DOI:
10.1039/C5RA27576H
(Paper)
RSC Adv., 2016,
6, 26441-26450
Global picture of isomerization and dissociation of CN2O2: new metastable isomers†
Received
23rd December 2015
, Accepted 29th February 2016
First published on 1st March 2016
Abstract
Molecules containing carbon (C), nitrogen (N) and oxygen (O) atoms have received considerable attention due to their great relevance in astrophysical, atmospheric and high-energy-density material (HEDM) realms. Greatly differing from most studies that mainly center on the low-lying isomers, study of C, N and O systems appeals for the additional consideration of high-energy species. Thus, understanding the thermodynamic and kinetic stability of diversified isomers is of vital importance to assess their role in these processes. In this work, we investigated a pentatomic CN2O2 system, the isomeric study of which was initiated 20 years ago and up to now its three isomers OCNNO, CNNO2 and NCNO2 have been experimentally characterized. Based on our global search strategies for both isomers and transition states, we constructed hitherto the most comprehensive potential energy surface of CN2O2, covering 15 new isomers and 29 new transition states. The ring-containing isomers, i.e., 14, 22 and 29, were shown to possess considerable rate-determining Gibbs free energy barriers with respect to the radical–radical (P3 NCO + NO, P6 3NCN + 3O2 or P10 3NNC + 3O2) and lowest-energy product (P1 CO2 + N2) at the (U)CCSD(T)/CBS level. Thus, they are expected to be experimentally observable. After the experimentally known OCNNO, CNNO2 and NCNO2, the presently found three isomers 14, 22 and 29 warmly welcome future laboratory investigations. In addition, for CNNO2 09, we located a previously unreported transition state, which provides a new viewpoint on its kinetic stability.
1. Introduction
Carbon (C), nitrogen (N) and oxygen (O) are key elements both on Earth and in space. They can form diversified organic or inorganic species that are important intermediates during atmospheric and interplanetary processes. Detailed isomerism and decomposition knowledge of the involved C, N, O-containing molecules is thus very useful for assessment of their relevance in these processes. On the other hand, C, N and O can form simple yet very stable molecules with sp-hybridized multiple bonding, e.g., CO, N2 and CO2, with bond energies of 256.5, 225.4 and 382.3 kcal mol−1, respectively, whereas their corresponding single bond energies are 85.6 and 40.0 kcal mol−1.1 This feature has made C, N, O very attractive building elements in designing the so-called “high-energy density material (HEDM)” molecules, which upon decomposition to the stable products, i.e., N2, CO2, CO with a big amount of energy can be released.
Due to the importance in diverse fields, molecules comprising the three elements (C, N, and O) have been extensively explored either in theory or in experiment.2–27 However, successful characterization of such species in laboratory have still been very limited. The possible obstacles should be (1) absence of the promising research targets, and (2) the difficulty in finding suitable synthetic precursors due to the exotic nature (i.e., most structures are high energy and would easily cause explosion). Fortunately, the modern computational chemistry has played increasing roles in resolving the first problem. By exploring the structures and stability (intrinsic and intermolecular) of various [Cx, Ny, Oz] isomers, chemists can usually tell which isomer deserves to be studied or not, or which isomer is of little likelihood to be observed. Such information should be very useful for synthetic chemists since costing limited time on a sound target is always economic. Nice examples include diazirinone CN2O-A and nitryl cyanide CN2O2-A (see Scheme 1), both of which were initially predicted by computations and about 17 years later verified by experiments.5,6,9,23,24
 |
| Scheme 1 The reported isomers of [C, N2, O] and [C, N2, O2] in literatures. | |
We have very recently developed a global search strategy for both isomers and transition states of molecular systems, which aims to provide a very comprehensive potential energy surface (PES) for diverse necessities. In this work, we report its application on a penatomic molecule [C, N2, O2] system, whose isomerism was studied about 20 years ago.6,9 We constructed up to present the most extensive PES of CN2O2. Besides reproducing the previously studied one complex, 17 isomers and 31 transition states, our PES covers 15 new isomers and 29 new transition states. In addition to the experimentally known isomers CN2O2-A, CN2O2-B, and CN2O2-C (see Scheme 1), the new PES led us to predict three novel viable ring-containing isomers, i.e., 14, 22 and 29, each of which have the considerable rate-determining barriers at the (U)CCSD(T)/CBS level. For the experimentally known CNNO2 09, we found a new transition state that should provide a renewed view on its kinetic stability.
2. Computational methods
2.1 Structural search strategy for isomers and transition states
To construct a comprehensive potential-energy surface (PES) covering various isomeric and decomposition pathways, it's indispensable to obtain isomers and transition states as many as possible. The flow chart for constructing CN2O2 PES is shown in Scheme 2. The isomeric search was realized via our locally developed “grid-based isomeric search strategy”, which has proven very effective to help generating diverse structures including the global minimum point for small to medium-sized systems.28 The B3LYP/3-21G search was first carried out, followed by the B3LYP/6-31G(d) recalculation for geometries and frequencies on distinct structures. In addition, we have developed an automatic strategy that could help fast and comprehensively find transition states.29 For the target CN2O2 molecule, there are two kinds of transition states, both interconversion and fragmentation. For the first type transition states, we applied the “QST2” algorithm. In a typical QST2 calculation, elements in both isomers need to have the same atomic sequence. However, for isomers that are generated in global search and their atoms do not correspond to each other, we have to carefully adjust the sequential number of each atom within isomer. This is quite tedious especially for a number of isomeric pairs. For the second type transition states, we let the computer to automatically consider the fragmentation with extrusion of relatively stable fragments such as N2, CO2, NO and so on. Note that the large scale search for transition states between isomers would sometimes lead to the decomposition transition states. While transition state search is always challenging, we expect that such a hybrid strategy could help us find transition states as many as possible. The nature of each transition state structure was checked by the frequency calculations (at the same level) used in optimization (a transition state has only one imaginary frequency). The connection of each transition state was determined by the intrinsic reaction coordinate (IRC) calculations.30
 |
| Scheme 2 The strategy for constructing CN2O2 potential energy surface. | |
2.2 Electronic structure energy calculations
The geometries and frequencies of each isomeric and transition state structure in singlet state were refined at the B3LYP/aug-cc-pVTZ optimization level followed by the CCSD(T)/aug-cc-pVTZ single-point energy calculations. The main singlet isomers and transition states were subjected to the CCSD(T)/aug-cc-pVQZ single-point energy calculations and CBS extrapolation.31 For all the obtained singlet isomers, we carried out the “stability” analysis of the wavefunction at the B3LYP/6-31G(d) level. If an isomer is “unstable”, we used the “broken-symmetry” method for the refined geometry optimization and single-point energy calculation.32 All the calculations were carried out with the GAUSSIAN03 (ref. 33a) and GAUSSIAN09 (ref. 33b) program packages.
