Skip to main content
ACS Omega logoLink to ACS Omega
. 2021 Sep 6;6(37):24280–24288. doi: 10.1021/acsomega.1c04531

C–H Bond Activation by the Excited Zinc Atom: Gas-Phase Formation of Methylzinc Hydride (HZnCH3) Based on Multireference Second-Order Perturbation Theory and Coupled Cluster Calculations

Jerzy Moc 1,*
PMCID: PMC8459409  PMID: 34568705

Abstract

graphic file with name ao1c04531_0007.jpg

The pioneering spectroscopic observations of the methylzinc hydride [HZnCH3(X1A1)] molecule were reported previously by the Ziurys group [J. Am. Chem. Soc.2010, 132, 17186–17192], and the possible formation mechanisms were suggested therein, including those with the participation of excited zinc atoms in reaction with methane. Herein, the ground singlet state and the lowest excited triplet state potential energy surfaces of the Zn + CH4 reaction have been explored using high-level electronic structure calculations with multireference second-order perturbation theory and coupled cluster singles and doubles with perturbative triples (CCSD(T)) methods in conjunction with all-electron basis sets (up to aug-cc-pV5Z) and scalar relativistic effects incorporated via the second-order Douglas–Kroll–Hess (DK) method. Based on the ab initio results, a plausible scenario for the formation of HZnCH3(X1A1) is proposed involving the activation of the C–H bond of methane by the lowest excited 3P state atomic zinc. Calculations also highlight the importance of an agostic-like Zn···H–C interactions in the pre-activation complex and good agreement between the structure of the HZnCH3(X1A1) molecule predicted at the DK-CCSD(T)/aug-cc-pVQZ-DK level of theory and that derived from rotational spectroscopy, as well as the discrepancies between the ab initio and density functional theory predictions.

1. Introduction

The activation of C–X (X = H, halogen) bonds in homogeneous systems is a research field relevant to important industrial processes.1 For X = H, the catalytic activation of C(sp3)–H bonds presents a much bigger challenge compared to that of C(sp2)–H bonds.2,3 In particular, splitting of the strong and nonreactive carbon–hydrogen bond in methane is thought to be the key step for methane conversion;4 in the present work, we use the “organometallic definition” of C–H bond activation as referring to “the formation of a carbon–metal bond by cleavage of a carbon–hydrogen bond”.2,3

In 2010, a methylzinc hydride molecule, HZnCH3, was synthesized in the gas phase in a DC discharge by the reaction of zinc vapor with methane in the presence of Ar gas and identified by using millimeter/submillimeter direct-absorption and Fourier-transform microwave spectroscopic techniques as described by Ziurys and co-workers.5 From the rotational constants (B) of the seven HZnCH3 isotopologues, an ro structure of the HZnCH3 molecule in the ground X1A1 electronic state was derived.5 In addition to the gas-phase reaction of atomic zinc with methane,5 the Zn/CH4 system was the subject of the low-temperature matrix isolation infrared (IR)6,7 and vacuum ultraviolet spectroscopy8 investigations. These investigations employed irradiation to accomplish the activation of the C–H bond in methane by zinc atoms.68 On the theoretical side, the Zn + CH4 reaction was studied9,10 using mostly density functional theory (DFT) and with an ab initio multireference configuration interaction (MRCI) plus the multireference second-order Møller–Plesset perturbation (MR/MP2) (MRCI- MR/MP2) method employing effective core potentials11 (we will refer to the results of the previous theoretical studies of Zn + CH4 when appropriate).

The electric discharge-induced reaction of zinc vapor with methane studied by Ziurys and co-workers5 is of importance in organometallic synthesis and serves as a model system for C–H bond activation. Although the possible formation mechanisms of the methylzinc hydride (HZnCH3(X1A1)) have been discussed by Ziurys and co-workers5 and other workers,10,11 including that of insertion of the zinc atom in the electronically excited (1P or 3P) state into the carbon–hydrogen bond of methane, the detailed mechanism by which CH4 is activated by atomic zinc in the lowest excited 3P state along with the subsequent formation of HZnCH3(X1A1) has not been sufficiently addressed. To provide more insight into this issue, we have embarked on a theoretical study to investigate the ground singlet state and the lowest excited triplet state potential energy surfaces of the Zn + CH4 reaction by using high-level single-reference and multireference ab initio methods (detailed in Section 4). We have also examined the sensitivity of the energetics of the Zn + CH4 system to the scalar relativistic effects.

2. Results and Discussion

2.1. Ground (1S) State and Excited (3P and 1P) States of Atomic Zinc

To prove the accuracy of the methods used, the calculated relative energies of the ground 1S(3d104s2) state and the excited 3P(3d104s14p1) and 1P(3d104s14p1) states of the Zn atom are compared with the experimental data12,13 in Tables 1 and S1. These tables show that the scalar relativistic effects, included using the second-order Douglas–Kroll–Hess (DK) Hamiltonian, cause the significant increase in the energy separation between the 1S and 3P states, thereby improving the accuracy of the estimate. For instance, the scalar relativistic effects are found to be 14.6, 18.4, and 18.0 kJ/mol with MCSCF(2,4)/aug-cc-pVTZ, MCQDPT2(2,4)/aug-cc-pVTZ, and CASPT2(2,4)/aug-cc-pVTZ, respectively, where MCSCF denotes multiconfigurational self-consistent field wave function, and MCQDPT2 and CASPT2 are two different implementations of the multireference second-order perturbation theory (MRPT2); the last three methods use the active space consisting of 2 electrons in 4 orbitals, (2,4), arising from Zn 4s4p orbitals. Our most accurate estimate of the relative energy for the 1S and 3P states of the Zn atom (Table 1), derived from the DK-(R)CCSD(T)/aug-cc-pV5Z-DK calculations of 389.9 kJ/mol [CCSD(T) signifies coupled-cluster theory with single, double, and perturbative triple excitations, with “R” indicating its spin-restricted variant used for the triplet], is consistent with earlier report14 and the experimental value of 391.2 kJ/mol.12,13Table 1 further shows that, at the DK-CASPT2(2,4)/aug-cc-pVTZ-DK level, the energy separation between the 1S and 1P states of 569.9 kJ/mol is somewhat overestimated compared to the experimental12,13 result of 559.4 kJ/mol. It is, however, the lowest excited 3P state of atomic zinc of primary importance from the point of view of the reaction mechanism considered in the present work.

Table 1. Relative Energies (kJ/mol) of the Ground 1S(3d104s2) State and the Excited 3P(3d104s14p1) and 1P(3d104s14p1) States of the Zn Atom Calculated Using Single-Referencea and Multireference Methods.

method 1S(3d104s2) 3P(3d104s14p1) 1P(3d104s14p1)
Non-Relativistic
MCQDPT2(2,4)b/aug-cc-pVTZ 0.0 366.1 560.2
CASPT2(2,4)b/aug-cc-pVTZ 0.0 364.0 552.7
CCSD(T)/aug-cc-pV5Z 0.0 371.5  
With Scalar Relativistic Effects via Second-Order DK
DK-MCQDPT2(2,4)b/aug-cc-pVTZ-DK 0.0 384.5 578.2
DK-CASPT2(2,4)b/aug-cc-pVTZ-DK 0.0 382.0 569.9
DK-CCSD(T)/aug-cc-pV5Z-DK 0.0 389.9  
exp.c 0.0 391.2d 559.4
a

For Zn(3P), this implies RCCSD(T).

b

Active space used in the MCQDPT2 and CASPT2 calculations was 2 electrons in 4 orbitals, (2,4), arising from Zn 4s4p orbitals.

c

Reference (12).

d

The spin–orbit average excitation energy (derived from Moore’s tables12).

Before proceeding with the Zn + CH4 system, we note that the remainder of this paper is organized as follows. First, we report on a pathway of the reaction of ground-state atomic zinc with methane; this includes a comparison of the equilibrium structure and vibrational frequencies of the insertion product with the available experimental data. After that, we describe a pathway for the reaction of excited 3P state atomic zinc with methane. A plausible mechanism for the formation of HZnCH3(X1A1) is next presented, followed by conclusions. All the relative energies quoted below are zero-point vibrational energy (ZPVE)-corrected, except for those given in Table 6.

Table 6. MRPT2 Relative Energiesa,b of the Species Relevant to the Formation and Dissociation of HZnCH3(X1A1) 3 (in kJ/mol).

species MCQDPT2(10,12)/aug-cc-pVTZ CASPT2(10,12)/aug-cc-pVTZ DK-CASPT2(10,12)/aug-cc-pVTZ-DK
Zn(3P) + CH4 (1) 366.1 (0.0) 364.0 (0.0) 382.0 (0.0)
TST(3A′) 395.4 (29.3) 397.9 (33.9) 408.8 (26.4)
HZnCH3(3A′) (3T) 356.5 (−9.6) 355.2 (−8.8) 361.5 (−20.5)
MEX 356.5 (−9.6) 356.1 (−7.9) 362.3 (−19.7)
ZnH(2Σ+) + CH3(2A2) 370.7 (4.6) 363.2 (−0.8) 369.0 (−13.0)
ZnCH3(2A1) + H(2S) 380.7 (14.6) 381.6 (17.6) 387.0 (5.0)
HZnCH3 (X1A1) (3) 46.9 (−319.2) 46.4 (−317.6) 50.6 (−331.4)
a

With respect to Zn(1S) + CH4 (1), the energies indicated in parentheses are relative to Zn(3P) + CH4 (1). Notice that these energies do not include corrections for ZPVE and spin–orbit effects (because MEX is not a stationary point on the 3N-6-dimensional potential energy surface, the frequency calculations could not be performed in this case). At the three consecutive MRPT2 levels indicated, the energies of TS (assuming the CCSD(T)/aug-cc-pVTZ geometry) are 355.2 (−10.9), 354.8 (−9.2), and 359.8 (−22.2) kJ/mol, respectively.

b

At the geometries optimized at the MCSCF(10,12)/aug-cc-pVTZ level.

2.2. van der Waals Complex Formation and Insertion Step for the Reaction of Ground-State Atomic Zinc with Methane

The reaction of ground-state atomic zinc with methane 1 (Figure 1a) entails formation of a van der Waals (vdW) complex Zn···η3-H3CH 2 of C3v symmetry (Figure 1b) whose binding energy (1.5 kJ/mol, Table 2) is found to be comparable to that of the analogous complex Cd···H3CH.15 Another initial vdW complex predicted here with CCSD(T), Zn···η1-HCH32a of C3v symmetry (Figure S1 of Supporting Information), has a somewhat lower binding energy than 2 (1.2 kJ/mol). No vdW complexes were reported in previous DFT studies of the Zn/CH4 system,9,10 which can be partially attributed to the well-known16 inability of conventional Kohn–Sham (KS) DFT to properly describe the dispersion forces. Complex 2 can, in principle, rearrange through insertion of the zinc atom into the C–H bond, which formally leads to the formation of the methylzinc hydride (HZnCH3(X1A1)) 3 species (Figure 1c). Before discussing the associated potential energy profile, we look more closely at the transition state involved.

Figure 1.

Figure 1

Structures of (a) reactant CH4 (1), (b) vdW complex Zn···η3-H3CH (2), and (c) insertion product HZnCH3 (X1A1) (3) of the reaction of ground-state atomic zinc with methane optimized at the CCSD(T)/aug-cc-pVQZ level. For 3, structural parameters shown in parentheses are from the DK-CCSD(T)/aug-cc-pVQZ-DK geometry optimization. Bold font indicates the experimental bond distances from ref (27) (CH41) and ref (5) (HZnCH3(X1A1) 3). Distances are in Å and angles are in degrees.

Table 2. Relative Energiesa of the Stationary Points of the Reaction of Ground-State Atomic Zinc with Methane Calculated at the ZPVE-Corrected CCSD(T)/aug-cc-pV5Z and DK-CCSD(T)/aug-cc-pV5Z-DK Levels.

species aug-cc-pV5Zb aug-cc-pV5Z-DKb,c
Zn(1S) + CH4 (1) 0.0 0.0
Zn···η3-H3CH (2) –1.5 –1.5
Zn···η1-HCH3 (2a)d –1.2 –1.2
TS 335.1 341.0
HZnCH3 (X1A1) (3) 45.2 47.7
a

Relative to Zn(1S) + CH4 (1) (in kJ/mol).

b

At the geometries optimized at the CCSD(T)/aug-cc-pVQZ level and including the CCSD(T)/aug-cc-pVTZ ZPVE contribution.

c

Computed with the DK-CCSD(T) method.

d

The optimized geometry of vdW complex 2a is shown in Figure S1.

Nowadays, stationary points of the reaction of activation of methane by the small metal clusters are usually located using KS DFT, with the energetics possibly refined by CCSD(T).17 It is therefore interesting to compare the structures of the corresponding transition state (TS) of the reaction of ground-state atomic zinc with methane predicted using KS DFT with B97-118 and B3LYP19,20 exchange–correlation functionals21 to that found using the coupled-cluster22,23 methods, an affordable task for the 3d-metal-containing-six-atom reaction system. These results show that Cs-symmetric TS predicted by KS DFT (including B97-1-DK) and CCSD represents a “typical” transition state for oxidative insertion9 (Figure 2a, the C–H bond is lengthened to 2.12–2.23 Å, thus being essentially broken), with the corresponding imaginary frequency values ranging from 791i to 925i cm–1, and consistent with the transition state reported from the relativistic DFT ZORA-BLYP/TZ2P calculations.9 In contrast, the Cs-symmetric TS found using CCSD(T) and DK-CCSD(T) (see Figure 2b), characterized by one imaginary frequency of 220i cm–1 (at the CCSD(T)/aug-cc-pVTZ level), is strongly asynchronous, featuring essentially a new Zn–H bond and a long Zn–C distance between ZnH and CH3 moieties of about 3.8 Å. The singlet CCSD(T) TS (Figure 2b) exhibits a significant multireference character24 and thus cannot be described correctly with the single-reference methods. We have not pursued this structure at the MR level because in the mechanism of HZnCH3(X1A1) formation considered here, the C–H activation step occurs through the relevant triplet transition state (as described in Section 2.4).

Figure 2.

Figure 2

Comparison of TS structures for the reaction of ground-state atomic zinc with methane predicted with (a) B3LYP, B97-1, B97-1-DK KS DFT, and coupled-cluster CCSD using the aug-cc-pVTZ basis set and (b) coupled-cluster CCSD(T)/aug-cc-pVQZ method; structural parameters given in parentheses are from the DK-CCSD(T)/aug-cc-pVQZ-DK geometry optimization. Distances are in Å and angles are in degrees. Transition vectors corresponding to the imaginary frequency of the KS DFT TS (a) and CCSD(T) TS (b) structures are shown.

The ground-state singlet potential energy profile of the Zn + CH4 reaction calculated at the DK-CCSD(T)/aug-cc-pV5Z-DK//CCSD(T)/aug-cc-pVQZ level of theory (Table 2) indicates that (1) the process is endothermic by 47.7 kJ/mol, that is, by about 40 kJ/mol less than previously computed,9,11 and (2) it has an intractable energy barrier (of 341 kJ/mol), consistent with earlier reports;9,11 the TS barrier of similar magnitude is found here using MRPT2 (see below). The theoretical results are compatible with the experimental studies of the Zn + CH4 reaction57 wherein an external stimulus was always required to make the reaction happen, possibly in order to first induce the (4s14p1)3P ← (4s2)1S electronic transition of zinc atoms to the reactive5,7,25,26 triplet state.

In Section 2.4, we will focus on the gas-phase activation of methane by excited 3P state atomic zinc. The spin–orbit coupling (SOC) (treated in Section 2.5) splits the 3P state of atomic zinc to the Zn(3PJ) states (J = 0, 1, and 2), and it was the excitation of the ground-state zinc to the 3P1 state that caused insertion into the C–H bond of methane.7 The spin–orbit interaction results in a decrease in the energy of the Zn atom by 2.4 kJ/mol; the latter is the difference between the spin–orbit average value and the Zn 3P1 state.

2.3. Structure and Vibrational Frequencies of HZnCH3(X1A1)

Based on Figure 1c, the zinc insertion product HZnCH3(X1A1) 3 features a linear H–Zn–C backbone (C3v symmetry), in agreement with the structure determined using rotational spectroscopy5 and quantum mechanical methods811 and consistent with the IR spectra of the matrix-isolated methylzinc hydride.7 It is also seen from this figure that there is good agreement between the equilibrium structure (re) of 3 optimized at the DK-CCSD(T)/aug-cc-pVQZ-DK level (values in parentheses) and the ro structure5 derived from measurements of the rotational spectra of the isotopologues of HZnCH3 (values in bold), in particular for the Zn–C and Zn–H bond lengths. For the C–H bond lengths, the agreement between the re(C–H) (1.093 Å) and ro(C–H) (1.140 Å)5 values is less satisfactory; in fact, the former C–H bond distances are similar to those reported at the DFT level.810

Finally, in Table 3, the calculated harmonic (ωi) and anharmonic (νi) vibrational frequencies of 3 are listed along with the available experimental data7 and previous10 harmonic DFT B3PW91 results. As this table indicates, the scalar relativistic effects cause an increase in the CCSD(T)/aug-cc-pVQZ harmonic frequencies of ZnH and ZnC stretch modes by 28 and 10 cm–1, respectively, due to an associated decrease in the two bond lengths (cf. Figure 1c). When the anharmonic contributions are considered by using second-order perturbation theory28 (see the values in the column of Table 3 under the heading “Hybrid”), the absolute deviation from the experiment is reduced to 18 cm–1 for the CH3 s-stretch (“s” stands for symmetric) and to 2–21 cm–1 for the next four modes with the lower frequencies. As a result, this leads to a better accordance with the experimental data7 compared to the previous DFT B3PW91 harmonic frequencies10 (the latter are given in the column of Table 3 under the heading “Literature”). A notable exception is the lowest frequency mode of 3, CZnH a-deform (“a” stands for asymmetric) for which the largest difference (32 cm–1) between the “hybrid” and the experimental (harmonic DFT) values is observed.

Table 3. Harmonic (ωi) and Anharmonic (νi) Vibrational Frequencies Calculated for the HZnCH3(X1A1) 3 Molecule along with the Experimental and Harmonic Literature Values (in cm–1).

    CCSD(T)/aug-cc-pVQZ DK-CCSD(T)/aug-cc-pVQZ-DK “Hybrid”a literatureb  
description of mode sym. of vib. ωi ωi νi ωi exp.c
CH3 a-stretch e 3103 3101 2960 (8) 3117 d
CH3 s-stretch a1 3022 3016 2902 (9) 3037 2919.8
ZnH stretch a1 1926 1954 1888 (212) 1901 1866.1
CH3 a-deform e 1466 1453 1436 (0) 1452 d
CH3 s-deform a1 1213 1213 1181 (3) 1203 1179.3
CH3 rock e 700 684 666 (74) 712 686.8, 689.4
ZnC stretch a1 568 577 569 (14) 562 566.5
CZnH a-deform e 424 414 410 (46) 442 442.6
a

Obtained in this work by combining the DK-CCSD(T)/aug-cc-pVQZ-DK harmonic vibrational frequencies with the anharmonic28 frequency corrections (based on the MP2/aug-cc-pVQZ calculations29); values in parentheses are the harmonic IR intensities (km/mol) obtained from the MP2/aug-cc-pVQZ calculations.

b

Values in brackets are the DFT B3PW91 harmonic vibrational frequencies taken from ref (10).

c

Data taken from the low-temperature matrix isolation IR study of the Zn/CH4 system reported in ref (7).

d

Not observed in the matrix IR spectra.7

2.4. Activation of Methane by Excited 3P State Atomic Zinc

The profile of the lowest triplet state potential energy surface of the Zn(3P) + CH4 reaction calculated at the DK-(R)CCSD(T)/aug-cc-pV5Z-DK//(R)CCSD(T)/aug-cc-pVQZ level of theory is shown in Table 4, with structures of the relevant species displayed in Figure 3. As observed above for the ground-state atomic zinc/methane reaction system, the reaction between Zn(3P) and methane involves formation of the initial vdW complex, Zn···η1-HCH3(3A′) 2T (Figure 3a, “T” stands for “Triplet”). In contrast, however, to the singlet counterpart 2 of C3v symmetry, complex 2T (Figure 3a) adopts a Cs symmetry and exhibits a short Zn···H–C contact with the distance of 2.244 Å, indicating the occurrence of a Zn···H–C interaction similar to the agostic one.30 This interaction can be viewed as a pre-activation step as it is manifested by both the elongated C–H bond distance of 1.103 Å [the CCSD(T)/aug-cc-pVQZ (experimental27) bond length of methane is 1.088 (1.092 Å)] and the increase of the predicted binding energy of 2T in comparison to the singlet analogue 2, 7.6 kJ/mol versus 1.5 kJ/mol, respectively, relative to the respective reactants [the relative energies of the triplet species quoted in this section are determined with respect to the Zn(3P1) + CH4 asymptote; cf. the values indicated in brackets in Table 4].

Table 4. Relative Energiesa of the Stationary Points of the Reaction of Excited 3P State Atomic Zinc with Methane Calculated at the ZPVE-Corrected (R)CCSD(T)/aug-cc-pV5Z and DK-(R)CCSD(T)/aug-cc-pV5Z-DK Levels.b.

speciesc aug-cc-pV5Zd aug-cc-pV5Z-DKd,e,f
Zn(3P) + CH4 (1) 371.5 (0.0) 389.9 (0.0) [0.0]
Zn···η1-HCH3 (2T) 362.3 (−9.4) 379.9 (-10.0) [−7.6]
TST(3A′) 397.5 (25.9) 409.2 (19.2) [21.6]
HZnCH3(3A′) (3T) 340.2 (−31.4) 346.0 (−43.5) [−41.1]
TSdissT(3A′) 369.4 (−2.1) 374.9 (−14.6) [-12.2]
ZnH(2Σ+) + CH3(2A2) 343.5 (−28.0) 349.4 (−40.6) [−38.2]
ZnCH3(2A1) + H(2S) 366.9 (−4.6) 372.8 (−17.2) [−14.8]
a

Relative to Zn(1S) + CH4 (1) except for the energies indicated in parentheses which are relative to Zn(3P) + CH4 (1) (in kJ/mol).

b

For the open-shell species, this refers to the corresponding RCCSD(T) levels.

c

For the optimized geometries of the ZnCH3(2A1), ZnH(2Σ+), and CH3(2A2) radicals, see Figure S5.

d

At the geometries optimized at the CCSD(T)/aug-cc-pVQZ level and including the CCSD(T)/aug-cc-pVTZ ZPVE contribution.

e

Computed with the DK-CCSD(T) method.

f

The energies (in kJ/mol) indicated in brackets are also corrected for spin–orbit effects.

Figure 3.

Figure 3

(a) “pre-activation” vdW complex Zn···η1-HCH3 (2T), (b) transition state for the C–H bond activation (TST), (c) resulting intermediate HZnCH3 (3T), and (d) transition state for the H-atom dissociation (TSdissT) from 3T, located on the lowest triplet state potential energy surface of the reaction of Zn(3P) with methane using the RCCSD(T)/aug-cc-pVQZ method. For 3T, structural parameters shown in parentheses are from the DK-RCCSD(T)/aug-cc-pVQZ-DK geometry optimization. Due to the convergence problems in the numerical calculation of the nuclear hessian of 2T at the RCCSD(T)/aug-cc-pVnZ (n = D,T) levels, the actual hessian of 2T was computed analytically with the UMP2/aug-cc-pVTZ29 method (and found to be positive-definite). Distances are in Å and angles are in degrees.

On the triplet reaction potential surface, the C–H bond activation is found to be exothermic by 41.1 kJ/mol with respect to Zn(3P1) + CH4 (Table 4) and proceeds from 2T to the intermediate 3T(3A′) (Figure 3c) via the “early” transition state TST(3A′) (Figure 3b), the latter characterized by one imaginary frequency of 1332i cm–1 (at the RCCSD(T)/aug-cc-pVTZ level). Note that in the TST3T step, the CH3 group rotates about the Zn–C axis (as confirmed by the IRC following using the UMP2/aug-cc-pVTZ method29). Interestingly, the corresponding energy barrier of 21.6 kJ/mol [with respect to Zn(3P1) + CH4], derived from the DK-(R)CCSD(T)/aug-cc-pV5Z-DK calculations (Table 4), shows good agreement with the experimental activation energy of the Zn(3P1) + CH4 process (∼14.7 kJ/mol) inferred from the matrix-isolation IR study of the radiation-induced reaction of zinc atoms with methane of Downs and co-workers7 (the estimate referred to therein as an “approximate upper limit”7). The latter two barrier evaluations are at variance with that of a previous computational study of the Zn(3P) + CH4 reaction by Castillo et al.11 who reported a much higher energy barrier (75.3 kJ/mol).

Compared to the singlet 3, the triplet 3T differs vastly in the equilibrium geometry and the depth of the corresponding potential well, congruent with the DFT study of Alikhani.10 Namely, 3T (Figure 3c) features a bent H–Zn–C backbone with the acute H–Zn–C bond angle of 78.0° and the Zn–C distance being longer than that of 3 by as much as 0.75 Å (at the DK-RCCSD(T)/aug-cc-pVQZ-DK level). Consequently, 3T is predicted to be only marginally stable (by 2.9 kJ/mol) with respect to dissociation into ZnH(2Σ+) + CH3(2A2) (Table 4), and, therefore, it can be represented as HZn···CH3(3A′); there is no energy barrier above the dissociation energy, consistent with an earlier report.11 By examining a higher energy dissociation channel of 3T into ZnCH3(2A1) + H(2S), we have found that it takes place via the TSdiss(3A′) transition state (Figure 3d) that lies 28.9 kJ/mol above 3T, giving rise to the energy barrier of 2.6 kJ/mol beyond the dissociation energy (Table 4).

2.5. Plausible Mechanism for the Formation of HZnCH3(X1A1) in the Gas Phase Involving Zn(3P) and CH4. Is This a Viable Reaction Pathway?

Under the experimental conditions used in the gas-phase study,5 the reactant mixture containing zinc vapor (Zn(g)) and methane was exposed to a DC discharge, with the latter described5 as “not a selective excitation method”, meaning that the Zn (4s14p1)3P ← (4s2)1S and (4s14p1)1P ← (4s2)1S electronic transitions could have occurred. In this section, a plausible scenario for the formation of HZnCH3(X1A1) 3 is considered that involves activation of methane by excited 3P state atomic zinc [note that the activation of CH4 by excited Zn(1P) was the subject of a previous theoretical study11].

We looked first at the HZnCH3(X1A1) formation pathway that involves intersystem crossing. Using both coupled-perturbed MCSCF(10,12)/def2-SVP31 (implemented in MOLPRO32) and MCSCF(10,12)/aug-cc-pVTZ combined with the constrained optimization method33 (the latter implemented in GAMESS34), the minimum energy crossing point (MEX) between the lowest triplet and singlet electronic states has been located in the vicinity of the triplet intermediate 3T (Figure 4b). The qualitatively similar type of MEX has been found at the B3LYP/aug-cc-pVTZ level using GAMESS33,34 (Figure 4a), consistent with a previous DFT10 report. The major difference between the two MEX structures is the significantly longer Zn–C distance predicted with MCSCF compared to that found with B3LYP. This is due to insufficient dynamic correlation in the MCSCF method, which has implications for the location of the system’s MEX(s).35 The rate of intersystem crossing is known to depend on the SOC.36 To compute the SOC matrix element at the MEX, we first applied the approximate one-electron Breit–Pauli spin–orbit operator combined with the effective nuclear charge approach37 along with state-averaged MCSCF wave functions employing the relativistic Stevens–Basch–Krauss–Jasien–Cundari (SBKJC) effective core potentials and associated basis sets (MCSCF/SBKJC ECP).37 In addition, the Breit–Pauli spin–orbit operator including the one-electron and two-electron terms was used in the interacting-states approach38 with state-averaged MCSCF wave functions and for the purpose of testing, with MRCI39,40 wave functions. The performance of these various SOC computing procedures was evaluated through calculation of the spin–orbit splittings in the 3P state of atomic zinc (Table 5), which shows that the MCSCF(2,4)/SBKJC ECP and DK-MCSCF(2,4)/aug-cc-pVTZ-DK (cc-pVTZ-DK) results are in reasonable agreement with those obtained by the DK-MRCI(2,4)/aug-cc-pwCVTZ-DK (cc-pwCVTZ-DK) approach and experiment.12 The magnitude of the calculated SOC constant36 at the MEX is strongly dependent on the Zn–C distance. That is, at the MCSCF geometry, the SOC constant is found to be in the range of 1.05–1.83 cm–1 with the MCSCF(2,4)/SBKJC ECP and DK-MCSCF(2,4)/aug-cc-pVTZ-DK procedures and at the B3LYP geometry, it is predicted to be 28.56 cm–1 (MCSCF(2,4)/SBKJC ECP) and 38.73 cm–1 (DK-MCSCF(2,4)/aug-cc-pVTZ-DK). Recall that MEX actually constitutes the HZn···CH3 radical pair and for the radical pairs, SOC is known to decrease sharply as the separation of the radical centers increases.41

Figure 4.

Figure 4

MEX between the lowest singlet (1A′) and the lowest triplet (3A′) state potential energy surfaces of the Zn + CH4 reaction determined with (a) B3LYP using the aug-cc-pVTZ basis set and (b) MCSCF(10,12) using the def2-SVP and aug-cc-pVTZ basis sets; the structural parameters obtained with the latter basis set are given in brackets. The B3LYP/aug-cc-pVTZ and MCSCF/aug-cc-pVTZ MEX structures have been located with GAMESS, and the MCSCF/def2-SVP one has been found with MOLPRO. Distances are in Å and angles are in degrees.

Table 5. Spin–Orbit Splitting for Zn(3P) (in cm–1).

method |3P03P1| |3P13P2|
MCSCF(2,4)/SBKJC ECPa 180.20 360.43
DK-MCSCF(2,4)/aug-cc-pVTZ-DKb,c 165.93 (167.13) 331.87 (334.27)
DK-MRCI(2,4)/aug-cc-pVTZ-DKc,d 170.98 (171.22) 341.95 (342.45)
DK-MRCI(2,4)/aug-cc-pwCVTZ-DKe,f 188.01 (183.76) 376.01 (367.53)
exp.g 190.08 388.93
a

Computed with the state-averaged MCSCF method using SBKJC ECPs and an approximate one-electron Breit–Pauli spin–orbit operator combined with the effective nuclear charge (Zeff) approach; see ref (37) for the relevant references.

b

Computed with the state-averaged MCSCF method and Breit–Pauli spin–orbit operator (including the one-electron and two-electron terms) in the interacting-states approach.38

c

Values in parentheses are obtained with the cc-pVTZ-DK basis set.

d

Only valence electrons correlated in the MRCI calculations (see also footnoteb).

e

All electrons correlated in the MRCI calculations (see also footnoteb).

f

Values in parentheses are obtained with the cc-pwCVTZ-DK basis set.

g

Taken from Moore’s tables.12

Table 6 shows the MRPT2 energies of the species relevant to the formation and dissociation of HZnCH3(X1A1) 3. As can be seen, there is an overall good agreement between the predictions of the two MRPT2 methods. In addition, significant scalar relativistic effects are found by comparing the CASPT2/aug-cc-pVTZ and DK-CASPT2/aug-cc-pVTZ-DK energies (ranging from 4.6 to 18 kJ/mol), in line with the coupled-cluster calculations.

Next, we considered a somewhat different mechanistic scenario of the formation of HZnCH3(X1A1) 3 involving the Zn(3P) and CH4 reactants, viewed to be a more plausible mechanism42 in the gas phase. In addition to the Zn 3P ← 1S electronic excitation and activation of methane by excited zinc through the TST transition state to yield 3T, this scenario allows the dissociation of the triplet intermediate HZn···CH3(3A′) 3T into ZnH(2Σ+) and CH3(2A2) and involves the recombination of the resulting radicals followed by relaxation to 3 (Figure 5). This scenario corresponds basically to a radical rebound mechanism43 that does not involve MEX. The presence of a “third body” Ar in the reaction chamber5 could have provided a stabilization of the HZnCH3(X1A1) 3 product (as it was also mentioned by Ziurys and co-workers5). Although some experimental evidence in favor of a radical mechanism for the formation of HZnCH3(X1A1) has been provided because “weak ZnH signals” were detected,5 a reaction dynamics investigation is needed to further support this mechanism.43

Figure 5.

Figure 5

Relative energies of the key stationary points on the lowest singlet and triplet potential surfaces of the Zn + CH4 reaction relevant to the formation of HZnCH3(X1A1) 3 as calculated at the DK-CASPT2(10,12)/aug-cc-pVTZ-DK//MCSCF(10,12)/aug-cc-pVTZ level (note that these energies do not include corrections for ZPVE and spin–orbit effects). Singlet (S) potential surface, navy blue; triplet (T) potential surface, pink; and MEX is denoted by the black dot. The values in parentheses are in kJ/mol.

Because, as disclosed above, both excited 3P and 1P state atomic zinc could have been produced under the experimental conditions employed in the gas-phase study,5 the zinc atom in the 1P excited state could have contributed to the C–H bond activation as well.42 That was spectroscopically observed in low-temperature matrix studies of the reaction of Zn(1P) with methane.8

3. Conclusions

We have probed potential energy surfaces of the reaction of ground 1S and excited 3P state atomic zinc with methane [using the CCSD(T) method] along with intersystem crossing (studied with MCSCF/MRPT2) and the importance of the scalar relativistic effects (included via the second-order DK method) on the reaction’s energetics and the structure of the methylzinc hydride (HZnCH3(X1A1)) product molecule. The latter molecule was previously5 synthesized in the gas phase in a DC discharge by the reaction of zinc vapor with methane and observed using rotational spectroscopy. The main results are as follows.

  • (1)

    The reaction of ground-state atomic zinc and methane to afford the methylzinc hydride (HZnCH3(X1A1)) is predicted to be endothermic relative to the separated reactants (by 48 kJ/mol) and requires overcoming an intractable energy barrier (of 340 kJ/mol), these results being compatible with the experimental studies57 and earlier calculations.9,11 There is good agreement between the equilibrium geometry (re) of the HZnCH3(X1A1) molecule predicted with the DK-CCSD(T)/aug-cc-pVQZ-DK method and the corresponding ro structure derived5 from rotational spectroscopy, especially for the Zn–C and Zn–H bond lengths.

  • (2)

    The reaction of excited 3P state atomic zinc with methane entails the formation of a “pre-activation” vdW complex that exhibits agostic-like Zn···H–C interaction. Involvement of Zn(3P) in the gas-phase reaction with methane has led to a dramatic decrease in the activation barrier compared to that found for ground-state atomic zinc. Moreover, the magnitude of the energy barrier of 21.6 kJ/mol as determined at the ZPVE and spin–orbit-corrected DK-(R)CCSD(T)/aug-cc-pV5Z-DK//(R)CCSD(T)/aug-cc-pVQZ level of theory is supported by the experimental activation energy of the Zn(3P1) + CH4 reaction (∼14.7 kJ/mol) inferred from the matrix-isolation IR study of the radiation-induced reaction of zinc atoms with methane.7 Significant scalar relativistic effects on the exothermicity of the overall Zn(3P) + CH4 → HZnCH3 (3A′) process and related energy barrier are found, that is, an increase by 12.1 kJ/mol and a decrease by 6.7 kJ/mol, respectively.

  • (3)

    A plausible mechanism for the formation of HZnCH3(X1A1) in the gas phase5 is suggested, which involves activation of the C–H bond of methane by the lowest excited 3P state atomic zinc to give the loosely bound intermediate HZn···CH3(3A′), which dissociates into ZnH and CH3, followed by the recombination of the resulting radicals. A reaction dynamics investigation is necessary to provide further support of the mechanism proposed.

4. Computational Methods

Potential energy surfaces of the reaction of atomic zinc with methane were probed using CCSD(T)23 in conjunction with the aug-cc-pVnZ (n = D,T,Q) basis sets.4447 For the open-shell species, the spin-restricted CCSD(T) method,48 denoted as RCCSD(T), was employed. The relevant stationary points were confirmed by harmonic vibrational frequency analyses carried out at the (R)CCSD(T)/aug-cc-pVnZ levels with n = D,T. Single-point (R)CCSD(T)/aug-cc-pV5Z calculations were also reported. The scalar relativistic effects were considered using the second-order Douglas–Kroll–Hess (DK) Hamiltonian49,50 in combination with the aug-cc-pVnZ-DK (n = T,Q,5) basis sets.4447,51 In all of the coupled-cluster calculations, Zn 1s2s2p3s3p and C 1s atomic orbitals were kept frozen (not correlated). Only the (R)CCSD(T)/aug-cc-pVQZ (and DK-(R)CCSD(T)/aug-cc-pVQZ-DK) structures have been presented in the main text (for a comparison of the (R)CCSD(T)/aug-cc-pVnZ structures with n = D,T,Q, see Figures S2, S8 and S9 of the Supporting Information).

The multireference treatment was accomplished by MCSCF wave function calculations of the complete active space (CAS)52,53 or fully optimized reaction space54 type, with the active space consisting of 10 electrons in 12 orbitals, (10,12), arising from Zn 4s4p, C 2s2p, and H 1s orbitals, followed by calculations using second-order multiconfiguration quasi-degenerate perturbation theory (MCQDPT2)55 and CASPT256 implementation of MRPT2 to account for dynamic correlation effects (when applied to one state only, MCQDPT2 is equivalent to multireference second-order Møller–Plesset perturbation theory, MRMP257,58).

MOLPRO2012.132 quantum chemistry code was used for the CC and CASPT2 computations and GAMESS34 code was exploited for the MCQDPT2 computations. Both programs were employed for the spin–orbit computations. The MP2 (UMP2) and KS DFT calculations were performed with the GAUSSIAN1659 program (more computational details are provided in the Supporting Information).

Acknowledgments

The author is thankful to the reviewers for their constructive criticisms and suggestions, which helped in improving this manuscript. This work is supported by the Wroclaw Centre for Networking and Supercomputing, WCSS.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.1c04531.

  • Full citation for refs (32, 34, 59), additional vdW complex Zn···η1-HCH32a formed due to interaction between ground-state atomic zinc and methane as optimized at the CCSD(T)/aug-cc-pVnZ (n = D,T,Q) levels; “full version” of Figure 3; third-order saddle point TOSPabstT located on the lowest triplet state potential energy surface of the reaction of Zn(3P) with methane using the RCCSD(T)/aug-cc-pVnZ (n = D,T) methods; structures of the triplet intermediate HZnCH3(3A′) 3T optimized at the MCSCF(10,12)/aug-cc-pVTZ, MCSCF(10,12)/def2-SVP, and RB3LYP/aug-cc-pVTZ levels; structures of the ZnCH3(2A1), CH3(2A2), and ZnH(2Σ+) radicals optimized at the RCCSD(T)/aug-cc-pVnZ (n = D,T,Q) levels; vdW complex Zn···η1-HCH32T located on the lowest triplet state potential energy surface of the reaction of Zn(3P) with methane using the UMP2/aug-cc-pVTZ method; structures of HZnCH3(X1A1) 3 and transition state TST(3A′) for the C–H bond activation optimized at the MCSCF(10,12)/aug-cc-pVTZ level; “full versions” of Figures 1 and 2 and of Tables 1, 2 and 4; and “Additional Computational Details” section (PDF)

The author declares no competing financial interest.

Supplementary Material

ao1c04531_si_001.pdf (894.7KB, pdf)

References

  1. Crespo M.; Martínez M.; Nabavizadeh S. M.; Rashidi M. Kinetico-Mechanistic Studies on C-X (X=H, F, Cl, Br, I) Bond Activation Reactions on Organoplatinum (II) Complexes. Coord. Chem. Rev. 2014, 279, 115–140. 10.1016/j.ccr.2014.06.010. [DOI] [Google Scholar]
  2. Labinger J. A.; Bercaw J. E. Understanding and Exploiting C-H Bond Activation. Nature 2002, 417, 507–514. 10.1038/417507a. [DOI] [PubMed] [Google Scholar]
  3. Jazzar R.; Hitce J.; Renaudat A.; Sofack-Kreutzer J.; Baudoin O. Functionalization of Organic Molecules by Transition-Metal-Catalyzed C(sp3)-H Activation. Chem.—Eur. J. 2010, 16, 2654–2672. 10.1002/chem.200902374. [DOI] [PubMed] [Google Scholar]
  4. Li J.; Zhou S.; Zhang J.; Schlangen M.; Weiske T.; Usharani D.; Shaik S.; Schwarz H. Electronic Origins of the Variable Efficiency of Room-Temperature Methane Activation by Homo- and Heteronuclear Cluster Oxide Cations [XYO2]+ (X,Y=Al, Si, Mg): Competition between Proton-Coupled Electron Transfer and Hydrogen-Atom Transfer. J. Am. Chem. Soc. 2016, 138, 7973–7981. 10.1021/jacs.6b03798. [DOI] [PubMed] [Google Scholar]
  5. Flory M. A.; Apponi A. J.; Zack L. N.; Ziurys L. M. Activation of Methane by Zinc: Gas-Phase Synthesis, Structure, and Bonding of HZnCH3. J. Am. Chem. Soc. 2010, 132, 17186–17192. 10.1021/ja106121v. [DOI] [PubMed] [Google Scholar]
  6. Billups W. E.; Konarski M. M.; Hauge R. H.; Margrave J. L. Activation of Methane with Photoexcited Metal Atoms. J. Am. Chem. Soc. 1980, 102, 7393–7394. 10.1021/ja00544a054. [DOI] [Google Scholar]
  7. Greene T. M.; Andrews L.; Downs A. J. The Reaction of Zinc, Cadmium, and Mercury Atoms with Methane: Infrared Spectra of the Matrix-Isolated Methylmetal Hydrides. J. Am. Chem. Soc. 1995, 117, 8180–8187. 10.1021/ja00136a015. [DOI] [Google Scholar]
  8. Henchy C.; Kilmartin U.; McCaffrey J. G. Vacuum Ultraviolet Spectroscopy and Photochemistry of Zinc Dihydride and Related Molecules in Low-Temperature Matrices. J. Phys. Chem. A 2013, 117, 9168–9178. 10.1021/jp404695a. [DOI] [PubMed] [Google Scholar]
  9. De Jong G. T.; Visser R.; Bickelhaupt F. M. Oxidative Addition to Main Group versus Transition Metals: Insights from the Activation Strain Model. J. Organomet. Chem. 2006, 691, 4341–4349. 10.1016/j.jorganchem.2006.03.006. [DOI] [Google Scholar]
  10. Alikhani M. E. Theoretical Study of the Insertion Reaction of Zinc, Cadmium, and Mercury Atoms with Methane and Silane. Chem. Phys. Lett. 1999, 313, 608–616. 10.1016/s0009-2614(99)01129-x. [DOI] [Google Scholar]
  11. Castillo S.; Ramírez-Solís A.; Díaz D.; Poulain E.; Novaro O. Theoretical Study of the Activation of Methane by Photoexcited Zinc Atoms. Mol. Phys. 1994, 81, 825–836. 10.1080/00268979400100551. [DOI] [Google Scholar]
  12. Moore C. E.Atomic Energy Levels; NSRDS-NBS Circular No 35; Office of Standard Reference Data; National Bureau of Standards: Washington, DC, 1971. [Google Scholar]
  13. National Institute of Standards and Technology . NIST Atomic Spectra Database, version 4: Gaithersburg, MD, 2013. http://www.nist.gov/pml/data/asd.cfm.
  14. Determan J. J.; Omary M. A.; Wilson A. K. Modeling the Photophysics of Zn and Cd Monomers, Metallophilic Dimers, and Covalent Excimers. J. Phys. Chem. A 2011, 115, 374–382. 10.1021/jp108384s. [DOI] [PubMed] [Google Scholar]
  15. Ramirez-Solis A.; Castillo S. C3v versus C2v Cd(1S,3P,1P)-CH4 van der Waals Complexes: A Variational and Perturbational Multireference Configuration Interaction Study. J. Chem. Phys. 1993, 98, 8065–8069. 10.1063/1.464562. [DOI] [Google Scholar]
  16. Krystyán S.; Pulay P. Can (Semi)Local Density Functional Theory Account for the London Dispersion Forces?. Chem. Phys. Lett. 1994, 229, 175–180. 10.1016/0009-2614(94)01027-7. [DOI] [Google Scholar]
  17. Schwarz H.; González-Navarrete P.; Li J.; Schlangen M.; Sun X.; Weiske T.; Zhou S. Unexpected Mechanistic Variants in the Thermal Gas-Phase Activation of Methane. Organometallics 2017, 36, 8–17. 10.1021/acs.organomet.6b00372. [DOI] [Google Scholar]
  18. Hamprecht F. A.; Cohen A. J.; Tozer D. J.; Handy N. C. Development and Assessment New Exchange-Correlation Functionals. J. Chem. Phys. 1998, 109, 6264–6271. 10.1063/1.477267. [DOI] [Google Scholar]
  19. Becke A. D. Density-Functional Exchange-Energy Approximation with Correct Asymptotic Behavior. Phys. Rev. A 1988, 38, 3098–3100. 10.1103/physreva.38.3098. [DOI] [PubMed] [Google Scholar]
  20. Lee C.; Yang W.; Parr R. G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B: Condens. Matter Mater. Phys. 1988, 37, 785–789. 10.1103/physrevb.37.785. [DOI] [PubMed] [Google Scholar]
  21. Both B97-1 and B3LYP exchange-correlation functionals were indicated recently to be among the “top twenty” best-performing functionals for the prediction of the bond dissociation energies of 3d transition metal diatomics, see:Xu X.; Zhang W.; Tang M.; Truhlar D. G. Do Practical Standard Coupled Cluster Calculations Agree Better than Kohn-Sham Calculations with Currently Available Functionals When Compared to the Best Available Experimental Data for Dissociation Energies of Bonds to 3d Transition Metals?. J. Chem. Theory Comput. 2015, 11, 2036–2052. 10.1021/acs.jctc.5b00081. [DOI] [PubMed] [Google Scholar]
  22. Purvis G. D.; Bartlett R. J. A Full Coupled-Cluster Singles and Doubles Model: the Inclusion of Disconnected Triples. J. Chem. Phys. 1982, 76, 1910–1918. 10.1063/1.443164. [DOI] [Google Scholar]
  23. Raghavachari K.; Trucks G. W.; Pople J. A.; Head-Gordon M. A Fifth-Order Perturbation Comparison of Electron Correlation Theories. Chem. Phys. Lett. 1989, 157, 479–483. 10.1016/s0009-2614(89)87395-6. [DOI] [Google Scholar]
  24. The dominant determinants of the MCSCF(10,12)/aug-cc-pVTZ wavefunction calculated at the CCSD(T)/aug-cc-pVTZ TS have CI coefficients of 0.78 and 0.58.
  25. Cartland H. E.; Pimentel G. C. Photosensitized Reaction of 3P Hg, Cd, and Zn Atoms with Difluorochloroethene in Krypton Matrix. J. Phys. Chem. 1986, 90, 1822–1827. 10.1021/j100400a017. [DOI] [Google Scholar]
  26. Breckenridge W. H. Activation of H-H, Si-H, and C-H Bonds by nsnp Excited States of Metal Atoms. J. Phys. Chem. 1996, 100, 14840–14855. 10.1021/jp960700j. [DOI] [Google Scholar]
  27. Hehre W. J.; Radom L.; Schleyer P. v. R.; Pople J. A.. Ab Initio Molecular Orbital Theory; John Wiley & Sons, Inc.: New York, 1986. [Google Scholar]
  28. Barone V. Anharmonic Vibrational Properties by a Fully Automated Second-Order Perturbative Approach. J. Chem. Phys. 2005, 122, 014108. 10.1063/1.1824881. [DOI] [PubMed] [Google Scholar]; ; this second-order perturbation theory method is based on quadratic, cubic and semidiagonal quartic force constants
  29. Møller C.; Plesset M. S. Note on an Approximation Treatment for Many-Electron Systems. Phys. Rev. 1934, 46, 618–622. 10.1103/PhysRev.46.618. [DOI] [Google Scholar]
  30. Lu Q.; Neese F.; Bistoni G. Formation of Agostic Structures Driven by London Dispersion. Angew. Chem., Int. Ed. 2018, 57, 4760–4764. 10.1002/anie.201801531. [DOI] [PubMed] [Google Scholar]
  31. Weigend F.; Ahlrichs R. Balanced Basis Sets of Split Valence, Triple Zeta Valence and Quadruple Zeta Valence Quality for H to Rn: Design and Assessment of Accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. 10.1039/b508541a. [DOI] [PubMed] [Google Scholar]
  32. Werner H.-J.; Knowles P. J.; Knizia G.; Manby F. R.; Schütz M.; Celani P.; Korona T.; Lindh R.; Mitrushenkov A.; Rauhut G.. et al. MOLPRO, version 2012.1; Cardiff University: Cardiff, U.K., 2012.
  33. Farazdel A.; Dupuis M. On the Determination of the Minimum on the Crossing Seam of Two Potential Energy Surfaces. J. Comput. Chem. 1991, 12, 276–282. 10.1002/jcc.540120219. [DOI] [Google Scholar]
  34. Schmidt M. W.; Baldridge K. K.; Boatz J. A.; Elbert S. T.; Gordon M. S.; Jensen J. H.; Koseki S.; Matsunaga N.; Nguyen K. A.; Su S.; et al. General Atomic and Molecular Electronic Structure System. J. Comput. Chem. 1993, 14, 1347–1363. 10.1002/jcc.540141112. [DOI] [Google Scholar]
  35. Harvey J. N.; Aschi M.; Schwarz H.; Koch W. The Singlet and Triplet States of Phenyl Cation. A Hybrid Approach for Locating Minimum Energy Crossing Points Between Non-Interacting Potential Energy Surfaces. Theor. Chem. Acc. 1998, 99, 95–99. 10.1007/s002140050309. [DOI] [Google Scholar]
  36. See, e.g.Moc J.; Fedorov D. G.; Gordon M. S. A Theoretical Study of the Reaction of Ti+ with Ethane. J. Chem. Phys. 2000, 112, 10247–10258. 10.1063/1.481666. [DOI] [Google Scholar]; Root-mean-square (rms) SOC constant defined36,41 as SOCrms = [∑Ms|SOC(MS)|2]1/2 is used, where SOC(MS) = <1Ψ|HSO^|3Ψ(MS) > and HSO^ is the Breit–Pauli spin–orbit Hamiltonian
  37. Koseki S.; Matsunaga N.; Asada T.; Schmidt M. W.; Gordon M. S. Spin-Orbit Coupling Constants in Atoms and Ions of Transition Elements: Comparison of Effective Core Potentials, Model Core Potentials, and All-Electron Methods. J. Phys. Chem. A 2019, 123, 2325–2339. 10.1021/acs.jpca.8b09218. [DOI] [PubMed] [Google Scholar]; , and references cited therein; especially, the references on the SBKJC ECPs are 73–76 in ref (37), and the references of these authors’ previous papers on the determination of the effective nuclear charges (Zeff) are 54–58, and 63–64 in ref (37)
  38. Berning A.; Schweizer M.; Werner H.-J.; Knowles P. J.; Palmieri P. Spin-Orbit Matrix Elements for Internally Contracted Multireference Configuration Interaction Wavefunctions. Mol. Phys. 2000, 98, 1823–1833. 10.1080/00268970009483386. [DOI] [Google Scholar]
  39. Werner H. J.; Knowles P. J. An Efficient Internally Contracted Multiconfiguration-Reference Configuration Interaction Method. J. Chem. Phys. 1988, 89, 5803–5814. 10.1063/1.455556. [DOI] [Google Scholar]
  40. Knowles P. J.; Werner H.-J. An Efficient Method for the Evaluation of Coupling Coefficients in Configuration Interaction Calculations. Chem. Phys. Lett. 1988, 145, 514–522. 10.1016/0009-2614(88)87412-8. [DOI] [Google Scholar]
  41. Carlacci L.; Doubleday C. Jr.; Furlani T. R.; King H. F.; McIver J. W. Jr. Spin-Orbit Coupling in Biradicals. Ab Initio MCSCF Calculations on Trimethylene and the Methyl-Methyl Radical Pair. J. Am. Chem. Soc. 1987, 109, 5323–5329. 10.1021/ja00252a004. [DOI] [Google Scholar]
  42. As Reviewer1 remarked, “the lifetime of the triplet intermediate 3T is too short to be collisionally stabilized.” This reviewer also anticipated that, compared to Zn(3P), “reactions of Zn(1P) might still be the more active species in the formation of HZnCH3”.
  43. That was Reviewer2 who first called this mechanism this way. This reviewer has also suggested that a reaction dynamics investigation is needed to further support the mechanism proposed herein.
  44. Dunning T. H. Gaussian Basis Sets for Use in Correlated Molecular Calculations 0.1. The Atoms Boron through Neon and Hydrogen. J. Chem. Phys. 1989, 90, 1007–1023. 10.1063/1.456153. [DOI] [Google Scholar]
  45. Kendall R. A.; Dunning T. H.; Harrison R. J. Electron Affinities of the First-Row Atoms Revisited. Systematic Basis Sets and Wave Functions. J. Chem. Phys. 1992, 96, 6796–6806. 10.1063/1.462569. [DOI] [Google Scholar]
  46. Woon D. E.; Dunning T. H. Gaussian Basis Sets for Use in Correlated Molecular Calculations. 3. The Atoms Aluminum Through Argon. J. Chem. Phys. 1993, 98, 1358–1371. 10.1063/1.464303. [DOI] [Google Scholar]
  47. Balabanov N. B.; Peterson K. A. Systematically convergent basis sets for transition metals. I. All-electron correlation consistent basis sets for the 3d elements Sc-Zn. J. Chem. Phys. 2005, 123, 064107. 10.1063/1.1998907. [DOI] [PubMed] [Google Scholar]
  48. Knowles P. J.; Hampel C.; Werner H. J. Coupled Cluster Theory for High Spin, Open Shell Reference Wave Functions. J. Chem. Phys. 1993, 99, 5219–5227. 10.1063/1.465990. [DOI] [Google Scholar]; Erratum J. Chem. Phys. 2000, 112, 3106-3107. 10.1063/1.480886. [DOI] [Google Scholar]
  49. Douglas M.; Kroll N. M. Quantum Electrodynamical Corrections to the Fine Structure of Helium. Ann. Phys. 1974, 82, 89–155. 10.1016/0003-4916(74)90333-9. [DOI] [Google Scholar]
  50. Hess B. A. Applicability of the No-Pair Equation with Free-Particle Projection Operators to Atomic and Molecular Structure Calculations. Phys. Rev. A: At., Mol., Opt. Phys. 1985, 32, 756–763. 10.1103/physreva.32.756. [DOI] [PubMed] [Google Scholar]
  51. De Jong W. A.; Harrison R. J.; Dixon D. A. Parallel Douglas-Kroll Energy and Gradients in NWChem: Estimating Scalar Relativistic Effects using Douglas-Kroll Contracted Basis Sets. J. Chem. Phys. 2001, 114, 48–53. 10.1063/1.1329891. [DOI] [Google Scholar]
  52. Werner H. J.; Knowles P. J. A Second Order Multiconfiguration SCF Procedure with Optimum Convergence. J. Chem. Phys. 1985, 82, 5053–5063. 10.1063/1.448627. [DOI] [Google Scholar]
  53. Knowles P. J.; Werner H.-J. An Efficient Second-Order MC SCF Method for Long Configuration Expansions. Chem. Phys. Lett. 1985, 115, 259–267. 10.1016/0009-2614(85)80025-7. [DOI] [Google Scholar]
  54. Schmidt M. W.; Gordon M. S. The Construction and Interpretation of MCSCF Wavefunctions. Annu. Rev. Phys. Chem. 1998, 49, 233–266. 10.1146/annurev.physchem.49.1.233. [DOI] [PubMed] [Google Scholar]
  55. Nakano H. Quasidegenerate Perturbation Theory with Multiconfigurational Self-Consistent-Field Reference Functions. J. Chem. Phys. 1993, 99, 7983–7992. 10.1063/1.465674. [DOI] [Google Scholar]
  56. Celani P.; Werner H.-J. Multireference Perturbation Theory for Large Restricted and Selected Active Space Reference Wave Functions. J. Chem. Phys. 2000, 112, 5546–5557. 10.1063/1.481132. [DOI] [Google Scholar]
  57. Hirao K. Multireference Møller-Plesset Method. Chem. Phys. Lett. 1992, 190, 374–380. 10.1016/0009-2614(92)85354-d. [DOI] [Google Scholar]
  58. Rintelman J. M.; Adamovic I.; Varganov S.; Gordon M. S. Multireference Second-Order Perturbation Theory: How Size Consistent is “Almost Size Consistent.”. J. Chem. Phys. 2005, 122, 044105. 10.1063/1.1817891. [DOI] [PubMed] [Google Scholar]
  59. 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.

Associated Data

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

Supplementary Materials

ao1c04531_si_001.pdf (894.7KB, pdf)

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES