Disclaimer: This is an example of a student written essay.
Click here for sample essays written by our professional writers.

Any scientific information contained within this essay should not be treated as fact, this content is to be used for educational purposes only and may contain factual inaccuracies or be out of date.

Influence of Pressure on H2s Oxidation: Experiments and Kinetic Modeling

Paper Type: Free Essay Subject: Chemistry
Wordcount: 5688 words Published: 8th Feb 2020

Reference this

ABSTRACT

The oxidation of H2S at different manometric pressures (0.6-40 bar), in the temperature range of 500-1000 K and under slightly oxidizing conditions (λ=2), has been studied. Experiments have been performed in a quartz tubular flow reactor. The results have shown that H2S conversion is shifted to lower temperatures as the pressure increases. The kinetic model used in this work is based on a previous one proposed by the authors to describe H2S oxidation at atmospheric pressure, which was updated with a H2/O2 reaction subset for high pressures. Model results match fairly well the experimental ones both from the present work and from the literature. The reaction pathways of H2S oxidation analyzed are similar to the ones at atmospheric pressure. The differences are found in the radicals that are involved in the oxidation process at the different pressures.

Keywords: H2S, oxidation, high pressure, sour gas, kinetic modeling

1. INTRODUCTION

As the energy demand increases worldwide, an efficient utilization of available natural resources is needed. Natural gas production is expected to peak near 2035, taking into consideration a scenario of cumulative production, plus remaining reserves, plus undiscovered resources [1]. The abundance of natural gas reserves can facilitate the transition from fossil derived to fully renewable fuels and chemical generation [2, 3]. The total long-term recoverable conventional gas resource base is more than 400 trillion cubic metres (tcm), and another 400 tcm are estimated for unconventional sources (sour gas) [4]. The increasing importance of unconventional fuel sources, such as sour and shale gas (natural gas with significant amounts of H2S and CO2, up to 30% content in volume each one [5]), brings interest to the direct use of these fuels, developing technologies and combustion processes focused on their knowledge and understanding under high pressure conditions [6]. The technology most used for the treatment of H2S is the Claus process, by producing sulfur from it [7, 8]. However, due to the large offer of sulfur worldwide [9] and the variable feedstock of H2S in natural gas reserves, different processes for treating sour gas and H2S streams are emerging [2, 10-13]. Thus, Langè et al. [14] proposed a detailed description of the phase behavior of the CH4+H2S system, in a wide range of temperatures and pressures, with the goal of performing the correct process design of new gas purification technologies. Another interesting process for exploitation of sour gas reserves is the direct gas combustion, for example, through oxy-combustion [15], which has been studied at high pressures to increase efficiency in power plants [16, 17].

Get Help With Your Essay

If you need assistance with writing your essay, our professional essay writing service is here to help!

Essay Writing Service

Some studies about mixtures of sulfur species and carbon species oxidation have been performed, with the final goal of knowing more about sour gas conversion. For example, Gersen et al. [18] experimentally studied the autoignition and oxidation of CH4/H2S mixtures in a rapid compression machine (RCM) and a flow reactor at high pressure. They showed prediction results with their model that agree well with the measured autoignition delay times. On the other hand, they also indicated that the H2S oxidation chemistry and the interaction of CH4 and H2S at high pressure are not well understood. Zeng et al. [19], in their work on the co-oxidation of CH4 and CS2 (a known impurity of the Claus process) in a flow reactor, experimentally observed an inhibiting effect by CS2 in the oxidation of methane at atmospheric pressure. They became aware of the complexity of the C-H-O-S combustion chemistry and claimed that their current model could not reflect all potentially significant reactions. Other recent studies have been devoted to validate their kinetic mechanisms on sulfur compounds oxidation with experiments from the literature. Bongartz et al. [15] and Bongartz and Ghoniem [20, 21] developed an optimized mechanism to make qualitative and even quantitative predictions on the combustion behavior of sour gas under oxy-fuel conditions. Salisu and Abhijeet [22] carried out kinetic simulations of acid gas (H2S and CO2) pyrolysis and oxidation for simultaneous syngas (H2+CO) and sulfur recovery, which results will assist in the design and optimization of acid gas conversion reactors. Despite these efforts, there is still a need for more accurate direct determination of several important rate constants as well as more validation data in order to improve modeling predictions of sour gas conversion [15].

The reaction steps in the H2S oxidation process remain unknown in many aspects and the available experimental data are limited. Some works have been carried out in the last years trying to understand H2S kinetic behavior under combustion conditions. The experimental and theoretical work of Zhou et al. [23] about H2S oxidation at atmospheric pressure established several kinetic parameters for different reactions involved in the process. Later, the exploratory work from Song et al. [24] on H2S oxidation under high pressures showed that, under oxidizing and stoichiometric conditions, the H2S oxidation depends strongly on the stoichiometry and pressure, claiming that experimental results for H2S oxidation at elevated pressures are scarce.

As H2S conversion can be affected by the operating conditions, such as temperature, pressure and combustion atmosphere, the present work addresses the oxidation of H2S at high pressures, with the basis of the work from the authors at atmospheric pressure [25], where H2S oxidation was studied in a flow reactor from reducing to oxidizing conditions and different temperatures (700-1400 K). In that study, H2S conversion was reasonably predicted by the kinetic model proposed, which included the addition of HSOO⇌HSO2 isomerization reaction, supported by recent theoretical works, as a key step in a faster reaction path of SH oxidation.

In this context, the present work studies the influence of manometric pressure (0.6-40 bar) at different temperatures (500-1000 K) and at slightly oxidizing conditions (approximately λ=2), on the conversion of hydrogen sulfide, performing flow reactor experiments, under laboratory controlled conditions. Moreover, a kinetic model capable of describing the oxidation process of H2S at high pressures has been used to interpret the experimental results. The results obtained in this work can be also useful for different industrial processes, such as the Claus process [26] or oxy-combustion of the sour gas [15, 20, 21, 27].

2. EXPERIMENTAL METHODOLOGY

The experimental set-up used to perform the high-pressure H2S oxidation experiments has been previously described in detail elsewhere [28]. Therefore, only a brief description of the main features is provided here. Reactants: H2S (approximately 500 ppm), O2 (approximately 1500 ppm) and N2 as carrier gas, were supplied from gas cylinders through mass flow controllers with an uncertainty in the flow rate measurements of approximately 0.5%. The oxygen required to carry out each oxidation experiment is determined by the air excess ratio (λ, defined as inlet oxygen divided by stoichiometric oxygen). Slightly oxidizing conditions (λ=2) were selected to study the oxidation of hydrogen sulfide at different manometric pressures. Table 1 contains the conditions for the different experiments performed. The moderate concentration of oxygen used in this work was chosen to avoid deposition of sulfur in the experimental set-up, if reducing conditions were used, and to minimize SO3 formation, which is enhanced at oxidizing conditions and could lead to corrosion problems. The reactant gases were premixed before entering the reactor, which consists of a quartz tube (inner diameter of 6 mm and 1500 mm in length) designed to approximate plug flow conditions [29]. The reactor is enclosed in a stainless-steel tube that acts as a pressure shell. The steel tube is placed horizontally in a tubular oven, with three individually controlled electrical heating elements that ensure an isothermal reaction zone of approximately 500 mm, with a uniform temperature profile (±5 K). Gas residence time depends on pressure and temperature and it can be expressed as tr(s) = 232*P(bar)/T(K) in the isothermal part of the reactor. Previously to the gas analysis systems, gases pass through a filter and a condenser to ensure gas cleaning. Products are analyzed by a gas chromatograph equipped with a thermal conductivity detector (TCD) to quantify H2S and O2, and a continuous UV analyzer to quantify SO2. The uncertainty of the measurements is estimated within 5%.

Table 1. Experimental conditions. N2 as bath gas. tr(s)=232*P(bar)/T(K).

Set

λ

Manometric pressure (bar)

H2S (ppm)

O2 (ppm)

1

1.99

0.6

505

1509

2

2.1

5

480

1520

3

2.06

10

485

1510

4

2.04

20

497

1520

5

2.06

40

500

1545

3. KINETIC MODEL

The kinetic model used in this work has been taken from a recent work by the authors, where H2S oxidation at atmospheric pressure was studied in a flow reactor, from reducing to oxidizing conditions [25]. The simulations have been carried out with the software Chemkin-Pro and the plug flow reactor code [30]. The oxidation mechanism proposed in [25] supported the evolution of the SH+O2 reaction (key reaction step in H2S oxidation) through isomerization from HSOO to HSO2, leading to the final product SO2, as proposed by Garrido et al. [31], in a high level ab initio study of the HSO2 system. The main reactions for H2S oxidation belong to the work from Zhou et al. [23] and Alzueta et al. [32]. The H2/O2 subset was updated for high pressures [33] and has been included in the present study with no modifications. However, no big differences have been observed in the simulations in this work using the updated H2/O2 subset.

The oxidation of H2S under high pressure conditions might be a source of O3, as mentioned by Song et al. [24], according to the SH+O2+O2⇌HSO+O3 reaction, and ozone may promote H2S conversion, as it is a much more reactive molecule than molecular oxygen. Thus, a subset for O3 reactions has been added from their work [24], which is mainly based on the work from Atkinson et al. [34]. Reactions involving O3 have not been found to be important under the present conditions.

Hydrogen sulfide oxidation proceeds via its reaction with the O/H radical pool to form SH, as can be seen in the reaction pathways shown in Figure 1. Under slightly oxidizing conditions, SH radicals react mainly with O2 to form the HSOO peroxide, which isomerizes to HSO2 and forms by decomposition the SO2+H final products, feeding the radical pool with H radicals. The HSO2 radicals can also react with oxygen, if it is available, or isomerize again to the radical HOSO, which will react in the same way as HSO2, decomposing or reacting with O2 to form SO2. The changes in reactivity during the H2S conversion process at different pressures are attributed to different radicals that participate depending on pressure. In the case of relatively low pressures (0.6, 5 and 10 bar manometric pressure), H2S will react mainly with radicals H (R1).

H2S + H ⇌ SH + H2   (R1)

However, as the pressure increases (20 and 40 bar), other pathways become also important, involving HO2 and OH radicals (R2 and R3).

H2S + HO2 ⇌ SH + H2O2  (R2)

H2S + OH ⇌ SH + H2O  (R3)

Formation of HO2 radicals is favored at high pressures (R4) [e.g. 18, 24, 35]. These radicals react with H2S to form SH and H2O2 (R2), and promote the oxidation via the branching reaction of H2O2 to give OH radicals (R5), which then interact again with H2S (R3).

O2 + H (+M) ⇌ HO2 (+M)  (R4)

H2O2 (+M) ⇌ OH + OH (+M) (R5)

Figure 1: Reaction pathways for the oxidation of H2S.

4. RESULTS AND DISCUSSION

The experimental (symbols) and simulated (lines) results of the concentrations of H2S, SO2 and O2 as a function of temperature are plotted from Figure 2 to Figure 6. All the experiments were carried out at slightly oxidizing conditions (λ=2), being SO2 the main product from the oxidation of H2S. The sulfur balance closes in all cases within ±5%. The conversion of H2S is shifted to lower temperatures as the pressure increases. Thus, the onset temperature for H2S conversion is 775 K in the case of 0.6 bar and 600 K at the highest pressure studied (40 bar). The consumption of O2 follows the same trend as hydrogen sulfide.

The simulated results obtained with the kinetic model used in this work reproduce well the experimental data, mainly, those obtained at near atmospheric pressure (0.6 bar manometric pressure). This fact confirms that the kinetic model is capable of predicting H2S conversion at near atmospheric pressure. For the rest of the pressures studied, the kinetic model underpredicts slightly the oxidation of H2S at the beginning of the reaction and overpredicts to some extent the H2S conversion at the end of the conversion. The exception would be the case at the highest pressure (40 bar), where a shift of 50 K temperature between experimental H2S concentrations and model predictions is seen.

Figure 2: Experimental results of H2S oxidation at 0.6 bar (Set 1 of Table 1).

Figure 3: Experimental results of H2S oxidation at 5 bar (Set 2 of Table 1).

Figure 4: Experimental results of H2S oxidation at 10 bar (Set 3 of Table 1).

.

Figure 5: Experimental results of H2S oxidation at 20 bar (Set 4 of Table 1).

Figure 6: Experimental results of H2S oxidation at 40 bar (Set 5 of Table 1).

If the results near atmospheric pressure of this work (0.6 bar manometric pressure) are compared with the ones obtained by the authors in their study of H2S oxidation at atmospheric pressure in other flow reactor [25], Figure 7, it can be observed that the results are quite similar, although a difference in the conversion onset temperature of 50 K exists, which might be due to the pressure difference (0.6 bar) or the residence time, as both experiments are carried out in different experimental installations. Previously, these set-ups operating at atmospheric pressure have already shown a similar behavior for CH3SH oxidation under stoichiometric conditions [36].

Figure 7: Concentrations of H2S vs. temperature at λ=2 and atmospheric pressure. Open symbols from [25] and solid symbols from the present work (Set 1 of Table 1).

It is worth to mention that the gas residence time in the reactor increases with the pressure. An additional analysis, through modeling results, of the effect of the residence time at two different pressures (1 and 21 bar absolute pressure) on H2S conversion can be seen in Figure 8.  In this figure, at 1 bar and 21 bar, the solid lines on the right correspond, respectively, to the gas residence time of the atmospheric pressure set-up used in [25] (tr = 194/T(K)) and the gas residence time of the present work high-pressure set-up (tr = 4872/T(K)), which are used as the reference cases. For both pressures, their reference residence time values are subsequently increased up to being approximately doubled. . As it can be observed, at 1 bar a 50 K temperature difference in H2S concentration profiles exists when the residence time is doubled. At high pressure (21 bar), the effect of the residence time is less important; in this case the solid line represents the residence time of set 4 in Table 1. Thus, under these conditions, the influence of gas residence time is less significant as the pressure increases. However, it is worth to highlight that the specific effect of the gas residence time depends on the given process studied. For example, Marrodán et al. [37], in their study about DME oxidation at high pressure, observed an important effect of the residence time at pressures even higher than those presented in Figure 8 on DME conversion when doubling the residence time at stoichiometric conditions.

Figure 8: Concentration of H2S vs. temperature at different pressures and different residence times.

A sensitivity analysis of model calculations has been performed for the highest pressure studied (40 bar), at the temperature when H2S is starting to react (675 K). In Figure 9, the sensitivity analysis for H2S shows the isomerization reaction (R6) as the most sensitive one:

HSOO ⇌ HSO2    (R6)

followed by the branching reaction of H2O2 (R5), which is comparatively a well stablished reaction. The next most important reaction is (R7):

SH + SH (+M) ⇌ H2S2 (+M)  (R7)

which has also been seen in the top 3 most sensitive reactions in works about H2S oxidation [24, 35]. In this work, the observed sensitivity has been related to the capability of H2S2 to convert HO2 radicals into H2O2, which would enhance the oxidation process and reaction (R8) is found as the fourth most sensitive reaction.

H2S2 + HO2 ⇌ HS2 + H2O2  (R8)

The formation of these species (H2S2 and HS2) is usually more important at reducing conditions. As mentioned by Zhou et al. [23], disulfur interactions as (R9) are extremely important in the ignition and propagation reactions during H2S oxidation even under fuel lean conditions, being (R7) chain terminating reaction, while (R9) is chain propagating reaction, due to the subsequent reaction of S with oxygen to form SO and then SO2.

SH + SH ⇌ H2S + S   (R9)

The same kinetic parameters for (R9) as in the work from Gao et al. [38] have been used. However, these authors mention that care should be taken in applying (R9) above 1 bar. Gao et al. also mention that H2S2 is more stable at high pressures and low temperatures than SH. Discrepancies in the model results compared to experimental ones might be related to the chemistry of disulfur species and claim for a better characterization of these reactions at high pressures.

Figure 9: Sensitivity analysis for H2S conversion, at λ=2, 40 bar and 675 K.

With the purpose of evaluating the performance of the kinetic model used in the present work, it has been tested against literature ignition delay time measurements at different pressures. Some cases from the work of Mathieu et al. [35], about the effects of H2S addition on hydrogen ignition behind reflected shock waves, have been taken, and the results are shown in Figure 10. In that study, ignition delay times were measured behind reflected shock waves for mixtures of 1% H2/1% O2, diluted in Ar and doped with various concentrations of H2S (100, 400, and 1600 ppm), over large pressure (around 1.6, 13, and 33 atm) and temperature (1045–1860 K) ranges. Their results showed a significant increase in the ignition delay time due to the addition of H2S, in some cases by a factor of 4 or more, over the baseline mixtures without H2S. This behavior was explained because H2S initially reacts before the H2 fuel does, mainly through the reaction H2S+H⇌SH+H2 (R1), thus taking H atoms away from the main branching reaction (R10), which would produce OH radicals, and thus inhibit the ignition process.

 H + O2 ⇌ OH + O  (R10)

 However, an increase in the reactivity was observed at the highest pressure investigated (33 atm) and at the temperature of 1100 K, using the highest H2S concentration investigated (1600 ppm). This fact would be in accordance with the discrepancy found in this work at high pressure (40 bar), where a higher experimental reactivity is observed. The present kinetic model fits well the ignition delay time measurements under different conditions of pressure and H2S concentration. The results from the present work and the previous study at atmospheric pressure in a flow reactor [25] suggest that the update in the kinetic model, adding the isomerization of HSOO to HSO2 as a possible step in H2S oxidation until the final product SO2, should be taken into consideration for future works on the oxidation of H2S.

Figure 10: Ignition delay time measurements vs. temperature for different experimental conditions, using a mixture of 1% H2/1% O2, diluted in Ar and doped with H2S. Experimental data are taken from the work of Mathieu et al. [35].

5. CONCLUSIONS

The present work addresses the oxidation of H2S at different manometric pressures (0.6-40 bar), in the temperature range of 500-1000 K, using a quartz tubular flow reactor. The experiments were performed at slightly oxidizing conditions (λ=2). The results show that the oxidation of H2S is shifted to lower temperatures as the pressure increases. The onset for H2S conversion starts at 775 K in the case of 0.6 bar and 600 K at the highest pressure studied (40 bar). The kinetic model about H2S oxidation at atmospheric pressure developed by the group in a previous work, together with an updated H2/O2 subset for high pressures, seems to work fairly well at different pressures. The kinetic model matches the experimental trends, except for the experiment at 40 bar, where a gap of 50 K in temperature between experimental concentrations and model predictions is observed. The sensitivity analysis performed at 40 bar indicates that the H2S conversion is mainly sensitive to the HSOO⇌HSO2 isomerization reaction, and some branching reactions involving H2S2 and H2O2.

REFERENCES

[1] Maggio G., Cacciola G., When will oil, natural gas, and coal peak?, Fuel 98 (2012) 111-123.

[2] Taifan W., Baltrusaitis J., Minireview: direct catalytic conversion of sour natural gas (CH4 + H2S + CO2) components to high value chemicals and fuels, Catal. Sci. Technol. 7 (2017) 2919-2929.

[3] Mac Kinnon M.A., Brouwer J., Samuelsen S., The role of natural gas and its infrastructure in mitigating greenhouse gas emissions, improving regional air quality, and renewable resource integration, Prog. Energy Combust. Sci. 64 (2018) 62-92.

[4] IEA International Energy Agency, Are we entering a golden age of gas?, World Energy Outlook, 2011.

[5] Hammer G., Lübcke T., Kettner R., Pillarella M.R., Recknagel H., Commichau A., Neumann H.-J., Paczynska-Lahme B., Ullmann’s Encyclopedia of Industrial Chemistry, Wiley-VCH: Weinheim, Germany, 2012, Vol. 23; Chapter Natural Gas, 739-792.

[6] US Department of Energy, Report of basic research needs for clean and efficient combustion of 21st century transportation fuels, 2006.

[7] Zarei S., Ganji H., Sadi M., Rashidzadeh M., Thermo-kinetic modeling and optimization of the sulfur recovery unit thermal stage, Appl. Therm. Eng. 103 (2016) 1095-1104.

[8] Nabikandi N.J., Fatemi S., Kinetic modelling of a commercial sulfur recovery unit based on Claus straight through process: Comparison with equilibrium model, J. Ind. Eng. Chem. Res. 30 (2015) 50-63.

[9] Rappold T.A., Lackner K.S., Large scale disposal of waste sulfur: From sulfide fuels to sulfate sequestration, Energy 35 (2010) 1368-1380.

[10] AuYeung N., Yokochi A.F.T., Steam reformation of hydrogen sulfide, Int. J. Hydrogen Energy 38 (2013) 6304-6313.

[11] Villasmil W., Steinfeld A., Hydrogen production by hydrogen sulfide splitting using concentrated solar energy – Thermodynamics and economic evaluation, Energy Convers. Manag. 51 (2010) 2353-2361.

[12] Todd Kappauf, Edward A. Fletcher, Hydrogen and sulfur from hydrogen sulfide-VI. Solar thermolysis, Energy 14 (1989) 443-449.

[13] Salman O.A., Bishara A., Marafi A., An alternative to the claus process for treating hydrogen sulfide, Energy 12 (1987) 1227-1232.

[14] Langè S., Campestrini M., Stringari P., Phase behavior of system methane + hydrogen sulfide, AIChE Journal 62 (2016) 4090-4108.

[15] Bongartz D., Shanbhogue S.J., Ghoniem A.F., Formation and control of sulfur oxides in sour gas oxy-combustion: prediction using a reactor network model, Energy Fuels 29 (2015) 7670-7680.

[16] Gopan A., Kumfer B.M., Axelbaum R.L., Effect of operating pressure and fuel moisture on net plant efficiency of a staged, pressurized oxy-combustion power plant, Int. J. Greenh. Gas Control 39 (2015) 390-396.

[17] Wang X., Adeosun A., Yablonsky G., Gopan A., Du P., Axelbaum R.L., Synergistic SOx/NOx chemistry leading to enhanced SO3 and NO2 formation during pressurized oxy-combustion, React. Kinet. Mech. Catal. 123 (2018) 313-322.

[18] Gersen S., van Essen M., Darmeveil H., Hashemi H., Rasmussen C.T., Christensen J.M., Glarborg P., Levinsky H., Experimental and modeling investigation of the effect of H2S addition to methane on the ignition and oxidation at high pressures, Energy Fuels 31 (2017) 2175-2182.

[19] Zeng Z., Dlugogorski B.Z., Oluwoye I., Altarawneh M., Co-oxidation of methane (CH4) and carbon disulfide (CS2), Proc. Combust. Inst. 37 (2019) 677-685.

[20] Bongartz D., Ghoniem A.F., Impact of sour gas composition on ignition delay and burning velocity in air and oxy-fuel combustion, Combust. Flame 162 (2015) 2749-2757.

[21] Bongartz D., Ghoniem A.F., Chemical kinetics mechanism for oxy-fuel combustion of mixtures of hydrogen sulfide and methane, Combust. Flame 162 (2015) 544-553.

[22] Salisu I., Abhijeet R., Kinetic simulation of acid gas (H2S and CO2) destruction for simultaneous syngas and sulfur recovery, Ind. Eng. Chem. Res. 55 (2016) 6743-6752.

[23] Zhou C.R., Sendt K., Haynes B.S., Experimental and kinetic modelling study of H2S oxidation, Proc. Combust. Inst. 34 (2013) 625-632.

[24] Song Y., Hashemi H., Christensen J.M., Zou C., Haynes B.S., Marshall P., Glarborg P., An exploratory flow reactor study of H2S oxidation at 30-100 bar, Int. J. Chem. Kinet. 49 (2017) 37–52.

[25] Colom-Díaz J.M., Abián M., Ballester M.Y., Millera Á., Bilbao R., Alzueta M.U., H2S conversion in a tubular flow reactor: Experiments and kinetic modeling, Proc. Combust. Inst. 37 (2019) 727-734.

[26] Haynes B.S., Combustion research for chemical processing, Proc. Combust. Inst. 37 (2019) 1-32.

[27] Toftegaard M.B., Brix J., Jensen P.A., Glarborg P., Jensen A.D., Oxy fuel combustion of solid fuels, Prog. Energy Combust. Sci. 36 (2010) 581-625.

[28] Marrodán L., Millera Á., Bilbao R., Alzueta M.U., High-pressure study of methylformate oxidation and its interaction with NO, Energy Fuels 28 (2014) 6107-6115.

[29] Rasmussen C.L., Hansen J., Marshall P., Glarborg P., Experimental measurements and kinetic modeling of CO/H2/O2/NOx conversión at high pressure, Int. J. Chem. Kinet. 40 (2008) 454-580.

[30] CHEMKIN-PRO 15151, Reaction Design, San Diego, (2013).

[31] Garrido J.D., Ballester M.Y., Orozco-González Y., Canuto S., CASPT2 Study of the Potential Energy Surface of the HSO2 System, J. Phys. Chem. A 115 (2011) 1453-1461.

[32] Alzueta M.U., Bilbao R., Glarborg P., Inhibition and sensitization of fuel oxidation by SO2, Combust. Flame 127 (2001) 2234-2251.

[33] Colom-Díaz J.M., Millera Á., Bilbao R., Alzueta M.U., High pressure study of H2 oxidation and its interaction with NO, Int. J. Hydrog. Energy 44 (2019) 6325-6332.

[34] Atkinson R., Baulch D.L., Cox R.A., Crowley J.N., Hampson R.F., Hynes R.G., Jenkin M.E., Rossi M.J., Troe J., Evaluated kinetic and photochemical data for atmospheric chemistry: Volume I – gas phase reactions of Ox, HOx, NOx and SOx species, Atmos. Chem. Phys. 4 (2004) 1461-1738.

[35] Mathieu O., Deguillaume F., Petersen E.L., Effects of H2S addition on hydrogen ignition behind reflected shock waves: Experiments and modeling, Combust. Flame 161 (2014) 23-36.

[36] Alzueta M.U., Pernía R., Abián M., Millera Á., Bilbao R., CH3SH conversion in a tubular flow reactor. Experiments and kinetic modelling, Combust. Flame 203 (2019) 23-30.

[37] Marrodán L., Arnal A.J., Millera Á., Bilbao R., Alzueta M.U., The inhibiting effect of NO addition on dimethyl ether high-pressure oxidation, Combust. Flame 197 (2018) 1-10.

[38] Gao Y., Zhou C.R., Sendt K., Haynes B.S., Marshall P., Kinetic and modeling studies of the reaction S+H2S, Proc. Combust. Inst. 33 (2011) 459-465.

 

Cite This Work

To export a reference to this article please select a referencing stye below:

Reference Copied to Clipboard.
Reference Copied to Clipboard.
Reference Copied to Clipboard.
Reference Copied to Clipboard.
Reference Copied to Clipboard.
Reference Copied to Clipboard.
Reference Copied to Clipboard.

Related Services

View all

DMCA / Removal Request

If you are the original writer of this essay and no longer wish to have your work published on UKEssays.com then please: