Open access peer-reviewed chapter

Infrared Spectra and Density Functional Theoretical Calculation of Transition Metal Oxide Reaction with Monochloromethane

Written By

Yanying Zhao, Xin Liu and Shuang Meng

Submitted: 18 November 2015 Reviewed: 31 May 2016 Published: 05 October 2016

DOI: 10.5772/64437

From the Edited Volume

Applications of Molecular Spectroscopy to Current Research in the Chemical and Biological Sciences

Edited by Mark T. Stauffer

Chapter metrics overview

1,786 Chapter Downloads

View Full Metrics

Abstract

In this chapter, we presented a short review of past and present experimental and theoretical work on the reactions of the transition metal monoxide and dioxide molecules with monochloromethane in excess argon matrices. A series of infrared absorption spectra combining with density functional theoretical (DFT) calculation characterized that the transition metal monoxide molecules produced by laser-ablated higher oxides activated C─H and C─Cl bonds of CH3Cl to first form the weakly bound MO(CH3Cl) (M = Sc, Y, Nb, Ta, Ti, Zr, Mn, Fe) complexes, which further photoisomerized to the more stable chlorine-transfer (Cl-transfer) CH3OMCl (M = Sc, Y), CH3M(O)Cl (M = Ti, Zr), CH3MOCl (M = Mn, Fe), and agostic hydrogen-transfer (H-transfer) CH2ClMOH (M = Sc, Y, Nb, Ta) products upon limited light excitation. Transition metal dioxides reaction with CH3Cl also formed MO2(CH3Cl) (M = Ti, Zr, Nb, Ta) complexes, which were further rearranged to the more stable Cl-transfer CH3OM(O)Cl (M = Ti, Zr) and agostic H-transfer CH2ClM(O)OH (M = Nb, Ta) molecules between the metal center atom and the chlorine atom upon ultraviolet light irradiation. Their different reactivity was interpreted according to the different valence electrons of metal center.

Keywords

  • monochloromethane
  • chlorine transfer
  • hydrogen transfer
  • transition metal oxides
  • agostic interaction

1. Introduction

Monochloromethane, as the one of the simplest halohydrocarbons, also called methyl chloride, plays an important role in the industrial, synthetic, materials chemistry. It is always regarded that monochloromethane is the largest natural source of ozone-depleting chlorine compounds and accounts for about 15% of the present atmospheric chlorine content as one kind of chlorinated volatile organic compounds (CVOCs). At present, monochloromethane is observed in the dry leaf with the content of 0.1–0.3 μg/g/h, and large emissions of monochloromethane are observed from some common certain types of ferns and dipterocarpaceae [1, 2]. Monochloromethane is also industrially produced by the oxidation and chlorination reaction of methane in the presence of metal chloride catalyst, and drying monochloromethane conversed to gasoline and olefins on the methanol to gasoline (MTG) and the methanol to olefins (MTO) catalysts [3, 4]. The conversion of methyl chloride to hydrocarbons has been investigated since the mid-1980s [5]. The product distribution of methyl chloride to hydrocarbons is strikingly similar to methanol conversion over the same topology [6]. Recently several ZSM-5 zeolites and SAPO sieves catalysts were reported the high performances on the catalytic conversion of monochloromethane to light olefins [79]. Modified SAPO-34 catalysts were also chosen to enhance its catalytic performance for the conversion of chloromethane to light olefins [1013]. The oxidation addition of metal into carbon-halogen bonds is a key step in many stoichiometric and catalytic reactions. Activation of compounds containing C─X (X = Cl, Br, I) bonds attracts widespread interest due to the underactive organic functional group and the inherent chemical properties. Predominantly alkyl and aryl halides are extensively applied as electrophiles in the transition metal-catalyzed cross-coupling reactions [1416]. It has a far-reaching significance on carbon-chlorine (C─Cl) bond catalytic oxidation on the conversion of monochloromethane to gasoline and olefins.

Scheme 1.

The reactivity of transition metal monoxide and dioxide with monochloromethane in argon from Refs. [2730].

The reactions of transition metal centers with chloromethane may serve as the simplest model for understanding the intrinsic mechanism of the organic halides catalytic oxidation processes. The reactions on transition metal atoms with monochloromethane have been intensively studied in solid noble gas matrices. Investigations have reported that C─X bond of CH3X (X = F, Cl, Br, I) are activated by transition metal atoms [1722]. The higher valence of group 6 metals can form the methylidyne complexes CH = MH2X (M = Mo, W, X = H, F, Cl, Br) [2326]. In this chapter, the reactions of simple transition-metal oxide molecules with monochloromethane in solid argon were reviewed using matrix infrared absorption spectroscopy and density functional theoretical (DFT) calculations. As shown in Scheme 1, the ground-state transition metal monoxide molecules activated carbon-hydrogen (C─H) and C─Cl bond of CH3Cl upon a certain wavelength excitation in argon matrices. The weakly bound MO(CH3Cl) (x = 1, 2; M = Sc, Y, Nb, Ta, Ti, Zr, Mn, Fe) complexes were initially formed and then isomerized to the more stable Cl-transfer CH3OMCl (M = Sc, Y) and CH3M(O)Cl (M = Ti, Zr, Nb, Ta, Mn, Fe), and agostic H-transfer CH2ClMOH (M = Sc, Y, Nb, Ta) isomers upon limited visible light excitation. The MO2(CH3Cl) (M = Ti, Zr, Nb, Ta), which were formed from the reactions on MO2 with CH3Cl, were further rearranged to the more stable Cl-transfer CH3OM(O)Cl (M = Ti, Zr) and H-transfer CH2ClM(O)OH (M = Nb, Ta) molecules with agostic interactions between the chlorine and the metal center under ultraviolet light irradiation.

Advertisement

2. Experimental and computational methods

The experimental setup for pulsed laser-ablated and matrix isolation Fourier transform infrared (FTIR) spectroscopic technique has been previously described in detail [31]. Briefly, the 1064 nm Nd:YAG laser fundamental (Spectra Physics, DCR 150, 20 Hz repetition rate, and 8 ns pulse width) was focused onto the rotating bulk metal oxide targets, which were prepared by sintered metal oxide powders. Laser-evaporation of bulk higher metal oxide targets has been proved to be an extensively available technique to prepare pure metal oxides in noble gas matrices [3234]. Using standard manometric technique, the CH3Cl/Ar samples were mixed at a proper proportion in a stainless steel vacuum line. The CH3Cl sample was subjected to several freeze-pump-thaw cycles at 77 K before use. The laser-evaporated species were co-deposited with chloromethane in excess argon onto a CsI window cooled normally to 6 K by a closed-cycle helium refrigerator (ARS, 202N). The matrix samples were deposited at a rate of approximately 5 mmol/h for 1–2 h. Isotopic-labeled 13CH3Cl and CD3Cl (ISOTEC, 99%) were used without further purification. Infrared spectra between 450 and 4000 cm−1 were recorded on a Bruker IFS 66v/s spectrometer using HgCdTe (MCT) detector cooled by liquid N2 at 0.5 cm−1 resolution. Samples were annealed to different temperatures and cooled back to 6 K to acquire the spectra, and selected samples were subjected to visible or broadband irradiation using a 250 W high-pressure mercury arc lamp with selected wavelength glass filters.

Density functional theoretical calculations were performed by using Gaussian 03 programs [35] to identify the experimental assignments. The three-parameter hybrid functional, according to Becke with additional correlation corrections from Lee, Yang, and Parr (B3LYP), was utilized [36, 37] to optimize ground geometries, calculate frequencies, and derive the zero-point vibrational energies. Transition-state optimizations were performed with the Berny geometry optimization algorithm at the B3LYP level. The 6-311++G(d, p) basis set was used for the H, C, O, Cl, Sc, Ti, Mn, and Fe atoms [38, 39], DGDZVP basis set for Y, Zr, and Nb atoms [40, 41], and the scalar-relativistic SDD pseudopotential and basis set for Ta atom [42, 43]. In addition, the CCSD(T) method was also applied to accurately calculate the single-point energies of the B3LYP-optimized structures with the same basis sets [44].

Advertisement

3. Transition metal monoxides reaction with CH3Cl

Reaction of transition metal monoxides (ScO, YO, TiO, ZrO, NbO, TaO, MnO, FeO) with monochloromethane was investigated in solid argon by infrared absorption spectroscopy, combining with isotopic substituted experiments and theoretical calculations. The initial reaction step is the formation of the MO(CH3Cl) (M = Sc, Y, Ti, Zr, Nb, Ta, Mn, Fe) complex with metal atom bound with chlorine atom and/or oxygen atom with H atoms on annealing. Upon a certain wavelength photolysis, the MO(CH3Cl) complex was isomerized by the insertion of the M═O to C─H and/or C─Cl/Cl─C bond. Selected region of infrared spectra is illustrated in Figures 14.

Figure 1.

Difference spectra in the selected regions scandium monoxide with isotopic substituted chloromethane in excess argon. (Spectrum taken after 15 min of broadband irradiation minus spectrum taken after 25 K annealing). (a) 0.5% CH3Cl, (b) 0.5% 13CH3Cl, and (c) 0.5% CD3Cl. (Reprinted with the permission from Ref. [27]. Copyright 2013 American Chemical Society).

Figure 2.

Difference spectra in the selected regions from co-deposition of a laser-ablated TiO2 target in excess argon. Spectrum taken after 15 min of full-arc broadband photolysis irradiation (λ < 300 nm) followed by the 25 K annealing minus spectrum taken after sample annealing at 25 K. (a) 0.5% CH3Cl, (b) 0.5% 13CH3Cl, and (c) 0.5% CD3Cl. (Reprinted with the permission from Ref. [28]. Copyright 2013 American Chemical Society).

Figure 3.

Infrared spectra in the selected region from co-deposition of laser-ablated MnO2 and Fe2O3 target with isotopically substituted CH3Cl in excess argon. Spectra were taken after 1.5 h of sample deposition followed by 25 K annealing and 15 min of irradiation (300 < λ < 580 nm) and 25 K annealing. (a) 0.5% CH3Cl, (b) 0.5% 13CH3Cl, and (c) 0.5% CD3Cl. (Reprinted with permission from Ref. [30]. Copyright 2013, with permission from Ref. [28] from Elsevier).

Figure 4.

Difference spectra in the selected regions from co-deposition of laser-evaporated niobium oxides with isotopic-substituted monochloromethane in excess argon. Spectrum taken after 15 min of full-arc photolysis minus spectrum taken after sample annealing at 25 K. (a) 0.5% CH3Cl, (b) 0.5% 13CH3Cl, and (c) 0.5% CD3Cl. (Reprinted with the permission from Ref. [29]. Copyright 2013 American Chemical Society).

In both the scandium and yttrium experiments, two MO(CH3Cl) (M = Sc, Y) complex isomers were formed spontaneously on annealing [27]. These absorptions of MO(CH3Cl) (M = Sc, Y) complex are observed at 898.4 and 919.1 cm−1 for Sc, and 1050.9, 805.9, and 784.8 cm−1 for Y, as shown in Table 1, which are corresponding to the Sc─O and Y─O vibration frequencies. The CH3OMCl and CH2ClMOH (M = Sc, Y) molecules were produced from the weakly bound MO(CH3Cl) complexes through photoinduced isomerization reactions on 250–300 nm wavelength irradiation, as shown in Figure 1. The CH3OMCl (M = Sc, Y) isomer observed at 1171.5 and 565.6 cm−1 for Sc and 1149.2 and 490.9 cm−1 for Y can be regarded as being formed through the addition of the C─Cl bond to the O═M bond, whereas the CH2ClMOH (M = Sc, Y) isomer observed at 3775.0 and 738.4 for Sc, and 3774.2 and 627.6 for Y is formed through the addition of the C─H bond to the M═O bond. On the basis of DFT calculations, the MO(CH3Cl) (M = Sc, Y) complex with Cs structure is more stable than the complex with C3v structure by 25.5 (Sc) or 24.0 (Y) kJ/mol. Both CH3OMCl and CH2ClMOH (M = Sc, Y) molecules are more stable than the MO(CH3Cl) complex isomers. The CH3OMCl (M = Sc, Y) molecule was predicted to proceed through a transition state with an energy barrier of 17.7 for Sc and 8.4 kJ/mol for Y from the MO(CH3Cl) complex, whereas the CH2ClMOH isomer also proceeded through a transition state with a much higher energy barrier of 160.1 for Sc and 178.5 kJ/mol for Y from the MO(CH3Cl) complex. The CH3OMCl (M = Sc, Y) structure is about 173.0 for Sc and 180.6 kJ/mol for Y lower in energy than the CH2ClScOH and CH2ClYOH isomer. The CH2ClMOH (M = Sc, Y) molecule was also calculated to involve agostic interaction observed between the metal atom and the chlorine atom due to short bond distances of 2.598 Å for Sc─Cl and 2.821 Å for Y─Cl. Such interaction is quite similar to the agostic interactions generally defined to characterize the distortion of an organometallic moiety, which brings an appended C─H bond into close proximity with the metal center [17, 21, 45].

Molecule Ground state Point group Vibrational frequencyb Binding energyc Ref.
ScO(CH3Cl) 2A′ C s 898.4 −34.9 [27]
2A1 C 3v 919.1 −9.4
YO(CH3Cl) 2A′ C s 1050.9, 783.5 −37.0 [27]
2A1 C 3v 805.9 −13.0
TiO(CH3Cl) 3A′ C s 961.8 −41.4 [28]
ZrO(CH3Cl) 3A C 1 898.2 −40.5 [28]
NbO(CH3Cl) 4A′ C s 935.6 −29.5 [29]
TaO(CH3Cl) 2A C 1 991.5 −5.0 [29]
MnO(CH3Cl) 6A′ C s 843.4 −38.9 [30]
FeO(CH3Cl) 5A′ Cs 882.7 −97.4 [30]

Table 1.

Ground electronic states, symmetry point groups, vibrational frequencies (cm−1) and binding energies (kJ/mol) for the MO(CH3Cl) species in solid argona.

a Only the values for the most abundant metal isotope are listed.


b The mode assignments of the experimental vibrational frequencies are discussed in the cited literature.


c Relative to the energy sum of ground metal oxide and CH3Cl.


For IVB metal monoxides, the ground-state MO(CH3Cl) (M = Ti, Zr) complexes correlate to the ground-state TiO (3Δ) and ZrO (1Σ). The binding energies are predicted to be 20.5 (Ti) and 12.2 kcal/mol (Zr), which are larger than the corresponding values of TiO(CH4) and ZrO(CH4) [45, 46]. The MO(CH3Cl) (M = Ti, Zr) complexes can rearrange to the CH3M(O)Cl isomers by metal terminal insertion to C─Cl bond upon UV light irradiation (λ < 300 nm), which are observed at 999.5 and 526.2 cm−1 for Ti, and 915.2 and 488.3 cm−1 for Zr, as shown in Figure 2. Theoretical calculations also indicated that the electronic state crossings exist from the MO (M = Ti, Zr) + CH3Cl reaction to the more stable CH3M(O)Cl molecules through the MO(CH3Cl) complexes traverse their corresponding transition states. The CH3M(O)Cl (M = Ti, Zr) molecule was predicted to have a singlet ground state without symmetry. According to CCSD(T) single-point calculations on B3LYP optimization geometry, the singlet ground state is 44.3 kcal/mol for CH3Ti(O)Cl and 52.2 kcal/mol for CH3Zr(O)Cl lower in energy than its corresponding triplet state. The triplet MO(CH3Cl) (M = Ti, Zr) isomerized to the singlet CH3M(O)Cl (M = Ti, Zr) molecule through their corresponding transition states, which indicated that these reactions related to the spin crossing under UV light irradiation.

Molecule Ground state Point group Vibrational frequencyb Binding energyc Ref.
CH3OScCl 2A C 1 1171.5, 565.6 −268.8 [27]
CH2ClScOH 2A C 1 3775.0, 738.4 −95.8 [27]
CH3OYCl 2A C 1 1149.2, 490.9 −300.3 [27]
CH2ClYOH 2A C 1 3774.2, 627.6 −119.7 [27]
CH3Ti(O)Cl 1A C 1 999.5, 526.2 −324.1 [28]
CH3Zr(O)Cl 1A C 1 915.2, 488.3 −349.0 [28]
CH2ClNb(O)H 2A C 1 1698.0, 985.0 −136.7 [29]
CH2ClTa(O)H 2A C 1 1760.0, 984.8 −182.7 [29]
CH3MnOCl 6A C 1 569.6, 542.2 −55.2 [30]
CH3FeOCl 5A C1 570.4, 561.5 −50.2 [30]

Table 2.

Ground electronic states, symmetry point groups, vibrational frequencies (cm−1), and binding energies (kJ/mol) for the isomers of MO(CH3Cl) in solid argona.

a Only the values for the most abundant metal isotope are listed.


b The mode assignments of the experimental vibrational frequencies are discussed in the cited literature.


c Relative to the energy sum of ground metal oxide and CH3Cl.


The ground-state NbO(CH3Cl) and TaO(CH3Cl) molecules are related to the ground-state NbO (4Σ) and TaO (2∆). The predicted binding energies of 29.5 (Nb) and 5.0 kJ/mol (Ta) are larger than the corresponding values of NbO(CH4) and TaO(CH4) complexes [46], which were predicted to be very weakly interaction with the metal atom being bound to three hydrogen atoms of CH4. The MO(CH3Cl) (M = Nb, Ta) complexes rearranged to the more stable doublet CH2ClM(O)H isomer upon visible light excitation, as shown in Table 2. Thus, some excited states may be involved during the reaction process. The CH2ClM(O)H molecules were predicted to involve agostic interactions between the chlorine atom and the metal center. It is quite interesting to note that the CH2ClM(O)H (M = Nb, Ta) molecules involve agostic interactions between the chlorine atom and the metal atom. It is notable that the group 5 metal methylidene complexes are more agostically distorted than the group 4 metal complexes. Taking CH2ClNb(O)H as an example, the ∠ClCNb was predicted to be only 80.4° with a Cl─Nb distance of 2.624 Å. Agostic distortion interaction is a universal phenomenon in the structures of the early transition metal alkylidene complexes and even more popular in the structures of the small methylidene complexes, in which agostic interactions are also observed between the group 4–6 transition metal atom and one of the R-hydrogen atoms.

The reactions of FeO and MnO with CH3Cl first formed the MO(CH3Cl) (M = Mn, Fe) complexes when annealing, which can isomerize to CH3MOCl (M = Mn, Fe) upon 300 < λ < 580 nm irradiation. The products were characterized by isotopic IR studies with CD3Cl and 13CH3Cl and density functional calculations, as shown in Figure 3. Based on theoretical calculations, the MO(CH3Cl) (M = Mn, Fe) complexes have 5A′ for Fe and 6A′ ground state for Mn with Cs symmetry, respectively, as listed in Table 1. The binding energies of MO(CH3Cl) (M = Mn, Fe) are 9.3 and 23.3 kcal/mol lower than MO + CH3Cl, which are higher in energy than MO(CH4) and MO(Ng) (Ng = Ar, Kr, Xe) at the same calculation level [4648]. The accurate CCSD(T) single-point calculations illustrate the CH3MOCl isomerism are 13.8 and 3.1 kcal/mol lower in energy than the MO(CH3Cl) (M = Mn, Fe) complexes.

The different reactivity of metal monoxide with CH3Cl can be rationalized in terms of changes in valence electron structures accompanied by electronic spin state crossing. In the scandium and yttrium reactions, the ground ScO and YO molecules reacted with CH3Cl to form two isomeric MO(CH3Cl) (M = Sc, Y) complexes spontaneously on annealing. Broad-band irradiation produced either the addition of the C─Cl bond to the O═M (M = Sc, Y) bond to form the CH3OMCl (M = Sc, Y) molecules with +II oxidation state of center metal or the addition of the C─H bond to the M═O bond to give the CH2ClMOH isomer with the valence of metal remaining in +II oxidation state. The CH2ClMOH (M = Sc, Y) include one α-chlorine atom to form agostic molecules between chlorine atom and metal center atom with less than 90° of ∠ClCM and short Cl---M (M = Sc, Y) distances. No α-H and/or α-Cl atom for the MO(CH3Cl) complex exist, so no agostic interaction is observed. Sc and Y have only three valence electrons, and hence they are not able to form high oxidation state structures. However, Mn and Fe have five and six valence electrons. Because their d orbitals are fully half-filled and hence are not easily lost, upon 300 < k < 580 nm irradiation the MO(CH3Cl) (M = Mn, Fe) complexes triggered the addition of the C─Cl bond to the M═O bond to form the CH3MOCl molecules with +II valence state. The Ti and Zr metals have four valence electrons, and their oxidation states increase from +II to +IV during the addition of MO insertion into the C─Cl bond to the metal to form CH3M(O)Cl molecules. For Nb and Ta, visible light irradiation triggered the H-atom transfer of the MO(CH3Cl) complexes from CH3Cl to the metal center to form the more stable CH2ClM(O)H isomers with the oxidation states of the metal increasing from the +II to +IV. However, the Nb and Ta have five valence electrons, and they cannot form +V oxidation structures, but possessing one valence electron characteristic of the agnostic chlorine effect.

Advertisement

4. Transition metal dioxides reaction with CH3Cl

The ground-state MO2 (M = Ti, Zr, Nb, Ta) molecules react with CH3Cl to first form the weakly bound MO2(CH3Cl) complexes with O···H and M···Cl bonds. For Ti and Zr, the MO2(CH3Cl) complexes can isomerize to the more stable CH3OM(O)Cl molecules with the addition of the C─Cl bond of CH3Cl to one of the O═M bond of MO2 on annealing after broadband light irradiation (λ < 300 nm), as shown in Figures 2 and 4. And the reaction potential energy profile interpreted the chemical reaction mechanism of C─Cl activation by MO2 (M = Ti, Zr). The photoisomerization reaction of MO2(CH3Cl) (M = Nb, Ta) is quite different from those of MO2(CH3Cl) (M = Ti, Zr). The MO2(CH3Cl) (M = Nb, Ta) complexes were initiated H-transfer under ultraviolet light irradiation to isomerize the more stable CH2ClM(O)OH molecules. The CH2ClM(O)OH (M = Nb, Ta) molecules were predicted to involve agostic interactions between the chlorine atom and the metal center. During the photoisomerization process, no electronic spin state crossings were found, as shown in Table 3, different from the reaction of metal monoxides with CH3Cl.

Molecule Ground state Point group Vibrational frequencyb Binding energyc Ref.
TiO2(CH3Cl) 1A C 1 940.3, 906.2 −95.7 [28]
ZrO2(CH3Cl) 1A′ C s 874.1, 804.2 −84.4 [28]
NbO2(CH3Cl) 2A C 1 948.1, 890.9 −75.66 [29]
TaO2(CH3Cl) 2A′ C s 948.4, 890.8 −52.9 [29]
CH3OTi(O)Cl 1A C 1 1173.0, 1102.6, 634.1 −326.0 [28]
CH3OZr(O)Cl 1A C 1 1153.1, 901.2 −331.9 [28]
CH2ClNb(O)OH 2A C 1 3678.4, 979.4, 712.8 −185.5 [29]
CH2ClTa(O)OH 2A C1 3690.8, 975.9, 605.8, 495.7 −171.8 [29]

Table 3.

Ground electronic states, symmetry point groups, vibrational frequencies (cm−1), and binding energies (kJ/mol) for the product from MO2 + CH3Cl in solid argona.

a Only the values for the most abundant metal isotope are listed.


b The mode assignments of the experimental vibrational frequencies are discussed in the cited literature.


c Relative to the energy sum of ground metal oxide and CH3Cl.


Advertisement

5. Conclusion and outlook

C─Cl and/or C─H bond of monochloromethane activation by transition metal monoxide and dioxide molecules has been investigated using matrix infrared spectroscopy in excess argon and density functional theoretical calculations. The metal monoxide and dioxide molecules prepared by laser-ablated bulk higher oxide targets reacted with monochloromethane to form the weakly bound MO(CH3Cl) (x = 1, 2; M = Sc, Y, Nb, Ta, Ti, Zr, Mn, Fe) complexes, which isomerized to the more stable CH3OMCl (M = Sc, Y), agostic CH2ClMOH (M = Sc, Y, Nb, Ta) and CH3M(O)Cl (M = Ti, Zr, Nb, Ta, Mn, Fe) isomers upon limited visible light excitation. Metal dioxides also reacted with CH3Cl to form MO2(CH3Cl) (M = Ti, Zr, Nb, Ta), which was rearranged to the more stable CH3OM(O)Cl (M = Ti, Zr) and CH2ClM(O)OH (M = Nb, Ta) molecules under ultraviolet light irradiation. Agostic interactions were observed in CH2ClMOH (M = Sc, Y, Nb, Ta) and CH2ClM(O)OH (M = Nb, Ta) between the chlorine atom and the metal center atom.

Advertisement

Acknowledgments

We gratefully acknowledge the financial support from National Natural Science Foundation of China (Grants No. 21273202 and 21473162). Y. Zhao is grateful to the Project Grants 521 Talents Cultivation of Zhejiang Sci-Tech University and China Scholarship Council (CSC) Foundation. This work is also supported by Zhejiang Provincial Top Key Academic Discipline of Chemical Engineering and Technology.

References

  1. 1. Yokouchi Y, Noijiri Y, Barrie LA, Toom-Sauntry D, Machida T, Inuzuka Y, Akimoto H, Li HJ, Fujinuma Y, Aoki SA. Strong source of methyl chloride to the atmosphere from tropical coastal land. Nature. 2000;403:295–298. DOI: 10.1038/35002049
  2. 2. Yokouchi Y, Ikeda M, Inuzuka Y, Yukawa T. Strong emission of methyl chloride from tropical plants. Nature. 2002;416:163–165. DOI: 10.1038/416163a
  3. 3. Tajima N, Tsuneda T, Toyama F, Hirao K. A new mechanism for the first carbon−carbon bond formation in the MTG process: a theoretical study. J. Am. Chem. Soc. 1998;120:8222–8229. DOI: 10.1021/ja9741483
  4. 4. Hazari N, Lglesia E, Labinger JA, Simonetti, DA. Selective homogeneous and heterogeneous catalytic conversion of methanol/dimethyl ether to triptane. Acc. Chem. Res. 2012;45: 653−662. DOI: 10.1021/ar2002528
  5. 5. Olah GA, Gupta B, Felberg JD, et al. Electrophilic reactions at single bonds. 20. Selective monohalogenation of methane over supported acid or platinum metal catalysts and hydrolysis of methyl halides over γ-alumina-supported metal oxide/hydroxide catalysts. A feasible path for the oxidative conversion of methane into methyl alcohol/dimethyl ether. J. Am. Chem. Soc. 1985;107:7097–7105. DOI: 10.1021/ja00310a057
  6. 6. Olsbye U, Svelle S, Bjørgen M, Beato P, Janssens TVW, Joensen F, Bordiga S, Lillerud KP. Conversion of methanol to hydrocarbons: how zeolite cavity and pore size controls product selectivity. Angew. Chem. Int. Ed. 2012;51:5810–5831. DOI: 10.1002/anie.201103657
  7. 7. Xu T, Zhang Q, Song H, Wang Y. Fluoride-treated H-ZSM-5 as a highly selective and stable catalyst for the production of propylene from methyl halides. J. Catal. 2012;295:232–241. DOI: 10.1016/j.jcat.2012.08.014
  8. 8. Zhang R, Zhang B, Shi Z, Liu N, Chen B. Catalytic behaviors of chloromethane combustion over the metal-modified ZSM-5 zeolites with diverse SiO2/Al2O3 ratios. J. Mol. Catal. A: Chem. 2015;398:223–230. DOI: 10.1016/j.molcata.2014.11.019
  9. 9. Kong L, Shen B, Zhao J, Liu J. Comparative study on the chloromethane to olefins reaction over SAPO-34 and HZSM-22. Ind. Eng. Chem. Res. 2014;53:16324–16331. DOI:10.1021/ie5028155
  10. 10. Olsbye U, Saure OV, Muddada NB, Bordiga S, Lamberti C, Nilsen M H, Lillerud KP, Svelle S. Methane conversion to light olefins – how does the methyl halide route differ from the methanol to olefins (MTO) route? Catal. Today. 2011;171:211–220. DOI: 10.1016/j.cattod.2011.04.020
  11. 11. Wei Y, Zhang D, Chang F, Xia Q, Su BL, Liu Z. Ultra-short contact time conversion of chloromethane to olefins over precoked SAPO-34: direct insight into the primary conversion with coke deposition. Chem. Commun. 2009;40:5999–60001. DOI: 10.1039/b909218h
  12. 12. Xu L, Du AP, Wei YX. Synthesis of SAPO-34 with only Si(4Al) species: effect of Si contents on Si incorporation mechanism and Si coordination environment of SAPO-34. Micropor. Mesopor. Mater. 2008;115:332–337. DOI: 10.1016/j.micromeso.2008.02.001
  13. 13. Jiang Z, Shen B, Zhao J, Wang L, Kong L, Xiao W. Enhancement of catalytic performances for the conversion of chloromethane to light olefins over SAPO-34 by modification with metal chloride. Ind. Eng. Chem. Res. 2015;54(49):12293–12302. DOI: 10.1021/acs.iecr.5b03586
  14. 14. Yang CT, Zhang ZQ, Liang J, Liu JH, Lu XY, Chen HH, Liu L. Copper-catalyzed cross-coupling of nonactivated secondary alkyl halides and tosylates with secondary alkyl Grignard reagents. J. Am. Chem. Soc. 2012;134:11124–11127. DOI: 10.1021/ja304848n
  15. 15. Rudolph A, Lautens M. Secondary alkyl halides in transition metal-catalyzed cross-coupling reactions. Angew. Chem. Int. Ed. 2009;48:2656–2670. DOI: 10.1002/anie.200803611
  16. 16. Xiao B, Liu ZJ, Liu L, Fu Y. Palladium-catalyzed C–H activation/cross-coupling of pyridine N-oxides with nonactivated secondary alkyl bromides. J. Am. Chem. Soc. 2013;135:616–619. DOI: 10.1021/ja3113752
  17. 17. Cho HG, Andrews L. Persistent photo-reversible transition-metal methylidene system generated from reaction of methyl fluoride with laser-ablated zirconium atoms and isolated in a solid argon matrix. J. Am. Chem. Soc. 2004;126(33):10485–10492. DOI: 10.1021/ja0486115
  18. 18. Cho HG, Andrews L. Infrared spectra of CH3–MX and CH2X–MH prepared in reactions of laser-ablated gold, platinum, palladium, and nickel atoms with CH3Cl and CH3Br. Organometallics. 2013;32(9):2753–2759. DOI: 10.1021/om400192v
  19. 19. Cho HG, Andrews L. Infrared spectra of CX3─MnX and CX2═MnX2 (X = H, F, Cl) prepared in reactions of laser-ablated manganese atoms with halomethanes. Organometallics. 2011;30(3):477–486. DOI: 10.1021/om100791h
  20. 20. Cho HG, Andrews L. Infrared spectra of simple methylidyne, methylidene, and insertion complexes generated in reactions of laser-ablated rhodium atoms with halomethanes and ethane. Organometallics. 2010;29(10):2211–2222. DOI: 10.1021/om900902p
  21. 21. Cho HG, Andrews L. Infrared spectra of the CH3─MX, CHMH2X─ complexes formed by reaction of methyl halides with laser-ablated group 5 metal atoms. J. Phys. Chem. A. 2006;110:10063–10077. DOI: 10.1021/jp0629644
  22. 22. Cho HG, Andrews L. Infrared spectra of the CH3─MX and CH2─MHX complexes formed by reactions of laser-ablated group 3 metal atoms with methyl halides. J. Phys. Chem. A. 2007;111:2480–2491. DOI: 10.1021/jp067662g
  23. 23. Cho HG, Andrews L. Formation of CH3TiX, CH2═TiHX, and (CH3)2TiX2 by reaction of methyl chloride and bromide with laser-ablated titanium atoms: photoreversible α-hydrogen migration. Inorg. Chem. 2005;44(4):979–988. DOI: 10.1021/ic048615a
  24. 24. Cho HG, Andrews L. Infrared spectra of methylidynes formed in reactions of Re atoms with methane, methyl halides, methylene halides, and ethane: methylidyne C─H stretching absorptions, bond lengths, and s character. Inorg. Chem. 2008;47:1653–1662. DOI: 10.1021/ic701505w
  25. 25. Cho HG, Andrews L. Infrared spectra of platinum insertion and methylidene complexes prepared in oxidative C─H(X) reactions of laser-ablated pt atoms with methane, ethane, and halomethanes. Organometallics. 2009;28:1358–1368. DOI: 10.1021/om801077x
  26. 26. Jonathan TL, Cho HG, Andrews L. Methylidyne XC⋮MX3 (M = Cr, Mo, W; X = H, F, Cl) diagnostic C─H and C─X stretching absorptions and methylidene CH2═MX2 analogues. Organometallics. 2007;26(25):6373–6387. DOI: 10.1021/om700689e
  27. 27. Zhao Y. C–Cl activation by group IV metal oxides in solid argon matrixes: matrix isolation infrared spectroscopy and theoretical investigations of the reactions of MOx (M = Ti, Zr; x = 1, 2) with CH3Cl. J. Phys. Chem. A. 2013;117(27):5664–5674.
  28. 28. Zhao Y, Huang Y, Zheng X, Zhou M. Preparation and characterization of the agostic bonding molecules between metal and chlorine from the reactions of niobium and tantalum monoxide and dioxide molecules with monochloromethane in solid argon. J. Phys. Chem. A. 2010;114(18):5779–5786. DOI: 10.1021/jp102199c
  29. 29. Huang Y, Zhao Y, Zheng X, Zhou M. Matrix isolation infrared spectroscopic and density functional theoretical study of the reactions of scandium and yttrium monoxides with monochloromethane. J. Phys. Chem. A. 2010;114(7):1476–2482. DOI: 10.1021/jp9101948
  30. 30. Zhao Y, Fan K, Huang Y, Zheng X. Matrix isolation infrared spectra, assignment and DFT investigation on reactions of iron and manganese monoxides with CH3Cl. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2013;116:96–101. DOI: 10.1016/j.saa.2013.07.010
  31. 31. Gong Y, Zhou M, Andrews L. Spectroscopic and theoretical studies of transition metal oxides and dioxygen complexes. Chem. Rev. 2009;109(12): 6765–6808. DOI: 10.1021/cr900185x
  32. 32. Zhao Y, Gong Y, Chen M, Zhou M. Noble gas-transition-metal complexes: coordination of VO2 and VO4 by Ar and Xe atoms in solid noble gas matrixes. J. Phys. Chem. A. 2006;110(5):1845–1849. DOI: 10.1021/jp056476s
  33. 33. Wang C, Zhuang J, Wang G, Chen M, Zhao Y, Zheng X, Zhou M. Tantalum dioxide complexes with dinitrogen. Formation and characterization of the side-on and end-on bonded TaO2(NN)x (x = 1–3) complexes. J. Phys. Chem. A. 2010;114(31):8083–8089. DOI: 10.1021/jp103866r
  34. 34. Zhao Y, Gong Y, Chen M, Ding C, Zhou M. Coordination of ScO+ and YO+ by multiple Ar, Kr, and Xe atoms in noble gas matrixes: a matrix isolation infrared spectroscopic and theoretical study. J. Phys. Chem. A. 2005;109(51):11765–11770. DOI: 10.1021/jp054517e
  35. 35. Frisch MJ, Trucks GW, Schlegel HB, Scuseria GE, Robb MA, Cheeseman JR, Montgomery JA, Vreven T, Kudin KN, Burant JC, Millam JM, Iyengar SS, Tomasi J, Barone V, Mennucci B, Cossi M, Scalmani G, Rega N, Petersson GA, Nakatsuji H, Hada M, Ehara M, Toyota K, Fukuda R, Hasegawa J, Ishida M, Nakajima T, Honda Y, Kitao O, Nakai H, Klene M, Li X, Knox JE, Hratchian HP, Cross JB, Adamo C, Jaramillo J, Gomperts R, Stratmann RE, Yazyev O, Austin AJ, Cammi R, Pomelli C, Ochterski JW, Ayala PY, Morokuma K, Voth GA, Salvador P, Dannenberg JJ, Zakrzewski VG, Dapprich S, Daniels AD, Strain MC, Farkas O, Malick DK, Rabuck AD, Raghavachari K, Foresman JB, Ortiz JV, Cui Q, Baboul AG, Clifford S, Cioslowski J, Stefanov BB, Liu G, Liashenko A, Piskorz P, Komaromi I, Martin RL, Fox DJ, Keith T, Al-Laham MA, Peng CY, Nanayakkara A, Challacombe M, Gill PMW, Johnson B, Chen W, Wong MW, Gonzalez C, Pople JA. Gaussian 03, Revision B.05, Gaussian. Inc.: Pittsburgh, PA, 2003.
  36. 36. Becke AD. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993;98:5648–5652. DOI: 10.1063/1.464913
  37. 37. Lee C, Yang W, Parr RG. Development of the Colle–Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B. 1988;B37:785–789. DOI: 10.1103/PhysRevB.37.785
  38. 38. McLean AD, Chandler GS. Contracted Gaussian basis sets for molecular calculations. I. Second row atoms, Z = 11 − 18. J. Chem. Phys. 1980;72:5639–5648. DOI: 10.1063/1.438980
  39. 39. Krishnan R, Binkley JS, Seeger R, Pople JA. Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 1980;72:650–654. DOI: 10.1063/1.438980
  40. 40. Godbout N, Salahub DR, Andzelm J, Wimmer E. Optimization of Gaussian-type basis sets for local spin density functional calculations. Part I. Boron through Neon, optimization technique and validation. Can. J. Chem. 1992;70:560−571. DOI: 10.1139/v92-079
  41. 41. Sosa C, Andzelm BC, Wimmer E, Dobbs KD, Dixon DA. A local density functional study of the structure and vibrational frequencies of molecular transition-metal compounds. J. Phys. Chem. 1992;96:6630–6636. DOI: 10.1021/j100195a022
  42. 42. Dolg M, Stoll H, Preuss H. Energy-adjusted abinitio pseudopotentials for the rare earth elements. J. Chem. Phys. 1989;90(3):1730–1734. DOI: 10.1063/1.456066
  43. 43. Andrae D, Haeussermann U, Dolg M, Stoll H, Preuss H. Energy-adjusted ab initio pseudopotentials for the second and third row transition elements. J. Theor. Chim. Acta, 1990;77(2):123–141. DOI:10.1007/BF01114537
  44. 44. Pople JA, Head-Gordon M, Raghavachari K. Quadratic configuration interaction. a general technique for determining electron correlation energies. J. Chem. Phys. 1987;87:5968–5975. DOI: 10.1063/1.453520
  45. 45. Wang G, Gong Y, Chen M, Zhou M. Methane activation by titanium monoxide molecules: a matrix isolation infrared spectroscopic and theoretical study. J. Am. Chem. Soc. 2006;128(17):5974–5980. DOI: 10.1021/ja0604010
  46. 46. Wang G, Zhou M. Probing the intermediates in the MO + CH4 ↔ M + CH3OH reactions by matrix isolation infrared spectroscopy. Int. Rev. Phys. Chem. 2008;27(1):1–25. DOI: 10.1080/01442350701685946
  47. 47. Zhao Y, Gong Y, Zhou M. Matrix isolation infrared spectroscopic and theoretical study of NgMO (Ng = Ar, Kr, Xe; M = Cr, Mn, Fe, Co, Ni) complexes. J. Phys. Chem. A. 2006;110(37):10777–10782. DOI: 10.1021/jp064100o
  48. 48. Zhao Y, Zhou M. Are matrix isolated species really “isolated”? Infrared spectroscopic and theoretical studies of noble gas-transition metal oxide complexes. Sci. China Chem. 2010;53(2):327–336. DOI: 10.1007/s11426-010-0044-9

Written By

Yanying Zhao, Xin Liu and Shuang Meng

Submitted: 18 November 2015 Reviewed: 31 May 2016 Published: 05 October 2016