Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2024 Jul 8;128(28):5548–5555. doi: 10.1021/acs.jpca.4c02433

Ab Initio Chemical Kinetics for Oxidation of CH3OH by N2O4: Elucidation of the Mechanism for Major Product Formation and Its Relevancy to Tropospheric Chemistry

Hue-Phuong Trac 1, Ming-Chang Lin 1,*
PMCID: PMC11264261  PMID: 38973582

Abstract

graphic file with name jp4c02433_0007.jpg

Next to CH4, CH3OH is the most abundant C1 organics in the troposphere. The redox reaction of CH3OH with N2O4 had been shown experimentally to produce CH3ONO, instead of CH3ONO2. The mechanism for the reaction remains unknown to date. We have investigated the reaction by ab initio MO calculations at the UCCSD(T)/6-311+G(3df,2p)//UB3LYP/6-311+G(3df,2p) level. The result indicates that the reaction takes place primarily by the isomerization of N2O4 to ONONO2 through a very loose transition state within the N2O4–CH3OH collision complex with a 14.3 kcal/mol barrier, followed by the rapid attack of ONONO2 at CH3OH producing CH3ONO and HNO3. The predicted mechanism for the redox reaction compares closely with the hydrolysis of N2O4. The computed rate constant, k1 = 1.43 × 10–8 T1.96 exp (−9092/T) (200–2000 K) cm3molecule–1s–1, for the formation of CH3ONO and HNO3 agrees reasonably with available low-temperature kinetic data and is found to be similar to that of the isoelectronic N2O4 + CH3NH2 reaction. We have also estimated the kinetics for the termolecular reaction, 2 NO2 + CH3OH, and compared it with the direct bimolecular process; the latter was found to be 4.4 × 105 times faster under the troposphere condition. On the basis of the known pollution levels of NO2, N2O4, and CH3OH, both processes were estimated to be of negligible importance to tropospheric chemistry, however.

1. Introduction

CH3OH is known to be an important organic pollutant in the troposphere with concentrations averaged to be 600 pptv, next to that of CH4.1 The origin of CH3OH is primarily of terrestrial rather than anthropogenic sources. In the troposphere, CH3OH may be removed by reactions with various known radicals such as OH and NO3; among them, the destruction by OH is dominant on account of its high reactivity and concentration in the troposphere.2 In addition, oxidation by larger nitrogen oxides, N2Ox (x = 4 and 5), is also possible under the tropospheric condition as discussed recently by Trac et al.3 and by Sarkar and Bandyopadhyay4 on the reactivities of N2O4 and N2O5 toward NH3, respectively. The rate of the former redox process was predicted to be about 100 times faster than that of the latter under the tropospheric condition, although the concentration of N2O4 is known to be lower than that of N2O5.3

The facile reaction of N2O4 with CH3OH at low temperatures was first reported by Harris and Siegel who noted the disappearance of NO2 upon mixing.5 Joffe and Gray identified methyl nitrite (CH3ONO) and nitric acid as the products of the reaction but not methyl nitrate (CH3ONO2).6 The first kinetic measurement for the N2O4 + CH3OH reaction and several other small alcohols was carried out by Fairlie et al.7 at 273–298 K following the NO2 decay kinetics by visible light absorption; they reported a negative temperature dependence, obeying the third-order rate law, – d[NO2]/dt = 2 k [NO2]2 [CH3OH], with k = 4.8 × 10–37 cm6/molecule2-s at 298 K. Nikki and co-workers8 investigated the reactions of NO2 with CH3OH and C2H5OH by FTIR spectroscopy monitoring the growth of RONO which followed the rate law, d[RONO]/dt = k [NO2]2 [ROH]. The rate constant for both alcohol reactions was reported to be the same, k = 5.7 × 10–37 cm6/molecule2-s at 298 K. Koda et al.9 employed a reactor with spray-mixing to improve the mixing of the two reactants and monitored the kinetics of NO2 decay photometrically; they reported the termolecular rate constant to be k = 1.79 × 10–36 cm6/molecule2-s at 294 K. Based on their mechanistic analysis of the NO2 decay kinetics, they postulated that the asymmetric N2O4, or ONONO2, might be involved in the reaction giving CH3ONO + HNO3, instead of the more abundant symmetric N2O4. More recently, Wojcik-Pastuszka et al.10 studied the reactions of N2O4 with CH3OH and several other small alcohols between 293 and 358 K by UV–vis spectroscopy, which allowed them to detect not only NO2 and N2O4 but also the RONO products. The kinetics of these reactions were determined by measurement of NO2 decay at 450 nm. They kinetically simulated the NO2 decay–time profiles with the mechanism: 2 NO2 ⇌ N2O4 and N2O4 + ROH ⇌ RONO + HNO3. The second-order rate constants for CH3ONO formation and its reverse reaction were reported in detail for the CH3OH reaction. To date, the mechanism for the N2O4reaction with CH3OH and other small alcohols remains unknown, however.

The reactivity of N2O4 toward NH3 and RNH2, including methyl amine3 (R = CH3) and hydrazines11,12 [R = NH2, CH3NH and (CH3)2N], has been investigated recently by quantum-chemical and statistical-theory calculations in our laboratory. The high reactivity of these reactions was attributed to the unique property of N2O4, which can undergo isomerization producing a highly reactive isomer ONONO2 via a roaming-like transition state (TS) during the course of bimolecular collisions by lengthening of the N–N bond (O2N–NO2) and rotating one of the O2N groups13 with barriers below the dissociation limit, ranging from 12.8 to 13.1 kcal/mol depending on collision partners.3,1114 Significantly, the roaming-like TS was also found to exist in the condensed phases including H2O13 and solid N2H4,14 in which the isomerization reaction produces the reactive ONONO2 isomer within the solid lattice producing N2H5+ONO2 with a large 70 kcal/mol exothermicity. The half-life of an embedded N2O4 in solid N2H4 at 218 K was predicted to be 0.5 s, which could reasonably explain the explosion observed when the N2O4–N2H4 solid mixture was warmed up slowly from 77 to 218 K during an experiment.14 Parenthetically, it should be mentioned that the conventional isomerization energies for N2O4 → ONONO2 reported in the literature lie in the range of 30–45 kcal/mol,13 too high for the hypergolic reactions of N2O4 with hydrazines to occur.11,12,14

In the present study, we investigate the mechanism for the redox reaction of N2O4 with CH3OH producing the experimentally observed products, CH3ONO and HNO3 by quantum-chemical calculations. If the mechanism of this reaction is similar to those of its isoelectronic analogues, CH3NH23 and NH2NH2,11 then one would expect the production of CH3ONO + HNO3 and CH3NO3 + HONO via a fast and a slow reaction channel, respectively. The significance of CH3ONO formation under the tropospheric condition will be examined on the basis of the predicted kinetics.

2. Computational Methods

2.1. Ab Initio Calculations

The electronic structures of all species involved in the oxidation of CH3OH by N2O4 were optimized with UB3LYP/6-311+G(3df,2p),15 including the association step 2NO2 + CH3OH. To improve energy prediction accuracy, the UCCSD(T)/6-311+G(3df,2p) method was employed with the UB3LYP/6-311 + G(3df,2p) optimized structures. For the heats of reactions forming different products, we have also compared the results obtained with the CCSD(T)/aug-cc-pVTZ//M06-2X/aug-cc-pVTZ method based on the structures optimized with the M06-2X/aug-cc-pVTZ method. Vibrational frequencies of all species involved were calculated at the same level employed for optimization; they were utilized to characterize the stationary points and for ZPE corrections. Unless specified otherwise, all energies cited below are obtained with the UCCSD(T)/6-311+G(3df,2p)//UB3LYP/6-311 + G(3df,2p) method. IRC analyses16 have been carried out to confirm the connectivity of TSs with reactants and products. In the present work, all the calculations were performed using the Gaussian 16 program.17

2.2. Kinetic Calculations

The kinetics for the reaction of N2O4 with CH3OH was calculated using the Variflex program.18 We employed the canonical transition-state theory (TST)19 to predict rate constants for a simple exchange reaction, while the RRKM theory20 was employed to predict the kinetics of a reaction taking place via a long-lived intermediate by solving the 1-D master equation. We utilized the microcanonical TST21 to determine the association rate for a barrierless channel. The potential energy function for the barrierless step N2O4 + CH3OH → N2O4/CH3OH (LM1) was estimated to cover the separation range of N2O4 and CH3OH from 3.19 to13.19 Å with a step size of 0.1 Å at the UB3LYP/6-311 + G(3df,2p) level. The Morse function Inline graphic represents the minimum energy path obtained by fully optimizing structures along the dissociation coordinate. Here, De, R, and Re have their usual definitions. The calculated Morse function for N2O4:CH3OH (LM1) → N2O4 + CH3OH can be represented with β = 1.38 Å–1, and the corresponding value of De is shown in Figure 1.

Figure 1.

Figure 1

Potential energy profile for the N2O4 (D2h) + CH3OH reaction calculated at the UCCSD(T)/6-311+G(3df,2p)//UB3LYP/6-311+G(3df,2p) level with ZPE corrections. Relative energies at 0 K are given in kcal/mol.

3. Results and Discussion

3.1. PES of the N2O4 (D2h) + CH3OH Reaction

We have established the PES for the reaction between N2O4 and CH3OH in the gas phase using the UCCSD(T)/6-311+G(3df,2p)//UB3LYP/6-311 + G(3df,2p) level. In Table 1, we first compare the heats of reaction for 2 potential product pairs predicted at the CCSD(T)/6-311 + G(3df,2p) level based on 2 different DFT optimization methods, UB3LYP/6-311 + G(3df,2p) and M06-2X/aug-cc-pVTZ; the result presented in the table indicates that both approaches agree closely, although the former appears to give values of ΔrHo with a slightly better agreement with available experimental data.

Table 1. Comparison of the Predicted Heats of Reaction at 0 K Based on 2 Different Optimization Methods with Experimental Values.

reaction ΔrHo (kcal/mol)
  predicted (I)a predicted (II)a literature
N2O4+CH3OH = HNO3+CH3ONO –2.5 –2.7 –2.428
N2O4+CH3OH = HONO + CH3ONO2 –2.5 –2.9 –2.528
a

Predicted values (I) by UCCSD(T)/6-311+G(3df,2p)//UB3LYP/6-311+G(3df,2p) and (II) by CCSD(T)/aug-cc-pVTZ//M06-2X/aug-cc-pVTZ.

Figure 1 shows the PES for the title reaction predicted using the UCCSD(T)/6-311+G(3df,2p)//UB3LYP/6-311+G(3df,2p) method. Their related structures optimized with the UB3LYP/6-311+G(3df,2p) method are presented in Figure 2. The relative energies and corresponding molecular parameters (vibrational frequencies and moments of inertia) for all involved species are summarized in Tables 2 and S2, respectively.

Figure 2.

Figure 2

Electronic structures of the key intermediates and TSs involved in the N2O4 (D2h) + CH3OH reaction optimized at the UB3LYP/6-311+G(3df,2p) level. Bond angles and bond lengths are in degree (°) and angstroms (Å), respectively.

Table 2. Relative Energies of Species in the N2O4 + CH3OH Reactiona.

  ZPEb UB3LYP/6-311+G(3df,2p)c UCCSD(T)/6-311+G(3df,2p)c
N2O4+CH3OH 46.7 –526.031 –525.103
LM1 47.4 –3.1 –6.4
LM2 47.3 6.5 2.9
LM3 47.6 –3.6 –9.6
LM4 47.5 –1.4 –7.3
TS1 44.3 8.7 7.9
TS2 45.4 6.5 4.3
TS3 44.9 25.6 28.1
a

The energies are in kcal/mol, relative to that of N2O4 + CH3OH, whose total energy is in Hartree/molecule as given.

b

The ZPE in kcal/mol was calculated at the UB3LYP/6-311+G(3df,2p) level.

c

The single-point energies are based on electronic structures calculated using UB3LYP/6-311+G(3df,2p) with ZPE corrections.

In Figure 1, the reaction occurs through the N2O4:CH3OH (LM1) complex with a binding energy of 6.4 kcal/mol, which is slightly more stable than that of the N2O4/H2O complex, 5.0 kcal/mol,13 but is very close to its isoelectronic N2O4/NH2NH2 complex, 6.7 kcal/mol11, and N2O4/CH3NH2 complex, 6.8 kcal/mol.3 In the present reaction, the LM1 complex can undergo further reaction by lengthening N2O4’s N–N bond to 3.343 Å concurrently rotating one of its NO2 groups at TS1 forming the ONONO2/CH3OH (LM2) complex, which lies 2.9 kcal/mol above the reactants or 9.3 kcal/mol above LM1. The bond ON–ONO2 in the LM2 complex is lengthened to 1.854 Å, which is similar to those in the ONONO2/NH2NH2 complex, 2.282 Å11, and ONONO2/CH3NH2 complex, 2.092 Å.3 Both are longer than those in cis- and trans-ON-ONO2 isomers, 1.685, and 1.622 Å, respectively,13 again suggesting that CH3OH can induce ionization of ON-ONO2 to form [ON+] [NO3]. Mulliken charge analysis shows that the NO and NO3 group charges are +0.367 and −0.507e, respectively (Figure 3).

Figure 3.

Figure 3

Mulliken charges of species involved in the low-energy reaction path.

In the present case, the same roaming-like TS TS1 exists for the isomerization of N2O4 to trans-ONONO2, with a barrier of 7.9 kcal/mol (or 14.3 kcal/mol above LM1), which is again very similar to those of the two isoelectronic systems: N2O4 + NH2NH2 5.9 kcal/mol and N2O4 + CH3NH2 8.0 kcal/mol. From LM2, the NO3 group can abstract the H atom of the terminal OH group in CH3OH via TS2 with a very small barrier of 1.4 kcal/mol to give the postreaction complex LM3 (HNO3:CH3ONO) with 9.6 kcal/mol exothermicity. The complex can readily separate to produce the product pair CH3ONO + HNO3 barrierlessly. To illustrate the effect of CH3OH complexation with N2O4, we have added the potential energy profile for the isomerization of N2O4 to ONONO2 without CH3OH predicted by Raghunath and Lin13 at the UCCSD(T)/CBS level with Dunning’s correlation consistent basis set (cc-pVXZ, where X = D, T, and Q) based on the optimized structures using UB3LYP/6-311+G(3df). The barrier for the isomerization, 12.8 kcal/mol, is lower than that from N2O4:CH3OH to ONONO2:CH3OH by 1.5 kcal/mol, which may be attributed to the complexation effect. Other examples on the effect can be found in Table 3.

Table 3. Comparison of LM1, TS1, LM2, and TS2 for Various N2O4 Reactions (Energies in kcal/mol)a.

reaction LM1 TS1 LM2 TS2
N2O4= ONONO2   12.813    
N2O4+ CH3OH = CH3ONO + HNO3 –6.4 7.9 –2.9 –4.3
N2O4+ CH3NH2= CH3NHNO + HNO3 –6.8 8.0 2.0 –1.63
N2O4+ NH3= NH2NO + HNO3 –5.3 8.7 4.6 9.93
N2O4+ H2O = HONO + HNO3 –5.0 8.7 2.7 12.613
N2O4+ N2H4= N2H3NO + HNO3 –6.7 5.9 –3.8 –4.911
N2O4+ N2H4= N2H3NO + HNO3(in CCl4solvent) –5.9 7.4 –8.6 –6.729
N2O4@HZ23= NO3+ N2H3NO + N2H5+ 0.0 13.1 –45.0 –43.614
N2O4+ CH3NHNH2= CH3NHN(H)NO + HNO3 –6.9 6.3 1.4 1.211
N2O4+ (CH3)2NNH2= (CH3)2NN(H)NO + HNO3 –6.0 7.2 0.6 –0.112
N2O4+ CH3NHNHCH3= CH3N(H)N(CH3)NO + HNO3 –6.1 9.9 –1.9 –2.112
a

UCCSD(T)/6-311+G(3df,2p)//UB3LYP/6-311+G(3df,2p) was employed in refs (3),11, and13 and this work, CCSD(T)/6-311+G(3df,2p)//B3LYP/6-311+G(3df,2p) was employed for the reaction in refs (12) and29 while UCCSD(T)/CBS//UB3LYP/6-311+G(3df) was employed in the N2O4 isomerization reaction ref (13), and the reaction occurring in N2H4 solid was predicted with the DFT method in the VASP code.14

From LM1, another reaction path involves an H-transfer from CH3OH to N2O4 via a tight five-member-ring TS (TS3) with a 28.1 kcal/mol barrier (or 34.5 kcal/mol above LM1) producing CH3ONO2 + trans-HONO. The reaction was predicted to be exothermic by 2.5 kcal/mol, which is exactly the same as the formation of CH3ONO + HNO3. This product channel cannot compete with the RTS-like channel producing CH3ONO and HNO3; it is kinetically unimportant.

In Table 2, the relative energies of all TSs and intermediates computed at UB3LYP/6-311+G(3df,2p) and UCCSD(T)/6-311+G(3df,2p)//UB3LYP/6-311+G(3df,2p) are summarized for a more convenient inspection. In Table 3, we compare the energies of LM1, TS1, LM2, and TS2 relative to each pair of the reactants, from N2O4 hydrolysis13 to N2O4–hydrazine rocket propellants,11,12,14 and to the reactions relevant to tropospheric chemistry.3 The implication of these data will be remarked later.

3.2. Kinetics of the Reaction

The kinetics of the N2O4 + CH3OH reaction are primarily controlled by TS1 because TS2 lies 3.6 kcal/mol below TS1 with a small 1.4 kcal/mol barrier, producing the postreaction complex with 9.6 kcal/mol exothermicity. Their rate constants were computed with the RRKM theory using the Variflex program, in the 200–2000 K range. For the reaction of N2O4 with CH3OH, the dominant product channel produces CH3ONO + HNO3, while the second channel is not considered due to its high barrier at TS3. The low-energy reaction path shown in Figure 1 can be given as follows

3.2.

The prereaction complex LM1 with the small 6.4 kcal/mol well is expected to bring about a minor P effect below room temperature as shown in Figure 1S; the formation of CH3ONO + HNO3 is, however, not affected by pressure as indicated in the figure. Because TS1 (7.9 kcal/mol) is much higher than TS2 (4.3 kcal/mol) and the well depth of LM2 is only 1.4 kcal/mol, as alluded to above, the kinetics of the N2O4 + CH3OH producing CH3ONO + HNO3 is solely controlled by TS1. The rate expression for the formation of CH3ONO + HNO3 in cm3molecule–1s–1 is determined by fitting the computed result with the 3 parameter Arrhenius equation at T = 200–2000 K

3.2.

Figure 4 compares the predicted bimolecular rate constant for CH3ONO + HNO3 production with the result of Wojcik-Pastuszka et al.10 evaluated by kinetic modeling of NO2 decay–time profiles using the 4 reaction mechanism including the forward and reverse processes: 2 NO2 ⇌ N2O4 and N2O4 + CH3OH ⇌ CH3ONO + HNO3 as mentioned in the Introduction. In the figure, we also present the results of Nikki8 and Koda,9 whose original third-order rate constants were kinetically modeled for N2O4 + CH3OH by Wojcik-Pastuszka et al.10 with the 4-reaction mechanism. The agreement between the predicted value and 3 sets of experimental results appears to be reasonable.

Figure 4.

Figure 4

Comparison of the predicted bimolecular rate constant for N2O4 + CH3OH→CH3ONO + HNO3 reaction with the values evaluated by Wojcik-Pastuszka et al.10 through kinetic modeling of NO2-time profiles. The third-order rate constants of both Nikki8 and Koda9 were kinetically modeled by Wojcik-Pastuszka et al.10 to give the second-order rate constants as shown.

The modeling of Wojcik-Pastuszka et al.10 also gave the rate constant for the reverse reaction, CH3ONO + HNO3 → N2O4 + CH3OH, k–1. In Figure 5, we compare the predicted value of k–1 with those obtained by the modeling based on the 4 reactions. The predicted value exhibits a substantial T dependence reflecting the reverse 10.4 kcal/mol barrier as indicated in Figure 1, whereas the model results showed no T dependence. This may reflect the result of poor initial static mixing of the reactants at low temperatures in a tubular reactor typically employed in a photometric measurement.

Figure 5.

Figure 5

Comparison of the predicted bimolecular rate constant for the reverse reaction CH3ONO + HNO3→N2O4 + CH3OH with the values evaluated by Wojcik-Pastuszka et al.10 through kinetic modeling of NO2-time decay profiles.

Estimation of the Contribution from the Termolecular 2NO2 + CH3OH Reaction

The relative average concentration of NO2 and N2O4 in the troposphere in ppbv is known to be about 100:7 × 10–5.1 We, therefore, attempted to estimate the termolecular rate constant for its potential contribution to the oxidation of CH3OH under the tropospheric condition. The method employed is similar to the one we used earlier for our quantitative interpretation of the termolecular kinetics for the 2NO2 + H2O reaction,22 which had been reliably measured in laboratories23 for comparison. In the present system, we can approximately account for the formation of CH3ONO + HNO3 by the stepwise mechanism

3.2. 2
3.2. -2
3.2. 3

The mechanism is similar to that invoked by Koda et al.,9 who initially assumed N2O4 instead of ONONO2 as the intermediate. However, based on their analysis of the measured NO2 decay kinetics, they concluded that ONONO2, rather than N2O4, was actually involved in the reaction.9 Interestingly, this conclusion is consistent with the PES presented in Figure 1 predicted by our high-level quantum calculations.

The steady-state assumption for the unstable ONONO2 intermediate leads to the rate for removal of CH3OH

3.2. 4

The 3 rate constants in eq 4 can be reliably predicted based on the energetics and structures computed at the UCCSD(T)/6-311+G(3df,2p)//UB3LYP/6-311+G(3df,2p) level. Our computed results at atmospheric pressure can be represented by the 3-parameter fitted Arrhenius equations

3.2.
3.2.
3.2.

Under the tropospheric condition, it can be shown that k–2k3 [CH3OH]; eq 4 becomes

3.2.

where K2 = k2/k–2 is the equilibrium constant for reaction (2). Take 200 K and 760 Torr, for example, the concentration of CH3OH in the troposphere is 9.63 × 1011 molecules cm–3, k2 = 2.93 × 10–14 cm3molecule–1s–1, k–2 = 5.41 × 106 s–1, and k3 = 2.56 × 10–14 cm3 molecule–1 s–1; therefore, k–2 = 5.41 × 106 s–1k3 x [CH3OH] = 2.47 × 10–2 s–1.

Therefore, at 200 K in the troposphere, we can show that the rate of CH3OH reaction by the termolecular mechanism presented above is about 4.4 × 105 times smaller than that of the bimolecular N2O4 + CH3OH reaction as estimated below by taking [NO2]/[N2O4] = 100/7 × 10–5,24

[bimolecular rate constant; k1] × [N2O4] = 8.85 × 10–21 × [7 × 10–5] = 6.20 × 10–25;

[3-step termolecular rate constant; ktf ] × [NO2]2 = 1.40 × 10–34 × [100]2 = 1.40 × 10–30, where the termolecular forward rate constant

3.2.

Accordingly, k1 [N2O4]/ktf [NO2]2 = 4.4 × 105 under the tropospheric condition; CH3OH can therefore be more effectively removed by the bimolecular N2O4 reaction, rather than the termolecular process.

3.3. Comparison of the Predicted Termolecular Rate Constant with Experimental Data

As the termolecular kinetics for 2NO2 + CH3OH → CH3ONO + HNO3 had been reported experimentally by Fairlie et al.,7 Nikki et al.,8 and Koda and co-workers9 as aforementioned, we compare the predicted value of ktf with the experimental results as shown in Figure 6. The agreement between the theory and the available experimental values is seen to be quite reasonable. In the figure, we also compare the ktf for the 2NO2 + CH3OH reaction with that of the analogous 2NO2 + H2O → HONO + HNO3 reaction measured experimentally by England and Corcoran23 together with our previously computed result based on the analogous 3-step mechanism22 shown above. In the case of the H2O reaction, the agreement between the theory and experiment is seen to be excellent, reflecting the validity of the predicted PES similar to the one shown in Figure 1.

Figure 6.

Figure 6

Comparison of the predicted third-order rate constant for 2NO2 + CH3OH with the experimental values of Fairlie et al.,7 Nikki et al.,8 and Koda and co-workers,9 as well as with that of the analogous 2NO2 + H2O reaction reported by Zhu et al.22 who compared the predicted third-order rate constant with the experimental result of England and Corcoran.23

The very different temperature dependences of the termolecular rate constants, as is evident in Figure 6, can be attributed to the strong positive temperature effect on k3 for the ONONO2 + H2O reaction (due to its high exit barrier) and the weak T effect on the ONONO2 + CH3OH reaction as shown in Figure S3 because of its low exit barrier.

3.4. Relevancy to the Tropospheric Chemistry

Based on the rate constant k1 given above for the formation of CH3ONO and the concentration levels of NO2, N2O4, and CH3OH in the polluted urban atmosphere, [NO2] ∼ 2 × 1011 molecules cm–3,25 [N2O4] ∼ 2 × 105 molecules cm–3,24 and [CH3OH] ∼ 1 × 1012 molecules cm–3,26 we can estimate the effective rate constant for removal of CH3OH by N2O4 at the lower troposphere at 298 K: k1 × [N2O4] = 2.7 × 10–18 × [1.6 × 105] = 4.4 × 10–13 s–1. The result suggests that the reaction is too slow to be significant.

4. Concluding Remarks

In this study, we have investigated the mechanism for the redox reaction of N2O4 with CH3OH by quantum-chemical calculations. The result of the calculations carried out at the UCCSD(T)/6-311+G(3df,2p)//UB3LYP/6-311 + G(3df,2p) level indicates that the favored reaction path affording the major products CH3ONO + HNO3 as reported in 1951 by Joffe and Gray6 was controlled by the isomerization process forming trans-ONONO2 from N2O4 in the presence of CH3OH during the bimolecular collision. The highly polar and reactive trans-ONONO2 rapidly attacks the CH3OH molecule producing CH3ONO and HNO3 via a 6-member-ring TS with a negligible barrier.

For the N2O4 → ONONO2 isomerization via the roaming-like TS within the N2O4–CH3OH complex, the barrier was predicted to be 7.9 kcal/mol above the reactants (or 14.3 kcal/mol from the prereaction complex). This barrier is significantly lower than the typical tight TS for the unimolecular isomerization (∼30–45 kcal/mol).27 The predicted rate constant for CH3ONO + HNO3 formation can be given by k1 = 1.43 × 10–8T1.96 exp(−9092/T) cm3molecule–1s–1 at T = 200–2000 K, independent of pressure. The result agrees very closely with that of the isoelectronic reaction N2O4 + CH3NH2 as shown in Figure S2.

We have also predicted the kinetics of the 2NO2 + CH3OH termolecular reaction based on the mechanism employed for the analogous 2NO2 + H2O → HONO + HNO3 reaction.22 Comparing our predicted second- and third-order rate constants with available, but scarce, experimental data in the literature for CH3OH reactions with N2O4 and NO2, respectively, the agreement between theory and experiments by and large appears to be reasonable.

The unusual reaction mechanism revealed from this series of studies, starting from the hydrolysis of N2O4 in the gas phase and in the H2O solution13 to the hypergolic ignition of N2O4 in contact with hydrazine propellants,11,12,14 and the processes relevant to the tropospheric chemistry involving potential pollutants such as NH33 and CH3OH indicate that the redox reactions occur via prereaction complexes with about 5 ± 1 kcal/mol binding energies which have only a negligible kinetic consequence except at low temperatures. The results were summarized in Table 3. The redox process starts from the isomerization of the symmetric N2O4 to trans-ONONO2 via a very loose, roaming-like TS1, lying above the complex well at about 14 ± 2 kcal/mol. The highly polar ONONO2 isomer is much more reactive than N2O4 toward the collision partner as is evident from the small TS2 barrier above LM2 in the present case (see Figure 1). However, the barrier at TS2 above LM2 for H2O was predicted to be about 10 kcal/mol, which may be compared with that of its isoelectronic reaction with NH3, 5.3 kcal/mol, and the very small value of 1.4 kcal/mol for CH3OH, reflecting entirely the strength of the bond to be broken by the abstraction reaction of the NO3 group producing HNO3.

In view of the fact that both N2O4 and CH3OH are known pollutants in the lower troposphere, we have examined the potential effect of the formation of CH3ONO and HNO3 from their reaction based on the predicted kinetics given above. Based on the known concentration levels of N2O424 and CH3OH,26 the rate of their reaction at 298 K was found to be too slow to be relevant to the troposphere chemistry.

Acknowledgments

This work is supported by the Ministry of Science and Technology, Taiwan (grant no. MOST 107-3017-F009-003), and the Center for Emergent Functional Matter Science of National Yang Ming Chiao Tung University from The Featured Areas Research Center Program within the framework of the Higher Education Sprout Project by the Ministry of Education (MOE) in Taiwan. We also want to thank the reviewers and the editor for their valuable comments and suggestions.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpca.4c02433.

  • Studying the impact of pressure on the N2O4 + CH3OH reaction and comparing rate constants for product formation in the N2O4 reaction with CH3OH and CH3NH2, as well as comparing bimolecular rate constants predicted for ONONO2 reactions with CH3OH and H2O (PDF)

The authors declare no competing financial interest.

Supplementary Material

jp4c02433_si_001.pdf (239.7KB, pdf)

References

  1. Heikes B. G.; Chang W.; Pilson M. E. Q.; Swift E.; Singh H. B.; Guenther A.; Jacob D. J.; Field B. D.; Fall R.; Riemer D.; et al. Atmospheric Methanol Budget and Ocean Implication. Global Biogeochem. Cycles 2002, 16, 1–23. 10.1029/2002gb001895. [DOI] [Google Scholar]
  2. Speight J. G.Environmental Organic Chemistry for Engineers; Butterworth-Heinemann: Oxford, U.K., 2017. [Google Scholar]
  3. Trac H. P.; Le Huyen T.; Lin M.-C. A Computational Study on the Redox Reactions of Ammonia and Methylamine with Nitrogen Tetroxide. J. Phys. Chem. A 2020, 124, 9923–9932. 10.1021/acs.jpca.0c08665. [DOI] [PubMed] [Google Scholar]
  4. Sarkar S.; Bandyopadhyay B. Reaction between N2O5 and NH3 under Tropospheric Conditions: A Quantum Chemical and Chemical Kinetic Investigation. J. Phys. Chem. A 2020, 124, 3564–3572. 10.1021/acs.jpca.0c00580. [DOI] [PubMed] [Google Scholar]
  5. Harris L.; Siegel B. M. Reactions of Nitrogen Dioxide with Other Gases. J. Am. Chem. Soc. 1941, 63, 2520–2523. 10.1021/ja01854a061. [DOI] [Google Scholar]
  6. Yoffe A. D.; Gray P. 325. Esterification by dinitrogen tetroxide. J. Chem. Soc. 1951, 1412–1414. 10.1039/jr9510001412. [DOI] [Google Scholar]
  7. Fairlie J.; A M.; Carberry J. J.; Treacy J. C. A Study of the Kinetics of the Reaction between Nitrogen Dioxide and Alcohols. J. Am. Chem. Soc. 1953, 75, 3786–3789. 10.1021/ja01111a054. [DOI] [Google Scholar]
  8. Niki H.; Maker P. D.; Savage C. M.; Breitenbach L. P. An FTIR Study of the Reaction between Nitrogen Dioxide and Alcohols. Int. J. Chem. Kinet. 1982, 14, 1199–1209. 10.1002/kin.550141104. [DOI] [Google Scholar]
  9. Koda S.; Yoshikawa K.; Okada J.; Akita K. Reaction Kinetics of Nitrogen Dioxide with Methanol in the Gas Phase. Environ. Sci. Technol. 1985, 19, 262–264. 10.1021/es00133a008. [DOI] [PubMed] [Google Scholar]
  10. Wójcik-Pastuszka D.; Gola A.; Ratajczak E. Gas Phase Kinetics of the Reaction System of 2NO2 ↔ N2O4 and Simple Alcohols between 293–358 K. Polym. J. Chem. 2005, 79, 1301–1313. [Google Scholar]
  11. Trinh L. H.; Raghunath P.; Lin M. C. Ab Initio Chemical Kinetics for Hypergolic Reactions of Nitrogen Tetroxide with Hydrazine and Methyl Hydrazine. Comput. Theor. Chem. 2019, 1163, 112505. 10.1016/j.comptc.2019.112505. [DOI] [Google Scholar]
  12. Trinh L. H.; Raghunath P.; Lin M. C. Ab Initio Chemical Kinetics for Nitrogen Tetroxide Reactions with 1,1- and 1,2-Dimethylhydrazines. Propellants, Explos., Pyrotech. 2020, 45, 1478–1486. 10.1002/prep.201900426. [DOI] [Google Scholar]
