Ultraviolet Absorption Cross Sections of KOH and KCl for Nonintrusive

Mar 5, 2019 - An understanding of potassium chemistry in energy conversion processes supports the development of complex biomass utilization with high...
0 downloads 0 Views 731KB Size
Subscriber access provided by ECU Libraries

Article

UV absorption cross-sections of KOH and KCl for nonintrusive species-specific quantitative detection in hot flue gases Wubin Weng, Christian Brackmann, Tomas Leffler, Marcus Aldén, and Zhongshan Li Anal. Chem., Just Accepted Manuscript • DOI: 10.1021/acs.analchem.9b00203 • Publication Date (Web): 05 Mar 2019 Downloaded from http://pubs.acs.org on March 8, 2019

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 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 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.

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 28 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

Analytical Chemistry

2

UV absorption cross-sections of KOH and KCl for nonintrusive species-specific quantitative detection in hot flue gases

3

Wubin Weng,1 Christian Brackmann,1 Tomas Leffler,2 Marcus Aldén1 and Zhongshan Li1, *

4

1 Division

1

5 6

of Combustion Physics, Lund University, P.O. Box 118, SE-221 00, Lund, Sweden

2 R&D,

Strategic Development, Vattenfall AB, 814 26 Älvkarleby, Sweden * corresponding

author: [email protected]

7

Abstract Understanding of potassium chemistry in energy conversion processes supports

8

the development of complex biomass utilization with high efficiency and low pollutant

9

emissions. Potassium exists mainly as potassium hydroxide (KOH), potassium chloride

10

(KCl) and atomic potassium (K) in combustion and related thermochemical processes.

11

We report, for the first time, the measurement of the UV absorption cross-sections of

12

KOH and KCl at temperatures between 1300 and 1800 K using a newly developed

13

method. Using the spectrally-resolved UV absorption cross-sections, the concentrations

14

of KOH and KCl were measured simultaneously. Additionally, we measured the

15

concentrations of K atom using tunable diode laser absorption spectroscopy both at 404.4

16

nm and 769.9 nm. The 404.4 nm line was utilized to expand the measurement dynamic

17

range to higher concentrations. A constant amount of KCl was seeded into premixed

18

CH4/air flames with equivalence ratios varied from 0.67 to 1.32, and the concentrations

19

of KOH, KCl and K atoms in the hot flue gas were measured nonintrusively. The results

20

indicate that these techniques can provide comprehensive data for quantitative

21

understanding of the potassium chemistry in biomass combustion/gasification.

22 23

1 ACS Paragon Plus Environment

Analytical Chemistry 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

24

Biomass fuels, especially herbaceous, usually have a high content of potassium which

25

will released as potassium hydroxide (KOH), potassium chloride (KCl) and potassium

26

atom (K) during combustion and gasification.1-3 These potassium species, especially KCl,

27

can cause severe operation problems of combustors and gasifiers, such as slagging,

28

fouling and high temperature corrosion.4 Thus, knowledge and understanding of the

29

formation and release of potassium species is desired.

30

Optical diagnostics have been widely employed for concentration measurements of

31

potassium species in many related studies.5-8 Laser-induced breakdown spectroscopy

32

(LIBS) was adopted to measure the total amount of potassium in flames9, 10 while excimer

33

laser-induced fragmentation fluorescence (ELIF) was used to detect the total amount of

34

KOH and KCl compounds.11, 12 Planar laser-induced fluorescence (PLIF) has been used

35

in measuring K atoms released from a burning biomass pellet.8 To obtain reliable

36

quantitative results, the application of absorption spectroscopic techniques is essential.

37

Broadband UV absorption spectroscopy is a powerful tool for measurement of KCl and

38

KOH,13-15 and tunable diode laser absorption spectroscopy (TDLAS) has been applied to

39

measure K atoms.7, 16 Collinear photofragmentation and atomic absorption spectroscopy

40

(CPFAAS) was developed by Sorvajärvi et al.17 for simultaneous detection of KCl, KOH

41

and K atoms.

42

In both UV absorption spectroscopy and the CPFAAS technique, absolute values of the

43

UV absorption cross-section of KOH and KCl are indispensable in obtaining absolute

44

concentrations. The UV absorption cross-sections of KCl have been determined in

2 ACS Paragon Plus Environment

Page 2 of 28

Page 3 of 28 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

Analytical Chemistry

45

previous investigations,18-20 but these values are only available for temperatures up to

46

1200 K. The UV absorption cross-sections of KOH and KCl at elevated temperatures,

47

valid in combustion environments, are still lacking. Rowland and Makide21 estimated the

48

absorption cross-sections of NaOH in flames based on the measurement of Daidoji,22 but

49

the values relied on the calculated amount of NaOH seeded into the flame, which

50

introduced a severe uncertainty. Recently, Weng et al.23 measured the UV absorption

51

spectrum of KOH and KCl in flame environments, and the UV absorption cross-sections

52

were evaluated based on the KCl absorption data presented by Leffler et al.20, which were

53

obtained at 1073 K. In the study of Weng et al.23, the influence of even higher temperature

54

on the absorption cross-section values was assumed to be negligible.

55

In the present work, a new method is developed to directly evaluate the UV absorption

56

cross-sections of KOH and KCl in environments with temperatures of around 1300 K and

57

1800 K. Using the obtained cross-sections, quantitative measurements of potassium

58

species were made in flames seeded with a constant amount of KCl, but with equivalence

59

ratios varied from 0.67 to 1.32 and a post-flame temperature of around 1800 K.

60

Comprehensive experimental data were obtained as a substantial improvement in the

61

study of the potassium chemistry compared the previous researches based on the

62

measurement of K atoms25, 26 or the total amount of KCl and KOH.15 The concentration

63

variation with equivalence ratios is discussed. The chemical balance between potassium

64

species in different flame environments is presented.

65

3 ACS Paragon Plus Environment

Analytical Chemistry 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

66

Methodology

67

Burner and flame conditions. Homogenous hot flue gas environments in a series of

68

well-defined conditions are prepared using an in-house developed multi-jet burner,27

69

which has a rectangular outlet sized 85 mm × 47 mm. As shown in Fig. 1, the burner

70

consists of two chambers, namely the jet chamber and the co-flow chamber. The jet

71

chamber supplies premixed gases to 181 cone-shaped laminar flames stabilized on an

72

array of jet nozzles, and the co-flow chamber supplies co-flow gases through a perforated

73

plate with holes surrounding each premixed jet flame evenly. The temperature and

74

composition of the mimicked hot flue gas can be controlled by adjusting the composition

75

and flow rates of the inlet gases, which are controlled via well-calibrated mass flow

76

controllers (Bronkhorst). The flame conditions used in this study are presented in Table

77

1.

78

Table 1. Flame conditions. Flame Case Flame 1 Flame 2 Flame 3 Flame 4 Flame 5 Flame 6 Flame 7 Flame 8 Flame 9 Flame 10

CH4 2.66 2.66 2.66 2.66 3.05 3.14 3.23 2.28 2.09 2.66

Gas flow rate (sl/min) Global Gas product Jet-flow Co-flow equivalence temperature T (K) ratio ϕ Air O2 N2 Air 17.34 1.89 6.95 11.61 0.67 1770 17.34 1.89 10.84 7.74 0.74 1750 17.34 1.89 14.21 4.37 0.83 1760 17.34 1.89 18.60 0 0.96 1790 17.11 1.86 13.95 0 1.12 1840 15.53 1.91 12.09 0 1.22 1890 14.16 1.93 9.30 0 1.32 1750 11.89 2.26 22.69 9.82 0.67 1390 10.90 2.07 26.51 10.66 0.63 1260 9.12 2.15 18.60 0 1.31 1470

79 80

4 ACS Paragon Plus Environment

Page 4 of 28

Page 5 of 28 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

Analytical Chemistry

81 82 83

Figure 1. Schematic of the setup for the burner system (a), the TDLAS system and the UV absorption system with the burner from the top view (b).

84

The investigated flames have different global equivalence ratios and temperatures (cf.

85

Table 1). The global equivalence ratio is calculated based on the composition of the total

86

inlet gases, indicating whether an oxidative (ϕ 1.0) environment

87

of the hot flue gas. The temperature of the hot flue gas was measured using two-line

88

atomic fluorescence thermometry (TLAF) which will be described in the following

89

section.

90

Water solutions with 0.5 mol/l for potassium carbonate (K2CO3) or 1.0 mol/l for KCl

91

were prepared and seeded into the flames. A fog of the solution, generated by an

92

ultrasonic generator, was transported by an air flow with a flow rate of 0.41 sl/min into

93

the jet-chamber and thus seeded into the premixed jet flames together with the reactant

94

gases. The K2CO3 or KCl in the fog was vaporized and decomposed as it passed through

95

the flame front region, and gas-phase potassium species, mainly KOH, KCl and K atoms,

96

were formed in the hot flue gases. Chloroform (CHCl3) was used as another source of

5 ACS Paragon Plus Environment

Analytical Chemistry 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

97

chlorine in the hot flue gas, besides potassium chloride, to control the ratio between

98

potassium and chlorine. Liquid chloroform was filled in a gas bubbler bottle stored in a

99

temperature bath (PolySicence) regulated at 273 K. The chloroform vapor was

100

transported by a nitrogen flow into the jet-chamber with a known concentration (a vapor

101

pressure of 0.0779 bar at 273 K).28 Assuming that all chloroform was decomposed

102

through the flame fronts,29,

103

calculated, and the value was controlled by adjusting the flow rate of the carrier gas.

30

the concentration of chlorine in the hot flue gas was

104 105

Temperature measurements. The temperature of the hot flue gas about 5 mm above the

106

burner outlet was measured with the TLAF technique and the results are presented in

107

Table 1 with an uncertainty of ~3%.31 The details of the TLAF technique using indium

108

atoms have been described by Borggren et al.31, 32. The present study used a similar setup,

109

which mainly consisted of three parts, i.e. the indium seeding system, the laser system

110

and the imaging system. For indium seeding, an indium chloride (InCl3) solution was

111

used in the same system employed for potassium seeding. After passage through the flame,

112

free indium atoms were produced to be used as the temperature marker. The laser system

113

consists of two external cavity diode lasers (Toptica, DL100pro and DL 100) controlled

114

by two separated analog control units, each including a temperature control module (DTC

115

110) and a current control module (DCC 110). Two continuous-wave laser beams were

116

produced at wavelengths 410 nm and 451 nm. Both lasers have a power of about 5 mW

117

and a beam size of about 1 mm2. The laser beams were overlapped using a dichroic mirror,

6 ACS Paragon Plus Environment

Page 6 of 28

Page 7 of 28 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

Analytical Chemistry

118

and passed through the hot flue gas at a height of about 5 mm above the burner outlet.

119

The indium transitions, 52P1/2⟶62S1/2 and 52P3/2⟶62S1/2, were probed and the

120

fluorescence at 451 nm was captured by an ICCD camera (PI-MAX 3, Princeton

121

Instruments) through a band-pass filter (450 nm ± 10 nm). Using the fluorescence signal,

122

a temperature profile above the burner outlet was obtained.

123 124

Concentration of potassium atoms. The concentration of K atoms was measured using

125

the TDLAS system based on the Beer-Lambert law,7, 16, 33, 34

126

𝐼(ν) = 𝐼𝑛 + 𝐼𝑒 + 𝐼0(ν)exp [ ― α(ν)]

(1)

127

where ν is the frequency of the light, 𝐼0(ν) and 𝐼(ν) are the intensity of the light before

128

and after passing through the absorbers, respectively, and 𝐼𝑛 and 𝐼𝑒 are the frequency-

129

independent backgrounds from the detector dark current and broadband emission,

130

respectively. It was found that the broadband emission 𝐼𝑒 from the investigated methane

131

flames was negligible especially as our detector was placed about 1.5 m away from the

132

flame and an aperture was used in the beam path. 𝐼𝑛 was assumed to be constant and

133

obtained under the condition without laser beam in the flame. The frequency-dependent

134

absorbance, α(ν), is proportional to the optical path length L and the number density N

135

of K atoms,7, 35

136 137 138

α(ν) = σ(ν)𝐿𝑁

(2)

where σ(ν) is the absorption cross section at frequency ν, which can be expressed as, σ(ν) =

ℎν0𝐵12𝜒(ν, 𝑁,𝑇) 𝑐

7 ACS Paragon Plus Environment

(3)

Analytical Chemistry 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 8 of 28

139

where h is the Planck’s constant, ν0 is the central frequency, 𝐵12 is the Einstein

140

absorption coefficient, c is the speed of light, 𝜒(ν, 𝑁,𝑇) is the area-normalized line shape

141

function and T is the temperature. The line shape function can be described by a Voigt

142

profile in the atmospheric environment. The Einstein absorption coefficient 𝐵12 can be

143

related to the Einstein spontaneous emission coefficient A21 and expressed as,

144 145

𝐵21 = (𝑐3/8πℎ𝜈3)𝐴21

(4)

𝐵12 = (𝑔2/𝑔1)𝐵21

(5)

146

where 𝑔1 and 𝑔2 are the degeneracy of state 1 and 2, respectively.

147

In the calculation of the number density of K atoms, firstly, the measured absorbance αm

148

(ν) was obtained based on equation (1) using measured intensities. Then, according to

149

equation (2) and (3), a Voigt profile was used to fit the measured absorbance αm(ν). With

150

the fitted absorbance αf(ν), the number density N can be derived by integrating over

151

equation (2), and expressed as: 𝑐

152

𝑁 = ℎ𝜈0𝐵12𝐿∫αf(𝜈)𝑑ν

153

In the present study, the concentration of K atoms was measured using two TDLAS

154

systems with tunable continuous-wave diode lasers centered at 769.9 nm and 404.4 nm.

155

The schematic diagram of the TDLAS system is presented in Fig. 1 (b).

156

In previous studies of potassium,7, 16, 33, 34 TDLAS with 769.9 nm lasers has been adopted

157

in favor of its high sensitivity. In our 769.9 nm TDLAS system, the beam with a power

158

of about 3 mW and size of about 1 mm2 was provided by an external cavity laser (Toptica,

159

DL100). Similar to the TLAF system, an analog control unit was used to control the

8 ACS Paragon Plus Environment

(6)

Page 9 of 28 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

Analytical Chemistry

160

temperature and the current of the laser. Additionally, a scan control module (SC 110)

161

was used for scanning of the wavelength with a range over 35 GHz at a repetition rate of

162

100 Hz. The scanning range was monitored by a high-finesse confocal Fabry-Perot etalon

163

(Topoca, FPI 100), and the intensity of the light was monitored by two photodiodes of

164

the same model (PDA100A, Thorlabs). The laser beam passed through the hot gas at a

165

height of ~5 mm above the burner outlet with a path length of 8.5 cm.

166

167 168 169 170 171 172

Figure 2. Absorption curves and Voigt function fittings of the 42S1/2⟶42P1/2 potassium atom transition at 769.9 nm for concentrations of 0.05 ppm (a) and 0.52 ppm (b). The measured curve in (b) is truncated due to strong absorption under optically thick conditions. The insets show the scanning signal as function of laser frequency with and without potassium absorption.

173

Typical scanning signals with and without K atoms are shown in the insets of Fig. 2. They

174

were used to derive the measured absorbance αm(ν) presented in the main graphs. For

175

the measurement with low concentration of K atoms (Fig. 2 (a)), the fitted absorbance αf

176

(ν) was obtained through the fitting of αm(ν) using a Voigt profile, which has a full

177

width at half maximum (FWHM) of 5.34 GHz, similar to values obtained by Qu et al.16

178

and Schlosser et al.33 The number density N was calculated based on equation (6), and

9 ACS Paragon Plus Environment

Analytical Chemistry 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

179

the concentration was determined to be 0.05 ppm.

180 181

Figure 3. Proposed correction curve of the detector response used for TDLAS data

182

analysis.

183

However, as the concentration increases, an absorption spectrum with regions of

184

complete signal extinction appears indicating optically thick conditions as shown in the

185

inset of Fig. 2 (b). Correspondingly, the absorbance peak was truncated and reached a

186

maximum plateau level, which cannot be fitted by a Voigt profile. However, Qu et al.16

187

reported that the Beer-Lambert law could still be applied to this condition, and the

188

concentration of K atoms could be determined through a reconstruction of the absorption

189

profile by fitting to the wings of the line profile. In the work of Qu et al.16, a corrected

190

absorbance was used to fit the measured absorbance αm(ν) under optically thick

191

conditions. In their correction, a simulated light intensity after absorption, 𝐼f(ν), was

192

derived based on equation (1), using a simulated absorbance αf(ν) with a Voigt profile

193

from equation (2) and the light intensity without absorption 𝐼0(ν). Then, under optically

194

thick conditions, considering the response limit of the detector, values of 𝐼f(ν) smaller

10 ACS Paragon Plus Environment

Page 10 of 28

Page 11 of 28 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

Analytical Chemistry

195

than 𝐼𝑛 were set to 𝐼𝑛, through which 𝐼f(ν) was corrected to 𝐼fc(ν), which was used to

196

mimic the actual signal obtained from the detector. This correction process can be

197

illustrated by the detector response curve shown in Fig. 3 with a red dotted line. Then, the

198

absorbance under optically thick condition, αfc(ν), was derived from 𝐼fc(ν) and 𝐼0(ν),

199

which was fitted to the measured absorbance αm(ν) through adjusting the absorbance αf

200

(ν). In the present study, a similar correction process was used. However, a new

201

correction equation was adopted for the conversion of 𝐼f(ν) to 𝐼fc(ν), to consider the

202

nonlinear response of the detector to the very weak signal. The proposed non-linear

203

correction trend is shown in Fig. 3, and the equation is expressed as,

204

𝐼fc(ν) = ln [exp [𝛽𝐼f(ν)] + exp [𝛽𝐼𝑛] ― 1]/𝛽

205

where β is a constant to fix the curvature of the correction curve. As shown in Fig. 2 (b),

206

the αfc(ν) was well fitted to the measured absorbance αm(ν) with a suitable αf(ν)

207

which was used to calculate the number density of K atoms using equation (6). The

208

concentration of the K atoms in Fig.2 (b) was determined to 0.52 ppm.

209

It is clear that by using the method described by Qu et al.16 for optical thick conditions,

210

the measurement dynamic range can be increased dramatically, i.e. from 40 ppt to 40 ppm

211

for a 1 cm path length for the 769.9 nm potassium 42S1/2⟶42P1/2 line.16 However, in the

212

present study, the path length is 8.5 cm, and the concentration of K atoms can be over 20

213

ppm, which thus made it impossible to achieve a proper measurement using the 769.9 nm

214

line, as most part of the measured absorption peak would be truncated. Therefore, another

215

TDLAS system probing the 404.4 nm 42S1/2⟶52P3/2 transition was introduced. The 404.4

11 ACS Paragon Plus Environment

(7)

Analytical Chemistry 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

216

nm absorption has an Einstein spontaneous emission coefficient (1.16·106 s-1)36 much

217

smaller than the 769.9 nm line (3.75·107 s-1)36, which allows measurements of much

218

higher concentrations (cf. equation (2) and (3)).

219

For the 404.4 nm TDLAS system, a beam with a power of about 5.5 mW and size of

220

about 1 mm2 was provided by another external cavity laser (Toptica, DL100pro)

221

controlled by an analog control unit. The wavelength was scanned with a range of about

222

21 GHz with a repetition rate of 100 Hz.

223

224 225 226 227 228 229

Figure 4. Absorption curves and the Voigt function fitting of the 42S1/2⟶52P3/2 potassium atom transition at 404.4 nm for concentrations of 0.25 ppm (a) and 19.6 ppm (b). The measured curve in (b) is truncated due to strong absorption under optically thick conditions. The insets show the scanning signal as the function of laser frequency with and without potassium absorption.

230

The scanning signals of the transmitted light with and without absorption by K atoms are

231

presented in the insets of Fig. 4, which were used to derive the measured absorbance

232

profiles. It can be seen that the scanning range of 21 GHz cannot cover the whole

233

absorption peak. For a low concentration situation, similar to the calculation process for

234

the 769.9 nm case, the measured absorbance 𝛼m(ν) was fitted by the absorbance αf(ν)

12 ACS Paragon Plus Environment

Page 12 of 28

Page 13 of 28 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

Analytical Chemistry

235

using a Voigt profile. The absorbance profile shown in Fig. 4 (a) has a full width at half

236

maximum (FWHM) of 13.3 GHz. In order to check the whole absorption profile,

237

measurements with different fixed laser wavelengths were conducted, and the absorption

238

values showed good overlap with the fitting curve from the data measured using laser

239

scanning. Hence, the fitted absorbance αf(ν) were used in the concentration evaluation

240

with equation (6). The concentration of K atoms in Fig. 4 (a) was determined to be 0.25

241

ppm using this approach.

242

As the concentration of K atoms increased, the optical thick condition occurred as shown

243

in the inset of Fig. 4 (b). The corresponding measured absorbance αm(ν) with truncation

244

of the absorption peak is shown in Fig. 4 (b), and was well fitted by the absorbance after

245

the correction on αf(ν) considering the optically thick absorption feature. Using the

246

obtained αf(ν), the concentration of the K atoms in Fig. 4 (b) was determined to be 19.6

247

ppm. The introduction of the 404.4 nm TDLAS system thus further expanded our

248

measurement dynamic range of K atoms, which fulfilled the requirement in this study.

249 250

Concentration of KOH and KCl. The concentrations of KOH and KCl were measured

251

using UV absorption spectroscopy with the optical setup presented in Fig. 1 (b). A

252

deuterium lamp was used as the broadband UV light source and the light was collimated

253

by a parabolic mirror to form a beam of 10 mm in diameter through an aperture, and then

254

guided to the hot flue gas region. With five UV-enhanced aluminum mirrors, the

255

absorption path length above the burner was increased to 522 mm. Finally, the UV light

13 ACS Paragon Plus Environment

Analytical Chemistry 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 14 of 28

256

is collected to a spectrometer (USB 2000+, Ocean Optics). To determine the path length

257

and check the homogeneity over the line of sight, the distribution of KOH in the flue gas

258

was visualized by laser-induced photofragmentation fluorescence using the fifth

259

harmonic of a pulsed Nd:YAG laser (Brilliant B, Quantel) at wavelength 213 nm.11, 12 A

260

homogeneous distribution of KOH in the interrogated region was confirmed in this

261

visualization measurement. As shown in Fig. 5 (a), spectrally resolved absorbance A(λ)

262

of different hot flue gases is obtained through the analysis of the collected spectrum using

263

the Beer-Lambert law,23

264

𝐴(λ) = ― ln

( ) = 𝑁 σ (λ)𝐿 𝐼s(λ)

A A

𝐼0(λ)

(8)

265

where 𝐼0(λ) and 𝐼s(λ) are the UV light intensity before and after the seeding of

266

potassium to the hot gas, respectively, 𝑁A is the number density of the alkali species in

267

the hot gas, σA(λ) is the absorption cross-section of alkali species, and L is the

268

absorption path length. In obtaining the absorption spectral curve shown in Fig. 5 (a) (also

269

in Fig. 6), the USB spectrometer was set to have a gate width of 1 second, and 10

270

measurements were averaged for each case, hence, it took 10 seconds to have the UV

271

spectra of KOH and KCl. In Fig. 5 (a), the spectrum mainly consists of broadband

272

absorption by KOH in the hot gas,23 but absorption lines of potassium atoms were also

273

observed at wavelengths 404 nm, 345 nm and 322 nm. Some dips on the absorbance curve

274

near 283 nm and 310 nm indicate the reduction of OH radicals in the hot gas with seeding

275

of potassium species.23 UV absorption cross-sections of potassium species, required to

276

retrieve the concentration of KOH and KCl, are evaluated from the absorbance data and

14 ACS Paragon Plus Environment

Page 15 of 28 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

Analytical Chemistry

277

presented in Fig. 5 (b). The detail of the evaluation process will be described in the

278

following section.

279

280 281 282

Figure 5. The absorbance of KOH in the fuel lean and rich flames (a) and the derived UV absorption cross-sections of KOH at different temperatures (b).

283 284

Results and discussion

285

UV absorption cross-sections of KOH and KCl. In the present study, the UV absorption

286

cross-section of KOH in combustion atmosphere was determined. In this evaluation, a

287

constant amount of K2CO3 was introduced into one fuel lean and one fuel rich flame. The

288

potassium chemistry results in the generation of KOH and K atoms as the dominant

289

potassium species in the hot flue gas.24 In the fuel lean case, KOH is the dominant species,

290

while in the fuel rich case, nearly half of the potassium exists as K atoms. When changing

291

from fuel lean to fuel rich flame, the increase of the concentration of K atoms will be

292

equal to the reduction of KOH. Based on the measured K-atom concentrations for the

293

lean and rich cases, the absorption cross-sections of KOH were evaluated using the

294

measured absorbance difference of KOH from the UV absorption spectroscopy. Derived

15 ACS Paragon Plus Environment

Analytical Chemistry 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

295 296

Page 16 of 28

from equation (8), the evaluation process can be described with the following equation, NK_rich–NK_lean=NKOH_lean–NKOH_rich = (AKOH_lean(λ)-AKOH_rich(λ))/(σKOH(λ)×L)

(9)

297

where N indicates the number density of K atoms and KOH molecules in different flame

298

cases. A(λ) indicates the absorbance of KOH in different flames. σKOH(λ) represents the

299

absorption cross-section of KOH and L indicates the optical path length. The subscript

300

‘K’ and ‘KOH’ indicate the potassium species while ‘rich’ and ‘lean’ indicates the

301

corresponding flame case.

302

Flame 1 (ϕ=0.67 and T=1768 K) and Flame 5 (ϕ=1.12 and T=1843 K) were used to

303

evaluate the absorption cross-section of KOH around 1800 K. The absorbance curves in

304

these two flames obtained with seeding of the same amount of K2CO3 are presented in

305

Fig. 5 (a). A detailed description of the absorption spectrum of KOH is provided by Weng

306

et al.23 Absorption of the OH radical (around 283 nm and 310 nm) and K atoms (around

307

404 nm, 345 nm and 322 nm) can be observed, but are much narrower than that of KOH.

308

The absorbance of KOH in the lean case nearly doubles the one in the rich case (cf. Fig.

309

5 (a)) and the difference was mainly caused by the difference in KOH concentration

310

between these two cases. The influence of the temperature difference between these two

311

cases (1768 K vs 1843 K) is negligible as reported by Weng et al.23 The measured number

312

density of K atoms in Flame 1 and Flame 5 was 1.7·1018 m-3 and 4.87·1019m-3,

313

respectively. Thus, the absorption cross-section of KOH around temperature 1800 K, as

314

shown in Fig. 5 (b), was obtained with equation (9) and equal to 1.1·10-21 m2/molecule at

315

327.3 nm and 1.29·10-21 m2/molecule at 246.2 nm. It is worth to note that by combining

16 ACS Paragon Plus Environment

Page 17 of 28 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

Analytical Chemistry

316

equations (8) and (9), one can find that the path length L can be canceled in obtaining the

317

cross-section values. Additionally, the absorption cross-section of KOH around

318

temperature 1400 K was evaluated from measurements made in Flame 8 (ϕ=0.67 and

319

T=1390 K) and Flame 10 (ϕ=1.31 and T=1471 K). The obtained absorption cross-section

320

is presented in Fig. 5 (b) and has a value of 1.05·10-21 m2/molecule at 327.3 nm and

321

1.26·10-21 m2/molecule at 246.2 nm. It is about twice of the one used by Sorvajärvi et al.17

322

near 320 nm, where it was approximated to the cross-section of NaOH reported by Self

323

et al.37 To the best of our knowledge, Fig. 5(b) presents the first experimental data on the

324

UV absorption cross-section of KOH. Comparing the two absorption cross-section

325

profiles of KOH at temperatures around 1400 K and 1800 K, it can be seen that the

326

influence of temperature is small. The average value of the cross-sections at temperatures

327

1400 K and 1800 K was therefore applied in the calculation of the concentration of KOH

328

in the following evaluation.

329

330 331 332 333

Figure 6. The absorbance of KOH and KCl in flame 9 with a global equivalence ratio of 0.63 and a temperature of 1260 K (a) and the derived UV absorption cross-sections of KCl at different temperatures (b). 17 ACS Paragon Plus Environment

Analytical Chemistry 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 18 of 28

334

Furthermore, the UV absorption cross-section of KCl in the high-temperature gas was

335

obtained. In order to generate KCl in the hot flue gas, CHCl3 was seeded into the premixed

336

flame as the source of chlorine. In the fuel lean cases, with the seeding of K and Cl, the

337

dominant potassium species in the hot flue gas were KOH and KCl, while less than 2.5%

338

of the potassium formed K atoms according to the measured concentration of K atoms

339

and the total potassium concentration obtained in the case without the seeding of Cl.

340

Especially in the low temperature case, Flame 9, there were almost no K atoms present.

341

Hence, the potassium balance is mainly between KOH and KCl. In the present

342

investigations, about 100 ppm of Cl was introduced into the flame while the concentration

343

of K element was kept around 20 ppm. It was observed that there was almost no KOH

344

formed with the seeding of 100 ppm Cl according to the structure of the absorption

345

spectrum as shown in Fig. 6 (a). Similar result has been presented by Schofield38 on

346

sodium chemistry and showed that NaCl became the dominant species as chlorine was

347

available in excess compared with sodium. The absorption was solely from KCl, as it

348

shows the same profile as the ones presented in previous studies on the UV absorption of

349

KCl.18-20, 23 Consequently, the UV absorption cross-section of KCl was obtained based on

350

the Beer-Lambert law with the measured absorbance of KCl in the flame (Fig. 6 (a)) and

351

its corresponding concentration from the measured concentration of KOH under the

352

condition without Cl seeding.

353 354

NKOH_lean = NKCl_lean = AKCl_lean(λ)/(σKCl(λ) × L)

(10)

Fig. 6 (b) shows the absorption cross-section of KCl at temperatures around 1800 K and

18 ACS Paragon Plus Environment

Page 19 of 28 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

Analytical Chemistry

355

1300 K with values of 2.3·10-21 m2/molecule and 2.26·10-21 m2/molecule, respectively, at

356

wavelength 246.2 nm. The variation of the cross-section with temperature is shown to be

357

small as concluded by Leffler et al.20 In the following, the value of 2.28·10-21 m2/molecule

358

at wavelength 246.2 nm was used in the KCl concentration evaluation. The present value

359

is about 20% lower than the value from Leffler et al.20 measured at a temperature of about

360

1000 K, and about 28% higher than the values from Davidovits et al.19 and Forsberg et

361

al.18 at a temperature of around 1100 K. In the measurements by Leffler et al.20 and

362

Davidovits et al.19, sealed quartz cells generating controlled amounts of KCl vapor were

363

used, in which the evaluation of the KCl cross-sections strongly relied on the

364

determination of the KCl vapor pressure in the cell. The accuracy of the calculated KCl

365

vapor pressure was directly affected by the accuracy of the salt reservoir temperature.20

366

A possible uncertainty in measured temperature of 15 K could cause an uncertainty in

367

vapor pressure of more than 20%.20 Moreover, since a nearly saturated KCl vapor was

368

generated in the cell, there might be some additional uncertainty induced by the formation

369

of K2Cl2 as discussed by Leffler et al.20 Compared with previous investigations, there are

370

thus some advantages in the present study. Firstly, the measurement was made directly

371

on the KCl vapor in the combustion environment, in which the formation of the K2Cl2

372

dimer was negligible at temperatures over 1300 K. Also, the evaluation of the absorption

373

cross-section was based on the measurement of the concentration of K atoms, which had

374

a relatively low uncertainty. According to the estimation of the error in the absorbance

375

curve fitting process, the uncertainty was about ±2%. In addition, about ±3% uncertainty

19 ACS Paragon Plus Environment

Analytical Chemistry 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

376

can be introduced by the uncertainty of around ±50 K in the TLAF temperature

377

measurement of the hot gas. Additionally, uncertainty might originate from the

378

inhomogeneity of the potassium distribution in the hot flue gas. The inhomogeneity was

379

estimated through the distribution of the photofragmenation fluorescence of KOH/KCl in

380

the hot flue gas. The fluctuation of the concentration of potassium in the measurement

381

region was estimated to be below ±10%. Moreover, in the rich condition, the secondary

382

reactions occurred with the ambient air to form a thin layer of flame front which might

383

change the balance between KOH and K atoms. This influence should be small

384

considering the fact that the reaction zone was very small with a thickness less than 1%

385

of the whole homogenous reaction zone. However, combining Equations (8) and (9), it

386

can be found that the path length L can be canceled and there was almost no influence on

387

the determination of the absorption cross section. Hence, the total error of the absorption

388

cross section was estimated to be about ±5% in the present study, without considering

389

uncertainties in path length, inhomogeneity of temperature and possible secondary

390

reactions.

391 392

Concentration measurements of K atoms, KOH and KCl in flames. Quantitative

393

measurements of KOH, KCl and K atoms were carried out to study the potassium-

394

chlorine chemistry. To mimic different oxidation and reduction environments, hot flue

395

gases were prepared from flames (Flame 1 – Flame 7) with global equivalence ratios that

396

varied from 0.67 to 1.32. A constant amount of KCl was seeded into the flames via KCl

20 ACS Paragon Plus Environment

Page 20 of 28

Page 21 of 28 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

Analytical Chemistry

397

water solution. The concentrations of KOH, KCl and K atoms measured through

398

broadband UV absorption spectroscopy and TDLAS are shown in Fig. 7 with a total

399

amount of potassium at around 22 ppm.

400

401 402 403 404 405 406

Figure 7. The concentration of K atoms, KOH, KCl and total K in the hot flue gas versus global equivalence ratio with seeding of KCl. The concentration of K atoms was measured by TDLAS both at 404.4 nm and 769.9 nm in the fuel lean cases. Note: the concentration in different cases was corrected due to the difference in total flow rate shown in Table 1.

407

It is clear that there is a significant change of the concentration for all the potassium

408

species as the equivalence ratio varies from lean to rich cases. For the lean cases, the

409

concentrations of different potassium species are almost kept constant. KCl had a

410

concentration of around 12 ppm and KOH had a concentration around 10 ppm. For the

411

lean cases the concentration of K atoms was measured both by the 404.4 nm and the 769.9

412

nm TDLAS system. The results obtained from these two methods agreed well, both giving

413

a value of around 0.3 ppm, which was much smaller than the concentrations of KCl and

414

KOH. It can be seen that there was only around 1% of potassium in the form of K atoms

415

in the oxidation environment at temperature 1800 K, while KCl and KOH constituted

21 ACS Paragon Plus Environment

Analytical Chemistry 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

416

about 55% and 43%, respectively, of the total potassium. In the reduction environments,

417

the concentration of K atoms increased to 8 ppm, and continually increased with

418

equivalence ratio. The fraction of K atoms became over 35% in reduction environments,

419

in agreement with results presented by Leffler et al.15 and indicating that K atoms should

420

be considered as one of the dominant species in the reduction environment together with

421

KCl and KOH, which had a fraction of 37% and 28%, respectively. The formation of

422

large amounts of K atoms might be caused by the existence of a large amount of H radicals

423

in the reduction environments, which enhanced reactions between KCl/KOH and H

424

radicals.26, 39 The ratio between the concentration of KOH and the sum of KOH and KCl

425

values was almost constant, even in the fuel rich flames, which indicates that chemical

426

equilibrium between KOH and KCl can be easily established.

427 428

Conclusions

429

In the present work, spectrally resolved UV absorption cross-sections of KOH and KCl

430

in the wavelength range 200 - 480 nm at temperatures 1300 K and 1800 K were, for the

431

first time, obtained directly through a reliable experimental method. The well-recognized

432

spectral features make it possible to perform simultaneous KOH and KCl measurement

433

in combustion environments. Meanwhile, the concentration of K atoms was measured

434

using TDLAS combining the absorption lines at 769.9 nm and 404.4 nm to cover a

435

dynamic range of concentrations from several ppb up to 20 ppm for a 10 cm path length.

436

Hence, all the major potassium species, i.e. KOH, KCl and K atoms, present in most

22 ACS Paragon Plus Environment

Page 22 of 28

Page 23 of 28 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

Analytical Chemistry

437

biomass combustion/gasification atmospheres, can be measured quantitatively. Using our

438

experimental setup, a constant amount of KCl was seeded into flames with varying

439

equivalence ratios and the concentrations of KOH, KCl and K were measured. The

440

chemical balance between different potassium species in different environments was

441

accurately determined. These quantitative techniques can be used for accurate potassium

442

chemistry and for potassium species monitoring in biomass combustors and gasifiers.

443 444

Acknowledgements

445

This work was financed by the Swedish Energy Agency (GRECOP, 38913-2), the Knut

446

& Alice Wallenberg foundation (KAW 2015.0294, ALADIN), the Swedish Research

447

Council (VR) and the European Research Council (ERC, 669466/TUCLA).

448 449

References

450

1.

451

during Pyrolysis and Combustion of High-Chlorine Biomass. Energy Fuels 2011, 25, 4961-4971.

452

2.

453

during the Fast-Pyrolysis Process of Rice Straw. Energy Fuels 2015, 29, 6404-6411.

454

3.

455

Transformation Pathways of Various K Species during Thermal Conversion of Agricultural Straw.

456

Part 1: Devolatilization Stage. Energy Fuels 2018.

457

4.

Johansen, J. M.; Jakobsen, J. G.; Frandsen, F. J.; Glarborg, P. Release of K, Cl, and S

Zhao, H.; Song, Q.; Wu, X.; Yao, Q. Study on the Transformation of Inherent Potassium

Huang, Y.; Liu, H.; Yuan, H.; Zhuang, X.; Yuan, S.; Yin, X.; Wu, C. Release and

Nielsen, H. P.; Frandsen, F. J.; Dam-Johansen, K.; Baxter, L. L. The implications of chlorine-

23 ACS Paragon Plus Environment

Analytical Chemistry 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

458

associated corrosion on the operation of biomass-fired boilers. Prog. Energy Combust. Sci. 2000,

459

26, 283-298.

460

5.

461

phase potassium during the combustion of single particles of biomass. Fuel 2016, 182, 110-117.

462

6.

463

and numerical studies of potassium release during biomass gasification. Proc. Combust. Inst.

464

2015, 35, 2389-2396.

465

7.

466

Measurement of Atomic Potassium in Plumes over Burning Solid Fuels Using Infrared-Diode

467

Laser Spectroscopy. Energy Fuels 2017, 31, 2831-2837.

468

8.

469

Measurement and kinetics of elemental and atomic potassium release from a burning biomass

470

pellet. Proc. Combust. Inst. 2018.

471

9.

472

Released from Burning Particles of Brown Coal and Pine Wood in a Laminar Premixed Methane

473

Flame Using Quantitative Laser-Induced Breakdown Spectroscopy. Appl. Spectrosc. 2011, 65,

474

684-691.

475

10. He, Y.; Zhu, J.; Li, B.; Wang, Z.; Li, Z.; Aldén, M.; Cen, K. In-situ Measurement of Sodium

476

and Potassium Release during Oxy-Fuel Combustion of Lignite using Laser-Induced Breakdown

477

Spectroscopy: Effects of O2 and CO2 Concentration. Energy Fuels 2013, 27, 1123-1130.

478

11. Oldenborg, R. C.; Baughcum, S. L. Photofragment fluorescence as an analytical technique:

Mason, P. E.; Darvell, L. I.; Jones, J. M.; Williams, A. Observations on the release of gas-

Fatehi, H.; He, Y.; Wang, Z.; Li, Z. S.; Bai, X. S.; Aldén, M.; Cen, K. F. LIBS measurements

Weng, W.; Gao, Q.; Wang, Z.; Whiddon, R.; He, Y.; Li, Z.; Aldén, M.; Cen, K. Quantitative

Liu, Y.; Wang, Z.; Xia, J.; Vervisch, L.; Wan, K.; He, Y.; Whiddon, R.; Bahai, H.; Cen, K.

Hsu, L.; Alwahabi, Z. T.; Nathan, G. J.; Li, Y.; Li, Z. S.; Aldén, M. Sodium and Potassium

24 ACS Paragon Plus Environment

Page 24 of 28

Page 25 of 28 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

Analytical Chemistry

479

application to gas-phase alkali chlorides. Anal. Chem. 1986, 58, 1430-1436.

480

12. Leffler, T.; Brackmann, C.; Aldén, M.; Li, Z. Laser-Induced Photofragmentation Fluorescence

481

Imaging of Alkali Compounds in Flames. Appl. Spectrosc. 2017, 71, 1289-1299.

482

13. Leffler, T.; Brackmann, C.; Berg, M.; Li, Z.; Aldén, M. On-Line alkali measurement for fuel

483

quality control in biomass-Operated boilers. IFRF Combustion Journal 2016, 2016, 1-19.

484

14. Leffler, T.; Brackmann, C.; Berg, M.; Aldén, M.; Li, Z. Online Alkali Measurement during Oxy-

485

fuel Combustion. Energy Procedia 2017, 120, 365-372.

486

15. Leffler, T.; Brackmann, C.; Weng, W.; Gao, Q.; Aldén, M.; Li, Z. Experimental investigations

487

of potassium chemistry in premixed flames. Fuel 2017, 203, 802-810.

488

16. Qu, Z.; Steinvall, E.; Ghorbani, R.; Schmidt, F. M. Tunable Diode Laser Atomic Absorption

489

Spectroscopy for Detection of Potassium under Optically Thick Conditions. Anal. Chem. 2016,

490

88, 3754-3760.

491

17. Sorvajärvi, T.; DeMartini, N.; Rossi, J.; Toivonen, J. In Situ Measurement Technique for

492

Simultaneous Detection of K, KCl, and KOH Vapors Released during Combustion of Solid

493

Biomass Fuel in a Single Particle Reactor. Appl. Spectrosc. 2014, 68, 179-184.

494

18. Forsberg, C.; Broström, M.; Backman, R.; Edvardsson, E.; Badiei, S.; Berg, M.; Kassman,

495

H. Principle, calibration, and application of the in situ alkali chloride monitor. Rev. Sci. Instrum.

496

2009, 80, 023104.

497

19. Davidovits, P.; Brodhead, D. C. Ultraviolet Absorption Cross Sections for the Alkali Halide

498

Vapors. J. Chem. Phys. 1967, 46, 2968-2973.

499

20. Leffler, T.; Brackmann, C.; Berg, M.; Aldén, M.; Li, Z. S. Development of an alkali chloride

25 ACS Paragon Plus Environment

Analytical Chemistry 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

500

vapour-generating apparatus for calibration of ultraviolet absorption measurements. Rev. Sci.

501

Instrum. 2017, 88, 023112.

502

21. Sherwood, R. F.; Yoshihiro, M. Upper stratospheric photolysis of NaOH. Geophys. Res. Lett.

503

1982, 9, 473-475.

504

22. Daidoji, H. Molecular absorption spectra of some sodium salts in flames. Bunseki Kagaku

505

1979, 28, 77-82.

506

23. Weng, W.; Leffler, T.; Brackmann, C.; Aldén, M.; Li, Z. Spectrally Resolved UV Absorption

507

Cross-Sections of Alkali Hydroxides and Chlorides Measured in Hot Flue Gases. Appl. Spectrosc.

508

2018, 0003702818763819.

509

24. Babushok, V. I.; Linteris, G. T.; Hoorelbeke, P.; Roosendans, D.; van Wingerden, K. Flame

510

Inhibition by Potassium-Containing Compounds. Combust. Sci. Technol. 2017, 189, 2039-2055.

511

25. Carabetta, R.; Kaskan, W. E. Oxidation of sodium, potassium, and cesium in flames. J. Phys.

512

Chem. 1968, 72, 2483-2489.

513

26. Slack, M.; Cox, J. W.; Grillo, A.; Ryan, R.; Smith, O. Potassium kinetics in heavily seeded

514

atmospheric pressure laminar methane flames. Combust. Flame 1989, 77, 311-320.

515

27. Weng, W.; Borggren, J.; Li, B.; Aldén, M.; Li, Z. A novel multi-jet burner for hot flue gases of

516

wide range of temperatures and compositions for optical diagnostics of solid fuels

517

gasification/combustion. Rev. Sci. Instrum. 2017, 88, 045104.

518

28. Stull, D. R. Vapor Pressure of Pure Substances. Organic and Inorganic Compounds. Ind.

519

Eng. Chem. 1947, 39, 517-540.

520

29. Xieqi, M.; Cicek, B.; Senkan, S. M. Chemical structures of fuel-rich and fuel-lean flames of

26 ACS Paragon Plus Environment

Page 26 of 28

Page 27 of 28 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

Analytical Chemistry

521

CCl4/CH4 mixtures. Combust. Flame 1993, 94, 131-145.

522

30. Li, Z. S.; Sun, Z. W.; Li, B.; Aldén, M.; Försth, M. Spatially resolved trace detection of HCl in

523

flames with mid-infrared polarization spectroscopy. Opt. Lett. 2008, 33, 1836-1838.

524

31. Borggren, J.; Weng, W.; Hosseinnia, A.; Bengtsson, P.-E.; Aldén, M.; Li, Z. Diode laser-

525

based thermometry using two-line atomic fluorescence of indium and gallium. Appl. Phys. B 2017,

526

123, 278.

527

32. Borggren, J.; Burns, I. S.; Sahlberg, A.; Aldén, M.; Li, Z. Temperature imaging in low-

528

pressure flames using diode laser two-line atomic fluorescence employing a novel indium seeding

529

technique. Appl. Phys. B 2016, 122, 58.

530

33. Schlosser, E.; Fernholz, T.; Teichert, H.; Ebert, V. In situ detection of potassium atoms in

531

high-temperature coal-combustion systems using near-infrared-diode lasers. Spectrochim. Acta,

532

Part A 2002, 58, 2347-2359.

533

34. Gao, Q.; Weng, W.; Li, B.; Li, Z. Quantitative and Spatially Resolved Measurement of Atomic

534

Potassium in Combustion Using Diode Laser. Chin. Phys. Lett. 2018, 35, 024202.

535

35. van Eyk, P. J.; Ashman, P. J.; Alwahabi, Z. T.; Nathan, G. J. Quantitative measurement of

536

atomic sodium in the plume of a single burning coal particle. Combust. Flame 2008, 155, 529-

537

537.

538

36. Sansonetti, J. E. Wavelengths, Transition Probabilities, and Energy Levels for the Spectra of

539

Potassium (KI through KXIX). J. Phys. Chem. Ref. Data 2008, 37, 7-96.

540

37. Self, D. E.; M. C. Plane, J. Absolute photolysis cross-sections for NaHCO3 , NaOH, NaO,

541

NaO2 and NaO3   : implications for sodium chemistry in the upper mesosphere. Phys. Chem.

27 ACS Paragon Plus Environment

Analytical Chemistry 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

542

Chem. Phys. 2002, 4, 16-23.

543

38. Schofield, K. The chemical nature of combustion deposition and corrosion: The case of alkali

544

chlorides. Combust. Flame 2012, 159, 1987-1996.

545

39. Li, B.; Sun, Z.; Li, Z.; Aldén, M.; Jakobsen, J. G.; Hansen, S.; Glarborg, P. Post-flame gas-

546

phase sulfation of potassium chloride. Combust. Flame 2013, 160, 959-969.

547

28 ACS Paragon Plus Environment

Page 28 of 28