Oleg B. Gadzhiev*ab,
Victor A. Dodonov*a and
Alexander I. Petrovcd
aDepartment of Chemistry, N.I. Lobachevsky State University of Nizhny Novgorod, 23 Gagarin Avenue, Nizhny Novgorod, 603950, Russia. E-mail: euriscomail@mail.ru; vadodonov@gmail.com
bG.G. Devyatykh Institute of Chemistry of High-Purity Substances, Russian Academy of Sciences, 49 Troponina Street, Nizhny Novgorod, 603950, Russia
cInstitute of Non-Ferrous Metals and Materials Science, Siberian Federal University, 81 Svobodny Prospect, Krasnoyarsk, 660041, Russia
dInstitute of Chemistry and Chemical Technology, Siberian Branch of Russian Academy of Sciences, 42 K. Marx Street, Krasnoyarsk, 660049, Russia
First published on 29th September 2016
The quantum chemical study of the MeOOH/(MeO)3Al model system has been carried out in order to predict the mechanism of the catalytic decomposition of t-BuOOH under mild conditions for the t-BuOOH/(t-BuO)3Al system being a powerful synthetic tool for selective oxidation. To elucidate the chemical excitation of O2 eliminated in the catalytic reaction and to predict the electronic state of O2, the topology of the potential energy surface (PES), the structures of intermediates and transition states, the activation and reaction energies were obtained at the B3LYP/cc-pVTZ theory level. It was shown that the peroxide, (MeO)2AlOOMe, corresponding to the experimentally obtained (t-BuO)2AlOOBu-t, is formed in the first step of the reaction. After that, in the main pathway, the aluminum-containing peroxide reacts with the second MeOOH molecule through the nucleophilic substitution of the second methoxy group forming the MeOAl(OOMe)2 diperoxide. The diperoxide rearranges to aluminum-containing ozonide MeOAlOOOMe. The ozonide isomerizes in the mononuclear-metal dioxygen intermediate (MeO)3Al·O2. The latter decomposes through the adiabatic ((MeO)3Al + O2(b1Σ+g)) and non-adiabatic ((MeO)3Al + O2(X3Σ−g)) pathways, which corresponds to experimental data about the incomplete conversion of O2 to O2(b1Σ+g). The generation of O2(b1Σ+g) was revealed by the analysis of the energy diagram calculated with the CCSD(T), CCSDT(Q), and CASSCF methods. It was suggested that the η1-(MeO)3Al·O2 and, thus, (t-BuO)3Al·O2 complexes are new sources of O2(b1Σ+g).
The investigations by V. A. Dodonov et al.5,6 of the early 1990s summarized in review7 point out that organoaluminum peroxides (RO)2AlOOR′ can be synthesized individually only with tertiary alkyl (aryl) peroxo groups (see ref. 8 and ESI†). The nucleophilic substitution reactions of aluminum tri-tert-butoxide with tertiary hydroperoxides (t-BuOOH, PhC(CH3)2OOH, Ph3COOH) have been studied further in ref. 5, 6 and 9–11 (Scheme 1). The corresponding organoaluminum peroxides have been obtained in the reactions with cumene hydroperoxide and trityl hydroperoxide in benzene at room temperature (Scheme 1).
![]() | ||
Scheme 1 The high-yield synthesis of individual organoaluminum peroxides (t-BuO)2AlOOR′ from ready available (t-BuO)3Al. |
Unlike aryl-containing hydroperoxides, tert-butyl hydroperoxide reacts with aluminum tert-butoxide yielding dioxygen and tert-butanol. The reaction of aluminum tert-butoxide with tert-butyl hydroperoxide at the mole ratio of 1:
2 in benzene or CCl4 yields 80–90% of O2. The elimination of O2 remains the main reaction5,6 when the compound ratio is increased to 1
:
10. Thus, the catalytic reaction occurs in the (t-BuO)3Al–2t-BuOOH system, where (t-BuO)3Al is a catalyst and t-BuOOH is a reagent. The catalytic systems based on aluminum or titanium tert-butoxides and tert-butyl hydroperoxide oxidize organic sulfides to sulfones with the yield of 82–99% at room temperature and at the mole ratio of 1
:
2 for (t-BuO)3Al and t-BuOOH12–14 (see also ref. 15 and 16, review17 and references therein). Other examples of aluminum catalysts in a combination with H2O2 or alkyl hydroperoxide as oxidants are relatively scarce: BINOL-Al in the Bayer–Villiger reaction developed by C. Bolm et al.,18,19 Al(salelen) for sulfides (chiral)oxidation developed by T. Katsuki.20,21
In order to understand the processes in the (t-BuO)3Al–2t-BuOOH catalytic system, reactions in non-reactive solvents (benzene and carbon tetrachloride) and in oxidizable solvents (various hydrocarbons) were studied in ref. 5, 6, 9 and 11. As was shown,5,6,9,11 the dioxygen formed in the catalytic reaction decays practically completely oxidizing geminal C–H bonds in alkanes and alkylaromatic hydrocarbons. The selective oxidation of C–H bonds in alkanes and alkylarenes was studied8 by means of EPR spin trapping with 2-methyl-2-nitrosopropane (MNP) and C-phenyl-N-tert-butyl nitrone (PBN). The following oxygen-centered radicals were identified:8 t-BuO˙, t-BuOO˙, (t-BuO)2AlOO˙, and (t-BuO)2AlO˙. On the basis of the experimental results, the existence of di-tert-butoxy-tert-butyltrioxyaluminum, (t-BuO)2AlOOOBu-t, and its thermolysis mechanism were suggested (Scheme 2).6–8
![]() | ||
Scheme 2 Tentative mononuclear aluminum–oxygen intermediates in the catalytic dissociation of t-BuOOH with singlet O2 generation: (t-BuO)3Al is a catalyst, t-BuOOH is a reagent. |
Generation of singlet dioxygen is a subject of numerous reviews,22–26 books,27,28 and ongoing researches. Indispensable is the role of singlet oxygen in photodynamic therapy of cancer, photomedicine and photobiology. Reactive oxygen substances (ROS) in ground and excited electronic states are of general importance for biochemistry and biomedical applications29–33 as well as environmental and radiochemical applications.34–36 While O2(1Δg) photosynthesis and O2 activation on transition metals' compounds are studied worldwide in many groups, the presented here catalytic chemical excitation of dioxygen up to O2(1Σ+g) on non-transition metal centre, that is, as was proposed earlier, in reactions of aluminum-containing peroxides, represents an essential novelty.
However, the proposed earlier catalytic mechanism (Scheme 2) of t-BuOOH dissociation with the active O2 generation cannot be considered as a proved mechanism because it is impossible to exclude high activation barriers for the specified reactions. It was not established earlier whether the assumed reactions are elementary ones and whether other short-lived and highly reactive intermediates containing aluminum dioxygen bond exist. Thus, the goal of the present study is to predict reaction mechanism in the (MeO)3Al + 2MeOOH system modeling the (t-BuO)3Al–2t-BuOOH system where the catalytic activation of dioxygen with the participation of a metal center was exhibited earlier. Here, we use quantum chemical methods to study structures, vibrational frequencies of intermediates and transition states and to predict reaction paths and sources of chemically excited and chemically activated O2.
To investigate the electronic structure of some intermediates that are primarily important for the generation of 1O2 and the determination of the electronic state of 1O2, state-of-the art quantum chemical methods were employed. On the model system, it becomes possible to use the coupled cluster methods (up to CCSDT(Q)) and the complete active space method (CASSCF).
2MeOOH → 2MeOH + 3O2 | (1) |
2MeOOH → 2MeOH + O2(1Δg) | (2) |
2MeOOH → 2MeOH + O2(b1Σ+g) | (2a) |
(MeO)3Al·O2 → (MeO)3Al + O2(1Δg) | (3) |
(MeO)3Al·O2 → (MeO)3Al + O2(b1Σ+g) | (3a) |
(MeO)3Al·O2 → (MeO)3Al + 3O2 | (4) |
The energies of the reactions (1) and (2) were estimated at the CCSDT(Q,fc)/cc-pVTZ theory level using the focal point analysis based on the CCSD(T,fc)/cc-pVTZ and CCSDT(Q,fc)/cc-pVDZ calculations. The energy of the reaction (2a) was calculated on the basis of the estimated reaction energy (2) and the experimental energy of the O2(b1Σ+g) ← O2(a1Δg) transition.
As will be shown in the present study, the (MeO)3Al·O2 intermediate is of especial importance for O2 generation. Thus, to establish the electronic state of 1O2 eliminated in the (MeO)3Al·O2 decomposition and to reveal the possibility of chemical electron excitation, i.e., to distinguish between the reactions (3) and (3a), the PES profile corresponding to the reaction ((MeO)3Al·O2 → (MeO)3Al + 1O2) was determined while the energy of the reaction was calculated using the CCSD(T)/6-31+G(2df,p)//B3LYP/cc-pVTZ, CCSD(T)/cc-pVTZ//B3LYP/cc-pVTZ composite approaches, and the CAS(14,10)/6-311G(d) theory level. The energies of the quasi-isolated system [(MeO)3Al + 1O2] were obtained by relaxed scans at the CAS(14,10)/6-311G(d) theory level. These reaction paths correspond to the adiabatic reaction. In these calculations, the aluminum–adjacent oxygen interatomic distance in the Al–O2 moiety was elongated to 1.5 Å above the equilibrium bond length in the (MeO)3Al·O2. Moreover, to distinguish between the reactions (3) and (3a) and, thus, to reveal an electronic term of O2, the energy of the reaction (4) was calculated at the CCSD(T)/cc-pVTZ theory level while the energies of the asymptotes corresponding to the reactions (3) and (3a), i.e., the levels of products, were estimated on the basis of the CCSDT(Q)/cc-pVTZ data for the electronic states of O2 and the coupled cluster data for the reaction (4). This procedure was applied since the calculations of (MeO)3Al·O2 with the CCSDT(Q) method are prohibitive time- and resource demanding.
To take into account the possibility of chemically activated O2 generation (reaction (4)) and, thus, to predict non-adiabatic reaction pathways in the MeOOH decomposition of the MeOOH/(MeO)3Al catalytic system, intersections between PESs were searched at the B3LYP/cc-pVTZ theory level.
The calculations were conducted with the Gaussian 03 software42 (for the B3LYP and the CASSCF methods), the GAMESS43–47 and CFOUR programs48 (for the CCSD(T) method). For the CCSDT(Q,fc)/cc-pVTZ energy estimations, the focal point analysis was carried out on the results obtained with the CFOUR and MRCC49–51 suites of programs. Minimum energy crossing point (MECP) search was conducted with ORCA suite of programs.52–54 The Moltran,55 GaussView03,56 and ChemCraft57 molecular editors were used to analyze the results.
![]() | ||
Fig. 2 Structures calculated by full geometry optimization at the B3LYP/cc-pVTZ theory level for the reaction 2MeOOH + (MeO)3Al → (MeO)Al(OOMe)2 + 2MeOH. Bond lengths are in Å. |
The reaction of the second MeOOH molecule with 3 or 3a or their complexes with the MeOH molecule can occur similarly to the interaction of the first MeOOH molecule with (MeO)3Al, i.e., according to the nucleophilic substitution mechanism contrary to the rearrangement 5 → TS2 → 6 (Fig. 1) that includes the cleavage of the O–O bond in the MeOO group and successive elimination of MeOH. Another reaction pathway is determined by 4c which an isomer of 4. Complex 4c with the relative energy (Erel = −125.5 kJ mol−1) that is similar to Erel of 4 (Erel = −124.8 kJ mol−1) eliminates MeOH. In this no transition state reaction with the activation energy of 63.5 kJ mol−1 8 is formed that is the isomer of 5. Structure 8 (Fig. 1 and 3) with a four-membered quasi-cycle consisting of the coordination and hydrogen bonds is 14.3 kJ mol−1 less stable than 5 (Fig. 1 and 2) formed by the coordination of the MeOOH molecule which is involved in the six-membered quasi-cycle. The isomerization 8 → TS3 → 9 occurs with the activation energy (Ea) of 34.9 kJ mol−1 and is a weakly exothermic reaction (Er = −29.9 kJ mol−1). Then, during the elimination of MeOH from 9, one or two peroxo groups may close in three-membered cycle or two cycles, respectively, with simultaneous formation the hydrogen-bonded complexes similar to 6.
![]() | ||
Fig. 3 Structures calculated by full geometry optimization at the B3LYP/cc-pVTZ theory level for the reactions in the (MeO)2AlOOMe + MeOOH + MeOH system. Bond lengths are in Å. |
![]() | ||
Fig. 5 Structures calculated by full geometry optimization at the B3LYP/cc-pVTZ theory level for the reaction of (MeO)2AlOOMe. Bond lengths are in Å. |
For 7 (Fig. 2), several isomers can be obtained by opening one or two three-membered quasi-cycles AlOO (Fig. 4). Mono-η2-methyl peroxides, except for conformation isomers formed by methyl group rotation, have conformation isomers 12 and 12a (Fig. 4) which are different in the dihedral angle ∠O–Al–O–O, where O–O is a peroxo group corresponding to the opened cycle. This angle is −11.6° for 12 and −179.0° for 12a (Fig. 4). The 12 and 12a conformers are characterized by a comparable energetic stability. Indeed, the relative energy (Erel) of 12 and two MeOH monomers is 7.5 kJ mol−1 and Erel = 4.6 kJ mol−1 was calculated for 12a and two MeOH monomers (Fig. 4). The two OOAl cycles (7 → 12b) are opened with Er = 29.3 kJ mol−1 (Fig. 4). In all the cases (7.4 kJ mol−1 for 7 → 12 and 4.5 kJ mol−1 for 7 → 12a), the activation energy is many times less than Ea of the reaction 7 → TS4 → 10 (Fig. 4). If two peroxo groups in MeOAl(OOMe)2 are not closed in a cycle, the conformation isomer 12b with two MeOH monomers (Fig. 4) is 29.4 kJ mol−1 less energetically stable than the initial reagents ((MeO)3Al + 2MeOOH).
Mono-η2-methyl peroxide 12 (Erel = 7.5 kJ mol−1) is isomerized with Er = 13.5 kJ mol−1 to ozonide-1,3 (13) via transition state TS6 with the relative energy (Erel) of 129.3 kJ mol−1 (Fig. 3 and 4). Ozonide-1,3 (10) is more energetically favorable than 13. Indeed, Er calculated for the reaction 13 → 10 is 61.4 kJ mol−1; however, isomerization through TS7 (Fig. 4) with the methyl group shift is a reaction with a rather high barrier (Ea = 185.9 kJ mol−1). This activation energy value is insignificantly higher than Ea = 182.0 kJ mol−1 for the reaction pathway to 7 → 10.
12a is slightly different in reactivity from the conformer 12 (Fig. 5). In this case, the transfer of the O atom of the OO-η2 group to the peroxo group and the increase of the oxygen chain length occurs with the activation energy of 159.2 kJ mol−1 for TS6a, i.e., 12 is more reactive because Ea for 7 → 10 is 22.8 kJ mol−1 more. The OOOCH3 chain is stabilized in the form of the η2-donor–acceptor coordination complex 13a that is 41.4 kJ mol−1 more energetically stable than 13. Ozonide-1,2 (structure 13a, Fig. 4 and 5) turns out to be 20 kJ mol−1 less stable than ozonide-1,3 (10, Fig. 5). Contrary to the reaction 12 → 13, the isomerization 12a → 13a is an exothermic (Er = 25.0 kJ mol−1).
![]() | ||
Fig. 7 Structures calculated by full geometry optimization at the B3LYP/cc-pVTZ theory level for the reaction of (MeO)2AlOOMe with MeOOH. Bond lengths are in Å. |
However, the activation energy (182.0 kJ mol−1) for the intramolecular transfer of the O atom in 7 with the formation of ozonide-1,3 (10) is by 48.6 kJ mol−1 higher than the highest barrier for the net reaction 5 → 7 (Fig. 1) while the synchronous formation of two η2-groups (5 → 6, Fig. 1) is the reaction with the highest activation barrier in the PES region corresponding to the pathway of the successive substitution of two MeO in (MeO)3Al (Fig. 1). Indeed, the activation energy (Ea = 133.4 kJ mol−1) of the reaction 5 → 6 (Fig. 1) is more than two times higher than Ea = 56.9 kJ mol−1 of the reaction 1 → 2 (Fig. 1) and the total activation energy of the stepwise reaction 4c → 9 (Fig. 1).
Another reaction path to ozonide to be compared is a path with the intermediate 9. Intermediate 9 (Fig. 1) is a precursor of conformers 12 and 12a (Fig. 4) that are formed from 9 after the elimination of the coordinative bonded MeOH and the closing of one AlO2Me ring with the activation energy of 96.5 kJ mol−1 with no transition state. The reaction 12 → 13 results in the formation of the OOO chain stabilized by including in the four-membered cycle AlO3 with almost equivalent lengths of the corresponding bonds (the difference is not more than 0.0004 Å, Fig. 5). Structure 13 (Fig. 5) insignificantly differs in energy (0.1 kJ mol−1 more stable) and in geometric parameters (the value of root mean square deviation for atomic positions is 0.039 Å) from a more symmetric structure that is characterized by the Cs symmetry point group. However, the symmetric structure has one (soft) imaginary vibrational frequency which breaks the symmetry. 13 (Fig. 4) is less energetically stable than 10 (Er = −61.4 kJ mol−1 for this rearrangement), the methyl group shift requires overcoming of the barrier of 185.9 kJ mol−1 (Fig. 4). This Ea value is insignificantly higher than Ea (182.0 kJ mol−1) of the rate-determining step for the minor pathway with TS4 (Fig. 4). While the barrier of 185.9 kJ mol−1 for the two-step reaction 12 → 13 → 10 is comparable to the barrier height of 182.0 kJ mol−1 in the single-step reaction 7 → TS4 → 10 determined relative to 7, the 12 → 13 reaction is endothermic and, thus, less thermodynamically favorable. Therefore, the pathway through TS4 is more probable than the two-step reaction with TS6 and TS7.
The ozonide-1,2 (structure 13a, Fig. 4 and 5) is formed with the activation energy (Ea) of 159.2 kJ mol−1, which exceeds the activation energies in the reaction pathways of the successive (two-step) nucleophilic substitution of one methoxy group and the subsequent O–O bond cleavage closing the AlOOOMe ring. The activation barrier (7.8 kJ mol−1) of the ozonide-1,2 isomerization, (13a) → ozonide-1,3 (10), is negligible in comparison with the activation energies calculated for 12a → TS6a → 13a, the isomerization 10 → TS5 → 11, and the 1O2 elimination from the (MeO)3Al·O2 complex (11) (Fig. 3). Therefore, the pathway 12a → TS6a → 13a → TS7a → 10 → TS5 → 11 (Fig. 3) is a more probable one among the three pathways of the ozonide synthesis. Indeed, the activation energy of 159.2 kJ mol−1 for the rate-determining step (12a → TS6a → 13a), and the reaction pathway 7 → TS4 → 10 → TS5 → 11 → 1O2 + (MeO)3Al with the rate-determining step 7 → TS4 → 10 for which Ea = 182.0 kJ mol−1 is a minor channel (Fig. 4).
The reaction 14 → TS8 → 15 (Fig. 6) is a pathway to the ozonide 15 through the direct attack of MeOOH onto the mono-η2-methyl peroxo group with building the OOO chain. The activation energy (Ea) of the reaction 14 → 15 is 173.7 kJ mol−1 (Fig. 6). This value is 14.5 kJ mol−1 higher than Ea for the reaction pathway with TS6a, but it is 8.3 kJ mol−1 less than Ea of the reaction through TS4, that is, the minor pathway with TS4 is less effective than this one with TS6a (Fig. 4).
Thus, the effectively competing reaction pathways to the ozonides and their complexes were predicted at the B3LYP/cc-pVTZ theory level. In each case, it was shown that the formation of a three-oxygen-atom chain (OOO) is the rate-determining step of the net reaction in the 2MeOOH + (MeO)3Al catalytic system. It is worth noting that the reaction pathway including the isomerization of mono-η2-peroxo aluminum (12a) to ozonide-1,2 (13a) is characterized by the lowest barrier height (Ea = 159.2 kJ mol−1, Fig. 4) and, thus, it is a major adiabatic channel.
As was shown in the scrupulous study of oxygen allotropes in the form of molecules and complexes,58 the energy of the transition O2(a1Δg) ← O2(X3Σ−g) is 99.5 ± 6.9 kJ mol−1 as calculated with the CCSDT(Q) method that agrees well with the experimental value59 of 94.7 kJ mol−1. In this case, it is necessary to use the CCSDT(Q) method to calculate the energy of the reaction with O2(1Δg) formation, i.e., for the reaction (2). The energy of the reaction (2a) will be obtained if we know the high accurate energy estimations of the reactions (1) and (2) and also the experimental energy of O2(b1Σ+g) ← O2(X3Σ−g).
In the present study, the energies of the reactions (1) and (2) were estimated at the CCSDT(Q,fc)/cc-pVTZ theory level using the focal point analysis based on the CCSD(T,fc)/cc-pVTZ and CCSDT(Q,fc)/cc-pVDZ calculations (ESI, Table 1S†). The energies of the reactions (1) and (2) are given as following −144.9 kJ mol−1 and −45.4 kJ mol−1 at the CCSDT(Q,fc)/cc-pVTZ theory level. The energy of the reaction (2a) was calculated to be 13.0 kJ mol−1 taking into account the energy of the reaction (1) obtained here and the experimental value of the O2(b1Σ+g) ← O2(X3Σ−g) transition energy. Thus, MeOOH is a high energetic molecule. Both pathways (2) and (2a) are energetically accessible because the decomposition of MeOOH with the O2(1Δg) elimination is weakly exothermic (Er = −45.4 kJ mol−1) while, in the generation of O2(b1Σ+g), the decomposition reaction is weakly endothermic (Er = 13.0 kJ mol−1).
Both theory levels (composite and multi-reference) are in agreement with the results of the relaxed scan of the Al–O(O) bond in 11 at the B3LYP/cc-pVTZ theory level: (1) there is no maximum on the potential curve, i.e., the decomposition of 11 on the singlet PES occurs with no transition state; (2) the barrier height is 19.0 kJ mol−1 (CCSD(T)//B3LYP with cc-pVTZ basis set), 34.7 kJ mol−1 (CASSCF). It is worthwhile to note that no intruder states were detected during the orbital update for the CASSCF calculation because the potential curve for the relaxed scan was smooth. The optimized structure of (MeO)3Al·O2 was not distorted, i.e., it was similar to the structure 11 optimized with the B3LYP method, but the Al–O-bond was longer. No η1-superoxo-η2-peroxo isomerism reported60–65 for the compounds of transition metals was revealed for (MeO)3Al·O2.
Next, to distinguish between the reactions (3) and (3a), the energy of the reaction (4) was calculated with the composite CCSD(T,fc)/cc-pVTZ//B3LYP/cc-pVTZ approach: Er is −104.6 (−115.1) kJ mol−1, the value for B3LYP is given in parentheses for comparison. One can conclude that the electronic configuration of the 3O2 molecule is correctly described by the CCSD(T) method (see the comparison of the results in ref. 58). Here, one can note that, for 11, the non-dynamic electron correlation is taken into account by the CCSD(T) method66 as was revealed by the T1 diagnostic (0.019) and the largest T2 amplitude with absolute value of 0.15 calculated at the CCSD(T)/6-31+G(2df,p) theory level. The use of the flexible cc-pVTZ basis set for coupled cluster method provides the result which is nearly free from the basis set incompleteness error. Thus, one can believe that the Er value of −104.6 kJ mol−1 is a reliable estimation.
However, for reactions with the (MeO)3Al·O2 intermediate, due to the especial time and resource consumption of calculations with the CCSDT(Q) method even with the moderate cc-pVDZ basis set, it is impossible to use this method for estimations with the focal point analysis. The energy of the intermediate 11 was calculated at the CCSD(T)/cc-pVTZ theory level as well as the energy of the reaction (4). On the basis of these results, the relative energy of the (MeO)3Al + O2(1Δg) level is estimated with the CCSDT(Q)/cc-pVTZ approach for the O2(1Δg)/O2(X3Σ−g) system. The relative energy of the (MeO)3Al + O2(b1Σ+g) level is obtained on the basis of the experimental energy of the excitation O2(b1Σ+g) ← O2(a1Δg) and the calculated energy of the reaction (3).
One can believe that the chosen combination of the CCSD(T) and CCSDT(Q) methods provides reliable results competing with the experimental data. Thus, the calculated reaction energies for (MeO)3Al·O2 and experimentally obtained values for the excitation O2(b1Σ+g) ← O2(1Δg) will be assembled on one energetic diagram (Fig. 8) and analyzed jointly. The level of isolated (MeO)3Al + O2(X3Σ−g) is selected as the initial point.
The main features of the determined location of the levels are: (1) the energy of the (MeO)3Al + O2(1Δg) system is lower than the level of (MeO)3Al·O2 (complex 11); (2) the quasi-isolated system [(MeO)3Al + 1O2] is energetically higher than 11; (3) the energy of the [(MeO)3Al + 1O2] coincides with the energy of (MeO)3Al + O2(b1Σ+g). Assuming that the (MeO)3Al·O2 dissociation corresponds to the (MeO)3Al + O2(1Δg), it is necessary to conclude that if the dissociation proceeds without a transition state (Fig. 8), the potential energy curve of the relaxed PES will exhibit decrease of the energy or if a saddle point or a conical intersection is located in the reaction pathway (Fig. 8), the curve will show an increase with subsequent decrease of the energy. However, as was calculated with quantum chemical methods, the energy curve smoothly increases, that is, the features of the relaxed scan curve for the assuming reactions are totally opposite. Therefore, one can conclude that the dissociation asymptotics of (MeO)3Al·O2 corresponds to (MeO)3Al + O2(b1Σ+g) rather than to (MeO)3Al + O2(1Δg).
The position of the η1-(MeO)3Al·OO level (Fig. 8) in the system of energy levels (MeO)3Al + O2(b1Σ+g), (MeO)3Al + O2(1Δg), and (MeO)3Al + O2(X3Σ−g) calculated with the CCSD(T) and CCSDT(Q) methods and the position of the quasi-isolated [(MeO)3Al + 1O2] level determined by the CASSCF method is of crucial importance for the generation of O2(b1Σ+g). The energy of chemical bonds of two MeOOH is transferred through η1-(MeO)3Al·OO into the energy of the excited O2 eliminated in the adiabatic decomposition reaction. The key factor to implement the quite unusual non-photoinduced process, i.e., the chemical excitation O2(b1Σ+g) ← O2(X3Σ−g), is the multi-pathway sequence of elementary reactions leading for the model system to (MeO)2AlOOOMe and, next, to η1-(MeO)3Al·OO. The (MeO)3Al·OO complex is the key intermediate for the O2(b1Σ+g) generation. On the basis of the conducted energy analysis for the diagram (Fig. 8), the hypothesis about the elimination of O2(b1Σ+g) in adiabatic reaction on the singlet PES and in the catalytic reaction of the MeOOH decomposition, where (MeO)3Al is the catalyst, can be considered to be justified with quantum chemical methods.
To our knowledge, the established in the present study the electronically excited state of the dioxygen, O2(b1Σ+g), eliminated in the catalytic reaction of the MeOOH decomposition (the model system) is the first report about the chemical excitation of O2 to the so-called second singlet state, O2(b1Σ+g). Here, it should be noted that the O2(1Δg) state is doubly degenerate. This phenomenon is new for the chemistry of peroxides and for catalysis on transition and non-transition metal centers. The catalytic oxidation of various substrates by the system t-BuOOH-transition metal compound67–73 conducted in the atmospheric O2, the activation of O2(X3Σ−g) and the formation of 1:
1 adducts of catalyst with O2(1Δg) as well as the O2(1Δg) photoinduced elimination74,75 have been shown previously. While the reactivity of O2(1Δg) has been studied for about 80 years,76–80 the chemistry of O2(b1Σ+g) is not essentially defined. Indeed, firstly, the lifetime of O2(b1Σ+g) determined by the spin-allowed transition O2(b1Σ+g) → O2(1Δg)75,79,81,82 is considerably shorter than the lifetime of O2(1Δg), i.e., the quenching79,83–86 is quantitative. Second, O2(b1Σ+g) is quantitatively deactivated through the electron–vibrational (e–v) mechanism unambiguously determined in ref. 83, 84 and 87. Moreover, as was shown in the earlier studies,83,84,87 the oxidation of substrates by O2(b1Σ+g) with C–H bonds becomes evading in the presence of the highly effective pathway of quenching.
Probably, the η1-(MeO)3Al·OO complex can show reactivity which could mimic O2(b1Σ+g), but in reaction pathway with bypass of O2(b1Σ+g) quenching. The selective oxidation12,13 of C–H bonds by the reactive forms of oxygen generated on the atoms of non-transition metals is a promising alternative to the developed catalytic systems on transition metals.88
The singlet spin coupling is the ground state for (MeO)3Al since, as was found at the B3LYP/cc-pVTZ theory level, the energy of the singlet–triplet transition (T1 ← S0) is (564.5 kJ mol−1). Thus, the asymptotics corresponding to the isolated monomers 3(MeO)3Al + 3O2 is not reachable from 11 with any remarkable thermodynamic probability. However, the singlet state is not the ground electronic state for 11 (Fig. 4), the triplet state is −79.0 kJ mol−1 more stable energetically than the lowest lying singlet state of the structure 11. Structure 16 (ESI, Fig. 1S†) is obtained in the geometry optimization of 11 (Fig. 4) on the triplet PES. The singlet state is the ground state for mono-η2-, bis-η2-methyl peroxides, and ozonides as was predicted at the B3LYP/cc-pVTZ theory level.
The search for the degeneracy of the singlet and triplet PESs for 3a, 7, 10, and 11 (Fig. 2 and 4) revealed minimum energy crossing points and, thus, non-stationary points of PES which determine shape of PES and non-adiabatic pathways for the net catalytic reaction. Structures corresponding to the MECPs are denoted as 3a-MECP, 7-MECP, 10-MECP, and 11-MECP (Fig. 9). Their energies are given: 98.5, 91.5, 2.4, and 10.5 kJ mol−1 at the B3LYP/cc-pVTZ theory level relative to 3a, 7, 10, and 11, respectively.
![]() | ||
Fig. 9 Structures corresponding to the minimum energy crossing points for the singlet and triplet PESs optimized at the B3LYP/cc-pVTZ theory level. Bond lengths are in Å. |
3a-MECP is a critical point which is not a stationary point of the singlet or triplet PESs for the AlOO cycle opening in 3a (Fig. 2) with the formation of structure 3b (Fig. 10). The structure 3b is 58.1 kJ mol−1 less stable than 3a (the singlet ground state). The decomposition of 3b with the formation of O(3P) and (MeO)3Al occurs with the reaction energy (Er) of 50.5 kJ mol−1. However, for the structure with two AlOO cycles, 7, the singlet–triplet spin recoupling does not lead to the cycle opening, i.e., in 7b (Fig. 10) with the energy of 75.8 kJ mol−1 above the level of 7 (Fig. 1) was located. The O–O bond lengths optimized at the B3LYP/cc-pVTZ theory level are 2.301 Å and 1.501 Å while in 7 the bonds are of 2.020 Å and 1.506 Å. Spin conversions via 10-MECP (Ea,T←S = 2.4 kJ mol−1) and 11-MECP (Ea,T←S = 10.5 kJ mol−1) result in the strongly exothermic decomposition reaction 11 → (MeO)3Al + 3O2 (Er = −119.8 kJ mol−1). The stabilization of weakly bound complex 16 (ESI, Fig. 1S†) is unlikely under such conditions and the reaction leads to the elimination of the (MeO)3Al and monomers. The asterisk denotes here and below a dioxygen molecule with the excess in energy. In this case, the O2 molecule is in the ground electronic state, but it is also vibrationally and/or translationally excited, i.e.,
can show higher or probably uncommon reactivity as compared to O2 thermalized under the (mild) conditions of the synthesis.12,13,89 Moreover, these non-adiabatic reactions show the possibility of incomplete conversion of 11 in (MeO)3Al and O2(b1Σ+g); thus, in addition to physical quenching in O2(b1Σ+g) → O2(1Δg) and O2(1Δg) → O2(X3Σ−g), it is a reaction pathway without the generation of 1O2. It agrees with the experimentally determined13 incomplete (about 50%) conversion of dioxygen in the reaction with singlet dioxygen trap, 9,10-methylanthracene, in weakly quenching solvent. The probability of low-energy singlet–triplet transition (the activation energies Ea,T←S(10-MECP) = 2.4 kJ mol−1 and Ea,T←S(11-MECP) = 10.5 kJ mol−1, and Ea(11 → (MeO)3Al + O2(b1Σ+g)) = 53.3 kJ mol−1 are given for comparison) makes it possible to assume the possibility of the process controlling by selection of the solvent, i.e., stimulation of the adiabatic decomposition reaction and, thus, increase of the O2(b1Σ+g) yield.
![]() | ||
Fig. 10 Structures for triplet PES optimized at the B3LYP/cc-pVTZ theory level. Bond lengths are in Å. |
The reaction η1-(MeO)3Al·OO → (MeO)3Al + O2 is a neither single-pathway nor product spin-pure process, so it is possible that the reactivity of O2(b1Σ+g)/O2(1Δg) (the latter one is the product of the spin-allowed quenching of O2(b1Σ+g)) and chemically activated (translationally–vibrationally–rotationally hot) 3O2, which is the product of the spin-forbidden reaction, may be observed for the t-BuOOH/(t-BuO)3Al system of experimental interest.
The conducted simulation makes it possible to propose that the catalytic decomposition of t-BuOOH leads to a complicated oxidizer system being the source of the reactive oxygen species: mono-η2-, bis-η2-peroxides, isomeric ozonides, η1-peroxide, chemically activated and chemically excited (O2(b1Σ+g) and O2(1Δg)) dioxygen. O2(1Δg) is the product of the decay of the higher excited state. O2(b1Σ+g) is characterized by the significantly shorter lifetime than O2(1Δg). The latter is the product of the spin-allowed transition decomposition of O2(b1Σ+g). One can believe that quenching through the e–v mechanism for O2(b1Σ+g) can create a pathway to exhibit an uncommon reactivity of O2(1Δg) which results from the chemical activation of the substrate by energy excess released in the O2(1Δg) ← O2(b1Σ+g) relaxation. The formation of the (t-BuO)3Al·O2 complex analogous to (MeO)3Al·O2 predicted here on the basis of quantum chemical study can determine the reactivity of chemically immobilized O2(b1Σ+g). Both pathways make the new chemistry of 1O2 possible, that is, the oxidizing activity results directly or indirectly from the reactivity of O2(b1Σ+g). Thus, it may determine important enhancement to famous Fenton and Haber–Weiss chemistry.
Footnote |
† Electronic supplementary information (ESI) available: Optimized Cartesian coordinates for reported intermediates, transition states, and minimum energy crossing points; the data for the focal point analysis; the reaction pathways for the peroxides dissociation, and additional comments on chemical reactivity of the catalytic system. See DOI: 10.1039/c6ra21471a |
This journal is © The Royal Society of Chemistry 2016 |