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 currently serves as the Samuel Candler Dobbs Professor of Theoretical Chemistry at Emory University.[1]

Publications, Awards and Affiliations[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 in Theory[2] and an Alexander von Humboldt Research Award.[citation needed] 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]

Joel Mark Bowman is well known for his extensive contribution in simulating Potential Energy Surfaces (PESs) for polyatomic molecules and clusters. Over the previous decade, approximately fifty (50) potential energy surfaces for polyatomic molecules ranging from 4-11 atoms and for clusters have been simulated employing his profound developed method, permutationally invariant polynomial method.[4]

There have also been an enormous improvement in his developed methodology in recent times. This method of simulating potential energy surfaces have been applied recently to formic acid dimer, H2-H2O complex, the four protonated water clusters [H+(H2O)n=2,3,4,6]. It has also been extended to large clusters using the many-body representation

In subsequent paragraphs, we introduce the permutationally invariant polynomial method and its applicability in simple systems.

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 for reactive and non-reactive systems is of crucial importance in theoretical and computational chemistry. Development of global PESs, or surfaces spanning a broad range of nuclear coordinates, has become necessary as a result.

Rather than utilizing all of the internuclear distances, theorists have created analytical equations for PESs by using the 3N − 6 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 molecule with N atoms.[5][6]

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. Though effective, this approach could be considered somewhat brute-force.[7]

In general, the PIP approach 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]

Potential Energy Fitting (PES) is based on the choice of coordinates. The choice of coordinates is related to the number of coordinates needed to simulate the PES. The number of coordinates is historically assumed to be 3N-6 and 3N-5 for nonlinear and linear molecules respectively. In view of that, most of the chosen coordinates are bond stretches, valence and dihedral angles or other curvilinear coordinates such as the Jacobi coordinates or polyspherical coordinates. There are advantages to each of these choices; however, none is truly general and universal.[4]

In the PIP approach, the N(N − 1)/2 internuclear distances are utilized. Interestingly, this number of variables is equal to 3N −6 (or 3N − 5 for diatomic) for N = 2, 3, 4 and differs for N ≥ 5. Thus, N = 5 is an important boundary that affects the choice of coordinates. Clearly, using all the internuclear distances or Morse variables is general and universal for all PESs. An additional 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 integrity of the Potential Energy Surface (PES) remains invariant under such transformations.[4]

Summary[edit]

Potential energy curve of the internal rotation of CH3OH from a full-dimensional, permutationally invariant potential energy surface[8]

In summary, PIP utilizing Morse variables offer a comprehensive method for mathematically characterizing high-dimensional Potential Energy Surfaces (PESs). By fixing a range parameter a in the Morse variable, the PES can be effectively determined through simple linear least-squares fitting of electronic energies ranging from thousands to hundreds of thousands.

The adoption of a permutationally invariant fitting basis, whether in the form of all internuclear distances or transformed variables like Morse variables, has facilitated the attainment of accurate fits for molecules and clusters containing up to 10 atoms. Numerous PESs for both reactive and nonreactive systems have been documented in the literature as a result. Furthermore, leveraging a comprehensive molecular many-body representation has enabled the recent development of ab initio PESs for sizable clusters such as water and protonated water.

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 {{citation|title=Selected Academic Highlights|publisher=Emory University|date=Fall [[2005|url=http://www.emory.edu/PROVOST/IPR/documents/selectedacademichighlights/2005_highlights_fall.pdf%7Caccess-date=2009-04-14%7Carchive-url=https://web.archive.org/web/20091128042922/http://www.emory.edu/PROVOST/IPR/documents/selectedacademichighlights/2005_highlights_fall.pdf%7Carchive-date=2009-11-28%7Curl-status=dead}}.
  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 Qu, Chen; Yu, Qi; Bowman, Joel M. (2018-04-20). "Permutationally Invariant Potential Energy Surfaces". Annual Review of Physical Chemistry. 69 (1): 151–175. doi:10.1146/annurev-physchem-050317-021139. ISSN 0066-426X.
  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). doi:10.1063/1.4817187. ISSN 0021-9606.
  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. doi:10.1063/1.476259. ISSN 0021-9606.
  8. ^ Qu, Chen; Yu, Qi; Bowman, Joel M. (2018-04-20). "Permutationally Invariant Potential Energy Surfaces". Annual Review of Physical Chemistry. 69 (1): 151–175. doi:10.1146/annurev-physchem-050317-021139. ISSN 0066-426X.

External links[edit]