Elsevier

Chemical Physics Letters

Volume 318, Issues 4–5, 25 February 2000, Pages 291-297
Chemical Physics Letters

Ab initio molecular orbital study of the mechanism of photodissociation of formamide

https://doi.org/10.1016/S0009-2614(00)00021-XGet rights and content

Abstract

In the present Letter, the reaction pathways for H2NCHO decomposition in the ground and excited states are traced with the use of the MP2 and CASSCF energy gradient techniques. The intersection point of the S1 and T1 surfaces are determined by means of the state-average CASSCF optimization. The n→π excitation leads to the first excited singlet state, where a radical pair of NH2 and CHO is formed. Then a hydrogen atom transfers from the CHO radical to the NH2 radical, producing CO and NH3 in the ground state. This is in good agreement with recent experimental findings. The calculated results also show that H2 and HNCO can be formed along a two-step T1 pathway. A recent experimental study on the photochemistry of the H2NCHO suggested that H2 elimination from H2NCHO takes place on the T1 surface.

Introduction

Formamide contains a prototype HN–CO peptide linkage. As an especially important model in the study of biological macromolecules, numerous experimental and theoretical investigations have been devoted to its ground-state structure and barrier to rotation around the C–N bond, which has been reviewed in previous studies 1, 2, 3, 4, 5, 6. Either the planar or non-planar structure of formamide has been reported in these studies. There are also numerous experimental 7, 8, 9, 10and theoretical 11, 12, 13, 14investigations on the electronic excitation of formamide, which have been reviewed in the previous studies 1, 5.

The pyrolysis and photolysis of formamide have received less attention from both experiment and theory. Early in 1974, the pyrolysis of formamide vapor was investigated by Back and Boden [15]. CO and NH3 were observed to be the major products, while H2 is the minor one. Rates of production of CO and H2 were approximately first order at pressures ranging from 6 to 33 Torr at temperatures from 262 to 500°C. Kakumoto et al. [16]have studied the thermal decomposition of formamide diluted in Ar from shock waves over the temperature ranging from 1690 to 2180 K. The rate was monitored by means of the IR emission of the carbon monoxide produced. As part of their investigations, an ab initio study on decomposition of NH2COH into CO+NH3 was performed at the HF level with the 3-21G, 4-31G and 6-31G basis sets. The rate constants of this reaction in low- and high-pressure limits were determined on the basis of the experimental data and the results from ab initio calculations. Boden and Back 15, 17also investigated the photolysis of formamide. Products detected were CO, H2 and NH3 with quantum yields at 150°C of 0.8, 0.6 and 0.2, respectively. It was deduced that there were three major primary processes, forming the products of NH2+CO+H, HNCHO+H and NH3+CO, respectively. Another 193-nm-induced photolysis of formamide in matrices have been recently reported by Lundell et al. [5]. They suggested that in solid Ar, the n–π excitation leads to the first exited singlet state, where the radical pair NH2⋯CHO is formed. The hot CHO radical transfers hydrogen to the NH2 radical, forming CO+NH3. In Xe, the n–π excitation occurs as well with 193 nm photolysis. However, when the radical pair is formed in the excited state, a strong external heavy-atom effect due to Xe is present, which induces intersystem crossing to a triplet surface, yielding HNCO+H2. In order to test the validity of the mechanism proposed above, Lundell and co-workers have carried out ab initio studies on the potential energy profiles of NH2CHO dissociating into CHO+NH2 in the low-lying electronic states. Since only the C–N bond distance was scanned, with all other structural parameters kept at ground-state equilibrium values, their potential curves are only qualitatively reliable.

The characterization of the mechanism of a photochemical reaction requires detailed knowledge of ground and excited-state reaction pathways and potential crossing regions where the system decays from one state to another. In the present Letter, the possible dissociation pathways of formamideH2NCHONH3+CO, →HNCO+H2, →NH2+CHO, →HNCHO+H, →H2NCO+Hin the ground, first excited singlet, and lowest triplet states, have been investigated with more advanced quantum chemical techniques. The crossing points of the different potential energy surfaces were determined using the state-averaged CASSCF method. The mechanism leading to different photoproducts was characterized on the basis of the potential energy profiles and their crossing points.

Section snippets

Computational details

All stationary structures of the ground state (S0), first excited singlet state (S1) and lowest triplet state (T1) have been fully optimized with the CASSCF (the complete active space SCF) method. The selection of the active space is the crucial step of the CASSCF calculations, which require some comments. The obvious choice for describing low-lying states of formamide would be six electrons distributed in four orbitals originating from the π and π orbitals of the CO double bond and n

Conclusions

The favorable reaction path can be summarized as follows:

H2NCHO molecules are excited first to the S1 state at 193 nm. Then they dissociate into H2N and CHO radicals on S1 surface, finally forming NH3 and CO ground-state products. This result is in very good agreement with experiment [5]. If there exists a heavy-atom effect, the H2NCHO(S1) molecules should make a transition to the T1 state through the S1/T1 crossing point. On the T1 state surface, the dissociation of formamide into NH2 and CHO

Acknowledgements

This work was supported by NSFC (Grant No. 29873008).

References (22)

  • G. Fogarasi et al.

    J. Phys. Chem. A

    (1997)
  • K.E. Laidig et al.

    Can. J. Chem.

    (1993)
  • X.C. Wany et al.

    Int. J. Quantum Chem.

    (1993)
  • N.A. Burton et al.

    J. Chem. Soc., Faraday Trans.

    (1993)
  • J. Lundell et al.

    J. Phys. Chem. A

    (1998)
  • Y. Li et al.

    J. Am. Chem. Soc.

    (1991)
  • H.D. Hunt et al.

    J. Am. Chem. Soc.

    (1953)
  • D.L. Peterson et al.

    J. Am. Chem. Soc.

    (1954)
  • K. Kaya et al.

    Theor. Chim. Acta

    (1967)
  • M. Basch et al.

    J. Chem. Phys.

    (1968)
  • L.B. Harding et al.

    J. Am. Chem. Soc.

    (1975)
  • Cited by (0)

    View full text