Quantum Yield of Nitrite from the Photolysis of Aqueous Nitrate above

Mar 24, 2017 - ... Rolf Müller , Reinhold Spang , Ines Tritscher , Tobias Wegner , Martyn P. Chipperfield , Wuhu Feng , Douglas E. Kinnison , Sasha M...
0 downloads 0 Views 729KB Size
Subscriber access provided by MacEwan University Libraries

Article

Quantum Yield of Nitrite from the Photolysis of Aqueous Nitrate above 300 nm Katherine Beem Benedict, Alexander S McFall, and Cort Anastasio Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b06370 • Publication Date (Web): 24 Mar 2017 Downloaded from http://pubs.acs.org on April 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.

Environmental Science & Technology 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 35

Environmental Science & Technology

1 2 3 4 5 6 7

8

Quantum Yield of Nitrite from the Photolysis of Aqueous Nitrate above 300 nm

9 10 Katherine B. Benedict1, Alexander S. McFall and Cort Anastasio*

11 12 13

Department of Land, Air, and Water Resources

14

University of California Davis

15

Davis, CA 95616

16 17 18

*Corresponding Author, Department of Land, Air, and Water Resources, University of California Davis; Tel: 530-754-6095; Email: [email protected]; Fax: 530-752-1552

19

1

now at Department of Atmospheric Science, Colorado State University, Fort Collins, CO 80523

20 21

Submitted in Revised Form to Environmental Science and Technology on March 22, 2017.

22

1

ACS Paragon Plus Environment

Environmental Science & Technology

Page 2 of 35

23

ABSTRACT: Photolysis of nitrate (NO3–) produces reactive nitrogen and oxygen species via

24

three different channels, forming: (1) nitrogen dioxide (NO2) and hydroxyl radical (•OH), (2)

25

nitrite (NO2-) and oxygen atom (O(3P)), and (3) peroxynitrite (ONOO–). These photoproducts are

26

important oxidants and reactants in surface waters, atmospheric drops, and snowpacks. While the

27

efficiency of the first channel, to form NO2, is well documented, a large range of values have

28

been reported for the second channel, nitrite, above 300 nm. In part, this disagreement reflects

29

secondary chemistry that can produce or destroy nitrite. In this study, we examine factors that

30

influence nitrite production and find that pH, nitrate concentration, and the presence of an •OH

31

scavenger can be important. We measure an average nitrite quantum yield (Φ(NO2–)) of (1.1 ±

32

0.2)% (313 nm, 50 µM nitrate, pH ≥ 5), which is at the upper end of past measurements and an

33

order of magnitude above the smallest – and most commonly cited – value reported for this

34

channel. Nitrite production is often considered a very minor channel in nitrate photolysis, but our

35

results indicate it is as important as the NO2 channel. In contrast, at 313 nm we observe no

36

formation of peroxynitrite, corresponding to Φ(ONOO–) < 0.26%.

37 38

2

ACS Paragon Plus Environment

Page 3 of 35

39

Environmental Science & Technology

TOC/Abstract Art

40

41 42

3

ACS Paragon Plus Environment

Environmental Science & Technology

Page 4 of 35

43

Introduction

44

Nitrate (NO3–) is an important, photochemically active contaminant in surface waters,

45

atmospheric water drops and particles, waste water, and snow. As shown in Figure 1, the longer-

46

wavelength absorption maximum of aqueous nitrate is at 302 nm, which enables direct

47

photolysis in sunlight. This photolysis proceeds via three main channels1–3 (Figure 2): (1)

48

formation of NO2 and •OH (which is formed via protonation of •O–), (2) formation of NO2– and

49

O(3P), and (3) at least at wavelengths below 300 nm, formation of peroxynitrite, ONOO–.

50

Nitrate photolysis in channel 11,4,5 can be a major source of •OH in atmospheric and surface

51

waters as well as a source of nitrogen oxides.5–7 There is good agreement regarding the quantum

52

yield of this channel, with an average value of Φ(•OH) of (1.35 ± 0.3)% at 298 K for the 302-nm

53

absorption band.4,5,8,9 This quantum yield is higher at short wavelengths, e.g., Φ(•OH) is 9% at

54

254 nm,3 but decreases steadily with increasing wavelength to a nearly constant value of about

55

1% above 300 nm.4,9,10

56

The second nitrate photolysis channel, which forms nitrite, has been examined in several

57

studies recently10–12 and has been a subject of investigation for decades.4,5,9,13 Nitrite and its

58

protonated form, nitrous acid (HONO), are important photochemical sources of •OH and nitric

59

oxide (NO) in a variety of environments. Measurements of Φ(NO2–) in aqueous solutions near

60

room temperature at tropospherically relevant wavelengths (> 290 nm) vary by nearly an order of

61

magnitude, from 0.5% to 4%.8,10,11,13,18,19 The wide range in literature values make it difficult to

62

model this channel and understand the environmental impacts of nitrite and O(3P) formation

63

from nitrate photolysis. While these studies measured the photoformation of nitrite, the other

64

product from channel 2, O(3P), should have the same quantum yield. However, the one study

65

that monitored O(3P) reported a much lower quantum yield of 0.11%,8 which is commonly cited 4

ACS Paragon Plus Environment

Page 5 of 35

Environmental Science & Technology

66

to suggest the nitrite channel is unimportant.12,20–22 We summarize the wide variation in

67

conditions and results from previous studies in Table S1 of the Supplement.

68

The large range in previously reported quantum yields for channel 2 might be due to the

69

different conditions used in past studies. For example, the pH values of illumination solutions

70

ranged from 3 to 11, with some studies not reporting a pH.19 The pH can affect the observed

71

quantum yield due to HONO formation and volatilization to the gas phase.12 In addition, most

72

studies used an organic compound, often formate, to scavenge •OH (reaction 4 in Figure 2),

73

suppressing its concentration and thus protecting nitrite from •OH oxidation (reaction 5).

74

However, reaction of •OH with organic compounds can form superoxide (•O2–; reaction 4),

75

which can convert NO2 to NO2– (reaction 6) and interfere with determinations of Φ(NO2–), as

76

observed recently by Sharko et al.12

77

concentrations, from 10 mM to 5 M;8,10–13,19 although aerosol concentrations can be in this high

78

range, nitrate concentrations in most atmospheric and surface waters are much lower.7,23–25

79

Nitrite quantum yields from Goldstein and Rabani10 were independent of nitrate concentration,

80

but yields from Roca et al.11 varied by over a factor of two with nitrate concentration (Table S1).

81

In short, the influence of experimental conditions, including nitrate concentration, pH, and

82

organic scavenger, has not been fully investigated to clarify which conditions are best used to

83

determine Φ(NO2–).

Past studies have also used a wide range of nitrate

84

Finally, the third channel of nitrate photolysis (Reaction 3), to form peroxynitrite, is poorly

85

understood at environmental wavelengths. ONOO– is difficult to measure due to its rapid

86

isomerization to nitrate (at pH < 6.6), decomposition,26,27 secondary formation from the reaction

87

of NO with •O2–,28 and because it absorbs light in the same region as nitrate and nitrite (Figure

88

1). ONOO– is formed from nitrate photolysis at 254 nm (Φ(OONO–) = 6.5 - 10%),3,10 but its 5

ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 35

89

importance as a product during illumination in the 300-nm absorption band is uncertain.1

90

Goldstein and Rabani10 observed no ONOO– formation during 300-nm illumination (Φ(ONOO–)

91

< 0.2%), but we are aware of no other studies of peroxynitrite formation at environmentally

92

relevant wavelengths.

93

Our overarching goals are to understand the impacts of several environmental variables on

94

the nitrite quantum yield and use our results to assess past reports of Φ(NO2–). To do this, we

95

explore the impacts of pH, the presence of organic scavenger, nitrate concentration, and the

96

nitrate salt cation on the nitrite quantum yield. We use a nitrate concentration, 50 µM, that is the

97

lowest concentration ever used in a reported Φ(NO2–) determination and that is much more

98

similar to typical environmental levels. Finally, we also examine peroxynitrite formation from

99

nitrate illumination at 254 and 313 nm.

100 101

Experimental Methods

102

Materials

103

We used ACS-certified sodium nitrite, sodium nitrate, hydrochloric acid (trace metal grade),

104

and sodium formate from Fisher. 2-nitrobenzaldehyde (98%), potassium nitrate (ACS Reagent

105

>99%), magnesium nitrate (ACS Reagent >99%), ammonium nitrate (>99%), sulfanilamide

106

(>99%), N-1-napthylethylene diamine (ACS Reagent 98%), sodium hydroxide (>98%), and L-

107

cysteine (>98%) were from Sigma-Aldrich. Calcium nitrate was from ACROS (99%) and

108

potassium phosphate monobasic was from EM Science (99%). Solutions were prepared in air-

109

saturated, deionized water from a Milli-Q Plus system (>18.2 MΩ). Stock solutions of nitrite

110

(0.010 M) and nitrate (0.10 M) were prepared annually and stored in an amber glass bottle in a

111

refrigerator. Absorbance spectra were taken periodically to check the stock concentrations using 6

ACS Paragon Plus Environment

Page 7 of 35

Environmental Science & Technology

112

a UV-vis spectrophotometer (Shimadzu). Dilutions for standards and illuminations were made

113

daily.

114 115

Measurement of Nitrite

116

Within 1 minute of stopping or pausing illumination, we initiated nitrite determination using 29–32

117

the Griess Method (e.g., refs.

), a colorimetric method that forms a strongly absorbing azo-

118

dye complex with nitrite. To 1.0 mL of sample (or nitrite standard) we added 25 µL of 1.0%

119

sulfanilamide in 10% (w/v) HCl solution, allowed the sample to react for 10 min, then added 25

120

µL of 0.10% N-1-napthylethylene diamine solution, and waited 10 min. We then measured light

121

absorption at 540 nm using a TIDAS II spectrophotometer (World Precision Instruments,

122

Sarasota, FL) with a liquid waveguide capillary cell (LWCC; length of 100 cm, effective path

123

length of 94 cm, 250 µL volume), and tungsten lamp. The TIDAS contains two lamps but the

124

deuterium lamp (200-350-nm) causes an artifact, so we turned it off during our measurements

125

(see section S1 in the supporting information for details). Absorption was measured from 350 to

126

700 nm to correct for any baseline shifts. The peak height between 530 and 550 nm was

127

determined as the difference between the maximum absorbance in this wavelength range relative

128

to a baseline determined from local absorption minima between 400 and 500 nm and between

129

550 and 700 nm. With the long pathlength of the TIDAS, we were able to measure nitrite very

130

sensitively, with a detection limit based on replicate blank analyses of 3 nM. We made fresh

131

standards of sodium nitrite each week and calibrated the spectrophotometer using concentrations

132

from 0 to 100 nM (see supplement section S2c). Samples and other solutions were pulled

133

through the LWCC at approximately 1 mL min-1 using a peristaltic pump. Manual syringe

134

injections can also be used, but we found that pulling with the pump gave more consistent results 7

ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 35

135

and fewer air bubbles in the LWCC. We cleaned the LWCC between each measurement with a

136

full cell volume of three separate cleaning solutions (1 M NaOH, 1 M HCl, and 50%

137

methanol/50% water), each separated by an air bubble and followed by rinsing with Milli-Q

138

water.

139

While NO2 can also react in the Griess method to form the same azo dye as NO2–,33 our

140

control experiments with N2 purging to remove NO2 during illumination show no statistically

141

significant effect on the nitrite quantum yield, as discussed in section S3 of the supplemental

142

information. In addition, calculations of partitioning for the sparingly soluble NO2 in our vials

143

indicate that the dissolved fraction of NO2 is small (20%) and should have at most a modest

144

effect on our nitrite quantum yields, in agreement with the control experiments. It is also

145

possible that NO2 is undergoing hydrolysis during illumination to make NO2–: Scharko et al.

146

recently suggested that NO2 produced from nitrate photolysis is more easily hydrolyzed to nitrite

147

compared to NO2 bubbled into solution.12 However, because our relatively rapid N2 flow (15 mL

148

min–1) during purging did not cause a statistically significant decrease in the NO2– quantum yield

149

(Section S3), it seems unlikely that NO2 hydrolysis is significantly interfering with our nitrite

150

determination.

151 152

Sample Illumination

153

Illumination solutions generally contained 50 µM NaNO3. Solutions with a pH of 6 or above

154

were adjusted using a potassium dihydrogen phosphate-sodium hydroxide buffer, while sulfuric

155

acid was used for pH values below 5.5. When used, the concentration of organic scavenger was

156

generally 100 µM for 50 µM nitrate solutions, and was at least 10% of the nitrate concentration.

8

ACS Paragon Plus Environment

Page 9 of 35

Environmental Science & Technology

157

Illuminations were carried out with 313-nm light from a 1000-W Hg/Xe lamp in a

158

monochromatic illumination system (Spectral Energy). Solutions were contained in stirred quartz

159

cuvettes (1-cm or 5-cm path length) or 2.0-mL HPLC vials (low impurity Type I Class A

160

borosilicate glass, 12 × 32 mm; Shimadzu P/N 228-45450-91) where the entire 1 mL of solution

161

was illuminated but not stirred. The sample temperature was controlled at 25°C using a custom

162

Peltier-cooled copper chamber (Paige Instruments). Quartz cuvettes, which were capped during

163

illuminations, were only opened to remove an aliquot at each time point and deliver it to a vial

164

already containing the first Griess reagent. Vials (each representing one time point) were capped

165

during illumination and were only opened to add the Griess reagents. Dark controls contained the

166

same solution as the illuminated sample but were not exposed to light and were analyzed

167

periodically throughout the illumination series for an experiment; there was no production of

168

nitrite in the dark. Periodic controls of deionized water were illuminated for 1 hour; no nitrite

169

was formed. Based on reported molar absorptivities9,34 and measured photon fluxes (see below),

170

the lifetimes of nitrite and aqueous HONO were 1.2 days and 4.5 hours, respectively, in our

171

illuminated solutions.

172 173

Detection of Peroxynitrite

174

As in several past studies,10,28,35,36 we measured light absorption at 302 nm to determine the

175

concentration of peroxynitrite. While peroxynitrite has a published molar absorptivity of 1670

176

M–1cm–1 at 302 nm,37 there is some uncertainty in the value due to the instability of ONOO– in

177

solution and the difficulty in making peroxynitrite. In addition, the absorption spectra of nitrate,

178

nitrite, and peroxynitrite overlap at 302 nm (Figure 1), which confounds light absorption

179

measurements. We tried two different methods to determine the concentration of ONOO– formed 9

ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 35

180

during 254-nm and 313-nm illumination in a pH > 10 solution. A high pH was used to prevent

181

peroxynitrite from protonating and forming peroxynitrous acid (pKa = 6.6), which undergoes

182

rapid isomerization to HNO3.27 In the main ONOO– method, we measured light absorption in the

183

illuminated sample at 302 nm, removed the expected absorbance contributions from nitrite and

184

nitrate, and used the ONOO– molar absorptivity to calculate the amount of peroxynitrite at each

185

time step (Supplemental section S4). We also tried the peroxynitrite method of Plumb and

186

Edwards,38 but had no success getting this method to work (Supplemental section S5).

187 188

Calculation of Quantum Yields

189

We measured the photon flux (Iλ) daily (or more often if experimental conditions changed

190

significantly) using the chemical actinometer 2-nitrobenzaldehyde (2NB)35 under the same

191

conditions (container, light, and temperature) as the nitrate photolysis experiment; for

192

experiments with high nitrate and scavenger concentrations we also added these species to the

193

actinometry solution. Aliquots of a 10 µM 2NB solution were illuminated for varying lengths of

194

time, then the concentration of remaining 2NB was measured using HPLC (C18 Column,

195

60%/40% acetonitrile/water eluent at 0.7 mL min–1, absorbance at 258 nm).39 Under the low

196

light-absorbing conditions of our actinometer, the measured rate constant for 2NB loss (j2NB,λ) is

197

equal to:40

198

j2NB,λ =2.303 × 10 Iλ × ,  ,

(1)

199

where Iλl is the surface-area-normalized photon flux (mol-photon cm-2 s-1 nm-1), ,  ,

200

(= 640 M–1 cm–1)40 is the product of the base-10 molar absorptivity and quantum efficiency for

201

2NB at 313 nm, and 2.303 converts from base-10 to base-e. Average j2NB values were 0.018 s–1

202

for vials and 0.005 s–1 for quartz cuvettes under typical conditions. Under the low light-absorbing 10

ACS Paragon Plus Environment

Page 11 of 35

Environmental Science & Technology

203

conditions of our nitrate samples (absorbance at 313 nm < 0.16), the rate constant of nitrite

204

formation is:

205

    (2) NO  → NO  = 2.303 × 10 ΦNO    , 

206

where ΦNO is

 is the quantum yield of nitrite formation from nitrate photolysis and   ,

207

the base-10 molar absorptivity of nitrate at the illumination wavelength (  , =5.29 M-1cm-1,

208

 , !" = 3.52 M–1cm–1)9.

209 210

The rate of nitrite formation during illumination is equal to: #[ %] #'

  = NO  → NO  [NO ]

(3)

211

where t is time. We illuminate samples over short times where the change in nitrite concentration

212

is linear and the formation rate is determined by a simple linear regression (see supplemental

213

section S1d). This is in contrast to some past methods where experiments were sometimes

214

performed over long time scales, resulting in non-linear plots of nitrite versus time.13,18

215 216

Combining equations 1 through 3 we solve for the quantum yield of NO2-: ΦNO

 =

#[ %] #'

(%)*,+ ,%-.,+

×/

[ %-.,+ ()0 ]  ,+

217

Similarly, the quantum yield for peroxynitrite is

218

ONOO  =

219

#[ ] #'

×/

(%)*,+ ,%)*,+

[ %)*,+ ()0 ]  ,+

(4)

(5)

For a given experiment, the relative standard error ranges from 1-10% for

#[ %] #'

and 1-7%

220

for  , . In comparison, the relative error for ,  , is 8% and for  , is 1%.

221

Errors on individual quantum yields (e.g., in figures) represent 1 standard error determined by

222

propagating measured or reported errors in all terms in equation 4. Errors on means were

223

determined from multiple individual experiments and are 1 standard deviation. 11

ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 35

224 225

Results and Discussion

226

Influence of Organic Scavengers and pH

227

While organic compounds can alter the apparent quantum yield of nitrite, presumably by

228

scavenging •OH and thus minimizing NO2– oxidation, it is unclear if they can also lead to

229

secondary formation of nitrite from the NO2 + •O2– reaction (Figure 2). To investigate this issue,

230

we first examine the expected fate of •OH in a solution containing nitrite and one or more

231

scavengers, either formate, cysteine, or bicarbonate (HCO3–)/carbonate (CO32–) from the

232

dissolution of atmospheric CO2 (supplemental section S2). We examine the calculated fates of

233



234

to the product of the second-order rate constant of scavenger with •OH and the scavenger

235

concentration (e.g., for formate (Fo) this is k’OH = kFo+OH[Fo]). With no added organics, HCO3–

236

/CO32– and NO2– are the dominant •OH scavengers (Figure 3a). The strength of the nitrite sink

237

for

238

bicarbonate/carbonate sink depends on pH, which governs CO2 partitioning from the gas phase

239

and its aqueous speciation. Between approximately pH 6 and 7 there is a transition zone where

240

the dominant •OH sink changes depending on the amount of NO2– in solution. At higher nitrite

241

concentrations (500 nM), bicarbonate and carbonate are the dominant sinks for •OH above

242

approximately pH 7, while at a low nitrite concentration (20 nM), bicarbonate and carbonate are

243

the dominant sinks for •OH above pH 6.2. Understanding where the experimental conditions fall

244

on Figure 3a is important for determining if •OH is an important sink for nitrite, which would

245

(artificially) reduce the nitrite quantum yield.

OH by quantifying its pseudo-first-order rate constant with each scavenger, k’OH, which is equal



OH depends on the amount of nitrite formed from nitrate photolysis, while the

12

ACS Paragon Plus Environment

Page 13 of 35

Environmental Science & Technology

246

To explore how the theoretical calculations compare to our experimental observations, we

247

performed experiments determining the quantum yield of nitrite, Φ(NO2–), for pH 2 to 10.5

248

solutions containing either 50 µM or 10 mM nitrate (Figure 3b). By the end of illumination, we

249

formed approximately 50 and 500 nM of NO2– in the 50 µM and 10 mM NO3– solutions,

250

respectively. At high pH, where we expect bicarbonate/carbonate to be the dominant •OH

251

scavenger, nitrite quantum yields were approximately constant, though noisy, with a higher value

252

for 50 µM NO3– (Figure 3b). In more acidic solutions, below approximately pH 5, Φ(NO2–)

253

decreases with decreasing pH, but there is quite a bit of scatter in the data. This decrease likely

254

has two causes in solutions with no added organic: (1) increased •OH reaction with NO2– in more

255

acidic solutions as bicarbonate/carbonate scavenging of •OH is less effective (Figure 3a) and, (2)

256

at pH values near and below the pKa of HONO (2.8),41 protonation of nitrite to form HONO,

257

which is likely lost to the gas phase.12 The latter effect has been observed by Scharko et al.,12

258

where HONO production from nitrate photolysis increased with solution acidity. Thus, our

259

values at higher pH are probably most representative of the primary quantum yield for nitrite:

260

average values for pH ≥ 5 in Figure 3b are 1.0% and 0.93% for 50 µM and 10 mM nitrite,

261

respectively. Since these solutions did not contain an organic •OH scavenger, it is possible that

262

our values are lower bounds; we address this issue below. Results from the literature for similar

263

conditions without scavenger (illumination wavelength above 300 nm, pH near 4, 0.01-5 M

264

sodium nitrate, room temperature) are less than our 50 µM nitrate quantum yield, ranging from

265

0.25% to 0.6% 11. Nitrite quantum yields measured at pH values near 4 without a scavenger are

266

likely underestimated because of NO2– scavenging by •OH and volatilization of HONO.

267

To determine if the decrease in Φ(NO2–) in acidic solutions was more influenced by •OH

268

scavenging of nitrite or by evaporation of HONO, we next performed experiments with an 13

ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 35

269

added organic scavenger. Experiments were performed at 25 °C and pH 3-8 in 50 µM NaNO3

270

with three scavenger conditions: (1) no added scavenger, (2) 50 µM sodium formate, or (3) 50

271

µM L-cysteine. As shown in Figure 4, for solutions with a pH of 5 or above, the average (± 1 σ)

272

value of Φ(NO2–) is similar for each of the three conditions: no scavenger (1.11 ± 0.17)%,

273

formate (1.18 ± 0.14)%, and cysteine (1.16 ± 0.04)%. The similar results for the three conditions

274

suggest that carbonate/bicarbonate scavenging of •OH is sufficiently protective of nitrite at these

275

pH values; it is also possible that trace organic contaminants in the solutions or containers are

276

helping to scavenge •OH.42 The similarity in the no scavenger, formate, and cysteine results also

277

suggest there is no significant secondary production of nitrite via reduction of NO2 by •O2– at

278

these low concentrations of nitrate and scavenger. To further test this idea, we performed a set of

279

experiments where we attempted to suppress •O2– and/or NO2 in order to prevent the secondary

280

formation of nitrite via reaction 6 (Figure 2). As described in supplemental section S3, these

281

attempts made no significant difference in Φ(NO2–), suggesting that secondary formation of

282

nitrite is negligible in our experiments with 50 µM NO3–.

283

Although there is significant scatter in the results of Figure 4, at lower pH values (< 4.5)

284

quantum yields decrease both with and without scavenger, consistent with loss of HONO to the

285

gas phase rather than a change in scavenging. Warneck and Wurzinger8 also found that the nitrite

286

quantum yield decreases with increasing acidity, starting somewhere between pH 9 and 5.6.

287

Based on our results, experiments should be performed at pH 5 or above so that HONO

288

evaporation is negligible. In contrast, at least two studies have used pH values below this (Table

289

S1).

290 291

Impact of Nitrate Concentration on Φ(NO2–) 14

ACS Paragon Plus Environment

Page 15 of 35

Environmental Science & Technology

292

To further explore the difference in Φ(NO2–) between 50 µM and 10 mM nitrate solutions

293

(Figure 3b), we examined the impact of nitrate concentration for solutions with and without

294

organic scavengers at pH 7. The concentration of organic scavenger varied from 100 to 1000 µM

295

depending on the initial nitrate concentration; organic:nitrate ratios were 0.1 to 2, similar to

296

previous studies.10,11 The results of these experiments are shown in Figure 5 along with results

297

from several past studies. At our lowest nitrate concentration of 50 µM, Φ(NO2–) is the same

298

with and without an organic scavenger, consistent with the results in Figure 3. We are the first to

299

show this, likely because we are working at much lower nitrate concentrations than past work. In

300

contrast to the scavenger having no effect at low nitrate concentrations, our quantum yields with

301

and without scavenger diverge at higher nitrate concentrations. In 5 and 10 mM NO3– solutions

302

containing formate or cysteine, Φ(NO2–) is up to 40% larger than at 50 µM NO3– (with or

303

without scavenger). This enhancement is likely because of secondary formation of nitrite via

304

reaction of nitrogen dioxide and superoxide (reaction 6 in Figure 2), consistent with recent

305

observations by Scharko et al.12

306

scavenger at 5 or 10 mM NO3– is 20 – 60% smaller than the 50µM NO3– value. Experiments at

307

50 µM NO3– form approximately 40 nM of nitrite over the course of illumination (10 min) while

308

10 mM nitrate solutions form approximately 500 nM of nitrite after 1 minute. Thus the lower

309

nitrite quantum yield in the high concentration nitrate solutions without scavenger is consistent

310

with our hypothesis that the dominant •OH sink changes from HCO3–/CO32– to NO2– as the nitrite

311

concentration increases (Figure 3a).

In contrast, the nitrite quantum yield in the absence of

312

At 10 mM NO3–, our no-scavenger experiments agree well with results from Roca et al.11

313

while our with-formate results are significantly higher, suggesting losses of nitrite in their pH 4

314

solutions (Figure 5). Roca et al.11 did not observe a trend in Φ(NO2–) over their nitrate 15

ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 35

315

concentration range (0.01 - 5 M), although the quantum yields vary significantly. Goldstein and

316

Rabani10 report an average Φ(NO2–) of 0.94% for an unknown number of 0.02 - 1 M nitrate

317

solutions containing 10 mM formate at pH 4, somewhat lower than our results. Older work from

318

Daniels et al.13, using solutions containing 1.0 M nitrate without a scavenger at neutral pH,

319

observed a quantum yield of 4%, which is far outside the range of other results (see Table S1).

320

Our results are most similar to Warneck and Wurzinger,8 who examined Φ(NO2–) in the 300-nm

321

region (room temperature, pH 5.6, 10 mM NaNO3, 2-propanol scavenger) and reported an

322

average quantum yield of (1.02 ± 0.07)%.

323 324

Influence of Nitrate Salt Cation

325

We next examined the effect of nitrate salt on Φ(NO2–). In all experiments up to this point,

326

we used sodium nitrate, but here we present Φ(NO2–) measurements for four additional salts:

327

calcium nitrate (Ca(NO3)2), magnesium nitrate (Mg(NO3)2), ammonium nitrate (NH4NO3), and

328

potassium nitrate (KNO3) at 25°C, pH 7, and 50 µM NO3¯. Values of Φ(NO2–) range from (0.9 ±

329

0.1)% for NH4NO3 to (1.16 ± 0.13)% for KNO3 (Figure 6), but none of the results are

330

statistically different at p < 0.05 and the average quantum yield is (1.13 ± 0.15)% for our data.

331

Two past studies have reported nitrite quantum yields from nitrate salts other than NaNO3

332

(Figure 6 and Table S1): Roca et al.11 reported values for Φ(NO2–) from Ca(NO3)2 that are

333

significantly higher (with scavenger) or lower (without scavenger) than our results, while Alif

334

and Boule19 reported a quantum yield for nitrite from KNO3 that is approximately half our value.

335

Finally, our finding that the partner cation of nitrate has no effect on Φ(NO2–) in solution is

336

consistent with recent results showing the •OH quantum yield (reaction 1 in Figure 1) in solution

337

is also independent of cation.43 In contrast, this same work found that nitrate photolysis in thin 16

ACS Paragon Plus Environment

Page 17 of 35

Environmental Science & Technology

338

aqueous films is affected by cation, with rates of NO2 release for KNO3, Mg(NO3)2, and NaNO3

339

approximately 3.5, 2.3, and 2.3 times higher, respectively, than Ca(NO3)2.

340

Wavelength Dependence

341

Since previous explorations of the nitrite quantum yield above 300 nm have used differing

342

wavelengths, we also tested for any wavelength dependence of Φ(NO2–). We conducted a series

343

of experiments over one month, varying only the illumination wavelength. As shown in Figure

344

S5, while there is some noise in our results, the quantum yield is essentially the same at 302, 305,

345

313, and 325 nm.

346

wavelengths, but the molar absorptivity of nitrate at 334 nm is very small (0.4 M–1 cm–1; Figure

347

S5) and difficult to measure; we hypothesize this parameter might be leading to an incorrectly

348

large quantum yield. We also performed experiments to confirm that the nitrite quantum yield is

349

independent of light intensity, as shown in Figure S6.

Our result at 334 nm is approximately 65% higher than at the other

350 351

Peroxynitrite Formation

352

To test the peroxynitrite (ONOO–) detection technique described in the methods section, we

353

first performed experiments at 254 nm to compare with reported quantum yields. Example

354

absorbance spectra for 254-nm illuminations at t = 0 min (black line) and t = 70 min (dashed

355

blue line) are shown in Figure 7a. The peak at time zero is from the initial 10 mM nitrate in

356

solution. After illumination we observe an increase in absorbance at all wavelengths (250 to

357

approximately 420 nm), which could be a result of photoformed nitrite or peroxynitrite (see

358

Figure 1). To determine the concentration of ONOO–, we subtracted the contributions of both

359

nitrate and nitrite from the absorbance spectrum, and assumed the residual absorbance was only

360

from ONOO– (Supplemental Section S4). The time profiles of the residual absorbance at 302, 17

ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 35

361

370, and 390 nm during 254-nm illumination are shown in the filled symbols in Figure 7b. The

362

ratio of the residual absorbance at the two wavelengths where ONOO– absorbs (for example,

363

A370nm:A390nm) is equal to the ratio of molar absorptivities between the same wavelengths (εONOO-

364

,370nm:εONOO-,390nm),

365

four experiments with 1 to 10 mM nitrate at pH 12 and illumination at 254 nm: the average (± 1

366

σ) quantum yield for ONOO– is (6.5 ± 1.5)%, which agrees well with measurements by

367

Goldstein and Rabani10 (6.5%) but is lower than that from Mark et al.3 (10%).

consistent with the residual absorbance being due to ONOO–. We performed

368

We then performed the same experiment with 313-nm illumination. In comparison to the

369

254-nm results, the main peak of absorbance (250 to 330 nm) did not change after 130 minutes

370

of illumination at 313 nm (Figure 7a). There is an increase in the absorbance between 350 and

371

390-nm, corresponding to photoformed nitrite. The residual absorbances at 302, 370, and 390 nm

372

show no increase during 313-nm illumination (Figure 7b). Instead, there is a slight decrease in

373

the residual absorbance, which is likely due to the loss of nitrate, since we assume the nitrate

374

concentration is constant over the course of illumination. (For comparison, we calculate that 4%

375

of the initial nitrate should have been lost during the 130 min of illumination, corresponding to

376

decreases in the residual absorption coefficient of 0.0003, 0.002, and 2.6E-6 cm–1 at 302, 370,

377

and 390 nm, respectively.) We performed additional experiments with 50 µM NaNO3 and

378

obtained similar outcomes. Based on these results, and the good agreement between our

379

peroxynitrite quantum yield at 254 nm and the literature, it appears there is no significant

380

ONOO– formation during 313-nm illumination. Based on our limit of detection, we calculated

381

Φ(ONOO–) to be less than 0.26%, similar to results from Goldstein and Rabani,10 who report an

382

upper bound of 0.2%.

383 18

ACS Paragon Plus Environment

Page 19 of 35

Environmental Science & Technology

384

Implications and Environmental Impacts

385

Our recommended value for Φ(NO2–) is (1.1 ± 0.2)% at 25°C, which is the average of our 61

386

experiments with 50 µM NO3–, pH ≥ 5, with and without scavengers. This value is 17% higher

387

than the recently reported average from Goldstein and Rabani.10 In contrast, our recommended

388

value is 10 times higher than the quantum yield for channel 2 determined from O(3P)

389

measurements by Warneck and Wurzinger.8 This is significant because a number of studies refer

390

to this low quantum yield (e.g., refs 12,20–22,44,45) as the nitrite yield. Perhaps because of this

391

perception, a number of modeling studies include only the production of NO2 – and not N(III)

392

(i.e., NO2– + HONO) production – from nitrate photolysis.46–49 However, our results, along with

393

those from past work,8,10–12,19 indicate that the rate of production of NO2–and HONO from nitrate

394

photolysis is comparable to the rate of NO2 formation.

395

This often neglected source of NO2– might help reconcile model underpredictions of daytime

396

concentrations of gaseous HONO in the atmosphere, which is an important source of •OH.50–53

397

Several explanations have been offered for this underprediction, including heterogeneous

398

chemistry54 and surface photolysis.55–57 Including multiple HONO formation channels has

399

improved the ability of models to reproduce observations,58–60 but discrepancies remain. The

400

photolysis of particulate nitrate to form nitrite is often not included when HONO sources are

401

analyzed58 or only the NO2 production channel is considered as a source.46 There is evidence that

402

aerosol nitrate is associated with an increase in HONO59,61 but many attribute the HONO source

403

to heterogeneous chemistry,60,62,63 even though others suggest the unknown source of nitrate is

404

correlated with other photolysis terms.33,61,64–67 In some cases, the unknown photolytic HONO

405

source is the dominant daytime term.68,69

19

ACS Paragon Plus Environment

Environmental Science & Technology

Page 20 of 35

406

To estimate the contribution of aerosol nitrate photolysis to HONO(g) production, we assume

407

an acidic, aqueous aerosol such that every NO2– formed leads to release of HONO. Under Davis,

408

CA summer solstice conditions with an aerosol nitrate concentration of 0.5 µg/m3 (8 nmol/m3),

409

photolysis will form approximately 0.3 ppt/hr of HONO. In winter in the Central Valley of

410

California, aerosol nitrate concentrations can be much higher,70 up to 30 µg/m3, corresponding to

411

a HONO source of 6 ppt/hr under winter solstice radiation. However, this is not enough to

412

provide closure to atmospheric measurements, which tend to differ several hundred ppt/hr64,69,71

413

or more,65–68,72, but it is a step in the right direction. Other studies suggest that photolysis of

414

nitrate on surfaces can be faster than in bulk solution. For example, it has been reported that the

415

molar absorptivity of surface-adsorbed nitrate is approximately 75 times greater than that of

416

aqueous nitrate73 and that photolysis of nitrate (measured by •OH production) occurs 4 orders of

417

magnitude faster in urban grime than in aqueous solution.74 However, the type of surface may be

418

important: Ye et al.56 found that while the photolysis rate for surface HNO3/nitrate was up to 3

419

times larger than gas-phase HNO3, the major product released was different for natural surfaces

420

(where HONO dominated) and artificial surfaces (where NOx dominated). Properly accounting

421

for these surface reactions in models will require measurements of both molar absorptivities and

422

quantum yields and an understanding of how these values vary with sample environment.

423 424

Acknowledgements

425

We thank Emily Lucic and Alex Funderburk for assistance with the experiments. This work

426

was funded by the Arctic Natural Sciences Program of the National Science Foundation (ANS-

427

1204169). 20

ACS Paragon Plus Environment

Page 21 of 35

428 429

Environmental Science & Technology

Supporting Information. Additional experiment details, calculation details, effect of some experimental variables, Tables S1-S5, Figure S1-S6

430

21

ACS Paragon Plus Environment

Environmental Science & Technology

431

Page 22 of 35

Figures

432 1800

28

ONOO–

Illumination

εNO3–, εNO2– (M–1cm–1)

24

NO2–

20

1400 1200 1000

16

800

12

600

8 4

1600 εONOO– (M–1cm–1)

32

400

NO3–

200 0

0 250 275 300 325 350 375 400 λ (nm)

433 434

Figure 1. Base-10 molar absorptivity of nitrate (NO3–)9, nitrite (NO2–)34, and peroxynitrite

435

(ONOO–)35. The spectral distribution of the illumination system output at 313 nm, in arbitrary

436

units, is included in purple.

437

22

ACS Paragon Plus Environment

Page 23 of 35

Environmental Science & Technology

438 NO3– + hν 3 2 1

3

ONOO–

NO2– + O( P) hν

NO2 + •OH

NO + •OH org.

5

+

H

4 6 •

NO2 + OH–

HONO (aq)

O2– HONO (g)

439 440 441

Figure 2. Primary and secondary reactions occurring during nitrate photolysis. The secondary formation of nitrite from reaction of NO2 with •O2– is indicated by the dashed lines (reaction 6).

442 443 444

23

ACS Paragon Plus Environment

Environmental Science & Technology

Page 24 of 35

a) 1.0E+07 Cysteine

1.0E+06

k'OH (s–1)

1.0E+05 Formate



2–

HCO3 + CO3

1.0E+04 –

500 nM NO2

1.0E+03



100 nM NO2



20 nM NO2

1.0E+02 1.0E+01 1 445

3

5

7

9

11

pH

b) –



H2CO3 ↔

HONO ↔ NO2

1.6%

HCO3

2–

↔ CO3

1.4%

Φ(NO2‾)

1.2% 1.0% 0.8% 0.6% 10 mM

0.4%

50 uM

0.2% 0.0% 1 446

3

5

7

9

11

pH

447 448

Figure 3: a) Calculated pseudo-first-order rate constants for •OH scavenging by nitrite, 50 µM

449

formate (orange line), 50 µM cysteine (purple line), and bicarbonate/carbonate (green line) in

450

equilibrium with atmospheric CO2 at a given pH. The three horizontal black lines represent the

451

rate constants for •OH scavenging for three different nitrite concentrations (20, 100, and 500 24

ACS Paragon Plus Environment

Page 25 of 35

Environmental Science & Technology

452

nM). Nitrite will be protected from •OH under conditions where k’OH for nitrite is significantly

453

below the value for another scavenger at a given pH.b) Results from initial experiments to

454

examine the relationship between Φ(NO2–) and pH for both 50 µM and 10 mM NO3– solutions

455

(313 nm, 25°C, no organic scavenger, pH adjusted with H2SO4 or a phosphate sodium hydroxide

456

buffer). The three thicker vertical lines represent the pKa values for HONO (2.8)41, H2CO3

457

(6.3)75, and HCO3– (10.33)75.

458 459 460 1.6%

1.4

1.4%

1.2 1.0

1.0%

0.8

0.8%

0.6

0.6%

Formate Cysteine 0.4 No Org (Vial) No Org (Cuvette) 0.2 Average pH > 5 Mole Fraction

0.4% 0.2% 0.0%

0.0

2 461

Mole Fraction of NO2-

Φ(NO2–)

1.2%

3

4

5

6

7

8

9

pH

462

Figure 4. Quantum yield of nitrite (50 µM NaNO3, 313 nm, 25°C) as a function of pH

463

without and with 50 µM organic scavenger. The open circles and squares represent experiments

464

performed in glass HPLC vials and quartz cuvettes, respectively. The dashed line is the mole

25

ACS Paragon Plus Environment

Environmental Science & Technology

Page 26 of 35

465

fraction of N(III) that is present as nitrite. pH was adjusted with either sulfuric acid (pH ≤ 5.6) or

466

potassium dihydrogen phosphate (pH ≥ 6).

467

26

ACS Paragon Plus Environment

Page 27 of 35

Environmental Science & Technology

468 469 This Work

470

Roca et al.

Warneck & Wurzinger

471

1.6%

Φ(NO2‾)

1.2%

0.8%

0.4%

0.0% 0.00001

0.0001

0.001 0.01 0.1 Initial Nitrate Concentration (M)

1

10

472 473 474

Figure 5. Quantum yield of nitrite, Φ(NO2–), as a function of nitrate concentration at pH 7

475

(313 nm, 25°C) with and without cysteine or formate (50-1000 µM, depending on [NO3–]) .

476

Triangles represent data from Roca et al.11 for pH 4 solutions with (orange filled) and without

477

(open) 10 mM formate (Fo), illuminated at 310 nm and 25°C. Squares represent data from

478

Warneck and Wurzinger8 for pH =5.6 solutions (305 nm and 0.13 M acetone (open) or 0.001M

479

formate (solid)) where the quantum yield for reaction 2 was determined either from NO2– (dark 27

ACS Paragon Plus Environment

Environmental Science & Technology

Page 28 of 35

480

green) or oxygen atom, O(3P) (light green line). The blue line is the average of data from

481

Goldstein and Rabani10 for pH 4.2-4.5 solutions containing 10 mM formate at 24°C with 300-nm

482

illumination. The dark purple diamond is the average of data from Alif and Boule19 for solutions

483

of 0.1 M KNO3 with formate at 310 nm (no temperature or pH are reported). The red “x” is the

484

average of two data points from Dubowski et al.18 for solutions of 10 mM NaNO3 illuminated at

485

313 nm and 20°C. Lines are qualitative fits to our data.

486 487

28

ACS Paragon Plus Environment

Page 29 of 35

Environmental Science & Technology

1.6%

This Study

Roca et al. no Fo

Roca et al. w/Fo

Goldstein and Rabani

Warneck and Wurzinger

Dubowski et al

Alif and Boule

Φ(NO2‾)

1.4% 1.2% 1.0% 0.8% 0.6% 0.4% 0.2% 0.0%

488

NaNO3 NaNO 3

Ca(NO3)2 Ca(NO3)2

Mg(NO3)2 Mg(NO3)2

NH4NO3 NH4NO3

KNO3 KNO 3

489 490 491

Figure 6. The influence of nitrate salt on the quantum yield of nitrite for experiments from

492

this work (solid black bars: 313 nm, 25°C, 50 µM NO3–, no scavenger, pH > 7), Roca et al.11

493

(310 nm, 25°C, 10 mM NO3–, no formate (white bar) or 10 mM formate (light orange bar)),

494

Goldstein and Rabani10 (striped blue bar; 24°C, 300 nm, 0.02-1.0 M NO3–, 10 mM formate, pH

495

4.2-4.5), Warneck and Wurzinger8 (green bar; 305 nm, 22°C, 0.01 M NO3¯, 0.13 M 2-propanol,

496

pH 5.6) Dubowski et al.18 (red-hatched bar, 313 nm, 0.01 M NO3¯, 293K, no pH reported) and

497

Alif and Boule19 (purple bar; 310 nm, 0.1 M KNO3, 0.5 M formate, no temperature or pH

498

reported). Error bars for our data are ± 1 σ, while bars for other data are the reported errors.

499 500 501 29

ACS Paragon Plus Environment

Environmental Science & Technology

Page 30 of 35

502 a)

0.50

Absorbance

TBefore = 0 Illumination 0.40

λ=313nm 313 nm, t t=130min =130 min

0.30

λ=254nm 254 nm, tt=70min =70 min

0.20

0.10

0.00 250

300

503 b)

350 Wavelength (nm)

400

450

Residual Abs. Coefficient (cm-1)

0.030 λ(ill): 254 λ(msmt): 302

0.025 0.020 0.015

λ(ill): 254 λ(msmt): 370

0.010 0.005

λ(ill): 254 λ(msmt): 390

0.000 λ(ill): 313

-0.005

0 504

50

100

150

Time (min)

505

Figure 7. Determination of peroxynitrite by UV absorbance measurements in solution (10

506

mM NaNO3, pH 13, 25 °C) illuminated with either 254 or 313 nm radiation. Panel a) shows UV

507

absorbance scans (5-cm pathlength) taken before illumination (t = 0; black line), after 70 min of

508

254-nm illumination (dashed blue line), and after 130 min of 313-nm illumination (dashed gray

509

line). Panel b) shows residual UV absorption coefficients (cm–1) as a function of time during 30

ACS Paragon Plus Environment

Page 31 of 35

Environmental Science & Technology

510

illumination (λ(ill)) at 254 nm (filled symbols) or 313 nm (open symbols). The residual

511

absorption coefficient is the pathlength-normalized absorbance after subtracting contributions

512

from nitrate and nitrite (see supplemental section S4 for calculation details). Absorption

513

coefficient values are shown for three different measurement wavelengths, λ(msmt): 302 nm

514

(circles), 370 nm (diamonds), and 390 nm (squares).

515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549

References (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13)

Mack, J.; Bolton, J. R. Photochemistry of nitrite and nitrate in aqueous solution: a review. J. Photochem. Photobiol. Chem. 1999, 128 (1–3), 1–13. Sharpless, C. M.; Linden, K. G. UV Photolysis of Nitrate: Effects of Natural Organic Matter and Dissolved Inorganic Carbon and Implications for UV Water Disinfection. Environ. Sci. Technol. 2001, 35 (14), 2949–2955. Mark, G.; Korth, H.-G.; Schuchmann, H.-P.; von Sonntag, C. The photochemistry of aqueous nitrate ion revisited. J. Photochem. Photobiol. Chem. 1996, 101 (2), 89–103. Zellner, R.; Exner, M.; Herrmann, H. Absolute OH quantum yields in the laser photolysis of nitrate, nitrite and dissolved H2O2 at 308 and 351 nm in the temperature range 278–353 K. J. Atmospheric Chem. 1990, 10 (4), 411–425. Zepp, R. G.; Hoigne, J.; Bader, H. Nitrate-induced photooxidation of trace organic chemicals in water. Environ. Sci. Technol. 1987, 21 (5), 443–450. Haag, W. R.; Hoigné, J. Photo-sensitized oxidation in natural water via. OH radicals. Chemosphere 1985, 14 (11–12), 1659–1671. Brezonik, P. L.; Fulkerson-Brekken, J. Nitrate-Induced Photolysis in Natural Waters:  Controls on Concentrations of Hydroxyl Radical Photo-Intermediates by Natural Scavenging Agents. Environ. Sci. Technol. 1998, 32 (19), 3004–3010. Warneck, P.; Wurzinger, C. Product quantum yields for the 305-nm photodecomposition of nitrate in aqueous solution. J. Phys. Chem. 1988, 92 (22), 6278–6283. Chu, L.; Anastasio, C. Quantum Yields of Hydroxyl Radical and Nitrogen Dioxide from the Photolysis of Nitrate on Ice. J. Phys. Chem. A 2003, 107 (45), 9594–9602. Goldstein, S.; Rabani, J. Mechanism of Nitrite Formation by Nitrate Photolysis in Aqueous Solutions:  The Role of Peroxynitrite, Nitrogen Dioxide, and Hydroxyl Radical. J. Am. Chem. Soc. 2007, 129 (34), 10597–10601. Roca, M.; Zahardis, J.; Bone, J.; El-Maazawi, M.; Grassian, V. H. 310 nm Irradiation of Atmospherically Relevant Concentrated Aqueous Nitrate Solutions: Nitrite Production and Quantum Yields. J. Phys. Chem. A 2008, 112 (51), 13275–13281. Scharko, N. K.; Berke, A. E.; Raff, J. D. Release of Nitrous Acid and Nitrogen Dioxide from Nitrate Photolysis in Acidic Aqueous Solutions. Environ. Sci. Technol. 2014, 48 (20), 11991–12001. Daniels, M.; Meyers, R. V.; Belardo, E. V. Photochemistry of the aqueous nitrate system. I. Excitation in the 300-mµ band. J. Phys. Chem. 1968, 72 (2), 389–399. 31

ACS Paragon Plus Environment

Environmental Science & Technology

550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595

Page 32 of 35

(14) Elshorbany, Y. F.; Crutzen, P. J.; Steil, B.; Pozzer, A.; Tost, H.; Lelieveld, J. Global and regional impacts of HONO on the chemical composition of clouds and aerosols. Atmospheric Chem. Phys. 2014, 14 (3), 1167–1184. (15) Jacobi, H.-W.; Hilker, B. A mechanism for the photochemical transformation of nitrate in snow. J. Photochem. Photobiol. Chem. 2007, 185 (2–3), 371–382. (16) Kleffmann, J. Daytime sources of nitrous acid (HONO) in the atmospheric boundary layer. Chemphyschem Eur. J. Chem. Phys. Phys. Chem. 2007, 8 (8), 1137–1144. (17) Minero, C.; Maurino, V.; Bono, F.; Pelizzetti, E.; Marinoni, A.; Mailhot, G.; Carlotti, M. E.; Vione, D. Effect of selected organic and inorganic snow and cloud components on the photochemical generation of nitrite by nitrate irradiation. Chemosphere 2007, 68 (11), 2111–2117. (18) Dubowski, Y.; Colussi, A. J.; Boxe, C.; Hoffmann, M. R. Monotonic Increase of Nitrite Yields in the Photolysis of Nitrate in Ice and Water between 238 and 294 K. J. Phys. Chem. A 2002, 106 (30), 6967–6971. (19) Alif, A.; Boule, P. Photochemistry and environment Part XIV. Phototransformation of nitrophenols induced by excitation of nitrite and nitrate ions. J. Photochem. Photobiol. Chem. 1991, 59 (3), 357–367. (20) Shi, G.; Buffen, A. M.; Hastings, M. G.; Li, C.; Ma, H.; Li, Y.; Sun, B.; An, C.; Jiang, S. Investigation of post-depositional processing of nitrate in East Antarctic snow: isotopic constraints on photolytic loss, re-oxidation, and source inputs. Atmos Chem Phys 2015, 15 (16), 9435–9453. (21) Boxe, C. S.; Colussi, A. J.; Hoffmann, M. R.; Perez, I. M.; Murphy, J. G.; Cohen, R. C. Kinetics of NO and NO 2 Evolution from Illuminated Frozen Nitrate Solutions. J. Phys. Chem. A 2006, 110 (10), 3578–3583. (22) Grannas, A. M.; Jones, A. E.; Dibb, J.; Ammann, M.; Anastasio, C.; Beine, H. J.; Bergin, M.; Bottenheim, J.; Boxe, C. S.; Carver, G.; et al. An overview of snow photochemistry: evidence, mechanisms and impacts. Atmos Chem Phys 2007, 7 (16), 4329–4373. (23) Hendry, C. D.; Brezonik, P. L. Chemistry of precipitation at Gainesville, Florida. Environ. Sci. Technol. 1980, 14 (7), 843–849. (24) Watanabe, K.; Ishizaka, Y.; Takenaka, C. Chemical characteristics of cloud water over the Japan Sea and the Northwestern Pacific Ocean near the central part of Japan: airborne measurements. Atmos. Environ. 2001, 35 (4), 645–655. (25) Eklund, T. J.; Mcdowell, W. H.; Pringle, C. M. Seasonal variation of tropical precipitation chemistry: La Selva, Costa Rica. Atmos. Environ. 1997, 31 (23), 3903–3910. (26) Kissner, R.; Koppenol, W. H. Product Distribution of Peroxynitrite Decay as a Function of pH, Temperature, and Concentration. J. Am. Chem. Soc. 2002, 124 (2), 234–239. (27) Goldstein, S.; Lind, J.; Merényi, G. Chemistry of Peroxynitrites as Compared to Peroxynitrates. Chem. Rev. 2005, 105 (6), 2457–2470. (28) Løgager, T.; Sehested, K. Formation and decay of peroxynitrous acid: a pulse radiolysis study. J. Phys. Chem. 1993, 97 (25), 6664–6669. (29) Moorcroft, M. J.; Davis, J.; Compton, R. G. Detection and determination of nitrate and nitrite: a review. Talanta 2001, 54 (5), 785–803. (30) Ridnour, L. A.; Sim, J. E.; Hayward, M. A.; Wink, D. A.; Martin, S. M.; Buettner, G. R.; Spitz, D. R. A Spectrophotometric Method for the Direct Detection and Quantitation of Nitric Oxide, Nitrite, and Nitrate in Cell Culture Media. Anal. Biochem. 2000, 281 (2), 223–229. 32

ACS Paragon Plus Environment

Page 33 of 35

596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641

Environmental Science & Technology

(31) Huang, G.; Zhou, X.; Deng, G.; Qiao, H.; Civerolo, K. Measurements of atmospheric nitrous acid and nitric acid. Atmos. Environ. 2002, 36 (13), 2225–2235. (32) Pratt, P. F.; Nithipatikom, K.; Campbell, W. B. Simultaneous determination of nitrate and nitrite in biological samples by multichannel flow injection analysis. Anal. Biochem. 1995, 231 (2), 383–386. (33) Villena, G.; Bejan, I.; Kurtenbach, R.; Wiesen, P.; Kleffmann, J. Interferences of commercial NO2; instruments in the urban atmosphere and in a smog chamber. Atmospheric Meas. Tech. 2012, 5 (1), 149–159. (34) Chu, L.; Anastasio, C. Temperature and Wavelength Dependence of Nitrite Photolysis in Frozen and Aqueous Solutions. Environ. Sci. Technol. 2007, 41 (10), 3626–3632. (35) Kissner, R.; Beckman, J. S.; Koppenol, W. H. Peroxynitrite studied by stopped-flow spectroscopy. Methods Enzymol. 1999, 301, 342–352. (36) Li, D.; Wang, L.; Zeng, X.; Zou, G. Spectrophotometric Determination of Peroxynitrite Using o ‐Phenylenediamine as a Probe. Anal. Lett. 2004, 37 (14), 2949–2963. (37) Hughes, M. N.; Nicklin, H. G. The chemistry of pernitrites. Part I. Kinetics of decomposition of pernitrous acid. J. Chem. Soc. Inorg. Phys. Theor. 1968, 450. (38) Plumb, R. C.; Edwards, J. O. Color centers in UV-irradiated nitrates. J. Phys. Chem. 1992, 96 (8), 3245–3247. (39) Galbavy, E. S.; Ram, K.; Anastasio, C. 2-Nitrobenzaldehyde as a chemical actinometer for solution and ice photochemistry. J. Photochem. Photobiol. Chem. 2010, 209 (2–3), 186– 192. (40) Anastasio, C.; Faust, B. C.; Allen, J. M. Aqueous phase photochemical formation of hydrogen peroxide in authentic cloud waters. J. Geophys. Res. Atmospheres 1994, 99 (D4), 8231–8248. (41) Anastasio, C.; Chu, L. Photochemistry of Nitrous Acid (HONO) and Nitrous Acidium Ion (H2ONO+) in Aqueous Solution and Ice. Environ. Sci. Technol. 2009, 43 (4), 1108–1114. (42) Chen, Z.; Chu, L.; Galbavy, E. S.; Ram, K.; Anastasio, C. Hydroxyl radical in/on illuminated polar snow: formation rates, lifetimes, and steady-state concentrations. Atmos Chem Phys 2016, 16 (15), 9579–9590. (43) Richards-Henderson, N. K.; Anderson, C.; Anastasio, C.; Finlayson-Pitts, B. J. The effect of cations on NO2 production from the photolysis of aqueous thin water films of nitrate salts. Phys. Chem. Chem. Phys. 2015, 17 (48), 32211–32218. (44) Berhanu, T. A.; Meusinger, C.; Erbland, J.; Jost, R.; Bhattacharya, S. K.; Johnson, M. S.; Savarino, J. Laboratory study of nitrate photolysis in Antarctic snow. II. Isotopic effects and wavelength dependence. J. Chem. Phys. 2014, 140 (24), 244306. (45) Legrand, M.; Preunkert, S.; Frey, M.; Bartels-Rausch, T.; Kukui, A.; King, M. D.; Savarino, J.; Kerbrat, M.; Jourdain, B. Large mixing ratios of atmospheric nitrous acid (HONO) at Concordia (East Antarctic Plateau) in summer: a strong source from surface snow? Atmos Chem Phys 2014, 14 (18), 9963–9976. (46) Leriche, M.; Voisin, D.; Chaumerliac, N.; Monod, A.; Aumont, B. A model for tropospheric multiphase chemistry: application to one cloudy event during the CIME experiment. Atmos. Environ. 2000, 34 (29–30), 5015–5036. (47) Zatko, M.; Geng, L.; Alexander, B.; Sofen, E.; Klein, K. The impact of snow nitrate photolysis on boundary layer chemistry and the recycling and redistribution of reactive nitrogen across Antarctica and Greenland in a global chemical transport model. Atmospheric Chem. Phys. 2016, 16 (5), 2819–2842. 33

ACS Paragon Plus Environment

Environmental Science & Technology

642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687

Page 34 of 35

(48) Herrmann, H.; Ervens, B.; Jacobi, H.-W.; Wolke, R.; Nowacki, P.; Zellner, R. CAPRAM2.3: A Chemical Aqueous Phase Radical Mechanism for Tropospheric Chemistry. J. Atmospheric Chem. 2000, 36 (3), 231–284. (49) Frey, M. M.; Roscoe, H. K.; Kukui, A.; Savarino, J.; France, J. L.; King, M. D.; Legrand, M.; Preunkert, S. Atmospheric nitrogen oxides (NO and NO2) at Dome C, East Antarctica, during the OPALE campaign. Atmos Chem Phys 2015, 15 (14), 7859–7875. (50) Li, G.; Lei, W.; Zavala, M.; Volkamer, R.; Dusanter, S.; Stevens, P.; Molina, L. T. Impacts of HONO sources on the photochemistry in Mexico City during the MCMA2006/MILAGO Campaign. Atmospheric Chem. Phys. 2010, 10 (14), 6551–6567. (51) Perner, D.; Platt, U. Detection of nitrous acid in the atmosphere by differential optical absorption. Geophys. Res. Lett. 1979, 6 (12), 917–920. (52) Czader, B. H.; Rappenglück, B.; Percell, P.; Byun, D. W.; Ngan, F.; Kim, S. Modeling nitrous acid and its impact on ozone and hydroxyl radical during the Texas Air Quality Study 2006. Atmospheric Chem. Phys. 2012, 12 (15), 6939–6951. (53) Spataro, F.; Ianniello, A. Sources of atmospheric nitrous acid: State of the science, current research needs, and future prospects. J. Air Waste Manag. Assoc. 2014, 64 (11), 1232– 1250. (54) Ramazan, K. A.; Syomin, D.; Finlayson-Pitts, B. J. The photochemical production of HONO during the heterogeneous hydrolysis of NO2. Phys. Chem. Chem. Phys. 2004, 6 (14), 3836. (55) Zhou, X.; Gao, H.; He, Y.; Huang, G.; Bertman, S. B.; Civerolo, K.; Schwab, J. Nitric acid photolysis on surfaces in low-NOx environments: Significant atmospheric implications. Geophys. Res. Lett. 2003, 30 (23), 2217. (56) Ye, C.; Gao, H.; Zhang, N.; Zhou, X. Photolysis of Nitric Acid and Nitrate on Natural and Artificial Surfaces. Environ. Sci. Technol. 2016, 50 (7), 3530–3536. (57) Zhang, N.; Zhou, X.; Bertman, S.; Tang, D.; Alaghmand, M.; Shepson, P. B.; Carroll, M. A. Measurements of ambient HONO concentrations and vertical HONO flux above a northern Michigan forest canopy. Atmos Chem Phys 2012, 12 (17), 8285–8296. (58) Sarwar, G.; Roselle, S. J.; Mathur, R.; Appel, W.; Dennis, R. L.; Vogel, B. A comparison of CMAQ HONO predictions with observations from the Northeast Oxidant and Particle Study. Atmos. Environ. 2008, 42 (23), 5760–5770. (59) Wong, K. W.; Tsai, C.; Lefer, B.; Grossberg, N.; Stutz, J. Modeling of daytime HONO vertical gradients during SHARP 2009. Atmospheric Chem. Phys. 2013, 13 (7), 3587– 3601. (60) Ziemba, L. D.; Dibb, J. E.; Griffin, R. J.; Anderson, C. H.; Whitlow, S. I.; Lefer, B. L.; Rappenglück, B.; Flynn, J. Heterogeneous conversion of nitric acid to nitrous acid on the surface of primary organic aerosol in an urban atmosphere. Atmos. Environ. 2010, 44 (33), 4081–4089. (61) Liu, Z.; Wang, Y.; Costabile, F.; Amoroso, A.; Zhao, C.; Huey, L. G.; Stickel, R.; Liao, J.; Zhu, T. Evidence of Aerosols as a Media for Rapid Daytime HONO Production over China. Environ. Sci. Technol. 2014, 48 (24), 14386–14391. (62) Wang, S.; Ackermann, R.; Spicer, C. W.; Fast, J. D.; Schmeling, M.; Stutz, J. Atmospheric observations of enhanced NO2-HONO conversion on mineral dust particles. Geophys. Res. Lett. 2003, 30 (11), 1595. (63) Saliba, N. A.; Chamseddine, A. Uptake of acid pollutants by mineral dust and their effect on aerosol solubility. Atmos. Environ. 2012, 46, 256–263. 34

ACS Paragon Plus Environment

Page 35 of 35

688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727

Environmental Science & Technology

(64) Kleffmann, J.; Gavriloaiei, T.; Hofzumahaus, A.; Holland, F.; Koppmann, R.; Rupp, L.; Schlosser, E.; Siese, M.; Wahner, A. Daytime formation of nitrous acid: A major source of OH radicals in a forest. Geophys. Res. Lett. 2005, 32 (5), L05818. (65) Michoud, V.; Colomb, A.; Borbon, A.; Miet, K.; Beekmann, M.; Camredon, M.; Aumont, B.; Perrier, S.; Zapf, P.; Siour, G.; et al. Study of the unknown HONO daytime source at a European suburban site during the MEGAPOLI summer and winter field campaigns. Atmospheric Chem. Phys. 2014, 14 (6), 2805–2822. (66) Elshorbany, Y. F.; Kurtenbach, R.; Wiesen, P.; Lissi, E.; Rubio, M.; Villena, G.; Gramsch, E.; Rickard, A. R.; Pilling, M. J.; Kleffmann, J. Oxidation capacity of the city air of Santiago, Chile. Atmos Chem Phys 2009, 9 (6), 2257–2273. (67) Tang, Y.; An, J.; Wang, F.; Li, Y.; Qu, Y.; Chen, Y.; Lin, J. Impacts of an unknown daytime HONO source on the mixing ratio and budget of HONO, and hydroxyl, hydroperoxyl, and organic peroxy radicals, in the coastal regions of China. Atmos Chem Phys 2015, 15 (16), 9381–9398. (68) Spataro, F.; Ianniello, A.; Esposito, G.; Allegrini, I.; Zhu, T.; Hu, M. Occurrence of atmospheric nitrous acid in the urban area of Beijing (China). Sci. Total Environ. 2013, 447, 210–224. (69) Sörgel, M.; Regelin, E.; Bozem, H.; Diesch, J.-M.; Drewnick, F.; Fischer, H.; Harder, H.; Held, A.; Hosaynali-Beygi, Z.; Martinez, M.; et al. Quantification of the unknown HONO daytime source and its relation to NO2. Atmospheric Chem. Phys. 2011, 11 (20), 10433– 10447. (70) Lurmann, F. W.; Brown, S. G.; McCarthy, M. C.; Roberts, P. T. Processes Influencing Secondary Aerosol Formation in the San Joaquin Valley during Winter. J. Air Waste Manag. Assoc. 2006, 56 (12), 1679–1693. (71) Zhou, X.; Civerolo, K.; Dai, H.; Huang, G.; Schwab, J.; Demerjian, K. Summertime nitrous acid chemistry in the atmospheric boundary layer at a rural site in New York State. J. Geophys. Res. Atmospheres 2002, 107 (D21), 4590. (72) Li, X.; Brauers, T.; Häseler, R.; Bohn, B.; Fuchs, H.; Hofzumahaus, A.; Holland, F.; Lou, S.; Lu, K. D.; Rohrer, F.; et al. Exploring the atmospheric chemistry of nitrous acid (HONO) at a rural site in Southern China. Atmospheric Chem. Phys. 2012, 12 (3), 1497– 1513. (73) Zhu, C.; Xiang, B.; Zhu, L.; Cole, R. Determination of absorption cross sections of surface-adsorbed HNO3 in the 290–330nm region by Brewster angle cavity ring-down spectroscopy. Chem. Phys. Lett. 2008, 458 (4–6), 373–377. (74) Baergen, A. M.; Donaldson, D. J. Photochemical Renoxification of Nitric Acid on Real Urban Grime. Environ. Sci. Technol. 2013, 47 (2), 815–820. (75) Seinfeld, John H. H; Pandis, S. N. N. Atmospheric Chemistry and Physics - From Air Pollution to Climate Change (2nd Edition)., 2nd Edition.; John Wiley & Sons: Hoboken, N.J, 2006.

35

ACS Paragon Plus Environment