Joel Bowman

From Wikipedia the free encyclopedia

Picture of Prof. Emeritus Joel Mark Bowman
Joel M. Bowman
BornJan. 16, 1948
EducationUniversity of California, Berkeley
California Institute of Technology
Scientific career
InstitutionsEmory University
Doctoral advisorAron Kuppermann

Joel Mark Bowman is an American physical chemist and educator. He is the Samuel Candler Dobbs Professor of Theoretical Chemistry at Emory University.[1]

Publications, honors and awards[edit]

Bowman is the author or co-author of more than 600 publications and is a member of the International Academy of Quantum Molecular Sciences. He received the Herschbach Medal[2]. He is a fellow of the American Physical Society[3] and of the American Association for the Advancement of Science.[1]

Research interests[edit]

His research interests are in basic theories of chemical reactivity.[1] His AAAS fellow citation cited him “for distinguished contributions to reduced dimensionality quantum approaches to reaction rates and to the formulation and application of self-consistent field approaches to molecular vibrations.”[1]

Bowman is well known for his contributions in simulating potential energy surfaces for polyatomic molecules and clusters. Approximately fifty potential energy surfaces for molecules and clusters have been simulated employing his permutationally invariant polynomial method.[4]

Permutationally Invariant Polynomial (PIP) method[edit]

Linear least-squares polynomial fits of indicated order n and r-value in the variables r and y to a Morse potential.[4]

Simulating potential energy surfaces (PESs) for reactive and non-reactive systems is of broad utility in theoretical and computational chemistry. Development of global PESs, or surfaces spanning a broad range of nuclear coordinates, is particularly necessary for certain applications, including molecular dynamics and Monte Carlo simulations and quantum reactive scattering calculations.

Rather than utilizing all of the internuclear distances, theoretical chemists often analytical equations for PESs by using a set of internal coordinates. For systems containing more than four atoms, the count of internuclear distances deviates from the equation 3N−6 (which represents the degrees of freedom in a three-dimensional space for a nonlinear molecule with N atoms).[5][6] As an example, Collins and his team developed a sophisticated method employing different sets of 3N−6 internal coordinates, which they applied to analyze the H+ CH4 reaction. They addressed permutational symmetry by replicating data for permutations of the H atoms.[7] In contrast to this approach, the PIP method uses the linear least-square method to accurately match tens of thousands of electronic energies for both reactive and non-reactive systems mathematically.

Methodology[edit]

Generally, the functions used in fitting potential energy surfaces to experimental and/or electronic structure theory data are based on the choice of coordinates. Most of the chosen coordinates are bond stretches, valence and dihedral angles, or other curvilinear coordinates such as the Jacobi coordinates or polyspherical coordinates.[citation needed] There are advantages to each of these choices.[4] In the PIP approach, the N(N − 1)/2 internuclear distances are utilized. This number of variables is equal to 3N −6 (or 3N − 5 = 1 for diatomic molecules) for N = 3, 4 and differs for N ≥ 5. Thus, N = 5 is an important boundary that affects the choice of coordinates. An advantage of employing this variable set is its inherent closure under all permutations of atoms. This implies that regardless of the order in which atoms are permuted, the resulting set of variables remains unchanged. However, the main focus pertains to permutations involving identical atoms, as the PES must be invariant under such transformations.[4]

Potential energy curve of the internal rotation of CH3OH from a full-dimensional, permutationally invariant potential energy surface[4]

PIP utilizing Morse variables of the form , where is the distance between atoms and and is a range parameter) offers a method for mathematically characterizing high-dimensional PESs. By fixing the range parameter in the Morse variable, the PES can be determined through linear least-squares fitting of computed electronic energies for the system at various structural arrangements. The adoption of a permutationally invariant fitting basis, whether in the form of all internuclear distances or transformed variables like Morse variables, facilitates the attainment of accurate fits for molecules and clusters.[citation needed]

Selected publications[edit]

  • Bowman, J. M. (2000), "Beyond Platonic Molecules (An Invited "Perspective")", Science, 290 (5492): 724–725, doi:10.1126/science.290.5492.724, PMID 11184203, S2CID 93762491.
  • Townsend, D.; Lahankar, S. A.; Lee, S. K.; Chambreau, S. D.; Suits, A. G.; Zhang, X.; Rheinecker, J.; Harding, L. B.; Bowman, J. M. (2004), "The roaming atom: Straying from the reaction path in formaldehyde decomposition", Science, 306 (5699): 1158–61, Bibcode:2004Sci...306.1158T, doi:10.1126/science.1104386, PMID 15498970, S2CID 31464376.
  • Huang, X.; McCoy, A. B.; Bowman, J. M.; Johnson, L. M.; Savage, C.; Dong, F.; Nesbitt, D. J. (2006), "Quantum deconstruction of the infrared spectrum of CH5+", Science, 311 (5757): 60–3, Bibcode:2006Sci...311...60H, doi:10.1126/science.1121166, PMID 16400143, S2CID 26158108.
  • Yin, H. M.; Kable, S. H.; Zhang, X.; Bowman, J. M. (2006), "Signatures of H2CO Photodissociation from two electronic states", Science, 311 (5766): 1443–6, Bibcode:2006Sci...311.1443Y, doi:10.1126/science.1123397, PMID 16527976, S2CID 37885013.
  • Vibrational Dynamics of Molecules, World Scientific Publishing, 2022.

References[edit]

  1. ^ a b c d Selected Academic Highlights (PDF), Emory University, Fall 2005, archived from the original (PDF) on 2009-11-28, retrieved 2009-04-14.
  2. ^ Esciencecommons (2013-08-22). "Joel Bowman's view from the top of theoretical chemistry". eScienceCommons. Retrieved 2024-04-07.
  3. ^ APS Membership listing, Division of Atomic, Molecular & Optical Physics, 2008 Archived 2008-11-21 at the Wayback Machine.
  4. ^ a b c d e Qu, Chen; Yu, Qi; Bowman, Joel M. (2018-04-20). "Permutationally Invariant Potential Energy Surfaces". Annual Review of Physical Chemistry. 69 (1): 151–175. Bibcode:2018ARPC...69..151Q. doi:10.1146/annurev-physchem-050317-021139. ISSN 0066-426X. PMID 29401038.
  5. ^ Chen, Jun; Xu, Xin; Xu, Xin; Zhang, Dong H. (2013-06-14). "Communication: An accurate global potential energy surface for the OH + CO → H + CO2 reaction using neural networks". The Journal of Chemical Physics. 138 (22). doi:10.1063/1.4811109. ISSN 0021-9606.
  6. ^ Jiang, Bin; Guo, Hua (2013-08-06). "Permutation invariant polynomial neural network approach to fitting potential energy surfaces". The Journal of Chemical Physics. 139 (5). Bibcode:2013JChPh.139e4112J. doi:10.1063/1.4817187. ISSN 0021-9606. PMID 23927248.
  7. ^ Thompson, Keiran C.; Jordan, Meredith J. T.; Collins, Michael A. (1998-05-22). "Polyatomic molecular potential energy surfaces by interpolation in local internal coordinates". The Journal of Chemical Physics. 108 (20): 8302–8316. Bibcode:1998JChPh.108.8302T. doi:10.1063/1.476259. ISSN 0021-9606.

External links[edit]