Next Article in Journal
RPiRLS: Quantitative Predictions of RNA Interacting with Any Protein of Known Sequence
Next Article in Special Issue
Synthesis of Nanometer Sized Bis- and Tris-trityl Model Compounds with Different Extent of Spin–Spin Coupling
Previous Article in Journal
Bivariate Correlation Analysis of the Chemometric Profiles of Chinese Wild Salvia miltiorrhiza Based on UPLC-Qqq-MS and Antioxidant Activities
Previous Article in Special Issue
Radical Chemistry in a Femtosecond Laser Plasma: Photochemical Reduction of Ag+ in Liquid Ammonia Solution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mechanistic Insight into the Degradation of Nitrosamines via Aqueous-Phase UV Photolysis or a UV-Based Advanced Oxidation Process: Quantum Mechanical Calculations

Department of Civil and Environmental Engineering, Michigan Technological University, 1400 Townsend Drive, Houghton, MI 49931, USA
*
Author to whom correspondence should be addressed.
Molecules 2018, 23(3), 539; https://doi.org/10.3390/molecules23030539
Submission received: 8 February 2018 / Revised: 21 February 2018 / Accepted: 23 February 2018 / Published: 28 February 2018
(This article belongs to the Special Issue Radical Chemistry)

Abstract

:
Nitrosamines are a group of carcinogenic chemicals that are present in aquatic environments that result from byproducts of industrial processes and disinfection products. As indirect and direct potable reuse increase, the presence of trace nitrosamines presents challenges to water infrastructures that incorporate effluent from wastewater treatment. Ultraviolet (UV) photolysis or UV-based advanced oxidation processes that produce highly reactive hydroxyl radicals are promising technologies to remove nitrosamines from water. However, complex reaction mechanisms involving radicals limit our understandings of the elementary reaction pathways embedded in the overall reactions identified experimentally. In this study, we perform quantum mechanical calculations to identify the hydroxyl radical-induced initial elementary reactions with N-nitrosodimethylamine (NDMA), N-nitrosomethylethylamine, and N-nitrosomethylbutylamine. We also investigate the UV-induced NDMA degradation mechanisms. Our calculations reveal that the alkyl side chains of nitrosamine affect the reaction mechanism of hydroxyl radicals with each nitrosamine investigated in this study. Nitrosamines with one- or two-carbon alkyl chains caused the delocalization of the electron density, leading to slower subsequent degradation. Additionally, three major initial elementary reactions and the subsequent radical-involved reaction pathways are identified in the UV-induced NDMA degradation process. This study provides mechanistic insight into the elementary reaction pathways, and a future study will combine these results with the kinetic information to predict the time-dependent concentration profiles of nitrosamines and their transformation products.

Graphical Abstract

1. Introduction

Nitrosamines, which contain N–NO functional groups, are a group of chemicals that pose mutagenicity, teratogenicity, and carcinogenicity [1]. Nitrosamines are the byproducts of various manufacturing, agricultural, and natural processes and have been found in natural aquatic environments and in the effluent of wastewater treatment processes [2]. As a type of nitrosamine, N-nitrosodimethyl amine (NDMA, (CH3)2N–NO) is a low-molecular-weight, neutral, organic contaminant that has also been found to be present in aquatic environments. The California Department of Health Services has set notification levels of 10 ng/L for NDMA and other nitrosamines in drinking water [3].
Ultraviolet (UV) photolysis and UV-based advanced oxidation processes (AOPs) that produce highly reactive hydroxyl radicals (HO) are attractive and promising water treatment technologies, which can inactivate pathogens and destroy a wide variety of organic chemical contaminants [4,5]. UV photolysis and UV-AOPs have been employed in wastewater reclamation processes for indirect or direct potable reuse of treated wastewater to increase water security and address water scarcity issues in many arid regions [6]. Wastewater reclamation processes use multiple barriers to physically remove pathogens and chemical contaminants via microfiltration/ultrafiltration, followed by nanofiltration (NF)/reverse osmosis (RO). After the NF/RO process, UV photolysis or UV-AOPs inactivate pathogens and destroy chemicals present in the NF/RO permeate stream. Over 50% of NDMA has been found to be present in the NF/RO permeate, and the use of UV photolysis or UV-AOPs are necessary to remove NDMA and other nitrosamines [7].
UV photolysis using a low-pressure UV lamp that emits photons at a wavelength of 254 nm is very effective at destroying NDMA due to the high molar absorptivity (1650 M−1cm−1 at 253.7 nm) and highly reactive HO produced in AOPs rapidly react with many nitrosamines to effectively destroy the initial contaminants (the second order reaction rate constants of HO; k = 108–109 M−1s−1) [8,9]. However, complex chemical reactions involving radicals produce a number of transformation byproducts, and hence, detailed reactivity and reaction pathways for NDMA and other nitrosamines have not been elucidated yet. For example, Mezyk’s group studied the kinetics of HO with various structurally different nitrosamines, and found that NDMA, N-nitrosomethylethylamine (NMEA) and N-nitrosodiethylamine (NDEA) showed different reactivity and degradation efficiency from other nitrosamines that have longer alkyl chains adjacent to the N–NO functional group. They proposed that radical delocalization caused the differences in the degradation efficiency, but the detailed reaction pathway has not been identified yet [8]. Stefan and Bolton (2002) investigated reaction pathways for NDMA degradation based on laboratory-scale batch photolysis experiments and explained the initial photolysis mechanisms based on the reaction pathways previously identified by studies in the 1960s and 1970s [10,11,12,13,14]. UV-induced NDMA degradation pathways were studied at both pH 3 and pH 7 to identify the transformation products, such as methylamine, dimethylamine, formaldehyde, formic acid, nitrite ion and nitrate ion [15,16]. Their careful experiments and measurement of transformation products proposed several key reaction pathways that were induced by UV photolysis at a wavelength of 253.7 nm at different pH values [15,16]. However, some of the pathways involved in the formation of transformation products are still unknown. UV-induced NDMA degradation was also studied and identified previously unknown reactive species in the NDMA degradation pathways [17,18,19]. The HO-induced NDMA degradation mechanisms were studied in an ozone-based AOP, and general reaction mechanisms were proposed [20,21]. The major transformation mechanisms were proposed based on experimental studies of the products, but the elementary reaction pathways are not known due to difficulties in identifying the embedded reactions that were involved in the overall reaction.
Quantum mechanical (QM) calculations using ab initio methods or density functional theory (DFT) are attractive approaches to identify elementary reaction pathways and the kinetics of complex fast radical reactions [22]. QM calculations have been used to support experimentally identified reaction pathways by calculating the reaction energy using statistical thermodynamics. Aqueous-phase enthalpy and free energies of activation and reaction were calculated to determine the dominant degradation pathway of dimethyl phthalate [23]. Elementary reactions involved in the HO-induced mineralization of flutriafol were identified [24]. DFT calculations were used to determine the NDMA formation mechanism from N,N-dimethylsulfamide via ozonation in water [25]. A high-level multi-point energy method was used to calculate the aqueous-phase free energies of activation for HO-induced reactions of a wide variety of organic compounds, including aliphatic compounds, alkenes, and aromatic compounds [26,27,28]. These studies highlight the usefulness of QM-based calculations to provide insight into reaction mechanisms that cannot be obtained by experiments. In addition, the findings from QM-based calculations also provide potential transformation products that can be identified in future experiments.
In this study, we use QM-based calculations to identify the HO-induced initial elementary reactions with NDMA and other nitrosamines as well as the UV-induced NDMA degradation pathways at 254 nm of wavelength. We investigate NDMA, NMEA, and N-nitrosomethylbutylamine (NMBA), which have different alkyl side chains that are adjacent to the nitroso functional group (–N–NO), to elucidate the effect of the alkyl side chain on the overall reactivity with HO. We also investigate UV-induced NDMA degradation using time-dependent (TD)-DFT to understand the molecular orbitals responsible for electron excitation and the nitrogen-containing radical reactions during the photolysis of NDMA.

2. Results and Discussion

2.1. HO-Induced Degradation

2.1.1. N-Nitrosodimethylamine (NDMA) Degradation Pathways Induced by HO

NDMA has three potential initial degradation mechanisms: (1) H atom abstraction from a C–H bond of the methyl group (pathway 1–1 in Figure 1), (2) HO addition to amine nitrogen (pathway 1–2 in Figure 2), and (3) HO addition to nitrosyl nitrogen (pathway 1–3 in Figure 3). Our QM calculations obtained Δ G aq , calc act values of 9.7 kcal/mol, 6.8 kcal/mol, and 9.6 kcal/mol for the respective pathways. H abstraction from a C–H bond forms a C-centered radical that reacts with the triplet state of molecular oxygen dissolved in water. Our previous studies indicate that the addition of molecular oxygen to a C-centered radical is a barrierless reaction with a Δ G aq , calc act of −20–30 kcal/mol, which enabled us to consistently predict the experimentally measured reaction rate constants [28]. The Δ G aq , calc act value obtained for the CH2NNOCH3 radical was 2.3 kcal/mol, which is significantly larger than those of typically observed reactions. This indicates that the N–NO functional group significantly affects molecular addition to the C-centered radical. The second-order reaction rate constant for the addition of molecular oxygen to a C-centered radical of NDMA was determined to be (5.3 ± 0.6) × 106 M−1s−1 [9], which is three orders of magnitude smaller than the typically observed rate constants (~5 × 109 M−1s−1) [29]. A more detailed discussion on the unique reactivity of molecular oxygen to C-centered radicals will be given in a later section. According to our calculations, the C-centered radical also undergoes electron transfer to produce CH3NNO=CH2 ( Δ G aq , calc act of −2.0 kcal/mol), followed by the loss of NO ( Δ G aq , calc act of −11.3 kcal/mol) to produce N-methylidenemethylamine (CH2=NHCH3). This latter pathway involves several barrierless reactions, and is dominant over the pathway involving the addition of molecular oxygen. The formation of N-methylidenemethylamine was also postulated in a previous report [18,19].
The second pathway is HO addition to the amine nitrogen, followed by the loss of an OH group. Although initial HO addition has a lower free energy of activation ( Δ G aq , calc act of 6.8 kcal/mol) than the H abstraction identified in pathway 1–1, the subsequent reaction has a larger activation barrier ( Δ G aq , calc act of 3.1 kcal/mol) to produce a N-centered radical (i.e., CH3NCH3). The N-centered radical undergoes either molecular oxygen addition or an H shift. The H shift has a significantly smaller Δ G aq , calc act of −1.9 kcal/mol than molecular oxygen addition to the N-centered radical ( Δ G aq , calc act of 9.8 kcal/mol). Thus, C-centered radical formation resulting from an H shift is the dominant pathway via TS8. The significantly large Δ G aq , calc act for the addition of molecular oxygen to a N-centered radical via TS7 can be verified by the experimentally obtained reaction rate constant for hydrazyl (k = 3.9 × 108 M−1s−1) [30].
Pathway 1–3 involves initial HO addition to the nitrosyl nitrogen with a Δ G aq , calc act of 9.6 kcal/mol. Although this reaction has an almost identical Δ G aq , calc act to that of pathway 1–1, the initial HO addition reaction that produces an alkoxyl radical (i.e., CH3NNO(OH)CH3) is not thermodynamically favored ( Δ G aq , calc react of 6.4 kcal/mol). This alkoxyl radical undergoes two pathways to produce (1) a N-centered radical with a Δ G aq , calc act of 3.1 kcal/mol and (2) methyl diamine (CH3NHCH3) with a Δ G aq , calc act of −8.0 kcal/mol.
The above investigation confirms that H abstraction from a C–H bond of the methyl functional group of NDMA is the dominant initial reaction pathway as induced by HO, which is consistent with the experimental investigation using the electron paramagnetic resonance (ESR) technique [9]. The experimentally determined second-order rate constant was (4.3 ± 0.12) × 108 M−1s−1, and this relatively slow H abstraction from a C–H bond by HO results from the electron-withdrawing effect of the neighboring N–NO functional group and the abnormally stable C-centered radical [9]. In the following sub-sections, the reactivity of NDMA will be compared to two other nitrosamines that have longer alkyl side chains (i.e., -CH2CH3 and -(CH2)2CH3) to investigate the unique reactivity of NDMA.

