Next Article in Journal
An Evaluation of Relationships between Radar-Inferred Kinematic and Microphysical Parameters and Lightning Flash Rates in Alabama Storms
Next Article in Special Issue
Aqueous-Phase Brown Carbon Formation from Aromatic Precursors under Sunlight Conditions
Previous Article in Journal
Implicit Definition of Flow Patterns in Street Canyons—Recirculation Zone—Using Exploratory Quantitative and Qualitative Methods
Previous Article in Special Issue
Electrochemical Evidence of non-Volatile Reduced Sulfur Species in Water-Soluble Fraction of Fine Marine Aerosols
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Aqueous Reactions of Sulfate Radical-Anions with Nitrophenols in Atmospheric Context

by
Krzysztof J. Rudziński
* and
Rafał Szmigielski
Institute of Physical Chemistry of the PAS, Kasprzaka 44/52, 01-224 Warsaw, Poland
*
Author to whom correspondence should be addressed.
Atmosphere 2019, 10(12), 795; https://doi.org/10.3390/atmos10120795
Submission received: 14 November 2019 / Revised: 1 December 2019 / Accepted: 4 December 2019 / Published: 9 December 2019
(This article belongs to the Special Issue Atmospheric Aqueous-Phase Chemistry)

Abstract

:
Nitrophenols, hazardous environmental pollutants, react promptly with atmospheric oxidants such as hydroxyl or nitrate radicals. This work aimed to estimate how fast nitrophenols are removed from the atmosphere by the aqueous-phase reactions with sulfate radical-anions. The reversed-rates method was applied to determine the relative rate constants for reactions of 2-nitrophenol, 3-nitrophenol, 4-nitrophenol, 2,4-dinitrophenol, and 2,4,6-trinitrophenol with sulfate radical-anions generated by the autoxidation of sodium sulfite catalyzed by iron(III) cations at ~298 K. The constants determined were: 9.08 × 108, 1.72 × 109, 6.60 × 108, 2.86 × 108, and 7.10 × 107 M−1 s−1, respectively. These values correlated linearly with the sums of Brown substituent coefficients and with the relative strength of the O–H bond of the respective nitrophenols. Rough estimation showed that the gas-phase reactions of 2-nitrophenol with hydroxyl or nitrate radicals dominated over the aqueous-phase reaction with sulfate radical-anions in deliquescent aerosol and haze water. In clouds, rains, and haze water, the aqueous-phase reaction of 2-nitrophenol with sulfate radical-anions dominated, provided the concentration of the radical-anions was not smaller than that of the hydroxyl or nitrate radicals. The results presented may be also interesting for designers of advanced oxidation processes for the removal of nitrophenol.

Graphical Abstract

1. Introduction

Nitrophenols are well known environmental trace compounds and pollutants [1,2], which have been detected in various environmental matrices including air [3,4,5,6], rainwater [3,7,8,9], cloud water [10], fog [10,11], snow [1], atmospheric aerosol [4,5,6,12,13,14,15,16,17,18,19,20,21,22,23,24,25], soils [26,27], and surface waters [10,28,29,30,31,32]. They originate from many anthropogenic and natural sources including: the incineration of wastes [33], industrial chemical processes [34], combustion of coal and biomass as well as vehicle and aviation fuels [35,36,37], degradation of pesticides [34,38,39], release of wood preservatives [34], and atmospheric chemical reactions. 4-nitrophenol (4-NP) and 2,4-dinitrophenol (2,4-DNP), along with sugar anhydrides such as levoglucosan, serve as markers of biomass burning in ambient aerosol [40,41,42,43]. These compounds and 2-nitrophenol are recognized components of atmospheric brown carbon, i.e., a collection of light absorbing organic compounds in the atmosphere [22,44,45]
Atmospheric reactions that yield nitrophenols take place both in the gas phase and in the aqueous phase. For instance, the gas-phase nitration of phenol involves hydroxyl radicals OH and NO2 in the daytime or nitrate radical NO3 and NO2 in the night to respectively produce 2-nitrophenol (2-NP) or 2-NP and 4-NP. The chemical mechanisms of both processes were thoroughly reviewed [1]. Formation of nitrophenols in atmospheric waters of all kinds is at least equally important but less understood. The possible pathways include oxidation of phenols with NO2 and OH or NO3 radicals [46], electrophilic nitration initiated by N2O5 and ClNO2 [1,47], photolytic and dark reactions involving nitrate radicals, inorganic nitrates, nitrites and nitrous acid HONO [48,49], and photolytic reactions with nitrogen dioxide in the presence of iron oxide and oxygen [49].
Atmospheric sinks for nitrophenols include photolysis [50] and the gas-phase reactions with OH radicals and NO3 radicals [51,52] which are characterized by the estimated residence time of several days. More efficient sinks may include partitioning to atmospheric aqueous phases followed by reactions with various radicals and/or photolysis [1,46]. More recently, Barsotti, et al. [53] demonstrated that the irradiation of aqueous solutions or viscous films containing several nitrophenols (2-NP, 4-NP, 2,4-DNP, and 2,6-DNP, i.e., 2,6-dinitrophenol) was an efficient source of HONO and NO2 ions. Vione, et al. [54] showed that OH radicals reacted faster than NO3 radicals with 2-NP and 4-NP in aqueous solutions to lower the atmospheric levels of 2-NP below those of 4-NP. Hems and Abbatt [55] studied the aqueous-phase photo-oxidation of 2,4-DNP by OH radicals, identified numerous intermediate products thereof and showed the corresponding evolution of UV absorbance of the reacting solutions. In addition, many laboratories studied the aqueous-phase reactions of other substituted phenols of atmospheric interest like guaiacol, nitro-guaiacol, vanillin, or syringol [55,56,57,58,59].
Much of the nitrophenol chemistry has been studied for the sake of advanced oxidation processes aimed at mitigation of nitrophenols in aquatic and industrial environments [60]. The technologies considered include: Fenton and photo-Fenton reactions based on H2O2 [34,61], TiO2 based photocatalysis [62,63,64], electrocatalysis [65], photo-electrocatalysis [66,67], and wet catalysis [68,69]. Among the latter, a promising process was proposed which utilized reactions of nitrophenols with sulfate radical-anions generated by the cobalt-mediated decomposition of peroxymonosulfate anions [70].
For years, some nitrophenols (2-NP, 4-NP, 2,4-DNP) have been listed as priority or hazardous pollutants [71,72,73]. Generally, mono- and di-nitrophenols are considered toxic in plants and mammals [74], while 4-nitrophenol is highly toxic in humans [75]. Although EPA USA has not considered 4-nitrophenol carcinogenic [38], a laboratory experiment showed the compound can destroy DNA in vitro [76].
This work was aimed at elucidating how fast nitrophenols are removed from the atmospheric waters by reaction with sulfate radical-anions, which are important atmospheric oxidants known to react fast with numerous atmospheric pollutants [77,78,79,80,81,82,83].

2. Experiments

2.1. Chemicals

The following chemicals were used as purchased: 2-nitrophenol (R.G.), 3-nitrophenol (REAGENTPLUS™, 99%), 4-nitrophenol and 2,4,6-trinitrophenol (1 wt % solution in water) from Sigma Aldrich, 2,4-dinitrophenol (97% + 15% H2O) from Alfa Aesar, Fe(ClO4)3·9H2O (purum) from Fluka, Na2S2O5 (EMSURE® ACS, Reag. Ph Eur. > 98%) and HClO4 (pro analysis) from Merck, argon (99.999%) from Multax. For each experiment, aqueous solutions of reactants were prepared freshly using Milli-Q water (18.2 MΩ cm, Milli-Q Advantage System from Merck Millipore). Buffer standards used for the calibration of pH electrodes were from Thermo Fisher Scientific. To avoid the contact with the atmospheric oxygen, Milli-Q water was deoxygenated by a stream of argon bubbled through for 20 min. Solutions of sodium bisulfite were obtained by dissolving Na2S2O5 in deoxygenated Milli-Q water
Na 2 S 2 O 5 + H 2 O 2 Na + + 2 HSO 3 ,
The acidity of solutions was adjusted to pH = 3.1 with 0.1 M HClO4 so the species in solutions were predominantly Na+ and HSO3 ions.

2.2. Estimation of the Rate Constants

The relative rate constants for reactions of nitrophenols with sulfate radical-anions were estimated using the reversed-rates method developed by Ziajka and Pasiuk-Bronikowska [84] and successfully applied to several organic compounds [83,85,86]. Briefly, the sulfate radical-anions are generated during chain autoxidation of sulfite anions catalyzed by Fe(III) cations. The mechanism of the autoxidation was presented in detail by Ziajka and Rudziński [83] and recalled in the SI. One runs several experiments with the autoxidation of S(IV) inhibited by two different compounds, inh1 and inh2, used at several different initial concentrations (e.g., Figure 1). Each experiment should attain a pseudo-stationary phase during which the autoxidation proceeds at a constant rate (e.g., Figure 2). Then, one plots the reciprocal stationary rates observed against the initial concentrations of the inhibitor used (Figure 3). If both plots are linear, the ratio of rate constants for reactions of the inhibitors with sulfate radical-anions is equal to the ratio of the slopes of the linear plots (Equation (2)). If one knows the rate constant for reaction of one inhibitor with sulfate radical-anions, one can calculate the rate constant for the other inhibitor.
k inh 1 + SO 4 = slope inh 1 slope inh 2 k inh 2 + SO 4 ,
In the present work, ethanol was used as a reference inhibitor against which the rate constants for nitrophenols were calculated.

2.3. Experimental Runs

The experimental setup and procedure for carrying out the stationary autoxidation of S(VI) inhibited by organic compounds was described in detail elsewhere [83]. Briefly, the experiments were carried out in a well-mixed glass reactor of 60 cm3 volume, closed with a Teflon cover and thermostatted at 298 K within a water jacket. For each run, the reactor was filled with aqueous solution of sodium bisulfite and oxygen so that it contained no gas phase. The pH of solution was adjusted to 3.1 with HClO4. The pH of the solution was recorded using a SenTix Mic combination pH electrode from WTW. Then, a small aliquot of aqueous solution of Fe(ClO4)3 catalyst was injected to start the reaction. Table 1 shows the initial concentrations of reactants. The autoxidation of S(IV) was followed by recording the concentration of oxygen using an Orion 97-08 from Thermo Fisher Scientific and a home-designed pH/oxygen meter and software. Equation (3a) shows the overall stoichiometry of uninhibited SIV autoxidation. Assuming the conversion of inhibitors was small, the stoichiometry of the autoxidation inhibited by a nitrophenol was defined by the same equation so that the rate of the autoxidation was defined by Equation (3b) [83].
2SO32−/2HSO3 + O2 = SO42−/HSO4,
r a u t o x i d a t i o n = d [ S ( IV ) ] d t = d [ S ( VI ) ] d t = 2 d [ O 2 ] d t ,

2.4. Correction of the Diffusional Limitations of the Rate Constants

Since the reactions examined were very fast, we corrected the rate constants determined for diffusional limitations using a simple resistance-in-series model [87,88,89,90]
k o b s e r v e d 1 = k r e a c t i o n 1 + k d i f f u s i o n 1 ,
k d i f f u s i o n = 4 π ( D A + D B ) ( r A + r B ) N × 10 3 ,
where all k are second order rate constants (M−1 s−1), D are diffusion coefficients of reactants A and B (m2 s−1), r are reaction radii of reactant molecules A and B (m), and N is the Avogadro number (mol−1). Details of the calculations are summarized in Section S3 of the SI.

3. Results

In this section, we present the experimental results obtained for 4-NP. The results for other nitrophenols were similar so we present them in the SI. Figure 1 shows consumption of oxygen during the autoxidation of S(IV) in the presence of 4-NP. The higher was the initial concentration of nitrophenol, the slower was the consumption of O2.
In each experiment, the autoxidation attained a quasi-stationary rate, as shown in Figure 2 for the run with [4-NP] = 0.28 mM.
Figure 3 shows the plots of reciprocal stationary rates for autoxidation of S(IV) in the presence of 4-NP or a reference compound ethanol versus initial concentrations of each inhibitor. The plots were linear, so their slopes were used in Equation (2) to calculate the relative rate constant for the reaction of 4-NP with sulfate radical-anions
k 4 NP + SO 4 = k EtOH + SO 4 slope 4 NP slope EtOH = 4.3 × 10 7 2.665 × 10 9 1.735 × 10 8 = 6.636 × 10 8   M 1 s 1 ,
Plots for other nitrophenols, all of them linear, were placed in the SI. The results of all experiments are collected in Table 2 and include the slopes of linear plots and the rate constants for reactions of nitrophenols with sulfate radical-anions, both observed and corrected for diffusional limitations. The uncertainties of the observed rate constants were estimated using the total differential method applied to Equation (6) with individual errors equal to the standard errors of the linear slopes and kEtOH (Table 2). The uncertainties of the corrected rate constants were estimated in a similar way from Equation (4), assuming arbitrarily the uncertainty of kdifffusion was 10%.

4. Discussion

4.1. Hammett’s Correlations

The reactions of nitrophenols with sulfate radical-anions appeared quite fast, with observed second order rate constants of the order 109 s−1 M−1 (Table 2). However, the diffusional limitations were not significant because the rate constants corrected for diffusional limitations were higher for a few percent only. The rate constants decreased with the number of NO2 groups in the molecule with the exception of 3-NP. Figure 4a shows the uncorrected rate constants for reactions of sulfate radical-anions with phenol and substituted phenols—chlorophenols [83] and nitrophenols (this work)—correlate well with sums of Brown substituent coefficients for the compounds. The Brown coefficients for chlorophenols were taken after [83], while those for nitrophenols were: σm+ = 0.71, σp+ = 0.79 [91,92] and σo+ = 0.66 σp+ = 0.52 [93] for meta, para and orto substituents, respectively. The straight line in Figure 4a was obtained by linear regression covering all data (Equation (7)).
log ( k S O 4 ) = ( 9.9006 ± 0.0785 ) ( 1.1513 ± 0.0970 ) σ + ,           R 2 = 0.9663 ,
Equation (7) can be used to estimate the second order rate constants for reactions of sulfate radical-anions with substituted phenols that had not been determined experimentally.
The Brown substituent coefficients can estimate the relative strength of the O–H bonds in substituted phenols with a linear correlation developed by Jonsson et al. [93] (Equation (8)).
Δ D O H = 2 + 29.9 ( σ o 2 + + σ m 3 + + σ p 4 + + σ m 5 + + σ o 6 + ) ,           kJ   mol 1 ,
Therefore, the rate constants for reactions of sulfate radical-anions with phenol and substituted phenols can also be correlated against ∆DO–H (Figure 4b). The corresponding linear regression is given by Equation (9).
log ( k S O 4 ) = ( 9.9498 ± 0.0828 ) ( 0.0508 ± 0.0043 ) Δ D O H ,         R 2 = 0.9665 ,
Equation (9) can estimate the second order rate constants for reactions of sulfate radical-anions with substituted phenols in place of Equation (7).

4.2. Atmospheric Significance

The atmospheric significance of the aqueous-phase reactions of nitrophenols with sulfate radical-anions was evaluated using the approach developed in [81]. The rate of the total conversion of a nitrophenol NP by a reactant X (OH or NO3) in the gas phase (rX,g) and in the aqueous phase (rX,aq × ω) was compared to the rate of conversion of this NP by sulfate radical-anions in the aqueous phase within the gas phase (rSO4,aq ω) as the ratio RX,tot-aq defined by Equation (10). The concentrations of the gas-phase and aqueous-phase reactants were assumed to follow the Henry’s Law.
R X , t o t - a q = r X , g + r X , a q ω r S O 4 , a q ω = k X , g [ X ] g [ N P ] g + k X , a q [ X ] a q [ N P ] a q ω k S O 4 , a q [ X ] a q [ S O 4 · ] a q ω = k X , g H d , X H d , NP + k X , a q ω k SO 4 , a q ω · [ X ] a q [ S O 4 · ] a q ,
where ω m3 m−3 is the atmospheric liquid water contents; kX,g and kX,aq dm3 mol−1 s−1 are the rate constant for the reaction of X with NP in the gas phase and the aqueous phase, respectively; kSO4,aq dm3 mol−1 s−1 is the rate constant for the reaction of SO4•− with NP in the aqueous phase; Hd is the dimensionless Henry’s constant (Hd = H × gas constant × absolute temperature) for X or for NP; [X]aq and [SO4•−]aq are the aqueous-phase concentrations of X and SO4•−.
Figure 5 compares the total conversion of 2-NP due to the gas-phase and the aqueous-phase reactions with OH radicals to the aqueous-phase reaction with SO4•− radical-anions (a) as well as the conversion of 2-NP due to the gas-phase and aqueous-phase reactions with NO3 radicals to the aqueous-phase reaction with SO4•− radical-anions (b). Data required for the calculations behind the plots are given in Table 2 and Table 3 while more details are available in the SI. The ranges of radical concentrations considered in Figure 5 fit well within the realistic ranges estimated by modeling for the atmospheric systems (Table 4) [94].
In most cases, the total OH radical sink for 2-NP dominates over the SO4•− aqueous sink in all atmospheric aqueous phases. Sulfate radical-anions take the lead only if they are in significant excess in clouds, rains, and storms (red line in Figure 5a, [OH]aq/[SO4•−]aq < 0.16). The total NO3 radical sink for 2-NP dominates over the SO4•− aqueous sink in aerosol and haze waters and in clouds and rains if in significant excess (red line in Figure 5b, [NO3]aq/[ SO4•−]aq > 36). In other cases, the SO4•− aqueous sink prevails. The above rationale is based on the assumption that the hydroxyl and nitrate radicals as well as 2-NP are at Henry’s equilibria in the gas phase and in the aqueous phase while the sulfate radical-anions exist only in the aqueous phase.
The rate of aqueous-phase reaction of a nitrophenol NP with sulfate radical-anions can also be compared to the rates of gas-phase reactions with OH or NO3 sinks alone, assuming the gas-phase and aqueous-phase reactants follow the Henry’s Law
R X , g a q = r X , g r S O 4 , a q ω = k X , g [ X ] g [ N P ] g k S O 4 , a q [ N P ] a q [ S O 4 · ] a q ω = k X , g k S O 4 , a q H d , X H d , NP ω · [ X ] a q [ S O 4 · ] a q ,
The results obtained for 2-NP (Figure 6) show that the gas-phase sinks dominate in the aerosol and haze waters and in clouds and rains provided hydroxyl or nitrate radicals are in excess. Surprisingly, there is little difference observed between the hydroxyl radicals and nitrate radicals in spite of significant difference between the rate constants for their gas-phase reactions with 2-NP (Table 2). This is explained by lower solubility of NO3 in water. When the aqueous-phase concentrations of both radicals are equal, the gas-phase concentration of NO3 is higher than the concentration of OH and compensates for its lower rate constant. For reference, one can compare the gas-phase and the aqueous phase conversions of 2-NP by hydroxyl radicals or nitrite radicals alone (Figure S3). Conversion by OH in the gas phase dominates in over conversion in aerosol and haze water but not over that in clouds and rain. Conversion of 2-NP by NO3 in the gas phase always dominates over conversion in atmospheric waters.
Rate of conversions of a NP by X (OH or NO3) and by SO4•− in the aqueous phases alone were compared using Equation (12).
r X , a q r SO 4 , a q = k X , a q k SO 4 , a q · [ X ] a q [ SO 4 · ] a q ,
Figure 7 shows that the aqueous-phase conversion of 2-NP by OH radicals dominates over that by SO4•− radical-anions when the ratio of the radicals concentrations is higher than ~0.15. A similar domination of NO3 radicals over SO4•− radical-anions requires the corresponding concentration ratio is greater than ~40.
We expect the comparison of gas-phase and aqueous-phase conversions for other nitrophenols studied would provide similar results when the gas-phase rate constants for these nitrophenols are available.

5. Conclusions

Nitrophenols (2-NP, 3-NP, 4-NP, and 2,4-NP) react fast with SO4•− radical-anions in aqueous solutions. Rate constants for these reactions, along with rate constants of several chlorophenols and phenol, correlate linearly with Brown substituent coefficients and with the relative strength of the O–H bonds in the molecules. The correlation allows estimation of rate constants for reactions of other substituted phenols with sulfate radical-anions.
The aqueous-phase reaction of 2-NP with sulfate radical-anions dominates over the aqueous-phase conversion of 2-NP by OH radicals only when SO4•‒ radicals are at least 10 times more abundant than the OH radicals. Similar domination over NO3 radical requires the concentration of sulfate radicals is at least a quarter of the concentration of nitrate radicals.
The comparison of gas-phase conversion of 2-NP by OH or NO3 radicals against the aqueous-phase conversion by sulfate radical-anions depends on the liquid water contents of a particular atmospheric system considered. In deliquescent aerosol and haze water (ω < 10−10 m3 m−3), gas-phase reactions always prevail over the aqueous-phase reactions. In cloud, rain and fog water (10−8 < ω < 10−6 m3 m−3), the aqueous-phase reaction of 2-NP dominates over the gas-phase conversion of 2-NP by hydroxyl or nitrate radicals provided the aqueous-phase concentration of sulfate radical-anions is not smaller than the aqueous-phase concentration of hydroxyl or nitrate radicals. These conclusions are based on the assumption that the gas-phase and aqueous-phase concentrations of OH, NO3, and 2-NP are bound by Henry’s equilibria.
The gas-phase and aqueous-phase conversions of other nitrophenols are expected to follow similar patterns. However, this expectation should be confirmed by calculations when constants of the gas-phase reactions of the nitrophenols with hydroxyl and nitrate radicals are available.
Last not least, we hope that the rate constants determined in the present work for atmospheric purposes may appear useful for designers of advanced oxidation processes aimed at removal of nitrophenols from various waste effluents utilizing sulfate radical-anions.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4433/10/12/795/s1, Table S1: Chain mechanism of SIV autoxidation catalyzed by FeIII, Figure S1: Concentration of oxygen recorded during autoxidation of NaHSO3 solution in the presence of nitrophenols, Figure S2: Linear plots of reciprocal quasi-stationary rates for autoxidation of S(IV) in the presence of nitrophenols, Table S2: Properties of nitrophenols and sulfate radical anion, Figure S3: The ratio of the gas-phase and the aqueous-phase conversions of 2-NP by OH and NO3 radicals.

Author Contributions

Conceptualization, methodology, investigation: K.J.R.; Data curation: K.J.R. and R.S.; Writing—original draft preparation: K.J.R.; Writing—review and editing: K.J.R. and R.S.

Funding

This research received no external funding.

Acknowledgments

The authors are grateful to Irena Grgić for the invitation to contribute to the Special Issue “Atmospheric Aqueous-Phase Chemistry” and to Józef Ziajka for extensive support in the experimental work behind this presentation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Harrison, M.A.J.; Barra, S.; Borghesi, D.; Vione, D.; Arsene, C.; Iulian Olariu, R. Nitrated phenols in the atmosphere: A review. Atmos. Environ. 2005, 39, 231–248. [Google Scholar] [CrossRef]
  2. Michałowicz, J.; Duda, W. Phenols—Sources and Toxicity. Pol. J. Environ. Stud. 2007, 16, 347–362. [Google Scholar]
  3. Belloli, R.; Bolzacchini, E.; Clerici, L.; Rindone, B.; Sesana, G.; Librando, V. Nitrophenols in air and rainwater. Environ. Eng. Sci. 2006, 23, 405–415. [Google Scholar] [CrossRef]
  4. Delhomme, O.; Morville, S.; Millet, M. Seasonal and diurnal variations of atmospheric concentrations of phenols and nitrophenols measured in the Strasbourg area, France. Atmos. Pollut. Res. 2010, 1, 16–22. [Google Scholar] [CrossRef] [Green Version]
  5. Morville, S.; Scheyer, A.; Mirabel, P.; Millet, M. A multiresidue method for the analysis of phenols and nitrophenols in the atmosphere. J. Environ. Monit. 2004, 6, 963–966. [Google Scholar] [CrossRef]
  6. Morville, S.; Scheyer, A.; Mirabel, P.; Millet, M. Spatial and geographical variations of urban, suburban and rural atmospheric concentrations of phenols and nitrophenols. Environ. Sci. Pollut. Res. 2006, 13, 83–89. [Google Scholar] [CrossRef]
  7. Huston, R.; Chan, Y.C.; Gardner, T.; Shaw, G.; Chapman, H. Characterisation of atmospheric deposition as a source of contaminants in urban rainwater tanks. Water Res. 2009, 43, 1630–1640. [Google Scholar] [CrossRef]
  8. Jaber, F.; Schummer, C.; Al Chami, J.; Mirabel, P.; Millet, M. Solid-phase microextraction and gas chromatography—Mass spectrometry for analysis of phenols and nitrophenols in rainwater, as their t-butyldimethylsilyl derivatives. Anal. Bioanal. Chem. 2007, 387, 2527–2535. [Google Scholar] [CrossRef]
  9. Asman, W.A.H.; Jørgensen, A.; Bossi, R.; Vejrup, K.V.; Bügel Mogensen, B.; Glasius, M. Wet deposition of pesticides and nitrophenols at two sites in Denmark: Measurements and contributions from regional sources. Chemosphere 2005, 59, 1023–1031. [Google Scholar] [CrossRef]
  10. Hofmann, D.; Hartmann, F.; Herrmann, H. Analysis of nitrophenols in cloud water with a miniaturized light-phase rotary perforator and HPLC-MS. Anal. Bioanal. Chem. 2008, 391, 161–169. [Google Scholar] [CrossRef] [Green Version]
  11. Boris, A.J.; Lee, T.; Park, T.; Choi, J.; Seo, S.J.; Collett, J.L., Jr. Fog composition at Baengnyeong Island in the eastern Yellow Sea: Detecting markers of aqueous atmospheric oxidations. Atmos. Chem. Phys. 2016, 16, 437–453. [Google Scholar] [CrossRef] [Green Version]
  12. Chow, K.S.; Huang, X.H.H.; Yu, J.Z. Quantification of nitroaromatic compounds in atmospheric fine particulate matter in Hong Kong over 3 years: Field measurement evidence for secondary formation derived from biomass burning emissions. Environ. Chem. 2016, 13, 665–673. [Google Scholar] [CrossRef] [Green Version]
  13. Duong, H.T.; Kadokami, K.; Trinh, H.T.; Phan, T.Q.; Le, G.T.; Nguyen, D.T.; Nguyen, T.T.; Nguyen, D.T. Target screening analysis of 970 semi-volatile organic compounds adsorbed on atmospheric particulate matter in Hanoi, Vietnam. Chemosphere 2019, 219, 784–795. [Google Scholar] [CrossRef] [PubMed]
  14. Fleming, L.T.; Lin, P.; Laskin, A.; Laskin, J.; Weltman, R.; Edwards, R.D.; Arora, N.K.; Yadav, A.; Meinardi, S.; Blake, D.R.; et al. Molecular composition of particulate matter emissions from dung and brushwood burning household cookstoves in Haryana, India. Atmos. Chem. Phys. 2018, 18, 2461–2480. [Google Scholar] [CrossRef] [Green Version]
  15. Giorio, C.; Bortolini, C.; Kourtchev, I.; Tapparo, A.; Bogialli, S.; Kalberer, M. Direct target and non-target analysis of urban aerosol sample extracts using atmospheric pressure photoionisation high-resolution mass spectrometry. Chemosphere 2019, 224, 786–795. [Google Scholar] [CrossRef] [PubMed]
  16. Irei, S.; Stupak, J.; Gong, X.; Chan, T.-W.; Cox, M.; McLaren, R.; Rudolph, J. Molecular marker study of particulate organic matter in Southern Ontario air. J. Anal. Methods Chem. 2017, 2017, 19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Kitanovski, Z.; Grgic, I.; Vermeylen, R.; Claeys, M.; Maenhaut, W. Liquid chromatography tandem mass spectrometry method for characterization of monoaromatic nitro-compounds in atmospheric particulate matter. J. Chromatogr. A 2012, 1268, 35–43. [Google Scholar] [CrossRef]
  18. Kitanovski, Z.; Shahpoury, P.; Samara, C.; Voliotis, A.; Lammel, G. Composition and mass size distribution of nitrated and oxygenated aromatic compounds in ambient particulate matter from southern and central Europe—implications for origin. Atmos. Chem. Phys. Discuss. 2019, 2019, 1–29. [Google Scholar] [CrossRef]
  19. Krivácsy, Z.; Gelencsér, A.; Kiss, G.; Mészáros, E.; Molnár, Á.; Hoffer, A.; Mészáros, T.; Sárvári, Z.; Temesi, D.; Varga, B.; et al. Study on the chemical character of water soluble organic compounds in fine atmospheric aerosol at the Jungfraujoch. J. Atmos. Chem. 2001, 39, 235–259. [Google Scholar] [CrossRef]
  20. Li, X.; Jiang, L.; Hoa, L.P.; Lyu, Y.; Xu, T.; Yang, X.; Iinuma, Y.; Chen, J.; Herrmann, H. Size distribution of particle-phase sugar and nitrophenol tracers during severe urban haze episodes in Shanghai. Atmos. Environ. 2016, 145, 115–127. [Google Scholar] [CrossRef]
  21. OÖzel, M.Z.; Hamilton, J.F.; Lewis, A.C. New sensitive and quantitative analysis method for organic nitrogen compounds in urban aerosol samples. Environ. Sci. Technol. 2011, 45, 1497–1505. [Google Scholar] [CrossRef] [PubMed]
  22. Teich, M.; Pinxteren, D.; Herrmann, H. Determination of nitrophenolic compounds from atmospheric particles using hollow-fiber liquid-phase microextraction and capillary electrophoresis/mass spectrometry analysis. Electrophoresis 2014, 35, 1353–1361. [Google Scholar] [CrossRef] [PubMed]
  23. Wang, L.; Wang, X.; Gu, R.; Wang, H.; Yao, L.; Wen, L.; Zhu, F.; Wang, W.; Xue, L.; Yang, L.; et al. Observations of fine particulate nitrated phenols in four sites in northern China: Concentrations, source apportionment, and secondary formation. Atmos. Chem. Phys. 2018, 18, 4349–4359. [Google Scholar] [CrossRef] [Green Version]
  24. Williams, B.J.; Goldstein, A.H.; Kreisberg, N.M.; Hering, S.V.; Worsnop, D.R.; Ulbrich, I.M.; Docherty, K.S.; Jimenez, J.L. Major components of atmospheric organic aerosol in southern California as determined by hourly measurements of source marker compounds. Atmos. Chem. Phys. 2010, 10, 11577–11603. [Google Scholar] [CrossRef] [Green Version]
  25. Zhang, Y.Y.; Müller, L.; Winterhalter, R.; Moortgat, G.K.; Hoffmann, T.; Pöschl, U. Seasonal cycle and temperature dependence of pinene oxidation products, dicarboxylic acids and nitrophenols in fine and coarse air particulate matter. Atmos. Chem. Phys. 2010, 10, 7859–7873. [Google Scholar] [CrossRef] [Green Version]
  26. Kristanti, R.A.; Kanbe, M.; Toyama, T.; Tanaka, Y.; Tang, Y.; Wu, X.; Mori, K. Accelerated biodegradation of nitrophenols in the rhizosphere of Spirodela polyrrhiza. J. Environ. Sci. 2012, 24, 800–807. [Google Scholar] [CrossRef]
  27. Vozňáková, Z.; Podehradská, J.; Kohličková, M. Determination of nitrophenols in soil. Chemosphere 1996, 33, 285–291. [Google Scholar] [CrossRef]
  28. Chiron, S.; Minero, C.; Vione, D. Occurrence of 2,4-dichlorophenol and of 2,4-dichloro-6-nitrophenol in the Rhone River Delta (Southern France). Environ. Sci. Technol. 2007, 41, 3127–3133. [Google Scholar] [CrossRef]
  29. Cho, E.; Khim, J.; Chung, S.; Seo, D.; Son, Y. Occurrence of micropollutants in four major rivers in Korea. Sci. Total Environ. 2014, 491–492, 138–147. [Google Scholar] [CrossRef]
  30. Musilová, J.; Barek, J.; Pecková, K. Determination of nitrophenols in drinking and river water by differential pulse voltammetry at boron-doped diamond film electrode. Electroanalysis 2011, 23, 1236–1244. [Google Scholar] [CrossRef]
  31. Schmidt-Bäumler, K.; Heberer, T.; Stan, H.-J. Occurrence and distribution of organic contaminants in the aquatic system in Berlin.Part II: Substituted phenols in Berlin surface water. Acta Hydroch. Hydrob. 1999, 27, 143–149. [Google Scholar] [CrossRef]
  32. Balasubramanian, P.; Balamurugan, T.S.T.; Chen, S.-M.; Chen, T.-W. Simplistic synthesis of ultrafine CoMnO3 nanosheets: An excellent electrocatalyst for highly sensitive detection of toxic 4-nitrophenol in environmental water samples. J. Hazard. Mater. 2018. [Google Scholar] [CrossRef]
  33. Jay, K.; Stieglitz, L. Identification and quantification of volatile organic components in emissions of waste incineration plants. Chemosphere 1995, 30, 1249–1260. [Google Scholar] [CrossRef]
  34. Goi, A.; Trapido, M. Comparison of advanced oxidation processes for the destruction of 2,4-dinitrophenol. Proc. Est. Acad. Sci. Chem. 2001, 50, 5–17. [Google Scholar]
  35. Inomata, S.; Fushimi, A.; Sato, K.; Fujitani, Y.; Yamada, H. 4-Nitrophenol, 1-nitropyrene, and 9-nitroanthracene emissions in exhaust particles from diesel vehicles with different exhaust gas treatments. Atmos. Environ. 2015, 110, 93–102. [Google Scholar] [CrossRef]
  36. Lu, C.; Wang, X.; Dong, S.; Zhang, J.; Li, J.; Zhao, Y.; Liang, Y.; Xue, L.; Xie, H.; Zhang, Q.; et al. Emissions of fine particulate nitrated phenols from various on-road vehicles in China. Environ. Res. 2019, 179, 108709. [Google Scholar] [CrossRef] [PubMed]
  37. Rubio, M.A.; Bustamante, P.; Vásquez, P. Atmospheric phenolic derivatives as tracers in an urban area. J. Chil. Chem. Soc. 2019, 64, 4407–4411. [Google Scholar] [CrossRef] [Green Version]
  38. Environmental Protection Agency. Interim Reregistration Eligibility Decision for Methyl Parathion, Case No. 0153; Office of Prevention, Pesticides and Toxic Substances, Ed.; United States Environmental Protection Agency: Washington, DC, USA, 2006; p. 20460.
  39. World Health Organization. Parathion in Drinking-Water; Background document for development of WHO guidelines for drinking-water quality; World Health Organization: Geneva, Switzerland, 2004. [Google Scholar]
  40. Bluvshtein, N.; Lin, P.; Flores, J.M.; Segev, L.; Mazar, Y.; Tas, E.; Snider, G.; Weagle, C.; Brown, S.S.; Laskin, A.; et al. Broadband optical properties of biomass-burning aerosol and identification of brown carbon chromophores. J. Geophys. Res. Atmos. 2017, 122, 5441–5456. [Google Scholar] [CrossRef]
  41. Priestley, M.; Le Breton, M.; Bannan, T.J.; Leather, K.E.; Bacak, A.; Reyes-Villegas, E.; De Vocht, F.; Shallcross, B.M.A.; Brazier, T.; Anwar Khan, M.; et al. Observations of Isocyanate, Amide, Nitrate, and Nitro Compounds From an Anthropogenic Biomass Burning Event Using a ToF-CIMS. J. Geophys. Res. Atmos. 2018, 123, 7687–7704. [Google Scholar] [CrossRef]
  42. Xie, M.; Chen, X.; Hays, M.D.; Holder, A.L. Composition and light absorption of N-containing aromatic compounds in organic aerosols from laboratory biomass burning. Atmos. Chem. Phys. 2019, 19, 2899–2915. [Google Scholar] [CrossRef] [Green Version]
  43. Wang, X.; Gu, R.; Wang, L.; Xu, W.; Zhang, Y.; Chen, B.; Li, W.; Xue, L.; Chen, J.; Wang, W. Emissions of fine particulate nitrated phenols from the burning of five common types of biomass. Environ. Pollut. 2017, 230, 405–412. [Google Scholar] [CrossRef] [PubMed]
  44. Hinrichs, R.Z.; Buczek, P.; Trivedi, J.J. Solar Absorption by Aerosol-Bound Nitrophenols Compared to Aqueous and Gaseous Nitrophenols. Environ. Sci.Technol. 2016, 50, 5661–5667. [Google Scholar] [CrossRef] [PubMed]
  45. Lin, P.; Bluvshtein, N.; Rudich, Y.; Nizkorodov, S.A.; Laskin, J.; Laskin, A. Molecular Chemistry of Atmospheric Brown Carbon Inferred from a Nationwide Biomass Burning Event. Environ. Sci.Technol. 2017, 51, 11561–11570. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Barzaghi, P.; Herrmann, H. A mechanistic study of the oxidation of phenol by OH/NO2/NO3 in aqueous solution. Phys. Chem. Chem. Phys. 2002, 4, 3669–3675. [Google Scholar] [CrossRef]
  47. Heal, M.R.; Harrison, M.A.J.; Neil Cape, J. Aqueous-phase nitration of phenol by N2O5 and ClNO2. Atmos. Environ. 2007, 41, 3515–3520. [Google Scholar] [CrossRef] [Green Version]
  48. Vione, D.; Maurino, V.; Minero, C.; Pelizzetti, E. Aqueous atmospheric chemistry: Formation of 2,4-dinitrophenol upon nitration of 2-nitrophenol and 4-nitrophenol in solution. Environ. Sci. Technol. 2005, 39, 7921–7931. [Google Scholar] [CrossRef]
  49. Vione, D.; Maurino, V.; Minero, C.; Pelizzetti, E.; Harrison, M.A.J.; Olariu, R.; Arsene, C. Photochemical reactions in the tropospheric aqueous phase and on particulate matter. Chem. Soc. Rev. 2006, 35, 441–453. [Google Scholar] [CrossRef]
  50. Yuan, B.; Liggio, J.; Wentzell, J.; Li, S.M.; Stark, H.; Roberts, J.M.; Gilman, J.; Lerner, B.; Warneke, C.; Li, R.; et al. Secondary formation of nitrated phenols: Insights from observations during the Uintah Basin Winter Ozone Study (UBWOS) 2014. Atmos. Chem. Phys. 2016, 16, 2139–2153. [Google Scholar] [CrossRef] [Green Version]
  51. Atkinson, R. Gas-phase tropospheric chemistry of organic compounds. J. Phys. Chem. Ref. Data Monogr. 1994, Monoghraph No. 2. 1–216. [Google Scholar]
  52. Atkinson, R.; Aschmann, S.M.; Arey, J. Reactions of hydroxyl and nitrogen trioxide radicals with phenol, cresols, and 2-nitrophenol at 296 ± 2 K. Environ. Sci.Technol. 1992, 26, 1397–1403. [Google Scholar] [CrossRef]
  53. Barsotti, F.; Bartels-Rausch, T.; De Laurentiis, E.; Ammann, M.; Brigante, M.; Mailhot, G.; Maurino, V.; Minero, C.; Vione, D. Photochemical Formation of Nitrite and Nitrous Acid (HONO) upon Irradiation of Nitrophenols in Aqueous Solution and in Viscous Secondary Organic Aerosol Proxy. Environ. Sci.Technol. 2017, 51, 7486–7495. [Google Scholar] [CrossRef] [PubMed]
  54. Vione, D.; Maurino, V.; Minero, C.; Duncianu, M.; Olariu, R.-I.; Arsene, C.; Sarakha, M.; Mailhot, G. Assessing the transformation kinetics of 2-and 4-nitrophenol in the atmospheric aqueous phase. Implications for the distribution of both nitroisomers in the atmosphere. Atmos. Environ. 2009, 43, 2321–2327. [Google Scholar] [CrossRef]
  55. Hems, R.F.; Abbatt, J.P.D. Aqueous Phase Photo-oxidation of Brown Carbon Nitrophenols: Reaction Kinetics, Mechanism, and Evolution of Light Absorption. ACS Earth Space Chem. 2018, 2, 225–234. [Google Scholar] [CrossRef]
  56. Li, Y.J.; Huang, D.D.; Cheung, H.Y.; Lee, A.K.Y.; Chan, C.K. Aqueous-phase photochemical oxidation and direct photolysis of vanillin—A model compound of methoxy phenols from biomass burning. Atmos. Chem. Phys. 2014, 14, 2871–2885. [Google Scholar] [CrossRef] [Green Version]
  57. Wei, B.; Sun, J.; Mei, Q.; He, M. Mechanism and kinetic of nitrate radical-initiated atmospheric reactions of guaiacol (2-methoxyphenol). Comp. Theor. Chem. 2018, 1129, 1–8. [Google Scholar] [CrossRef]
  58. Yu, L.; Smith, J.; Laskin, A.; Anastasio, C.; Laskin, J.; Zhang, Q. Chemical characterization of SOA formed from aqueous-phase reactions of phenols with the triplet excited state of carbonyl and hydroxyl radical. Atmos. Chem. Phys. 2014, 14, 13801–13816. [Google Scholar] [CrossRef] [Green Version]
  59. Yu, L.; Smith, J.; Laskin, A.; George, K.M.; Anastasio, C.; Laskin, J.; Dillner, A.M.; Zhang, Q. Molecular transformations of phenolic SOA during photochemical aging in the aqueous phase: Competition among oligomerization, functionalization, and fragmentation. Atmos. Chem. Phys. 2016, 16, 4511–4527. [Google Scholar] [CrossRef] [Green Version]
  60. Litter, M.I. Introduction to Photochemical Advanced Oxidation Processes for Water Treatment. In Environmental Photochemistry Part II; Boule, P., Bahnemann, D.W., Robertson, P.K.J., Eds.; Springer: Berlin/Heidelberg, Germany, 2005; Volume 2, Pt M, pp. 325–366. [Google Scholar]
  61. Goi, A.; Trapido, M. Hydrogen peroxide photolysis, Fenton reagent and photo-Fenton for the degradation of nitrophenols: A comparative study. Chemosphere 2002, 46, 913–922. [Google Scholar] [CrossRef]
  62. Ahn, W.-Y.; Sheeley, S.A.; Rajh, T.; Cropek, D.M. Photocatalytic reduction of 4-nitrophenol with arginine-modified titanium dioxide nanoparticles. Appl. Catal. B Environ. 2007, 74, 103–110. [Google Scholar] [CrossRef]
  63. Ali, I.; Hassan, A.; Shabaan, S.; El-Nasser, K. Synthesis and characterization of composite catalysts Cr/ZSM-5 and their effects toward photocatalytic degradation of p-nitrophenol. Arab. J. Chem. 2017, 10, S2106–S2114. [Google Scholar] [CrossRef] [Green Version]
  64. Priya, M.H.; Madras, G. Kinetics of photocatalytic degradation of chlorophenol, nitrophenol, and their mixtures. Ind. Eng. Chem. Res. 2006, 45, 482–486. [Google Scholar] [CrossRef]
  65. Liu, Y.; Liu, H.; Li, Y. Comparative study of the electrocatalytic oxidation and mechanism of nitrophenols at Bi-doped lead dioxide anodes. Appl. Catal. B Environ. 2008, 84, 297–302. [Google Scholar] [CrossRef]
  66. Christensen, P.A.; Egerton, T.A.; Kosa, S.A.M.; Tinlin, J.R.; Scott, K. The photoelectrocatalytic oxidation of aqueous nitrophenol using a novel reactor. J. Appl. Electrochem. 2005, 35, 683–692. [Google Scholar] [CrossRef]
  67. Zhou, M.; Lei, L. An improved UV/Fe3+ process by combination with electrocatalysis for p-nitrophenol degradation. Chemosphere 2006, 63, 1032–1040. [Google Scholar] [CrossRef] [PubMed]
  68. Apolinário, Â.C.; Silva, A.M.T.; Machado, B.F.; Gomes, H.T.; Araújo, P.P.; Figueiredo, J.L.; Faria, J.L. Wet air oxidation of nitro-aromatic compounds: Reactivity on single-and multi-component systems and surface chemistry studies with a carbon xerogel. Appl. Catal. B Environ. 2008, 84, 75–86. [Google Scholar] [CrossRef]
  69. Diaz, E.; Polo, A.M.; Mohedano, A.F.; Casas, J.A.; Rodriguez, J.J. On the biodegradability of nitrophenols and their reaction products by catalytic hydrogenation. J. Chem. Technol. Biotechnol. 2012, 87, 1263–1269. [Google Scholar] [CrossRef]
  70. Anipsitakis, G.P.; Dionysiou, D.D.; Gonzalez, M.A. Cobalt-Mediated Activation of Peroxymonosulfate and Sulfate Radical Attack on Phenolic Compounds. Implications of Chloride Ions. Environ. Sci. Technol. 2006, 40, 1000–1007. [Google Scholar] [CrossRef]
  71. EPA. The Original List of Hazardous Air Pollutants. Available online: http://www.epa.gov/ttn/atw/188polls.html (accessed on 29 October 2014).
  72. EPA. Priority Pollutants. Available online: http://water.epa.gov/scitech/methods/cwa/pollutants.cfm (accessed on 27 October 2014).
  73. Keith, L.; Telliard, W. ES&T Special Report: Priority pollutants: I-a perspective view. Environ. Sci.Technol. 1979, 13, 416–423. [Google Scholar] [CrossRef]
  74. Gramatica, P.; Santagostino, A.; Bolzacchini, E.; Rindone, B. Atmospheric monitoring, toxicology and QSAR modelling of nitrophenols. Fresenius Environ. Bull. 2002, 11, 757–762. [Google Scholar]
  75. Schafer, K.S.; Reeves, M.; Spitzer, S.; Kegley, S.E. Chemical Trespass Pesticides in our Bodies and Corporate Accountability; Pesticide Action Network North America: San Francisco, CA, USA, 2004. [Google Scholar]
  76. Zhang, D.-P.; Wu, W.-L.; Long, H.-Y.; Liu, Y.-C.; Yang, Z.-S. Voltammetric Behavior of o-Nitrophenol and Damage to DNA. Int. J. Mol. Sci. 2008, 9, 316–326. [Google Scholar] [CrossRef] [Green Version]
  77. Buxton, G.V.; Salmon, G.A.; Williams, J.E. The reactivity of biogenic monoterpenes towards OH·and SO4 radicals in de-oxygenated acidic solution. J. Atmos. Chem. 2000, 36, 111–134. [Google Scholar] [CrossRef]
  78. Clifton, C.L.; Huie, R.E. Rate constants for hydrogen abstraction reactions of the sulfate radical, SO4. Alcohols. Int. J. Chem. Kinet. 1989, 21, 677–687. [Google Scholar] [CrossRef]
  79. George, C.; Rassy, H.E.; Chovelon, J.M. Reactivity of selected volatile organic compounds (VOCs) toward the sulfate radical (SO4). Int. J. Chem. Kinet. 2001, 33, 539–547. [Google Scholar] [CrossRef]
  80. Herrmann, H.; Hoffmann, D.; Schaefer, T.; Bräuer, P.; Tilgner, A. Tropospheric Aqueous-Phase Free-Radical Chemistry: Radical Sources, Spectra, Reaction Kinetics and Prediction Tools. ChemPhysChem 2010, 11, 3796–3822. [Google Scholar] [CrossRef] [PubMed]
  81. Rudzinski, K.J. Degradation of Isoprene in the Presence of Sulphoxy Radical Anions. J. Atmos. Chem. 2004, 48, 191–216. [Google Scholar] [CrossRef]
  82. Rudziński, K.J.; Gmachowski, L.; Kuznietsova, I. Reactions of isoprene and sulphoxy radical-anions—A possible source of atmospheric organosulphites and organosulphates. Atmos. Chem. Phys. 2009, 9, 2129–2140. [Google Scholar] [CrossRef] [Green Version]
  83. Ziajka, J.; Rudzinski, K. Autoxidation of S-IV inhibited by chlorophenols reacting with sulfate radicals. Environ. Chem. 2007, 4, 355–363. [Google Scholar] [CrossRef]
  84. Ziajka, J.; Pasiuk-Bronikowska, W. Autoxidation of sulphur dioxide in the presence of alcohols under conditions related to the tropospheric aqueous phase. Atmos. Environ. 2003, 37, 3913–3922. [Google Scholar] [CrossRef]
  85. Ziajka, J.; Pasiuk-Bronikowska, W. Rate constants for atmospheric trace organics scavenging SO4 in the Fe-catalysed autoxidation of S(IV). Atmos. Environ. 2005, 39, 1431–1438. [Google Scholar] [CrossRef]
  86. Grgić, I.; Podkrajšek, B.; Barzaghi, P.; Herrmann, H. Scavenging of SO4 radical anions by mono-and dicarboxylic acids in the Mn(II)-catalyzed S(IV) oxidation in aqueous solution. Atmos. Environ. 2007, 41, 9187–9194. [Google Scholar] [CrossRef]
  87. Elliot, A.J.; McCracken, D.R.; Buxton, G.V.; Wood, N.D. Estimation of rate constants for near-diffusion-controlled reactions in water at high temperatures. J. Chem. Soc. Faraday Trans. 1990, 86, 1539–1547. [Google Scholar] [CrossRef]
  88. Zhu, L.; Nicovich, J.M.; Wine, P.H. Temperature-dependent kinetics studies of aqueous phase reactions of SO4 radicals with dimethylsulfoxide, dimethylsulfone, and methanesulfonate. J. Photochem. Photobiol. A 2003, 157, 311–319. [Google Scholar] [CrossRef]
  89. Noyes, R.M. Effects of diffusion rates on chemical kinetics. In Progress in Reaction Kinetics; Porter, G., Stevens, B., Eds.; Pergamon Press: Oxford, UK, 1961; Volume 1, pp. 129–160. [Google Scholar]
  90. North, A.M. The Collision Theory of Chemical Reactions in Liquids; Methuen: London, UK, 1964. [Google Scholar]
  91. Hansch, C.; Leo, A.; Taft, R.W. A survey of Hammett substituent constants and resonance and field parameters. Chem. Rev. 1991, 91, 165–195. [Google Scholar] [CrossRef]
  92. Jones, R.A.Y. Physical and Mechanistic Organic Chemistry; Cambridge University Press: Cambridge, MA, USA, 1979. [Google Scholar]
  93. Jonsson, M.; Lind, J.; Eriksen, T.E.; Merényi, G. O–H bond strengths and one-electron reduction potentials of multisubstituted phenols and phenoxyl radicals. Predictions using free energy relationships. J. Chem. Soc. Perkin Trans. 2 1993, 1567–1568. [Google Scholar] [CrossRef]
  94. Herrmann, H.; Schaefer, T.; Tilgner, A.; Styler, S.A.; Weller, C.; Teich, M.; Otto, T. Tropospheric aqueous-phase chemistry: Kinetics, mechanisms, and its coupling to a changing gas phase. Chem. Rev. 2015, 115, 4259–4334. [Google Scholar] [CrossRef]
  95. Atkinson, R. Kinetics and mechanisms of the gas-phase reactions of the hydroxyl radical with organic compounds. Chem. Ref. 1989, Monograph No. 1. 1–246. [Google Scholar]
  96. Herrmann, H. Kinetics of Aqueous Phase Reactions Relevant for Atmospheric Chemistry. Chem. Rev. 2003, 103, 4691–4716. [Google Scholar] [CrossRef]
  97. Guo, X.X.; Brimblecombe, P. Henry’s law constants of phenol and mononitrophenols in water and aqueous sulfuric acid. Chemosphere 2007, 68, 436–444. [Google Scholar] [CrossRef]
Figure 1. Concentration of oxygen recorded during autoxidation of NaHSO3 inhibited by 4-NP at various initial concentrations.
Figure 1. Concentration of oxygen recorded during autoxidation of NaHSO3 inhibited by 4-NP at various initial concentrations.
Atmosphere 10 00795 g001
Figure 2. Rate of oxygen consumption during autoxidation of NaHSO3 inhibited by 4-NP (initially 0.28 mM) evaluated from data in Figure 1.
Figure 2. Rate of oxygen consumption during autoxidation of NaHSO3 inhibited by 4-NP (initially 0.28 mM) evaluated from data in Figure 1.
Atmosphere 10 00795 g002
Figure 3. Linear plots of reciprocal quasi-stationary rates of the autoxidation of NaHSO3 inhibited by 4-NP or by the reference ethanol versus initial concentrations of each inhibitor. The slope uncertainties are equal to standard errors of regression coefficients.
Figure 3. Linear plots of reciprocal quasi-stationary rates of the autoxidation of NaHSO3 inhibited by 4-NP or by the reference ethanol versus initial concentrations of each inhibitor. The slope uncertainties are equal to standard errors of regression coefficients.
Atmosphere 10 00795 g003
Figure 4. Correlation of rate constants for reactions of sulfate radical-anions with phenol (P), chlorophenols (green circles: 4-CP; 2,4-DCP; 2,5-DCP; 2,4,5-TCP; 2,4,6-TCP; 2,3,5,6-TTCP) [83] and nitrophenols (blue squares: 2-NP; 3-NP; 4-NP; 2,4-DNP; 2,4,6-TNP) [this work] against (a) sums of Brown substituent coefficients and (b) the relative strength of the O–H bond.
Figure 4. Correlation of rate constants for reactions of sulfate radical-anions with phenol (P), chlorophenols (green circles: 4-CP; 2,4-DCP; 2,5-DCP; 2,4,5-TCP; 2,4,6-TCP; 2,3,5,6-TTCP) [83] and nitrophenols (blue squares: 2-NP; 3-NP; 4-NP; 2,4-DNP; 2,4,6-TNP) [this work] against (a) sums of Brown substituent coefficients and (b) the relative strength of the O–H bond.
Atmosphere 10 00795 g004
Figure 5. Total rate of 2-NP conversion in the atmosphere due the gas-phase and aqueous-phase reactions with OH radicals (a) or NO3 radicals (b) compared to the rate of the aqueous-phase reaction of 2-NP with sulfate radical-anions for various ratios of radicals in the aqueous phases ([OH]aq/[SO4•−]aq) and various liquid water contents ω (based on Equation (10)).
Figure 5. Total rate of 2-NP conversion in the atmosphere due the gas-phase and aqueous-phase reactions with OH radicals (a) or NO3 radicals (b) compared to the rate of the aqueous-phase reaction of 2-NP with sulfate radical-anions for various ratios of radicals in the aqueous phases ([OH]aq/[SO4•−]aq) and various liquid water contents ω (based on Equation (10)).
Atmosphere 10 00795 g005
Figure 6. Rate of 2-NP conversion in the atmosphere due to the gas-phase reaction with OH radicals (a) or with NO3 radicals (b) compared to the rate of the aqueous-phase reaction of 2-NP with sulfate radical-anions for various proportions of radicals in the aqueous phase [OH]aq/[SO4•−]aq and varying liquid water contents (ω) (based on Equation (11)).
Figure 6. Rate of 2-NP conversion in the atmosphere due to the gas-phase reaction with OH radicals (a) or with NO3 radicals (b) compared to the rate of the aqueous-phase reaction of 2-NP with sulfate radical-anions for various proportions of radicals in the aqueous phase [OH]aq/[SO4•−]aq and varying liquid water contents (ω) (based on Equation (11)).
Atmosphere 10 00795 g006
Figure 7. Comparison of the rates of aqueous-phase conversion of 2-NP due to reaction with SO4•− radicals and OH radicals (blue line) or NO3 radicals (red line).
Figure 7. Comparison of the rates of aqueous-phase conversion of 2-NP due to reaction with SO4•− radicals and OH radicals (blue line) or NO3 radicals (red line).
Atmosphere 10 00795 g007
Table 1. Initial concentrations of reactants used in the experiments.
Table 1. Initial concentrations of reactants used in the experiments.
CompoundAbbreviationConcentration Range, mM
2-nitrophenol2-NP0.032–0.347
3-nitrophenol3-NP0.019–0.089
4-nitrophenol4-NP0.073–0.711
2,4-dinitrophenol2,4-DNP0.169–1.325
2,4,6-trinitrophenol2,4,6-TNP0.197–0.788
HSO3 2
O2 ~0.25
Fe(ClO4)3 0.01
Table 2. Experimental slopes of linear plots (Figure 3 and Figure S2) and rate constants for reactions of nitrophenols with sulfate radical-anions (observed and corrected for diffusional limitations).
Table 2. Experimental slopes of linear plots (Figure 3 and Figure S2) and rate constants for reactions of nitrophenols with sulfate radical-anions (observed and corrected for diffusional limitations).
CompoundSlope, s M−2kobserved, M−1s−1kdiffusionA, M−1s−1kreaction, M−1s−1(kr − ko)/ko, %
EtOH (reference)(1.735 ± 0.001) × 108 (4.30 ± 0.86) × 107
2-NP(3.662 ± 0.235) × 109(9.08 ± 2.40) × 1082.01 × 1010(9.50 ± 4.58) × 1084.74
3-NP(6.957 ± 0.277) × 109(1.72 ± 0.41) × 1092.01 × 1010(1.89 ± 0,89) × 1099.53
4-NP(2.665 ± 0.985) × 109(6.60 ± 3.77) × 1082.01 × 1010(6.83 ± 5.42) × 1083.41
2,4-DNP(1.154 ± 0.152) × 109(2.86 ± 0.95) × 1082.13 × 1010(2.90 ± 1.58) × 1081.51
2,4,6-TNP(2.865 ± 0.180) × 108(7.10 ± 1.87) × 1072.18 × 1010(7.12 ± 3.31) × 1070.33
A—assumed unceratinty 2 × 109 M−1 s−1(10%).
Table 3. Rate and Henry’s constants for selected atmospheric reactants at 296–298 K.
Table 3. Rate and Henry’s constants for selected atmospheric reactants at 296–298 K.
Reaction X + NPkX,gkX,aqHdXHdNP
cm3 molecule−1 s−1dm3 mol−1 s−1dm3 mol−1 s−1
OH + 2-NP9.00 × 10−13 (a)5.42 × 1085.90 × 109 (b)6.11 × 102 (d)5.17 × 103 (e)
NO3 + 2-NP2.00 × 10−14 (a)1.20 × 1072.30 × 107 (c)1.47 × 101 (d)
(a) [52,95]; (b) [54]; (c) [96]; (d) [81]; (e) [97].
Table 4. Ranges of radical concentrations in gas phase, clouds and deliquescent particles [94].
Table 4. Ranges of radical concentrations in gas phase, clouds and deliquescent particles [94].
Gas Phase (a)Aqueous Phase
OHNO3OHNO3SO4•−[OH]/[SO4•−][NO3]/[SO4•−]
molecule cm−3mol dm−3
Minimal1.4 × 1027.9 × 1061.4 × 10−161.6 × 10−165 × 10−171.6 × 10−41.8 × 10−3
Maximal6.6 × 1031.2 × 1078.0 × 10−123.0 × 10−139 × 10−131.6 × 1056.0 × 103
(a) calculated using Henry’s constants from Table 3.

Share and Cite

MDPI and ACS Style

Rudziński, K.J.; Szmigielski, R. Aqueous Reactions of Sulfate Radical-Anions with Nitrophenols in Atmospheric Context. Atmosphere 2019, 10, 795. https://doi.org/10.3390/atmos10120795

AMA Style

Rudziński KJ, Szmigielski R. Aqueous Reactions of Sulfate Radical-Anions with Nitrophenols in Atmospheric Context. Atmosphere. 2019; 10(12):795. https://doi.org/10.3390/atmos10120795

Chicago/Turabian Style

Rudziński, Krzysztof J., and Rafał Szmigielski. 2019. "Aqueous Reactions of Sulfate Radical-Anions with Nitrophenols in Atmospheric Context" Atmosphere 10, no. 12: 795. https://doi.org/10.3390/atmos10120795

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop