Subscriber access provided by TULANE UNIVERSITY
Article
Quantum chemical calculations resolved identification of methylnitrocatechols in atmospheric aerosols Sanja Frka, Martin Šala, Ana Krofli#, Matej Hus, Alen Cusak, and Irena Grgi# Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b00823 • Publication Date (Web): 02 May 2016 Downloaded from http://pubs.acs.org on May 4, 2016
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 34
Environmental Science & Technology
1
Quantum chemical calculations resolved identification of methylnitrocatechols in
2
atmospheric aerosols
3
Sanja Frka*†┴, Martin Šala†, Ana Kroflič†, Matej Huš‡, Alen Čusak§║ and Irena Grgić†
4 5
†
6
┴
Analytical Chemistry Laboratory, National Institute of Chemistry, 1000 Ljubljana, Slovenia Division for Marine and Environmental Research, Ruđer Bošković Institute, 10000 Zagreb, Croatia
7 8
‡
Laboratory of Catalysis and Chemical Reaction Engineering, National Institute of Chemistry, 1000 Ljubljana, Slovenia
9 §
10 ║
11
Alkemika Ltd., 3000 Celje, Slovenia
Acies Bio Ltd., 1000 Ljubljana, Slovenia
12 13
Abstract
14
Methylnitrocatechols (MNCs) are secondary organic aerosol (SOA) tracers and major
15
contributors to atmospheric brown carbon; however, their formation and ageing processes in
16
atmospheric waters are unknown. To investigate the importance of aqueous-phase
17
electrophilic substitution of 3-methylcatechol with nitronium ion (NO2+), we performed
18
quantum calculations of their favorable pathways. The calculations predicted the formation of
19
3-methyl-5-nitrocatechol (3M5NC), 3-methyl-4-nitrocatechol (3M4NC), and a negligible
20
amount of 3-methyl-6-nitrocatechol (3M6NC). MNCs in atmospheric PM2 samples were
21
further inspected by LC/(-)ESI-MS/MS using commercial as well as de novo synthesized
22
authentic standards. We detected 3M5NC and, for the first time, 3M4NC. In contrast to
23
previous reports, 3M6NC was not observed. Agreement between calculated and observed
24
3M5NC/3M4NC ratios cannot unambiguously confirm the electrophilic mechanism as the
25
exclusive formation pathway of MNCs in aerosol water. However, the examined nitration by 1 ACS Paragon Plus Environment
Environmental Science & Technology
26
NO2+ is supported by (1) the absence of 3M6NC in the ambient aerosols analysed and (2) the
27
constant 3M5NC/3M4NC ratio in field aerosol samples, which indicates their common
28
formation pathway. The magnitude of error one could make by incorrectly identifying
29
3M4NC as 3M6NC in ambient aerosols was also assessed, suggesting the importance of
30
evaluating the literature regarding MNCs with special care.
31 32
Introduction
33
Organic aerosols (OA) are essential to many areas of science ranging from Earth’s climate to
34
human health. Despite fundamental advances in understanding how OA form and evolve
35
during atmospheric ageing, the specific ways in which they impact the atmosphere and
36
climate are only beginning to be understood. A significant and highly variable fraction of OA
37
named brown carbon (BrC), a light-absorbing form of particulate matter (PM), has recently
38
attracted much interest because of its potentially large effect on Earth's climate.1–4 Yet, in
39
many global models, organic particulates are still considered to be only light scattering.5 By
40
absorbing solar radiation in (sub)visible wavelengths, BrC heats the atmosphere, significantly
41
affecting the overall atmospheric energy distribution. The literature often refers to BrC as a
42
primary organic aerosol (POA) because it is consistently observed in areas influenced by
43
combustion of both biomass and fossil fuels.6–8 However, there is growing evidence that BrC
44
also results from secondary aging processes. For example, volatile organic compounds
45
(VOCs) in biomass-burning smoke have been oxidized to secondary organic aerosols (SOA)
46
including brown carbon SOA.9–12 Such secondary aging processes often involve nitrogen, and
47
brown SOA have been detected during photooxidation of volatile aromatic precursors (e.g.,
48
toluene, benzene, or xylene) in the presence of NOx.13–16 The resulting product distribution
49
strongly depends on the structure of the aromatic precursor, which substantially affects its
50
reactivity and absorption of light in the near UV and visible regions of the spectrum.
2 ACS Paragon Plus Environment
Page 2 of 34
Page 3 of 34
Environmental Science & Technology
51
Nitroaromatic compounds (NACs) are among the largest and most important groups of
52
industrial chemicals in use today. Especially catechol derivatives form one of the most
53
ubiquitous families of natural antioxidants present in a variety of natural products, flavors,
54
and pharmaceuticals. However, reactions involving NACs in the atmosphere are a matter of
55
concern, as their interactions with DNA and the resulting mutagenicity render them hazardous
56
to human health and wildlife.17 The mutagenicity and carcinogenicity of NACs have been
57
specifically addressed, as influenced by the position of the nitro group on the aromatic ring,
58
considering also presence of other functional groups.17 It has been suggested that NACs and,
59
in particular, abundant methylnitrocatechols (MNCs), which are formed through atmospheric
60
oxidation of biomass-burning plumes, are the major contributors to atmospheric BrC.18–20
61
Recently, yellow-colored MNCs have drawn much attention due to their high concentrations
62
in ambient aerosols,20,21 contributing (with other less abundant NACs) ~1% to the total
63
organic aerosol mass.22,23 Moreover, MNCs rank among the most abundant chromophoric,
64
organic, water-soluble constituents of PM from biomass-burning emissions.19 Based on
65
laboratory and aerosol chamber experiments, MNCs are considered SOA contributors
66
originating from NOx photooxidation of m-cresol, which is released during biomass burning
67
as a thermal degradation product of lignin.20 MNCs have also been suggested as tracer
68
compounds of anthropogenic SOA because the photooxidation of VOCs, such as toluene in
69
the presence of NOx, can lead to the formation of MNCs as multigenerational SOA
70
products.21, 24, 25
71
Cloud droplets, rain, fogs, and water films associated with aerosol particles (wet aerosols)
72
represent different forms of atmospheric liquid water. Concentrated mixtures of salts and
73
organics, and low (acidic) pH are typical of wet aerosols and fogs, in general leading to a
74
favorable chemical environment for the aqueous processing. Aqueous-phase reactions have
75
been the subject of increased interest in the scientific community as they represent potential
3 ACS Paragon Plus Environment
Environmental Science & Technology
76
pathways contributing to the formation and processing of SOA, yet their magnitude in
77
tropospheric models is often underestimated.26 Although aromatic nitration ranks among the
78
most widely studied organic reactions with frequent occurrences in the biosphere, the exact
79
mechanisms remain poorly understood. In aqueous solution, aromatic nitration is proposed to
80
occur via a mechanistically well-defined electrophilic substitution reaction. It has been
81
generally accepted that, under acidic conditions, the reaction involves the nitronium ion,
82
NO2+, as a common nitrating agent, and the formation of Wheland intermediate (σ-
83
complex).27 Alternatively, a free radical mechanism is often proposed to explain the
84
mechanism of aromatic nitration by nitrous acid (HNO2) in the environment with aromatic
85
pollutants, such as substituted phenols, forming the corresponding nitrophenols.28–31
86
However, aqueous-phase nitration of aromatic molecules with nitrite/nitrous acid (NO2-
87
/HNO2) and hydrogen peroxide (H2O2) has also been shown to lead to the formation of
88
NACs,29,30 suggesting that nitrating electrophiles are possible reactive species in the acidic
89
atmospheric waters. Electrophilic substitution has also been used to explain nitration of the
90
wood-burning pollutant guaiacol, under atmospherically relevant conditions (pH = 4.5),
91
providing additional evidence that electrophiles are important reactive species in acidic
92
atmospheric aqueous environments and may contribute substantially to the transformation of
93
aromatic compounds especially during the night time.32 Recent chamber observation33
94
indicates that 4-methyl-5-nitrocatechol (4M5NC) is only formed from 4-methylcatechol after
95
irradiation has stopped, thus, further highlighting that night-time chemistry may be important
96
for this group of atmospheric pollutants.
97
Processing of aromatic compounds is known to form SOA, but subsequent oxidation products
98
are largely unknown or only tentatively identified. To our knowledge, there are no studies on
99
the (photo)nitration of methylcatechols in atmospheric waters nor have their possible nitration
100
products been unambiguously identified. Over the past decades, with increasing computing 4 ACS Paragon Plus Environment
Page 4 of 34
Page 5 of 34
Environmental Science & Technology
101
power and a better understanding of the dynamic processes that govern chemical reactions,
102
computational quantum chemistry has become a reliable tool to support experimental
103
atmospheric chemistry, particularly with regard to reaction mechanisms, conformational
104
energies, thermodynamic properties, and transition states.34–36 Assuming that electrophilic
105
nitration is also relevant for atmospheric methylcatechols (structural isomers of guaiacols), we
106
performed quantum chemical calculations of energetically favorable pathways of electrophilic
107
substitution of 3-methylcatechol (3MC) with NO2+ in aqueous medium and assessed the most
108
favorable reaction products. Due to interest in NACs as important yellow-colored and
109
potentially toxic SOA constituents and to better elucidate their night-time chemistry in the
110
troposphere, we supplemented our investigation with a detailed chemical analysis of MNC
111
compounds in ambient aerosols.
112 113
EXPERIMENTAL SECTION
114 115
Computational method
116
Ab initio quantum chemical calculations were performed with Gaussian 09 program37 at the
117
MP2/6-31++g(d,p).38–42 Preliminary calculations were also performed with B3LYP density
118
functional and yielded comparable energies and geometries. However, B3LYP is notorious
119
for not identifying all energy minima in organic reactions. For aromatic nitration, B3LYP
120
fails to locate certain energy minima, as previously also noted by Esteves et al.43 Taking this
121
into account, MP2 was ultimately chosen as the best compromise between accuracy and
122
computational cost. Aqueous environment was modeled with polarizable continuum model
123
with effective dielectric constant 78.5. Thermodynamic properties were obtained from
124
frequency calculations. Energies and thermodynamic quantities are reported relative to the
125
infinitely separated nitronium ion and catechol molecule at 1 bar and 298 K, correcting for
5 ACS Paragon Plus Environment
Environmental Science & Technology
126
zero-point energy. This is a reasonable assumption because the reaction energies are large
127
enough for small temperature variations not to influence results noticeably. Charges were
128
calculated with the natural bond order analysis (NBO).
129
Geometries of intermediates and transition states were fully optimized without constraints.
130
The nature of each species was determined with vibrational analysis on the optimized
131
geometry. Intermediates were characterized by not having any imaginary frequencies as
132
minima of the potential energy surface (PES), whereas transition states (TS) had exactly one
133
imaginary frequency, corresponding to their saddle position on PES. TS were located with
134
synchronous transit-guided quasi-Newton method (STQN).
135 136
Field samples
137
Size-segregated aerosol samples were collected for 48–72 h on prebaked Al-foils at an urban
138
background site of Ljubljana, Slovenia (approx. 270,000 inhabitants, 298 m a.s.l.) with a
139
Berner cascade impactor (HAUKE, LPI 25/0,015/2; nominal flow rate of 25.8 L min-1)
140
between December 3, 2014 and March 2, 2015. Although the impactor has 10 collection
141
stages of the nominal size ranges of aerodynamic diameter (Dae) from 0.038 to 15.6 µm, for
142
the purpose of this study only stages with PM of Dae up to 2.1 µm were further processed and
143
represented as PM2 values obtained by summing up the corresponding deposits of stages from
144
1 to 6. To ensure a sufficient amount of material for chemical analysis, corresponding stages
145
of 2–4 samples were extracted jointly with high purity Milli-Q water (Millipore, Bedford,
146
MA, USA) and further diluted in the injection solvent containing 50 µg L-1 picric acid
147
(internal standard, IS; Sigma-Aldrich), 7.5 mM ammonium formate buffer pH 3 (ammonium
148
formate: puriss p.a., eluent additive for LC-MS, Fluka; and formic acid: eluent additive for
149
LC-MS, Fluka), and 200 µM ethylenediaminetetraacetic acid (EDTA; 99.995%; Sigma-
6 ACS Paragon Plus Environment
Page 6 of 34
Page 7 of 34
Environmental Science & Technology
150
Aldrich) for the analysis of MNCs by LC-MS/MS. For details regarding the sample collection
151
and pretreatment please refer to SI material.
152 153
LC-MS/MS analysis
154
The water extracts of aerosol samples (as well as commercially available standards and the de
155
novo synthesized standard isolated with the semi-preparative HPLC, see below) were
156
analyzed by UltiMate™ 3000 UHPLC system (Thermo Scientific, USA) coupled with a triple
157
quadrupole/linear ion trap mass spectrometer (4000 QTRAP LC-MS/MS System; Applied
158
Biosystems/MDS Sciex, Ontario, Canada). An analytical HPLC column Atlantis T3 column
159
(3.0 × 150 mm, 3 µm particle size; Waters) with Atlantis T3 guard column (3.0 × 10 mm, 3
160
µm particle size; Waters) was used with the flow rate of 0.3 ml min-1. A mobile phase
161
consisted of methanol (MeOH, Chromasolv® LC-MS grade, ≥ 99.9%, Fluka)/tetrahydrofuran
162
(THF, Chromasolv® Plus, for HPLC, ≥ 99.9%, Fluka)/water (30/15/55, V/V/V) with 7.5 mM
163
ammonium formate buffer, pH 3. The elution program was adopted from Kitanovski et al.44
164
and is given in detail in Table S1. The injection volume and the column temperature were 10
165
µL and 30 °C, respectively.
166
After the separation, MNCs were detected by using negative polarity electrospray ionization
167
((-)ESI) in the selected reaction monitoring (SRM) mode. The SRM transition m/z 168 → m/z
168
138, characteristic of MNCs, was monitored as well as the MS/MS product ion spectra. The
169
operating conditions of mass spectrometer were as follows: -45.0 V for declustering potential
170
(DP), -22.4 V for collision energy (CE), -9.6 V for collision cell exit potential (CXP). Data
171
were acquired and processed using Analyst 1.5 Software (Applied Biosystems/MDS
172
Analytical Technologies Instruments).
173
The quantification of MNCs present in field samples was performed by running a series of
174
commercially available authentic standards: 4-methyl-5-nitrocatechol (4M5NC, 0.5–50 µg 7 ACS Paragon Plus Environment
Environmental Science & Technology
Page 8 of 34
175
L-1, 6 calibration points, linear fitting, R2 > 0.999; Santa Cruz Biotechnology, INC., CA,
176
USA), 3-methyl-5-nitrocatechol (3M5NC, 0.5–50 µg L-1, 6 calibration points, linear fitting,
177
R2 > 0.998; Atomax Chemicals, Shenzhen, China), and 3-methyl-6-nitrocatechol (3M6NC,
178
50–1500 µg L-1, 6 calibration points, linear fitting, R2 > 0.989; Atomax Chemicals, Shenzhen,
179
China) as well as de novo synthesized 3-methyl-4-nitrocatechol (3M4NC, 0.5–50 µg L-1, 6
180
calibration points, linear fitting, R2 > 0.998; synthesized, purified, and identified within this
181
study) containing fixed concentration of 50 µg L-1 of picric acid (aqueous solution 1.0%;
182
Sigma-Aldrich) used as an internal standard. Method limits of detection (LODs) were
183
determined as analyte concentrations that correspond to a signal-to-noise ratio (S/N) of 3. The
184
recovery efficiencies of MNCs water extraction from the foils was also taken into account and
185
varied between 95.3 and 99.6%.
186 187
Synthesis and identification of the 3-methylcatechol nitration product
188
Nitration of 3MC in aqueous solution using a NaNO2/H2SO4 system led to the formation of a
189
MNC reaction mixture.20,44 An Agilent 1100 HPLC system with a quaternary pump and a
190
UV–vis diode array detector (DAD) coupled to an analytical-scale fraction collector was used
191
for semi-preparative isolation of the minor reaction product extract. For details regarding the
192
synthesis and product isolation please refer to SI material. The structure of the purified
193
reaction product was elucidated by combining high-resolution mass spectrometric (HRMS)
194
data and NMR spectra. HRMS measurements were performed with a hybrid quadrupole
195
orthogonal acceleration time-of-flight mass spectrometer (QTOF Premier, Waters, Milford,
196
MA, USA) and the compound was analyzed in negative mode. The capillary voltage was set
197
to 3.0 kV, whereas the sampling cone voltage was 40 V. 1H-NMR,
198
gHMBC were recorded on a Varian Unity Inova 300 MHz NMR spectrometer in the
199
deuterated methanol (CD3OD) at ambient temperature. Chemical shifts (δ) are given in ppm 8 ACS Paragon Plus Environment
13
C-NMR, gHSQC, and
Page 9 of 34
Environmental Science & Technology
200
and calibrated using the signal of undeuterated solvent or internal reference (CD3OD/0.03%
201
Me4Si). Chemical shifts of residual undeuterated solvent are δH = 3.31 ppm and δC = 49.15
202
ppm. 13C-NMR signals with complete proton decoupling are described with a chemical shift
203
in ppm.
204 205
RESULTS AND DISCUSSION
206 207
Theoretical predictions of 3-methylcatechol nitration products in aqueous phase
208
As we were less interested in the formation of reactive species (NO2+) per se, and more in
209
different affinity of all available sites on the aromatic ring for the electrophilic attack of a
210
nitrating agent, we limited theoretical calculations to the interplay of 3MC and NO2+ which
211
has been long known as the key electrophile in the process.27 In addition to the original
212
Ingold-Hughes mechanism, a single-electron transfer process45,46 and several other
213
mechanisms have been proposed (e.g., the Olah mechanism47) which all feature the arenium
214
ion as the key intermediate for reaction rate and selectivity. Different pathways that lead to 4-,
215
5-, and 6-nitrated products were studied. We identified six different stable intermediates
216
(Figure 1) and five transition states (Figure 2), thermodynamic and geometric properties of
217
which are presented in Table S2.
218
9 ACS Paragon Plus Environment
Environmental Science & Technology
219 220 221
Figure 1. Geometries of intermediates for NO2+/3-methylcatechol (3MC) complexes, obtained with
222
energy minimization and verified with vibrational analysis. All distances are in Å.
223
224 225 226
Figure 2. Geometries of transition states for NO2+/3-methylcatechol (3MC) complexes, obtained with
227
STQN and verified with vibrational analysis. All distances are in Å
228
10 ACS Paragon Plus Environment
Page 10 of 34
Page 11 of 34
Environmental Science & Technology
229
The geometries of the reactants (and products) were optimized, finding the most stable
230
structures for the different species. Thus, isolated nitronium cation is perfectly linear with the
231
calculated equilibrium O–N distance 1.15 Å. The positive charge is mostly retained on the
232
nitrogen atom (+1.06) with oxygen atoms being only slightly negative (-0.03). Isolated
233
catechol adopts the expected structure with the aromatic ring, hydroxyl groups, and methyl
234
carbon lying in the same plane. Structure 1 is a shallow energetic minimum, corresponding to
235
the perpendicular approach of NO2+ to the delocalized electronic system of catechol. This
236
early intermediate represents a weakly bounded π-complex of NO2+ and 3MC, and is common
237
to all studied pathways. Nitronium ion is situated 2.52 Å above the center of the aromatic
238
ring, oriented perpendicularly to it. Its geometry remains unperturbed when compared to that
239
of the isolated NO2+, as well as atomic charges. Formation of this complex is energetically
240
and kinetically favorable. As it has already been shown for general aromatic nitration,48 there
241
is no activation barrier for this first step. Instead, the relative energy of the complex
242
monotonically decreases to -28.1 kJ mol-1 as the nitronium species approaches the 3MC
243
molecule.
244
After this step, the reaction mechanism bifurcates into two pathways. Intermediate structure 1
245
can convert into two different δ-reactant complexes, shown as structures 4 and 5 where the
246
nitronium is located approximately parallel to the aromatic ring plane. In structures 4 and 5,
247
the nitrogen atom resides 2.38 Å and 2.43 Å above the ring, halfway between C1–C2 or C4–C5
248
atoms, respectively. The O–N–O angle is reduced to 147°, while the O–N bond is slightly
249
elongated to 1.18 Å. In these δ-reactant complexes, the interaction between the nitronium
250
group and the aromatic ring is stronger, as evidenced by the aromatic ring compensation for
251
the electron deficiency at the nitronium. The nitrogen atom has a charge +0.75 and the oxygen
252
atoms -0.26 each. Thermodynamically, the formation of structure 4 is more favorable than the
253
formation of structure 5 by 13.3 kJ mol-1. Kinetically, however, the reverse is true. The
11 ACS Paragon Plus Environment
Environmental Science & Technology
254
structure 2, corresponding to the transition state between 1 and 4, has a higher energy than the
255
structure 3, corresponding to the transition state between 1 and 5, by 1–2 kJ mol-1.
256
Nevertheless, the activation energy in both cases is small (less than 10 kJ mol-1) and is not the
257
rate-determining step. It should be noted that it is also possible for the nitronium ion to attack
258
the 3MC molecule from a different orientation, effectively skipping 1 and equilibrating
259
directly at 4 or 5. In all cases, the reaction rate was found to be determined by the subsequent
260
step, where an arenium ion is formed. This step is also known to influence the
261
regioselectivity.43 A δ-reactant complex is converted into the corresponding δ-intermediate
262
complex, with nitro group bound to the carbon atom, yielding structures 9, 10, and 11, which
263
are attained through the transition states 6, 7, and 8, respectively. The hydrogen atom remains
264
bound to the carbon and is only cleaved afterwards in a fast, non-rate-limiting step, leading to
265
the formation of respective 4-nitro (3M4NC), 5-nitro (3M5NC), and 6-nitro (3M6NC)
266
substituted 3-methylcatechols. Thermodynamically, the formation of 9 and 10 is favored, as
267
both have similar lower energies (-86.7 and -91.4 kJ mol-1, respectively) in comparison to the
268
nitration on site 6 yielding the structure 11 of higher energy -70.3 kJ mol-1. Kinetic
269
considerations show the same trends, with transition states 6 and 7 having considerably lower
270
energies (-41 and -43 kJ mol-1, respectively) in comparison to 8 (-22 kJ mol-1).
271
The potential energy surface change during nitration of 3-methylcatechol (3MC) is shown in
272
Figure 3 while discussed reaction pathways are schematically presented in Figure 4. Due to
273
lower activation barriers for nitration of 3MC on sites 4 and 5 (+19 and +16 kJ mol-1,
274
respectively) than for nitration on site 6 (+51 kJ mol-1), only negligible amounts of 3M6NC
275
are expected to be formed through this reaction mechanism. Conversely, 3M4NC and 3M5NC
276
should both be produced in larger quantities.
277
12 ACS Paragon Plus Environment
Page 12 of 34
Page 13 of 34
Environmental Science & Technology
278 279 280
Figure 3. The potential energy surface change during nitration of 3-methylcatechol (3MC)
281
with NO2+. Three different pathways corresponding to the formation of 3-methyl-4-
282
nitrocatechol (3M4NC, 9), 3-methyl-5-nitrocatechol (3M5NC, 10), and 3-methyl-6-
283
nitrocatechol (3M6NC, 11) are shown.
284
285 286
13 ACS Paragon Plus Environment
Environmental Science & Technology
Page 14 of 34
287
Figure 4. Schematic presentation of proposed electrophilic nitration pathways of 3-
288
methylcatechol in aqueous solution.
289 290
To our knowledge, neither the energetics of the considered reaction nor the reaction products
291
have been reported previously. According to theoretical predictions and presuming that
292
electrophilic nitration is relevant for atmospheric methylcatechols, as it was previously shown
293
for guaiacols,49 we should expect a dominant presence of 3M5NC and 3M4NC isomers in
294
ambient aerosol samples. Namely, ignoring the effects of fast hydrogen abstraction, a rough
295
estimate for the relative proportion of the final products can be made from Gibbs free energies
296
of the transition states in the rate-determining step.
297
Using Eyring equation ,
298
(1)
299
kB, h, and R representing Boltzman, Planck and gas constants, T is the reference temperature
300
298.15 K, and G# Gibbs free energy (kJ mol-1) taken from Table S2, we obtain the following
301
reaction
rate
constants:
,
,
(corresponding to the relative reaction rates
302 303
: 1). One should note that the reported values agree with a
304
microscopic description of the studied reaction pathways according to the following kinetic
305
expression
306
aware that these kinetic rate constants cannot be used when considering the same reactions
307
macroscopically because macroscopic first order reaction rate constants are usually reported
308
for equilibrium concentrations of measurable reactant species, resulting in the kinetic equation
309
(subscript M refers to macroscopic). Still, due to low temperatures and
310
concentrations in the atmosphere, the actual ratio of products is expected to be in agreement
311
with the presented kinetic analysis (3M4NC:3M5NC=1:2.4; the amount of 3M6NC formed is
(subscript m refers to microscopic). One should be also
14 ACS Paragon Plus Environment
Page 15 of 34
Environmental Science & Technology
312
negligible). Due to necessary approximations and inherent limitations of the method, the ratio
313
should be considered accurate within an order of magnitude.
314
There is a limited number of papers referring to the presence of MNCs in field samples,19, 20,
315
22, 23
316
al.22 and Iinuma et al.20 Previous studies report on the overall domination of 4M5NC,
317
3M5NC, as well as 3M6NC in nitroaromatic fraction of atmospheric aerosols, together with
318
4-nitrocatechol (4NC). However, to date there has been no evidence of 3M4NC in field
319
aerosols; only a chamber study identified the 3M4NC isomer as a third generation SOA
320
product of toluene photooxidation, based on the knowledge of gas-phase mechanisms and the
321
structural elucidation of the mass spectral fragmentation patterns.33 These facts motivated our
322
further evaluation of theoretical predictions through a detailed chemical analysis of MNC
323
compounds in ambient PM.
whereas an effort to quantify specific MNC isomers has been only made by Kitanovski et
324 325
Identification of methylnitrocatechols in atmospheric particulate matter
326
Identification of MNC isomers present in water extracts of aerosol samples from Ljubljana,
327
Slovenia was performed by LC-(-)ESI-MS/MS. The typical SRM chromatogram of MNCs
328
(m/z 168 → m/z 138) in the field sample is given in Figure 5a.
329
15 ACS Paragon Plus Environment
Environmental Science & Technology
330 331 332
Figure 5. SRM chromatograms of (a) ambient PM sample, Ljubljana, Slovenia (collected 14–
333
25/1/2015), (b) 30 µg L-1 reference mixture of 4-methyl-5-nitrocatechol (4M5NC), 3-methyl-
334
6-nitrocatechol (3M6NC) and 3-methyl-5-nitrocatechol (3M5NC), and (c) reaction mixture
335
after 3-methylcatechol (3MC) nitration as well as MS/MS data of m/z 168 compounds in (d)
336
ambient PM sample peak eluted at 20.1 min, (e) 30 µg L-1 3M6NC reference compound
337
eluted at 20.6 min and (f) minor peak in nitration reaction mixture eluted at 20.1 min.
338 339
The corresponding fragmentation patterns, obtained from aerosol MS/MS experiment, of the
340
minor (retention time, RT 20.1 min) and two major (RT 15.7 and 23.1 min) m/z 168 peaks are
341
presented in Figures 5d and S1a,b, respectively. In order to structurally distinguish between
342
the three distinct peaks attributed to positional isomers of MNCs in the field samples, their
16 ACS Paragon Plus Environment
Page 16 of 34
Page 17 of 34
Environmental Science & Technology
343
chromatographic behaviors and fragmentation patterns were first compared to commercially
344
available reference compounds 4M5NC, 3M6NC, and 3M5NC.
345
The SRM chromatogram of the reference compounds is presented in Figure 5b. Only the
346
corresponding fragmentation pattern of 3M6NC is shown in Figure 5e because fragmentations
347
of reference compounds 4M5NC and 3M5NC agree well with published data.20,44 The two
348
major peaks in the field samples were assigned to 4M5NC (RT 15.7 min, Figure S1a) and
349
3M5NC (RT 23.1 min, Figure S1b) isomers. As it has also been previously documented, their
350
MS/MS product ion spectra differ only in relative abundances of ions corresponding to the
351
neutral losses of NO (m/z 138) and HNO (m/z 137), the loss of NO2 (m/z 122) as well as the
352
combined loss of NO and CO (m/z 109).20,44 However, the retention time of 3M6NC reference
353
compound (RT 20.6 min) did not satisfactory match with the minor peak occurring in the field
354
samples at 20.1 min. Furthermore, the fragmentation patterns were distinctly different as well,
355
indicating the presence of another, closely eluting, isobaric MNC in the atmospheric aerosols.
356
3M6NC has a quite similar fragmentation pattern (Figure 5e) as the other two investigated
357
positional isomers, containing all of the fragments also found in the fragmentation patterns of
358
4M5NC and 3M5NC (vide supra). However, two additional peaks at m/z 151 and 150,
359
corresponding to neutral losses of •OH and H2O, respectively, can be found in the
360
fragmentation pattern of 3M6NC. All of the above agrees with the fact that the compounds
361
are closely related, whereas the latter difference can be attributed to the ortho effect, i.e. the
362
nitro group in 3M6NC is adjacent to one of the hydroxyl groups.50 However, MS/MS
363
spectrum of the compound eluting at 20.1 min in the field sample extracts (Figure 5d) shows
364
no peaks for neutral losses of •OH and H2O, only fragments with different relative
365
abundances, seen also in 4M5NC and 3M5NC fragmentation patterns (Figures S1a,b).
366
Therefore, it can be certainly concluded that there is no 3M6NC in the field samples, but
367
rather another possible MNC species. Noteworthy, in previous studies of ambient MNCs,
17 ACS Paragon Plus Environment
Environmental Science & Technology
368
chromatographic peaks have been assigned to 4M5NC, 3M5NC, and 3M6NC. According to
369
Iinuma et al.,20 the presence of 3M6NC was postulated based on the comparison of field
370
sample LC-MS characteristics with those of synthesized reaction products isolated after the
371
nitration of 3MC. Although Kitanovski et al.44 unambiguously identified the major reaction
372
product of the same nitration reaction as 3M5NC, by combining NMR and LC-MS
373
techniques, the identification of the minor reaction product is still lacking detailed analysis.
374
Figure 5c illustrates SRM chromatogram of a reaction mixture after the similar nitration
375
procedure as mentioned above (see Experimental section). The synthesized compounds are
376
eluting at 20.1 and 23.1 min, agreeing with the field samples. The fragmentation of the major
377
peak (RT 23.1 min) is shown in Figure S1c and corresponds to the already identified 3M5NC
378
isomer,44 while the retention time and the fragmentation pattern of the peak corresponding to
379
the minor reaction product (Figure 5f) agree with those of the yet unidentified peak in the
380
ambient samples eluting at 20.1 min and not to the 3M6NC standard. Therefrom it follows
381
that the minor reaction product is not 3M6NC as it has previously been thought, but can be,
382
taking into account its synthesis from 3MC, only assigned to 3M4NC isomer.
383
To unambiguously identify the unknown MNC isomer of our concern, the minor product was
384
isolated from the reaction mixture and structurally characterized by HRMS (deprotonated
385
molecule: [C7H6NO4]- exact mass measured: 168.0300 Da; exact mass calculated: 168.0297
386
Da; ppm 1.8) and 1H-NMR, 13C-NMR, and 2D correlation (gHSQC and gHMBC) NMR. The
387
combination of 1D- and 2D-NMR experiments unequivocally confirmed that the exact
388
molecular structure of the isolated compound corresponds to 3M4NC (for details please refer
389
to SI material). All the resonances, multiplicities, and coupling constants in 1H-NMR (300
390
MHz) and
391
(Table S3). In addition, excellent agreement with experimental values was observed for
392
calculated 13C-NMR resonances. The structural assignment of 3M4NC was finally supported
13
C-NMR (75 MHz) spectra were in accordance with the structure of 3M4NC
18 ACS Paragon Plus Environment
Page 18 of 34
Page 19 of 34
Environmental Science & Technology
393
by gHSQC and gHMBC 2D correlation spectroscopy (1H-13C), see Figures S4, S5 and S6.
394
Thus, precise structural identification of 3M4NC isomer, synthesized as a minor nitration
395
product of 3MC, together with its chromatographic and fragmentation properties described
396
above unambiguously confirm the presence of 3M4NC in the atmospheric aerosols, which has
397
never been done before.
398 399
Quantification of methylnitrocatechols in atmospheric particulate matter
400
Atmospheric concentrations of 4M5NC, 3M4NC, and 3M5NC in urban winter aerosols were
401
determined by LC-MS analysis of ambient PM2 samples using commercially available
402
authentic standards (4M5NC and 3M5NC) and de novo synthesized 3M4NC standard
403
compound by means of an internal calibration method and are shown in Table S4. The
404
concentrations of 3M5NC and 3M4NC ranged from 10.2 to 43.2 ng m-3 and from 2.0 to 8.2
405
ng m-3, respectively. It should be noted that 3M6NC was not detected in any of the PM2
406
samples although the ESI source was also exchanged by an atmospheric pressure chemical
407
ionization interface, which resulted in lower LOD of 3M6NC. In addition, we quantified
408
4M5NC isomer despite its formation pathway is different from those of 3M5NC, 3M4NC and
409
3M6NC. Namely, in contrast to the studied isomers which contain the methyl group attached
410
to the position 3-, 4M5NC compound has the methyl group attached to the position 4- of the
411
catechol ring. Thus, the formation mechanism of the 4M5NC should include analogue
412
nitration of 4-methylcatechol (4MC). However, beside 3M5NC and 3M6NC, 4M5NC
413
concentrations considerably contributed to the overall MNCs concentration in our samples.
414
Thus, during the sampling period, the average total concentration of MNCs in ambient PM2
415
was 43 ng m-3, contributing on average 0.15% (range: 0.11–0.21%) to the PM2 mass. For
416
comparison, Iinuma et al.20 determined on average 5.2 ng m-3 MNCs in wintertime PM10 from
417
a rural village Seiffen, Germany, which is for an order of magnitude lower than the values
19 ACS Paragon Plus Environment
Environmental Science & Technology
418
obtained within this work. On the other hand, Kitanovski et al.22 measured a comparable
419
amount of MNCs in winter PM10 at the same sampling site of Ljubljana (69.2 ng m-3).
420
We next assessed the magnitude of error one could make by the incorrect assignment of
421
3M6NC instead of 3M4NC in an aerosol sample. If authentic 3M6NC standard compound
422
was used for quantification of 3M4NC peak by LC-MS, about two orders of magnitude higher
423
concentration would be determined, introducing also a significant error (of about the factor of
424
10) to the total ambient MNCs concentration. 3M6NC obviously ionizes poorly under ESI
425
conditions leading to the significantly lower response with respect to the other considered
426
isomers. The same can be deduced from the obtained LODs, being 0.15 mg L-1 for 4M5NC,
427
3M5NC, and 3M4NC and 40 mg L-1 for 3M6NC. As MNCs, along with 4NC, are the major
428
nitroaromatic compounds in ambient aerosols, contributing together more that 95% to the
429
total nitroaromatic mass,22 the substantial error due to the incorrect MNCs identification
430
would further lead to the overestimation of the overall nitroaromatic contribution to the
431
ambient PM and OC. Calculated for Ljubljana winter samples, contribution of total MNCs to
432
PM2 would increase to l.2% (from 0.15%). Assuming the 40% fraction of organic material in
433
total aerosol mass as relevant for this study,22,51 the actual 0.4% ratio of OC would be grossly
434
overestimated as >3%, if we used the value from the incorrect quantification of MNC. As
435
reported for several urban sites, the latter value is in the range of (di)carboxylic acids in
436
ambient PM,52,53 which are well known as major SOA components formed in both the gas and
437
aerosol phases.54-56
438 439
Environmental significance
440
We theoretically predicted the possible reaction products of the electrophilic nitration of 3MC
441
in aqueous solution and further assessed the relevance of this reaction pathway for the
442
formation of atmospheric MNCs by comparing the theoretical calculations with field
20 ACS Paragon Plus Environment
Page 20 of 34
Page 21 of 34
Environmental Science & Technology
443
measurements. According to quantum chemical calculations, energetically favorable
444
pathways of electrophilic substitution of 3MC by NO2+ lead to the preferential formation of
445
3M5NC and 3M4NC isomers, whereas negligible amounts of 3M6NC are expected to form.
446
After a detailed reassessment of MNCs in ambient aerosols, both 3M5NC and 3M4NC were
447
determined in winter PM2 from Ljubljana, Slovenia; at the same time, no 3M6NC could be
448
detected in these samples. The presented theoretical calculations anticipate the ratio of
449
3M5NC to 3M4NC according to the electrophilic formation pathway in favor of 3M5NC (the
450
ratio 1:2.4 was obtained and should be considered accurate within one order of magnitude).
451
On average, concentrations 5.6 ± 0.8 times higher of 3M5NC to 3M4NC were observed in
452
ambient aerosols, which agrees well with the theoretical calculations. Although the predicted
453
kinetics of aqueous-phase formation of MNCs cannot point directly to the prevailing
454
electrophilic nitration of 3MC in the atmospheric wet aerosols, a very good correlation (R2 =
455
0.99) obtained between 3M5NC and 3M4NC concentrations (Figure 6) indicates their
456
common formation pathway(s) and supports this assumption.
457
458 459
Figure 6. Correlation between 3M5NC and 3M4NC concentrations determined for PM2
460
samples (n=7) collected during winter 2014/2015 from Ljubljana, Slovenia.
21 ACS Paragon Plus Environment
Environmental Science & Technology
461 462
Aqueous-phase nitration of 3MC by nitronium ion is corroborated as a formation pathway of
463
MNCs in the wet atmospheric aerosols as (1) no 3M6NC could be detected in any of the
464
ambient samples and (2) the fixed ratio of 3M5NC to 3M4NC agrees well with theoretical
465
predictions and indicates their common formation pathway. A potential discrepancy between
466
the theoretical predictions and the field measurements could be explained by distinct vapor
467
pressures and/or solubilities of the isomers under consideration, which would affect their
468
Henry's law constants and consequently their partitioning between the atmospheric aqueous
469
and gaseous phases. Unfortunately, none of these data (neither vapor pressures or solubilities
470
nor Henry's law constants) are available for the investigated compounds to support our
471
explanation. As estimation methods have already been recognized as one of the major
472
contributors to introducing errors into predictive atmospheric models, they will be avoided.24
473
Still, formation pathways and the multiphase partitioning are not the sole factor to be
474
discussed when considering the ratios between organics in the atmospheric samples; losses of
475
organic compounds through their ageing and degradation processes, influenced by their
476
individual chemical stabilities, are also governing the chemical equilibrium of the
477
atmosphere. However, this study cannot rule out any of other possible condensed- or gas-
478
phase MNC formation pathways, proceeding in the dark or being photochemically induced
479
and involving different reactive species (e.g., •NO3, •NO2, HOONO, and •OH; organic radicals
480
or excited triplet states; H2O2; and metal ions) that could result in different product
481
yields.20,28,57–59
482
A review paper by Noziere et al.4 comprehensively discussed and highlighted the importance
483
of maximum level identification of organic compounds in the atmosphere. Accurate
484
identification is needed to reveal correct sources and precursors, to properly represent
485
formation pathways of specific SOA tracers in atmospheric models, and to elucidate
22 ACS Paragon Plus Environment
Page 22 of 34
Page 23 of 34
Environmental Science & Technology
486
corresponding climate and health effects.4 Moreover, the position of aromatic substitution was
487
found to be crucial for controlling not only the formation but also the losses of MNCs in the
488
aerosols.33 Such findings underscore the relevance of studying SOA formation and evolution
489
on a molecular level and the importance of identifying tracer organic compounds of interest.
490
Semi-volatile MNCs are contributing to the water-soluble organic fraction of atmospheric
491
PM; they are capable of forming SOA and regarded as tracers of processed biomass-burning
492
emissions as well as aged anthropogenic aerosols.20,33 They are prone to inappropriate
493
analysis by the most commonly used analytical techniques, which could introduce a
494
significant misconception about their sources, atmospheric chemistry leading to their
495
formation and aging, and/or their potential hazardous impact on the environment. We
496
evaluated the eventual errors in quantification of particular MNCs and consequently in
497
calculating NACs contributions to the OC and PM due to the incorrect identification of
498
3M4NC as 3M6NC. Moreover, MNCs rank among the most abundant chromophoric
499
substances in the atmospheric aerosols.19 As relative positions of –OH and –NO2 functional
500
groups on the aromatic ring in particular affect the absorbance of (methyl)nitrocatechols,60
501
incorrect MNCs quantification could also lead to incorrect estimations of their contribution to
502
the overall absorption of water soluble BrC. As already pointed out by Noziere et al.,4
503
speculative or careless analytics resulting in misidentification of chemical species in
504
atmospheric samples could lead to misunderstanding of basic environmental processes and
505
thus inappropriate evaluation of state-of-the-art predictive models, which could consequently
506
introduce important fundamental errors to beliefs of the scientific community. Bearing the
507
survey of published data on MNCs and discussion above in mind, we find it of utmost
508
importance to consider the up-to-date literature regarding MNCs notation with special care.
509 510
Corresponding Author
23 ACS Paragon Plus Environment
Environmental Science & Technology
511
*Address: Analytical Chemistry Laboratory, National Institute of Chemistry, Hajdrihova 19,
512
SI-1001 Ljubljana, Slovenia
513
Division for Marine and Environmental Research, Ruđer Bošković Institute, Bijenička 54,
514
PO-10000 Zagreb, Croatia
515
Phone: +386 (1) 4760 361; Fax: +386 (1) 4760 300; E-mail:
[email protected] 516 517
Acknowledgments
518
Financial support provided from the European Commission and the Croatian Ministry of
519
Science, Education and Sports through Marie Curie FP7-PEOPLE-2011-COFUND project
520
NEWFELPRO (author SF), and from the Slovenian Research Agency (Contract Nos. P1-0034
521
and P2-0152; authors SF, MŠ, AK, MH, IG), is gratefully acknowledged.
522 523
Supporting Information
524
Detailed procedures of the ambient aerosol collection and pretreatment; synthesis and
525
isolation of the 3MC nitration product; NMR characterization of the purified reaction product,
526
4 Tables and 6 Figures. This information is available free of charge via the Internet at
527
http://pubs.acs.org/.
528 529
References:
530
1.
531
atmospheric absorber of solar radiation? Atmos. Chem. Phys. 2013, 13, (17), 8607–8621.
532
2.
533
of organic aerosol in the atmosphere and on snow: Effects of SOA and brown carbon. J.
534
Geophys. Res.-Atmospheres 2014, 119, (12), 7453–7476.
Feng, Y.; Ramanathan, V.; Kotamarthi, V. R., Brown carbon: a significant
Lin, G.; Penner, J. E.; Flanner, M. G.; Sillman, S.; Xu, L.; Zhou, C., Radiative forcing
24 ACS Paragon Plus Environment
Page 24 of 34
Page 25 of 34
Environmental Science & Technology
535
3.
Laskin, A.; Laskin, J.; Nizkorodov, S. A., Chemistry of Atmospheric Brown Carbon.
536
Chem. Rev. 2015, 115, (10), 4335-4382.
537
4.
538
Glasius, M.; Grgić, I.; Hamilton, J. F.; Hoffmann, T.; Iinuma, Y.; Jaoui, M.; Kahno, A.;
539
Kampf, C. J.; Kourtchev, I.; Maenhaut, W.; Marsden, N.; Saarikoski, S.; Schnelle-Kreis, J.;
540
Surratt, J. D.; Szidat, S.; Szmigielski, R.; Wisthaler, A., The Molecular Identification of
541
Organic Compounds in the Atmosphere: State of the Art and Challenges. Chem. Rev. 2015,
542
115, (10), 3919-3983.
543
5.
544
Winther, J-G. Introduction. In Climate Change 2013: The Physical Science Basis.
545
Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental
546
Panel on Climate Change, Cambridge University Press: Cambridge, United Kingdom and
547
New York, USA, 2013.
548
6.
549
Andreae, M. O., Optical properties of humic-like substances (HULIS) in biomass-burning
550
aerosols. Atmos. Chem. Phys. 2006, 6, 3563–3570.
551
7.
552
Soluble Organic Aerosol material and the light-absorption characteristics of aqueous extracts
553
measured over the Southeastern United States. Atmos. Chem. Phys. 2010, 10, (13), 5965–
554
5977.
555
8.
556
matter isolated from atmospheric aerosols collected under different atmospheric conditions.
557
Anal. Chim. Acta 2005, 530, (1), 7–14.
Noziere, B.; Kaberer, M.; Claeys, M.; Allan, J.; D'Anna, B.; Decesari, S.; Finessi, E.;
Cubasch, U.; Wuebbles, D.; Chen, D.; Facchini, M. C.; Frame, D.; Mahowald, N.;
Hoffer, A.; Gelencser, A.; Guyon, P.; Kiss, G.; Schmid, O.; Frank, G. P.; Artaxo, P.;
Hecobian, A.; Zhang, X.; Zheng, M.; Frank, N.; Edgerton, E. S.; Weber, R. J., Water-
Duarte, R.; Pio, C. A.; Duarte, A. C., Spectroscopic study of the water-soluble organic
25 ACS Paragon Plus Environment
Environmental Science & Technology
558
9.
Shapiro, E. L.; Szprengiel, J.; Sareen, N.; Jen, C. N.; Giordano, M. R.; McNeill, V. F.,
559
Light-absorbing secondary organic material formed by glyoxal in aqueous aerosol mimics.
560
Atmos. Chem. Phys. 2009, 9, (7), 2289–2300.
561
10.
562
Nizkorodov, S. A., Formation of nitrogen- and sulfur-containing light-absorbing compounds
563
accelerated by evaporation of water from secondary organic aerosols. J. Geophys. Res.-
564
Atmospheres 2012, 117.
565
11.
566
H.; Donahue, N. M.; Robinson, A. L., Absorptivity of brown carbon in fresh and photo-
567
chemically aged biomass-burning emissions. Atmos. Chem. Phys. 2013, 13, (15), 7683–7693.
568
12.
569
absorbing soluble organic aerosol in Los Angeles and Atlanta: A contrast in secondary
570
organic aerosol. Geophys. Res. Lett. 2011, 38.
571
13.
572
photooxidation of aromatic hydrocarbons: Molecular composition. Environ. Sci. Technol.
573
1997, 31, (5), 1345–1358.
574
14.
575
Laboratory studies on optical properties of secondary organic aerosols generated during the
576
photooxidation of toluene and the ozonolysis of alpha-pinene. Atmos. Chem. Phys. 2010, 115.
577
15.
578
irradiation of aqueous solutions containing H2O2 and phenolic compounds. Atmos. Environ.
579
2010, 44, (4), 541–551.
580
16.
581
Physico-chemical characterization of SOA derived from catechol and guaiacol - a model
Nguyen, T. B.; Lee, P. B.; Updyke, K. M.; Bones, D. L.; Laskin, J.; Laskin, A.;
Saleh, R.; Hennigan, C. J.; McMeeking, G. R.; Chuang, W. K.; Robinson, E. S.; Coe,
Zhang, X.; Lin, Y.-H.; Surratt, J. D.; Zotter, P.; Prevot, A. S. H.; Weber, R. J., Light-
Forstner, H. J. L.; Flagan, R. C.; Seinfeld, J. H., Secondary organic aerosol from the
Nakayama, T.; Matsumi, Y.; Sato, K.; Imamura, T.; Yamazaki, A.; Uchiyama, A.,
Chang, J. L.; Thompson, J. E., Characterization of colored products formed during
Ofner, J.; Krueger, H. U.; Grothe, H.; Schmitt-Kopplin, P.; Whitmore, K.; Zetzsch, C.,
26 ACS Paragon Plus Environment
Page 26 of 34
Page 27 of 34
Environmental Science & Technology
582
substance for the aromatic fraction of atmospheric HULIS. Atmos. Chem. Phys. 2011, 11, (1),
583
1–15.
584
17.
585
Toxicol. 2000, 13, (8), 673–692.
586
18.
587
Surratt, J. D.; Seinfeld, J. H., Formation of secondary organic aerosol from irradiated alpha-
588
pinene/toluene/NOx mixtures and the effect of isoprene and sulfur dioxide. J. Geophys. Res.-
589
Atmospheres 2008, 113, (D9).
590
19.
591
W.; Meszaros, T.; Salma, I., Chemical characterisation of humic-like substances from urban,
592
rural and tropical biomass burning environments using liquid chromatography with UV/vis
593
photodiode array detection and electrospray ionisation mass spectrometry. Environ. Chem.
594
2012, 9, (3), 273–284.
595
20.
596
Tracer Compounds for Biomass Burning Secondary Organic Aerosols. Environ. Sci. Technol.
597
2010, 44, (22), 8453–8459.
598
21.
599
Absorption Angstrom Exponent of Light-absorbing Organic Components in Aerosol Extracts
600
from the Los Angeles Basin. Environ. Sci. Technol. 2013, 47, (8), 3685–3693.
601
22.
602
chromatography tandem mass spectrometry method for characterization of monoaromatic
603
nitro-compounds in atmospheric particulate matter. J. Chrom. A 2012, 1268, 35–43.
604
23.
605
E.; Claeys, M.; Maenhaut, W., One-year study of nitro-organic compounds and their relation
Purohit, V.; Basu, A. K., Mutagenicity of nitroaromatic compounds. Chem. Res.
Jaoui, M.; Edney, E. O.; Kleindienst, T. E.; Lewandowski, M.; Offenberg, J. H.;
Claeys, M.; Vermeylen, R.; Yasmeen, F.; Gomez-Gonzalez, Y.; Chi, X.; Maenhaut,
Iinuma, Y.; Boege, O.; Graefe, R.; Herrmann, H., Methyl-Nitrocatechols: Atmospheric
Zhang, X.; Lin, Y.-H.; Surratt, J. D.; Weber, R. J., Sources, Composition and
Kitanovski, Z.; Grgić, I.; Vermeylen, R.; Claeys, M.; Maenhaut, W., Liquid
Kahnt, A.; Behrouzi, S.; Vermeylen, R.; Shalamzari, M. S.; Vercauteren, J.; Roekens,
27 ACS Paragon Plus Environment
Environmental Science & Technology
606
to wood burning in PM10 aerosol from a rural site in Belgium. Atmos. Environ. 2013, 81,
607
561–568.
608
24.
609
Auld, J., Aerosol speciation and mass prediction from toluene oxidation under high NOx
610
conditions. Atmos. Environ. 2010, 44, (3), 361–369.
611
25.
612
monolayers observed by Brewster angle microscopy and canal surface viscometry. Polym. J.
613
1996, 28, (9), 784–789.
614
26.
615
Dommen, J.; Donahue, N. M.; George, C.; Goldstein, A. H.; Hamilton, J. F.; Herrmann, H.;
616
Hoffmann, T.; Iinuma, Y.; Jang, M.; Jenkin, M. E.; Jimenez, J. L.; Kiendler-Scharr, A.;
617
Maenhaut, W.; McFiggans, G.; Mentel, T. F.; Monod, A.; Prevot, A. S. H.; Seinfeld, J. H.;
618
Surratt, J. D.; Szmigielski, R.; Wildt, J., The formation, properties and impact of secondary
619
organic aerosol: current and emerging issues. Atmos. Chem. Phys. 2009, 9, (14), 5155–5236.
620
27.
621
University Press: Ithaca, NY, 1953.
622
28.
623
environmental chemistry of nitrite: Nitration of phenol upon nitrite photoinduced oxidation.
624
Environ. Sci. Technol. 2002, 36, (4), 669–676.
625
29.
626
processes in the environmental chemistry of nitrite. 2. The role of hydrogen peroxide.
627
Environ. Sci. Technol. 2003, 37, (20), 4635–4641.
628
30.
629
hydroxylation of benzene in the presence of nitrite/nitrous acid in aqueous solution.
630
Chemosphere 2004, 56, (11), 1049–1059.
Kelly, J. L.; Michelangeli, D. V.; Makar, P. A.; Hastie, D. R.; Mozurkewich, M.;
Sato, N.; Ito, S.; Yamamoto, M., Interfacial behavior of acetalized poly(vinyl alcohol)
Hallquist, M.; Wenger, J. C.; Baltensperger, U.; Rudich, Y.; Simpson, D.; Claeys, M.;
Ingold, C. K., Chapter 6. In Structure and Mechanism in Organic Chemistry, Cornell
Vione, D.; Maurino, V.; Minero, C.; Pelizzetti, E., New processes in the
Vione, D.; Maurino, V.; Minero, C.; Borghesi, D.; Lucchiari, M.; Pelizzetti, E., New
Vione, D.; Maurino, V.; Minero, C.; Lucchiari, M.; Pelizzetti, E., Nitration and
28 ACS Paragon Plus Environment
Page 28 of 34
Page 29 of 34
Environmental Science & Technology
631
31.
Beake, B. D.; Constantine, J.; Moodie, R. B., Nitration and oxidation of 4-
632
methoxyphenol by nitrous-acid in aqueous acid-solution. J. Chem. Soc. Perk T 2 1994, (2),
633
335–340.
634
32.
635
remote atmospheric conditions? Sci. Rep. 2015, 5.
636
33.
637
M.; Ward, M. W.; Wyche, K. P.; Munoz, A.; Vera, T.; Vazquez, M.; Borras, E.; Rodenas, M.,
638
Insights into the Formation and Evolution of Individual Compounds in the Particulate Phase
639
during Aromatic Photo-Oxidation. Environ. Sci. Technol. 2015, 49, (22), 13168–13178.
640
34.
641
Academic Press: Burlington, MA, 2006.
642
35.
643
S., Experimental and Theoretical Study of Aqueous cis-Pinonic Acid Photolysis. J. Phys.
644
Chem. A 2013, 117, (48), 12930–12945.
645
36.
646
Tropospheric Oxidation of Vinyl Acetate Initiated by OH Radical: A Theoretical Study. J.
647
Phys. Chem. A 2013, 117, (18), 3739–3750.
648
37.
649
R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li,
650
X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.;
651
Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao,
652
O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.;
653
Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.;
654
Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.;
655
Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.;
Kroflič, A.; Grilc, M.; Grgić, I., Does toxicity of aromatic pollutants increase under
Pereira, K. L.; Hamilton, J. F.; Rickard, A. R.; Bloss, W. J.; Alam, M. S.; Camredon,
Wallace, J.; Hobbs, P., Atmospheric Science: An Introductory Survey. 2nd ed.;
Lignell, H.; Epstein, S. A.; Marvin, M. R.; Shemesh, D.; Gerber, R. B.; Nizkorodov,
Mandal, D.; Sahu, C.; Bagchi, S.; Das, A. K., Kinetics and Mechanism of the
Frisch, M. J. T., G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J.
29 ACS Paragon Plus Environment
Environmental Science & Technology
656
Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski,
657
J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.;
658
Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.;
659
Cioslowski, J.; Fox, D. J. Gaussian, Gaussian 09, Revision E.01, Gaussian, Inc: Wallingford
660
CT, 2009.
661
38.
662
Chem. Phys. Lett. 1988, 153, (6), 503-–506.
663
39.
664
of electron correlation. Chem. Phys. Lett. 1989, 154, (1), 83–89.
665
40.
666
Phys. Lett. 1990, 166, (3), 275–280.
667
41.
668
and gradient. Chem. Phys. Lett. 1990, 166, (3), 281–289.
669
42.
670
- Theory and application to bifurcated hydrogen-bonds in the water hexamer. Chem. Phys.
671
Lett. 1994, 220, (1-2), 122–128.
672
43.
673
Rasul, G.; Prakash, G. K. S.; Olah, G. A., Unified mechanistic concept of electrophilic
674
aromatic nitration: Convergence of computational results and experimental data. J. Am.
675
Chem. Soc. 2003, 125, (16), 4836–4849.
676
44.
677
chromatographic method based on ultraviolet-visible and electrospray ionization mass
678
spectrometric detection for the identification of nitrocatechols and related tracers in biomass
679
burning atmospheric organic aerosol. Rap. Comm. Mass Spec. 2012, 26, (7), 793–804.
680
45.
Headgordon, M.; Pople, J. A.; Frisch, M. J., MP2 energy evaluation by direct methods
Saebo, S.; Almlof, J., Avoiding the integral storage bottleneck in LCAO calculations
Frisch, M. J.; Headgordon, M.; Pople, J. A., A direct MP2 gradient - method. Chem.
Frisch, M. J.; Headgordon, M.; Pople, J. A., Semidirect algorithms for the MP2 energy
Headgordon, M.; Headgordon, T., Analytic MP2 frequencies without 5th-order storage
Esteves, P. M.; Carneiro, J. W. D.; Cardoso, S. P.; Barbosa, A. G. H.; Laali, K. K.;
Kitanovski, Z.; Grgić, I.; Yasmeen, F.; Claeys, M.; Čusak, A., Development of a liquid
Kenner, J., Oxidation and Reduction in Chemistry. Nature 1945, 156, 369–370.
30 ACS Paragon Plus Environment
Page 30 of 34
Page 31 of 34
Environmental Science & Technology
681
46.
Weiss, J., Simple electron transfer processes in systems of conjugated double bonds.
682
Trans. Faraday Soc. 1946, 42, 116–121.
683
47.
684
Nitronium Tetrafluoroborate Nitration of Alkylbenzenes in Tetramethylene Sulfone Solution.
685
Remarks on Certain Aspects of Electrophilic Aromatic Substitution 2. J. Am. Chem. Soc.
686
1961, 83, (22), 4571–4580.
687
48.
688
of benzene with nitronium ion. J.Phys. Chem. A 2003, 107, (51), 11440–11444.
689
49.
690
Atmospheric Waters: Nitrite, A Source of Reactive Nitronium Ion in the Atmosphere.
691
Environ. Sci. Technol. 2015, 49, (15), 9150–9158.
692
50.
693
Science: Mill Valley, 1980.
694
51.
695
C.; Van Dingenen, R.; Ervens, B.; Nenes, A.; Nielsen, C. J.; Swietlicki, E.; Putaud, J. P.;
696
Balkanski, Y.; Fuzzi, S.; Horth, J.; Moortgat, G. K.; Winterhalter, R.; Myhre, C. E. L.;
697
Tsigaridis, K.; Vignati, E.; Stephanou, E. G.; Wilson, J., Organic aerosol and global climate
698
modelling: a review. Atmos. Chem. Phys. 2005, 5, 1053–1123.
699
52.
700
acids and water-soluble organic carbon in aerosols in New Delhi, India, in winter:
701
Characteristics and formation processes. J. Geophys. Res.-Atmospheres 2009, 114.
702
53.
703
soluble organic acids in PM2.5 in Beijing, China. Atmos. Environ. 2005, 39, (16), 2819–2827.
Olah, G. A.; Kuhn, S. J.; Flood, S. H., Aromatic Substitution. VIII.1 Mechanism of the
Chen, L. T.; Xiao, H. M.; Xiao, J. J.; Gong, X. D., DFT study on nitration mechanism
Kroflič, A.; Grilc, M.; Grgić, I., Unraveling Pathways of Guaiacol Nitration in
McLaferty, F., Chapter 9. In Interpretation of Mass Spectra, 3rd ed.; University
Kanakidou, M.; Seinfeld, J. H.; Pandis, S. N.; Barnes, I.; Dentener, F. J.; Facchini, M.
Miyazaki, Y.; Aggarwal, S. G.; Singh, K.; Gupta, P. K.; Kawamura, K., Dicarboxylic
Huang, X. F.; Hu, M.; He, L. Y.; Tang, X. Y., Chemical characterization of water-
31 ACS Paragon Plus Environment
Environmental Science & Technology
704
54.
Tolocka, M. P.; Jang, M.; Ginter, J. M.; Cox, F. J.; Kamens, R. M.; Johnston, M. V.,
705
Formation of oligomers in secondary organic aerosol. Environ. Sci. Technol. 2004, 38, (5),
706
1428–1434.
707
55.
708
Fisseha, R.; Weingartner, E.; Frankevich, V.; Zenobi, R.; Baltensperger, U., Identification of
709
polymers as major components of atmospheric organic aerosols. Science 2004, 303, (5664),
710
1659–1662.
711
56.
712
Weingartner, E.; Baltensperger, U., Identification of organic acids in secondary organic
713
aerosol and the corresponding gas phase from chamber experiments. Anal. Chem. 2004, 76,
714
(22), 6535–6540.
715
57.
716
Arsenec, C., Photochemical reactions in the tropospheric aqueous phase and on particulate
717
matter. Chem. Soc. Rev. 2006, (35), 441–453.
718
58.
719
production from aqueous reactions of atmospheric phenols with an organic triplet excited
720
state. Environ. Sci. Technol. 2014, 48, (2), 1049–1057.
721
59.
722
presence of nitrous acid ionization equilibria. J. Hazard. Mater. 2010, 174, 801–806.
723
60.
724
characterization of brown carbon (BrC) chromophores in secondary organic aerosol generated
725
from photo-oxidation of toluene. Phys. Chem. Chem. Phys. 2015, 17, (36), 23312–23325.
Kalberer, M.; Paulsen, D.; Sax, M.; Steinbacher, M.; Dommen, J.; Prevot, A. S. H.;
Fisseha, R.; Dommen, J.; Sax, M.; Paulsen, D.; Kalberer, M.; Maurer, R.; Hofler, F.;
Vione, D.; Maurino, V.; Minero, C,; Pelizzetti, E.; Harrison, M. A. J.; Olariuc, R-I.;
Smith, J. D.; Sio, V.; Yu, L.; Zhang, Q.; Anastasio, C., Secondary organic aerosol
Khalafi, L.; Rafiee, M., Kinetic study of the oxidation and nitration of catechols in the
Lin, P.; Liu, J.; Shilling, J. E.; Kathmann, S. M.; Laskin, J.; Laskin, A., Molecular
32 ACS Paragon Plus Environment
Page 32 of 34
Page 33 of 34
Environmental Science & Technology
726 727
TOC/Abstract Art
33 ACS Paragon Plus Environment
Environmental Science & Technology Page 34 of 34
ACS Paragon Plus Environment