290
Oxidation chemistry of carbon disulfide (CS 2 ) and its interaction with hydrocarbons in combustion processes Zhe Zeng, BEng This thesis is presented for the degree of Doctor of Philosophy School of Engineering and Information Technology, Murdoch University, Western Australia 2017

Oxidation chemistry of carbon disulfide (CS and its ... · Kamal Siddique, Arif Abdullah, Jomana Al-Nu’airat, Sidra Jabeen, Dr Anam Saeed, Dr Juita and Dr Jakub Skut. Courtesy of

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

  • Oxidation chemistry of carbon disulfide (CS2)

    and its interaction with hydrocarbons in

    combustion processes

    Zhe Zeng, BEng

    This thesis is presented for the degree of

    Doctor of Philosophy

    School of Engineering and Information Technology,

    Murdoch University, Western Australia

    2017

  • i

    Statement of originality

    I declare that this thesis is my own account of my research and contains as its main content

    work which has not previously been submitted for a degree at any tertiary education institution.

    The thesis contains no material previously published or written by another person, except

    where due reference has been made in text.

    Zhe Zeng

    June, 2017

  • ii

    Supervisory statement

    We, the undersigned, attest that Higher Research Degree candidate, Zhe Zeng, has devised and

    synthesised the experimental program, conducted experiments, analysed data, performed

    computational quantum-mechanical calculations and has written all papers included in this

    thesis. Professor Bogdan Z. Dlugogorski and Dr Mohammednoor Altarawneh provided the

    necessary advice on the experimental program, project direction and assisted with the editing

    of the papers, consistent with normal supervisors-candidate relations.

    Professor Bogdan Z. Dlugogorski

    June, 2017

    Dr Mohammednoor Altarawneh

    June, 2017

  • iii

    Acknowledgment

    I would like to thank my supervisors Professor Bogdan Z. Dlugogorski and Dr Mohammednoor

    Altarawneh for their unwavering and exceptional support. Their understanding, mentorship,

    encouragement, dedication and magnanimity have provided a good fundamental driving force

    for the present study. Thank you again, for your quick correspondence, corrections and

    gestures of support.

    I am grateful to Murdoch University for the award of a postgraduate research scholarship which

    provided a valuable financial support for this research. This study has also been funded by the

    Australian Research Council (ARC), with grants of computing time from the National

    Computational Infrastructure (NCI) and the Pawsey Supercomputing Centre in Perth,

    Australia.

    I acknowledge my gratitude to my fellow student colleagues and staffs of the Fire Safety and

    Combustion Kinetics Research Laboratory: Ibukun Oluwoye, Niveen Assaf, Nassim Zeinali,

    Kamal Siddique, Arif Abdullah, Jomana Al-Nu’airat, Sidra Jabeen, Dr Anam Saeed, Dr Juita

    and Dr Jakub Skut. Courtesy of you all, I experienced an exciting and fun-filled years of

    computation and experiments.

    Special thanks to my wife Anqi Wang, who supported me through the most difficult period of

    my research. This experience is the most memorable and valuable days in my life.

  • iv

    Abstract

    This thesis presents a series of scientific studies exploring the oxidation chemistry of carbon

    disulfide (CS2) and its interaction with hydrocarbons in combustion systems. The results

    illustrate the extreme flammability of CS2 even at low temperature (self-ignition temperature

    at 363 K under ambient pressure). The thesis proposes a comprehensive oxidation mechanism

    that works over a wide range of atmospheric and combustion conditions with and without

    moisture. The thesis also provides experimental validation on the promotion of CS2 on the

    ignition of methane.

    Experiments involving CS2 oxidation in combustion processes have been conducted with

    tubular-flow (TFR) and jet-stirred (JSR) reactors to provide experimental validation for the

    proposed mechanism. Online Fourier transform infrared (FTIR) spectroscopy served to

    identify and quantitate the product species and to obtain the detailed species conversion profiles

    during the combustion process. Low ignition temperature of CS2 of 860 K in TFR and 710 K

    JSR, with residence time at 0.3 s under ambient pressure, implies the extreme flammability of

    CS2. The presence of moisture exhibits no effect on the oxidation of CS2 for ignition

    temperature and species profile at below 1100 K, although moisture converts CO into CO2 at

    higher temperature (> 1200 K). Co-oxidation experiments of CH4/CS2/O2 in JSR illustrate the

    promotion effect of CS2 on the ignition of methane.

    Quantum calculations afford the investigation of primary steps governing the OH-initiated

    oxidation of CS2 in the atmosphere. We also propose a comprehensive oxidation mechanism

    of CS2 in combustion systems. The thesis suggests the intersystem crossing (ISC) between

    triplet and singlet pathways to explain the extreme flammability of CS2, in analogy to the

  • v

    controlling steps operating for other reduced sulfur species. DFT calculations examine the

    interplay between the SH radical and C1 - C4 hydrocarbons to demonstrate the distinct

    inhibition and promotion effects of H2S/SH on alkanes and alkenes/alkynes in pyrolysis

    processes. The findings reported in this study apply to both atmospheric and combustion

    systems. Especially, the developed mechanism provides improved understanding of the

    oxidation of fossil fuels containing sulfur species.

  • vi

    Table of contents

    Statement of originality ............................................................................................................ i

    Supervisory statement ............................................................................................................. ii

    Acknowledgment .................................................................................................................... iii

    Abstract .................................................................................................................................... iv

    Table of contents ..................................................................................................................... vi

    List of publications ................................................................................................................. xii

    Chapter 1. Introduction .......................................................................................................... 1

    1.1. Research Background ...................................................................................................... 2

    1.2. Project objectives and thesis outline ............................................................................... 3

    Reference ................................................................................................................................ 7

    Chapter 2. Literature review ................................................................................................ 10

    2.1. Introduction ................................................................................................................... 11

    2.2. Oxidation of H2S ........................................................................................................... 14

    2.2.1. Atmospheric oxidation of H2S ................................................................................ 14

    2.2.2 Combustion mechanism of H2S ............................................................................... 20

    2.3. Oxidation of CS2 ........................................................................................................... 27

    2.3.1. Atmospheric oxidation of CS2 ................................................................................ 28

    2.3.2. Combustion mechanism of CS2 .............................................................................. 32

    2.3.3. Production of S atom in excited state in photolysis of reduced sulfur species ....... 40

  • vii

    2.4. Interaction between reduced sulfur species and hydrocarbons ..................................... 43

    2.4.1. Influence of H2S/SH on the combustion process of hydrocarbons ........................ 44

    2.4.2. Impact of SO2 on ethylene pyrolysis ...................................................................... 46

    2.5. Conclusion ..................................................................................................................... 48

    Reference .............................................................................................................................. 52

    Chapter 3. Methodology ........................................................................................................ 66

    3.1. Experimental methodology ........................................................................................... 67

    3.1.1. Gas feeding system ................................................................................................. 67

    3.1.2. Tubular-flow reactor (TFR) .................................................................................... 68

    3.1.3. Jet-stirred reactor (JSR) .......................................................................................... 69

    3.1.4. Online analytical technique .................................................................................... 71

    3.2. Computational methodology ......................................................................................... 72

    3.2.1. Quantum chemistry calculation .............................................................................. 72

    3.2.2. Kinetic modelling ................................................................................................... 77

    Reference .............................................................................................................................. 79

    Chapter 4. Atmospheric oxidation of carbon disulfide (CS2) ............................................ 80

    Abstract ................................................................................................................................ 81

    4.1. Introduction ................................................................................................................... 82

    4.2. Computational Methodology ........................................................................................ 84

    4.3. Results and Discussion .................................................................................................. 85

    4.3.1. CS2 + OH ................................................................................................................ 85

  • viii

    4.3.2. S-adduct SCS(OH) + O2 ......................................................................................... 92

    4.3.3. C-adduct SC(OH)S + O2 ........................................................................................ 96

    4.4. Conclusion ..................................................................................................................... 98

    Acknowledgement ................................................................................................................ 98

    Reference .............................................................................................................................. 99

    Chapter 5. Inhibition and promotion of pyrolysis by hydrogen sulfide (H2S) and

    sulfanyl radical (SH) ............................................................................................................ 102

    Abstract .............................................................................................................................. 103

    5.1. Introduction ................................................................................................................. 104

    5.2. Computational Details ................................................................................................. 106

    5.3. Results and discussion ................................................................................................. 107

    5.3.1. Overview of labile H abstraction sites in C1-C4 hydrocarbons ............................ 107

    5.3.2. H abstraction from alkanes ................................................................................... 111

    5.3.3. H abstraction from alkenes and alkynes ............................................................... 118

    5.3.4. Validation of CBS-QB3 calculations ................................................................... 121

    5.3.5. Relationship between BDH of abstracted C‒H bond and activation enthalpy of SH

    + hydrocarbon reaction ................................................................................................... 123

    5.3.6. Activity of SH radical compared with those of OH, NH2 and HO2 ..................... 125

    5.4. Conclusion ................................................................................................................... 126

    Appendix ............................................................................................................................ 127

    Acknowledgement .............................................................................................................. 127

    Reference ............................................................................................................................ 128

  • ix

    Chapter 6. Flammability of CS2 and other reduced sulfur species ................................. 132

    Abstract .............................................................................................................................. 133

    6.1. Introduction ................................................................................................................. 134

    6.2. Methodology ............................................................................................................... 137

    6.2.1. Experiments with tubular flow reactor ................................................................. 137

    6.2.2. Kinetic modelling ................................................................................................. 140

    6.3. Results and discussion ................................................................................................. 141

    6.3.1. Experimental results from tubular flow reactor .................................................... 141

    6.3.2. Sensitivity analysis of CS2 consumption .............................................................. 146

    6.3.3. CS2/O2 subset ........................................................................................................ 147

    6.3.4. S/O2 subset ............................................................................................................ 153

    6.3.5. Updated mechanism and modelling validation of our experimental results ........ 155

    6.4. Conclusion ................................................................................................................... 156

    Acknowledgement .............................................................................................................. 157

    Reference ............................................................................................................................ 158

    Chapter 7. Combustion chemistry of carbon disulfide (CS2) .......................................... 164

    Abstract .............................................................................................................................. 165

    7.1. Introduction ................................................................................................................. 166

    7.2. Experimental ............................................................................................................... 168

    7.3. Kinetic Modelling ....................................................................................................... 171

    7.4. Results and Discussion ................................................................................................ 176

  • x

    7.4.1 Experimental results with JSR ............................................................................... 176

    7.4.2. Kinetic modelling of dry oxidation of CS2 and revision of COS/O2 subset ......... 180

    7.4.3. The influence of moisture on CS2 oxidation ........................................................ 187

    7.5. Conclusion ................................................................................................................... 191

    Acknowledgement .............................................................................................................. 192

    Appendix ............................................................................................................................ 192

    Reference:........................................................................................................................... 193

    Chapter 8. Enhanced oxidation of CH4 in presence of CS2 ............................................. 198

    Abstract .............................................................................................................................. 199

    8.1 Introduction .................................................................................................................. 200

    8.2 Methodology ................................................................................................................ 202

    8.2.1 Experimental set-up ............................................................................................... 202

    8.2.2 Kinetic modelling .................................................................................................. 204

    8.3 Results and discussion .................................................................................................. 204

    8.3.1 Sole oxidation of CS2/O2 and CH4/O2 ................................................................... 204

    8.3.2 Co-oxidation of CS2/CH4/O2 ................................................................................. 206

    8.3.3 Kinetic modelling of the co-oxidation of CS2/CH4/O2 .......................................... 209

    8.4 Conclusion .................................................................................................................... 213

    Reference ............................................................................................................................ 214

    Chapter 9. Conclusion and recommendation .................................................................... 217

    9.1. Conclusion ................................................................................................................... 218

  • xi

    9.2 Recommendation .......................................................................................................... 220

    Appendix A for chapter 5 .................................................................................................... 223

    Appendix B for chapter 7 .................................................................................................... 243

    Appendix C for risk assessment.......................................................................................... 274

  • xii

    List of publications

    Journal articles

    1. Z. Zeng, M. Altarawneh, B. Z. Dlugogorski, Atmospheric oxidation of carbon

    disulfide (CS2), Chem. Phys. Lett., 669 (2017) 43-48.

    2. Z. Zeng, M. Altarawneh, I. Oluwoye, P. Glarborg, B. Z. Dlugogorski, Inhibition and

    promotion of pyrolysis by hydrogen sulfide (H2S) and sulfanyl radical (SH), J. Phys.

    Chem. A, 120 (2016) 8941-8948.

    3. Z. Zeng, B. Z. Dlugogorski, M. Altarawneh, Flammability of CS2 and other reduced

    sulfur species, Fire Saf. J., in press, DOI: https://doi.org/10.1016/j.firesaf.2017.03.073.

    Conference papers

    1. Z. Zeng, M. Altarawneh, B. Z. Dlugogorski, “Reactions of SH radical with C1-C4

    Hydrocarbons.” Proceedings of the Australian Combustion Symposium, 2015,

    Melbourne, Australia

    2. Z. Zeng, B. Z. Dlugogorski, M. Altarawneh, “Flammability of CS2 and Other

    Reduced Sulfur Species.” 12th International Symposium on Fire Safety Science (the

    12th IAFSS Symposium), 2017, Lund, Sweden

    3. Z. Zeng, M. Altarawneh, B. Z. Dlugogorski, “Enhanced oxidation of methane by

    CS2.” Asia-Pacific Conference on Combustion, 2017, Sydney, Australia.

  • Chapter 1. Introduction

    1

    Chapter 1

    Introduction

  • Chapter 1. Introduction

    2

    1.1. Research Background

    Sulfur bearing compounds represent a group of major impurities in most organic fuels [1, 2].

    In combustion, sulfur transforms primarily into oxide species, i.e., SOx, acting as the critical

    source for a series of atmospheric air contaminants. The typical sulfur content in coal varies

    from 0.4 - 4.0 wt. % [3]. In commercial diesel and gasoline, sulfur is limited to 500 ppm and

    30 ppm, respectively [4]. Likewise, pipeline-quality natural gas usually contains sulfur at

    levels of around 2,000 ppm [5]. In typical biomass, the content of sulfur varies from 0.02 - 0.3

    wt. % [2]. The enormous consumption of fossil fuels in oxidation processes, e.g., in power

    plants, engines and stoves, results in significant emission of sulfur oxides, especially in

    countries that do not require or enforce the emissions of SOx, limiting the exploitation of fossil

    and biomass fuels.

    Sulfur dioxide (SO2) constitutes the most notorious sulfur compound emitted from combustion

    of fossil fuels. However, under fuel rich combustion conditions or during thermal pyrolysis of

    fuels, the product gases may also contain reduces sulfur species, such as hydrogen sulfide

    (H2S), carbon disulfide (CS2) and carbonyl sulfide (COS). Gas processing plants and oil

    purification and gasification processes produce H2S as a side product. While CS2 and COS

    evolve as major by-products in Claus gas-desulfurising process [6, 7], which converts H2S into

    elementary solid sulfur, volcanic eruptions and bushfires [8] also contribute to the total

    atmospheric budget of CS2 and COS. CS2 exists at levels of parts per trillion (ppt) in the earth’s

    troposphere; 15 to 30 ppt in the nonurban troposphere and 100 to 200 ppt in polluted urban

    areas [9]. On the other hand, COS stands as the most abundant sulfur compound in the

    atmosphere (> 400 ppt) with a relatively long life-time of up to years [10].

  • Chapter 1. Introduction

    3

    Existing as a volatile liquid at room temperature (25 oC), CS2 represents a non-polar solvent

    which has found numerous applications in laboratories and industry, such as extractive

    metallurgy. Likewise, the production of synthetic fibres, rubbers and pesticides requires CS2

    and COS as feedstocks. Unfortunately, CS2 displays extreme flammability and high explosion

    propensity [11], properties responsible for fires in laboratories and chemical warehouses [12,

    13].

    Apart from H2S [14], the oxidation mechanism and combustion properties of CS2 and COS

    remain poorly understood. Lack of appropriate reaction mechanisms prevent detailed

    modelling of important industrial plants, such as the Claus process or separating sulfur

    impurities in the feedstocks in oil refineries [15]. To gain insight into the extreme flammability

    of CS2 and the role of sulfur compounds in oxidation of fuels, one needs not only the

    mechanisms of oxidation of CS2 and COS, but also the mechanisms describing the interaction

    between sulfur and hydrocarbon species.

    1.2. Project objectives and thesis outline

    The general objective of this study is to construct a comprehensive oxidation mechanism for

    CS2 and to investigate the reactions between sulfur species and hydrocarbons. Specifically,

    this thesis intends to:

    1. Perform accurate quantum chemistry calculations to improve the existing mechanism

    of atmospheric oxidation of CS2

  • Chapter 1. Introduction

    4

    2. Collect precise measurements of CS2 oxidation under combustion condition using

    tubular-flow (TFR) and jet-stirred (JSR) reactors

    3. Investigate the extreme flammability and low ignition temperature of CS2 with quantum

    chemistry calculations and propose further enhancements to the oxidation mechanism

    for CS2

    4. Validate the proposed mechanism against the experimental results from TFR and JSR

    through kinetic modelling

    5. Study the interaction between SH radical and C1 – C4 hydrocarbons with density

    function theory (DFT) calculations, and

    6. Explore the effect of CS2 addition on the ignition of methane with careful JSR

    experiments.

    To address the above aims, the thesis has the following structure:

    Chapter 2 reviews the literature pertinent to experiments and reaction kinetics of reduced sulfur

    species. We highlight the occurrence of intersystem crossing (ISC) in H2S and S/SO oxidation

    processes. ISC is a transition between two electronic states with different spin multiplicities

    but same energy level. In the case of sulfur chemistry, the oxidation reaction will start with

    triplet oxygen molecule and transit to singlet reaction pathway through the cross-over point on

    the potential energy surfaces, offering an alternative reaction corridor with much lower

    activation barriers. For CS2 oxidation, the high activation energy of triplet oxidation pathway

    fails to explain the low ignition temperature for the experiments conducted with the tubular-

    flow reactor, later prompting us to introduce the ISC to the CS2 combustion mechanism.

    Photolysis of reduced sulfur species (H2S, CS2 and COS) also leads to the formation of singlet

    sulfur atom (1S). This chapter also examines the interaction between sulfur species and

  • Chapter 1. Introduction

    5

    hydrocarbons, summarising 1) influence of H2S/SH on hydrocarbons combustion process and

    2) impact of SO2 on the pyrolysis of ethylene.

    Chapter 3 covers the experimental and theoretical methodologies applied in this project. It

    depicts the experimental set-up constructed in this work to study the oxidation of CS2 and its

    interaction with hydrocarbons, including the reactant feeding system, details of TFR and JSR

    and online analytical measurements. This chapter also comprises a description of the quantum

    chemistry calculations and kinetic modelling, including the software used in this thesis:

    Gaussian 09, MESMER 3.0, ChemRate, KiSTheIP and Chemkin Pro.

    Chapter 4 investigates primary steps governing the OH-initiated atmospheric oxidation of CS2

    with theoretical calculations. In the absence of surface effects, we find the overall reaction OH

    + CS2 → COS + SH too slow to account for the formation of the reported experimental

    products. The S-adduct represents the most plausible product formed by addition of OH to

    CS2. The adduct then undergoes a bimolecular reaction with atmospheric O2 yielding COS,

    OH and SO. The kinetic analysis developed in this study explains the atmospheric fate of

    reduced sulfur species, an important group of compounds in the global S cycle.

    Chapter 5 resolves the interaction of sulfanyl radical (SH) with aliphatic (C1 - C4)

    hydrocarbons, using CBS-QB3 based quantum chemistry calculations. Our findings

    demonstrate that, the documented inhibition effect of hydrogen sulfide (H2S) on pyrolysis of

    alkanes does not apply to alkenes and alkynes. During interaction with hydrocarbons, the

    inhibitive effect of H2S and promoting interaction of SH radical depend on the reversibility of

    the H abstraction processes. In the case of methane, we conclude that, the reactivity of SH

  • Chapter 1. Introduction

    6

    radicals towards abstracting H atoms exceeds that of HO2 but falls below those of OH and NH2

    radicals.

    Chapter 6 discusses the experimental results of CS2 oxidation using a tubular-flow reactor

    under dry conditions for an oxygen-fuel equivalence ratio of 0.7, 1.0 and 1.2 and for a

    temperature range of 700 to 1200 K. Compared with moist oxidation, the reaction forms CO

    as a final product, as we detected no CO2 in the exhaust gases. A sensitivity analysis confirmed

    a significant influence of the CS2/O2 and S/O2 sub-mechanisms on the oxidation process.

    Subsequent theoretical calculations identify the intersystem crossing between triplet and singlet

    pathways that lowers the activation energy of the key elementary reactions, as the reason for

    the extreme flammability of CS2. This chapter also presents the analysis of the nature of the

    intersystem crossing and its impact on the low ignition temperature of CS2, and updates the

    CS2/O2 and S/O2 reaction subsets for the dry oxidation mechanism of CS2. The results of the

    kinetic modelling and the experimental measurements demonstrate good agreement.

    Chapter 7 introduces an experimental study of dry and moist oxidation of CS2 in JSR and

    develops a comprehensive oxidation mechanism with updated COS/O2 subset. The dry and

    moist oxidation of CS2 exhibit the same conversion profile for each species in experiments

    performed in the temperature range of 650 – 1100 K under the atmospheric pressure. In

    comparison, kinetic modelling with the moist oxidation mechanism proposed in this work

    predicts the conversion of CO to CO2 at temperatures in excess of 1200 K. The fast conversion

    of COS in experiments, in which this species arises as an intermediate, forces us to study in

    detail the consumption channels for COS. We propose the intersystem crossing process to

    occur in COS/O2 subset, in analogy to the ISC appearing for other species in the CS2/S/O2

  • Chapter 1. Introduction

    7

    system, to reduce the reaction barrier. Good agreement emerges between measured and

    modelled onset temperature and CS2 consumption, confirming the robustness of the model.

    Chapter 8 explores the impact of added CS2 in the oxidation of CH4. Co-oxidation experiments

    of CH4/CS2/O2 in JSR illustrate the promotion effect of CS2 on the ignition of methane in

    natural gas. However, the presence of CH4 significantly elevates the oxidation temperature of

    CS2, indicating the inhibition effect of CH4 on CS2 oxidation.

    Chapter 9 presents the concluding remarks of the thesis and provides suggestions for future

    studies on chemistry of reduced sulfur species.

    Reference

    [1] C.F. Cullis, M.F.R. Mulcahy, The kinetics of combustion of gaseous sulphur compounds,

    Combust. Flame 18 (1972) 225-292.

    [2] A. Williams, J.M. Jones, L. Ma, M. Pourkashanian, Pollutants from the combustion of solid

    biomass fuels, Prog. Energy Combust. Sci. 38 (2012) 113-137.

    [3] G.P. Huffman, S. Mitra, F.E. Huggins, N. Shah, S. Vaidya, F. Lu, Quantitative analysis of

    all major forms of sulfur in coal by x-ray absorption fine structure spectroscopy, Energy &

    Fuels 5 (1991) 574-581.

    [4] Department of Environment and Energy, Fuel Standard Determination 2001, Government

    of Australia, 2001.

  • Chapter 1. Introduction

    8

    [5] P. Jaramillo, W.M. Griffin, H.S. Matthews, Comparative life-cycle air emissions of coal,

    domestic natural gas, LNG, and SNG for electricity generation, Environ. Sci. Technol. 41

    (2007) 6290-6296.

    [6] K. Karan, A.K. Mehrotra, L.A. Behie, A high-temperature experimental and modeling

    study of homogeneous gas-phase COS reactions applied to Claus plants, Chem. Eng. Sci. 54

    (1999) 2999-3006.

    [7] K. Karan, L.A. Behie, CS2 formation in the Claus reaction furnace:  a kinetic study of

    methane−sulfur and methane−hydrogen sulfide reactions, Ind. Eng. Chem. Res., 43 (2004)

    3304-3313.

    [8] P. Warneck, The biogeochemical cycling of sulfur and nitrogen in the remote atmosphere,

    EOS, Trans. Am. Geophys. Union, 68 (1987) 240-240.

    [9] V. Gardiner, Global Tropospheric Chemistry—A Plan for Action, National Academy Press,

    1985. 110-111.

    [10] M.P. Barkley, P.I. Palmer, C.D. Boone, P.F. Bernath, P. Suntharalingam, Global

    distributions of carbonyl sulfide in the upper troposphere and stratosphere, Geophys. Res. Lett.

    35 (2008) L14810.

    [11] F.R. Taylor, A.L. Myerson, First limit induction time studies of CS2−O2 explosions, Symp.

    (Int.) Combust. [Proc.] 7 (1958) 72-79.

    [12] American Industrial Hygiene Association, Carbon Disulfide (CS2) Fire.

    https://www.aiha.org/get-involved/VolunteerGroups/LabHSCommittee/Pages/Carbon-

    Disulfide-Fire.aspx (accessed 05/05 2017).

    [13] G.M. Bodner, Lecture demonstration accidents from which we can learn, J. Chem. Educ.

    62 (1985) 1105.

    [14] F.G. Cerru, A. Kronenburg, R.P. Lindstedt, Systematically reduced chemical mechanisms

    for sulfur oxidation and pyrolysis, Combust. Flame 146 (2006) 437-455.

  • Chapter 1. Introduction

    9

    [15] F. Viteri, M. Abián, Á. Millera, R. Bilbao, M.U. Alzueta, Ethylene–SO2 interaction under

    sooting conditions: PAH formation, Fuel 184 (2016) 966-972.

  • Chapter 2. Literature review

    10

    Chapter 2

    Literature review

  • Chapter 2. Literature review

    11

    2.1. Introduction

    Sulfur stands as an essential impurity in fossil fuels, biofuels and municipal wastes [1, 2].

    Taking up to several percentage by weight in fuels, sulfur gives rise to industrial problems on

    a large scale, especially as an atmospheric pollutant [3]. In nature, sulfur occurs in the

    pure element form of sulfide and sulfate in minerals. Once released into air through oxidation,

    the mainly gaseous product - sulfur oxide (SOx) is in the form of sulfur dioxide (SO2) [4],

    which is an important greenhouse gas as well as the source of acid rain. Anthropogenic

    emissions of sulfur into the troposphere peaked during the year 1972 at about 131 million

    tonnes [5]. By 2011, due to the world-wide efforts to control air pollution, the emission of

    sulfur species has been reduced to 106 million tonnes [6]. More than 85 % of sulfur compounds

    originate from fuel combustions in power plants or other devices [7]. As a result, sulfur

    chemistry draws a lot of attention among researchers in both energy and environmental

    engineering.

    Most of the current research on sulfur combustion focuses on the industrial thermal conversion,

    oxidation of H2S and emission of SO2 as air pollutant. H2S emerges as the principal sulfur

    carrier, in the gasification process of coal [8]. In oil industries, sulfur compounds such as thiols,

    thiophenes, sulphides and disulfides (0 – 3 wt. %), are converted to H2S, before they could

    poison the cracking catalysts for hydrocarbons [9]. H2S is also commonly present in natural

    gas [10]. To remove H2S, amine treatment is applied to crude oil [11] while cation containing

    materials, such as limestone (Ca2+) and Magnesium oxide (Mg2+), are sprayed into coals [12].

    The multi-step Claus process can convert sulfur into the solid state from gaseous H2S in natural

    gas and other exhausts containing H2S, derived from refining crude oil and other industrial

    processes [13]. The oxidation of H2S releases SO2, known as the most notorious sulfur

  • Chapter 2. Literature review

    12

    compound emitted from fuel combustion [14]. However, a comprehensive oxidation

    mechanism for H2S still needs further studies, with key elementary reactions plagued by

    significant uncertainties [15].

    Apart from H2S, other less-known reduced sulfur species – carbon disulfide (CS2) and carbonyl

    sulfide (COS) – are also produced during combustion or thermal pyrolysis of fuels. H2S, SO2,

    CS2 and COS are major sulfur-containing species produced during coal pyrolysis [16]. In the

    Claus process, CS2 and COS contribute up to 20 % of the total sulfur emissions in exhaust [17,

    18]. The formation of CS2 and COS is confirmed in the oxidation of methane seeded with H2S

    [19, 20], which is prevalent in natural gas as an impurity. The presence of SO2 in ethylene

    pyrolysis also leads to the formation of CS2 [21]. However, studies on CS2 and COS oxidation

    are scarce, leading to uncertainties in modelling combustion processes in the real-wold

    scenario.

    Pure CS2 exists as a colourless liquid, with a low boiling point at 46.3°C under atmospheric

    pressure. As an excellent non-polar solvent, CS2 has found numerous applications in both

    laboratories and industry. Additionally, the production of synthetic fibres, rubbers and

    pesticides requires CS2 as feedstock. In industry, CS2 is manufactured with methane and solid

    sulfur, in the presence of silica gel or alumina catalyst under 600 °C [22]:

    2 CH4 + S8 catalyst, 600°C→ 2 CS2 + 4 H2S

    However, CS2 is characterized by extreme flammability [23] and high explosion propensity

    [24], which has led to several fire accidents in laboratories [25, 26] and chemical warehouses

    [27].

  • Chapter 2. Literature review

    13

    Natural production of CS2 and COS occurs from volcanic eruptions and bushfires [28], while

    anthropogenic emission contributes to most of the total atmospheric content of CS2 and COS

    [29]. CS2 exists at levels of parts per trillion (ppt) in the earth’s troposphere; 15 to 30 ppt in

    the non-urban troposphere and 100 to 200 ppt in polluted urban areas [30]. On the other hand,

    COS stands as the most abundant sulfur compound in the atmosphere (> 400 ppt) [31].

    Despite the important role of reduced sulfur species in the environmental and energy fields,

    studies on combustion chemistry of CS2 and COS are still limited. This study aims to review

    the literature related to experiments and oxidation kinetics of reduced sulfur species (H2S, CS2

    and COS), and their interaction with hydrocarbons in combustion processes. Firstly, we

    examine the pre-established oxidation processes for H2S, which contains S/SO sub-mechanism

    SH radical as the key intermediate. The occurrence of intersystem crossing (ISC) in S/SO and

    H2S combustion processes is highlighted. Next, we survey the oxidation of CS2 and COS. In

    the atmosphere, CS2 will convert itself to COS with the aid of OH radical. COS represents the

    most stable and most abundant of sulfur carrier in the atmosphere. For CS2 oxidation, the high

    activation energy of triplet oxidation pathway fails to explain the low ignition temperature for

    the experiments conducted with the tubular-flow reactor. The ISC is highlighted to occur in

    the CS2 combustion mechanism. COS is produced as an intermediate in CS2 oxidation process,

    i.e., COS/O2 mechanism is involved in CS2 oxidation as a subset. We also discuss the

    production of S atom in both ground state (3S) and excited state (1S) in photolysis of reduced

    sulfur species. Finally, the interaction between sulfur species and hydrocarbons has been

    investigated for 1) Influence of H2S/SH on combustion process of hydrocarbons and 2) impact

    of SO2 on ethylene pyrolysis. Future research directions for sulfur chemistry will be suggested

    in the conclusion section, for both experimental and theoretical aspects.

  • Chapter 2. Literature review

    14

    2.2. Oxidation of H2S

    As the most significant sulfur carrier in natural gas, oxidation and pyrolysis of H2S has been a

    subject of many studies focusing on its formation in coal pyrolysis [32], conversion in Claus

    process [13] and application as potential hydrogen source [33, 34]. As a flammable gas, the

    oxidation of H2S has been investigated with batch reactor to reveal its explosion limit [35, 36],

    and shock wave tube to decide its ignition delay time [37]. With the aid of a mass spectrometer,

    SO, SO2 and SO3 are sampled from the flat flame of H2S [38].

    The recent study by Zhou et al. proposes a comprehensive oxidation mechanism for H2S with

    tubular flow reactor [14] and accurate theoretical calculations [39, 40]. Song et al. extended

    both the experimental and kinetic modelling of H2S oxidation in plug flow reactor at elevated

    pressure up to 100 bar [15]. Owing to the uncertainties in reactions related to SH radical,

    further work is necessary to improve the mechanism of H2S oxidation.

    In this section, H2S oxidation is reviewed under atmospheric and combustion processes. We

    also make recommendations for future work on H2S oxidation in each part.

    2.2.1. Atmospheric oxidation of H2S

    Reduced sulfur species such as H2S, CS2 and COS represent major atmospheric S-carriers and

    assume an important role in the global sulfur cycle. It is generally believed that the majority

  • Chapter 2. Literature review

    15

    of H2S is contributed by natural processes such as volcanic eruptions, bush fires, life-cycles in

    marshes and sea [28, 41]. However, human activities also result in a significant emission of

    H2S, especially in the energy industry [42]. In earth’s troposphere, H2S varies in the level of

    20 – 60 ppt [43].

    Direct interaction between H2S and O2 proceeds slowly under atmospheric thermal conditions.

    While no direct experimental measurement has been found, a quantum chemistry calculation

    was conducted by Montoya et al. [44] to demonstrate a high activation energy at 159.8 kJ/mol.

    The reaction rate for R2-1 is expressed as k1 = 4.6 × 10-19 × T2.76 × exp (-159.8/RT)

    [cm3/(molecule∙s)] in Arrhenius equation fitted from 300 to 2000 K. As extrapolated in this

    work, the rate constant for H2S + O2 corresponds to 2.7 × 10-40 cm3/(molecule∙s), which is too

    slow to account for the atmospheric conversion of H2S (even when the atmospheric oxygen

    concentration is considered).

    H2S + O2 → HO2 + SH R2-1

    With analytical techniques applied to atmospheric chemistry, researchers have investigated the

    oxidation agent of H2S in the atmosphere among atomic oxygen O [45], ozone O3 [46], atomic

    hydrogen H and hydroxyl radical OH. The reaction between H2S and O has been studied with

    the aid of electron spin resonance (ESR) applied detect and measure O concentration in a flow

    reactor [45]. UV-Vis absorption of O3 and resonance fluorescence of H facilitate researchers

    in measuring the reaction rate of H2S + O3 [46] and H2S + H [47]. The reaction between OH

    and H2S has been measured in discharge-flow reactor with laser induced fluorescence LIF, to

    quantitate the concentration of OH radical [48].

  • Chapter 2. Literature review

    16

    From Table 2.1, which contrasts the rate constant of H2S oxidized by O2, O, O3, H and OH in

    the atmosphere, it can be observed that the reaction between OH radical and H2S contributes

    to the most important consumption channel for atmospheric H2S. Moreover, the very low

    concentrations of O and H compared to OH in the atmosphere [49] clearly weakens the

    influence of R2-2 and R2-4.

    Table 2.1. Atmospheric oxidation of H2S, reaction rate k (cm3/(molecule∙s)) is under

    atmospheric condition at 298 K, 1 atm.

    Reaction Method k Source

    R2-1 H2S + O2 → HO2 + SH Theoretical

    calculations

    2.7 × 10-40 [44]

    R2-2 H2S + O → OH + SH Experimental

    measurement

    2.3 × 10-14 [45]

    R2-3 H2S + O3 → HO2 + SH Experimental

    measurement

    4.0 × 10-16 [46]

    R2-4 H2S + H → H2 + SH Experimental

    measurement

    9.6 × 10-13 [47]

    R2-5 H2S + OH → H2O + SH Experimental

    measurement

    3.6 × 10-12 [48]

    By locating OH to be the most potent oxidisers for H2S in the atmosphere, the reaction pathway

    of R2-5 has been studied intensively with quantum chemistry calculations by Mousavipour et

    al. [50]. Potential energy surface is explored at the MP2/6-311++G(d,p) level of theory [51].

    Rate constants were calculated using transition state theory (TST) [52] with one-dimensional

    correction for tunneling effect [53], resulting in an activation energy at 4.2 kJ/mol fitted from

    300 – 3000 K. The computed reaction rate amounts to 2.65 × 10-12 cm3/(molecule∙s) at 298 K,

    which is in good agreement with the experimental analogous value at 3.6 × 10-12

    cm3/(molecule∙s) [48], confirming the accuracy of the theoretical calculation.

  • Chapter 2. Literature review

    17

    Figure 2.1 demonstrates the reaction pathway and structure of the transition state structure

    calculated by Mousavipour et al. [50]. The highly reactive OH radical will extract one H from

    H2S to form H2O and SH radical. This reaction can be understood as H exchange between SH

    and OH radical. Since the bond dissociation enthalpy of H2O (498.8 kJ/mol) is much higher

    than that of H2S (380.1 kJ/mol) at 298 K [54], the OH radical is far more competitive than the

    SH radical, to acquire H.

    Figure 2.1. Calculated reaction pathway and transition state structure for H2S + OH → H2O +

    SH, plotted from calculation by Mousavipour et al. [50]. All enthalpy values are with reference

    to initial reactants H2S + OH, in kJ/mol at 0 K. All distance values (around chemical bonds)

    are in Å = 10-10 m.

    Further conversion of SH radicals, which involves O2, would proceed through R2-6: SH + O2

    → SO + OH. However, the reaction rate of R2-6 is measured to be less than 1.0 × 10-17

    cm3/(molecule∙s) [55] using resonance fluorescence technique, which is too slow to account

    for the consumption of SH in the atmosphere. O3 and NO2 have been found to be the major

    agents for SH conversion at a reasonable reaction rate, to produce HSO [56, 57] as products.

  • Chapter 2. Literature review

    18

    As shown in Table 2.2, NO2 acts as the most powerful scavenger for SH in the atmosphere. In

    spite of the important role of NO2 in atmospheric sulfur cycle, no quantum chemistry

    calculation has been conducted to reveal the reaction pathway of R2-8.

    Resende and Ornellas [58] applied theoretical calculations to explore the consumption of SH

    in the atmosphere for reaction R2-7: SH + O3 → HSO + O2. However, by conducting

    calculations in ground state with 2SH and 3O3, the computed activation energy attains a value

    of 26.5 kJ/mol. This energy barrier leads to a reaction rate for R2-7 at 2.3 × 10-17

    cm3/(molecule∙s) for atmospheric condition at 1 atm and 298 K, which is too slow compared

    to experimental measurement at 4.3 × 10-12 cm3/(molecule∙s). Since no spin forbidden channels

    were considered in this work, we suggest re-investigation of the reaction, proceeding through

    the excited state of reactants (1O3) or transition state. The intersystem crossing between triplet

    and singlet pathway may account for the fast conversion of SH as measured in R2-7.

    Table 2.2. Atmospheric conversion of SH, reaction rate k is under atmospheric condition at

    298 K, 101 kPa.

    Reaction Method k (cm3/(molecule∙s)) Source

    R2-6 SH + O2 → SO + OH Experimental

    measurement

    < 1.0 × 10-17 [55]

    R2-7 SH + O3 → HSO + O2 Experimental

    measurement

    4.3 × 10-12 [56]

    R2-8 SH + NO2 → HSO + NO Experimental

    measurement

    6.5 × 10-11 [57]

    The reaction rate of R2-9: HSO + O2 → HO2 + SO is measured to be less than 2.0 × 10-17

    cm3/(molecule∙s) [59] under atmospheric condition. While the interaction between HSO and

    O3 in the atmosphere (R2-10: HSO + O3 → HSO2 + O2) is also too slow to afford the

  • Chapter 2. Literature review

    19

    consumption of HSO in the atmosphere with a measured reaction rate at 5.0 × 10-14

    cm3/(molecule∙s) [60]. Reaction with NO2 (R2-11: HSO + NO2 → HSO2 + NO) has been

    found to be the main channel of HSO conversion, with a reaction rate measured to be 9.6 ×

    10−12 cm3/(molecule∙s) in the atmosphere [59]. To the best of our knowledge, literature

    presents no account of the reaction between HSO and OH. In this regard, the OH radical would

    extract the weakly bonded H atom from S-H bond in HSO to form H2O and SO.

    Table 2.3. Atmospheric conversion of HSO and HSO2, reaction rate k is under atmospheric

    condition at 298 K, 101 kPa.

    Reaction Method k (cm3/(molecule∙s)) Source

    R2-9 HSO + O2 → HO2 + SO Experimental

    measurement

    < 2.0 × 10-17 [59]

    R2-10 HSO + O3 → HSO2 + O2 Experimental

    measurement

    5.0 × 10-14 [60]

    R2-11 HSO + NO2 → HSO2 + NO Experimental

    measurement

    9.6 × 10−12 [59]

    R2-12 HSO2 + O2 → SO2 + HO2 Experimental

    measurement

    3.0 × 10−13 [59]

    Subsequent conversion of HSO2 is measured by Lovejoy et al. [59] to proceed with O2 with a

    reaction rate at 3.0 × 10−13 cm3/(molecule∙s). However, it must be noted that the methods used

    by Lovejoy et al. to obtain the rate coefficient was indirect, which may lead to substantial error

    due to the uncertainty of HO2 released from parallel reactions [61]. Table 2.3 lists the reaction

    rate for HSO and HSO2 consumption in the atmosphere.

    It is worthwhile mentioning that no theoretical studies on the atmospheric conversion of HSO

    and HSO2 have been found. Here, we would recommend further studies on atmospheric

    reaction involving HSO and HSO2 with computational chemistry, to reveal the detailed reaction

  • Chapter 2. Literature review

    20

    pathways. The proposed reaction HSO + OH may open another corridor for HSO consumption,

    which is operating independent of the NO2 species in the atmosphere.

    Figure 2.2 presents a summary for H2S oxidation in literature, as well as a recommended

    pathway, which is of potential importance. In a nutshell, NO2 is extremely important in the

    consumption of H2S in the atmosphere. However, in the absence of NO2, an alternative

    pathway is proposed here involving OH radical, which is relatively more prevalent in the

    atmosphere. We also highlight the failure of ground state calculation on SH + 3O3 to reproduce

    the experimental results. The intersystem crossing between triplet and singlet may assist in

    reducing the activation energy for the reaction SH + 3O3. Finally, more theoretical work on

    HSO and HSO2 is suggested to broaden our understanding on atmospheric oxidation of H2S.

    Figure 2.2. Atmospheric oxidation pathway for H2S and suggested study on alternative

    pathway.

    2.2.2 Combustion mechanism of H2S

    Early research on H2S oxidation focused on its ignition delay time [37], explosion limit [35,

    36] as well as the flame structures [20, 38], and the reaction mechanism was based on measured

  • Chapter 2. Literature review

    21

    product distributions. The first comprehensive review for the combustion of H2S was

    conducted by Cullis and Mulcahy in 1972 [3], based on low-temperature photolysis

    experiments and flame studies. The final product of H2S oxidation is identified as SO2 with

    SO and S2O performing as important intermediates. The proposed consumption channels

    involve intermediates such as SH, SO, SO3, S, S2, OH and HO2. However, due to the lack of

    reliable kinetic parameters for reactions and thermodynamic data for these species, the

    validation of the mechanism remains a very challenging task. Further study by Gargurevich

    [62] summarized some early work on kinetics of sulfur chemistry and also estimated the rate

    parameters for the missing steps, to resolve the role of H2S in Claus process. Further studies

    by Hughes et al. [63] discussed the uncertainties caused by key steps such as SO + OH and SO2

    + H. By realizing the importance of SH, S and SO radical in the mechanism, Tsuchiya et al.

    [64] measured the reaction rate for three key elementary reactions: SH + O2, S + O2 and SO +

    O2 with shock wave tube at high temperature range (1000 – 1600 K). A systematic reduced

    combustion mechanism for H2S has been proposed by Cerru et al [65, 66], extending the radical

    pool to contain HSS, HSSH, HSO and HSO2. However, the author attributes the uncertainties

    related to SH, HSS and HSSH reactions to insufficient studies, and highlights the paramount

    importance of related reactions on the ignition of H2S oxidation.

    Some recent studies by Zhou et al. [67, 68] and Gao et al. [39] have combined experimental

    work and high level theoretical calculations to resolve the role of H2/S2 system (including SH,

    HSS and HSSH) in H2S oxidation, in which they suggest a crossing process from triplet ground

    state to singlet excited state for transition structure. The updated mechanism results in a

    satisfactory agreement with further experimental validation, in a vertical tubular flow reactor

    at 950 – 1150 K under atmospheric pressure [14]. Song et al. further expanded the experiments

    with flow reactor under high pressure range between 30 and 99 atm (atmospheric pressure),

  • Chapter 2. Literature review

    22

    and interpreted the results with the mechanism of Zhou et al. [14] with certain modification.

    It is pointed out that the branch ratio of the reaction of SH + SH to give H2S + S (chain

    propagation) or HSSH (chain termination) is the key step to control the oxidation of H2S.

    We also intend to point out that the mechanism modelling with plug-flow reactor (PFR) in the

    work of Zhou et al. work [14] may not be consistent with their experiments. For tubular flow

    reactor, the flow condition can be approximately divided into the laminar [69] and the turbulent

    [70] flow regime. As controlled by residence time, the flow condition of inlet gases wouldn’t

    reach a turbulent flow with a low flow velocity and Reynold’s number. The species

    composition and temperature distribution in laminar flow condition may result in disagreement

    with plug flow modelling, in which species and temperature are homogeneous everywhere.

    The modelling of PFR is only an approximation to tubular-flow reactor under atmospheric

    pressure. This also explains the fact that Song et al. applied a high pressure tubular-flow reactor

    to reach turbulent flow condition as modelled with PFR.

    In this section, we review the key oxidation process of H2S based on the mechanism of Zhou

    et al. [14]. The occurrence of intersystem crossing for selected elementary reactions, especially

    in H/S system and S/O system, is the focal point of this part.

    The chain initiation of H2S oxidation has been studied by Montoya et al. [44] through

    theoretical calculation with G2 method [71]. However, direct H abstraction by O2 in ground

    state is featured with a sizable activation enthalpy up to 168.8 kJ/mol at 0 K, as shown in Fig.

    2.3. The reaction is also strongly endothermic by 169.5 kJ/mol, which is not favored in

    combustion processes. The backward reaction is more prevalent due to a very shallow reverse

    reaction barrier. This indicates that SH tends to abstract H from HO2.

  • Chapter 2. Literature review

    23

    Figure 2.3. Potential enthalpy surface for H2S + O2 calculated with G2 method at 0 K by

    Montoya et al. [44]. All enthalpy values are with reference to initial reactants H2S + 3O2, in

    kJ/mol.

    However, a crossing between singlet and triplet state is considered by the author to give us a

    singlet HOOSH adduct, which is even stabler than the initial triplet reactants H2S + 3O2 by

    15.8 kJ/mol. The intersystem crossing (ISC) from ground triplet state to singlet adduct opens

    another corridor to form HSO + OH, with a lower activation barrier through the crossing point

    (ISC point). However, the author has proposed the ISC process without locating the crossing

    point. Here, we would recommend calculation of the detailed potential energy surface (PES)

    for both triplet and singlet pathway separately. Then by overlapping the two PESs, we could

    approximately fix the crossing point.

    H2S + O2 → HSO + OH R2-13

    The consumption of HSO has not been investigated. The sole estimation from Zhou et al. [14]

    suggests an activation energy at 27.6 kJ/mol to produce SO2 and OH. Here, we would

    recommend a theoretical study on the formation of a four-membered ring of HOOSO, where

    the O-O bond breaks to give OH and SO2. The produced OH radical extracts one H from H2S

  • Chapter 2. Literature review

    24

    readily, through a trivial activation energy at 4.2 kJ/mol, as calculated by Mousavipour et al.

    [50]. The quantum calculation of R2-2 has been discussed in the atmospheric part to give SH

    as an important chain carrier.

    HSO + O2 → SO2 + OH R2-14

    H2S + OH → SH + H2O R2-2

    Direct measurement of SH + O2 has been conducted by Tsuchiya et al. [64] with shock wave

    tube. The activation energy is fitted at 75 kJ/mol in temperature range from 1400 – 1700 K.

    However, the theoretical calculation by Zhou et al. [72] at ground state (triplet pathway)

    predicted the formation of SO + OH and HSO + O via activation barriers at 81 kJ/mol and 89

    kJ/mol at 0 K, respectively. The production of HSO + O is unlikely to proceed with an

    endothermicity at 89 kJ/mol, which means the reverse reaction is featured with no activation

    barrier. However, the theoretical values derived from this work underestimate the reaction rate

    by a factor of around 1.5, compared to measured results from 1400 – 1700 K. A faster reaction

    channel is needed to reproduce the experimental results. By considering an electronically

    excited state HSO2 suggested by Freitas et al. [73], we recommend underpinning the singlet

    reaction pathway and a potential crossing point to ground triplet reactions. This may explain

    the relatively fast conversion between SH and O2 as measured in experiments.

    SH + O2 → OH + SO R2-15

    SH + O2 → HSO + O R2-16

    Due to the low reactivity between SH and O2, the recombination of SH + SH is considered to

    act as a major chain propagation step. Theoretical calculations were made by Zhou et al. [40,

  • Chapter 2. Literature review

    25

    68] with consideration of ISC. As depicted in Fig. 2.4, the triplet reaction pathway is featured

    with an activation barrier at 23.7 kJ/mol at 0 K. SH abstracts H from another SH to produce

    H2S and S atom. An alternative path is provided through singlet pathway. Direct addition of

    SH gives singlet 1HSSH, which further converts to singlet 1H2SS by H migration between S

    atoms. Subsequent fission of S-S bond and intersystem crossing from singlet to triplet surface

    result in 1H2S and 3S in ground state as products. Further experimental study by Gao et al. [39]

    confirms the ISC channel affords the predominant reaction pathway under low temperature

    range < 800 K, while triplet pathway would dominate the reaction above 1000 K.

    Compared to R2-15, the recombination of SH proceeds through singlet pathway plays an

    important role in chain propagating, to produce S atom in the system.

    SH + SH → H2S + S R2-17

    Figure 2.4. Potential enthalpy surface for SH + SH at 0 K calculated by Zhou et al. [68]. All

    enthalpy values are with reference to initial reactants 2SH + 2SH, in kJ/mol.

    Further interaction between S and O2 produces SO and O, which exhibits a substantial influence

    on the H2S oxidation, as the chain branching step. Lu et al. [74] reported results from combined

  • Chapter 2. Literature review

    26

    experimental measurements and theoretical calculations. The experiments reveal that the S +

    3O2 reaction demonstrates different temperature dependence for low (T < 1000 K, Ea < 3

    kJ/mol) and high (T > 1000 K, Ea > 30 kJ/mol) temperature range, as illustrated in Fig. 2.5.

    The behaviour of the S + 3O2 reaction is similar to that of the SH + SH reaction discussed

    above. Computational work of the same researchers, performed at the G2M level of theory,

    has also explored the possible reaction pathways. The reaction proceeding on a singlet surface

    represents the only channel that can account for the low-temperature oxidation. At

    temperatures higher than 1200 K, the triplet pathway dominates the overall S + 3O2 reaction.

    S + O2 → SO + O R2-18

    Tsuchiya et al. [64] observed similar behaviour of SO + 3O2, concluding that the activation

    energy of this reaction resides 34.0 kJ/mol above the separated reactants within a temperature

    range of 1130 – 1640 K. However, for lower temperatures of 250 – 585 K, the activation

    energy corresponds to 19.0 kJ/mol, as reported by Garland [75]. This significant discrepancy

    likely originates from ISC phenomenon operating in the low temperature window.

    SO + O2 → SO2 + O R2-19

  • Chapter 2. Literature review

    27

    Figure 2.5. Arrhenius plot for reaction S + 3O2 based on Lu et al.’s [74] experiments. The

    dashed lines represent the triplet (black) and singlet (red) pathways suggested in this work.

    In Fig. 2.6, the key consumption steps for H2S oxidation reviewed in this study are

    demonstrated. We highlight the ISC process in R2-13, R2-17 and R2-18, and highlight the

    possible occurrence of ISC in R2-19. Further investigation on reaction pathway for R2-14 is

    also recommended on both triplet and singlet surface.

    Figure 2.6. Key oxidation step for H2S in combustion process.

    2.3. Oxidation of CS2

    In this section, the oxidation pathway of CS2 is reviewed under both atmospheric and

    combustion processes. COS oxidation is also included as a sub-mechanism in CS2 combustion

  • Chapter 2. Literature review

    28

    mechanism. We also reviewed the photolysis study of reduced sulfur species: H2S, CS2 and

    COS, which produce S atom in its excited state (singlet 1S) instead of ground state (triplet 3S).

    2.3.1. Atmospheric oxidation of CS2

    Apart from H2S, reduced sulfur species of CS2 and COS represent major sulfur carriers in the

    atmosphere. Both natural [28] and human activities result in CS2 emission into atmosphere.

    However, the majority of atmospheric CS2 originates from industrial processes, such as

    production of insecticides and man-made fibres. Moreover, CS2 and COS correspond to an

    important by-product from the Claus gas de-sulfurising process [18], taking up nearly 20 % of

    the total sulfur emission from the Claus process. CS2 exists at levels of parts per trillion in the

    Earth’s troposphere; 15 to 30 ppt in the non-urban troposphere and 100 to 200 ppt in polluted

    urban areas [30]. COS stands as the most abundant sulfur compound in the atmosphere (> 400

    ppt) with a relatively long life-time of up to years [31].

    Similar to H2S, direct interaction between CS2 and O2 is featured by a significant activation

    energy, which renders the reaction less important in the atmosphere. With the very low

    concentrations of O(3P) in the atmosphere [49] and the absence of any decomposition of CS2

    near the UV-region [76], OH radicals constitute the sole initial atmospheric oxidiser. However,

    experimental results on the reaction of CS2 and OH reaction indicated an extremely slow rate

    constant:

    CS2 + OH → COS + SH R2-20

  • Chapter 2. Literature review

    29

    By producing OH with pulsed H2O photolysis technique, Kurylo [77] as well as Atkinson et

    al. [78] reported the rate coefficient of R2-20 as 1.9 × 10-13 cm3/(molecule∙s), while Atkinson

    and Pitts measured an upper limit for R2-20 to be 7.0 × 10-14 cm3/(molecule∙s) at 298.15 K.

    However, photolysis of H2O at 165-185 nm may also dissociate the CS2 molecule, resulting in

    an error in the reported rate constant of R2-20 due to loss of CS2. A CS2-filling cell was applied

    by Wine et al. to filter radiation that dissociates CS2 and report an even lower rate constant for

    R2-20, below 2.0 × 10-15 cm3/(molecule∙s) at 298.15 K [79]. Similar result has been achieved

    by Iyer and Rowland who produced OH radicals by a continuous wave (CW) photolysis of

    H2O2 at 254 nm [80], avoiding the dissociation energy of C-S bond.

    By introducing O2 into the system of CS2 and OH, the consumption of CS2 proceeds with a

    much higher reaction rate. Jones et al. [81, 82] measured the rate constant of R2-20 in the

    presence of O2 (i.e. CS2 + OH + O2) using photolysis of HONO, reporting a rather rapid rate

    constant of 2.0 × 10-12 cm3/ (molecule∙s) at 298.15 K. The promotion effect of O2 was also

    confirmed by Barnes et al. using CW photolysis [83], who obtained a very similar rate constant.

    Two important steps control the atmospheric oxidation of CS2, namely formation of a CS2OH

    and reactions of O2 with this adduct:

    CS2 + OH → CS2OH R2-21

    CS2OH + O2 → products R2-22

    Experiments by Murrells et al. [84] and Lovejoy et al. [85] tested a wide temperature range of

    249 – 318 K to confirm the formation of OCS, and SO2 as the sole experimental products.

  • Chapter 2. Literature review

    30

    Theoretically, calculations of Lunell et al. [86] at the HF/3-21G(d) level of theory predicted

    two distinct isomers of CS2OH, viz., the S-adduct – SCS(OH) and the C-adduct – SC(OH)S.

    However, an incorrect structure of the SCS(OH) adduct is considered in Lunell et al.’s work

    [86], resulting in a strongly endothermic reaction by 104.2 kJ/mol, with respect to the reaction

    of CS2 and OH in R2-21. Subsequent computations of McKee [87] at the modified G1 [88]

    level of theory, at 298.15 K, yielded reaction enthalpies of -24.7 kJ/mol and -126.8 kJ/mol,

    respectively, for the generation of the S- and C-adducts of CS2OH in R2-21. McKee and Wine

    [89] further investigated R2-22 with optimised geometries obtained at the B3LYP/6-311+G(d)

    level of theory and constructed a pathway for the formation of COS and HOSO. Figure 2.7

    depicts the detailed reaction pathway for the S-adduct to interact with O2. The insertion of O2

    on C atom of S-adduct results in an intermediate SCS (OH) O2 with a sizable activation barrier

    at 28.6 kJ/mol. Subsequent fissions of C-S and O-O bond yield COS and HOSO as products.

    However, the calculated well depth of S-adduct amounts to 26.9 kJ/mol, which underestimates

    the stability of SCS(OH) as measured at 45.6 ± 4.2 kJ/mol by Murrells et al. [84].

    Figure 2.7. Potential enthalpy surface for CS2 + OH + O2 calculated at 0 K by McKee and

    Wine [89]. All enthalpy values are with reference to initial reactants CS2 + OH + O2, in kJ/mol.

    The HOSO + O2 reaction has a near zero activation barrier, as calculated by McKee and Wine

    [89] The generation of HO2 + SO2 should proceed readily in the atmosphere. It explains the

  • Chapter 2. Literature review

    31

    observation of SO2 in experimental measurements. R2-22 is updated below to show the

    formation of HOSO, while R2-23 demonstrates its consumption channel in the atmosphere.

    CS2OH + O2 → COS + HOSO R2-22

    HOSO + O2 → HO2 + SO2 R2-23

    While SO2 will be rapidly converted to acid rain by OH and O2 as oxidiser [90-93] in the

    atmosphere, COS tends to accumulate as the most abundant sulfur carrier in the earth’s

    troposphere (> 400 ppt), with a relatively long life-time of up to years [31].

    Figure 2.8. Atmospheric oxidation pathway for CS2.

    Figure 2.8 depicts the atmospheric conversion of CS2 with COS and SO2 as products. Although

    a comprehensive quantum calculation has been carried out by McKee and Wine [89], the

    relative low level of theory may lead to significant error, as compared with experiments.

    Further calculation with an adequately high level of theory is desirable, to further improve the

    atmospheric oxidation mechanism of CS2.

  • Chapter 2. Literature review

    32

    2.3.2. Combustion mechanism of CS2

    Earlier work on the combustion of CS2 was reviewed by Cullis and Mulcahy [3] . Myerson

    and Taylor extensively studied the ignition of CS2-O2 mixtures as an extremely flammable

    chemical, as a representative case for branched-chain reactions [24, 94]. It was demonstrated

    that a mixture containing as little as 0.03 % (by mol) CS2 in oxygen ignites at 80 oC and 0.05

    atm, resulting in sustained cool flame propagation. Once ignited, a cool flame propagates

    through the mixture at 55 oC, whereas the temperature rise in the flame is less than 15 oC. The

    most important oxidation products were reported to be CO, SO2, CO2, CS and SO, as well as

    COS. Intermediates such as SO, S and O appear much more reactive than COS and CS [95],

    with relatively low accumulation.

    Studies investigated the effect of various factors on the ignition behaviour of CS2, including

    irradiation by ultraviolet light [96], the presence of inert gases [97] and the conditions of

    surfaces [24]. Results from these investigations pointed to a profound distinction between cool

    and hot flames of CS2 with reference to hydrocarbons [98, 99]. As evidence of the cool flame

    combustion of CS2, it was found that the flammability of CS2 increases with the concentration

    of oxygen in the combustible CS2 mixture [100]. This behaviour significantly differs from that

    observed in combustion of hydrocarbons, where the second ignition limit arises due to a three-

    body reaction H + O2 + M → HO2 + M [101].

    In CS2 oxidation, the two rapid steps R2-18 and R2-19 produce the reactive O radicals:

    S + O2 → SO + O R2-18

    SO + O2 → SO2 + O R2-19

  • Chapter 2. Literature review

    33

    Various chain inhibitors modify the duration of an induction period from a few seconds to

    several minutes [24].

    Hanst and Myerson et al. [102] were the first to study the kinetics of explosive combustion of

    CS2. They used kinetic absorption spectroscopy to follow the variations in the concentrations

    of radical and molecular species during the explosion of CS2. The CS2-O2 mixture, initially

    heated to between 190 oC and 300 oC, exploded spontaneously. Strong and continuous

    absorptions confirmed the formation of SO and CS with the appearance of SO2 in the early

    stages of the explosions [96]. Spectroscopic techniques were also utilised to study explosions

    of CS2 initiated by shock waves [103] and flash photolysis [104]. In both cases, CS2 was found

    to be oxidized mainly into SO, SO2 and CO. However, experimental measurements on slow

    combustion and explosive combustion of CS2 have not contributed much insight into

    mechanisms governing its oxidation, as only final products at particular low temperature ranges

    were determined.

    Reaction pathways contributing to the explosion behaviour of CS2 are summarised in Fig. 2.9.

    Central to these mechanisms is the slow build-up of CS via reaction of O atom with CS2 and

    its further oxidation into SO/SO2. The fate of produced S atom was assumed to be controlled

    by reactions with O atom and CS2, which performs as the initiation step for CS2 explosions. It

    is worthwhile to note that the formation of an O atom in this mechanism was attributed to

    “unknown” initiation reactions.

  • Chapter 2. Literature review

    34

    Figure 2.9. Oxidation pathways of CS2 derived from explosion behaviour of CS2 at early stage

    [96]. The initiation step is believed to be CS2 + O, while final products are CO and SO2.

    Along the same line of enquiry, a low pressure flame of CS2 was studied by Azatyan et al.

    [105] to address the high-temperature combustion mechanism, with the electron spin resonance

    (ESR) technique to detect O, CS and SO as intermediates. By analysing the relationship

    between O and SO in experiments, the authors proposed the formation of COS as a major

    conversion channel between O and SO with the reaction: COS + O → CO + SO.

    To confirm the occurrence of these reactions, Azatyan et al. [105] added small concentrations

    of COS to the CS2 flame at a temperature interval of 350 oC to 600 oC. A lower accumulation

    of O and the maximum yield of SO were observed, validating the influence of COS in the

    oxidation of CS2.

    Homan et al. [106] deployed the isothermal flow reactor to study combustion of CS2 in an

    excess of O2 at 927 oC and 0.4 atm. They confirmed the presence of COS in the oxidation

    process of CS2. In order to describe their experimental product profiles, a reduced kinetic

    model was constructed including COS as intermediate. The reaction CS2 + O → COS + S

    affords the production of COS.

  • Chapter 2. Literature review

    35

    CS2 + O → CS + SO R2-24

    CS + O → CO + S R2-25

    CS + O2 → CO + SO R2-26

    COS + O → CO + SO R2-27

    CS2 + O → COS + S R2-28

    Glarborg and Marshall [107] have recently studied the oxidation mechanism of COS with

    kinetic modelling. A detailed oxidation mechanism has been proposed, based on evaluation of

    data from literature. For the consumption of COS, the author considered three pathways:

    COS → CO + S R2-29

    COS + O → CO + SO R2-30

    COS + O2 → CO + SO2 R2-31

    For the thermal decomposition process (R2-29: COS → CO + S), a high activation energy at

    260 kJ/mol fitted from experimental work at high temperature (1900 – 3230 K) [108] makes

    it too slow to account for the conversion of COS. The COS interaction with atomic oxygen

    (R2-30: COS + O → CO + SO) presents as the major consumption channel for COS. The

    authors fitted the rate parameters from the experimental measurement by Homman et al. [16]

    around 1270 K, indicating an activation energy at 21.8 kJ/mol. The dominant products are CO

    + SO, while CO2 + S could form at higher temperature. With the absence of atomic O for chain

    initiation, the rate parameter of reaction between COS and O2, R2-31 (COS + O2 → CO + SO2),

    is estimated with an activation energy at 134.4 kJ/mol.

  • Chapter 2. Literature review

    36

    Furthermore, to evaluate their model, the authors have compared it in the oxidation of COS

    with analogous experimental data obtained in batch reactors [109], flow reactors [106], and

    shock tubes [110]. Although the proposed COS oxidation mechanism captured well the

    experimental profiles for major species measured with tubular flow reactor by Homann et al.

    [111], a slight correction for the onset of reaction has to be made to match the experiment

    initiation. The reaction COS + O has been recognized to be very sensitive to the overall

    mechanism and intersystem crossing from triplet to singlet surface would occur with high

    probability [112].

    The first attempt to construct a detailed kinetic model for CS2 oxidation was undertaken by

    Howgate and Barr [113] to model CS2-O2 flame. They include the direct interaction between

    CS2 and O2 to account for the chain initiation with an activation energy at 179.5 kJ/mol, as

    indicated: CS2 + O2 → CS + SO2. Further investigation by Hardy and Gardiner [114] adopted

    this chain initiation step in their mechanism and achieved satisfactory agreement with their

    experimental measurement for the ignition delay time of CS2 with shock wave tube.

    Subsequent measurement of Saito et al. [115] and Murakami et al. [116] revised the activation

    energy of R2-32 to be 134.5 kJ/mol and 130.0 kJ/mol, respectively, to match their experimental

    results from shock wave tube.

    CS2 + O2 → CS + SO2 R2-32

    Very recently, Glarborg et al. [117] formulated a kinetic model for oxidation of CS2. Kinetic

    data for key CS2/CS + O2 reactions have been investigated on the basis of ab initio calculations.

    For reactions CS2 + O2, the authors investigated the reaction pathway on both triplet and singlet

    surface, and concluded the products to be COS + SO, instead of formation of CS + SO2.

  • Chapter 2. Literature review

    37

    CS2 + O2 → COS + SO R2-33

    Figure 2.10. Potential energy surface for CS2 + O2 calculated by Glarborg et al. [117]. All

    energy values are with reference to initial reactants CS2 + O2, in kJ/mol.

    As demonstrated in Fig. 2.10, the triplet pathway of CS2 + O2 climbs through the CS2O2 adduct,

    which dissociates into COS and SO, through an overall reaction barrier of 223 kJ/mol.

    However, the singlet oxidation offers an optional pathway with a much lower activation barrier.

    Compared to the triplet CS2O2 structure that features a strong O-O bond, the singlet SOOCS

    adduct forms through the dissociative addition of oxygen atoms at C and S atom, with a much

    lower energy level. By locating the cross-over between the energy surface of triplet and singlet

    pathways, Glarborg et al. estimated the crossing point to reside 145 kJ/mol above the triplet

    reactants CS2 + 3O2, which coincides with the experimentally interpreted value in the shock

    wave tube system (130.0 – 179.5 kJ/mol) [114-116].

    With the formation of SO through R2-33, its interaction with O2 gives O as the most important

    chain carrier. The subsequent interactions between CS2 and O proceed through three reaction

    channels, resulting in CS + SO (R2-24), COS + S (R2-28) and CO + S2 (R2-34) as products.

  • Chapter 2. Literature review

    38

    CS2 + O → CO + S2 R2-34

    An experimental study using a fast flow reactor concluded that the formation of CS + SO

    dominates the overall process, with only a small generation of COS + S and CO + S2 [118].

    However, theoretical calculations on the triplet surface support only the formation of CS + SO,

    with a small activation energy at 4.2 kJ/mol. To produce COS + S or CO + S2, the reactions

    must overcome higher barriers of 41.3 kJ/mol and 33.2 kJ/mol, respectively. Thus, both

    reactions should be negligible, when contrasted with the formation of CS + SO [119]. Here,

    we would like to suggest the possible singlet reaction pathways to account for the formation of

    COS + S and CO + S2 [120], as detected in experiments.

    As calculated by Glarborg et al. [117], CS is consumed by O2 with an activation energy at

    117.5 kJ/mol on triplet surface. From shock-tube experiments, Murakami et al. reported the

    activation energy of R2-26 as 51.0 ± 61.7 kJ/mol [116]. The significant uncertainty from

    experiments and high activation barrier derived by triplet reaction pathway is to be noted; we

    would like to suspect the occurrence of ISC for R2-26.

    CS + O2 → CO + SO R2-26

    In general, the mechanism proposed by Glarborg et al. [117] displays good agreement with

    experimental results pertinent to ignition delays and explosion limits of CS2. However, their

    mechanism tends to over-predict the concentration of species under low temperature

    conditions, especially with respect to measurements from flow reactors [111] and shockwave

    tubes [115]. The uncertainty of the chain initiation reaction, CS2 + O2, and competition

    between the chain branching reactions, CS + O2, accounted for the discrepancy.

  • Chapter 2. Literature review

    39

    More recently, Abián et al. [121] studied the moist oxidation of CS2 in a flow reactor at 1 atm

    between 127 – 1127 oC under different oxygen-fuel equivalence ratios (λ) of 0.2, 0.7, 1.0, 2.0

    and 20.0. They compared their experimental profiles of consumed CS2 with the corresponding

    profiles calculated from the mechanism of Glarborg et al. [117] The modelling results from

    the mechanism of Glarborg et al. [117] over-predict the onset of the oxidation temperature for

    stoichiometry condition (λ = 1.0) by around 260 K. By adjusting the activation energies of the

    initiation reaction (CS2 + O2) and the chain propagation step (CS + O2), to lower values, Abián

    et al. reproduced their experimentally-measured concentration of species, as illustrated in Fig.

    2.11. However, it is worthwhile noting that the mechanism updated by Abián et al. [121] also

    underestimates the ignition of CS2 for around 100 K. The introduction of moisture, lack of

    control of residence time at different temperatures and plausible involvement of surface-

    mediated reactions on the reactor walls in Abian et al.’s experiments [121], make it difficult to

    interpret their experimental measurements and to differentiate plausible initiation by catalytic

    surfaces and water molecules. Further experiments on CS2 oxidation for kinetic modelling are

    required to develop a more accurate mechanism of CS2 oxidation.

    Figure 2.11. Comparison between experimental results for CS2 oxidation and the modelling

    results derived from the mechanisms of Glarborg et al.[117] and Abián et al. [121].

  • Chapter 2. Literature review

    40

    Figure 2.12. Oxidation pathways of CS2 with formation of COS. The initiation step is

    considered to be CS2 + O2, while final products are CO and SO2.

    To summarise, Fig. 2.12 demonstrates the key oxidation pathway for CS2, with formation of

    COS as an important intermediate. From literature, we highlight the occurrence of ISC in R2-

    33 as chain initiation step. More theoretical calculations are recommended on the key

    elementary reactions such as R2-26, R2-30 and R2-31, with consideration of singlet reaction

    surface. Experimental data for kinetic modelling is also desired to validate and update the

    oxidation mechanism for CS2.

    2.3.3. Production of S atom in excited state in photolysis of reduced sulfur species

    With consideration of the ISC process predicted from theoretical calculation for oxidation of

    reduced sulfur species, we review the direct experimental measurements of singlet/triplet

    species produced in pertinent studies. In the work of Nan et al. [122], the Doppler-broadening

  • Chapter 2. Literature review

    41

    fine spectrometry of S is measured by laser induced fluorescence (LIF) at 147 nm to identify

    the ratio of 3S/1S. As photolysis of COS at 222 nm light proceeded in a glass cell, the integrated

    intensity of 3S profile increased with time, due to the quenching of 1S produced in reaction.

    The production of singlet 1S appears predominant in photolysis of COS, with the produced

    3S/1S = 0.05/0.95. While further study on the photolysis of H2S gives combined 3S/1S =

    0.87/0.13 as products [123]. The mixed sulfur atoms with distinct electron spin state yielded

    from photolysis of reduced sulfur species reinforce the occurrence of ISC in sulfur-related

    reactions.

    Recently, by coupling mass spectrometry (MS) and photon-ionization techniques with

    synchrotron vacuum ultra-violet light as photon source, McGivern et al. [124] and Qi et al.

    [125] manage to distinguish the 3S and 1S atom with their ionization energy at 10.36 eV (3S)

    and 7.61 (1S), respectively. Compared to the traditional electron ionization applied with MS,

    photon ionization is featured with a high resolution at 0.1 eV, which is adequate to recognize

    the difference in the energy level of 3S and 1S. A study on CS2 photolysis with 193 nm light

    reveals that the combined formation of 3S/1S = 0.75/0.25 was conducted by McGivern et al.

    [124]. Figure 2.13 illustrates the energy level of CS2 photolysis reactions.

  • Chapter 2. Literature review

    42

    Figure 2.13. The energy level of CS2 photolysis reactions (in kJ/mol) with branch ratio

    calculated in this work by Gaussian 09 at CBS-QB3 composite [126].

    As demonstrated in Fig. 2.13, we observe that the ground state of CS2 is in singlet (1CS2), while

    that of S is in triplet (3S). If the photolysis of CS2 proceeds from ground state to ground state,

    an ISC point should exist to build a bridge to triplet S. As measured in experiments, 25 % of

    CS2 proceeds through the ISC to form ground state 3S [126]. The difference of electron

    quantum spin number of ground state 1CS2 and 3S has prompted consideration of the energy

    level of relevant reduced sulfur species. It is revealed that the ground state of species constituted

    by only S and O atoms are in triplet (3SO, 3S, 3O and 3O2), except for 1SO2, while the gourd

    state for species such as 1CS2, 1COS, 1CS and 1SO2 are in singlet. If we consider all oxidation

    processes to have proceeded on the ground state surface, we should have the reaction process

    with two ISC points as shown below:

    1CS2 → 3SO → 1SO2

  • Chapter 2. Literature review

    43

    Table 2.4 summarises the branching ratio of 3S/1S from photolysis of reduced sulfur species,

    as measured in experiments. Although the parent compounds are in their singlet (ground) state,

    the corresponding sulfurs atom arise in both singlet and triplet states. This has indicated that

    the ISC is predominant in sulfur related species, especially when it couples with O2.

    Table 2.4. The branching ratio of 3S/1S from photolysis of reduced sulfur species (H2S, COS

    and CS2).

    Species 3S/1S Wavelength (nm) Source

    H2S 0.87/0.13 193 [122]

    COS 0.95/0.05 222 [123]

    CS2 0.75/0.25 193 [124]

    However, the identification of electronic spin states of intermediates or prod