2.1.2. N-Nitrosomethylethylamine (NMEA) Degradation Pathways Induced by HO

NMEA has three potential H abstraction sites: (1) a C–H bond of the –CH2– functional group adjacent to the N–NO functional group by pathway 2–1; (2) a C–H bond of the terminal CH3 functional group in the ethyl chain by pathway 2–2; and (3) a C–H bond of the terminal CH3 functional group adjacent to the N–NO functional group by pathway 2–3. Figure 4, Figure 5 and Figure 6 show the free energy profiles per reaction coordinate for each pathway. Our calculations revealed similar Δ G aq , calc act values for H atom abstraction: 11.1 kcal/mol in pathway 2–1 and 11.7 kcal/mol in pathway 2–3), except 62.7 kcal/mol in pathway 2–2. It is still not clear why the pathway 2–2 had such a high barrier. All three pathways are thermodynamically favorable ( Δ G aq , calc react < 0). Each pathway produces a C-centered radical, i.e., CH3CHNNOCH3 in pathway 2–1, CH2CH2NNOCH3 in pathway 2–2, and CH3 CH2NNOCH2 in pathway 2–3. The Δ G aq , calc act values for the addition of molecular oxygen to CH3CHNNOCH3, CH2CH2NNOCH3, and CH3 CH2NNOCH2 are 3.8 kcal/mol, −13.9 kcal/mol, and −2.2 kcal/mol, respectively. As observed in pathway 1, the Δ G aq , calc act values of these three C-centered radicals are still larger than the typical values (−20 to −25 kcal/mol). This indicates that the functional group directly neighboring the N–NO functional group affects the slow reaction of molecular oxygen addition to CH2CH2NNOCH3. Given that the other reaction pathways of the three C-centered radicals have either a larger Δ G aq , calc act than that for molecular oxygen addition or include thermodynamically unfavorable reactions ( Δ G aq , calc react > 0), the formation of peroxyl radicals resulting from the addition of molecular oxygen is the dominant reaction pathway in the subsequent NMEA degradation mechanism.

2.1.3. N-Nitrosomethylbutylamine (NMBA) Degradation Pathways Induced by HO

NMBA has four potential H abstraction sites from C–H bonds by HO: (1) a C–H bond of the –CH2– functional group adjacent to the N–NO functional group by pathway 3–1; (2) a C–H bond of the –CH2 functional group adjacent to the –CH2– functional groups on both sides by pathway 3–2; (3) a C–H bond of the terminal CH3 functional group in a butyl chain by pathway 3–3; and (4) a C–H bond of the terminal CH3 functional group that is adjacent to the N–NO functional group by pathway 3–4. Figure 7, Figure 8, Figure 9 and Figure 10 show the free energy profiles per reaction coordinate for each pathway. The calculated Δ G aq , calc act values are 10.2 kcal/mol for pathway 3–1, 8.3 kcal/mol for pathway 3–2, 10.9 kcal/mol for pathway 3–3, and 11.9 kcal/mol for pathway 3–4. The smaller Δ G aq , calc act value for pathway 3–2 compared with those for NDMA and NDEA shows consistent reactivity with the experimentally obtained rate constants: 109 M−1s−1 for N-nitrosobutylamine, 4.3 × 108 M−1s−1 for NDMA and 4.95 × 108 M−1s−1 for NMEA [8]. The initial H abstraction reactions for all of the pathways are thermodynamically favorable.
Interestingly, we observed distinctive differences in the reactivity of molecular oxygen addition to different C-centered radicals for NMBA. The initial H abstraction from different C–H bonds in NMBA produced CH3NNOCHCH2CH3 by pathway 3–1, CH3NNOCH2CHCH3 by pathway 3–2, CH3NNO(CH2)2CH2 by pathway 3–3, and CH2NNO(CH2)2CH3 by pathway 3–4. While molecular oxygen addition to CH3NNOCHCH2CH3 and CH2NNO(CH2)2CH3 have larger Δ G aq , calc act values of 4.2 kcal/mol and −12.4 kcal/mol, the Δ G aq , calc act values for CH3NNOCH2CHCH3 (−25.6 kcal/mol) and CH3NNO(CH2)2CH2 (−23.9 kcal/mol) are very similar to those that were observed for typical molecular oxygen addition to C-centered radicals. Thus, the alkyl side chain affects the stability of the C-centered radicals and their subsequent reactivity. The significantly slower reaction of molecular oxygen addition to the C-centered radicals produced from NDMA and NMEA may be due to the delocalization of the radical spin density from the formed C-centered radicals onto the N–NO bond(s). This radical delocalization occurs only when a terminal CH2 is adjacent to N–NO or CH2 neighbors the N–NO functional group. When the alkyl chain contains three CH2 functional groups, the CH2 three positions away from the N–NO functional group does not seem to contribute to the radical delocalization. Thus, the molecular oxygen adds to the C-centered radical without being affected by the delocalization. The different extent of radical delocalization can also explain the lower degradation efficiencies that were observed for NDMA and NEMA (approximately 80~85% degradation efficiency) as compared with nitrosodibutylamine (100% degradation efficiency) [8].
To investigate the effect of the location of the C-centered radical on the occurrence of radical delocalization, we calculated the Δ G aq , calc act values for radical transfer from a C-centered radical to a neighboring C-/N-centered radical. For example, CH3NNOCHCH2CH3 undergoes radical transfer from a carbon to the amine nitrogen to produce CH3NNO=CHCH2CH3. This reaction has a Δ G aq , calc act of 0.41 kcal/mol, which indicates a low barrier for this radical delocalization (pathway 3–1). Similarly, CH2NNO(CH2)2CH3 requires 3.7 kcal/mol to produce CH2=NNO=CHCH2CH3 (pathway 3–4). In contrast, CH3NNOCH2CHCH3 requires a Δ G aq , calc act of 38.6 kcal/mol to produce CH3NNOCHCH2CH3 (pathway 3–2). A similar significantly larger Δ G aq , calc act value of 40.0 kcal/mol was also observed for the radical transfer reaction from CH2CH2NNOCH3 to CH3CHNNOCH3 via pathway 2–2. Thus, there is a significant barrier for radical transfer from the functional group neighboring the N–NO functional group to the nearest CH2 group. Therefore, a C-centered radical in CH2CH2NNOCH3 or CH3NNOCH2CHCH3 would rather undergo molecular oxygen addition than radical transfer to produce CH3CHNNOCH3 or CH3NNOCHCH2CH3, respectively.

2.2. UV-Induced Degradation

NDMA Degradation Pathways Induced by UV Photolysis

NDMA absorbs photons at a wavelength of 228 nm with a molar absorptivity of 7380 M−1cm−1 and quantum yield of 0.13 at pH 7 [7]. At a wavelength of 253.7 nm, where a typical low-pressure UV lamp emits photons, the molar absorptivity was reported to be 1650 M−1s−1, and the quantum yield was 0.24 at pH 7 [7]. Another smaller peak is observed at approximately 350 nm. Our TD-DFT calculation obtained one major and one minor peak at 212 nm and 341 nm, respectively. The molecular orbitals that were responsible for the π→π* and n→π transitions at 212 nm and 341 nm are shown in Figure 11. At 212 nm, the N–N bond comprises the highest occupied molecular orbital (HOMO), whereas the C–N bond comprises the HOMO at 341 nm. This analysis indicates that the N–N bond is susceptible breakage under photolysis with a low-pressure UV lamp. This finding is consistent with the experimental findings that were reported in the previous literature.
The UV photolysis-induced NDMA degradation pathways were extensively studied [15,16]. According to their studies, NDMA undergoes three major degradation pathways induced by UV photolysis: (1) formation of an aminium radical [(CH3)2N(+)H] and nitric oxide (NO) resulting from homolytic cleavage of the N–N bond (pathway 4–1 in Figure 12); (2) formation of dimethylamine [(CH3)2NH2+] and nitrous acid (HNO2) resulting from heterolytic photocleavage of the N–N bond facilitated by a water molecule (pathway 4–2 in Figure 13); and (3) formation of N-methylidenemethylamine [(CH2=N(+)HCH3], NO, and a superoxide anion radical (O2) in the presence of dissolved oxygen (i.e., triplet state of 3O2) (pathway 4–3 in Figure 14).
The products of (CH3)2N(+)H and NO in pathway 4–1 react in a solvent cage to produce N-methylidenemethylamine [(CH2=N(+)HCH3] and nitroxyl (HNO). Our calculation obtained a Δ G aq , calc act of 1.6 kcal/mol for this reaction. Then, N-methylidenemethylamine undergoes rapid hydrolysis to produce methylamine (CH3NH2+) and formaldehyde (HCHO). A total of 99% of the HCHO is hydrolyzed to form a germinal diol in the aqueous phase [30]; therefore, the hydrated form of HCHO (i.e., CH2(OH)2) exists in the aqueous phase. CH2(OH)2 reacts with HO via H abstraction to produce CH(OH)2 with a Δ G aq , calc act of 10.0 kcal/mol. As was examined in the HO-induced pathways, this C-centered radical reacts with molecular oxygen to produce a peroxyl radical (i.e., OOCH(OH)2) ( Δ G aq , calc act of −34.9 kcal/mol). The peroxyl radical undergoes uni/bimolecular decay to produce stable lower-molecular-weight products [31]. When OOCH(OH)2 undergoes unimolecular decay (i.e., HO2 elimination), formic acid (HCOOH) is produced ( Δ G aq , calc act of 31.6 kcal/mol), which has been experimentally observed [32].
One of the C–H bonds in the methyl group of the dimethylamine produced in pathway 4–2 undergoes H abstraction by HO to produce a C-centered radical with a Δ G aq , calc act of 13.9 kcal/mol. Molecular oxygen adds to the C-centered radical to produce a peroxyl radical with a Δ G aq , calc act of −15.0 kcal/mol, and the peroxyl radical undergoes subsequent uni/bimolecular decay.
The products of NO and O2- from pathway 4–3 react in a solvent cage to produce peroxynitrite (ONOO) with a Δ G aq , calc act of 1.72 kcal/mol. The rate constant for this reaction was determined to be (4.3 − 7.6) × 109 M−1s−1 [32,33,34]. Then, ONOO undergoes rearrangement with a Δ G aq , calc act of 57.8 kcal/mol to produce a nitrate ion (NO3). This rearrangement was proposed as isomerization by Anbar and Taube (1954) [35]. ONOO also reacts with HO2/O2•− via single electron transfer to produce an OONO radical. Our calculation indicates that this reaction is barrierless, with a Δ G aq , calc act of −16.2 kcal/mol, but the reaction is not thermodynamically favorable ( Δ G aq , calc react of 3.4 kcal/mol). Finally, the OONO radical undergoes cleavage with a Δ G aq , calc react of −0.56 kcal/mol to produce NO.
When nitrate undergoes UV photolysis, a nitrite ion (NO2) and NO2 are produced. Then, NO2 reacts with HO, O2•−, or NO2 with a Δ G aq , calc react of 48.3 kcal/mol, 40.2 kcal/mol, or 100.6 kcal/mol to produce ONOOH, NO2-/NO3, or N2O4, respectively (Figure 15). Although the disproportionation of NO2 has the largest free energy barrier, the reaction product, N2O4, undergoes hydrolysis to produce NO3 and NO2.

2.3. Environmental Implication and Future Study

Nitrosamines, and NDMA in particular, are extremely potent carcinogenic contaminants in water. The concentration at which NDMA shows potent carcinogenicity is extremely low (0.7 ng/L) [1]. Experimentally investigating the ng/L fate of many chemical contaminants during water treatment processes is time consuming and expensive. Our computational study highlights the usefulness of QM calculations to reveal the elementary reaction pathways that are embedded in the overall reaction pathways that are typically identified by analytical techniques. This technique becomes more useful when the contaminant concentrations are below the analytical detection limit.
Once the elementary reaction pathways are identified, the reaction rate constants should be determined or predicted to calculate the reaction rate of each molecule or species involved in each elementary reaction step. By combining the elementary reaction pathways and the reaction rate constants, one can predict the time-dependent concentration profiles of a target chemical contaminant and its transformation products. This elementary-reaction-based kinetic model could be used as an initial screening tool for many potentially toxic chemical contaminants to estimate the fate of degradation pathways. Our efforts towards the development of such elementary-reaction-based kinetic model are underway.

3. Materials and Methods

All of the QM calculations were performed with the Gaussian 09 revision D.02 program [36] using the Michigan Tech high-performance cluster “Superior” and homemade LINUX workstations. The M06-2X/cc-pVDZ [37] was used to optimize the electronic structures and calculate the frequencies in both the gas and aqueous phase for the HO-induced reaction pathways with NDMA, NMEA, and NMBA. The UV-induced reaction pathways with NDMA was calculated with the Gaussian-4 theory (G4) [38]. The aqueous-phase structures and frequencies were obtained using an implicit polarizable continuum model [universal solvation model (SMD)] [39]. Previously, we verified the combination of M06-2X/cc-pVDZ or G4 with the SMD model by successfully applying it to other aqueous-phase radical-involved reactions [27,28]. Theoretically calculated absorption spectra were obtained from a TD-DFT analysis [40,41] of the optimized aqueous-phase structure of NDMA at the level of M06-2X/cc-pVDZ with the SMD solvation model. To investigate the contributions from molecular orbitals to the peak of the spectra, molecular orbitals were determined using a natural population analysis at the level of M06-2X/cc-pVQZ with the SMD solvation model. The detailed calculation procedures for the transition state search, the aqueous-phase free energies of activation and reaction, and the associated computational methods are found in previous reports [29].

Supplementary Materials

Supplementary materials are available on line.

Acknowledgments

This work was supported by the National Science Foundation Award: CBET-1435926. Any opinions, findings, conclusions, or recommendations expressed in this publication are those of the authors and do not necessarily reflect the view of the supporting organization.

Author Contributions

D.M. and E.C. have the same contribution.

Conflicts of Interest

The authors declare no conflict of interest.

List of Symbols and Abbreviations

AOPsadvanced oxidation processes
CH2=N(+)HCH3N-methylidenemethylamine
CH3NHCH3methyl diamine
DFTdensity functional theory
Δ G aq , calc act theoretically calculated aqueous phase free energy of activation
Δ G aq , calc react theoretically calculated aqueous phase free energy of reaction
G4Gaussian-4 theory
HCHOformaldehyde
HCOOHformic acid
HNO2nitrous acid
HOhydroxyl radicals
HOMOhighest occupied molecular orbital
NFnanofiltration
NDEAN-nitrosodiethylamine
NDMAN-nitrosodimethylamine
NMEAN-nitrosomethylethylamine
NOnitric oxide
NO3nitrate ion
NO2nitrite ion
ONOOperoxynitrite
O2superoxide anion radical
QMquantum mechanical
ROreverse osmosis
SMDuniversal solvation model
TD-DFTtime-dependent density functional theory
TStransition state
UVultraviolet

References

  1. US EPA. Technical Fact Sheet N-Nitroso-Dimethylamine (NDMA); EPA 505-F-17-005, November 2017; EPA: Washington, DC, USA, 2017.
  2. Fine, D.H.; Rounbehler, D.P.; Rounbehler, A.; Silvergleid, A.; Sawicki, E.; Krost, K.; Demarrais, G.A. Determination of dimethylnitrosamine in air, water, and soil by thermal-energy analysis- measurements in baltimore, md. Environ. Sci. Technol. 1977, 11, 581–584. [Google Scholar] [CrossRef]
  3. California Environmental Protection Agency, State Water Resrouces Control Board. NDMA and Other Nitrosamines-Drinking Water Issues. Available online: https://www.waterboards.ca.gov/drinking_water/certlic/drinkingwater/NDMA.shtml (accessed on 6 February 2018).
  4. Glaze, W.H.; Kang, J.W. Advanced oxidation processes- test of a kinetic-model for the oxidation of organic-compounds with ozone and hydrogen-peroxide in a semibatch reactor. Ind. Eng. Chem. Res. 1989, 28, 1580–1587. [Google Scholar] [CrossRef]
  5. Glaze, W.H.; Kang, J.W.; Chapin, D.H. The chemistry of water-treatment processes involving ozone, hydrogen-peroxide and ultraviolet-radiation. Ozone Sci. Eng. 1987, 9, 335–352. [Google Scholar] [CrossRef]
  6. Tchobanoglous, G. Framework for Direct Potable Reuse; WateReuse Research Foundation: Alexandria, VA, USA, 2015. [Google Scholar]
  7. Mitch, W.A.; Sharp, J.O.; Trussell, R.R.; Valentine, R.L.; Alvarez-Cohen, L.; Sedlak, D.L. N-nitrosodimethylamine (NDMA) as a drinking water contaminant: A review. Environ. Eng. Sci. 2003, 20, 389–404. [Google Scholar] [CrossRef]
  8. Landsman, N.A.; Swancutt, K.L.; Bradford, C.N.; Cox, C.R.; Kiddle, J.J.; Mezyk, S.P. Free radical chemistry of advanced oxidation process removal of nitrosamines in water. Environ. Sci. Technol. 2007, 41, 5818–5823. [Google Scholar] [CrossRef] [PubMed]
  9. Mezyk, S.P.; Cooper, W.J.; Madden, K.P.; Bartels, D.M. Free radical destruction of N-nitrosodimethylamine in water. Environ. Sci. Technol. 2004, 38, 3161–3167. [Google Scholar] [CrossRef] [PubMed]
  10. Chow, Y.L. Nitrosamine photochemistry-reactions of aminium radicals. Acc. Chem. Res. 1973, 6, 354–360. [Google Scholar] [CrossRef]
  11. Chow, Y.L.; Tam, J.N.S.; Lau, M.P.; Perry, R.A. Photoreactions of nitroso-compounds in solution. XX. Photoreduction, photoelimination, and photoaddition of nitrosamines. Can. J. Chem. 1972, 50, 1044–1050. [Google Scholar] [CrossRef]
  12. Daiber, D.; Preussmann, R. Quantitative colorimetrische bestimmung organischer n-nitroso-verbindungen durch photochemische spaltung der nitrosaminbindung (quantitative colorimetric determination of organic n-nitroso compounds by photochemical splitting of the nitroso amine bond). Fresenius Z. Anal. Chem. 1964, 206, 344–352. [Google Scholar] [CrossRef]
  13. Grilli, S.; Tosi, M.R.; Prodi, G. Degradation of dimethylnitrosoamine catalyzed by physical and chemical agents. Gann 1975, 66, 481–488. [Google Scholar] [PubMed]
  14. Stefan, M.I.; Bolton, J.R. Uv direct photolysis of N-nitrosodimethylamine (NDMA): Kinetic and product study. Helv. Chim. Acta 2002, 85, 1416–1426. [Google Scholar] [CrossRef]
  15. Lee, C.; Choi, W.; Kim, Y.G.; Yoon, J. Uv photolytic mechanism of N-nitrosodimethylamine in water: Dual pathways to methylamine versus dimethylamine. Environ. Sci. Technol. 2005, 39, 2101–2106. [Google Scholar] [CrossRef] [PubMed]
  16. Lee, C.; Choi, W.; Yoon, J. Uv photolytic mechanism of N-nitrosodimethylamine in water: Roles of dissolved oxygen and solution ph. Environ. Sci. Technol. 2005, 39, 9702–9709. [Google Scholar] [CrossRef] [PubMed]
  17. Kwon, B.G.; Kim, J.-O.; Namkung, K.C. The formation of reactive species having hydroxyl radical-like reactivity from UV photolysis of N-nitrosodimethylamine (NDMA): Kinetics and mechanism. Sci. Total Environ. 2012, 437, 237–244. [Google Scholar] [CrossRef] [PubMed]
  18. Kwon, B.G.; Kim, J.-O.; Namkung, K.C. Formation of reactive species enhanced by H2O2 addition in the photodecomposition of N-nitrosodimethylamine (NDMA). Environ. Eng. Res. 2013, 18, 29–35. [Google Scholar] [CrossRef]
  19. Lee, C.; Yoon, J.; Von Gunten, U. Oxidative degradation of N-nitrosodimethylamine by conventional ozonation and the advanced oxidation process ozone/hydrogen peroxide. Water Res. 2007, 41, 581–590. [Google Scholar] [CrossRef] [PubMed]
  20. Lv, J.; Li, Y.M.; Song, Y. Reinvestigation on the ozonation of N-nitrosodimethylamine: Influencing factors and degradation mechanism. Water Res. 2013, 47, 4993–5002. [Google Scholar] [CrossRef] [PubMed]
  21. Xiao, R.Y.; Noerpel, M.; Luk, H.L.; Wei, Z.S.; Spinney, R. Thermodynamic and kinetic study of ibuprofen with hydroxyl radical: A density functional theory approach. Int. J. Quantum Chem. 2014, 114, 74–83. [Google Scholar] [CrossRef]
  22. An, T.C.; Gao, Y.P.; Li, G.Y.; Kamat, P.V.; Peller, J.; Joyce, M.V. Kinetics and mechanisms of •OH mediated degradation of dimethyl phthalate in aqueous solution: Experimental and theoretical studies. Environ. Sci. Technol. 2014, 48, 641–648. [Google Scholar] [CrossRef] [PubMed]
  23. Liu, S.Q.; Zhou, X.Z.; Han, W.Q.; Li, J.S.; Sun, X.Y.; Shen, J.Y.; Wang, L.J. Theoretical and experimental insights into the •OH-medited mineralization mechanism of flutriafol. Electrochim. Acta 2017, 235, 223–232. [Google Scholar] [CrossRef]
  24. Trogolo, D.; Mishra, B.K.; Heeb, M.B.; von Gunten, U.; Arey, J.S. Molecular mechanism of ndma formation from N,N-dimethylsulfamide during ozonation: Quantum chemical insights into a bromide-catalyzed pathway. Environ. Sci. Technol. 2015, 49, 4163–4175. [Google Scholar] [CrossRef] [PubMed]
  25. Minakata, D.; Crittenden, J. Linear free energy relationships between aqueous phase hydroxyl radical reaction rate constants and free energy of activation. Environ. Sci. Technol. 2011, 45, 3479–3486. [Google Scholar] [CrossRef] [PubMed]
  26. Minakata, D.; Song, W.H.; Crittenden, J. Reactivity of aqueous phase hydroxyl radical with halogenated carboxylate anions: Experimental and theoretical studies. Environ. Sci. Technol. 2011, 45, 6057–6065. [Google Scholar] [CrossRef] [PubMed]
  27. Minakata, D.; Song, W.H.; Mezyk, S.P.; Cooper, W.J. Experimental and theoretical studies on aqueous-phase reactivity of hydroxyl radicals with multiple carboxylated and hydroxylated benzene compounds. Phys. Chem. Chem. Phys. 2015, 17, 11796–11812. [Google Scholar] [CrossRef] [PubMed]
  28. Minakata, D.; Mezyk, S.P.; Jones, J.W.; Daws, B.R.; Crittenden, J.C. Development of linear free energy relationships for aqueous phase radical-involved chemical reactions. Environ. Sci. Technol. 2014, 48, 13925–13932. [Google Scholar] [CrossRef] [PubMed]
  29. Buxton, G.V.; Stuart, C.R. Radiation chemistry of aqueous solutions of hydrazine at elevated temperatures. J. Chem. Soc. Faraday Trans. 1997, 93, 1535–1538. [Google Scholar] [CrossRef]
  30. McMurry, J. Nucleophilic addition of H2O: Hydration. In Organic Chemistry; Cengage Learning: Boston, MA, USA, 2015. [Google Scholar]
  31. Bothe, E.; Schultefrohlinde, D. Reaction of dihydroxymethyl radical with molecular-oxygen in aqueous-solution. Z. Naturforsch. B 1980, 35, 1035–1039. [Google Scholar] [CrossRef]
  32. Von Sonntag, C.; Schuchmann, H.P. The elucidation of peroxyl radical reactions in aqueous-solution with the help of radiation-chemical methods. Angew. Chem. Int. Ed. 1991, 30, 1229–1253. [Google Scholar] [CrossRef]
  33. Padmaja, S.; Huie, R.E. The reaction of nitric-oxide with organic peroxyl radicals. Biochem. Biophys. Res. Commun. 1993, 195, 539–544. [Google Scholar] [CrossRef] [PubMed]
  34. Goldstein, S.; Czapski, G. The reaction of NO with O2 and HO: A pulse radiolysis study. Free Radic. Biol. Med. 1995, 19, 505–510. [Google Scholar] [CrossRef]
  35. Anbar, M.; Taube, H. Interaction of nitrous acid with hydrogen peroxide and with water. J. Am. Chem. Soc. 1954, 76, 6243–6247. [Google Scholar] [CrossRef]
  36. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09, Revision D.02; Gaussian Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  37. Zhao, Y.; Truhlar, D.G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar]
  38. Curtiss, L.A.; Redfern, P.C.; Raghavachari, K. Gaussian-4 theory. J. Chem. Phys. 2007, 126, 84108. [Google Scholar] [CrossRef] [PubMed]
  39. Marenich, A.V.; Cramer, C.J.; Truhlar, D.G. Universal solvation model based on solute electron density and a continuum model of the solvent defined by the bulk dielectric constant and atomic surface tensions. J. Phys. Chem. B 2009, 113, 6378–6396. [Google Scholar] [CrossRef] [PubMed]
  40. Adamo, C.; Jacquemin, D. The calculations of excited-state properties with Time-Dependent Density Functional Theory. Chem. Soc. Rev. 2013, 42, 845–856. [Google Scholar] [CrossRef] [PubMed]
  41. Laurent, A.D.; Adamo, C.; Jacquemin, D. Dye chemistry with time-dependent density functional theory. Phys. Chem. Chem. Phys. 2014, 16, 1434–14356. [Google Scholar] [CrossRef] [PubMed]
Sample Availability: Samples of the compounds are not available from the authors.
Figure 1. Free energy profile for pathway 1–1 of the HO-induced reaction pathways for N-nitrosodimethylamine (NDMA) via H abstraction. TS denotes the transition state, and P denotes the product. The numbers (kcal/mol) are the free energy of activation for the TS and free energy of reaction for the P relative to the corresponding reactant.
Figure 1. Free energy profile for pathway 1–1 of the HO-induced reaction pathways for N-nitrosodimethylamine (NDMA) via H abstraction. TS denotes the transition state, and P denotes the product. The numbers (kcal/mol) are the free energy of activation for the TS and free energy of reaction for the P relative to the corresponding reactant.
Molecules 23 00539 g001
Figure 2. Free energy profile for pathway 1–2 of the HO-induced reaction pathways for N-nitrosodimethylamine (NDMA) via HO addition to amine nitrogen. The numbers (kcal/mol) are the free energy of activation for the TS and free energy of reaction for the P relative to the corresponding reactant.
Figure 2. Free energy profile for pathway 1–2 of the HO-induced reaction pathways for N-nitrosodimethylamine (NDMA) via HO addition to amine nitrogen. The numbers (kcal/mol) are the free energy of activation for the TS and free energy of reaction for the P relative to the corresponding reactant.
Molecules 23 00539 g002
Figure 3. Free energy profile for pathway 1–3 of the HO-induced reaction pathways for N-nitrosodimethylamine (NDMA) via HO addition to the nitrosyl nitrogen. The numbers (kcal/mol) are the free energy of activation for the TS and free energy of reaction for the P relative to the corresponding reactant.
Figure 3. Free energy profile for pathway 1–3 of the HO-induced reaction pathways for N-nitrosodimethylamine (NDMA) via HO addition to the nitrosyl nitrogen. The numbers (kcal/mol) are the free energy of activation for the TS and free energy of reaction for the P relative to the corresponding reactant.
Molecules 23 00539 g003
Figure 4. Free energy profile for pathway 2–1 of the HO-induced reaction pathways for NMEA via H abstraction from a C–H bond of the –CH2– functional group adjacent to the N–NO functional group. The numbers (kcal/mol) are the free energy of activation for the TS and free energy of reaction for the P relative to the corresponding reactant.
Figure 4. Free energy profile for pathway 2–1 of the HO-induced reaction pathways for NMEA via H abstraction from a C–H bond of the –CH2– functional group adjacent to the N–NO functional group. The numbers (kcal/mol) are the free energy of activation for the TS and free energy of reaction for the P relative to the corresponding reactant.
Molecules 23 00539 g004
Figure 5. Free energy profile for pathway 2–2 of the HO-induced reaction pathways for NMEA via H abstraction from a C–H bond of the terminal CH3 functional group in the ethyl chain. The numbers (kcal/mol) are the free energy of activation for the TS and free energy of reaction for the P relative to the corresponding reactant.
Figure 5. Free energy profile for pathway 2–2 of the HO-induced reaction pathways for NMEA via H abstraction from a C–H bond of the terminal CH3 functional group in the ethyl chain. The numbers (kcal/mol) are the free energy of activation for the TS and free energy of reaction for the P relative to the corresponding reactant.
Molecules 23 00539 g005
Figure 6. Free energy profile for pathway 2–3 of the HO-induced reaction pathways for NMEA via H abstraction from a C–H bond of the terminal CH3 functional group adjacent to the N–NO functional group. The numbers (kcal/mol) are the free energy of activation for the TS and free energy of reaction for the P relative to the corresponding reactant.
Figure 6. Free energy profile for pathway 2–3 of the HO-induced reaction pathways for NMEA via H abstraction from a C–H bond of the terminal CH3 functional group adjacent to the N–NO functional group. The numbers (kcal/mol) are the free energy of activation for the TS and free energy of reaction for the P relative to the corresponding reactant.
Molecules 23 00539 g006
Figure 7. Free energy profile for pathway 3–1 of the HO-induced reaction pathways for NMBA via H abstraction from a C–H bond of the –CH2– functional group adjacent to the N–NO functional group. The numbers (kcal/mol) are the free energy of activation for the TS and free energy of reaction for the P relative to the corresponding reactant.
Figure 7. Free energy profile for pathway 3–1 of the HO-induced reaction pathways for NMBA via H abstraction from a C–H bond of the –CH2– functional group adjacent to the N–NO functional group. The numbers (kcal/mol) are the free energy of activation for the TS and free energy of reaction for the P relative to the corresponding reactant.
Molecules 23 00539 g007
Figure 8. Free energy profile for pathway 3–2 of the HO-induced reaction pathways for NMBA via H abstraction from a C–H bond of the -CH2 functional group adjacent to the –CH2– functional groups on both sides. The numbers (kcal/mol) are the free energy of activation for the TS and free energy of reaction for the P relative to the corresponding reactant.
Figure 8. Free energy profile for pathway 3–2 of the HO-induced reaction pathways for NMBA via H abstraction from a C–H bond of the -CH2 functional group adjacent to the –CH2– functional groups on both sides. The numbers (kcal/mol) are the free energy of activation for the TS and free energy of reaction for the P relative to the corresponding reactant.
Molecules 23 00539 g008
Figure 9. Free energy profile for pathway 3–3 of the HO-induced reaction pathways for NMBA via H abstraction from a C–H bond of the terminal CH3 functional group in a butyl chain. The numbers (kcal/mol) are the free energy of activation for the TS and free energy of reaction for the P relative to the corresponding reactant.
Figure 9. Free energy profile for pathway 3–3 of the HO-induced reaction pathways for NMBA via H abstraction from a C–H bond of the terminal CH3 functional group in a butyl chain. The numbers (kcal/mol) are the free energy of activation for the TS and free energy of reaction for the P relative to the corresponding reactant.
Molecules 23 00539 g009
Figure 10. Free energy profile for pathway 3–4 of the HO-induced reaction pathways for NMBA via H abstraction from a C–H bond of the terminal CH3 functional group adjacent to the N–NO functional group. The numbers (kcal/mol) are the free energy of activation for the TS and free energy of reaction for the P relative to the corresponding reactant.
Figure 10. Free energy profile for pathway 3–4 of the HO-induced reaction pathways for NMBA via H abstraction from a C–H bond of the terminal CH3 functional group adjacent to the N–NO functional group. The numbers (kcal/mol) are the free energy of activation for the TS and free energy of reaction for the P relative to the corresponding reactant.
Molecules 23 00539 g010
Figure 11. HOMO and lowest unoccupied molecular orbital (LUMO) of the π→π* (a) and n→π (b) transitions at 212 nm and 341 nm, respectively.
Figure 11. HOMO and lowest unoccupied molecular orbital (LUMO) of the π→π* (a) and n→π (b) transitions at 212 nm and 341 nm, respectively.
Molecules 23 00539 g011
Figure 12. Free energy profile for pathway 4–1 of the HO-induced reaction pathways for NDMA photolysis.
Figure 12. Free energy profile for pathway 4–1 of the HO-induced reaction pathways for NDMA photolysis.
Molecules 23 00539 g012
Figure 13. Free energy profile for pathway 4–2 of the HO-induced reaction pathways for NDMA photolysis.
Figure 13. Free energy profile for pathway 4–2 of the HO-induced reaction pathways for NDMA photolysis.
Molecules 23 00539 g013
Figure 14. Free energy profile for pathway 4–3 of the HO-induced reaction pathways for NDMA photolysis.
Figure 14. Free energy profile for pathway 4–3 of the HO-induced reaction pathways for NDMA photolysis.
Molecules 23 00539 g014
Figure 15. Free energy profile for the reaction of NO2• with HO, O2•−, and NO2.
Figure 15. Free energy profile for the reaction of NO2• with HO, O2•−, and NO2.
Molecules 23 00539 g015

Share and Cite

MDPI and ACS Style

Minakata, D.; Coscarelli, E. Mechanistic Insight into the Degradation of Nitrosamines via Aqueous-Phase UV Photolysis or a UV-Based Advanced Oxidation Process: Quantum Mechanical Calculations. Molecules 2018, 23, 539. https://doi.org/10.3390/molecules23030539

AMA Style

Minakata D, Coscarelli E. Mechanistic Insight into the Degradation of Nitrosamines via Aqueous-Phase UV Photolysis or a UV-Based Advanced Oxidation Process: Quantum Mechanical Calculations. Molecules. 2018; 23(3):539. https://doi.org/10.3390/molecules23030539

Chicago/Turabian Style

Minakata, Daisuke, and Erica Coscarelli. 2018. "Mechanistic Insight into the Degradation of Nitrosamines via Aqueous-Phase UV Photolysis or a UV-Based Advanced Oxidation Process: Quantum Mechanical Calculations" Molecules 23, no. 3: 539. https://doi.org/10.3390/molecules23030539

Article Metrics

Back to TopTop