Experimental and Numerical Study of the Effects of Steam Addition on

Aug 4, 2017 - The effect of H2O addition on the oxidation of methane and ammonia during oxy-fuel combustion was investigated both experimentally and n...
0 downloads 13 Views 815KB Size
Subscriber access provided by UNIVERSITY OF ADELAIDE LIBRARIES

Article

Experimental and numerical study of the effects of steam addition on NO formation during methane and ammonia oxy-fuel combustion Yizhuo He, Xiaochuan Zheng, Jianghui Luo, Hangfei Zheng, Chun Zou, Guangqian Luo, and Chuguang Zheng Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/acs.energyfuels.7b01550 • Publication Date (Web): 04 Aug 2017 Downloaded from http://pubs.acs.org on August 5, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Energy & Fuels is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 31

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

1

Experimental and numerical study of the effects

2

of steam addition on NO formation during

3

methane and ammonia oxy-fuel combustion

4

Yizhuo He, Xiaochuan Zheng, Jianghui Luo, Hangfei Zheng, Chun Zou*, Guangqian

5

Luo*, Chuguang Zheng

6

State Key Laboratory of Coal Combustion, Huazhong University of Science and

7

Technology, Wuhan, 430074, P. R. China

8

*Corresponding author. Tel: +86 2787542417-8314; fax: +86 2787545526.

9

E-mail address: [email protected] (C. Zou), [email protected](G.Luo)

10 11

Abstract: The effect of H2O addition on the oxidation of methane and ammonia during

12

oxy-fuel combustion was investigated both experimentally and numerically. Comparison

13

experiments between O2/CO2 and O2/CO2/H2O atmospheres were conducted in a flow

14

reactor at atmospheric pressure with equivalence ratios ranging from fuel-rich to

15

fuel-lean and temperature from 973 K to 1773 K. The experimental results indicate that

16

the effects of H2O addition shift the onset temperature of oxidation to the lower values,

17

inhibit CO formation significantly and enhance NO formation remarkably. The chemical

18

kinetic mechanism, which was hierarchically structured and updated in our previous work,

19

captured the main characteristics of CO and NO formation satisfactorily. The presence of

20

H2O leads to far higher OH radical concentrations in the CO2/H2O atmospheres. The

21

ultrahigh OH radical concentrations dramatically enhance the reactions between OH and 1

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

22

amine radicals, resulting in the significant enhancement of pathway NH2 → NH → HNO

23

→ NO and NH2 → NH → N → NO in CO2/H2O atmospheres. Meanwhile, NH2 →

24

CHxNHy/HNCO → NCO → NO is vastly demoted in CO2/H2O atmospheres. The

25

increase in pathways NH2 → NH → HNO → NO and NH2 → NH → N → NO is always

26

much more than the decline in pathway NH2 → CHxNHy/HNCO → NCO → NO. Hence,

27

H2O addition in oxy-fuel combustion enhances NO formation during the oxidation of

28

methane and ammonia. In addition, the effects of H2O addition become stronger on

29

enhancing NO formation with the increasing H2O concentration in CO2/H2O atmospheres

30

by further amplifying the amount of OH radicals.

31 32

Keywords: Oxy-fuel combustion; H2O addition; Reaction mechanism; Plug-flow reactor;

33

NO

34 35

1. .Introduction

36

Environmental crisis caused by greenhouse gases has drawn extensive attentions of

37

the international community in recent decades. Carbon dioxide (CO2) emitted from the

38

combustion of fossil fuels is the primary greenhouse gas at the present stage.1,2 In

39

response to the emission challenges, oxy-fuel combustion has been comprehensively

40

considered as a promising alternative technology for carbon capture and storage.3-8

41

Oxy-fuel combustion implies that the recycled flue gases are used to moderate the high

42

temperature generated by fuel combustion with pure oxygen, instead of air, ensuring that

43

the CO2 volume concentration exceeds 90% in the exhaust gas, which is almost

44

sequestration-ready. 2

ACS Paragon Plus Environment

Page 2 of 31

Page 3 of 31

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

45

It also has been found that NO emissions can be reduced during oxy-fuel

46

combustion compared with conventional air combustion for coal,9-14 hence, the

47

mechanisms and characteristics of NO formation impacted by CO2 has attracted

48

considerable attentions in recent years.15-18 Due to the nitrogen chemistry mechanisms for

49

coal combustion are extremely complicated, homogenous gas-phase chemistry was

50

chosen to provide theoretical foundation for the underlying mechanisms in many

51

investigations. Kim et al.19 numerically studied the effect of CO2 addition on NO

52

emission in H2/N2 laminar diffusion flame and they pointed out that the C-related

53

reactions affect the production of prompt NO in the case of CO2 addition. Park et al.20

54

numerically investigated the chemical effect of CO2 dilution on NO emission

55

characteristic in methane–air counterflow diffusion flame and results showed that the

56

mole production rates of nitrogenous species are prevented considerably. Park et al.21 also

57

computationally examined NO emission behavior in methane oxy-fuel combustion

58

recirculated with CO2 and they pointed out that the chemical effects of recirculated CO2

59

not only reduce the formation and destructions of NO through the Fenimore mechanism

60

but also suppress the NO formation through the thermal mechanism. The chemical effects

61

of CO2 addition in ethylene diffusion flame were demonstrated by Liu et al.22 and

62

numerical results showed that CO2 reduces NO emissions and reaction CO2 + H = CO +

