas, the same as the introducing of water vapor into the tube reactor. Furthermore, the total gas flow rate into the condenser can be adjusted through the three-way valve. All the experiments in this study, the total gas flow rate is set to be 1.5 L/min. Two measurements are conducted under one experiment condition, the average value of measurements is used as one data point.

The sulfuric acid present in the condenser was flushed by a known quantity of distilled deionized water. The conversion of SO2 to SO3 in the reactor was determined by the amount of H2SO4 in the condenser. An ion chromatogram analyzer (ICS-1100, Dionex) was used to determine the concentration of H2SO4 in the distilled deionized water. The measurement error of ion chromatogram analyzer is within 5%.

To evaluate the accuracy of the SO3 collecting system, a SO3 sampling system has been built in laboratory, as shown in Fig. 2, where dilute sulfuric acid has been evaporated to simulate the sulfuric acid mist in the reaction gas. In general, almost all the SO3 convert to sulfuric acid mist in the snake condenser. Therefore, the concentration of SO3 seem same as the concentration of sulfuric acid mist. The collecting efficiency of the condenser is defined as the ratio between the amount of sulfuric acid mist collect in the condenser and the amount of sulfuric acid introduced to the evaporator.

Fig. 2
figure 2

The diagram of SO3 collecting system built in laboratory

In the sampling experiments, the concentration of SO2, O2 and CO2 in the gas is set to be 1%, 4% and 20%, N2 is the balance gas. The flowrate of the gas is set at 1.5 L/min, temperature of evaporator is set around 400 ℃, the temperature of heating tape and condenser is set to be 250 and 85 ℃, the flow rate of dilute acid introduced into the evaporator is set at 0.24 ml/min. The adjustment of the SO3 concentration in the gas could be realize by changing the concentration of dilute acid introduced to the system. The concentration of dilute acid introduced into the evaporator is about 5 mM to 30 mM. The average value of two measurements under one experiment condition is used as one data point. As shown in Fig. 3, the collecting efficiency of sampling system is relatively stable with the change of SO3 concentration, the collecting efficiency is around 90%.

Fig. 3
figure 3

The collecting efficiency of SO3 collecting system with different SO3 concentration

Results and discussion

Effect of SO2 concentration

To study the effect of input SO2 concentration on the output SO3 concentration in the flue gas, the conversion of SO2 to SO3 and the order of reaction with respect to the SO2 concentration were obtained. Experiments were conducted at temperature 850 ℃ with a fixed oxygen concentration of 4 vol% and by varying the SO2 concentration among 0.52, 0.75, 0.87, and 1.31 vol%.

When the compositions in the reaction are far away from equilibrium, the reversible reaction can be ignored, and under such conditions, the rate of the generation of SO3 can be expressed by the following empirical equation:

$$r_{{{\text{so}}_{3} }} = k_{f} {\text{C}}_{{{\text{SO}}_{2} }}^{n} {\text{C}}_{{{\text{O}}_{{2}} }}^{m} = k_{f} P_{{{\text{SO}}_{{2}} }}^{n} P_{{{\text{O}}_{{2}} }}^{m} /[RT]^{m + n}$$
(4)

where n and m are reaction orders with respect to SO2 and O2 concentrations. The oxidation of SO2 was less than 2% under our experiment conditions, its concentration can be assumed to be constant (i.e.,\(P_{{{\text{SO}}_{{2}} }}\) = \(P_{{{\text{SO}}_{2} }}^{0}\)). Additionally, the consumed oxygen calculated from the generated SO3 were calculated to be less than 1%. Therefore, for the sake of simplicity, if it is assumed that both the reactants were in excess compared to the product, the relationship \({\text{ln}}(P_{{{\text{so}}_{3} }} ) = \ln (k{\prime} t_{res} ) + m\ln (P_{{{\text{so}}_{2} }}^{0} )\) can be derived, where k´ is equal to kf times \((P_{{{\text{O}}_{2} }}^{0} )^{{\text{n}}} /[RT]^{{{\text{m}} + n - 1}}\). Hence, the slope (m) of a plot of \({\text{ln}}(P_{{SO_{3} }} )\) versus \({\text{ln}}(P_{{{\text{SO}}_{2} }}^{0} )\) would result in the order of reaction with respect to the SO2 concentration.

In Fig. 4a, the concentration of SO3 at the outlet of the reactor and the conversion of SO2 is plotted against the input SO2 concentration. SO2 concentration varied from 5200 ppm to 13,100 ppm, corresponding to the high SO2 concentration in the tail gas incinerator. As expected, the concentration of SO3 was higher with the increasing SO2 concentration in the flue gas. The concentration of SO3 at the outlet of the reactor was measured from 24 to 60 ppm in the experiments conducted. The conversion of SO2 to SO3 ranged from 0.544% for 5200 ppm of SO2 input to 0.489% for 7548 ppm of SO2, 0.484% for 8747 ppm of SO2, and 0.429% for 13,100 ppm of SO2.

Fig. 4
figure 4

a Effect of the input SO2 concentration on the output SO3 concentration in the flue gas and conversion of SO2 to SO3, b order of dependency with respect to SO2 concentration for SO3 formation reaction. Inlet gas mixture for the experiments: variable ppm of SO2, 4 vol% O2, and balance N2, at temperature of 850 ℃

The oxygenation efficiency varied with the concentration of SO2 for the SO2 oxidation are according with previous works [10, 24]. It's worth noting that from Fig. 4a, with the increase of SO2 concentration in the flue gas, a decrease of SO2 conversion is observed, which is consistent with the conclusion in the previous studies [5, 10, 12, 24]. As shown in Fig. 4b, \({\text{ln}}(P_{{{\text{SO}}_{3} }} )\) is plotted against \(\ln (P_{{{\text{SO}}_{2} }}^{0} )\), the linear fitting of the data point results in a slope of 0.98 under the conducted experimental parameters. The slope is close to unity which implies that the order of the reaction with respect to the SO2 concentration is unity. In the previous work of Forzatti et al., where the oxidation of SO2 was explored on a honeycomb catalyst, the results indicated that a first-order dependence could give a reasonable approximation of the SO2 oxidation in the range from 0 to 1000 ppm of SO2 [25, 26]. The reaction order obtained in present study is the SO2 oxidation taking place in gas phase, which is different from the previous studies [25, 26].

Effect of O2 concentration

The effect of O2 concentration on the SO2 oxidation in gas phase was studied in a manner similar to that employed for studying the effect of the SO2 concentration on the SO3 formation. The conversion of SO2 to SO3 and the order of reaction with respect to the O2 concentration were obtained. According to combustion condition in the tail gas incinerator of the sulfur recovery unit, experiments were performed at temperature 850 ℃ with a fixed SO2 concentration of 1 vol% and by varying the O2 concentration in 1, 2, 3, 4 and 5 vol%.

As shown in Fig. 5a, the concentration of SO3 at the outlet of the reactor is plotted against the input O2 concentration. As anticipated, the oxidation of SO2 is enhanced with the enrichment of oxygen, due to more molecular oxygen existing in the flue gas. As O2 concentration increases from 1 vol% to 2 vol%, the concentration of SO3 increases significantly from 28 to 34 ppm. Further increasing O2 concentration from 4 vol% to 5 vol% does not result in significant increase of SO3 formation, which is consistent with the previous work from Fleig et al. [4]. Different form the statement from previous work that at least 1% excess O2 is needed for SO2 oxidation [27], the conversion of SO2 in the atmosphere of low excess O2 is observed. In Fig. 5b, \({\text{ln}}(P_{{{\text{SO}}_{3} }} )\) is plotted against \(\ln (P_{{{\text{O}}_{{2}} }}^{{0}} )\), and the straight line fitted through the data points results in a slope of 0.25 under the conducted experimental conditions. According to the slope of the fitting data, the reaction order for the SO2 oxidation with respect to the O2 is 0.25, which is similar to the research of Fan et al. [28]. The fractional order implies that the reaction between SO2 and O2 is not an elementary reaction.

Fig. 5
figure 5

a Effect of the input O2 concentration on the output SO3 concentration in the flue gas, b order of dependency with respect to O2 concentration for SO3 formation reaction. Inlet gas mixture for the experiments: variable ppm of O2, 1 vol% SO2, and balance N2, at temperature of 850 ℃

Effect of experimental temperature

Under the standard-state conditions (i.e., 25 ℃ and 1 atm), the change in Gibbs free energy of the reaction 1 is − 71 kJ/mol [29]. Therefore, thermodynamically, the oxidation of SO2 in gas phase is feasible. However, the chemical kinetics of reaction 1 limits the conversion of SO2. The oxidation of SO2 is too slow at low temperature. According to the thermodynamic model [18], thermodynamically, complete oxidation of SO2 is expected at below 500 °C and, as the temperature is increased to 900 and 1000 °C, the conversion decreased to about 6.75 and 3.21%, respectively. To get a deeper insight into the temperature effect, the temperature dependence of the oxidation of SO2 to SO3 was investigated using a reactant mixture consisting of 1 vol % SO2, 4 vol % O2 and by varying the experiment temperature among 600, 700, 800, 900 and 1000 ℃. Figure 6 shows the result of the temperature variation, and a positive correlation was found between the SO2 oxidation and temperature. The concentration of SO3 in the flue gas increased from 13 to 101 ppm as the temperature increased from 600 to 1000 ℃ in the performed experiments. It's worth noting that water vapor is absent in the tube reactor in the experiments conducted, and the O radicals combine with SO2 corresponding to the oxidation of SO2. The conversion of SO2 is 0.13% at 600 ℃.

Fig. 6
figure 6

Effect of the experiment temperature on the output SO3 concentration in the flue gas. Inlet gas mixture for the experiments: 4 vol % of O2, 1 vol % SO2, and balance N2

Higher concentration of O radicals at high temperature contributes to the SO2 oxidation which results in conversion of 1% at 1000 ℃. This is consistent with the finding by Bayless [27].

Effect of water vapor

Generally speaking, the presence of moisture is inevitable in the flue gas in the tail gas incinerator. To evaluate the effect of moisture content on the oxidation of SO2 in gas phase, the concentration of SO3 at the outlet of the reactor was measured at water vapor contents varying form 0 vol% to 8.7 vol% in the flue gas. As shown in Fig. 7, the introducing of moisture in the reactor results in a significantly enhance of SO2 oxidation. The SO3 concentration at the outlet of the reactor is increased from 41 to 85 ppmv with the presence of water vapor.

$${\text{HO}}_{2} ( + M) \Leftrightarrow {\text{H}} + {\text{O}}_{2}^{{}} ( + M)$$
(5)
$${\text{H}} + {\text{O}}_{2} \Leftrightarrow {\text{O}} + {\text{OH}}$$
(6)
$${\text{H}}_{2} {\text{O}} + {\text{O}} \Leftrightarrow {\text{OH}} + {\text{OH}}$$
(7)
Fig. 7
figure 7

Effect of the input moisture on the output SO3 concentration in the flue gas and the conversion of SO2. Inlet gas mixture for the experiments: variable concentration of H2O, 1 vol% SO2, 4 vol% O2 and balance N2, at temperature of 850 ℃

The water vapor exist in the flue gas increases the formation of H radicals through reaction 5 and thereby the formation of O and OH radicals by reaction 6. Furthermore, the concentration of OH radicals is increased through reaction 7. The increasing concentration of OH radicals enhances the oxidation of SO2 by reaction 2. However, in the range of moisture content from 0.3% to 8.7%, all the measurements show comparable values, and no obvious dependency of the SO2 oxidation on the water vapor content is visible. In the literature from Tobias and Belo [10, 24], a similar trend was observed with the increase of moisture content. It's worth noting that due to the need to condense the SO3 in the flue gas, all the experiments in the reference literature were conducted with the presence of moisture in the reactor. Due to the introducing of water vapor at outlet of the reactor in the present work, the experiments performed under conditions without water vapor in the reactor are feasible. The obvious enhancing effect of water vapor for SO2 oxidation in gas phase was demonstrated in this study. Further work is needed to elucidate the independence of the varying moisture content for the SO2 conversion in gas phase.

Effect of CO2

Due to the considerable partial pressure of CO2 in the tail gas incinerator of the sulfur recovery unit, the effect of CO2 on the oxidation of SO2 is considered. Figure 8 shows the SO2 conversion varying with CO2 concentration. As can be seen in Fig. 8, the increasing content of CO2 has a very slight effect on the SO2 oxidation at experimental temperature. The presence of CO2 can depress the availability of O/H radicals, since the reaction between CO2 and H radicals, as shown in reaction 8, is in competition with reaction 6. Therefore, the concentration of H radical can be decreased in the presence of high CO2 concentration. The H radical is important to the production of O radicals and OH production. Thus, the SO2 conversion would be decreased in the presence of CO2. However, on the other hand, the third body efficiency of reactions 1 and 2 for CO2 is higher than that for N2, which counteracts the depressed effect occurring through radical removal by CO2 [30]. To conclude, the presence of CO2 in the flue gas influences the SO2 oxidation in both directions (production/reduction). Overall, no significant effect of CO2 was found to the oxidation of SO2.This result is consistent with previous [12] works [5].

$${\text{CO}}_{2} + {\text{H}} \Leftrightarrow {\text{CO}} + {\text{OH}}$$
(8)
Fig. 8
figure 8

Effect of the input CO2 on the conversion of SO2. Inlet gas mixture for the experiments: variable concentration of CO2, 1 vol% SO2, 4 vol% O2, 3.8% moisture and balance N2, at temperature of 850 ℃

Conclusions

The homogeneous oxidation of SO2 to SO3 was investigated using a quartz flow reactor under the combustion conditions similar to those in the tail gas incinerator of the sulfur recovery unit. The conversion of SO2 decreased with the increase of SO2 concentration. The reaction order of SO3 formation was determined to be 0.98 and 0.25 with respect to SO2 and O2. The SO2/SO3 conversion increased from 0.13% to 1% as the temperature increased from 600 to 1000 ℃. The enhancing of SO2 conversion is the result of higher concentration of O radicals at high temperature. The presence of water vapor has a significant effect to promote the conversion of SO2. SO2 oxidation was also found to be independent of the CO2 in the flue gas.

Experiments were designed to determine the effect of different combustion conditions on SO3 formation. The results of this research would provide an idea to reduce the production of SO3 in tail gas combustor (such as the Claus installations). In the future, the effect of the combustion chamber structure combine with the operating parameter of combustor on the SO3 formation maybe the research direction.

Availability of data and materials

All data generated or analysed during this study are included in this published article.

References

  1. Laryea-Goldsmith R, Oakey J, Simms NJ. Gaseous emissions during concurrent combustion of biomass and non-recyclable municipal solid waste. Chem Cent J. 2011;5(1):4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Liang F-Y, Ryvak M, Sayeed S, Zhao N. The role of natural gas as a primary fuel in the near future, including comparisons of acquisition, transmission and waste handling costs of as with competitive alternatives. Chem Cent J. 2012;6(1):S4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Xiang B, Shen W, Zhang M, Yang H, Lu J. Effects of different factors on sulfur trioxide formations in a coal-fired circulating fluidized bed boiler. Chem Eng Sci. 2017;172:262–77.

    Article  CAS  Google Scholar 

  4. Fleig D, Andersson K, Normann F, Johnsson F. SO3 formation under oxyfuel combustion conditions. Ind Eng Chem Res. 2011;50(14):8505–14.

    Article  CAS  Google Scholar 

  5. Duan L, Duan Y, Sarbassov Y, Li Y, Anthony EJ. SO3 formation under oxy-CFB combustion conditions. Int J Greenhouse Gas Control. 2015;43:172–8.

    Article  CAS  Google Scholar 

  6. Zheng C, Wang Y, Liu Y, Yang Z, Qu R, Ye D, Liang C, Liu S, Gao X. Formation, transformation, measurement, and control of SO3 in coal-fired power plants. Fuel. 2019;241:327–46.

    Article  CAS  Google Scholar 

  7. Bahadori A. Estimation of combustion flue gas acid dew point during heat recovery and efficiency gain. Appl Therm Eng. 2011;31(8):1457–62.

    Article  CAS  Google Scholar 

  8. Breeze P. Chapter 4 - combustion plant emissions: sulfur dioxide, nitrogen oxides, and acid rain. In: Breeze P, editor. Electricity generation and the environment. Cambridge: Academic Press; 2017. p. 33–47.

    Chapter  Google Scholar 

  9. Hwang SM, Cooke JA, Witt KJD, Rabinowitz MJ. Determination of the rate coefficients of the SO2 + O + M → SO3 + M reaction. Int J Chem Kinet. 2010;42(3):168–80.

    Article  CAS  Google Scholar 

  10. Belo LP, Elliott LK, Stanger RJ, Spörl R, Shah KV, Maier J, Wall TF. High-temperature conversion of SO2 to SO3: homogeneous experiments and catalytic effect of fly ash from air and oxy-fuel firing. Energy Fuels. 2014;28(11):7243–51.

    Article  CAS  Google Scholar 

  11. Jørgensen TL, Livbjerg H, Glarborg P. Homogeneous and heterogeneously catalyzed oxidation of SO2. Chem Eng Sci. 2007;62(16):4496–9.

    Article  Google Scholar 

  12. Fleig D, Alzueta MU, Normann F, Abián M, Andersson K, Johnsson F. Measurement and modeling of sulfur trioxide formation in a flow reactor under post-flame conditions. Combust Flame. 2013;160(6):1142–51.

    Article  CAS  Google Scholar 

  13. Alzueta MU, Bilbao R, Glarborg P. Inhibition and sensitization of fuel oxidation by SO2. Combust Flame. 2001;127(4):2234–51.

    Article  CAS  Google Scholar 

  14. Wang X, Zhang J, Wang Z, Wang Y, Vujanović M, Li P, Tan H. Experimental and kinetics study on SO3 catalytic formation by Fe2O3 in oxy-combustion. J Environ Manag. 2019;236:420–7.

    Article  CAS  Google Scholar 

  15. Hindiyarti L, Glarborg P, Marshall P. Reactions of SO3 with the O/H radical pool under combustion conditions. J Phys Chem A. 2007;111(19):3984–91.

    Article  CAS  PubMed  Google Scholar 

  16. Fulle D, Hamann HF, Hippler H. The pressure and temperature dependence of the recombination reaction HO+SO2+M→HOSO2+M. Phys Chem Chem Phys. 1999;1(11):2695–702.

    Article  CAS  Google Scholar 

  17. Blitz MA, Hughes KJ, Pilling MJ. Determination of the high-pressure limiting rate coefficient and the enthalpy of reaction for OH + SO2. J Phys Chem A. 2003;107(12):1971–8.

    Article  CAS  Google Scholar 

  18. Sarbassov Y, Duan L, Jeremias M, Manovic V, Anthony EJ. SO3 formation and the effect of fly ash in a bubbling fluidised bed under oxy-fuel combustion conditions. Fuel Process Technol. 2017;167:314–21.

    Article  CAS  Google Scholar 

  19. Fleig D, Andersson K, Johnsson F, Leckner B. Conversion of sulfur during pulverized oxy-coal combustion. Energy Fuels. 2011;25(2):647–55.

    Article  CAS  Google Scholar 

  20. Tan Y, Croiset E, Douglas MA, Thambimuthu KV. Combustion characteristics of coal in a mixture of oxygen and recycled flue gas. Fuel. 2006;85(4):507–12.

    Article  CAS  Google Scholar 

  21. Stanger R, Wall T. Sulphur impacts during pulverised coal combustion in oxy-fuel technology for carbon capture and storage. Prog Energy Combust Sci. 2011;37(1):69–88.

    Article  CAS  Google Scholar 

  22. Croiset E, Thambimuthu KV. NOx and SO2 emissions from O2/CO2 recycle coal combustion. Fuel. 2001;80(14):2117–21.

    Article  CAS  Google Scholar 

  23. Reiner T, Arnold F. Laboratory flow reactor measurements of the reaction SO3 + H2O + M → H2SO4 + M: implications for gaseous H2SO4 and aerosol formation in the plumes of jet aircraft. Geophys Res Lett. 1993;20(23):2659–62.

    Article  CAS  Google Scholar 

  24. Schwaemmle T, Heidel B, Brechtel K, Scheffknecht G. Study of the effect of newly developed mercury oxidation catalysts on the DeNOx-activity and SO2–SO3-conversion. Fuel. 2012;101:179–86.

    Article  CAS  Google Scholar 

  25. Svachula J, Alemany LJ, Ferlazzo N, Forzatti P, Tronconi E, Bregani F. Oxidation of sulfur dioxide to sulfur trioxide over honeycomb DeNoxing catalysts. Ind Eng Chem Res. 1993;32(5):826–34.

    Article  CAS  Google Scholar 

  26. Forzatti P. Kinetic study of a simple chemical reaction. Catal Today. 1997;34(3):401–9.

    Article  CAS  Google Scholar 

  27. Bayless DJ, Jewmaidang J, Tanneer S, Birru R. Kinetics of low-temperature homogeneous SO3 formation for use in flue gas conditioning for improved electrostatic precipitator performance. Proc Combust Inst. 2000;28(2):2499–505.

    Article  CAS  Google Scholar 

  28. Shi Y, Fan M. Reaction kinetics for the catalytic oxidation of sulfur dioxide with microscale and nanoscale iron oxides. Ind Eng Chem Res. 2007;46(1):80–6.

    Article  CAS  Google Scholar 

  29. David RL, editor. CRC handbook of chemistry and physics. 84th ed. Boca Raton: CRC Press; 2003.

    Google Scholar 

  30. Glarborg P, Bentzen LLB. Chemical effects of a high CO2 concentration in oxy-fuel combustion of methane. Energy Fuels. 2008;22(1):291–6.

    Article  CAS  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

This work was financially supported by the Postdoctoral Project Foundation of PetroChina Southwest Oil & Gasfield Company (No. 20210306-10).

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization, ZX.; Methodology, ZX, DT; Funding acquisition, CH; Investigation, ZX; Resources, LJ; Writing—original draft, ZX, DT; Writing—review and editing, ZX and LJ; Supervision, CH. and LJ. All authors have read and agreed to the published version of the manuscript.

Corresponding author

Correspondence to Li Jinjin.

Ethics declarations

Ethics approval and content to participate

The manuscript does not contain studies with animal subjects.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Xi, Z., Tong, D., Honggang, C. et al. Homogeneous oxidation of SO2 in the tail gas incinerator of sulfur recovery unit. BMC Chemistry 17, 178 (2023). https://0-doi-org.brum.beds.ac.uk/10.1186/s13065-023-01096-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://0-doi-org.brum.beds.ac.uk/10.1186/s13065-023-01096-w

Keywords