  13. Putikam R.; Lin M. C. A Novel Mechanism for the Isomerization of N2O4 and Its Implication for the Reaction with H2O and Acid Rain Formation. Int. J. Quantum Chem. 2018, 118, 1–10. 10.1002/qua.25560. [DOI] [Google Scholar]
  14. Raghunath P.; Lin M. C. Isomerization of N2O4 in Solid N2H4 and Its Implication for the Explosion of N2O4–N2H4 Solid Mixtures. J. Phys. Chem. C 2018, 122, 23501–23505. 10.1021/acs.jpcc.8b07030. [DOI] [Google Scholar]
  15. Becke A. D. Density-functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648–5652. 10.1063/1.464913. [DOI] [Google Scholar]
  16. Gonzalez C.; Schlegel H. B. Improved Algorithms For Reaction Path Following: Higher-order Implicit Algorithms. J. Chem. Phys. 1991, 95, 5853–5860. 10.1063/1.461606. [DOI] [Google Scholar]
  17. Frisch M. J.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Petersson G. A.; Nakatsuji H.; et al. Gaussian 16. Revision C.01; Gaussian Inc.: Wallingford, CT, 2016.
  18. Klippenstein S.; Wagner A.; Dunbar R.; Wardlaw D.; Robertson S.. Variflex, version 1.00; Argonne National Laboratory: Argonne, IL, 1999.
  19. Steinfeld J. I.; Francisco J. S.; Hase W. L.. Chemical Kinetics and Dynamics; Prentice Hall: Upper Saddle River, N.J., 1999. [Google Scholar]
  20. Wardlaw D. M.; Marcus R. A. RRKM Reaction Rate Theory for Transition States of Any Looseness. Chem. Phys. Lett. 1984, 110, 230–234. 10.1016/0009-2614(84)85219-7. [DOI] [Google Scholar]
  21. Wardlaw D. M.; Marcus R. A.. In On the Statistical Theory of Unimolecular Processes; Prigogine I., Rice S. A., Eds.; Plenum Press: New York, 1988; Vol. 70, pp 231–263. [Google Scholar]
  22. Zhu R. S.; Lai K.-Y.; Lin M. C. Ab Initio Chemical Kinetics for the Hydrolysis of N2O4 Isomers in the Gas Phase. J. Phys. Chem. A 2012, 116, 4466–4472. 10.1021/jp302247k. [DOI] [PubMed] [Google Scholar]
  23. England C.; Corcoran W. H. Kinetics and Mechanisms of the Gas-Phase Reaction of Water Vapor and Nitrogen Dioxide. Ind. Eng. Chem. Fundam. 1974, 13, 373–384. 10.1021/i160052a014. [DOI] [Google Scholar]
  24. Kamboures M. A.; Raff J. D.; Miller Y.; Phillips L. F.; Finlayson-Pitts B. J.; Gerber R. B. Complexes of HNO3 and NO3– with NO2 and N2O4, and Their Potential Role in Atmospheric HONO Formation. Phys. Chem. Chem. Phys. 2008, 10, 6019–6032. 10.1039/b805330h. [DOI] [PubMed] [Google Scholar]
  25. Cooper M. J.; Martin R. V.; Hammer M. S.; Levelt P. F.; Veefkind P.; Lamsal L. N.; Krotkov N. A.; Brook J. R.; McLinden C. A. Global Fine-scale Changes in Ambient NO2 during COVID-19 Lockdowns. Nature 2022, 601, 380–387. 10.1038/s41586-021-04229-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Bates K. H.; Jacob D. J.; Wang S.; Hornbrook R. S.; Apel E. C.; Kim M. J.; Millet D. B.; Wells K. C.; Chen X.; Brewer J. F.; et al. The Global Budget of Atmospheric Methanol: New Constraints on Secondary, Oceanic, and Terrestrial Sources. J. Geophys. Res. 2021, 126, 1–23. 10.1029/2020jd033439. [DOI] [Google Scholar]
  27. Pimentel A. S.; Lima F. C. A.; da Silva A. B. F. The Isomerization of Dinitrogen Tetroxide: O2N– NO2 → ONO– NO2. J. Phys. Chem. A 2007, 111, 2913–2920. 10.1021/jp067805z. [DOI] [PubMed] [Google Scholar]
  28. Burcat A.; Ruscic B.. Third Millennium Ideal Gas and Condensed Phase Thermochemical Database for Combustion with update from Active Thermochemical Tables; Argonne National Laboratory: Argonne, Illinois, U.S.A. and Technion-IIT, Haifa, Israel, 2005.
  29. Trinh L. H.; Raghunath P.; Lin M. C. Quantum Chemical Modeling of Spontaneous Reactions of N2O4 with Hydrazines in CCl4 Solution at Low Temperature. Comput. Theor. Chem. 2020, 1188, 112951. 10.1016/j.comptc.2020.112951. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

jp4c02433_si_001.pdf (239.7KB, pdf)

Articles from The Journal of Physical Chemistry. a are provided here courtesy of American Chemical Society

RESOURCES