63

OH is primarily responsible for the chemical effects of CO2 addition. Meanwhile, due to

64

the absence of N2 in oxy-fuel combustion, NO emissions almost completely derive from

65

fuel-NO. Hence, the conversions of hydrogen cyanide and ammonia, which are main

66

precursors of NO in the coal combustion, to NO become the focus of researches is this

67

area. Giménez-López et al.23 studied the oxidation of HCN in O2/CO2 atmospheres in a

3

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

68

flow reactor both experimentally and numerically. It was found that CO2 + H = CO + OH

69

competes with O2 + H = O + OH, reducing the formation of chain carriers, which clearly

70

leads to inhibiting HCN oxidation. In the experimental and numerical investigations of

71

ammonia oxidation in O2/CO2 atmospheres, both Mendiara et al.24 and Watanabe et al.25

72

concluded that the increased OH/H ratio and high CO levels increase the probability of

73

forming N2 instead of NO.

74

The review of literature above is mainly concerned in CO2 addition and O2/CO2

75

combustion atmospheres. In fact, steam is also a main composition in flue gases during

76

oxy-fuel combustion.26 Therefore, oxy-fuel combustion with H2O addition has drawn

77

researchers’ attentions recently. The effects of steam addition on coal gasification,27 coal

78

ignition28-30 and methane oxidation31 during oxy-fuel combustion have been investigated,

79

however, little information is available for the effects of steam addition on NO formation

80

during oxy-fuel combustion.

81

In this work, a systematic experimental study of the oxidation of methane and

82

ammonia in O2/CO2/H2O atmospheres was carried out at atmospheric pressure with

83

equivalence ratios ranging from fuel-lean to fuel-rich conditions (i.e. 0.2, 1.0, 1.6) and

84

temperature from 973 to 1773 K. The experiments were performed in a laboratory plug

85

flow reactor. Ammonia was chosen because it is the main N-compound in the

86

devolatilization of biomass and low-rank coal,5,32 and its production is greater than that of

87

HCN in the presence of steam.27,33 The experimental results are analyzed in terms of a

88

chemical kinetic model with 168 species and 1208 reactions, which was hierarchically

89

structured and updated in our previous work.34 In addition, the effects of H2O

90

concentration on NO formation from NH3 were also addressed.

4

ACS Paragon Plus Environment

Page 4 of 31

Page 5 of 31

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

91 92

2. Experimental and Modeling

93

Fig. 1 schematically demonstrates the experimental apparatus used in the present

94

work. The apparatus is comprised of a gas feeding system, a VDM (Vapor Delivery

95

Module) system, a flow reactor, and a gas composition test system.

96

The flow reactor with an internal diameter of 12 mm and a length of 1100 mm was

97

constructed according to Skjøth-Rasmussen et al.35 for homogeneous gas-phase reactions.

98

Alumina was chosen as the material of the flow reactor in order to avoid the catalytic

99

effect on experimental results. The flow reactor was heated using an electrically heated

100

oven, which allows the maximum temperature up to 1800 K. The temperature profiles

101

within the reactor were measured using a type S thermocouple under inert conditions (1

102

L/min CO2). The uncertainty of the temperature measurement is ±4 K. Typical

103

temperature profiles are demonstrated in Fig. 2, which implies that the length of

104

isothermal reaction zone is approximately 700 mm. The temperature of the isothermal

105

zone is referred to as the reaction temperature in this work.

106

High-purity gases (99.99%) were separately supplied from gas cylinders and the

107

flow rates were controlled precisely by mass flow controllers. The Bronkhorst VDM

108

(Vapor Delivery Module) system, a compact integrated system to realize mass flow

109

control of vapor, was used to add H2O vapor to O2/CO2 atmospheres. The VDM system

110

was intended to generate a predefined vapor flow, accruing from an accurately controlled

111

distilled water mass flow injected into an accurately controlled carrier gas (CO2) flow

112

with subsequent evaporation inside a temperature controlled chamber. This predefined

113

vapor flow was preliminarily blended with the oxygen and remaining carbon dioxide in a

5

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

114

isothermal chamber at 473 K. Then the total flow was premixed sufficiently with

115

methane and ammonia in the mixer prior to entering the reactor. In order to avoid steam

116

condensation, the heating tapes were used to heat the pipeline connecting the isothermal

117

chamber, the mixer and the reactor to maintain the temperature at a constant value of 423

118

K.

119

A water cooler was installed at the outlet of the reactor to cool down the product gas

120

rapidly. In the case of steam addition, argon (Ar) of equal volume to steam was fed into

121

the product gas to compensate for volume loss due to steam condensation. The

122

concentrations of CO and NO in the product gas were measured on-line using Fourier

123

transform infrared spectroscopy (GASMET-DX4000) with a resolution of 8 cm−1 and

124

scanning speed of 10 scans per second. The uncertainty of the measurement is estimated

125

as ±1%.

126

In order to minimize the axial dispersion of the reactants and comply with a

127

reasonable plug-flow approximation, the total flow rate for all experiments was kept

128

constant at 1 L/min (Standard Temperature and Pressure), which has been investigated

129

and validated by Skjøth-Rasmussen35 and Glarborg et.al.24 Also, the reactants were

130

highly diluted with carbon dioxide or carbon dioxide and steam to minimize the influence

131

of temperature rise due to chemical reaction. The methane concentration was about 2500

132

ppm, the ammonia concentration was approximately 500 ppm, and the oxygen

133

concentration was calculated according to the defined equivalence ratios. In consideration

134

of the application of rich and lean equivalence ratios in staged combustion, the

135

experiments were conducted covering a wide range of equivalence ratios to investigate

136

the effect of H2O addition on the oxidation of methane and ammonia during oxy-fuel

6

ACS Paragon Plus Environment

Page 6 of 31

Page 7 of 31

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

137

combustion. Three equivalence ratios, i.e. 0.2, 1.0 and 1.6, were chosen on behalf of

138

fuel-lean, stoichiometric, and fuel-rich conditions, respectively. And the calculations of

139

the equivalence ratios were based on the oxidation reaction as follows:

140

xCH4 + yNH3 + (2x+1.25y)O2 → xCO2 + yNO + (2x+1.5y)H2O

141

ϕ=

142

Experiments were carried out in the temperature range of 973–1773 K at intervals of

143

(CH 4 + NH 3 ) / O2 [(CH 4 + NH3 ) / O2 ] stoic

(1) (2)

20 K. Detailed experimental conditions are listed in Table 1.

144

Numerical predictions adopting full experimental temperature profiles within the

145

flow reactor were performed using the plug flow reactor (PFR) module in conjunction

146

with CHEMKIN-PRO. The mechanism adopted in this work was hierarchically

147

established for the oxidation of methane and ammonia in O2/H2O atmospheres in our

148

previous study34, containing the comprehensive oxidation mechanism for hydrogen,

149

C1–C2 hydrocarbons, nitrogen-containing species (HCN, NH3) and the interactions of

150

these compositions. The mechanism contained 170 species and 1208 reactions. The more

151

details and validations of this mechanism can be found elsewhere34.

152 153

3. .Results and discussion

154

Fig. 3 shows the comparisons between the experimental and numerical results of the

155

CO and NO profiles as a function of reaction temperature at two different atmospheres

156

(O2/CO2 and O2/CO2/H2O) and equivalence ratios (fuel-rich, stoichiometric, and fuel-lean

157

conditions). Closed symbols and solid lines represent experimental and numerical results

158

in CO2 atmospheres, whereas open symbols and dashed lines represent those in CO2/H2O 7

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

159

atmospheres.

160

Under fuel-rich conditions, as displayed in Fig. 3a, the CO formation in CO2 case

161

occurs at 1113 K, increases sharply to peak (5985 ppm) at 1393 K, and then increases

162

again moderately with temperature above 1453 K. The NO formation increases extremely

163

slowly at the beginning until it steps to 56 ppm at 1413 K and then increases very slowly

164

up to 108 ppm at 1773 K. In CO2/H2O case, CO formation initiates at 1093 K, goes up to

165

peak (3830 ppm) at 1373 K, and then increases slightly with temperature above 1393 K.

166

Similar to that in CO2 case, the NO concentration jumps to 78 ppm at 1393 K and then

167

increases gradually up to 166 ppm at 1773 K.

168

Under stoichiometric conditions exhibited in Fig. 3b, the CO formation in CO2 case

169

starts to rise at 1133 K, exhibits a peak (3104 ppm) at 1333 K, and then increase again

170

gradually with reaction temperature above 1393 K. The NO formation suddenly steps to

171

85 ppm at 1353 K and then increases slightly with reaction temperature up to 142 ppm at

172

1773 K. In CO2/H2O case, CO formation exhibits a semblable tendency with a peak value

173

of 3088 ppm at 1313 K. The NO concentration jumps to 100 ppm at 1333 K, and then

174

increases gradually to a steady value of 191 ppm above 1633 K.

175

Under fuel-lean conditions, as shown in Fig. 3c, the CO formation increases and

176

decreases dramatically exhibiting a maximum value of 353 ppm at 1113 K, and then

177

increases again with temperature in CO2 case. The NO formation increases gradually up

178

to 199 ppm at 1453 K and then nearly levels off. In CO2/H2O case, the CO formation

179

demonstrates a maximum value of 673 ppm at 1113 K, decreases to zero and remains

180

undetected until it increases again extremely slowly with temperature above 1453 K. The

181

NO formation increases gradually up to 242 ppm at 1533 K and then remains nearly

8

ACS Paragon Plus Environment

Page 8 of 31

Page 9 of 31

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

182

Energy & Fuels

unchanged.

183

Synthesizing the comparisons of the results observed in both CO2 and CO2/H2O

184

atmospheres, the effects of H2O addition on the oxidation of methane and ammonia

185

during oxy-fuel combustion can be summarized as follows: (1) it shifts the onset

186

temperature of oxidation to the lower values; (2) it inhibits CO formation significantly; (3)

187

it enhances NO formation markedly.

188

It can be seen from Fig. 3 that the chemical kinetic model satisfactorily reproduces

189

the main features of CO and NO formation measured in experiments in both CO2 and

190

CO2/H2O atmospheres, although tolerant deviations exist especially under fuel-lean

191

conditions. Hence, the mechanism proposed previously is appropriate for revealing the

192

effects of H2O addition on the ammonia oxidation during oxy-fuel combustion of

193

methane.

194

In the process of fuel oxidation and pollutants formation the radical pool has been

195

generally considered as a critical participant, hence, it is always dispensable to analyze

196

the radical pool structure at first. Fig. 4 compares the H, O and OH mole fractions

197

profiles in CO2 and CO2/H2O atmospheres under fuel-rich, stoichiometric and fuel-lean

198

conditions at 1673 K. As shown in Fig. 4, OH radicals are consistently predominant in

199

radical pool under all three conditions for both CO2 and CO2/H2O atmospheres. It also

200

can be seen that OH radicals in CO2/H2O atmospheres are much higher while H and O

201

radicals are nearly tantamount compared with those in CO2 atmospheres for different

202

equivalence ratios. In CO2 cases, R33 (H + CO2 = CO + OH) competes for H radicals

203

with the main chain branching reaction R1 (H + O2 = O + OH) leading to the dominant

204

position of OH radicals. However, in CO2/H2O cases, the presence of H2O substantially

9

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 31

205

enhances the reactions R30 (H + H2O = OH + H2) and R14 (O + H2O = OH + OH),

206

yielding a large amount of OH radicals and simultaneously consuming large amounts of

207

H and O radicals. Hence, OH radicals are far higher in CO2/H2O atmospheres than those

208

in CO2 atmospheres. It also can be found that O radicals increase to some extent under

209

fuel-lean conditions for both CO2 and CO2/H2O atmospheres. And this can be attributed

210

to the fact that the increasing O2 concentration strengthens R1 (H + O2 = O + OH)

211

subsequently producing more O radicals under fuel-lean conditions.

212 213 214

In order to clarify the NO formation mechanism intuitively, the OPR, which is short for the overall production rate, is introduced and defined as follows: l

OPRi , j = ∫ ωi , j dx

(3)

0

215

where i denotes species, j denotes elementary reaction, ωi,j means the mole production

216

rate of species i through elementary reaction j and l is the length of reaction zone.

217

Fig. 5 compares the OPRNO of importantly contributing reaction steps between CO2

218

and CO2/H2O atmospheres covering from fuel-rich to fuel-lean conditions. The top ten

219

reactions contributing to NO formation for each case are extracted to be synthesized in

220

Fig. 5. It is compelling that significant differences exist in common among fuel-rich,

221

stoichiometric and fuel-lean conditions between CO2 and CO2/H2O atmospheres, which

222

are summarized as follows: (1) the contribution to NO of R755 (HNO + OH = NO + H2O)

223

is enhanced vastly in CO2/H2O cases; (2) the OPR of R857 (N + OH = NO + H) is

224

strengthened dramatically in CO2/H2O cases; (3) the NCO-related reactions R1005 (NCO

225

+ O = NO + CO) and R1009 (NCO + O2 = NO + CO2), which are dominant in CO2

226

atmospheres, are suppressed sharply in CO2/H2O cases. It can be seen that the presence of

227

H2O alters the structure of radical pool, resulting in a remarkable variation of oxidation 10

ACS Paragon Plus Environment

Page 11 of 31

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

228

pathway of NH3. Hence, it is indispensable to carry out an elaborate analysis on the

229

oxidation pathways of NH3.

230

Fig. 6 compares the critical NO formation pathways between the CO2 and CO2/H2O

231

atmospheres. As demonstrated in Fig. 6, NH3 is initially converted into NH2 through

232

hydrogen abstraction reaction. Then, the pathways from NH2 to NO can be generally

233

divided into five main pathways as follows: (a) NH2 → HNO → NO; (b) NH2 → NH →

234

NO; (c) NH2 → NH → HNO → NO; (d) NH2 → NH → N → NO; and (e) NH2 →

235

CHxNHy/HNCO → NCO → NO.

236

For NH2 conversion, R829 (NH2 + OH = NH + H2O) is enhanced by the abundant

237

OH radicals in CO2/H2O atmospheres, implying that NH2 converts more to NH radicals

238

instead of reacting with other species. Correspondingly, the channel of NH2 → CHxNHy

239

through R1090 (CH3 + NH2 = CH3NH2) and R1091 (CH3 + NH2 = CH2NH2 + H) is

240

weakened dramatically and ultimately the channel of NH2 → CHxNHy/HNCO → NCO is

241

suppressed. Meanwhile, the promotion of the channel of NH2 → NH in CO2/H2O

242

atmospheres alters the availability of NH2 and NH radicals, which is the reason for the

243

change of main NO reduction reactions R839 (NH2 + NO = N2 + H2O), R840 (NH2 + NO

244

= NNH + OH) and R853 (NH + NO = N2O + H) as displayed in Fig. 5. In addition, it is

245

noteworthy that the channel of NH2 → HNCO is reversed in CO2/H2O atmospheres. The

246

ultrahigh CO concentration in CO2 atmospheres promotes HNCO formation through

247

-R978 (NH2 + CO = HNCO + H). However, H2O addition decreases the CO

248

concentration sharply enough to reverse -R978 (NH2 + CO = HNCO + H) in CO2/H2O

249

atmospheres. Therefore, the channel of NH2 → HNCO → NCO is forbidden in CO2/H2O

250

atmospheres.

11

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 31

251

For NH conversion, the channels of NH → HNO and NH → N are facilitated

252

through R846 (NH + OH = HNO + H) and R847 (NH + OH = N + H2O) , yielding larger

253

amount of HNO and N radicals in CO2/H2O atmospheres. This can be attributed to more

254

sufficient NH and OH radicals in CO2/H2O atmospheres compared with those in CO2

255

atmospheres. Subsequently, sufficient HNO, N and OH radicals strengthen R775 (HNO +

256

OH = NO + H2O) and R857 (N + OH = NO + H) as shown in Fig. 5 , leading to the

257

enhancement of pathway (c) NH2 → NH → HNO → NO and (d) NH2 → NH → N →

258

NO in CO2/H2O atmospheres.

259

For NCO conversion, NCO is almost all derived from the channel of NH2 →

260

CHxNHy/HNCO → NCO. As mentioned above, the channel of NH2 → CHxNHy/HNCO

261

→ NCO is remarkably suppressed in CO2/H2O atmospheres, leading to the suppression

262

of NCO formation. Hence, R1005 (NCO + O = NO + CO) and R1009 (NCO + O2 = NO

263

+ CO2) are inhibited dramatically for NO formation, as show in Fig. 5, implying that

264

pathway (e) NH2 → CHxNHy/HNCO → NCO → NO is vastly demoted in CO2/H2O

265

atmospheres.

266

In order to reveal the NO formation mechanism quantitatively from pathway point

267

of view, the conversion rate of nitrogen was introduced to evaluate the conversion of NH3

268

to NO, which is defined as:

269 270

CRNO =

M NO MN

(4)

where MNO designates the mass of N in NO and MN represents the mass of N in NH3.

271

The conversion rate (CRNO) of each pathway, which is obtained based on the ORP of

272

all the relevant elementary reactions along the NO formation pathways, is compared

273

between CO2 and CO2/H2O atmospheres demonstrated in Fig. 7. It can be seen from Fig. 12

ACS Paragon Plus Environment

Page 13 of 31

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

274

7 that pathway (e) NH2 → CHxNHy/HNCO → NCO → NO and (c) NH2 → NH → HNO

275

→ NO are dominant for NO formation in CO2 and CO2/H2O atmosphere, respectively.

276

Although pathway (a) NH2 → HNO → NO is fairly critical under fuel-lean condition,

277

which is attributed to that the extremely high O2 concentration enhances the channel of

278

NH2 → HNO through R833 (NH2 + O2 = HNO + OH), the effect of H2O addition on

279

pathway (a) is limited as shown in Fig. 7. As discussed in the previous section, the H2O

280

addition in oxy-fuel combustion strengthens pathway (c) NH2 → NH → HNO → NO

281

significantly reflecting in that the increase value of CRNO through pathway (c) is 12.4%,

282

12.8% and 9.3% under fuel-rich, stoichiometric, and fuel-lean conditions, respectively. It

283

also enhances pathway (d) NH2 → NH → N → NO reflecting in that the increase value of

284

CRNO through pathway (d) is 4.1%, 4.5% and 2.8% under fuel-rich, stoichiometric, and

285

fuel-lean conditions, respectively. Meanwhile, the H2O addition weakens pathway (e)

286

NH2 → CHxNHy/HNCO → NCO → NO markedly embodied in that the decline value of

287

CRNO through pathway (e) is 10.8%, 10.6% and 4.3% under fuel-rich, stoichiometric, and

288

fuel-lean conditions, respectively. It can be found that the increase of pathway (c) and (d)

289

is always superior to the decline of pathway (e) in CO2/H2O atmospheres. Hence, H2O

290

addition in oxy-fuel combustion enhances NO formation during the oxidation of methane

291

and ammonia.

292

In addition, the effect of H2O concentration on NO formation will be discussed in

293

the following section. Fig. 8 shows the comparisons between the experimental and

294

numerical results of the CO and NO profiles as a function of H2O concentration in

295

CO2/H2O atmospheres at different equivalence ratios (fuel-rich, stoichiometric, and

296

fuel-lean conditions). As displayed in Fig. 8, the increasing H2O concentration inhibits

13

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

297

the CO formation and enhances the NO formation. Meanwhile, the prediction results are

298

in good agreements with the experiments data, and this offers a further validation for the

299

applicability of the present mechanism. Fig. 9 demonstrates that the variations of the

300

conversion rate (CRNO) of each pathway with the increasing H2O concentration (5%, 15%

301

and 30%) at different equivalence ratios (fuel-rich, stoichiometric and fuel-lean

302

conditions) in CO2/H2O atmospheres. It indicates that pathway (c) NH2 → NH → HNO

303

→ NO and (d) NH2 → NH → N → NO are dramatically strengthened with the increasing

304

H2O concentration while pathway (e) NH2 → CHxNHy/HNCO → NCO → NO is

305

evidently suppressed. Specially, pathway (a) NH2 → HNO → NO also decreases with the

306

increasing H2O concentration under fuel-lean conditions. However, the increase of

307

pathway (c) and (d) with the increasing H2O concentration is always overwhelmingly

308

dominant in CO2/H2O atmospheres which in responsible for the increasing NO formation.

309

The fundamental reason for this is that the increasing H2O concentration further enhances

310

the reactions R30 (H + H2O = OH + H2) and R14 (O + H2O = OH + OH) yielding a

311

larger amount of OH radicals, which amplifies the effect of H2O addition on the

312

oxidation of methane and ammonia during oxy-fuel combustion.

313 314

5. Conclusions

315

The effect of H2O addition on the oxidation of methane and ammonia during

316

oxy-fuel combustion was investigated both experimentally and numerically. Comparison

317

experiments between CO2 and CO2/H2O atmospheres were accomplished in a flow

318

reactor at atmospheric pressure with equivalence ratios ranging from fuel-rich to

319

fuel-lean and temperature from 973 K to 1773 K.

14

ACS Paragon Plus Environment

Page 14 of 31

Page 15 of 31

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

320

The comparative experiments results indicate that the effects of H2O addition on the

321

oxidation of methane and ammonia during oxy-fuel combustion shift the onset

322

temperature of oxidation to the lower values, inhibit CO formation significantly and

323

enhance NO formation markedly. The underlying mechanisms have been revealed using

324

a detailed chemical kinetic mechanism in the production rate and pathway analysis point

325

of view.

326

The H2O addition substantially enhances the reactions H + H2O = OH + H2 and O +

327

H2O = OH + OH, leading to far higher OH radical concentrations in the CO2/H2O

328

atmospheres than those in the CO2 atmospheres. The ultrahigh OH radical concentrations

329

dramatically enhance the reactions between OH and amine radicals (NH2, NH, and N),

330

resulting in the significant enhancement of pathway NH2 → NH → HNO → NO and NH2

331

→ NH → N → NO in CO2/H2O atmospheres. NH2 radicals are converted far more to NH

332

radicals inhibiting the channel of NH2 → CHxNHy. Meanwhile, the channel of NH2 →

333

HNCO → NCO is forbidden in CO2/H2O atmospheres by reversing NH2 + CO = HNCO

334

+ H. Consequently, pathway NH2 → CHxNHy/HNCO → NCO → NO is vastly demoted

335

in CO2/H2O atmospheres. However, the increase of pathway NH2 → NH → HNO → NO

336

and NH2 → NH → N → NO is always much more than the decline of pathway NH2 →

337

CHxNHy/HNCO → NCO → NO in CO2/H2O atmospheres. Hence, H2O addition in

338

oxy-fuel combustion enhances NO formation during the oxidation of methane and

339

ammonia. In addition, the effects of H2O addition become stronger on enhancing NO

340

formation with the increasing H2O concentration in CO2/H2O atmospheres by further

341

amplifying the amount of OH radicals.

342

15

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

343 344 345

Acknowledgements This work was supported by the National Key Research & Development Special Project (No. 2016YFB0600801) of the National Natural Science Foundation of China.

346 347

References

348

(1) Solomon, S.; Qin, D.; Manning, M.; Chen, Z.; Marquis, M.; Averyt, K.; Tignor, M.;

349

Miller, H. IPCC, Climate change 2007: the physical science basis. Contribution of

350

working group I to the fourth assessment report of the intergovernmental panel on

351

climate change. Cambridge University Press, Cambridge, United Kingdom and New

352

York, NY, USA: 2007.

353

(2) Buhre, B.; Elliott, L.; Sheng, C.; Gupta, R.; Wall, T., Oxy-fuel combustion

354

technology for coal-fired power generation. Progress in energy and combustion science

355

2005, 31, (4), 283-307.

356

(3) Scheffknecht, G.; Al-Makhadmeh, L.; Schnell, U.; Maier, J., Oxy-fuel coal

357

combustion—A review of the current state-of-the-art. International Journal of

358

Greenhouse Gas Control 2011, 5, Supplement 1, (0), S16-S35.

359

(4) Chen, L.; Yong, S. Z.; Ghoniem, A. F., Oxy-fuel combustion of pulverized coal:

360

Characterization, fundamentals, stabilization and CFD modeling. Progress in Energy and

361

Combustion Science 2012, 38, (2), 156-214.

362

(5) Toftegaard, M. B.; Brix, J.; Jensen, P. A.; Glarborg, P.; Jensen, A. D., Oxy-fuel

363

combustion of solid fuels. Progress in Energy and Combustion Science 2010, 36, (5),

364

581-625.

365

(6) Wall, T. F., Combustion processes for carbon capture. Proceedings of the

16

ACS Paragon Plus Environment

Page 16 of 31

Page 17 of 31

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

366

Combustion Institute 2007, 31, (1), 31-47.

367

(7) Wall, T.; Liu, Y.; Spero, C.; Elliott, L.; Khare, S.; Rathnam, R.; Zeenathal, F.;

368

Moghtaderi, B.; Buhre, B.; Sheng, C.; Gupta, R.; Yamada, T.; Makino, K.; Yu, J., An

369

overview on oxyfuel coal combustion—State of the art research and technology

370

development. Chemical Engineering Research and Design 2009, 87, (8), 1003-1016.

371

(8) Edge, P.; Gharebaghi, M.; Irons, R.; Porter, R.; Porter, R.; Pourkashanian, M.; Smith,

372

D.; Stephenson, P.; Williams, A., Combustion modelling opportunities and challenges for

373

oxy-coal carbon capture technology. Chemical Engineering Research and Design 2011,

374

89, (9), 1470-1493.

375

(9) Andersson, K.; Normann, F.; Johnsson, F.; Leckner, B., NO emission during

376

oxy-fuel combustion of lignite. Industrial & Engineering Chemistry Research 2008, 47,

377

(6), 1835-1845.

378

(10) Lupiáñez, C.; Guedea, I.; Bolea, I.; Díez, L. I.; Romeo, L. M., Experimental study of

379

SO2 and NOx emissions in fluidized bed oxy-fuel combustion. Fuel Processing

380

Technology 2013, 106, 587-594.

381

(11) Czakiert, T.; Bis, Z.; Muskala, W.; Nowak, W., Fuel conversion from oxy-fuel

382

combustion in a circulating fluidized bed. Fuel Processing Technology 2006, 87, (6),

383

531-538.

384

(12) Wang, G.; Zander, R.; Costa, M., Oxy-fuel combustion characteristics of

385

pulverized-coal in a drop tube furnace. Fuel 2014, 115, 452-460.

386

(13) Shaddix, C. R.; Molina, A., Fundamental investigation of NOx formation during

387

oxy-fuel combustion of pulverized coal. Proceedings of the Combustion Institute 2011,

388

33, (2), 1723-1730.

17

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

389

(14) Jia, L.; Tan, Y.; Anthony, E., Emissions of SO2 and NOx during Oxy−Fuel CFB

390

Combustion Tests in a Mini-Circulating Fluidized Bed Combustion Reactor. Energy &

391

Fuels 2009, 24, (2), 910-915.

392

(15) Normann, F.; Andersson, K.; Leckner, B.; Johnsson, F., Emission control of nitrogen

393

oxides in the oxy-fuel process. Progress in Energy and Combustion Science 2009, 35, (5),

394

385-397.

395

(16) Sung, C.; Law, C. In Dominant chemistry and physical factors affecting NO

396

formation and control in oxy-fuel burning, Symposium (International) on Combustion,

397

1998; Elsevier: 1998; pp 1411-1418.

398

(17) Krzywański, J.; Czakiert, T.; Muskała, W.; Nowak, W., Modelling of CO2, CO, SO2,

399

O2 and NOx emissions from the oxy-fuel combustion in a circulating fluidized bed. Fuel

400

Processing Technology 2011, 92, (3), 590-596.

401

(18) Seepana, S.; Jayanti, S., Flame structure and NO generation in oxy-fuel combustion

402

at high pressures. Energy Conversion and Management 2009, 50, (4), 1116-1123.

403

(19) Kim, S.-G.; Park, J.; Keel, S.-I., Thermal and chemical contributions of added H2O

404

and CO2 to major flame structures and NO emission characteristics in H2/N2 laminar

405

diffusion flame. International Journal of Energy Research 2002, 26, (12), 1073-1086.

406

(20) Park, J.; Kim, S.-G.; Lee, K.-M.; Kim, T. K., Chemical effect of diluents on flame

407

structure and NO emission characteristic in methane-air counterflow diffusion flame.

408

International Journal of Energy Research 2002, 26, (13), 1141-1160.

409

(21) Park, J.; Park, J. S.; Kim, H. P.; Kim, J. S.; Kim, S. C.; Choi, J. G.; Cho, H. C.; Cho,

410

K. W.; Park, H. S., NO Emission Behavior in Oxy-fuel Combustion Recirculated with

411

Carbon Dioxide. Energy & Fuels 2006, 21, (1), 121-129.

18

ACS Paragon Plus Environment

Page 18 of 31

Page 19 of 31

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

412

(22) Liu, F.; Guo, H.; Smallwood, G. J.; Gülder, Ö. L., The chemical effects of carbon

413

dioxide as an additive in an ethylene diffusion flame: implications for soot and NOx

414

formation. Combustion and Flame 2001, 125, (1), 778-787.

415

(23) Giménez-López, J.; Millera, A.; Bilbao, R.; Alzueta, M. U., HCN oxidation in an

416

O2/CO2 atmosphere: An experimental and kinetic modeling study. Combustion and

417

Flame 2010, 157, (2), 267-276.

418

(24) Mendiara, T.; Glarborg, P., Ammonia chemistry in oxy-fuel combustion of methane.

419

Combustion and Flame 2009, 156, (10), 1937-1949.

420

(25) Watanabe, H.; Marumo, T.; Okazaki, K., Effect of CO2 Reactivity on NOx

421

Formation and Reduction Mechanisms in O2/CO2 Combustion. Energy & Fuels 2012, 26,

422

(2), 938-951.

423

(26)Tan, Y. Oxy-fuel combustion for power generation and carbon dioxide (CO2)

424

capture. Woodhead Publishing Ltd press: 2011.

425

(27) Hecht, E. S.; Shaddix, C. R.; Geier, M.; Molina, A.; Haynes, B. S., Effect of CO2 and

426

steam gasification reactions on the oxy-combustion of pulverized coal char. Combustion

427

and Flame 2012, 159, (11), 3437-3447.

428

(28) Riaza, J.; Álvarez, L.; Gil, M.; Pevida, C.; Pis, J.; Rubiera, F., Effect of oxy-fuel

429

combustion with steam addition on coal ignition and burnout in an entrained flow reactor.

430

Energy 2011, 36, (8), 5314-5319.

431

(29) Marek, E.; Świątkowski, B., Reprint of “Experimental studies of single particle

432

combustion in air and different oxy-fuel atmospheres”. Applied Thermal Engineering

433

2015, 74, 61-68.

434

(30) Cai, L.; Zou, C.; Guan, Y.; Jia, H.; Zhang, L.; Zheng, C., Effect of steam on ignition

19

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

435

of pulverized coal particles in oxy-fuel combustion in a drop tube furnace. Fuel 2016,

436

182, 958-966.

437

(31) Wang, L.; Liu, Z.; Chen, S.; Zheng, C.; Li, J., Physical and Chemical Effects of CO2

438

and H2O Additives on Counterflow Diffusion Flame Burning Methane. Energy & Fuels

439

2013, 27, (12), 7602-7611.

440

(32) Becidan, M.; Skreiberg, Ø.; Hustad, J. E., NOx and N2O Precursors (NH3 and HCN)

441

in Pyrolysis of Biomass Residues. Energy & fuels 2007, 21, (2), 1173-1180.

442

(33) McKenzie, L. J.; Tian, F.-J.; Guo, X.; Li, C.-Z., NH3 and HCN formation during the

443

gasification of three rank-ordered coals in steam and oxygen. Fuel 2008, 87, (7),

444

1102-1107.

445

(34) He, Y.; Zou, C.; Song, Y.; Chen, W.; Jia, H.; Zheng, C., Experimental and Numerical

446

Study of the Effect of High Steam Concentration on the Oxidation of Methane and

447

Ammonia during Oxy-Steam Combustion. Energy & Fuels 2016, 30, (8), 6799-6807.

448

(35) Skjøth-Rasmussen, M. S.; Glarborg, P.; Østberg, M.; Johannessen, J.; Livbjerg, H.;

449

Jensen, A.; Christensen, T., Formation of polycyclic aromatic hydrocarbons and soot in

450

fuel-rich oxidation of methane in a laminar flow reactor. Combustion and Flame 2004,

451

136, (1), 91-128.

452 453 454 455 456 457 458 20

ACS Paragon Plus Environment

Page 20 of 31

Page 21 of 31

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

459

Figure captions

460

Fig. 1. Schematic diagram of the experimental apparatus.

461

Fig. 2. Temperature profiles within the reactor.

462

Fig. 3. Experimental data and numerical predictions for different equivalence ratios

463

(fuel-rich (a), stoichiometric (b), and fuel-lean (c) conditions) as functions of reaction

464

temperature.

465

Fig. 4. Comparison of H, O, and OH mole fraction profiles between O2/CO2 and

466

O2/CO2/H2O atmospheres for fuel-rich (a), stoichiometric (b) and fuel-lean (c) conditions

467

at 1673 K.

468

Fig. 5. Comparison of OPRNO between O2/CO2 and O2/CO2/H2O atmospheres for

469

fuel-rich (a), stoichiometric (b) and fuel-lean (c) conditions at 1673 K.

470

Fig. 6. Comparison of the NO formation pathways between the O2/CO2 (a) and

471

O2/CO2/H2O (b) atmospheres for stoichiometric condition at 1673 K.

472

Fig. 7. Comparison of the elemental N conversion rates (CRNO) obtained through each

473

pathway at 1673 K in O2/CO2 and O2/CO2/H2O atmospheres.

474

Fig. 8. Experimental data and numerical predictions for different equivalence ratios

475

(fuel-rich, stoichiometric, and fuel-lean conditions) as functions of H2O concentration at

476

1673 K in O2/CO2/H2O atmospheres.

477

Fig. 9. Comparison of the elemental N conversion rates (CRNO) among different H2O

478

concentrations at 1673 K in O2/CO2/H2O atmospheres.

479 480

Table captions

481

Table 1. Experimental conditions in the present work.

21

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Fig. 1. Schematic diagram of the experimental apparatus. 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 22

ACS Paragon Plus Environment

Page 22 of 31

Page 23 of 31

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

Fig. 2. Temperature profiles within the reactor. 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 23

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

a

b

c

Fig. 3. Experimental data and numerical predictions for different equivalence ratios (fuel-rich (a), stoichiometric (b), and fuel-lean (c) conditions) as functions of reaction temperature. 517

24

ACS Paragon Plus Environment

Page 24 of 31

Page 25 of 31

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

a

b

c

Fig. 4. Comparison of H, O, and OH mole fraction profiles between O2/CO2 and O2/CO2/H2O atmospheres for fuel-rich (a), stoichiometric (b) and fuel-lean (c) conditions at 1673 K. 518 519 520 521 522 523 524 525 25

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

526

a

b

c

Fig. 5. Comparison of OPRNO between O2/CO2 and O2/CO2/H2O atmospheres for fuel-rich (a), stoichiometric (b) and fuel-lean (c) conditions at 1673 K. 527 528 529 530 26

ACS Paragon Plus Environment

Page 26 of 31

Page 27 of 31

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

531

a

b

Fig. 6. Comparison of the NO formation pathways between the O2/CO2 (a) and O2/CO2/H2O (b) atmospheres for stoichiometric condition at 1673 K. 532 533 534 535 536 537 27

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Fig. 7. Comparison of the elemental N conversion rates (CRNO) obtained through each pathway at 1673 K in O2/CO2 and O2/CO2/H2O atmospheres. 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 28

ACS Paragon Plus Environment

Page 28 of 31

Page 29 of 31

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

Fig. 8. Experimental data and numerical predictions for different equivalence ratios (fuel-rich, stoichiometric, and fuel-lean conditions) as functions of H2O concentration at 1673 K in O2/CO2/H2O atmospheres. 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570

29

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Fig. 9. Comparison of the elemental N conversion rates (CRNO) among different H2O concentrations at 1673 K in O2/CO2/H2O atmospheres. 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 30

ACS Paragon Plus Environment

Page 30 of 31

Page 31 of 31

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

589 φ 1.6 1.0 0.2

Table 1. Experimental conditions in the present work. CH4/ppm NH3/ppm O2/ppm CO2/% H2O/% 2511 507 3531 99.35-H2O% 1,5,10,15,20,25,30 2505 505 5647 99.13-H2O% 1,5,10,15,20,25,30 2499 503 28276 96.86-H2O% 1,5,10,15,20,25,30

590

31

ACS Paragon Plus Environment