For the energetics of key structures, we applied a modified version of CASPT2 (Complete Active Space with Second-order Perturbation Theory, developed by Celani and Werner,34 referred to as ‘RS2C’ in Molpro), which accounts for dynamic correlation, using the CASSCF wave functions as references in the RS2C calculation, active spaces including 16 electrons and 13 active orbitals, namely CASPT2(16e,13o). The CASPT2 calculations were made without symmetry constraints on the wavefunction and carried out with the Molpro 2010 (ref. 35) program package.
3. Results and discussion
The energy and relative energy of isomers, transition states, decomposition products to P1 at the B3LYP/aug-cc-pVTZ + ZPVE CCSD(T)/aug-cc-pVTZ//B3LYP/aug-cc-pVTZ + ZPVE and CCSD(T)/aug-cc-pVTZ//B3LYP/aug-cc-pVTZ + GFE levels were shown in the Tables 1, 2 and 3, respectively. Note that the symbols “+ZPVE” and “+GFE” were used for the zero-point vibrational energy (ZPVE) correction and the Gibbs free energy (GFE) correction, respectively. In the present work, the GFE values were associated with the standard condition, i.e., temperature is 298.150 Kelvin and pressure is 1.00000 atm. The structures of CN2O2 isomers are shown in Scheme 3. Cartesian coordinates, energy and lowest frequency of all the considered isomers and transition states can be found in the ESI (SI1 and SI2†). The comprehensive potential energy surface of CN2O2 is shown in Fig. 1, based on the CCSD(T)/aug-cc-pVTZ//B3LYP/aug-cc-pVTZ + GFE energies. Note that the isomers and transition states with “u” indicate that they are performed with “broken-symmetry” method.32 For simplicity, the total energy of the products CO2 and N2 is set to be zero as reference. For ease of discussion, we only list the isomers and transition states relevant to 14, 22 and 29 in Fig. 2. The high-level calculations of main competitive isomerization pathways of 14, 22 and 29 are collected in Table 4. Note that unless specified, the following discussions are based on the Gibbs free energies corrected CCSD(T)/aug-cc-pVTZ//B3LYP/aug-cc-pVTZ single-point energy.
Table 1 The energy and relative energy of CN2O2 isomers to P1 (CO2 + N2) at the B3LYP/aug-cc-pVTZ level with inclusion of ZPVE correction, CCSD(T)/aug-cc-pVTZ//B3LYP/aug-cc-pVTZ level with inclusion of ZPVE correction and with Gibbs free energy correction. For concise, “AVTZ” and “AVQZ” are denoted for “aug-cc-pVTZ” and “aug-cc-pVQZ”, respectively
Isomers |
B3(AVTZ) + ZPVE (a.u.) |
RE (kcal mol−1) |
SP(AVTZ) + ZPVE (a.u.) |
RE (kcal mol−1) |
SP(AVTZ) + GFE (a.u.) |
RE (kcal mol−1) |
01 |
−298.054147 |
102.0 |
−297.5290574 |
109.5 |
−297.5567864 |
116.7 |
02 |
−298.047306 |
106.3 |
−297.5234721 |
113.0 |
−297.5519051 |
119.7 |
03 |
−298.030635 |
116.8 |
−297.5199457 |
115.2 |
−297.5467917 |
122.9 |
04 |
−298.021708 |
122.4 |
−297.5023993 |
126.3 |
−297.5303723 |
133.3 |
05 |
−298.020094 |
123.4 |
−297.4998182 |
127.9 |
−297.5278612 |
134.8 |
06 |
−297.986086 |
144.7 |
−297.4693048 |
147.0 |
−297.4965738 |
154.5 |
07 |
−297.988805 |
143.0 |
−297.4622678 |
151.4 |
−297.4917088 |
157.5 |
08 |
−297.949907 |
167.4 |
−297.4225379 |
176.4 |
−297.4501279 |
183.6 |
09 |
−297.94109 |
173.0 |
−297.4214629 |
177.0 |
−297.4500369 |
183.7 |
10 |
−297.936069 |
176.1 |
−297.4172773 |
179.7 |
−297.4451493 |
186.7 |
11 |
−297.931664 |
178.9 |
−297.4116656 |
183.2 |
−297.4397766 |
190.1 |
12 |
−297.919846 |
186.3 |
−297.4070408 |
186.1 |
−297.4341028 |
193.7 |
13 |
−297.91447 |
189.7 |
−297.4071216 |
186.0 |
−297.4335986 |
194.0 |
14 |
−297.913821 |
190.1 |
−297.4062288 |
186.6 |
−297.4330208 |
194.3 |
15 |
−297.913592 |
190.2 |
−297.4048519 |
187.5 |
−297.4326859 |
194.5 |
16 |
−297.910153 |
192.4 |
−297.4019661 |
189.3 |
−297.4317111 |
195.2 |
17 |
−297.900834 |
198.2 |
−297.3859067 |
199.4 |
−297.4132117 |
206.8 |
18 |
−297.897324 |
200.4 |
−297.3812452 |
202.3 |
−297.4083372 |
209.8 |
uL1 |
−297.904759 |
195.8 |
−297.3717207 |
208.3 |
−297.4032677 |
213.0 |
19 |
−297.894352 |
202.3 |
−297.3700767 |
209.3 |
−297.3993637 |
215.5 |
20 |
−297.88998 |
205.0 |
−297.368442 |
210.3 |
−297.395602 |
217.8 |
u21 |
−297.885358 |
207.9 |
−297.366762 |
211.4 |
−297.392639 |
219.7 |
22 |
−297.875816 |
213.9 |
−297.3654243 |
212.2 |
−297.3925703 |
219.7 |
23 |
−297.864732 |
220.9 |
−297.3399851 |
228.2 |
−297.3669611 |
235.8 |
24 |
−297.860692 |
223.4 |
−297.3386707 |
229.0 |
−297.3661417 |
236.3 |
25 |
−297.853936 |
227.7 |
−297.3382542 |
229.3 |
−297.3658962 |
236.5 |
26 |
−297.847845 |
231.5 |
−297.3350357 |
231.3 |
−297.3649687 |
237.0 |
u27 |
−297.863123 |
221.9 |
−297.3329303 |
232.6 |
−297.3607383 |
239.7 |
28 |
−297.847086 |
232.0 |
−297.3307853 |
233.9 |
−297.3580663 |
241.4 |
29 |
−297.830208 |
242.6 |
−297.3188541 |
241.4 |
−297.3462981 |
248.8 |
30 |
−297.7424 |
297.7 |
−297.2430254 |
289.0 |
−297.2694544 |
297.0 |
31 |
−297.740295 |
299.0 |
−297.222485 |
301.9 |
−297.250141 |
309.1 |
32 |
−297.712826 |
316.2 |
−297.2113642 |
308.9 |
−297.2378492 |
316.8 |
Table 2 The energy and relative energy of CN2O2 transition states to P1 (CO2 + N2) at the B3LYP/aug-cc-pVTZ level with inclusion of ZPVE correction, CCSD(T)/aug-cc-pVTZ//B3LYP/aug-cc-pVTZ level with inclusion of ZPVE correction and with Gibbs free energy correction. For concise, “AVTZ” and “AVQZ” are denoted for “aug-cc-pVTZ” and “aug-cc-pVQZ”, respectively
TSs |
B3(AVTZ) + ZPVE (a.u.) |
RE (kcal mol−1) |
SP(AVTZ) + ZPVE (a.u.) |
RE (kcal mol−1) |
SP(AVTZ) + GFE (a.u.) |
RE (kcal mol−1) |
The topomeric forms of 08 are not listed in PES but in SI2.† |
ts01/02 |
−298.045519 |
107.4 |
−297.5219222 |
114.0 |
−297.5492292 |
121.4 |
ts03/P1 |
−298.031282 |
116.4 |
−297.5195416 |
115.5 |
−297.5461446 |
123.4 |
ts02/03 |
−298.017494 |
125 |
−297.5016698 |
126.7 |
−297.5283608 |
134.5 |
ts04/05 |
−298.011543 |
128.8 |
−297.4923139 |
132.6 |
−297.5199559 |
139.8 |
ts02/P2 |
−298.008457 |
130.7 |
−297.4877493 |
135.5 |
−297.5157103 |
142.5 |
ts01/05 |
−298.007081 |
131.6 |
−297.4850535 |
137.1 |
−297.5124495 |
144.5 |
ts01/07 |
−297.985744 |
145 |
−297.4605825 |
152.5 |
−297.4887755 |
159.4 |
ts01/P2 |
−297.975001 |
151.7 |
−297.4493664 |
159.5 |
−297.4764404 |
167.1 |
ts01/P4 |
−297.949902 |
167.4 |
−297.4382686 |
166.5 |
−297.4667876 |
173.2 |
ts12/P3 |
−297.920068 |
186.2 |
−297.4038333 |
188.1 |
−297.4305543 |
195.9 |
ts04/10 |
−297.917268 |
187.9 |
−297.4041609 |
187.9 |
−297.4304909 |
195.9 |
ts01/15 |
−297.913552 |
190.3 |
−297.4026658 |
188.8 |
−297.4301268 |
196.2 |
ts10/11 |
−297.920225 |
186.1 |
−297.4015921 |
189.5 |
−297.4293521 |
196.6 |
ts08/08a |
−297.935651 |
176.4 |
−297.4024652 |
189.0 |
−297.4292532 |
196.7 |
ts15/16 |
−297.908406 |
193.5 |
−297.3989044 |
191.2 |
−297.4260644 |
198.7 |
1-uts13/P1 |
−297.900473 |
198.5 |
−297.3971805 |
192.3 |
−297.4237435 |
200.2 |
ts06/09 |
−297.915282 |
189.2 |
−297.3930043 |
194.9 |
−297.4202313 |
202.4 |
ts12/16 |
−297.901095 |
198.1 |
−297.3895949 |
197.0 |
−297.4165549 |
204.7 |
ts08/11 |
−297.907297 |
194.2 |
−297.3848706 |
200.0 |
−297.4124136 |
207.3 |
ts05/06 |
−297.901678 |
197.7 |
−297.3851160 |
199.9 |
−297.412422 |
207.3 |
1-uts01/14 |
−297.887675 |
206.5 |
−297.3901634 |
206.2 |
−297.4019544 |
213.8 |
ts12/14 |
−297.888037 |
206.3 |
−297.3750301 |
206.2 |
−297.4015961 |
214.1 |
ts16/21 |
−297.846176 |
232.5 |
−297.3732073 |
207.3 |
−297.3998253 |
215.2 |
ts01/09 |
−297.886567 |
207.2 |
−297.3717622 |
208.2 |
−297.3990172 |
215.7 |
ts20/P1 |
−297.890666 |
204.6 |
−297.3705111 |
209.0 |
−297.3979011 |
216.4 |
ts11/19 |
−297.892366 |
203.5 |
−297.3676326 |
210.8 |
−297.3957896 |
217.7 |
ts08/L1 |
−297.882507 |
209.7 |
−297.3594521 |
216.0 |
−297.3889251 |
222.0 |
ts08/15 |
−297.884127 |
208.7 |
−297.3605757 |
215.3 |
−297.3881257 |
222.5 |
2-ts13/P1 |
−297.86778 |
219 |
−297.3603760 |
215.4 |
−297.387228 |
223.1 |
2-ts01/14 |
−297.8667 |
219.7 |
−297.3716116 |
217.6 |
−297.3839016 |
225.2 |
ts14/P4 |
−297.870283 |
217.4 |
−297.3533869 |
219.8 |
−297.3807539 |
227.1 |
ts14/P1 |
−297.863369 |
221.7 |
−297.3540405 |
219.4 |
−297.3808835 |
227.1 |
ts08/16 |
−297.871989 |
216.3 |
−297.3483385 |
222.9 |
−297.3756935 |
230.3 |
ts17/P3 |
−297.858264 |
224.9 |
−297.3426760 |
226.5 |
−297.369327 |
234.3 |
ts08/17 |
−297.853795 |
227.8 |
−297.3422946 |
226.7 |
−297.3684606 |
234.9 |
ts02/25 |
−297.85286 |
228.3 |
−297.3392072 |
228.7 |
−297.3659232 |
236.4 |
ts11/26 |
−297.847035 |
232 |
−297.3353634 |
231.1 |
−297.3641744 |
237.5 |
ts17/P6 |
−297.870679 |
217.2 |
−297.3347392 |
231.5 |
−297.3628292 |
238.4 |
uts06/22 |
−297.852706 |
228.4 |
−297.3488285 |
231.3 |
−297.3629875 |
238.8 |
ts18/22 |
−297.835564 |
239.2 |
−297.3328865 |
232.6 |
−297.3601975 |
240.0 |
ts18/P6 |
−297.869776 |
217.7 |
−297.331627 |
233.4 |
−297.359644 |
240.4 |
ts13/24 |
−297.845785 |
232.8 |
−297.3299731 |
234.5 |
−297.3567531 |
242.2 |
ts17/22 |
−297.830701 |
242.2 |
−297.3295360 |
234.7 |
−297.356736 |
242.2 |
uts27/P8 |
−297.862916 |
222 |
−297.3287498 |
235.2 |
−297.3566748 |
242.2 |
ts17/18 |
−297.83949 |
236.7 |
−297.3176143 |
242.2 |
−297.3448003 |
249.7 |
uts28/P10 |
−297.847991 |
231.4 |
−297.3160832 |
243.2 |
−297.3442162 |
250.1 |
ts23/P8 |
−297.839435 |
236.8 |
−297.3156450 |
243.4 |
−297.343159 |
250.7 |
1-ts25/29 |
−297.782175 |
272.7 |
−297.2909349 |
259.0 |
−297.3182789 |
266.3 |
1-ts28/29 |
−297.795646 |
264.2 |
−297.2894133 |
259.9 |
−297.3169383 |
267.2 |
uts09/29 |
−297.80562 |
258 |
−297.2865597 |
261.7 |
−297.3144117 |
268.8 |
1-ts22/29 |
−297.777762 |
275.5 |
−297.2677147 |
273.5 |
−297.2953067 |
280.8 |
ts08/P6 |
−297.766969 |
282.2 |
−297.2664989 |
274.3 |
−297.2932849 |
282.0 |
ts02/29 |
−297.76347 |
284.4 |
−297.2529290 |
282.8 |
−297.279542 |
290.6 |
ts13/30 |
−297.742716 |
297.5 |
−297.2579355 |
289.0 |
−297.2691775 |
297.2 |
2-ts22/29 |
−297.741011 |
298.5 |
−297.2416527 |
299.2 |
−297.2536417 |
306.9 |
2-ts28/29 |
−297.738615 |
300 |
−297.2363483 |
302.5 |
−297.2483183 |
310.2 |
ts31/P3 |
−297.733389 |
303.3 |
−297.2321516 |
304.9 |
−297.2452716 |
312.2 |
2-ts25/29 |
−297.731033 |
304.8 |
−297.2272537 |
308.2 |
−297.2393027 |
315.9 |
ts32/P8 |
−297.704528 |
321.4 |
−297.2145292 |
315.2 |
−297.2286342 |
322.6 |
ts08/32 |
−297.706773 |
320 |
−297.2155115 |
314.9 |
−297.2282015 |
322.9 |
Table 3 The energy and relative energy of decomposition products to P1 (CO2 + N2) at the B3LYP/aug-cc-pVTZ level with inclusion of ZPVE correction, CCSD(T)/aug-cc-pVTZ//B3LYP/aug-cc-pVTZ level with inclusion of ZPVE correction and with Gibbs free energy correction. For concise, “AVTZ” and “AVQZ” are denoted for “aug-cc-pVTZ” and “aug-cc-pVQZ”, respectively
Products |
B3(AVTZ) + ZPVE (a.u.) |
RE (kcal mol−1) |
SP(AVTZ) + ZPVE (a.u.) |
RE (kcal mol−1) |
SP(AVTZ) + GFE (a.u.) |
RE (kcal mol−1) |
P1:CO2 + N2 |
−298.216738 |
0.0 |
−297.7036027 |
0.0 |
−297.7427197 |
0.0 |
P2:1NNO + CO |
−298.08066 |
85.4 |
−297.5581084 |
91.3 |
−297.5985224 |
90.5 |
P3:NCO + NO |
−297.998516 |
136.9 |
−297.4736023 |
144.3 |
−297.5155833 |
142.5 |
P4:c-1NNO + CO |
−297.972684 |
153.1 |
−297.4598074 |
153.0 |
−297.5020054 |
151.1 |
P5:1COO + N2 |
−297.956742 |
163.2 |
−297.4426592 |
163.7 |
−297.4828662 |
163.1 |
P6:3NCN + 3O2 |
−297.865818 |
220.2 |
−297.3871418 |
198.6 |
−297.4278338 |
197.6 |
P7:1NON + CO |
−297.896453 |
201.0 |
−297.3823147 |
201.6 |
−297.4227407 |
200.8 |
P8:2CN + 2NO2 |
−297.893737 |
202.7 |
−297.3739819 |
206.8 |
−297.4176549 |
204.0 |
P9:2CNO + NO |
−297.900427 |
198.5 |
−297.3760993 |
205.5 |
−297.4127913 |
207.0 |
P10:3NNC + 3O2 |
−297.865818 |
220.2 |
−297.3407606 |
227.7 |
−297.3830766 |
225.7 |
 |
| Scheme 3 The geometric structures of CN2O2 isomers. | |
 |
| Fig. 1 The comprehensive potential energy surface of CN2O2 which is based on the CCSD(T)/aug-cc-pVTZ//B3LYP/aug-cc-pVTZ + GFE energies. | |
 |
| Fig. 2 Schematic decomposition and isomerization pathways of stable isomers 14, 22 and 29 in singlet which is based on the CCSD(T)/aug-cc-pVTZ//B3LYP/aug-cc-pVTZ + GFE energies. | |
Table 4 The main competitive isomerization pathways of 14, 22 and 29 in singlet. For concise, “AVTZ” and “AVQZ” are denoted for “aug-cc-pVTZ” and “aug-cc-pVQZ”, respectively
|
T1Diag |
<S2> |
CCSD(T)/AVTZ//B3LYP/AVTZ |
CCSD(T)/AVQZ//B3LYP/AVTZ |
CCSD(T)/CBS//B3LYP/AVTZ |
CASPT2/AVTZ//B3LYP/AVTZ |
14 |
0.017 |
0 |
0.0 |
0.0 |
0.0 |
0.0 |
1-uts01/14 (to P1 (CO2 + N2)) |
0.039 |
0.640732 |
19.5 |
19.7 |
19.8 |
19.2 |
ts12/14 (to P3 (NCO + NO)) |
0.038 |
0 |
19.8 |
19.8 |
19.8 |
20.2 |
22 |
0.016 |
0 |
0.0 |
0.0 |
0.0 |
0.0 |
uts06/22 (to P8 (CN + NO2)) |
0.106 |
0.854292 |
19.1 |
18.9 |
19.2 |
12.6 |
ts17/22 (to P6 (3NCN + 3O2)) |
0.089 |
0 |
22.5 |
23.1 |
23.5 |
24.0 |
ts18/22 (to P6 (3NCN + 3O2)) |
0.079 |
0 |
20.3 |
20.9 |
21.3 |
20.4 |
29 |
0.017 |
0 |
0.0 |
0.0 |
0.0 |
0.0 |
uts09/29 (to P8 (CN + NO2)) |
0.060 |
0.875304 |
20.0 |
20.4 |
20.6 |
14.2 |
1-ts25/29 (to P1 (CO2 + N2)) |
0.115 |
0 |
17.5 |
18.8 |
19.7 |
25.3 |
1-ts28/29 (to P10 (3NNC + 3O2)) |
0.049 |
0 |
18.4 |
18.9 |
19.3 |
18.3 |
Prior to the discussion, it's desirable to show two basic “stability” terms of an isomer, i.e., thermodynamic stability and kinetic stability. The former is simply based on the isomer's relative energy (better in form of the Gibbs free energy energy). The high-energy isomer is apt to undergo the conversion to the lower-energy one in thermodynamics. The higher the relative energy, the lower thermodynamic stability an isomer has. By contrast, the “kinetic stability” is usually determined by the barrier height of the most feasible process of an isomer. The larger the barrier height, the higher kinetic stability an isomer possesses under certain conditions. It has been suggested that “activation energy of approximately 15 kcal mol−1 is needed to ensure a half-life of a day at room temperature”.36 This value is used in this work to evaluate the likeliness for the observation (at least at the spectroscopic level) of an energy-rich isomer. Surely, for the use of HEDM, the rate-determining barrier of more than 30 kcal mol−1 (in form of Gibbs free energy) is preferable if the isomer has a very large energy release upon fragmentation.37
3.1 Overview of the CN2O2 potential energy surface
Ever since the study of the CN2O2 system was initiated 20 years ago, it has been explored with much interest from various aspects of interest both theoretically and experimentally.2–6,8,9,12,14,18–20,26 For example, the CN2O2 isomers are important intermediates in the reactions between CN and NO2,3,6,18,20 NCN and O2 (ref. 19 and 26) and CNO and NO.2,12,14 In addition, due to the rather large energy release with respect to the stable product N2 + CO2, some CN2O2 isomers have been proposed as potential high-energy density materials.4,5,8,9 Up to now, one weakly bound complex L1 and a total of 17 CN2O2 isomers (all in singlet state) have been reported, i.e., 01, 02, 03, 04, 05, 06, 08, 09, 10, 11, 13, 15, 16, 17, 18, 24, 32. All of them are covered in the comprehensive potential energy surface of CN2O2 (see Fig. 1) obtained via our global isomeric search strategy. Moreover, we located 15 new singlet isomers (07, 12, 14, 19, 20, 21, 22, 23, 25, 26, 27, 28, 29, 30, 31), of which 14, 22 and 29 possess considerable kinetic stability that will be discussed in detail below.
The transition states are very important to connect the intermediate isomers or fragments. They provide a detailed kinetic picture for the associated isomers. Based on our global transition state search strategy, we reproduced all the previously reported 31 singlet TSs. In addition, we located 29 new singlet TSs. Of particular note, ts06/09 is a new transition state for an experimentally characterized isomer CNNO2 09, and will re-determine the kinetic stability of 09. For simplicity, in the following parts, we mainly discuss the key or new structures related to the CN2O2 PES.
Usually, there is only one transition state connecting two species. It is of great interest to find out that for CN2O2, the interconversion between the isomeric pairs (01, 14), (22, 29) and (25, 29) each involve two transition states (see Scheme 4). The general difference in the latter two isomeric pairs is that one TS is structurally looser (lower-energy) than the other. For (01, 14), the two TSs mainly differ in the dihedral angle of the O-atom that is associated with the ring opening/closure. The detailed IRC curves can be found in SI3.†
 |
| Scheme 4 The involved transition states of interconversion between the isomeric pairs (01, 14), (22, 29) and (25, 29). | |
3.2 New transition state for isomer CNNO2 09
CNNO2 09 is the isocyano counterpart of NCNO2 and has been previously studied both experimentally7 and computationally.19 For its kinetic stability, we located three singlet TSs, i.e., ts01/09 (32.0 kcal mol−1), ts06/09 (18.7 kcal mol−1) and ts09/29 (85.1 kcal mol−1). The three TSs are associated with the 1, 2-CN transfer, CN ↔ NC exchange, and ring-closure of NO2, respectively. The first TS has been reported previously.19 The latter two were newly found in this work. The energy of the direct C–N cleavage to P8 CN + NO2 is 20.3 kcal mol−1. Notably, among all the possible evolution pathways, the CN ↔ NC exchange via ts06/09 is the most feasible, and is thus the rate-determining step for determining the kinetic stability of 09. With ts06/09, the barrier height of 09 was determined to be 18.7 kcal mol−1 at the CCSD(T)/aug-cc-pvtz//B3LYP/aug-cc-pVTZ with Gibbs free energy correction. It is worth mentioning that the new ts06/09 has important implications. In the previous studies,19 ts01/09 is the only evolution transition state for 09. Since ts01/09 lies 11.7 kcal mol−1 higher than the direct C–N fission product P8 CN + NO2, the most favorable channel of 09 was reported to be the direct decomposition to P8 CN + NO2 rather than to the low-lying isomer ONNCO 01. The considerable rate-determining Gibbs free energy barrier 18.7 kcal mol−1 indicates that the branched isomer CNNO2 09 would be experimentally viable. In fact, it has been characterized via the spectroscopic techniques.7
3.3 Isomers 14, 22 and 29
The singlet isomers 14, 22 and 29 were found in the present work for the first time. Isomer 14 is butterfly-shaped structure consisting with a N–N cross bond. Except the exocyclic C
O bond, all the bonds within 14 are single bonding, which contributes to its high-energy feature (194.3 kcal mol−1 above P1). 14 can be structurally viewed as a product of the O-addition to the N
N bond of diazirinone CN2O-A. Turned from the sp2-hybridization to sp3-hybridization, the two N-atoms become pyramidal each with one lone pair of electrons. The main competition channels of 14 are as follows:
Path 14-1: 14 → 12 → P3 (NCO + NO) (19.8 kcal mol−1); |
Path 14-2: 14 → 01 → 02 → 03 → P1 (CO2 + N2) (19.8 kcal mol−1). |
Note that the value in parenthesis is the overall barrier height (the energy difference between the reactant isomer and the highest-energy transition state) of the given pathway. With the radical–radical product P3, the energy release of Path 14-1 (48.2 kcal mol−1) is much less than that (194.3 kcal mol−1) of Path 14-2. Note that 12 was also a newly located isomer in this work and contains the singlet diradical character. Unfortunately, its kinetic stability is very low, i.e., 2.6 kcal mol−1, with respect to the ring-opening to P3. Kinetically, both Path 14-1 and Path 14-2 are highly competitive with almost equal barrier heights 19.8 kcal mol−1 at the CCSD(T)/CBS level with the Gibbs free energy correction. The large value 19.8 kcal mol−1 well suggests its high viability for experimental detection. In comparison to CN2O-A, the O-addition converts the N
N bond to N–N bond, which considerably increases the energy release, i.e., 99.6 kcal mol−1 for CN2O-A5 and 194.3 kcal mol−1 for 14. The cost of such energy-gaining is that the rate-determining barrier height is decreased from 27 kcal mol−1 for CN2O-A to 19.8 kcal mol−1 for 14.
22 and 29 can be seen as the cyano and isocyano-substituted oxaziridines (RNO2), both of which possess the same moiety c-NO2. The main competition channels of 22 and 29 can be described as follows:
Path 22-1: 22 → 06 → P8 (CN + NO2) (19.2 kcal mol−1); |
Path 22-2: 22 → 18 → P6 (3NCN + 3O2) (21.3 kcal mol−1); |
Path 22-3: 22 → 17 → P6 (3NCN + 3O2) (23.5 kcal mol−1); |
Path 29-1: 29 → 25 → 02 → 03 → P1 (CO2 + N2) (19.7 kcal mol−1); |
Path 29-2: 29 → 28 → P10 (3NNC + 3O2) (19.3 kcal mol−1); |
Path 29-3: 29 → 09 → 06 → P8 (CN + NO2) (20.6 kcal mol−1). |
The isomers 22 and 29 both have three competitive channels involving the decomposition to the lower-energy (P3 (NCO + NO)) and high-energy radical–radical product (P6, P10). The most kinetically favored channel of 22 is Path 22-1 with the rate-determining barrier 19.2 kcal mol−1 at the CCSD(T)/CBS level with Gibbs free energy correction. At the CCSD(T)/CBS level, dissociation to P10 is slightly more favorable with the rate-determining barrier 19.3 kcal mol−1. Clearly, both 22 and 29 isomers are expected to be kinetically stable and viable in experiment. Of particular interest, the Gibbs free energy calculations showed that for all the three newly located isomers, the thermodynamically most favored P1 is in fact less competitive than the radical–radical products.
3.4 Possible synthetic routes to 14, 22 and 29
By means of the structural comparison, 14 can be obtained via the O-addition to the N
N bond of diazirinone CN2O-A that was characterized in 2011.23 The isomers 22 and 29 belong to the family of dioxiranes (XO2) that are important intermediates in photochemical and thermal reactions.38 Many such dioxiranes have been experimentally characterized.38 An approach to the generation of XO2 is often based on the reaction of an oxidant with mono-, di-, or trivalent (neutral) centers, such as nitrenes, silylenes, phosphines, sulfides, selenides, and so on. Oxidants that have been used in the generation of these high-energy XO2 dioxiranes include lowest excited singlet state oxygen (1O2), ground state triplet oxygen (3O2), ozone (O3), superoxide ion (O2˙−), and hydrogen peroxide. This will provide the possibility for the synthesis of isomers 22 and 29. To assist their future laboratory characterization, the vibrational frequencies, rotational constants and dipole moments are provided in Table 5.
Table 5 The key spectroscopic parameters including wave numbers Mwav (cm−1), rotational constants R (GHz) and dipole moment D (Debye) of 14, 22 and 29 at the B3LYP/aug-cc-pVTZ level
|
Mwav (cm−1) |
R (GHz) |
D (Debye) |
14 |
184.0221, 508.1086, 668.2793, 673.5074, 755.5372, 883.4728, 956.2877, 1113.6246, 2003.2457 |
18.38810, 5.08526, 4.75150 |
0.0829 |
22 |
232.8269, 243.2824, 596.8890, 599.3739, 747.4999, 808.6776, 952.6085, 1161.5485, 2327.1165 |
20.49199, 3.98262, 3.71273 |
1.3492 |
29 |
160.5204, 195.5691, 532.1401, 560.6040, 713.7855, 825.4605, 908.7385, 1140.4474, 2131.7311 |
21.05931, 4.24523, 3.91255 |
1.0235 |
Further, we computed the diagnostic values (T1Diag) for the three new isomers and associated key transition states at the CCSD/aug-cc-pVTZ level. Clearly, as listed in Table 4, the isomers 14, 22 and 29 all have the T1Diag values smaller than 0.02, a recommended threshold by Lee.39 Yet, the T1Diag values of the associated transition states are far larger than 0.02, indicating a need for the energy calculation based on the multi-reference wavefunctions (Table 4). Interestingly, at the CASPT2(16e,13o)/aug-cc-pVTZ//(U)B3LYP/aug-cc-pVTZ level with GFE correction, the predicted barrier heights of 14 are very similar to those at the CCSD(T)/CBS level, both around 20 kcal mol−1. Yet, for 22, the O–O cleavage barrier is to some degree decreased from 19.2 (at CCSD(T)/CBS) to 12.6 kcal mol−1 (at CASPT2), while the barriers of the other paths are little changed. For 29, CASPT2 induces contrastive effect on the two ring-opening processes of the c-NO2 moiety. The O–O cleavage barrier is decreased from 20.6 (at CCSD(T)/CBS) to 14.2 kcal mol−1 (at CASPT2), whereas the N–O cleavage barrier to isomer 25 is increased from 19.7 (at CCSD(T)/CBS) to 25.3 kcal mol−1 (at CASPT2). It seems that the nondynamic calculations have great influence on the c-NO2 ring-opening kinetics.
Despite the very “small” size and the 20 year isomeric research history of CN2O2, the present PES study still newly find 15 isomers (3 are kinetically stable) and 29 transition states, almost one times the reported species, i.e., 17 isomers and 31 transition states. This surely merits from the application of our global search strategy for both isomers and transition states. The successfulness of our strategy depends on the operation of isomeric optimization (“grid” method28) and the transition state search (QST2 and fragmentation29) by trying all possible combinative possibilities. It is worthy of note that an effective global reaction route mapping (GRRM) program has been recently developed by Morokuma et al.40 Basing on the anharmonic downward distortion following (ADDF) and artificial force induced reaction (AFIR) methods, the program can automatically follow reaction pathways starting from local minima toward transition structures (TSs) and dissociation channels and then constructs a global potential energy surface (PES). The GRRM strategies have been applied to a variety of chemical systems ranging from thermal- and photochemical-reactions in small systems to organometallic- and enzymecatalysis, on the basis of quantum chemical calculations.40 Formally, our strategy follows “all isomers → all transition states”, while GRRM follows the sequential “isomer/product → transition state → isomer → …”. It is our belief that for constructing the isomerization/fragmentation potential energy surface of simple systems, the GRRM and ours could generate very similar reaction pictures.
Due to the great relevance of C, N, O-containing systems in combustion, in space as well as the interest in molecular HEDMs, the presently constructed comprehensive isomerization and fragmentation picture should provide a useful base for further understanding of the chemistry of various processes. Especially, the kinetic behavior of various isomers at the Gibbs free energy level could help establish the kinetic model of relevant isomers or reactions. Notably, previous studies on the kinetic stability of [C, N2, O2] isomers have all neglected the influence of Gibbs free energies (GFE), which are yet very important for some key processes. We are currently investigating the kinetic properties of some important radical–radical reactions, e.g., CN + NO2, NCO/CNO + NO, based on the present PES by means of the master equation rate constant calculations.
4. Conclusions
Based on our global search strategies for both isomers and transition states, the hitherto most comprehensive potential energy surface of CN2O2, covering 15 new isomers and 29 new transition states, was constructed. For the experimentally known CNNO2 09, we newly found that a direct –CN ↔ NC conversion to NCNO2 06 followed by the subsequent sequential dissociation to P8 CN + NO2 is its rate-determining step. Besides, three newly located isomers 14, 22 and 29 each possess the considerable rate-determining Gibbs free energy barriers at the (U)CCSD(T)/CBS level. They are expected to be experimentally observable. In addition to the experimentally known OCNNO, CNNO2 and NCNO2 species, the presently found three isomers 14, 22 and 29 should enrich the CN2O2 family and warmly welcome future laboratory investigations. Besides, the potential energy surface constructed at the Gibbs free energy level could provide a useful base for understanding the kinetic behavior of various chemical processes.
Acknowledgements
This work was funded by the National Natural Science Foundation of China (No. 21273093, 21473069, 21073074). The reviewers' invaluable comments and suggestions are greatly acknowledged.
References
-
(a) T. L. Cottrell, The Strengths of Chemical Bonds, 2nd edn, Butterwoth, London, 1958 Search PubMed;
(b) B. deB. Darwent, National Standard Reference Data Series, National Bureau of Standards, no. 31, Washington, 1970 Search PubMed;
(c) S. W. Benson, J. Chem. Educ., 1965, 42, 502 CrossRef CAS.
- G. A. McGibbon, C. A. Kingsmill, J. K. Terlouw and P. C. Burgers, Int. J. Mass Spectrom. Ion Processes, 1992, 121, R11 CrossRef CAS.
- M. Lin, Y. He and C. Melius, J. Phys. Chem., 1993, 97, 9124 CrossRef CAS.
- J. Park and J. F. Hershberger, J. Chem. Phys., 1993, 99, 3488 CrossRef CAS.
- A. A. Korkin, P. von Ragué Schleyer and R. J. Boyd, Chem. Phys. Lett., 1994, 227, 312 CrossRef CAS.
- A. Averyanov, Y. G. Khait and Y. V. Puzanov, J. Mol. Struct., 1996, 367, 87 CrossRef CAS.
- T. M. Klapötke, G. McIntyre and A. Schulz, J. Chem. Soc., Dalton Trans., 1996, 3237 RSC.
- T. M. Klapötke and A. Schulz, Inorg. Chem., 1996, 35, 7897 CrossRef.
-
(a) A. A. Korkin, J. Leszczynski and R. J. Bartlett, J. Phys. Chem., 1996, 100, 19840 CrossRef CAS;
(b) F. F. He, S. M. Gao, G. de Petris, M. Rosi and Y. H. Ding, J. Phys. Chem. A, 2016 DOI:10.1021/acs.jpca.5b12275.
- A. A. Korkin, A. Lowrey, J. Leszczynski, D. B. Lempert and R. J. Bartlett, J. Phys. Chem. A, 1997, 101, 2709 CrossRef CAS.
- G. Maier, H. P. Reisenauer, J. Eckwert, M. Naumann and M. De Marco, Angew. Chem., Int. Ed., 1997, 36, 1707 CrossRef CAS.
- J. Bylykbashi and E. Lewars, J. Mol. Struct., 1999, 469, 77 CrossRef CAS.
- R. Zhu and M. Lin, J. Phys. Chem. A, 2000, 104, 10807 CrossRef CAS.
- S. Dua and J. H. Bowie, J. Phys. Chem. A, 2003, 107, 76 CrossRef CAS.
- R. Prasad, J. Chem. Phys., 2004, 120, 10089 CrossRef CAS PubMed.
- C. S. Jamieson, A. M. Mebel and R. I. Kaiser, Phys. Chem. Chem. Phys., 2005, 7, 4089 RSC.
- Y. J. Wu and Y. P. Lee, J. Chem. Phys., 2005, 123, 174301 CrossRef PubMed.
- X. Zeng, M. Ge, Z. Sun and D. Wang, Inorg. Chem., 2005, 44, 9283 CrossRef CAS PubMed.
- J. X. Zhang, Z. S. Li, J. Y. Liu and C. C. Sun, J. Phys. Chem. A, 2005, 109, 10307 CrossRef CAS PubMed.
- R. Zhu and M.-C. Lin, Int. J. Chem. Kinet., 2005, 37, 593 CrossRef CAS.
- W. Feng and J. F. Hershberger, J. Phys. Chem. A, 2009, 113, 3523 CrossRef CAS PubMed.
- P. R. Schreiner, H. P. Reisenauer, E. Mátyus, A. G. Császár, A. Siddiqi, A. C. Simmonett and W. D. Allen, Phys. Chem. Chem. Phys., 2009, 11, 10385 RSC.
- X. Zeng, H. Beckers, H. Willner and J. F. Stanton, Angew. Chem., Int. Ed., 2011, 50, 1720 CrossRef CAS PubMed.
- M. Rahm, G. Bélanger-Chabot, R. Haiges and K. O. Christe, Angew. Chem., Int. Ed., 2014, 53, 6893 CrossRef CAS PubMed.
- F. F. He, X. Y. Zhang and Y. H. Ding, RSC Adv., 2015, 5, 46648 RSC.
- R. Zhu and M. Lin, J. Phys. Chem. A, 2007, 111, 6766 CrossRef CAS PubMed.
- N. Faßheber and G. Friedrichs, Int. J. Chem. Kinet., 2015, 47, 586 CrossRef.
-
(a) C. B. Shao and Y. H. Ding, Grid-Based Comprenhensive Isomeric Search Algorithm, Jilin University, Changchun, China, 2010 Search PubMed;
(b) Z. H. Cui, M. Contreras, Y. H. Ding and G. Merino, J. Am. Chem. Soc., 2011, 133, 13228 CrossRef CAS PubMed;
(c) S. M. Gao and Y. H. Ding, RSC Adv., 2012, 2, 11764 RSC;
(d) X. Y. Tang, Z. H. Cui, C. B. Shao and Y. H. Ding, Int. J. Quantum Chem., 2012, 112, 1299 CrossRef CAS;
(e) S. M. Gao and Y. H. Ding, RSC Adv., 2012, 2, 11764 RSC;
(f) C. Guo, Z. H. Cui and Y. H. Ding, Struct. Chem., 2013, 24, 263 CrossRef CAS;
(g) C. Guo, Z. H. Cui and Y. H. Ding, Int. J. Quantum Chem., 2013, 113, 2213 CrossRef CAS;
(h) C. Guo, C. Wang and Y. H. Ding, Struct. Chem., 2014, 25, 1023 CrossRef CAS;
(i) X. Y. Zhang and Y. H. Ding, Comput. Theor. Chem., 2014, 1048, 18 CrossRef CAS;
(j) X. Y. Zhang and Y. H. Ding, RSC Adv., 2015, 5, 27134 RSC.
- Yi-hong Ding, A comprehensive transition-state search strategy, Jilin University, Changchun, 2015 Search PubMed.
- K. Fukui, Acc. Chem. Res., 1981, 14, 363 CrossRef CAS.
- A. Halkier, T. Helgaker, P. Jørgensen, W. Klopper, H. Koch, J. Olsen and A. K. Wilson, Chem. Phys. Lett., 1998, 286, 243 CrossRef CAS.
-
(a) L. Noodleman, J. Chem. Phys., 1981, 74, 5737 CrossRef CAS;
(b) L. Noodleman and E. R. Davidson, Chem. Phys., 1986, 109, 131 CrossRef;
(c) A. Ovchinnikov and J. K. Labanowski, Phys. Rev. A, 1996, 53, 3946 CrossRef CAS PubMed;
(d) C. Adamo, V. Barone, A. Bencini, F. Totti and I. Ciofini, Inorg. Chem., 1999, 38, 1996 CrossRef CAS PubMed.
-
(a) M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery Jr, J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian 03, Revision D.02, Gaussian, Inc., Wallingford, CT, 2009 Search PubMed;
(b) M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, J. A. Montgomery Jr T. Vreven, K. N. Kudin, J. C. Burant, J. M. Millam, S. S. Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G. A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kita, H. Nakai, M. Klene, X. Li, J. E. Knox, H. P. Hratchian, J. B. Cro, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, P. Y. Ayala, K. Morokuma, G. A. Voth, P. Salvador, J. J. Dannenbe, V. G. Zakrzewski, S. Dapprich, A. D. Daniels, M. C. Strain, O. Farkas, D. K. Malick, A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. V. Ortiz, Q. Cui, A. G. Baboul, S. Clifford, J. Cioslowski, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. L. Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, M. Challacombe, P. M. W. Gill, B. Johnson, W. Chen, M. W. Wong, C. Gonzalez and J. A. Pople, Gaussian 09, Revision D.01, Gaussian, Inc., Wallingford, CT, 2004 Search PubMed.
- P. Celani and H.-J. Werner, J. Chem. Phys., 2000, 112, 5546 CrossRef CAS.
- H. J. Werner, P. J. Knowles, G. Knizia, F. R. Manby and M. Schütz, Molpro: A General Purpose Quantum Chemical Program Package, WIREs Computational Molecular Science, 2012, 2, 242 CrossRef CAS.
- R. Hoffmann, P. v. R. Schleyer and H. F. Schaefer, Angew. Chem. Int. Ed., 2008, 47, 7164 CrossRef PubMed.
- M. Rahm and T. Brinck, Chem.–Eur. J., 2010, 16, 6590 CrossRef CAS PubMed.
- N. Sawwan and A. Greer, Chem. Rev., 2007, 107, 3247 CrossRef CAS PubMed.
- T. J. Lee and P. R. Taylor, Int. J. Quantum Chem., 1989, 36, 199 CrossRef.
- S. Maeda, K. Ohno and K. Morokuma, Phys. Chem. Chem. Phys., 2013, 15, 3683 RSC.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra27576h |
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.