A Cuban Campesino in Chemistry’s Academic Court
- Rigoberto Hernandez*Rigoberto Hernandez*Email: [email protected]Department of Chemistry, Johns Hopkins University, Baltimore, Maryland 21218, United StatesDepartments of Chemical & Biomolecular Engineering and Materials Science and Engineering, Johns Hopkins University, Baltimore, Maryland 21218, United StatesMore by Rigoberto Hernandez
1. Introduction
Figure 1

Figure 1. Then and now. At left, Hernandez is pictured in his high school yearbook in a stylish cotton suit, gray leather shoes, and open collar which was on trend if different from the fashion of polos and tweed that would be the norm just a few months later upon teleporting to Princeton (photo by Mark Maynard). At right, Hernandez is pictured in his current work uniform featuring a tie, white dress shirt, and a jacket (photo by Robert R. Felt).
Figure 2

Figure 2. A selected snapshot of Hernandez’s research efforts shown (clockwise from top left) through journal covers spanning across transition state theory (reprinted with permission from ref (1), copyright 2010 Elsevier), Janus and striped particles (reprinted with permission from ref (2), copyright 2014 AIP Publishing), diffusion through complex environments (reprinted with permission from ref (3), copyright 2010 American Chemical Society), stochastic hard collisions (reprinted with permission from ref (4), copyright 2018 Elsevier), sustainable nanoparticles (reprinted with permission from ref (5), copyright 2016 American Chemical Society), and diversity equity (reprinted with permission from ref (6), copyright 2018 American Chemical Society).
2. Preliminaries
3. Research Track in Theoretical and Computational Chemistry
4. Research Track in Diversity Equity
5. Concluding Remarks
Acknowledgments
My recent theoretical and computational chemistry work has been supported by the National Science Foundation (NSF) through Grant No. CHE-1700749, No. CHE 2001611, and No. OAC-1940152. My recent diversity equity work has been supported by the Sloan Foundation. I am grateful for the recognition and support from the Research Corporation for Science Advancement throughout my academic career. My deep thanks to my research collaborator, Alex Popov, who has been a leader in my theoretical and computational chemistry group since 2004. The OXIDE program would not have succeeded were it not for the work of Dontarie Stallings, Srikant Iyer, and Shannon Watt. Finally, in addition to those already called out by name in the main text, I thank the many mentors that helped me throughout my career, including Dorothy Takos (my sixth grade teacher), Robert Reichert (my junior high school chess coach and math teacher), Charlie Fefferman, Pablo Debenedetti, Rob Tycko, John Tully, Peter Rossky, Silvia Ronco, Linda Raber, Michael Doyle, Peter Saalfrank, Elizabeth Boylan, Celeste Rohlfing, Janice Hicks, Michelle Francl, George Shields, Peter Dorhout Jörg Main, Tom Connelly, and Luis Echegoyen. Many thanks also to my Hialeah High School colleagues and Isaura Delgado, in particular, for helping to identify the photographer who had the rights to the “then” photo in Figure 1.
References
This article references 89 other publications.
- 1Hernandez, R.; Bartsch, T.; Uzer, T. Transition State Theory in Liquids Beyond Planar Dividing Surfaces. Chem. Phys. 2010, 370, 270– 276, DOI: 10.1016/j.chemphys.2010.01.016Google Scholar1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXlslCku7Y%253D&md5=6fabe3c8f033cdc4f304c9e1dcebf14eTransition state theory in liquids beyond planar dividing surfacesHernandez, Rigoberto; Uzer, T.; Bartsch, ThomasChemical Physics (2010), 370 (1-3), 270-276CODEN: CMPHC2; ISSN:0301-0104. (Elsevier B.V.)The success of transition state theory (TST) in describing the rates of chem. reactions has been less-than-perfect in soln. (and sometimes even in the gas phase) because conventional dividing surfaces are only approx. free of recrossings between reactants and products. Recent advances in dynamical systems theory have helped to identify the interconnected manifolds-"superhighways"-leading from reactants to products. The existence of these manifolds has been proven rigorously, and explicit algorithms are available for their calcn. We now show that these extended structures can be used to obtain reaction rates directly in dissipative systems. We also suggest a treatment for the substantially more general case in which the mol. solvent is fully specified by the positions of all its atoms. Specifically, we can construct effective solvent configurations for which the exact TST manifolds can be constructed and used to sample the rates of an open system.
- 2Hagy, M. C.; Hernandez, R. Dynamical simulation of electrostatic striped colloidal particles. J. Chem. Phys. 2014, 140, 034701, DOI: 10.1063/1.4859855Google Scholar2https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXps1CisQ%253D%253D&md5=384c7aaa7f06c8f42d6136f890b0bf63Dynamical simulation of electrostatic striped colloidal particlesHagy, Matthew C.; Hernandez, RigobertoJournal of Chemical Physics (2014), 140 (3), 034701/1-034701/13CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The static and dynamic properties of striped colloidal particles are obtained using mol. dynamics computer simulations. Striped particles with n = 2 to n = 7 stripes of alternating elec. charge are modeled at a high level of detail through a pointwise (PW) representation of the particle surface. We also consider the extent to which striped particles are similar to comparable isotropically attractive particles-such as depletion attracting colloids-by modeling striped particles with an isotropic pair interaction computed by coarse-graining (CG) over orientations at a pair level. Surprisingly, the CG models reproduce the static structure of the PW models for a range of vol. fractions and interaction strengths consistent with the fluid region of the phase diagram for all n. As a corollary, different n-striped particle systems with comparable pair affinities (e.g., dimer equil. const.) have similar static structure. Stronger pair interactions lead to a collapsed structure in simulation as consistent with a glass-like phase. Different n-striped particle systems are found to have different phase boundaries and for certain n's no glass-like state is obsd. in any of our simulations. The CG model is found to have accelerated dynamics relative to the PW model for the same range of fluid conditions for which the models have identical static structure. This suggests striped electrostatic particles have slower dynamics than comparable isotropically attractive colloids. The slower dynamics result from a larger no. of long-duration reversible bonds between pairs of striped particles than seen in isotropically attractive systems. Higher n-striped particles systems generally have slower dynamics than lower n-striped systems with comparable pair affinities. (c) 2014 American Institute of Physics.
- 3Tucker, A. K.; Hernandez, R. Observation of a trapping transition in the diffusion of a thick needle through fixed point scatterers. J. Phys. Chem. A 2010, 114, 9628– 9634, DOI: 10.1021/jp100111yGoogle Scholar3https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXkvV2rtrc%253D&md5=657bd4b9d5ed5c635882a3940f5d73f2Observation of a Trapping Transition in the Diffusion of a Thick Needle through Fixed Point ScatterersTucker, Ashley K.; Hernandez, RigobertoJournal of Physical Chemistry A (2010), 114 (36), 9628-9634CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The long-time correlation functions of an infinitesimally thin needle moving through stationary point scatterers--a so-called Lorentz model--exhibits surprisingly long-time tails. These long-time tails are now seen to persist in a two-dimensional model even when the needle has a finite thickness. If the needles are too thick, then the needles are effectively trapped at all nontrivial densities of the scatterers. At needle widths approx. equal to or smaller than σ = ε/20 where ε is the av. spacing between scatterers, the needle diffuses and exhibits the crossover transition from the expected Enskog behavior to the enhanced translation diffusion seen earlier by Hofling, Frey and Franosch. At this needle width, an increase in its center-of-mass diffusion with respect to increasing d. is seen after a crossover d. of n* ≈ 5 is reached. (The reduced d. n* is defined as n* = nL2 where n is the no. d. of particles and L is the needle length.). The crossover transition for needles with finite thickness is spread over a range of densities exhibiting intermediate behavior. The asymptotic divergence of the center of mass diffusion is suppressed compared to that of infinitely thin needles. Finally, a new diminished diffusion regime, apparently due to the increased importance of head-on collisions, now appears at high scattering densities.
- 4Singh, R. S.; Hernandez, R. Modeling soft core-shell colloids using stochastic hard collision dynamics. Chem. Phys. Lett. 2018, 708, 233– 240, DOI: 10.1016/j.cplett.2018.08.032Google Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXhsFKqtbvP&md5=8fba26473e6e7617d800939900293bbfModeling soft core-shell colloids using stochastic hard collision dynamicsSingh, Rakesh S.; Hernandez, RigobertoChemical Physics Letters (2018), 708 (), 233-240CODEN: CHPLBC; ISSN:0009-2614. (Elsevier B.V.)The equil. structure and thermodn. behavior of a soft matter suspension of hard-core soft-shell (HCSS) colloids have been modeled using stochastic hard collision (SHC) dynamics [Craven et al., J. Chem. Phys. 2013, 138, 244901]. SHC dynamics includes the effects of inter-penetrability of soft particles while retaining the computational efficiency of hard collision dynamics. The HCSS colloids are characterized using two parameters: softness parameter (a measure of the propensity toward interpenetration), and core-to-shell ratio. In both, mol. dynamics and Monte-Carlo simulations, the softness parameter is seen to profoundly affect the effective packing fraction and the thermodn. behavior of the system. Once we accounted for the trivial effects from changes in the reduced vol. fraction on changing softness, we found that the inter-particle penetration continues to affect this thermodn. behavior. The structural origin of this nontrivial effect likely lies in the loss of translational order upon increasing softness.
- 5Cui, Q.; Hernandez, R.; Mason, S. E.; Frauenheim, T.; Pedersen, J. A.; Geiger, F. Mini-review. Sustainable nanotechnology: Opportunities and challenges for theoretical/computational studies. J. Phys. Chem. B 2016, 120, 7297– 7306, DOI: 10.1021/acs.jpcb.6b03976Google Scholar5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XhtFers7nE&md5=9ffc40bbd229be188cacd88cead01028Sustainable Nanotechnology: Opportunities and Challenges for Theoretical/Computational StudiesCui, Qiang; Hernandez, Rigoberto; Mason, Sara E.; Frauenheim, Thomas; Pedersen, Joel A.; Geiger, FranzJournal of Physical Chemistry B (2016), 120 (30), 7297-7306CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)For assistance in the design of the next generation of nanomaterials that are functional and have minimal health and safety concerns, it is imperative to establish causality, rather than correlations, in how properties of nanomaterials det. biol. and environmental outcomes. Due to the vast design space available and the complexity of nano/bio interfaces, theor. and computational studies are expected to play a major role in this context. In this minireview, we highlight opportunities and pressing challenges for theor. and computational chem. approaches to explore the relevant physicochem. processes that span broad length and time scales. We focus discussions on a bottom-up framework that relies on the detn. of correct intermol. forces, accurate mol. dynamics, and coarse-graining procedures to systematically bridge the scales, although top-down approaches are also effective at providing insights for many problems such as the effects of nanoparticles on biol. membranes.
- 6Hernandez, R.; Stallings, D.; Iyer, S. K. In National Diversity Equity Workshops in Chemical Sciences (2011–2017); Hernandez, R., Stallings, D., Iyer, S. K., Eds.; ACS Symposium Series; American Chemical Society; Oxford University Press: Washington DC, 2018; Vol. 1277.Google ScholarThere is no corresponding record for this reference.
- 7Hernandez, R. What’s in a name? (Part 1). 2013; http://everywherechemistry.blogspot.com/2013/04/whats-in-name-part-1.html, accessed January 15, 2021.Google ScholarThere is no corresponding record for this reference.
- 8Coy, S. L.; Hernandez, R.; Lehmann, K. K. Limits on the transition to Gaussian orthogonal ensemble behavior: Saturated radiationless transitions between strongly coupled potential surfaces. Phys. Rev. A: At., Mol., Opt. Phys. 1989, 40, 5935– 5949, DOI: 10.1103/PhysRevA.40.5935Google ScholarThere is no corresponding record for this reference.
- 9Nolan, S. A.; Buckner, J. P.; Kuck, V. J.; Marzabadi, C. H. Analysis by Gender of the Doctoral and Postdoctoral Institutions of Faculty Members at the Top-Fifty Ranked Chemistry Departments. J. Chem. Educ. 2004, 81, 356, DOI: 10.1021/ed081p356Google Scholar9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXhs1Cnurc%253D&md5=8a98162a965d871a7cb5eec68652b352Analysis by gender of the doctoral and postdoctoral institutions of faculty members at the top-fifty ranked chemistry departmentsKuck, Valerie J.; Marzabadi, Cecilia H.; Nolan, Susan A.; Buckner, Janine P.Journal of Chemical Education (2004), 81 (3), 356-363CODEN: JCEDA8; ISSN:0021-9584. (Journal of Chemical Education, Dept. of Chemistry)There is no expanded citation for this reference.
- 10Miller, W. H.; Hernandez, R.; Moore, C. B.; Polik, W. F. A transition state theory-based statistical distribution of unimolecular decay rates, with application to unimolecular decomposition of formaldehyde. J. Chem. Phys. 1990, 93, 5657– 5666, DOI: 10.1063/1.459636Google Scholar10https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3cXmtlyrsb4%253D&md5=0f49d9a550c645d84e5f668cf1568accA transition state theory-based statistical distribution of unimolecular decay rates with application to unimolecular decomposition of formaldehydeMiller, William H.; Hernandez, Rigoberto; Moore, C. Bradley; Polik, William F.Journal of Chemical Physics (1990), 93 (8), 5657-66CODEN: JCPSA6; ISSN:0021-9606.A statistical distribution of state-specific unimol. decay rates is derived (within the framework of random matrix theory) that is detd. completely by the transition state properties of the potential energy surface. It includes the std. χ-square distributions as a special case. Model calcns. are presented to show the extent to which it can differ from the χ-square distribution, and specific application is made to the state-specific unimol. decay rate data for D2CO → D2 + CO.
- 11Hernandez, R.; Miller, W. H.; Moore, C. B.; Polik, W. F. A Random Matrix/Transition State Theory for the Probability Distribution of State-Specific Unimolecular Decay Rates: Generalization to Include Total Angular Momentum Conservation and Other Dynamical Symmetries. J. Chem. Phys. 1993, 99, 950– 962, DOI: 10.1063/1.465360Google Scholar11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXmtlCqurw%253D&md5=6dba8c44383d72510019dd7f62290cacA random matrix/transition state theory for the probability distribution of state-specific unimolecular decay rates: generalization to include total angular momentum conservation and other dynamical symmetriesHernandez, Rigoberto; Miller, William H.; Moore, C. Bradley; Polik, William F.Journal of Chemical Physics (1993), 99 (2), 950-62CODEN: JCPSA6; ISSN:0021-9606.A previously developed random matrix/transition state theory (RM/TST) model for the probability distribution of state-sp. unimol. decay rates was generalized to incorporate total angular momentum conservation and other dynamic symmetries. The model is made into a predictive theory by using a semiclassical method to det. the transmission probabilities of a nonseparable rovibrational Hamiltonian at the transition state. The overall theory gives a good description of the state-sp. rates for the D2CO → D2 + CO unimol. decay; in particular, it describes the dependence of the distribution of rates on total angular momentum J. Comparison of the exptl. values with results of the RM/TST theory suggests that there is mixing among the rovibrational states.
- 12Miller, W. H. Autobiographical Sketch of William Hughes Miller. J. Phys. Chem. A 2001, 105, 2487– 2489, DOI: 10.1021/jp0101920Google ScholarThere is no corresponding record for this reference.
- 13Miller, W. H. Semi-classical theory for non-separable systems:. Construction of “good” action-angle variables for reaction rate constants. Faraday Discuss. Chem. Soc. 1977, 62, 40, DOI: 10.1039/DC9776200040Google Scholar13https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE1cXhsF2isrs%253D&md5=238838036ad290e52071a98b617f4aacSemiclassical theory for non-separable systems: construction of good action-angle variables for reaction rate constantsMiller, William H.Faraday Discussions of the Chemical Society (1977), 62 (Potential Energy Surf.), 40-6CODEN: FDCSB7; ISSN:0301-7249.A semiclassical expression for bimol. rate consts. for reactions which have a single activation barrier is obtained in terms of the good action variables of the classical Hamiltonian which are assocd. with the saddle point region of the potential energy surface. The formulas apply to nonseparable and separable saddle points.
- 14Miller, W. H.; Hernandez, R.; Handy, N. C.; Jayatilaka, D.; Willetts, A. Ab initio calculation of anharmonic constants for a transition state, with application to semiclassical transition state tunneling probabilities. Chem. Phys. Lett. 1990, 172, 62– 68, DOI: 10.1016/0009-2614(90)87217-FGoogle Scholar14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3cXlslyltLg%253D&md5=9933893dff69dead50201e0165060c7cAb initio calculation of anharmonic constants for a transition state, with application to semiclassical transition state tunneling probabilitiesMiller, William H.; Hernandez, Rigoberto; Handy, Nicholas C.; Jayatilaka, Dylan; Willetts, AndrewChemical Physics Letters (1990), 172 (1), 62-8CODEN: CHPLBC; ISSN:0009-2614."Good" (i.e. conserved) action variables exist in the vicinity of a saddle point (i.e. transition state) of a potential energy surface in complete analogy to those related to a min. on the surface. Transition state theory tunneling (or transmission) probabilities can be expressed semiclassically in terms of these "good" action variables, including the effects of non-separable coupling of all degrees of freedom with each other. This paper shows how ab initio quantum chem. methods recently developed for calcg. anharmonic consts. about a potential min. (i.e. for ordinary vibrational energy levels) can be readily adapted to obtain those related to a transition state, thus providing a rigorous and practical way to apply this non-separable transition state theory. Application is made to the transition state for the reaction D2CO → D2 + CO.
- 15Cohen, M. J.; Handy, N. C.; Hernandez, R.; Miller, W. H. Cumulative reaction probabilities for H + H2 → H2 + H from a knowledge of the anharmonic force field. Chem. Phys. Lett. 1992, 192, 407– 416, DOI: 10.1016/0009-2614(92)85491-RGoogle Scholar15https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK38XktFCltrk%253D&md5=941c2660fb9992c2485483bb7b07d700Cumulative reaction probabilities for atomic hydrogen + molecular hydrogen → molecular hydrogen + atomic hydrogen from a knowledge of the anharmonic force fieldCohen, Michael J.; Handy, Nicholas C.; Hernandez, Rigoberto; Miller, William H.Chemical Physics Letters (1992), 192 (4), 407-16CODEN: CHPLBC; ISSN:0009-2614.In an earlier publication, the authors showed how knowledge of a quartic force field expanded about a transition state can be used to obtain transition-state-theory tunneling probabilities. Thus coupling between the reaction mode and other modes is included in this second-order perturbation theory approach. The authors study the very anharmonic reaction H+H2→H2+H and show that even in this extreme case, there is reasonable agreement between the cumulative reaction probabilities calcd. by this semiclassical approach, and full quantum calcns.
- 16Hernandez, R.; Miller, W. H. Semiclassical Transition State Theory. A New Perspective. Chem. Phys. Lett. 1993, 214, 129– 136, DOI: 10.1016/0009-2614(93)90071-8Google Scholar16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2cXhs1Cqtg%253D%253D&md5=bd9ec72e0e1175247eb467b86ff67fc9Semiclassical transition state theory. A new perspectiveHernandez, Rigoberto; Miller, William H.Chemical Physics Letters (1993), 214 (2), 129-36CODEN: CHPLBC; ISSN:0009-2614.The semiclassical transition state theory (SCTST) introduced by W. H. Miller et al. (1990) requires the inversion of an (effectively integrable) Hamiltonian with respect to the action of the reactive coordinate. The inversion may be avoided in computing the thermal rate const.; the resulting expression provides an appealing link to conventional transition state theory. This reformulation of the SCTST rate is illustrated by application to the bimol. reaction, H + H2 → H2 + H, and to the unimol. dissocn., D2CO → D2 + CO.
- 17Hernandez, R. A Combined Use of Perturbation Theory and Diagonalization: Application to Bound Energy Levels and Semiclassical Rate Theory. J. Chem. Phys. 1994, 101, 9534– 9547, DOI: 10.1063/1.467985Google Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXisFChtLc%253D&md5=b157f8eb49cf25239abdcc314adfba62A combined use of perturbation theory and diagonalization: application to bound energy levels and semiclassical rate theoryHernandez, RigobertoJournal of Chemical Physics (1994), 101 (11), 9534-47CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)A new method, mixed diagonalization, is introduced in which an effective Hamiltonian operator acting on a reduced dimensional space is constructed using the similarity transformations of canonical Van Vleck perturbation theory (CVPT). This construction requires the characterization of modes into two categories, global and local, which in the bound vibrational problem are tantamount to the large and small amplitude vibrations, resp. The local modes in the Hamiltonian are projected out by CVPT, and the resulting Hamiltonian operator acts only on the space of global modes. The method affords the treatment of energy levels of bound systems in which some vibrational assignments are possible. It systematically provides a reduced dimensional Hamiltonian which is more amenable to exact numerical soln. that the original full-dimensional Hamiltonian. In recent work, a semiclassical transition state theory (SCTST) rate expression was written in terms of a Hamiltonian operator parameterized by the imaginary action along the local reaction path in the transition state region [Chem. Phys. Lett. 214, 129(1993)]. The Hamiltonian constructed by mixed diagonalization has this form, and can be used to obtain more accurate semiclassical rate expressions.
- 18Nguyen, T. L.; Stanton, J. F.; Barker, J. R. Ab Initio Reaction Rate Constants Computed Using Semiclassical Transition-State Theory: HO + H2 → H2O + H and Isotopologues. J. Phys. Chem. A 2011, 115, 5118– 5126, DOI: 10.1021/jp2022743Google Scholar18https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXlsFyrsro%253D&md5=bde4a2106d3e067acc5c90549fd685caAb Initio Reaction Rate Constants Computed Using Semiclassical Transition-State Theory: HO + H2 → H2O + H and IsotopologuesNguyen, Thanh Lam; Stanton, John F.; Barker, John R.Journal of Physical Chemistry A (2011), 115 (20), 5118-5126CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)A new algorithm for the semiclassical transition-state theory (SCTST) formulated by W.H. Miller and co-workers is used to compute rate consts. for the isotopologues of the title reaction, with no empirical adjustments. The SCTST and relevant results from second-order vibrational perturbation theory (VPT2) are summarized. VPT2 is used at the CCSD(T) level of electronic structure theory to compute the anharmonicities of the fully coupled vibrational modes (including the reaction coordinate) of the transition structure. The anharmonicities are used in SCTST to compute the rate consts. over the temp. range from 200 to 2500 K. The computed rate consts. are compared to exptl. data and theor. calcns. from the literature. The SCTST results for abs. rate consts. and for both primary and secondary isotope effects are in excellent agreement with the exptl. data for this reaction over the entire temp. range. The sensitivity of SCTST to various parameters is investigated by using a set of simplified models. The results show that multidimensional tunneling along the curved reaction path is important at low temps. and the anharmonic coupling among the vibrational modes is important at high temps. The theor. kinetics data are also presented as fitted empirical algebraic expressions.
- 19Barker, J. R.; Nguyen, T. L.; Stanton, J. F. Kinetic Isotope Effects for Cl + CH4 → HCl + CH3 Calculated Using ab Initio Semiclassical Transition State Theory. J. Phys. Chem. A 2012, 116, 6408– 6419, DOI: 10.1021/jp212383uGoogle Scholar19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhsVyksbY%253D&md5=9c2718ad57590e2e7d28f1ef1b412f97Kinetic Isotope Effects for Cl + CH4 .dblharw. HCl + CH3 Calculated Using ab Initio Semiclassical Transition State TheoryBarker, John R.; Nguyen, Thanh Lam; Stanton, John F.Journal of Physical Chemistry A (2012), 116 (24), 6408-6419CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Calcns. were carried out for 25 isotopologues of the title reaction for various combinations of 35Cl, 37Cl, 12C, 13C, 14C, H, and D. The computed rate consts. are based on harmonic vibrational frequencies calcd. at the CCSD(T)/aug-cc-pVTZ level of theory and Xij vibrational anharmonicity coeffs. calcd. at the CCSD(T) /aug-cc-pVDZ level of theory. For some reactions, anharmonicity coeffs. were also computed at the CCSD(T)/aug-cc-pVTZ level of theory. The classical reaction barrier was taken from Eskola et al. [J. Phys. Chem. A 2008, 112, 7391-7401], who extrapolated CCSD(T) calcns. to the complete basis set limit. Rate consts. were calcd. for temps. from ∼100 to ∼2000 K. The computed ab initio rate const. for the normal isotopologue is in good agreement with expts. over the entire temp. range (∼10% lower than the recommended exptl. value at 298 K). The ab initio H/D kinetic isotope effects (KIEs) for CH3D, CH2D2, CHD3, and CD4 are in very good agreement with literature exptl. data. The ab initio 12C/13C KIE is in error by ∼2% at 298 K for calcns. using Xij coeffs. computed with the aug-cc-pVDZ basis set, but the error is reduced to ∼1% when Xij coeffs. computed with the larger aug-cc-pVTZ basis set are used. Systematic improvements appear to be possible. The present SCTST results are found to be more accurate than those from other theor. calcns. Overall, this is a very promising method for computing ab initio kinetic isotope effects.
- 20Nguyen, T. L.; Stanton, J. F. Ab Initio Thermal Rate Calculations of HO + HO → O(3P) + H2O Reaction and Isotopologues. J. Phys. Chem. A 2013, 117, 2678– 2686, DOI: 10.1021/jp312246qGoogle Scholar20https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXjs1ynsL0%253D&md5=34768f5c4b7a916b8b7e87e664bb8cbbAb Initio Thermal Rate Calculations of HO + HO = O(3P) + H2O Reaction and IsotopologuesNguyen, Thanh Lam; Stanton, John F.Journal of Physical Chemistry A (2013), 117 (13), 2678-2686CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The forward and reverse reactions, HO + HO .dblharw. O(3P) + H2O, which play roles in both combustion and lab. studies, were theor. characterized with a master equation approach to compute thermal reaction rate consts. at both the low and high pressure limits. Our ab initio k(T) results for the title reaction and two isotopic variants agree very well with expts. (within 15%) over a wide temp. range. The calcd. reaction rate shows a distinctly non-Arrhenius behavior and a strong curvature consistent with the expt. This characteristic behavior is due to effects of pos. barrier height and quantum mech. tunneling. Tunneling is very important and contributes more than 70% of total reaction rate at room temp. A prereactive complex is also important in the overall reaction scheme.
- 21Komatsuzaki, T.; Berry, R. S. Regularity in chaotic reaction paths. I. Ar6. J. Chem. Phys. 1999, 110, 9160– 9173, DOI: 10.1063/1.478838Google Scholar21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXis1Sis7Y%253D&md5=1550a27d59b9b8af2073cc8e85bfd2ffRegularity in chaotic reaction paths. I. Ar6Komatsuzaki, Tamiki; Berry, R. StephenJournal of Chemical Physics (1999), 110 (18), 9160-9173CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)We scrutinize the saddle crossings of a simple cluster of six atoms to show (a) that it is possible to choose a coordinate system in which the transmission coeff. for the classical reaction path is unity at all energies up to a moderately high energy, above which the transition state is chaotic; (b) that at energies just more than sufficient to allow passage across the saddle, all or almost all the degrees of freedom of the system are essentially regular in the region of the transition state; and (c) that the degree of freedom assocd. with the reaction coordinate remains essentially regular through the region of the transition state, even to moderately high energies. Microcanonical mol. dynamics simulation of Ar6 bound by pairwise Lennard-Jones potentials reveals the mechanics of passage. We use Lie canonical perturbation theory to construct the nonlinear transformation to a hyperbolic coordinate system which reveals these regularities. This transform "rotates away" the recrossings and nonregular behavior, esp. of the motion along the reaction coordinate, leaving a coordinate and a corresponding dividing surface in phase space which minimize recrossings and mode-mode mixing in the transition state region. The action assocd. with the reactive mode tends to be an approx. invariant of motion through the saddle crossings throughout a relatively wide range of energy. Only at very low energies just above the saddle could any other approx. invariants of motion be found for the other, nonreactive modes. No such local invariants appeared at energies at which the modes are all chaotic and coupled to one another.
- 22Komatsuzaki, T.; Berry, R. S. Regularity in chaotic reactions paths II: Ar6. Energy dependence and visualization of the reaction bottleneck. Phys. Chem. Chem. Phys. 1999, 1, 1387– 1397, DOI: 10.1039/a809424aGoogle Scholar22https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXhslGgs7g%253D&md5=5ff61674899be175158a96ca9ec2c56aRegularity in chaotic reaction paths II: Ar6. Energy dependence and visualization of the reaction bottleneckKomatsuzaki, Tamiki; Stephen Berry, R.Physical Chemistry Chemical Physics (1999), 1 (6), 1387-1397CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)Reaction trajectories reveal regular behavior in the reactive degree of freedom and unit transmission coeffs. as the system crosses the saddle region sepg. reactants and products. The regularity persists up to moderately high energies, even when all other degrees of freedom are chaotic. This behavior is apparent in a representation obtained by transformation with Lie canonical perturbation theory. The dividing surface in this representation is analogous to the conventional dividing surface in the sense that it is the point set for which the reaction coordinate has the const. value it has at the saddle-point singularity. However the nonlinear, full phase-space character of the transformation makes the new crossing surface a complicated, abstr. object whose interpretation and visualization, the objective of this paper, can be realized by cataloging the recrossings as they disappear in successively higher orders of perturbation, and by projection into spaces of only a few dimensions. The result is a conceptual interpretation of how regular behavior persists in a reactive degree of freedom.
- 23Komatsuzaki, T.; Berry, R. S. Local regularity and non-recrossing path in transition state—a new strategy in chemical reaction theories. J. Mol. Struct.: THEOCHEM 2000, 506, 55– 70, DOI: 10.1016/S0166-1280(00)00402-4Google Scholar23https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXksFahsLs%253D&md5=bb91e969b136a0772a2798a29ce9808fLocal regularity and non-recrossing path in transition state - a new strategy in chemical reaction theoriesKomatsuzaki, T.; Berry, R. S.Journal of Molecular Structure: THEOCHEM (2000), 506 (), 55-70CODEN: THEODJ; ISSN:0166-1280. (Elsevier Science B.V.)We analyze local regularities in the regions of transition states of a 6-atom Lennard-Jones cluster to demonstrate how one can choose a non-recrossing reaction path in phase space along which the transmission coeff. for the classical reaction path is unity from threshold up to a moderately high energy, above which the transition state is chaotic, and how one can picture the nonlinear, full-phase-space character of the dividing hypersurface by projecting it into spaces of only a few dimensions. These overcome one of the long-standing ambiguities in chem. reaction theories, the recrossing problem, up to moderately high energies, and make transition state theory more generalized, applicable even in cases in which apparent recrossings spoil the conventional theory.
- 24Komatsuzaki, T.; Berry, R. S. Regularity in chaotic reaction paths III: Ar6 local invariances at the reaction bottleneck. J. Chem. Phys. 2001, 115, 4105– 4117, DOI: 10.1063/1.1385152Google Scholar24https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXmt1ajsLw%253D&md5=52c4cadf179b5d0bbe341a10e61552d8Regularity in chaotic reaction paths III: Ar6 local invariances at the reaction bottleneckKomatsuzaki, Tamiki; Berry, R. StephenJournal of Chemical Physics (2001), 115 (9), 4105-4117CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)We recently developed a new method to ext. a many-body phase-space dividing surface, across which the transmission coeff. for the classical reaction path is unity. The example of isomerization of a 6-atom Lennard-Jones cluster showed that the action assocd. with the reaction coordinate is an approx. invariant of motion through the saddle regions, even at moderately high energies, at which most or all the other modes are chaotic [J. Chem. Phys. 105, 10838 (1999); Phys. Chem. Chem. Phys. 1, 1387 (1999)]. In the present article, we propose a new algorithm to analyze local invariances about the transition state of N-particle Hamiltonian systems. The approx. invariants of motion assocd. with a reaction coordinate in phase space densely distribute in the sea of chaotic modes in the region of the transition state. Using projections of distributions in only two principal coordinates, one can grasp and visualize the stable and unstable invariant manifolds to and from a hyperbolic point of a many-body nonlinear system, like those of the one-dimensional, integrable pendulum. This, in turn, reveals a new type of phase space bottleneck in the region of a transition state that emerges as the total energy increases, which may trap a reacting system in that region.
- 25Komatsuzaki, T.; Berry, R. S. Dynamical hierarchy in transition states: Why and how does a system climb over the mountain?. Proc. Natl. Acad. Sci. U. S. A. 2001, 98, 7666– 7671, DOI: 10.1073/pnas.131627698Google Scholar25https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXlt1Kns7o%253D&md5=04d954cf2d5bd6bd942c1664d8e7356aDynamical hierarchy in transition states: why and how does a system climb over the mountain?Komatsuzaki, Tamiki; Berry, R. StephenProceedings of the National Academy of Sciences of the United States of America (2001), 98 (14), 7666-7671CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)How a reacting system climbs through a transition state during a reaction was an intriguing subject for decades. Here the authors present and quantify a technique to identify and characterize local invariance about the transition state of an N-particle Hamiltonian system, using Lie canonical perturbation theory combined with microcanonical mol. dynamics simulation. At least three distinct energy regimes of dynamical behavior occur in the region of the transition state, distinguished by the extent of their local dynamical invariance and regularity. Isomerization of a six-atom Lennard-Jones cluster illustrates this: up to energies high enough to make the system manifestly chaotic, approx. invariants of motion assocd. with a reaction coordinate in phase space imply a many-body dividing hypersurface in phase space that is free of recrossings even in a sea of chaos. The method makes it possible to visualize the stable and unstable invariant manifolds leading to and from the transition state, i.e., the reaction path in phase space, and how this regularity turns to chaos with increasing total energy of the system. This, in turn, illuminates a new type of phase space bottleneck in the region of a transition state that emerges as the total energy and mode coupling increase, which keeps a reacting system increasingly trapped in that region.
- 26Pollak, E.; Pechukas, P. Unified statistical model for “complex” and “direct” reaction mechanisms: A test on the collinear H + H2 exchange reaction. J. Chem. Phys. 1979, 70, 325– 333, DOI: 10.1063/1.437194Google Scholar26https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE1MXhtFSlsbo%253D&md5=068124475d995a1a85f774e33349c13aUnified statistical model for "complex" and "direct" reaction mechanisms: a test on the collinear atomic hydrogen + molecular hydrogen exchange reactionPollak, Eli; Pechukad, PhilipJournal of Chemical Physics (1979), 70 (1), 325-33CODEN: JCPSA6; ISSN:0021-9606.Miller's unified statistical theory for bimol. chem. reactions is tested on the collinear H + H2 exchange reaction, treated classically. The reaction probability calcd. from unified statistical theory is more accurate than that calcd. from ordinary transition state theory or from variational transition state theory; in particular, unified statistical theory predicts the high-energy falloff of the reaction probability, which transition state theory does not. A derivation of unified statistical theory emphasizes the dynamical and statistical assumptions that are the foundation of the theory. These assumptions unambiguously define the "collision complex" in unified statistical theory and are tested in detail on the H + H2 reaction. Finally, a lower bound on the reaction probability is derived; this bound complements the upper bound provided by transition state theory and is significantly more accurate, for the H + H2 reaction, than either transition state theory or unified statistical theory.
- 27Pollak, E. In Theory of Chemical Reaction Dynamics; Baer, M., Ed.; CRC Press: Boca Raton, FL, 1985; Vol. 3; p 123.Google ScholarThere is no corresponding record for this reference.
- 28Grobgeld, D.; Pollak, E.; Zakrzewski, J. A Numerical Method for Locating Stable Periodic Orbits in Chaotic Systems. Phys. D 1992, 56, 368– 380, DOI: 10.1016/0167-2789(92)90176-NGoogle ScholarThere is no corresponding record for this reference.
- 29Moix, J. M.; Hernandez, R.; Pollak, E. Momentum and velocity autocorrelation functions of a diatomic molecule are not necessarily proportional to each other. J. Phys. Chem. B 2008, 112, 213– 218, DOI: 10.1021/jp0730951Google Scholar29https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXhtVSjsbnP&md5=ff8195a43609d8c18e227241cc6a7d00Momentum and Velocity Autocorrelation Functions of a Diatomic Molecule Are Not Necessarily Proportional to Each OtherMoix, Jeremy M.; Hernandez, Rigoberto; Pollak, EliJournal of Physical Chemistry B (2008), 112 (2), 213-218CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)We present a computation of the classical momentum and velocity correlation functions of Br2 considered as an idealized mol. wire connecting dissipated lead atoms at each end of the dimer. It is demonstrated that coupling of the diat. relative momentum to the leads may result in momenta that are not equal to the mass-weighted velocity. These differences show up in numerical simulations of both the av. value and time correlations of the bond momentum and velocity. These observations are supported by anal. predictions for the av. temp. of the diat. They imply that the "std. recipes" for modeling the system with a generalized Langevin equation are insufficient.
- 30Hernandez, R.; Cao, J.; Voth, G. A. On the Feynman path centroid density as a phase space distribution in quantum statistical mechanics. J. Chem. Phys. 1995, 103, 5018– 5026, DOI: 10.1063/1.470588Google Scholar30https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXotFams78%253D&md5=e2a70c3406eeb3fdd44a51a21704c803On the Feynman path centroid density as a phase space distribution in quantum statistical mechanicsHernandez, Rigoberto; Cao, Jianshu; Voth, Gregory A.Journal of Chemical Physics (1995), 103 (12), 5018-26CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The phase space formulation of quantum statistical mechanics using the Feynman path centroid d. offers an alternative perspective to the std. Wigner prescription for the classical-like evaluation of equil. and/or dynamical quantities of statistical systems. The use of this formulation has been implicit in recent work on quantum rate theories, for example, in which the centroid d. distribution replaces the classical Boltzmann distribution. In order to further understand the approxns. involved in this and similar transcriptions, the present work elaborates and clarifies the issue of operator ordering in a rigorous centroid-based formulation. In particular, through the use of the Weyl correspondence, a precise definition of the centroid symbol of operators and their products is presented. Though we fall short of finding the algebraic structure tantamount to that found in the Weyl symbols-of which the Wigner distribution is an example- the resulting expressions have internal consistency and are amenable to approx. evaluation through cumulant expansions.
- 31Hernandez, R.; Voth, G. A. Quantum time correlation functions and classical coherence. Chem. Phys. 1998, 233, 243– 255, DOI: 10.1016/S0301-0104(98)00027-5Google Scholar31https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXkvVegt7s%253D&md5=26cc85ce6144fb34185db68125d0c9b5Quantum time correlation functions and classical coherenceHernandez, Rigoberto; Voth, Gregory A.Chemical Physics (1998), 233 (2,3), 243-255CODEN: CMPHC2; ISSN:0301-0104. (Elsevier Science B.V.)Quantum time correlation functions for electronically adiabatic and nonadiabatic processes have recently been evaluated directly using the semiclassical initial value representation [J. Cao and G.A. Voth, J. Chem. Phys. 104 (1996) 271]. In this paper, the approach to the classical limit of this theory is analyzed and the nature of coherence in such limits is explored.
- 32Verashchagina, A.; Bettio, F. Gender segregation in the labour market; Directorate-General for Employment, Social Affairs and Inclusion (European Commission), 2009.Google ScholarThere is no corresponding record for this reference.
- 33Hernandez, R.; Somer, F. L. Stochastic dynamics in irreversible nonequilibrium environments. 1. The fluctuation-dissipation relation. J. Phys. Chem. B 1999, 103, 1064– 1069, DOI: 10.1021/jp983625gGoogle Scholar33https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXntlahtQ%253D%253D&md5=d45c7c07622b0487a5507b9b5171af00Stochastic Dynamics in Irreversible Nonequilibrium Environments. 1. The Fluctuation-Dissipation RelationHernandez, Rigoberto; Somer, Frank L.,Jr.Journal of Physical Chemistry B (1999), 103 (7), 1064-1069CODEN: JPCBFK; ISSN:1089-5647. (American Chemical Society)A generalization of the generalized Langevin equation (stochastic dynamics) is introduced in order to model chem. reactions which take place in changing environments. The friction kernel representing the solvent response is given a nonstationary form with respect to which the instantaneous random solvent force satisfies a natural generalization of the fluctuation-dissipation relation. Theor. considerations, as well as numerical simulations, show that the dynamics of this construction satisfy the equipartition theorem beyond its equil. limits.
- 34Hernandez, R.; Somer, F. L. Stochastic dynamics in irreversible nonequilibrium environments. 2. A model for thermosetting polymerization. J. Phys. Chem. B 1999, 103, 1070– 1077, DOI: 10.1021/jp9836269Google Scholar34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXntlahug%253D%253D&md5=3474745581fe4d6fa7cd3014b0842d2dStochastic Dynamics in Irreversible Nonequilibrium Environments. 2. A Model for Thermosetting PolymerizationHernandez, Rigoberto; Somer, Frank L. Jr.Journal of Physical Chemistry B (1999), 103 (7), 1070-1077CODEN: JPCBFK; ISSN:1089-5647. (American Chemical Society)A generalization of the generalized Langevin equation (GLE), the so-called irreversible GLE (iGLE) [Hernandez, R.; Somer, F. L. J. Phys. Chem. 1999, 103, XXXX], is further extended to describe non-stationary environments in which the non-stationarity is induced by the macroscopic behavior of the ensemble itself, rather than an external force. Such a formalism lends itself to the dynamical study of the length distributions of growing polymers.
- 35Hernandez, R. The projection of a mechanical system onto the irreversible generalized Langevin equation (iGLE). J. Chem. Phys. 1999, 111, 7701– 7704, DOI: 10.1063/1.480160Google Scholar35https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXms1eqt74%253D&md5=8e9c710fe60dbe955bebf0d365c768a3The projection of a mechanical system onto the irreversible generalized Langevin equationHernandez, RigobertoJournal of Chemical Physics (1999), 111 (17), 7701-7704CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The so-called irreversible generalized Langevin equation [R. Hernandez and F. L. Somer, J. Phys. Chem. B 103, 1064 (1999)], which extends the generalized Langevin equation (stochastic dynamics) to include irreversible changes (nonstationarity) in the solvent response, is shown to be the projection of an explicit time-dependent Hamiltonian system.
- 36Locker, C. R.; Hernandez, R. A minimalist model protein with multiple folding funnels. Proc. Natl. Acad. Sci. U. S. A. 2001, 98, 9074– 9079, DOI: 10.1073/pnas.161438898Google Scholar36https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXlvFSkur8%253D&md5=f03eccf584b758291b24d5ebb39f76aaA minimalist model protein with multiple folding funnelsLocker, C. Rebecca; Hernandez, RigobertoProceedings of the National Academy of Sciences of the United States of America (2001), 98 (16), 9074-9079CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)Kinetic and structural studies of wild-type proteins such as prions and amyloidogenic proteins provide suggestive evidence that proteins may adopt multiple long-lived states in addn. to the native state. All of these states differ structurally because they lie far apart in configuration space, but their stability is not necessarily caused by cooperative (nucleation) effects. In this study, a minimalist model protein is designed to exhibit multiple long-lived states to explore the dynamics of the corresponding wild-type proteins. The minimalist protein is modeled as a 27-monomer sequence confined to a cubic lattice with three different monomer types. An order parameter-the winding index-is introduced to characterize the extent of folding. The winding index has several advantages over other commonly used order parameters like the no. of native contacts. It can distinguish between enantiomers, its calcn. requires less computational time than the no. of native contacts, and reduced-dimensional landscapes can be developed when the native state structure is not known a priori. The results for the designed model protein prove by existence that the rugged energy landscape picture of protein folding can be generalized to include protein "misfolding" into long-lived states.
- 37Servos, J. W. Physical Chemistry from Ostwald to Pauling: The Making of a Science in America; Princeton University Press, 1990.Google ScholarThere is no corresponding record for this reference.
- 38Hernandez, R. 6. Blurring of Physical Chemistry and Chemical Physics. 2013; http://everywherechemistry.blogspot.com/2013/08/6-blurring-of-physical-chemistry-and.html, accessed January 15, 2021.Google ScholarThere is no corresponding record for this reference.
- 39Schatz, G. C. Celebrating Our 120th Anniversary. J. Phys. Chem. A 2016, 120, 9679– 9681, DOI: 10.1021/acs.jpca.6b10935Google Scholar39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XitVenurnI&md5=22a9dc1151a1bc0877aaae145060ff55Celebrating Our 120th AnniversarySchatz, George C.Journal of Physical Chemistry A (2016), 120 (49), 9679-9681CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)There is no expanded citation for this reference.
- 40Bartsch, T.; Hernandez, R.; Uzer, T. Transition state in a noisy environment. Phys. Rev. Lett. 2005, 95, 058301, DOI: 10.1103/PhysRevLett.95.058301Google Scholar40https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXmvF2gsLc%253D&md5=9e8e781defda79067a698661b4e07907Transition State in a Noisy EnvironmentBartsch, Thomas; Hernandez, Rigoberto; Uzer, T.Physical Review Letters (2005), 95 (5), 058301/1-058301/4CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)Transition state theory overestimates reaction rates in soln. because conventional dividing surfaces between reagents and products are crossed many times by the same reactive trajectory. We describe a recipe for constructing a time-dependent dividing surface free of such recrossings in the presence of noise. The no-recrossing limit of transition state theory thus becomes generally available for the description of reactions in a fluctuating environment.
- 41Shukla, C.; Hallett, J. P.; Popov, A. V.; Hernandez, R.; Liotta, C.; Eckert, C. Molecular Dynamics Simulation of the Cybotactic Region in Gas-Expanded Methanol-Carbon Dioxide and Acetone-Carbon Dioxide Mixtures. J. Phys. Chem. B 2006, 110, 24101– 24111, DOI: 10.1021/jp0648947Google Scholar41https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XhtFCltLvP&md5=d6917e698291b04997a83f17183393abMolecular Dynamics Simulation of the Cybotactic Region in Gas-Expanded Methanol-Carbon Dioxide and Acetone-Carbon Dioxide MixturesShukla, Charu L.; Hallett, Jason P.; Popov, Alexander V.; Hernandez, Rigoberto; Liotta, Charles L.; Eckert, Charles A.Journal of Physical Chemistry B (2006), 110 (47), 24101-24111CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)Local solvation and transport effects in gas-expanded liqs. (GXLs) are reported based on mol. simulation. GXLs were found to exhibit local d. enhancements similar to those seen in supercrit. fluids, although less dramatic. This approach was used as an alternative to a multiphase atomistic model for these mixts. by utlilizing exptl. results to describe the necessary fixed conditions for a locally (quasi-) stable mol. dynamics model of the (single) GXL phase. The local anisotropic pair correlation function, orientational correlation functions, and diffusion rates are reported for two systems: CO2-expanded methanol and CO2-expanded acetone at 298 K and pressures up to 6 MPa.
- 42Hallett, J. P.; Kitchens, C. L.; Hernandez, R.; Liotta, C.; Eckert, C. Probing the Cybotactic Region in Gas-Expanded Liquids (GXLs). Acc. Chem. Res. 2006, 39, 531– 538, DOI: 10.1021/ar0501424Google Scholar42https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XlsFaktr8%253D&md5=070ffed95f106060adbb9a3fb26211bdProbing the Cybotactic Region in Gas-Expanded Liquids (GXLs)Hallett, Jason P.; Kitchens, Christopher L.; Hernandez, Rigoberto; Liotta, Charles L.; Eckert, Charles A.Accounts of Chemical Research (2006), 39 (8), 531-538CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)Gas-expanded liqs. (GXLs) are a new and benign class of liq. solvents, which may offer many advantages for sepns., reactions, and advanced materials. GXLs are intermediate in properties between normal liqs. and supercrit. fluids, both in solvating power and in transport properties. Other advantages include benign nature, low operating pressures, and highly tunable properties by simple pressure variations. The chem. community has only just begun to exploit the advantages of these GXLs for industrial applications. This account focuses on the synergism of exptl. techniques with theor. modeling resulting in a powerful combination for exploring chem. structure and transport in the cybotactic region of GXLs (at the nanometer lengthscale).
- 43Stockard, J.; Rohlfing, C. M.; Richmond, G. L. Equity for women and underrepresented minorities in STEM: Graduate experiences and career plans in chemistry. Proc. Natl. Acad. Sci. U. S. A. 2021, 118, e2020508118, DOI: 10.1073/pnas.2020508118Google Scholar43https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXitFaisL8%253D&md5=ada69c53528b1bf08e5d0de857fb553fEquity for women and underrepresented minorities in STEM: Graduate experiences and career plans in chemistryStockard, Jean; Rohlfing, Celeste M.; Richmond, Geraldine L.Proceedings of the National Academy of Sciences of the United States of America (2021), 118 (4), e2020508118CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)Recent events prompted scientists in the United States and throughout the world to consider how systematic racism affects the scientific enterprise. This paper provides evidence of inequities related to race-ethnicity and gender in graduate school experiences and career plans of PhD students in the top 100 ranked departments in one science, technol., engineering, and math (STEM) discipline, chem. Mixed-model regression analyses were used to examine factors that might moderate these differences. The results show that graduate students who identified as a member of a racial/ethnic group traditionally underrepresented in chem. (underrepresented minorities, URM) were significantly less likely than other students to report that their financial support was sufficient to meet their needs. They were also less likely to report having supportive relationships with peers and postdocs. Women, and esp. URM women, were significantly less likely to report supportive relationships with advisors. Despite their more neg. experiences in graduate school, students who identified as URM expressed greater commitment to finishing their degree and staying in the field. When there was at least one faculty member within their departments who also identified as URM they were also more likely than other students to aspire to a university professorship with an emphasis on research. Men were significantly more likely than women to express strong commitment to finishing the PhD and remaining in chem., but this difference was stronger in top-ranked departments. Men were also more likely than women to aspire to a professorship with an emphasis on research, and this difference remained when individual and departmental-level variables were controlled.
- 44Bowman, J. M. Autobiography of Joel M. Bowman. J. Phys. Chem. A 2013, 117, 6907– 6909, DOI: 10.1021/jp405529pGoogle ScholarThere is no corresponding record for this reference.
- 45Tucker, S. C. Solvent Density Inhomogeneities in Supercritical Fluids. Chem. Rev. 1999, 99, 391– 418, DOI: 10.1021/cr9700437Google Scholar45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXjt1Oquw%253D%253D&md5=6a00e344f9f2c3fd7c6d571935a79efaSolvent Density Inhomogeneities in Supercritical FluidsTucker, Susan C.Chemical Reviews (Washington, D. C.) (1999), 99 (2), 391-418CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review, with ∼184 refs., where fundamentals, evidence of local d. inhomogeneities, and addnl. characteristics of local d. inhomogeneities were discussed.
- 46Locker, C. R.; Hernandez, R. Folding behavior of model proteins with weak energetic frustration. J. Chem. Phys. 2004, 120, 11292– 11303, DOI: 10.1063/1.1751394Google Scholar46https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXksVCqurc%253D&md5=9f12cdcc8c3ef9f3115a3e21f1e224a6Folding behavior of model proteins with weak energetic frustrationLocker, C. Rebecca; Hernandez, RigobertoJournal of Chemical Physics (2004), 120 (23), 11292-11303CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The native structure of fast-folding proteins, albeit a deep local free-energy min., may involve a relatively small energetic penalty due to nonoptimal, though favorable, contacts between amino acid residues. The weak energetic frustration that such contacts represent varies among different proteins and may account for folding behavior not seen in unfrustrated models. Minimalist model proteins with heterogeneous contacts-as represented by lattice heteropolymers consisting of three types of monomers-also give rise to weak energetic frustration in their corresponding native structures, and the present study of their equil. and nonequil. properties reveals some of the breadth in their behavior. In order to capture this range within a detailed study of only a few proteins, four candidate protein structures (with their cognate sequences) have been selected according to a figure of merit called the winding index-a characteristic of the no. of turns the protein winds about an axis. The temp.-dependent heat capacities reveal a high-temp. collapse transition, and an infrequently obsd. low-temp. rearrangement transition that arises because of the presence of weak energetic frustration. Simulation results motivate the definition of a new measure of folding affinity as a sequence-dependent free energy-a function of both a reduced stability gap and high accessibility to non-native structures-that correlates strongly with folding rates.
- 47Leite, V. B. P.; Alonso, L. C. P.; Newton, M.; Wang, J. Single Molecule Electron Transfer Dynamics in Complex Environments. Phys. Rev. Lett. 2005, 95, 118301, DOI: 10.1103/PhysRevLett.95.118301Google Scholar47https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXpvFGktrc%253D&md5=b52f959f5825f207ddea56311f94e8e7Single Molecule Electron Transfer Dynamics in Complex EnvironmentsLeite, Vitor B. P.; Alonso, Luciana C. P.; Newton, Marshall; Wang, JinPhysical Review Letters (2005), 95 (11), 118301/1-118301/4CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)We propose a new theor. approach to study the kinetics of the electron transfer (ET) under the dynamical influence of the complex environments with the first passage times (FPT) of the reaction events. By measuring the mean and high order moments of FPT and their ratios, the full kinetics of ET, esp. the dynamical transitions across different temp. zones, is revealed. The potential applications of the current results to single mol. electron transfer are discussed.
- 48Sension, R.; Tokmakoff, A. Proceedings of “Optical Probes of Dynamics in Complex Environments”; OSTI, 2008; https://www.osti.gov/servlets/purl/1062182, accessed January 15, 2021.Google ScholarThere is no corresponding record for this reference.
- 49Virshup, A. M.; Punwong, C.; Pogorelov, T. V.; Lindquist, B. A.; Ko, C.; Martínez, T. J. Photodynamics in Complex Environments: Ab Initio Multiple Spawning Quantum Mechanical/Molecular Mechanical Dynamics. J. Phys. Chem. B 2009, 113, 3280– 3291, DOI: 10.1021/jp8073464Google Scholar49https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXhsFSqsLvL&md5=e0f0c87e8ccf752578878b4256ee89e3Photodynamics in Complex Environments: Ab Initio Multiple Spawning Quantum Mechanical/Molecular Mechanical DynamicsVirshup, Aaron M.; Punwong, Chutintorn; Pogorelov, Taras V.; Lindquist, Beth A.; Ko, Chaehyuk; Martinez, Todd J.Journal of Physical Chemistry B (2009), 113 (11), 3280-3291CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)Our picture of reactions on electronically excited states has evolved considerably in recent years, due to advances in our understanding of points of degeneracy between different electronic states, termed "conical intersections" (CIs). CIs serve as funnels for population transfer between different electronic states, and play a central role in ultrafast photochem. Because most practical photochem. occurs in soln. and protein environments, it is important to understand the role complex environments play in directing excited-state dynamics generally, as well as specific environmental effects on CI geometries and energies. In order to model such effects, we employ the full multiple spawning (FMS) method for multistate quantum dynamics, together with hybrid quantum mech./mol. mech. (QM/MM) potential energy surfaces using both semiempirical and ab initio QM methods. In this article, we present an overview of these methods, and a comparison of the excited-state dynamics of several biol. chromophores in solvent and protein environments. Aq. solvation increases the rate of quenching to the ground state for both the photoactive yellow protein (PYP) and green fluorescent protein (GFP) chromophores, apparently by energetic stabilization of their resp. CIs. In contrast, solvation in methanol retards the quenching process of the retinal protonated Schiff base (RPSB), the rhodopsin chromophore. Protein environments serve to direct the excited-state dynamics, leading to higher quantum yields and enhanced reaction selectivity.
- 50Hernandez, R.; Popov, A. Molecular dynamics out of equilibrium: Mechanics and measurables. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2014, 4, 541– 561, DOI: 10.1002/wcms.1190Google Scholar50https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhvVelu7jK&md5=83f3cee550944e4259712b398279c52eMolecular dynamics out of equilibrium: mechanics and measurablesHernandez, Rigoberto; Popov, Alexander V.Wiley Interdisciplinary Reviews: Computational Molecular Science (2014), 4 (6), 541-561CODEN: WIRCAH; ISSN:1759-0884. (Wiley-Blackwell)Mol. dynamics is fundamentally the integration of the equations of motion over a representation of an at. and mol. system. The most rigorous choice for performing mol. dynamics entails the use of quantum-mech. equations of motion and a representation of the mol. system through all of its electrons and atoms. For most mol. problems involving at least hundreds of atoms, but generally many more, this is simply computationally prohibitive. Thus the art of mol. dynamics lies in choosing the representation and the appropriate equations of motion capable of addressing the requisite measurables. When used adroitly, it can provide both equil. (averaged) and time-dependent properties of a mol. system. Many computational packages now exist that perform mol. dynamics simulations. They generally include force fields to represent the interactions between atoms and mols. (smoothing out electrons through the Born-Oppenheimer approxn.) and integrate the remaining particles classically. Despite these simplifications, all-atom mol. dynamics remains computationally inaccessible if one includes the no. of atoms required to simulate mesoscopic solvents. Here we use anal. models to demonstrate how mol. dynamics can be used to limit the solvent size in systems experiencing either equil. or nonequil. conditions. It is equally important to address the measurables (such as reaction rates) that are to be obtained prior to the generation of the data-intensive trajectories. WIREs Comput Mol Sci 2014, 4:541-561. doi: 10.1002/wcms.1190 For further resources related to this article, please visit the . Conflict of interest: The authors have declared no conflicts of interest for this article.
- 51Press release: Nobel Prize in Chemistry 2013. https://www.nobelprize.org/prizes/chemistry/2013/press-release/, accessed January 15, 2021.Google ScholarThere is no corresponding record for this reference.
- 52MacKerell, A. D., Jr. All-atom empirical potential for molecular modeling and dynamics studies of proteins. J. Phys. Chem. B 1998, 102, 3586– 3616, DOI: 10.1021/jp973084fGoogle Scholar52https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXivVOlsb4%253D&md5=ebb5100dafd0daeee60ca2fa66c1324aAll-Atom Empirical Potential for Molecular Modeling and Dynamics Studies of ProteinsMacKerell, A. D., Jr.; Bashford, D.; Bellott, M.; Dunbrack, R. L.; Evanseck, J. D.; Field, M. J.; Fischer, S.; Gao, J.; Guo, H.; Ha, S.; Joseph-McCarthy, D.; Kuchnir, L.; Kuczera, K.; Lau, F. T. K.; Mattos, C.; Michnick, S.; Ngo, T.; Nguyen, D. T.; Prodhom, B.; Reiher, W. E., III; Roux, B.; Schlenkrich, M.; Smith, J. C.; Stote, R.; Straub, J.; Watanabe, M.; Wiorkiewicz-Kuczera, J.; Yin, D.; Karplus, M.Journal of Physical Chemistry B (1998), 102 (18), 3586-3616CODEN: JPCBFK; ISSN:1089-5647. (American Chemical Society)New protein parameters are reported for the all-atom empirical energy function in the CHARMM program. The parameter evaluation was based on a self-consistent approach designed to achieve a balance between the internal (bonding) and interaction (nonbonding) terms of the force field and among the solvent-solvent, solvent-solute, and solute-solute interactions. Optimization of the internal parameters used exptl. gas-phase geometries, vibrational spectra, and torsional energy surfaces supplemented with ab initio results. The peptide backbone bonding parameters were optimized with respect to data for N-methylacetamide and the alanine dipeptide. The interaction parameters, particularly the at. charges, were detd. by fitting ab initio interaction energies and geometries of complexes between water and model compds. that represented the backbone and the various side chains. In addn., dipole moments, exptl. heats and free energies of vaporization, solvation and sublimation, mol. vols., and crystal pressures and structures were used in the optimization. The resulting protein parameters were tested by applying them to noncyclic tripeptide crystals, cyclic peptide crystals, and the proteins crambin, bovine pancreatic trypsin inhibitor, and carbonmonoxy myoglobin in vacuo and in a crystal. A detailed anal. of the relationship between the alanine dipeptide potential energy surface and calcd. protein φ, χ angles was made and used in optimizing the peptide group torsional parameters. The results demonstrate that use of ab initio structural and energetic data by themselves are not sufficient to obtain an adequate backbone representation for peptides and proteins in soln. and in crystals. Extensive comparisons between mol. dynamics simulation and exptl. data for polypeptides and proteins were performed for both structural and dynamic properties. Calcd. data from energy minimization and dynamics simulations for crystals demonstrate that the latter are needed to obtain meaningful comparisons with exptl. crystal structures. The presented parameters, in combination with the previously published CHARMM all-atom parameters for nucleic acids and lipids, provide a consistent set for condensed-phase simulations of a wide variety of mols. of biol. interest.
- 53Mackerell, A. D., Jr.; Feig, M.; Brooks, C. L., III Extending the treatment of backbone energetics in protein force fields: limitations of gas-phase quantum mechanics in reproducing protein conformational distributions in molecular dynamics simulations. J. Comput. Chem. 2004, 25, 1400– 1415, DOI: 10.1002/jcc.20065Google Scholar53https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXlsVOgt7c%253D&md5=b2451bb5df548447f8b172a211bc1848Extending the treatment of backbone energetics in protein force fields: Limitations of gas-phase quantum mechanics in reproducing protein conformational distributions in molecular dynamics simulationsMacKerell, Alexander D., Jr.; Feig, Michael; Brooks, Charles L., IIIJournal of Computational Chemistry (2004), 25 (11), 1400-1415CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)Computational studies of proteins based on empirical force fields represent a powerful tool to obtain structure-function relationships at an at. level, and are central in current efforts to solve the protein folding problem. The results from studies applying these tools are, however, dependent on the quality of the force fields used. In particular, accurate treatment of the peptide backbone is crucial to achieve representative conformational distributions in simulation studies. To improve the treatment of the peptide backbone, quantum mech. (QM) and mol. mech. (MM) calcns. were undertaken on the alanine, glycine, and proline dipeptides, and the results from these calcns. were combined with mol. dynamics (MD) simulations of proteins in crystal and aq. environments. QM potential energy maps of the alanine and glycine dipeptides at the LMP2/cc-pVxZ/MP2/6-31G* levels, where x = D, T, and Q, were detd., and are compared to available QM studies on these mols. The LMP2/cc pVQZ//MP2/6-31G* energy surfaces for all three dipeptides were then used to improve the MM treatment of the dipeptides. These improvements included addnl. parameter optimization via Monte Carlo simulated annealing and extension of the potential energy function to contain peptide backbone .vphi., ψ dihedral crossterms or a .vphi., ψ grid-based energy correction term. Simultaneously, MD simulations of up to seven proteins in their cryst. environments were used to validate the force field enhancements. Comparison with QM and crystallog. data showed that an addnl. optimization of the .vphi., ψ dihedral parameters along with the grid-based energy correction were required to yield significant improvements over the CHARMM22 force field. However, systematic deviations in the treatment of .vphi. and ψ in the helical and sheet regions were evident. Accordingly, empirical adjustments were made to the grid-based energy correction for alanine and glycine to account for these systematic differences. These adjustments lead to greater deviations from QM data for the two dipeptides but also yielded improved agreement with exptl. crystallog. data. These improvements enhance the quality of the CHARMM force field in treating proteins. This extension of the potential energy function is anticipated to facilitate improved treatment of biol. macromols. via MM approaches in general.
- 54Kalé, L.; Skeel, R.; Bhandarkar, M.; Brunner, R.; Gursoy, A.; Krawetz, N.; Phillips, J.; Shinozaki, A.; Varadarajan, K.; Schulten, K. NAMD2: Greater scalability for parallel molecular dynamics. J. Comput. Phys. 1999, 151, 283– 312, DOI: 10.1006/jcph.1999.6201Google Scholar54https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXivFejt7Y%253D&md5=f40c0fc219c6fef216fae5f0dc8c9003NAMD2: Greater Scalability for Parallel Molecular DynamicsKale, Laxmikant; Skeel, Robert; Bhandarkar, Milind; Brunner, Robert; Gursoy, Attila; Krawetz, Neal; Phillips, James; Shinozaki, Aritomo; Varadarajan, Krishnan; Schulten, KlausJournal of Computational Physics (1999), 151 (1), 283-312CODEN: JCTPAH; ISSN:0021-9991. (Academic Press)Mol. dynamics programs simulate the behavior of biomol. systems, leading to understanding of their functions. However, the computational complexity of such simulations is enormous. Parallel machines provide the potential to meet this computational challenge. To harness this potential, it is necessary to develop a scalable program. It is also necessary that the program be easily modified by application-domain programmers. The NAMD2 program presented in this paper seeks to provide these desirable features. It uses spatial decompn. combined with force decompn. to enhance scalability. It uses intelligent periodic load balancing, so as to maximally utilize the available compute power. It is modularly organized, and implemented using Charm++, a parallel C++ dialect, so as to enhance its modifiability. It uses a combination of numerical techniques and algorithms to ensure that energy drifts are minimized, ensuring accuracy in long running calcns. NAMD2 uses a portable run-time framework called Converse that also supports interoperability among multiple parallel paradigms. As a result, different components of applications can be written in the most appropriate parallel paradigms. NAMD2 runs on most parallel machines including workstation clusters and has yielded speedups in excess of 180 on 220 processors. This paper also describes the performance obtained on some benchmark applications. (c) 1999 Academic Press.
- 55Pearlman, D. A.; Case, D. A.; Caldwell, J.; Ross, W. R.; Cheatham, T. E., III; DeBolt, S.; Ferguson, D.; Seibel, G.; Kollman, P. AMBER, a computer program for applying molecular mechanics, normal mode analysis, molecular dynamics and free energy calculations to elucidate the structures and energies of molecules. Comput. Phys. Commun. 1995, 91, 1– 41, DOI: 10.1016/0010-4655(95)00041-DGoogle Scholar55https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXps1Wrtrw%253D&md5=8dc71939a46bbc17d5da08782d4e6ec8"AMBER", a package of computer programs for applying molecular mechanics, normal mode analysis, molecular dynamics and free energy calculations to stimulate the structural and energetic properties of moleculesPearlman, David A.; Case, David A.; Caldwell, James W.; Ross, Wilson S.; Cheatham, Thomas E. III; DeBolt, Steve; Ferguson, David; Seibel, George; Kollman, PeterComputer Physics Communications (1995), 91 (1-3), 1-42CODEN: CPHCBZ; ISSN:0010-4655. (Elsevier)We describe the development, current features, and some directions for future development of the AMBER package of computer programs. This package has evolved from a program that was constructed to do Assisted Model Building and Energy Refinement to a group of programs embodying a no. of the powerful tools of modern computational chem.-mol. dynamics and free energy calcns.
- 56Brown, W. M.; Petersen, M. K.; Plimpton, S. J.; Grest, G. S. Liquid crystal nanodroplets in solution. J. Chem. Phys. 2009, 130, 044901, DOI: 10.1063/1.3058435Google Scholar56https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhtVWntbg%253D&md5=fd0b4c29e8994e5da4aa8eec53178918Liquid crystal nanodroplets in solutionBrown, W. Michael; Petersen, Matt K.; Plimpton, Steven J.; Grest, Gary S.Journal of Chemical Physics (2009), 130 (4), 044901/1-044901/7CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The aggregation of liq. crystal nanodroplets from a homogeneous soln. is studied by mol. dynamics simulations. The liq. crystal particles are modeled as elongated ellipsoidal Gay-Berne particles while the solvent is modeled as spherical Lennard-Jones particles. Extending previous studies of , we find that liq. crystal nanodroplets are not stable and that after sufficiently long times the nanodroplets always aggregate into a single large droplet. Results describing the droplet shape and orientation for different temps. and shear rates are presented. The implementation of the Gay-Berne potential for biaxial ellipsoidal particles in a parallel mol. dynamics code is also briefly discussed. (c) 2009 American Institute of Physics.
- 57Durrant, J. D.; Kochanek, S. E.; Casalino, L.; Ieong, P. U.; Dommer, A. C.; Amaro, R. E. Mesoscale All-Atom Influenza Virus Simulations Suggest New Substrate Binding Mechanism. ACS Cent. Sci. 2020, 6, 189– 196, DOI: 10.1021/acscentsci.9b01071Google Scholar57https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXjtFKhsL0%253D&md5=84c3498ba325642e2389c06d4c118de5Mesoscale All-Atom Influenza Virus Simulations Suggest New Substrate Binding MechanismDurrant, Jacob D.; Kochanek, Sarah E.; Casalino, Lorenzo; Ieong, Pek U.; Dommer, Abigail C.; Amaro, Rommie E.ACS Central Science (2020), 6 (2), 189-196CODEN: ACSCII; ISSN:2374-7951. (American Chemical Society)Influenza virus circulates in human, avian, and swine hosts, causing seasonal epidemic and occasional pandemic outbreaks. Influenza neuraminidase, a viral surface glycoprotein, has two sialic acid binding sites. The catalytic (primary) site, which also binds inhibitors such as oseltamivir carboxylate, is responsible for cleaving the sialic acid linkages that bind viral progeny to the host cell. In contrast, the functional annotation of the secondary site remains unclear. Here, we better characterize these two sites through the development of an all-atom, explicitly solvated, exptl. based integrative model of the pandemic influenza A H1N1 2009 viral envelope, contg. ∼160 million atoms and spanning ∼115 nm in diam. Mol. dynamics simulations of this crowded subcellular environment, coupled with Markov state model theory, provide a novel framework for studying realistic mol. systems at the mesoscale and allow us to quantify the kinetics of the 150-loop transition between the open and closed states. An anal. of chloride ion occupancy along the neuraminidase surface implies a potential new role for the neuraminidase secondary site, wherein the terminal sialic acid residues of the linkages may bind before transfer to the primary site where enzymic cleavage occurs. Altogether, our work breaks new ground for mol. simulation in terms of the size, complexity, and methodol. analyses of the simulated components, as well as provides fundamental insights into the understanding of substrate recognition processes for this vital influenza drug target, suggesting a new strategy for the development of anti-influenza therapeutics.
- 58Hagy, M. C.; Hernandez, R. Dynamical simulation of dipolar Janus colloids: Dynamical properties. J. Chem. Phys. 2013, 138, 184903, DOI: 10.1063/1.4803864Google Scholar58https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXnsVyrsLw%253D&md5=f4d2faca35479eab47b17f30643d245dDynamical simulation of dipolar Janus colloids: Dynamical propertiesHagy, Matthew C.; Hernandez, RigobertoJournal of Chemical Physics (2013), 138 (18), 184903/1-184903/10CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The dynamical properties of dipolar Janus particles are studied through simulation using our previously-developed detailed pointwise (PW) model and an isotropically coarse-grained (CG) model. The CG model is found to have accelerated dynamics relative to the PW model over a range of conditions for which both models have near identical static equil. properties. Phys., this suggests dipolar Janus particles have slower transport properties (such as diffusion) in comparison to isotropically attractive particles. Time rescaling and damping with Langevin friction are explored to map the dynamics of the CG model to that of the PW model. Both methods map the diffusion const. successfully and improve the velocity autocorrelation function and the mean squared displacement of the CG model. Neither method improves the distribution of reversible bond durations f(tb) obsd. in the CG model, which is found to lack the longer duration reversible bonds obsd. in the PW model. We attribute these differences in f(tb) to changes in the energetics of multiple rearrangement mechanisms. This suggests a need for new methods that map the coarse-grained dynamics of such systems to the true time scale. (c) 2013 American Institute of Physics.
- 59Downey-Mavromatis, A.; Widener, A. Chemistry faculty’s diversity has changed little since 2011. Chem. Eng. News 2020, 98, 22– 25, DOI: 10.1021/cen-09843-feature2Google Scholar59https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXit12ju7fF&md5=c06a7f4abf3ff20ec34c69dcb7cc7575Chemistry faculty's diversity has changed little since 2011Downey-Mavromatis, ArmindaChemical & Engineering News (2020), 98 (43), 22-25CODEN: CENEAR; ISSN:1520-605X. (American Chemical Society)For the past decade, C&EN has worked with the Open Chem. Collaborative in Diversity Equity (OXIDE) to publish statistics on gender and racial and ethnic diversity of US chem. faculty. OXIDE collects the data, a task that's not as easy as it might sound, its leaders say. Nevertheless, it's important work, esp. with the increased focus from universities on campus diversity efforts after protests about racism and police brutality against Black people this year. "This is definitely a moment of reflection for our nation-and one cannot effectively reflect without knowing the facts," says Dontarie Stallings, assoc. director of OXIDE and a teaching professor of chem. at the University of California San Diego. The current facts: the racial and ethnic diversity of US chem. faculty has changed little since 2011. OXIDE began documenting diversity among chem. faculty because although some universities survey their faculty and students to understand their diversity and OXIDE hopes data will help departments improve faculty's diversity, inclusion, and equity 10.47287/cen-09843-feature2-meeting-gr1Rigoberto Hernandez (left) and Dontarie Stallings at an American Chem. Society meeting in 2016 (Credit: Courtesy of Rigoberto Hernandez) oxidesurveyRigoberto HernandezDontarie StallingsA photo of Rigoberto Hernandez and Dontarie Stallings.
- 60Hernandez, R. The private sector’s role in chemistry’s future. Chem. Eng. News 2015, 93, 33, DOI: 10.1021/cen-09337-commentGoogle ScholarThere is no corresponding record for this reference.
- 61Hernandez, R. OneChemistry in the marketplace of ideas. Chem. Eng. News 2017, 95, 41, DOI: 10.1021/cen-09517-commentGoogle ScholarThere is no corresponding record for this reference.
- 62Craven, G. T.; Junginger, A.; Hernandez, R. Lagrangian descriptors of driven chemical reaction manifolds. Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top. 2017, 96, 022222, DOI: 10.1103/PhysRevE.96.022222Google ScholarThere is no corresponding record for this reference.
- 63Schraft, P.; Junginger, A.; Feldmaier, M.; Bardakcioglu, R.; Main, J.; Wunner, G.; Hernandez, R. Neural network approach to time-dependent dividing surfaces in classical reaction dynamics. Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top. 2018, 97, 042309, DOI: 10.1103/PhysRevE.97.042309Google ScholarThere is no corresponding record for this reference.
- 64Feldmaier, M.; Reiff, J.; Benito, R. M.; Borondo, F.; Main, J.; Hernandez, R. Influence of external driving on decays in the geometry of the LiCN isomerization. J. Chem. Phys. 2020, 153, 084115, DOI: 10.1063/5.0015509Google Scholar64https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXhslWqtr7F&md5=8e4c634fd35a0574038dee134be46e23Influence of external driving on decays in the geometry of the LiCN isomerizationFeldmaier, Matthias; Reiff, Johannes; Benito, Rosa M.; Borondo, Florentino; Main, Joerg; Hernandez, RigobertoJournal of Chemical Physics (2020), 153 (8), 084115CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The framework of transition state theory relies on the detn. of a geometric structure identifying reactivity. It replaces the laborious exercise of following many trajectories for a long time to provide chem. reaction rates and pathways. In this paper, recent advances in constructing this geometry even in time-dependent systems are applied to the LiCN .dblharw. LiNC isomerization reaction driven by an external field. We obtain decay rates of the reactant population close to the transition state by exploiting local properties of the dynamics of trajectories in and close to it. We find that the external driving has a large influence on these decay rates when compared to the non-driven isomerization reaction. This, in turn, provides renewed evidence for the possibility of controlling chem. reactions, like this one, through external time-dependent fields. (c) 2020 American Institute of Physics.
- 65Bureau, H.; Quirk, S.; Hernandez, R. The relative stability of trpzip1 and its mutants determined by computation and experiment. RSC Adv. 2020, 10, 6520, DOI: 10.1039/D0RA00920BGoogle Scholar65https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXjtVeisbw%253D&md5=47af55f3a988711d077a6af9263cc27aThe relative stability of trpzip1 and its mutants determined by computation and experimentBureau, Hailey R.; Quirk, Stephen; Hernandez, RigobertoRSC Advances (2020), 10 (11), 6520-6535CODEN: RSCACL; ISSN:2046-2069. (Royal Society of Chemistry)Six mutants of the tryptophan zipper peptide trpzip1 have been computationally and exptl. characterized. We det. the varying roles in secondary structure stability of specific residues through a mutation assay. Four of the mutations directly effect the Trp-Trp interactions and two of the mutations target the salt bridge between Glu5 and Lys8. CD spectra and thermal unfolding are used to det. the secondary structure and stability of the mutants compared to the wildtype peptide. Adaptive steered mol. dynamics has been used to obtain the energetics of the unfolding pathways of the mutations. The hydrogen bonding patterns and side-chain interactions over the course of unfolding have also been calcd. and compared to wildtype trpzip1. The key finding from this work is the importance of a stabilizing non-native salt bridge pair present in the K8L mutation.
- 66Zhuang, Y.; Bureau, H.; Quirk, S.; Hernandez, R. Adaptive Steered Molecular Dynamics of Biomolecules. Mol. Simul. 2021, 47, 408, DOI: 10.1080/08927022.2020.1807542Google Scholar66https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXhs1Kks7nN&md5=c51977e6a907041213010ccd153ccfb8Adaptive steered molecular dynamics of biomoleculesZhuang, Yi; Bureau, Hailey R.; Quirk, Stephen; Hernandez, RigobertoMolecular Simulation (2021), 47 (5), 408-419CODEN: MOSIEA; ISSN:0892-7022. (Taylor & Francis Ltd.)Adaptive steered mol. dynamics (ASMD) is a variant of steered mol. dynamics (SMD) in which the driving of the auxiliary - viz steered - particle is performed in stages. In SMD, many nonequil. trajectories are generated allowing for fast sampling of the ensemble space and access to rare events. Equil. observables, such as the potential of mean force along the pathway, result from averaging over these trajectories using Jarzynski's Equality (JE). Unfortunately, in SMD, a large no. of trajectories are needed to cover all possible configurations in order to obtain converged quantities in the exponential av. of the JE, and this is computationally expensive. ASMD reduces the no. of trajectories that must be sampled by discarding those trajectories that have deviated far from the equil. path in stages. At the end of a stage, one chooses - or contracts - one (in naive ASMD) or some (in multi-branched ASMD or MB-ASMD) of the configurations produced from the previous stage to initiate the trajectories in the next stage. Alternatively, in full-relaxation ASMD (FR-ASMD), all generated structures are relaxed under the constraint of a fixed auxiliary particle exerting no net work on the system. We provide a direct comparison of the energetics and other observables obtained from these approaches. We find that FR-ASMD is preferred when the unfolding pathways follow up along a single funnel and the system is sufficiently small that computational resources are not a limiting concern. It gives the highest accuracy in such cases while avoiding the inefficiencies of SMD. However, for complex energy landscapes typical of most multi-domain proteins, MB-ASMD is preferred because it provides a mechanism to sample alternative pathways while suffering only a modest loss in accuracy compared to FR-ASMD.
- 67Zhuang, Y.; Bureau, H. R.; Lopez, C.; Bucher, R.; Quirk, S.; Hernandez, R. Energetics and structure of alanine-rich _-helices via Adaptive Steered Molecular Dynamics (ASMD). Biophys. J. 2021, 120, 2009, DOI: 10.1016/j.bpj.2021.03.017Google Scholar67https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXoslGkt74%253D&md5=b8e34535a458f3191440b0477bb2cc6cEnergetics and structure of alanine-rich α-helices via adaptive steered molecular dynamicsZhuang, Yi; Bureau, Hailey R.; Lopez, Christine; Bucher, Ryan; Quirk, Stephen; Hernandez, RigobertoBiophysical Journal (2021), 120 (10), 2009-2018CODEN: BIOJAU; ISSN:0006-3495. (Cell Press)The energetics and hydrogen bonding profiles of the helix-to-coil transition were found to be an additive property and to increase linearly with chain length, resp., in alanine-rich α-helical peptides. A model system of polyalanine repeats was used to establish this hypothesis for the energetic trends and hydrogen bonding profiles. Numerical measurements of a synthesized polypeptide Ac-Y(AEAAKA)kF-NH2 and a natural α-helical peptide a2N (1-17) provide evidence of the hypothesis's generality. Adaptive steered mol. dynamics was employed to investigate the mech. unfolding of all of these alanine-rich polypeptides. We found that the helix-to-coil transition is primarily dependent on the breaking of the intramol. backbone hydrogen bonds and independent of specific side-chain interactions and chain length. The mech. unfolding of the α-helical peptides results in a turnover mechanism in which a 310-helical structure forms during the unfolding, remaining at a near const. population and thereby maintaining additivity in the free energy. The intermediate partially unfolded structures exhibited polyproline II helical structure as previously seen by others. In summary, we found that the av. force required to pull alanine-rich α-helical peptides in between the endpoints-namely the native structure and free coil-is nearly independent of the length or the specific primary structure.
- 68Mahala, B.; Hernandez, R. Solvent softness effects on unimolecular chemical reaction rate constants. Chem. Phys. Lett. 2020, 744, 137182, DOI: 10.1016/j.cplett.2020.137182Google Scholar68https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXjtF2iu70%253D&md5=f44619c6871d6d5799fda42ba532e37cSolvent softness effects on unimolecular chemical reaction rate constantsMahala, Benjamin D.; Hernandez, RigobertoChemical Physics Letters (2020), 744 (), 137182CODEN: CHPLBC; ISSN:0009-2614. (Elsevier B.V.)The stochastic hard collision (SHC) model for a coarse-grained solvent has been used to solvate a class of model chem. reactions. Although there is precedent from our prior work that the use of nondeterministic interaction potentials can lead to accurate dynamics at long-enough time scales, this work provides evidence that this model has dynamical consistency specifically in the response to the reactant motion. That is, the agreement between the theor. rate consts. and those obtained from the stochastic simulations provides support for the SHC solvent without sacrificing the correlations in the dynamics of the reacting solutes.
- 69Murphy, C. J.; Vartanian, A. M.; Geiger, F. M.; Hamers, R. J.; Pedersen, J.; Cui, Q.; Haynes, C. L.; Carlson, E. E.; Hernandez, R.; Klaper, R. D.; Orr, G.; Rosenzweig, Z. Biological responses to engineered nanomaterials: Needs for the next decade. ACS Cent. Sci. 2015, 1, 117, DOI: 10.1021/acscentsci.5b00182Google Scholar69https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhtVaku7vM&md5=35006c55ffaa627f090c5f23d8ca8d8aBiological Responses to Engineered Nanomaterials: Needs for the Next DecadeMurphy, Catherine J.; Vartanian, Ariane M.; Geiger, Franz M.; Hamers, Robert J.; Pedersen, Joel; Cui, Qiang; Haynes, Christy L.; Carlson, Erin E.; Hernandez, Rigoberto; Klaper, Rebecca D.; Orr, Galya; Rosenzweig, Ze'evACS Central Science (2015), 1 (3), 117-123CODEN: ACSCII; ISSN:2374-7951. (American Chemical Society)A review. The interaction of nanomaterials with biomols., cells, and organisms is an enormously vital area of current research, with applications in nanoenabled diagnostics, imaging agents, therapeutics, and contaminant removal technologies. Yet the potential for adverse biol. and environmental impacts of nanomaterial exposure is considerable and needs to be addressed to ensure sustainable development of nanomaterials. In this Outlook four research needs for the next decade are outlined: (i) measurement of the chem. nature of nanomaterials in dynamic, complex aq. environments; (ii) real-time measurements of nanomaterial-biol. interactions with chem. specificity; (iii) delineation of mol. modes of action for nanomaterial effects on living systems as functions of nanomaterial properties; and (iv) an integrated systems approach that includes computation and simulation across orders of magnitude in time and space.
- 70Wu, M.; Vartanian, A. M.; Chong, G.; Pandiakumar, A. K.; Hamers, R. J.; Hernandez, R.; Murphy, C. J. Solution NMR analysis of ligand environment in quaternary ammonium-terminated self-assembled monolayers on gold nanoparticles: The effect of surface curvature and ligand structure. J. Am. Chem. Soc. 2019, 141, 4316, DOI: 10.1021/jacs.8b11445Google Scholar70https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXjtFCrt74%253D&md5=533e812bc895c8d2b559484e647f1d15Solution NMR Analysis of Ligand Environment in Quaternary Ammonium-Terminated Self-Assembled Monolayers on Gold Nanoparticles: The Effect of Surface Curvature and Ligand StructureWu, Meng; Vartanian, Ariane M.; Chong, Gene; Pandiakumar, Arun Kumar; Hamers, Robert J.; Hernandez, Rigoberto; Murphy, Catherine J.Journal of the American Chemical Society (2019), 141 (10), 4316-4327CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)We report a soln. NMR-based anal. of (16-mercaptohexadecyl)trimethylammonium bromide (MTAB) self-assembled monolayers on colloidal gold nanospheres (AuNSs) with diams. from 1.2 to 25 nm and gold nanorods (AuNRs) with aspect ratios from 1.4 to 3.9. The chem. shift anal. of the proton signals from the solvent-exposed headgroups of bound ligands suggests that the headgroups are satd. on the ligand shell as the sizes of the nanoparticles increase beyond ∼10 nm. Quant. NMR shows that the ligand d. of MTAB-AuNSs is size-dependent. Ligand d. ranges from ∼3 mols. per nm2 for 25 nm particles to up to 5-6 mols. per nm2 in ∼10 nm and smaller particles for in situ measurements of bound ligands; after I2/I- treatment to etch away the gold cores, ligand d. ranges from ∼2 mols. per nm2 for 25 nm particles to up to 4-5 mols. per nm2 in ∼10 nm and smaller particles. T2 relaxation anal. shows greater hydrocarbon chain ordering and less headgroup motion as the diam. of the particles increases from 1.2 nm to ∼13 nm. Mol. dynamics simulations of 4, 6, and 8 nm (11-mercaptoundecyl)trimethylammonium bromide-capped AuNSs confirm greater hydrophobic chain packing order and satn. of charged headgroups within the same spherical ligand shell at larger nanoparticle sizes and higher ligand densities. Combining the NMR studies and MD simulations, we suggest that the headgroup packing limits the ligand d., rather than the sulfur packing on the nanoparticle surface, for ∼10 nm and larger particles. For MTAB-AuNRs, no chem. shift data nor ligand d. data suggest that two populations of ligands that might correspond to side-ligands and end-ligands exist; yet T2 relaxation dynamics data suggest that headgroup mobility depends on aspect ratio and abs. nanoparticle dimensions.
- 71Daly, C. A., Jr.; Allen, C. R.; Rozanov, N. D.; Chong, G.; Melby, E. S.; Kuech, T. R.; Lohse, S. E.; Murphy, C. J.; Pedersen, J. A.; Hernandez, R. Surface Coating Structure and Its Interaction with Cytochrome c in EG6-Coated Nanoparticles Varies with Surface Curvature. Langmuir 2020, 36, 5030– 5039, DOI: 10.1021/acs.langmuir.0c00681Google Scholar71https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXnsVCns7g%253D&md5=9e71f49e51279feec665b49fd71e8e55Surface Coating Structure and Its Interaction with Cytochrome c in EG6-Coated Nanoparticles Varies with Surface CurvatureDaly, Clyde A.; Allen, Caley; Rozanov, Nikita; Chong, Gene; Melby, Eric S.; Kuech, Thomas R.; Lohse, Samuel E.; Murphy, Catherine J.; Pedersen, Joel A.; Hernandez, RigobertoLangmuir (2020), 36 (18), 5030-5039CODEN: LANGD5; ISSN:0743-7463. (American Chemical Society)The compn., orientation, and conformation of proteins in biomol. coronas acquired by nanoparticles in biol. media contribute to how they are identified by a cell. While numerous studies have investigated protein compn. in biomol. coronas, relatively little detail is known about how the nanoparticle surface influences the orientation and conformation of the proteins assocd. with them. The authors previously showed that the peripheral membrane protein cytochrome c adopts preferred poses relative to neg. charged 3-mercaptopropionic acid (MPA)-gold nanoparticles (AuNPs). Here, the authors employ mol. dynamics simulations and complementary expts. to establish that cytochrome c also assumes preferred poses upon assocn. with nanoparticles functionalized with an uncharged ligand, specifically ω-(1-mercaptounde-11-cyl)hexa(ethylene glycol) (EG6). The authors find that the display of the EG6 ligands is sensitive to the curvature of the surface-and, consequently, the effective diam. of the nearly spherical nanoparticle core-which in turn affects the preferred poses of cytochrome c.
- 72Bathe, M.; Hernandez, R.; Komiyama, T.; Machiraju, R.; Neogi, S. Autonomous computing materals. ACS Nano 2021, 15, 3586– 3592, DOI: 10.1021/acsnano.0c09556Google Scholar72https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXlt1Shs7k%253D&md5=f02169e68d5a3171fcfaf8b3f25cfa1cAutonomous Computing MaterialsBathe, Mark; Hernandez, Rigoberto; Komiyama, Takaki; Machiraju, Raghu; Neogi, SanghamitraACS Nano (2021), 15 (3), 3586-3592CODEN: ANCAC3; ISSN:1936-0851. (American Chemical Society)A review. Conventional materials are reaching their limits in computation, sensing, and data storage capabilities, ushered in by the end of Moore's law, myriad sensing applications, and the continuing exponential rise in worldwide data storage demand. Conventional materials are also limited by the controlled environments in which they must operate, their high energy consumption, and their limited capacity to perform simultaneous, integrated sensing, computation, and data storage and retrieval. In contrast, the human brain is capable of multimodal sensing, complex computation, and both short- and long-term data storage simultaneously, with near instantaneous rate of recall, seamless integration, and minimal energy consumption. Motivated by the brain and the need for revolutionary new computing materials, we recently proposed the data-driven materials discovery framework, autonomous computing materials. This framework aims to mimic the brain's capabilities for integrated sensing, computation, and data storage by programming excitonic, phononic, photonic, and dynamic structural nanoscale materials, without attempting to mimic the unknown implementational details of the brain. If realized, such materials would offer transformative opportunities for distributed, multimodal sensing, computation, and data storage in an integrated manner in biol. and other nonconventional environments, including interfacing with biol. sensors and computers such as the brain itself.
- 73Webb, D. R. National Research Council (Us) Chemical Sciences Roundtable. Minorities in the Chemical Workforce: Diversity Models That Work: A Workshop Report to the Chemical Sciences Roundtable; NRC, 2003; https://www.ncbi.nlm.nih.gov/books/NBK36322/, accessed March 31, 2018.Google ScholarThere is no corresponding record for this reference.
- 74Hernandez, R.; Stallings, D.; Iyer, S. In Diversity in the Scientific Community Vol. 1: Quantifying Diversity and Formulating Success; Cheng, H. N., Nelson, D., Eds.; ACS Symposium Series; American Chemical Society; Oxford University Press: Washington DC, 2017; Vol. 1255, pp 101– 112.Google ScholarThere is no corresponding record for this reference.
- 75Berryman, S. E. Who will Do Science? Minority and Female Attainment of Science and Mathematics Degrees: Trends and Causes; Rockefeller Foundation, 1983.Google ScholarThere is no corresponding record for this reference.
- 76Steele, C. M. A Threat in the Air: How Stereotypes Shape Intellectual Identity and Performance. Am. Psychol. 1997, 52, 613, DOI: 10.1037/0003-066X.52.6.613Google Scholar76https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADyaK2szhsleksw%253D%253D&md5=82b49b03cac3553c9fb53d20dbfadebaA threat in the air. How stereotypes shape intellectual identity and performanceSteele C MThe American psychologist (1997), 52 (6), 613-29 ISSN:0003-066X.A general theory of domain identification is used to describe achievement barriers still faced by women in advanced quantitative areas and by African Americans in school. The theory assumes that sustained school success requires identification with school and its subdomains; that societal pressures on these groups (e.g., economic disadvantage, gender roles) can frustrate this identification; and that in school domains where these groups are negatively stereotyped, those who have become domain identified face the further barrier of stereotype threat, the threat that others' judgments or their own actions will negatively stereotype them in the domain. Research shows that this threat dramatically depresses the standardized test performance of women and African Americans who are in the academic vanguard of their groups (offering a new interpretation of group differences in standardized test performance), that it causes disidentification with school, and that practices that reduce this threat can reduce these negative effects.
- 77Raber, L. R. Georgia Section Celebrates 75th Herty Medal. Chem. Eng. News 2009, 87, 43– 44, DOI: 10.1021/cen-v087n042.p043Google Scholar77https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhtlWktLvJ&md5=95fb2b73b17aafc1d0cb75ed247fefc8Georgia Section celebrates 75th Herty MedalRaber, Linda R.Chemical & Engineering News (2009), 87 (42), 43-44CODEN: CENEAR; ISSN:0009-2347. (American Chemical Society)The ACS Georgia Section celebrates the 75th awarding of the Charles Holmes Herty Medal for the year 2009. Craig L. Hill, a professor at Emory University, received the award for his contributions in catalytic oxidn. In particular, he is being recognized for his development of the most reactive known catalysts for removal of environmental pollutants, odors, and toxics from air; prepn. of terminal noble-metal oxo compds.; and creation of the only stable and sol. catalysts for oxidn. of water.
- 78Friend, C. M.; Houk, K. N. Workshop on Building Strong Academic Chemistry Departments through Gender Equity; American Chemical Society, 2006; https://www.acs.org/content/dam/acsorg/funding/awards/national/gender-equity-report-cover.pdf, accessed January 16, 2021.Google ScholarThere is no corresponding record for this reference.
- 79Ali, H. B. Workshop on Excellence Empowered by a Diverse Academic Workforce: Achieving Racial & Ethnic Equity in Chemistry; OSTI, 2008; https://www.osti.gov/biblio/952471, accessed March 31, 2018.Google ScholarThere is no corresponding record for this reference.
- 80Hernandez, R.; Watt, S. In Career Challenges and Opportunities in the Global Chemistry Enterprise; Cheng, H. N., Shah, S., Wu, M. L., Eds.; ACS Symposium Series; American Chemical Society; Oxford University Press: Washington DC, 2014; Vol. 1169, pp 207– 224.Google ScholarThere is no corresponding record for this reference.
- 81Stallings, D.; Iyer, S.; Hernandez, R. In National Diversity Equity Workshops in Chemical Sciences (2011–2017); Hernandez, R., Stallings, D., Iyer, S. K., Eds.; ACS Symposium Series; American Chemical Society; Oxford University Press: Washington DC, 2018; Vol. 1277, pp 1– 19.Google ScholarThere is no corresponding record for this reference.
- 82Iyer, S.; Stallings, D.; Hernandez, R. In National Diversity Equity Workshops in Chemical Sciences (2011–2017); Hernandez, R., Stallings, D., Iyer, S. K., Eds.; ACS Symposium Series; American Chemical Society; Oxford University Press: Washington DC, 2018; Vol. 1277, pp 21– 49.Google ScholarThere is no corresponding record for this reference.
- 83Stallings, D.; Iyer, S.; Hernandez, R. In National Diversity Equity Workshops in Chemical Sciences (2011–2017); Hernandez, R., Stallings, D., Iyer, S. K., Eds.; ACS Symposium Series; American Chemical Society; Oxford University Press: Washington DC, 2018; Vol. 1277, pp 51– 77.Google ScholarThere is no corresponding record for this reference.
- 84Iyer, S.; Stallings, D.; Hernandez, R. In National Diversity Equity Workshops in Chemical Sciences (2011–2017); Hernandez, R., Stallings, D., Iyer, S. K., Eds.; ACS Symposium Series; American Chemical Society; Oxford University Press: Washington DC, 2018; Vol. 1277, pp 79– 107.Google ScholarThere is no corresponding record for this reference.
- 85Stallings, D.; Iyer, S.; Hernandez, R. In National Diversity Equity Workshops in Chemical Sciences (2011–2017); Hernandez, R., Stallings, D., Iyer, S. K., Eds.; ACS Symposium Series; American Chemical Society; Oxford University Press: Washington DC, 2018; Vol. 1277, pp 109– 140.Google ScholarThere is no corresponding record for this reference.
- 86Jacobs, M. Reasons Sought for Lack of Diversity. Chem. Eng. News 2001, 79, 100– 103, DOI: 10.1021/cen-v079n040.p100Google ScholarThere is no corresponding record for this reference.
- 87Hernandez, R. Advancing the chemical sciences through diversity. Chem. Eng. News 2014, 92, 45Google ScholarThere is no corresponding record for this reference.
- 88Kotov, A. Think Like a Grandmaster; Batsford, 1970.Google ScholarThere is no corresponding record for this reference.
- 89Clark, A. E.; Adams, H.; Hernandez, R.; Krylov, A. I.; Niklasson, A. M. N.; Sarupria, S.; Wang, Y.; Wild, S. M.; Yang, Q. Y. ACS Cent. Sci. 2021, DOI: 10.1021/acscentsci.1c00685 .Google ScholarThere is no corresponding record for this reference.
Cited By
This article is cited by 4 publications.
- Rigoberto Hernandez. Discipline-Based Diversity Research in Chemistry. Accounts of Chemical Research 2023, 56
(7)
, 787-797. https://doi.org/10.1021/acs.accounts.2c00797
- Vicki H. Grassian. Physical Chemistry of Environmental Interfaces and the Environment in Physical Chemistry─A Career Perspective. The Journal of Physical Chemistry A 2022, 126
(30)
, 4874-4880. https://doi.org/10.1021/acs.jpca.2c04098
- Vicki H. Grassian. Physical Chemistry of Environmental Interfaces and the Environment in Physical Chemistry─A Career Perspective. The Journal of Physical Chemistry B 2022, 126
(30)
, 5598-5604. https://doi.org/10.1021/acs.jpcb.2c04099
- Vicki H. Grassian. Physical Chemistry of Environmental Interfaces and the Environment in Physical Chemistry─A Career Perspective. The Journal of Physical Chemistry C 2022, 126
(30)
, 12320-12326. https://doi.org/10.1021/acs.jpcc.2c04100
Figure 1
Figure 1. Then and now. At left, Hernandez is pictured in his high school yearbook in a stylish cotton suit, gray leather shoes, and open collar which was on trend if different from the fashion of polos and tweed that would be the norm just a few months later upon teleporting to Princeton (photo by Mark Maynard). At right, Hernandez is pictured in his current work uniform featuring a tie, white dress shirt, and a jacket (photo by Robert R. Felt).
Figure 2
Figure 2. A selected snapshot of Hernandez’s research efforts shown (clockwise from top left) through journal covers spanning across transition state theory (reprinted with permission from ref (1), copyright 2010 Elsevier), Janus and striped particles (reprinted with permission from ref (2), copyright 2014 AIP Publishing), diffusion through complex environments (reprinted with permission from ref (3), copyright 2010 American Chemical Society), stochastic hard collisions (reprinted with permission from ref (4), copyright 2018 Elsevier), sustainable nanoparticles (reprinted with permission from ref (5), copyright 2016 American Chemical Society), and diversity equity (reprinted with permission from ref (6), copyright 2018 American Chemical Society).
References
ARTICLE SECTIONSThis article references 89 other publications.
- 1Hernandez, R.; Bartsch, T.; Uzer, T. Transition State Theory in Liquids Beyond Planar Dividing Surfaces. Chem. Phys. 2010, 370, 270– 276, DOI: 10.1016/j.chemphys.2010.01.016Google Scholar1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXlslCku7Y%253D&md5=6fabe3c8f033cdc4f304c9e1dcebf14eTransition state theory in liquids beyond planar dividing surfacesHernandez, Rigoberto; Uzer, T.; Bartsch, ThomasChemical Physics (2010), 370 (1-3), 270-276CODEN: CMPHC2; ISSN:0301-0104. (Elsevier B.V.)The success of transition state theory (TST) in describing the rates of chem. reactions has been less-than-perfect in soln. (and sometimes even in the gas phase) because conventional dividing surfaces are only approx. free of recrossings between reactants and products. Recent advances in dynamical systems theory have helped to identify the interconnected manifolds-"superhighways"-leading from reactants to products. The existence of these manifolds has been proven rigorously, and explicit algorithms are available for their calcn. We now show that these extended structures can be used to obtain reaction rates directly in dissipative systems. We also suggest a treatment for the substantially more general case in which the mol. solvent is fully specified by the positions of all its atoms. Specifically, we can construct effective solvent configurations for which the exact TST manifolds can be constructed and used to sample the rates of an open system.
- 2Hagy, M. C.; Hernandez, R. Dynamical simulation of electrostatic striped colloidal particles. J. Chem. Phys. 2014, 140, 034701, DOI: 10.1063/1.4859855Google Scholar2https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXps1CisQ%253D%253D&md5=384c7aaa7f06c8f42d6136f890b0bf63Dynamical simulation of electrostatic striped colloidal particlesHagy, Matthew C.; Hernandez, RigobertoJournal of Chemical Physics (2014), 140 (3), 034701/1-034701/13CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The static and dynamic properties of striped colloidal particles are obtained using mol. dynamics computer simulations. Striped particles with n = 2 to n = 7 stripes of alternating elec. charge are modeled at a high level of detail through a pointwise (PW) representation of the particle surface. We also consider the extent to which striped particles are similar to comparable isotropically attractive particles-such as depletion attracting colloids-by modeling striped particles with an isotropic pair interaction computed by coarse-graining (CG) over orientations at a pair level. Surprisingly, the CG models reproduce the static structure of the PW models for a range of vol. fractions and interaction strengths consistent with the fluid region of the phase diagram for all n. As a corollary, different n-striped particle systems with comparable pair affinities (e.g., dimer equil. const.) have similar static structure. Stronger pair interactions lead to a collapsed structure in simulation as consistent with a glass-like phase. Different n-striped particle systems are found to have different phase boundaries and for certain n's no glass-like state is obsd. in any of our simulations. The CG model is found to have accelerated dynamics relative to the PW model for the same range of fluid conditions for which the models have identical static structure. This suggests striped electrostatic particles have slower dynamics than comparable isotropically attractive colloids. The slower dynamics result from a larger no. of long-duration reversible bonds between pairs of striped particles than seen in isotropically attractive systems. Higher n-striped particles systems generally have slower dynamics than lower n-striped systems with comparable pair affinities. (c) 2014 American Institute of Physics.
- 3Tucker, A. K.; Hernandez, R. Observation of a trapping transition in the diffusion of a thick needle through fixed point scatterers. J. Phys. Chem. A 2010, 114, 9628– 9634, DOI: 10.1021/jp100111yGoogle Scholar3https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXkvV2rtrc%253D&md5=657bd4b9d5ed5c635882a3940f5d73f2Observation of a Trapping Transition in the Diffusion of a Thick Needle through Fixed Point ScatterersTucker, Ashley K.; Hernandez, RigobertoJournal of Physical Chemistry A (2010), 114 (36), 9628-9634CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The long-time correlation functions of an infinitesimally thin needle moving through stationary point scatterers--a so-called Lorentz model--exhibits surprisingly long-time tails. These long-time tails are now seen to persist in a two-dimensional model even when the needle has a finite thickness. If the needles are too thick, then the needles are effectively trapped at all nontrivial densities of the scatterers. At needle widths approx. equal to or smaller than σ = ε/20 where ε is the av. spacing between scatterers, the needle diffuses and exhibits the crossover transition from the expected Enskog behavior to the enhanced translation diffusion seen earlier by Hofling, Frey and Franosch. At this needle width, an increase in its center-of-mass diffusion with respect to increasing d. is seen after a crossover d. of n* ≈ 5 is reached. (The reduced d. n* is defined as n* = nL2 where n is the no. d. of particles and L is the needle length.). The crossover transition for needles with finite thickness is spread over a range of densities exhibiting intermediate behavior. The asymptotic divergence of the center of mass diffusion is suppressed compared to that of infinitely thin needles. Finally, a new diminished diffusion regime, apparently due to the increased importance of head-on collisions, now appears at high scattering densities.
- 4Singh, R. S.; Hernandez, R. Modeling soft core-shell colloids using stochastic hard collision dynamics. Chem. Phys. Lett. 2018, 708, 233– 240, DOI: 10.1016/j.cplett.2018.08.032Google Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXhsFKqtbvP&md5=8fba26473e6e7617d800939900293bbfModeling soft core-shell colloids using stochastic hard collision dynamicsSingh, Rakesh S.; Hernandez, RigobertoChemical Physics Letters (2018), 708 (), 233-240CODEN: CHPLBC; ISSN:0009-2614. (Elsevier B.V.)The equil. structure and thermodn. behavior of a soft matter suspension of hard-core soft-shell (HCSS) colloids have been modeled using stochastic hard collision (SHC) dynamics [Craven et al., J. Chem. Phys. 2013, 138, 244901]. SHC dynamics includes the effects of inter-penetrability of soft particles while retaining the computational efficiency of hard collision dynamics. The HCSS colloids are characterized using two parameters: softness parameter (a measure of the propensity toward interpenetration), and core-to-shell ratio. In both, mol. dynamics and Monte-Carlo simulations, the softness parameter is seen to profoundly affect the effective packing fraction and the thermodn. behavior of the system. Once we accounted for the trivial effects from changes in the reduced vol. fraction on changing softness, we found that the inter-particle penetration continues to affect this thermodn. behavior. The structural origin of this nontrivial effect likely lies in the loss of translational order upon increasing softness.
- 5Cui, Q.; Hernandez, R.; Mason, S. E.; Frauenheim, T.; Pedersen, J. A.; Geiger, F. Mini-review. Sustainable nanotechnology: Opportunities and challenges for theoretical/computational studies. J. Phys. Chem. B 2016, 120, 7297– 7306, DOI: 10.1021/acs.jpcb.6b03976Google Scholar5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XhtFers7nE&md5=9ffc40bbd229be188cacd88cead01028Sustainable Nanotechnology: Opportunities and Challenges for Theoretical/Computational StudiesCui, Qiang; Hernandez, Rigoberto; Mason, Sara E.; Frauenheim, Thomas; Pedersen, Joel A.; Geiger, FranzJournal of Physical Chemistry B (2016), 120 (30), 7297-7306CODEN: JPCBFK; ISSN:1520-5207. (American Chemical Society)For assistance in the design of the next generation of nanomaterials that are functional and have minimal health and safety concerns, it is imperative to establish causality, rather than correlations, in how properties of nanomaterials det. biol. and environmental outcomes. Due to the vast design space available and the complexity of nano/bio interfaces, theor. and computational studies are expected to play a major role in this context. In this minireview, we highlight opportunities and pressing challenges for theor. and computational chem. approaches to explore the relevant physicochem. processes that span broad length and time scales. We focus discussions on a bottom-up framework that relies on the detn. of correct intermol. forces, accurate mol. dynamics, and coarse-graining procedures to systematically bridge the scales, although top-down approaches are also effective at providing insights for many problems such as the effects of nanoparticles on biol. membranes.
- 6Hernandez, R.; Stallings, D.; Iyer, S. K. In National Diversity Equity Workshops in Chemical Sciences (2011–2017); Hernandez, R., Stallings, D., Iyer, S. K., Eds.; ACS Symposium Series; American Chemical Society; Oxford University Press: Washington DC, 2018; Vol. 1277.Google ScholarThere is no corresponding record for this reference.
- 7Hernandez, R. What’s in a name? (Part 1). 2013; http://everywherechemistry.blogspot.com/2013/04/whats-in-name-part-1.html, accessed January 15, 2021.Google ScholarThere is no corresponding record for this reference.
- 8Coy, S. L.; Hernandez, R.; Lehmann, K. K. Limits on the transition to Gaussian orthogonal ensemble behavior: Saturated radiationless transitions between strongly coupled potential surfaces. Phys. Rev. A: At., Mol., Opt. Phys. 1989, 40, 5935– 5949, DOI: 10.1103/PhysRevA.40.5935Google ScholarThere is no corresponding record for this reference.
- 9Nolan, S. A.; Buckner, J. P.; Kuck, V. J.; Marzabadi, C. H. Analysis by Gender of the Doctoral and Postdoctoral Institutions of Faculty Members at the Top-Fifty Ranked Chemistry Departments. J. Chem. Educ. 2004, 81, 356, DOI: 10.1021/ed081p356Google Scholar9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXhs1Cnurc%253D&md5=8a98162a965d871a7cb5eec68652b352Analysis by gender of the doctoral and postdoctoral institutions of faculty members at the top-fifty ranked chemistry departmentsKuck, Valerie J.; Marzabadi, Cecilia H.; Nolan, Susan A.; Buckner, Janine P.Journal of Chemical Education (2004), 81 (3), 356-363CODEN: JCEDA8; ISSN:0021-9584. (Journal of Chemical Education, Dept. of Chemistry)There is no expanded citation for this reference.
- 10Miller, W. H.; Hernandez, R.; Moore, C. B.; Polik, W. F. A transition state theory-based statistical distribution of unimolecular decay rates, with application to unimolecular decomposition of formaldehyde. J. Chem. Phys. 1990, 93, 5657– 5666, DOI: 10.1063/1.459636Google Scholar10https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3cXmtlyrsb4%253D&md5=0f49d9a550c645d84e5f668cf1568accA transition state theory-based statistical distribution of unimolecular decay rates with application to unimolecular decomposition of formaldehydeMiller, William H.; Hernandez, Rigoberto; Moore, C. Bradley; Polik, William F.Journal of Chemical Physics (1990), 93 (8), 5657-66CODEN: JCPSA6; ISSN:0021-9606.A statistical distribution of state-specific unimol. decay rates is derived (within the framework of random matrix theory) that is detd. completely by the transition state properties of the potential energy surface. It includes the std. χ-square distributions as a special case. Model calcns. are presented to show the extent to which it can differ from the χ-square distribution, and specific application is made to the state-specific unimol. decay rate data for D2CO → D2 + CO.
- 11Hernandez, R.; Miller, W. H.; Moore, C. B.; Polik, W. F. A Random Matrix/Transition State Theory for the Probability Distribution of State-Specific Unimolecular Decay Rates: Generalization to Include Total Angular Momentum Conservation and Other Dynamical Symmetries. J. Chem. Phys. 1993, 99, 950– 962, DOI: 10.1063/1.465360Google Scholar11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3sXmtlCqurw%253D&md5=6dba8c44383d72510019dd7f62290cacA random matrix/transition state theory for the probability distribution of state-specific unimolecular decay rates: generalization to include total angular momentum conservation and other dynamical symmetriesHernandez, Rigoberto; Miller, William H.; Moore, C. Bradley; Polik, William F.Journal of Chemical Physics (1993), 99 (2), 950-62CODEN: JCPSA6; ISSN:0021-9606.A previously developed random matrix/transition state theory (RM/TST) model for the probability distribution of state-sp. unimol. decay rates was generalized to incorporate total angular momentum conservation and other dynamic symmetries. The model is made into a predictive theory by using a semiclassical method to det. the transmission probabilities of a nonseparable rovibrational Hamiltonian at the transition state. The overall theory gives a good description of the state-sp. rates for the D2CO → D2 + CO unimol. decay; in particular, it describes the dependence of the distribution of rates on total angular momentum J. Comparison of the exptl. values with results of the RM/TST theory suggests that there is mixing among the rovibrational states.
- 12Miller, W. H. Autobiographical Sketch of William Hughes Miller. J. Phys. Chem. A 2001, 105, 2487– 2489, DOI: 10.1021/jp0101920Google ScholarThere is no corresponding record for this reference.
- 13Miller, W. H. Semi-classical theory for non-separable systems:. Construction of “good” action-angle variables for reaction rate constants. Faraday Discuss. Chem. Soc. 1977, 62, 40, DOI: 10.1039/DC9776200040Google Scholar13https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE1cXhsF2isrs%253D&md5=238838036ad290e52071a98b617f4aacSemiclassical theory for non-separable systems: construction of good action-angle variables for reaction rate constantsMiller, William H.Faraday Discussions of the Chemical Society (1977), 62 (Potential Energy Surf.), 40-6CODEN: FDCSB7; ISSN:0301-7249.A semiclassical expression for bimol. rate consts. for reactions which have a single activation barrier is obtained in terms of the good action variables of the classical Hamiltonian which are assocd. with the saddle point region of the potential energy surface. The formulas apply to nonseparable and separable saddle points.
- 14Miller, W. H.; Hernandez, R.; Handy, N. C.; Jayatilaka, D.; Willetts, A. Ab initio calculation of anharmonic constants for a transition state, with application to semiclassical transition state tunneling probabilities. Chem. Phys. Lett. 1990, 172, 62– 68, DOI: 10.1016/0009-2614(90)87217-FGoogle Scholar14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3cXlslyltLg%253D&md5=9933893dff69dead50201e0165060c7cAb initio calculation of anharmonic constants for a transition state, with application to semiclassical transition state tunneling probabilitiesMiller, William H.; Hernandez, Rigoberto; Handy, Nicholas C.; Jayatilaka, Dylan; Willetts, AndrewChemical Physics Letters (1990), 172 (1), 62-8CODEN: CHPLBC; ISSN:0009-2614."Good" (i.e. conserved) action variables exist in the vicinity of a saddle point (i.e. transition state) of a potential energy surface in complete analogy to those related to a min. on the surface. Transition state theory tunneling (or transmission) probabilities can be expressed semiclassically in terms of these "good" action variables, including the effects of non-separable coupling of all degrees of freedom with each other. This paper shows how ab initio quantum chem. methods recently developed for calcg. anharmonic consts. about a potential min. (i.e. for ordinary vibrational energy levels) can be readily adapted to obtain those related to a transition state, thus providing a rigorous and practical way to apply this non-separable transition state theory. Application is made to the transition state for the reaction D2CO → D2 + CO.
- 15Cohen, M. J.; Handy, N. C.; Hernandez, R.; Miller, W. H. Cumulative reaction probabilities for H + H2 → H2 + H from a knowledge of the anharmonic force field. Chem. Phys. Lett. 1992, 192, 407– 416, DOI: 10.1016/0009-2614(92)85491-RGoogle Scholar15https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK38XktFCltrk%253D&md5=941c2660fb9992c2485483bb7b07d700Cumulative reaction probabilities for atomic hydrogen + molecular hydrogen → molecular hydrogen + atomic hydrogen from a knowledge of the anharmonic force fieldCohen, Michael J.; Handy, Nicholas C.; Hernandez, Rigoberto; Miller, William H.Chemical Physics Letters (1992), 192 (4), 407-16CODEN: CHPLBC; ISSN:0009-2614.In an earlier publication, the authors showed how knowledge of a quartic force field expanded about a transition state can be used to obtain transition-state-theory tunneling probabilities. Thus coupling between the reaction mode and other modes is included in this second-order perturbation theory approach. The authors study the very anharmonic reaction H+H2→H2+H and show that even in this extreme case, there is reasonable agreement between the cumulative reaction probabilities calcd. by this semiclassical approach, and full quantum calcns.
- 16Hernandez, R.; Miller, W. H. Semiclassical Transition State Theory. A New Perspective. Chem. Phys. Lett. 1993, 214, 129– 136, DOI: 10.1016/0009-2614(93)90071-8Google Scholar16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2cXhs1Cqtg%253D%253D&md5=bd9ec72e0e1175247eb467b86ff67fc9Semiclassical transition state theory. A new perspectiveHernandez, Rigoberto; Miller, William H.Chemical Physics Letters (1993), 214 (2), 129-36CODEN: CHPLBC; ISSN:0009-2614.The semiclassical transition state theory (SCTST) introduced by W. H. Miller et al. (1990) requires the inversion of an (effectively integrable) Hamiltonian with respect to the action of the reactive coordinate. The inversion may be avoided in computing the thermal rate const.; the resulting expression provides an appealing link to conventional transition state theory. This reformulation of the SCTST rate is illustrated by application to the bimol. reaction, H + H2 → H2 + H, and to the unimol. dissocn., D2CO → D2 + CO.
- 17Hernandez, R. A Combined Use of Perturbation Theory and Diagonalization: Application to Bound Energy Levels and Semiclassical Rate Theory. J. Chem. Phys. 1994, 101, 9534– 9547, DOI: 10.1063/1.467985Google Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXisFChtLc%253D&md5=b157f8eb49cf25239abdcc314adfba62A combined use of perturbation theory and diagonalization: application to bound energy levels and semiclassical rate theoryHernandez, RigobertoJournal of Chemical Physics (1994), 101 (11), 9534-47CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)A new method, mixed diagonalization, is introduced in which an effective Hamiltonian operator acting on a reduced dimensional space is constructed using the similarity transformations of canonical Van Vleck perturbation theory (CVPT). This construction requires the characterization of modes into two categories, global and local, which in the bound vibrational problem are tantamount to the large and small amplitude vibrations, resp. The local modes in the Hamiltonian are projected out by CVPT, and the resulting Hamiltonian operator acts only on the space of global modes. The method affords the treatment of energy levels of bound systems in which some vibrational assignments are possible. It systematically provides a reduced dimensional Hamiltonian which is more amenable to exact numerical soln. that the original full-dimensional Hamiltonian. In recent work, a semiclassical transition state theory (SCTST) rate expression was written in terms of a Hamiltonian operator parameterized by the imaginary action along the local reaction path in the transition state region [Chem. Phys. Lett. 214, 129(1993)]. The Hamiltonian constructed by mixed diagonalization has this form, and can be used to obtain more accurate semiclassical rate expressions.
- 18Nguyen, T. L.; Stanton, J. F.; Barker, J. R. Ab Initio Reaction Rate Constants Computed Using Semiclassical Transition-State Theory: HO + H2 → H2O + H and Isotopologues. J. Phys. Chem. A 2011, 115, 5118– 5126, DOI: 10.1021/jp2022743Google Scholar18https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXlsFyrsro%253D&md5=bde4a2106d3e067acc5c90549fd685caAb Initio Reaction Rate Constants Computed Using Semiclassical Transition-State Theory: HO + H2 → H2O + H and IsotopologuesNguyen, Thanh Lam; Stanton, John F.; Barker, John R.Journal of Physical Chemistry A (2011), 115 (20), 5118-5126CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)A new algorithm for the semiclassical transition-state theory (SCTST) formulated by W.H. Miller and co-workers is used to compute rate consts. for the isotopologues of the title reaction, with no empirical adjustments. The SCTST and relevant results from second-order vibrational perturbation theory (VPT2) are summarized. VPT2 is used at the CCSD(T) level of electronic structure theory to compute the anharmonicities of the fully coupled vibrational modes (including the reaction coordinate) of the transition structure. The anharmonicities are used in SCTST to compute the rate consts. over the temp. range from 200 to 2500 K. The computed rate consts. are compared to exptl. data and theor. calcns. from the literature. The SCTST results for abs. rate consts. and for both primary and secondary isotope effects are in excellent agreement with the exptl. data for this reaction over the entire temp. range. The sensitivity of SCTST to various parameters is investigated by using a set of simplified models. The results show that multidimensional tunneling along the curved reaction path is important at low temps. and the anharmonic coupling among the vibrational modes is important at high temps. The theor. kinetics data are also presented as fitted empirical algebraic expressions.
- 19Barker, J. R.; Nguyen, T. L.; Stanton, J. F. Kinetic Isotope Effects for Cl + CH4 → HCl + CH3 Calculated Using ab Initio Semiclassical Transition State Theory. J. Phys. Chem. A 2012, 116, 6408– 6419, DOI: 10.1021/jp212383uGoogle Scholar19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhsVyksbY%253D&md5=9c2718ad57590e2e7d28f1ef1b412f97Kinetic Isotope Effects for Cl + CH4 .dblharw. HCl + CH3 Calculated Using ab Initio Semiclassical Transition State TheoryBarker, John R.; Nguyen, Thanh Lam; Stanton, John F.Journal of Physical Chemistry A (2012), 116 (24), 6408-6419CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Calcns. were carried out for 25 isotopologues of the title reaction for various combinations of 35Cl, 37Cl, 12C, 13C, 14C, H, and D. The computed rate consts. are based on harmonic vibrational frequencies calcd. at the CCSD(T)/aug-cc-pVTZ level of theory and Xij vibrational anharmonicity coeffs. calcd. at the CCSD(T) /aug-cc-pVDZ level of theory. For some reactions, anharmonicity coeffs. were also computed at the CCSD(T)/aug-cc-pVTZ level of theory. The classical reaction barrier was taken from Eskola et al. [J. Phys. Chem. A 2008, 112, 7391-7401], who extrapolated CCSD(T) calcns. to the complete basis set limit. Rate consts. were calcd. for temps. from ∼100 to ∼2000 K. The computed ab initio rate const. for the normal isotopologue is in good agreement with expts. over the entire temp. range (∼10% lower than the recommended exptl. value at 298 K). The ab initio H/D kinetic isotope effects (KIEs) for CH3D, CH2D2, CHD3, and CD4 are in very good agreement with literature exptl. data. The ab initio 12C/13C KIE is in error by ∼2% at 298 K for calcns. using Xij coeffs. computed with the aug-cc-pVDZ basis set, but the error is reduced to ∼1% when Xij coeffs. computed with the larger aug-cc-pVTZ basis set are used. Systematic improvements appear to be possible. The present SCTST results are found to be more accurate than those from other theor. calcns. Overall, this is a very promising method for computing ab initio kinetic isotope effects.
- 20Nguyen, T. L.; Stanton, J. F. Ab Initio Thermal Rate Calculations of HO + HO → O(3P) + H2O Reaction and Isotopologues. J. Phys. Chem. A 2013, 117, 2678– 2686, DOI: 10.1021/jp312246qGoogle Scholar20https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXjs1ynsL0%253D&md5=34768f5c4b7a916b8b7e87e664bb8cbbAb Initio Thermal Rate Calculations of HO + HO = O(3P) + H2O Reaction and IsotopologuesNguyen, Thanh Lam; Stanton, John F.Journal of Physical Chemistry A (2013), 117 (13), 2678-2686CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)The forward and reverse reactions, HO + HO .dblharw. O(3P) + H2O, which play roles in both combustion and lab. studies, were theor. characterized with a master equation approach to compute thermal reaction rate consts. at both the low and high pressure limits. Our ab initio k(T) results for the title reaction and two isotopic variants agree very well with expts. (within 15%) over a wide temp. range. The calcd. reaction rate shows a distinctly non-Arrhenius behavior and a strong curvature consistent with the expt. This characteristic behavior is due to effects of pos. barrier height and quantum mech. tunneling. Tunneling is very important and contributes more than 70% of total reaction rate at room temp. A prereactive complex is also important in the overall reaction scheme.
- 21Komatsuzaki, T.; Berry, R. S. Regularity in chaotic reaction paths. I. Ar6. J. Chem. Phys. 1999, 110, 9160– 9173, DOI: 10.1063/1.478838Google Scholar21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXis1Sis7Y%253D&md5=1550a27d59b9b8af2073cc8e85bfd2ffRegularity in chaotic reaction paths. I. Ar6Komatsuzaki, Tamiki; Berry, R. StephenJournal of Chemical Physics (1999), 110 (18), 9160-9173CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)We scrutinize the saddle crossings of a simple cluster of six atoms to show (a) that it is possible to choose a coordinate system in which the transmission coeff. for the classical reaction path is unity at all energies up to a moderately high energy, above which the transition state is chaotic; (b) that at energies just more than sufficient to allow passage across the saddle, all or almost all the degrees of freedom of the system are essentially regular in the region of the transition state; and (c) that the degree of freedom assocd. with the reaction coordinate remains essentially regular through the region of the transition state, even to moderately high energies. Microcanonical mol. dynamics simulation of Ar6 bound by pairwise Lennard-Jones potentials reveals the mechanics of passage. We use Lie canonical perturbation theory to construct the nonlinear transformation to a hyperbolic coordinate system which reveals these regularities. This transform "rotates away" the recrossings and nonregular behavior, esp. of the motion along the reaction coordinate, leaving a coordinate and a corresponding dividing surface in phase space which minimize recrossings and mode-mode mixing in the transition state region. The action assocd. with the reactive mode tends to be an approx. invariant of motion through the saddle crossings throughout a relatively wide range of energy. Only at very low energies just above the saddle could any other approx. invariants of motion be found for the other, nonreactive modes. No such local invariants appeared at energies at which the modes are all chaotic and coupled to one another.
- 22Komatsuzaki, T.; Berry, R. S. Regularity in chaotic reactions paths II: Ar6. Energy dependence and visualization of the reaction bottleneck. Phys. Chem. Chem. Phys. 1999, 1, 1387– 1397, DOI: 10.1039/a809424aGoogle Scholar22https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXhslGgs7g%253D&md5=5ff61674899be175158a96ca9ec2c56aRegularity in chaotic reaction paths II: Ar6. Energy dependence and visualization of the reaction bottleneckKomatsuzaki, Tamiki; Stephen Berry, R.Physical Chemistry Chemical Physics (1999), 1 (6), 1387-1397CODEN: PPCPFQ; ISSN:1463-9076. (Royal Society of Chemistry)Reaction trajectories reveal regular behavior in the reactive degree of freedom and unit transmission coeffs. as the system crosses the saddle region sepg. reactants and products. The regularity persists up to moderately high energies, even when all other degrees of freedom are chaotic. This behavior is apparent in a representation obtained by transformation with Lie canonical perturbation theory. The dividing surface in this representation is analogous to the conventional dividing surface in the sense that it is the point set for which the reaction coordinate has the const. value it has at the saddle-point singularity. However the nonlinear, full phase-space character of the transformation makes the new crossing surface a complicated, abstr. object whose interpretation and visualization, the objective of this paper, can be realized by cataloging the recrossings as they disappear in successively higher orders of perturbation, and by projection into spaces of only a few dimensions. The result is a conceptual interpretation of how regular behavior persists in a reactive degree of freedom.
- 23Komatsuzaki, T.; Berry, R. S. Local regularity and non-recrossing path in transition state—a new strategy in chemical reaction theories. J. Mol. Struct.: THEOCHEM 2000, 506, 55– 70, DOI: 10.1016/S0166-1280(00)00402-4Google Scholar23https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXksFahsLs%253D&md5=bb91e969b136a0772a2798a29ce9808fLocal regularity and non-recrossing path in transition state - a new strategy in chemical reaction theoriesKomatsuzaki, T.; Berry, R. S.Journal of Molecular Structure: THEOCHEM (2000), 506 (), 55-70CODEN: THEODJ; ISSN:0166-1280. (Elsevier Science B.V.)We analyze local regularities in the regions of transition states of a 6-atom Lennard-Jones cluster to demonstrate how one can choose a non-recrossing reaction path in phase space along which the transmission coeff. for the classical reaction path is unity from threshold up to a moderately high energy, above which the transition state is chaotic, and how one can picture the nonlinear, full-phase-space character of the dividing hypersurface by projecting it into spaces of only a few dimensions. These overcome one of the long-standing ambiguities in chem. reaction theories, the recrossing problem, up to moderately high energies, and make transition state theory more generalized, applicable even in cases in which apparent recrossings spoil the conventional theory.
- 24Komatsuzaki, T.; Berry, R. S. Regularity in chaotic reaction paths III: Ar6 local invariances at the reaction bottleneck. J. Chem. Phys. 2001, 115, 4105– 4117, DOI: 10.1063/1.1385152Google Scholar24https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXmt1ajsLw%253D&md5=52c4cadf179b5d0bbe341a10e61552d8Regularity in chaotic reaction paths III: Ar6 local invariances at the reaction bottleneckKomatsuzaki, Tamiki; Berry, R. StephenJournal of Chemical Physics (2001), 115 (9), 4105-4117CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)We recently developed a new method to ext. a many-body phase-space dividing surface, across which the transmission coeff. for the classical reaction path is unity. The example of isomerization of a 6-atom Lennard-Jones cluster showed that the action assocd. with the reaction coordinate is an approx. invariant of motion through the saddle regions, even at moderately high energies, at which most or all the other modes are chaotic [J. Chem. Phys. 105, 10838 (1999); Phys. Chem. Chem. Phys. 1, 1387 (1999)]. In the present article, we propose a new algorithm to analyze local invariances about the transition state of N-particle Hamiltonian systems. The approx. invariants of motion assocd. with a reaction coordinate in phase space densely distribute in the sea of chaotic modes in the region of the transition state. Using projections of distributions in only two principal coordinates, one can grasp and visualize the stable and unstable invariant manifolds to and from a hyperbolic point of a many-body nonlinear system, like those of the one-dimensional, integrable pendulum. This, in turn, reveals a new type of phase space bottleneck in the region of a transition state that emerges as the total energy increases, which may trap a reacting system in that region.
- 25Komatsuzaki, T.; Berry, R. S. Dynamical hierarchy in transition states: Why and how does a system climb over the mountain?. Proc. Natl. Acad. Sci. U. S. A. 2001, 98, 7666– 7671, DOI: 10.1073/pnas.131627698Google Scholar25https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXlt1Kns7o%253D&md5=04d954cf2d5bd6bd942c1664d8e7356aDynamical hierarchy in transition states: why and how does a system climb over the mountain?Komatsuzaki, Tamiki; Berry, R. StephenProceedings of the National Academy of Sciences of the United States of America (2001), 98 (14), 7666-7671CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)How a reacting system climbs through a transition state during a reaction was an intriguing subject for decades. Here the authors present and quantify a technique to identify and characterize local invariance about the transition state of an N-particle Hamiltonian system, using Lie canonical perturbation theory combined with microcanonical mol. dynamics simulation. At least three distinct energy regimes of dynamical behavior occur in the region of the transition state, distinguished by the extent of their local dynamical invariance and regularity. Isomerization of a six-atom Lennard-Jones cluster illustrates this: up to energies high enough to make the system manifestly chaotic, approx. invariants of motion assocd. with a reaction coordinate in phase space imply a many-body dividing hypersurface in phase space that is free of recrossings even in a sea of chaos. The method makes it possible to visualize the stable and unstable invariant manifolds leading to and from the transition state, i.e., the reaction path in phase space, and how this regularity turns to chaos with increasing total energy of the system. This, in turn, illuminates a new type of phase space bottleneck in the region of a transition state that emerges as the total energy and mode coupling increase, which keeps a reacting system increasingly trapped in that region.
- 26Pollak, E.; Pechukas, P. Unified statistical model for “complex” and “direct” reaction mechanisms: A test on the collinear H + H2 exchange reaction. J. Chem. Phys. 1979, 70, 325– 333, DOI: 10.1063/1.437194Google Scholar26https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE1MXhtFSlsbo%253D&md5=068124475d995a1a85f774e33349c13aUnified statistical model for "complex" and "direct" reaction mechanisms: a test on the collinear atomic hydrogen + molecular hydrogen exchange reactionPollak, Eli; Pechukad, PhilipJournal of Chemical Physics (1979), 70 (1), 325-33CODEN: JCPSA6; ISSN:0021-9606.Miller's unified statistical theory for bimol. chem. reactions is tested on the collinear H + H2 exchange reaction, treated classically. The reaction probability calcd. from unified statistical theory is more accurate than that calcd. from ordinary transition state theory or from variational transition state theory; in particular, unified statistical theory predicts the high-energy falloff of the reaction probability, which transition state theory does not. A derivation of unified statistical theory emphasizes the dynamical and statistical assumptions that are the foundation of the theory. These assumptions unambiguously define the "collision complex" in unified statistical theory and are tested in detail on the H + H2 reaction. Finally, a lower bound on the reaction probability is derived; this bound complements the upper bound provided by transition state theory and is significantly more accurate, for the H + H2 reaction, than either transition state theory or unified statistical theory.
- 27Pollak, E. In Theory of Chemical Reaction Dynamics; Baer, M., Ed.; CRC Press: Boca Raton, FL, 1985; Vol. 3; p 123.Google ScholarThere is no corresponding record for this reference.
- 28Grobgeld, D.; Pollak, E.; Zakrzewski, J. A Numerical Method for Locating Stable Periodic Orbits in Chaotic Systems. Phys. D 1992, 56, 368– 380, DOI: 10.1016/0167-2789(92)90176-NGoogle ScholarThere is no corresponding record for this reference.
- 29Moix, J. M.; Hernandez, R.; Pollak, E. Momentum and velocity autocorrelation functions of a diatomic molecule are not necessarily proportional to each other. J. Phys. Chem. B 2008, 112, 213– 218, DOI: 10.1021/jp0730951Google Scholar29https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXhtVSjsbnP&md5=ff8195a43609d8c18e227241cc6a7d00Momentum and Velocity Autocorrelation Functions of a Diatomic Molecule Are Not Necessarily Proportional to Each OtherMoix, Jeremy M.; Hernandez, Rigoberto; Pollak, EliJournal of Physical Chemistry B (2008), 112 (2), 213-218CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)We present a computation of the classical momentum and velocity correlation functions of Br2 considered as an idealized mol. wire connecting dissipated lead atoms at each end of the dimer. It is demonstrated that coupling of the diat. relative momentum to the leads may result in momenta that are not equal to the mass-weighted velocity. These differences show up in numerical simulations of both the av. value and time correlations of the bond momentum and velocity. These observations are supported by anal. predictions for the av. temp. of the diat. They imply that the "std. recipes" for modeling the system with a generalized Langevin equation are insufficient.
- 30Hernandez, R.; Cao, J.; Voth, G. A. On the Feynman path centroid density as a phase space distribution in quantum statistical mechanics. J. Chem. Phys. 1995, 103, 5018– 5026, DOI: 10.1063/1.470588Google Scholar30https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXotFams78%253D&md5=e2a70c3406eeb3fdd44a51a21704c803On the Feynman path centroid density as a phase space distribution in quantum statistical mechanicsHernandez, Rigoberto; Cao, Jianshu; Voth, Gregory A.Journal of Chemical Physics (1995), 103 (12), 5018-26CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The phase space formulation of quantum statistical mechanics using the Feynman path centroid d. offers an alternative perspective to the std. Wigner prescription for the classical-like evaluation of equil. and/or dynamical quantities of statistical systems. The use of this formulation has been implicit in recent work on quantum rate theories, for example, in which the centroid d. distribution replaces the classical Boltzmann distribution. In order to further understand the approxns. involved in this and similar transcriptions, the present work elaborates and clarifies the issue of operator ordering in a rigorous centroid-based formulation. In particular, through the use of the Weyl correspondence, a precise definition of the centroid symbol of operators and their products is presented. Though we fall short of finding the algebraic structure tantamount to that found in the Weyl symbols-of which the Wigner distribution is an example- the resulting expressions have internal consistency and are amenable to approx. evaluation through cumulant expansions.
- 31Hernandez, R.; Voth, G. A. Quantum time correlation functions and classical coherence. Chem. Phys. 1998, 233, 243– 255, DOI: 10.1016/S0301-0104(98)00027-5Google Scholar31https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXkvVegt7s%253D&md5=26cc85ce6144fb34185db68125d0c9b5Quantum time correlation functions and classical coherenceHernandez, Rigoberto; Voth, Gregory A.Chemical Physics (1998), 233 (2,3), 243-255CODEN: CMPHC2; ISSN:0301-0104. (Elsevier Science B.V.)Quantum time correlation functions for electronically adiabatic and nonadiabatic processes have recently been evaluated directly using the semiclassical initial value representation [J. Cao and G.A. Voth, J. Chem. Phys. 104 (1996) 271]. In this paper, the approach to the classical limit of this theory is analyzed and the nature of coherence in such limits is explored.
- 32Verashchagina, A.; Bettio, F. Gender segregation in the labour market; Directorate-General for Employment, Social Affairs and Inclusion (European Commission), 2009.Google ScholarThere is no corresponding record for this reference.
- 33Hernandez, R.; Somer, F. L. Stochastic dynamics in irreversible nonequilibrium environments. 1. The fluctuation-dissipation relation. J. Phys. Chem. B 1999, 103, 1064– 1069, DOI: 10.1021/jp983625gGoogle Scholar33https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXntlahtQ%253D%253D&md5=d45c7c07622b0487a5507b9b5171af00Stochastic Dynamics in Irreversible Nonequilibrium Environments. 1. The Fluctuation-Dissipation RelationHernandez, Rigoberto; Somer, Frank L.,Jr.Journal of Physical Chemistry B (1999), 103 (7), 1064-1069CODEN: JPCBFK; ISSN:1089-5647. (American Chemical Society)A generalization of the generalized Langevin equation (stochastic dynamics) is introduced in order to model chem. reactions which take place in changing environments. The friction kernel representing the solvent response is given a nonstationary form with respect to which the instantaneous random solvent force satisfies a natural generalization of the fluctuation-dissipation relation. Theor. considerations, as well as numerical simulations, show that the dynamics of this construction satisfy the equipartition theorem beyond its equil. limits.
- 34Hernandez, R.; Somer, F. L. Stochastic dynamics in irreversible nonequilibrium environments. 2. A model for thermosetting polymerization. J. Phys. Chem. B 1999, 103, 1070– 1077, DOI: 10.1021/jp9836269Google Scholar34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXntlahug%253D%253D&md5=3474745581fe4d6fa7cd3014b0842d2dStochastic Dynamics in Irreversible Nonequilibrium Environments. 2. A Model for Thermosetting PolymerizationHernandez, Rigoberto; Somer, Frank L. Jr.Journal of Physical Chemistry B (1999), 103 (7), 1070-1077CODEN: JPCBFK; ISSN:1089-5647. (American Chemical Society)A generalization of the generalized Langevin equation (GLE), the so-called irreversible GLE (iGLE) [Hernandez, R.; Somer, F. L. J. Phys. Chem. 1999, 103, XXXX], is further extended to describe non-stationary environments in which the non-stationarity is induced by the macroscopic behavior of the ensemble itself, rather than an external force. Such a formalism lends itself to the dynamical study of the length distributions of growing polymers.
- 35Hernandez, R. The projection of a mechanical system onto the irreversible generalized Langevin equation (iGLE). J. Chem. Phys. 1999, 111, 7701– 7704, DOI: 10.1063/1.480160Google Scholar35https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXms1eqt74%253D&md5=8e9c710fe60dbe955bebf0d365c768a3The projection of a mechanical system onto the irreversible generalized Langevin equationHernandez, RigobertoJournal of Chemical Physics (1999), 111 (17), 7701-7704CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The so-called irreversible generalized Langevin equation [R. Hernandez and F. L. Somer, J. Phys. Chem. B 103, 1064 (1999)], which extends the generalized Langevin equation (stochastic dynamics) to include irreversible changes (nonstationarity) in the solvent response, is shown to be the projection of an explicit time-dependent Hamiltonian system.
- 36Locker, C. R.; Hernandez, R. A minimalist model protein with multiple folding funnels. Proc. Natl. Acad. Sci. U. S. A. 2001, 98, 9074– 9079, DOI: 10.1073/pnas.161438898Google Scholar36https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXlvFSkur8%253D&md5=f03eccf584b758291b24d5ebb39f76aaA minimalist model protein with multiple folding funnelsLocker, C. Rebecca; Hernandez, RigobertoProceedings of the National Academy of Sciences of the United States of America (2001), 98 (16), 9074-9079CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)Kinetic and structural studies of wild-type proteins such as prions and amyloidogenic proteins provide suggestive evidence that proteins may adopt multiple long-lived states in addn. to the native state. All of these states differ structurally because they lie far apart in configuration space, but their stability is not necessarily caused by cooperative (nucleation) effects. In this study, a minimalist model protein is designed to exhibit multiple long-lived states to explore the dynamics of the corresponding wild-type proteins. The minimalist protein is modeled as a 27-monomer sequence confined to a cubic lattice with three different monomer types. An order parameter-the winding index-is introduced to characterize the extent of folding. The winding index has several advantages over other commonly used order parameters like the no. of native contacts. It can distinguish between enantiomers, its calcn. requires less computational time than the no. of native contacts, and reduced-dimensional landscapes can be developed when the native state structure is not known a priori. The results for the designed model protein prove by existence that the rugged energy landscape picture of protein folding can be generalized to include protein "misfolding" into long-lived states.
- 37Servos, J. W. Physical Chemistry from Ostwald to Pauling: The Making of a Science in America; Princeton University Press, 1990.Google ScholarThere is no corresponding record for this reference.
- 38Hernandez, R. 6. Blurring of Physical Chemistry and Chemical Physics. 2013; http://everywherechemistry.blogspot.com/2013/08/6-blurring-of-physical-chemistry-and.html, accessed January 15, 2021.Google ScholarThere is no corresponding record for this reference.
- 39Schatz, G. C. Celebrating Our 120th Anniversary. J. Phys. Chem. A 2016, 120, 9679– 9681, DOI: 10.1021/acs.jpca.6b10935Google Scholar39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XitVenurnI&md5=22a9dc1151a1bc0877aaae145060ff55Celebrating Our 120th AnniversarySchatz, George C.Journal of Physical Chemistry A (2016), 120 (49), 9679-9681CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)There is no expanded citation for this reference.
- 40Bartsch, T.; Hernandez, R.; Uzer, T. Transition state in a noisy environment. Phys. Rev. Lett. 2005, 95, 058301, DOI: 10.1103/PhysRevLett.95.058301Google Scholar40https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXmvF2gsLc%253D&md5=9e8e781defda79067a698661b4e07907Transition State in a Noisy EnvironmentBartsch, Thomas; Hernandez, Rigoberto; Uzer, T.Physical Review Letters (2005), 95 (5), 058301/1-058301/4CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)Transition state theory overestimates reaction rates in soln. because conventional dividing surfaces between reagents and products are crossed many times by the same reactive trajectory. We describe a recipe for constructing a time-dependent dividing surface free of such recrossings in the presence of noise. The no-recrossing limit of transition state theory thus becomes generally available for the description of reactions in a fluctuating environment.
- 41Shukla, C.; Hallett, J. P.; Popov, A. V.; Hernandez, R.; Liotta, C.; Eckert, C. Molecular Dynamics Simulation of the Cybotactic Region in Gas-Expanded Methanol-Carbon Dioxide and Acetone-Carbon Dioxide Mixtures. J. Phys. Chem. B 2006, 110, 24101– 24111, DOI: 10.1021/jp0648947Google Scholar41https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XhtFCltLvP&md5=d6917e698291b04997a83f17183393abMolecular Dynamics Simulation of the Cybotactic Region in Gas-Expanded Methanol-Carbon Dioxide and Acetone-Carbon Dioxide MixturesShukla, Charu L.; Hallett, Jason P.; Popov, Alexander V.; Hernandez, Rigoberto; Liotta, Charles L.; Eckert, Charles A.Journal of Physical Chemistry B (2006), 110 (47), 24101-24111CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)Local solvation and transport effects in gas-expanded liqs. (GXLs) are reported based on mol. simulation. GXLs were found to exhibit local d. enhancements similar to those seen in supercrit. fluids, although less dramatic. This approach was used as an alternative to a multiphase atomistic model for these mixts. by utlilizing exptl. results to describe the necessary fixed conditions for a locally (quasi-) stable mol. dynamics model of the (single) GXL phase. The local anisotropic pair correlation function, orientational correlation functions, and diffusion rates are reported for two systems: CO2-expanded methanol and CO2-expanded acetone at 298 K and pressures up to 6 MPa.
- 42Hallett, J. P.; Kitchens, C. L.; Hernandez, R.; Liotta, C.; Eckert, C. Probing the Cybotactic Region in Gas-Expanded Liquids (GXLs). Acc. Chem. Res. 2006, 39, 531– 538, DOI: 10.1021/ar0501424Google Scholar42https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XlsFaktr8%253D&md5=070ffed95f106060adbb9a3fb26211bdProbing the Cybotactic Region in Gas-Expanded Liquids (GXLs)Hallett, Jason P.; Kitchens, Christopher L.; Hernandez, Rigoberto; Liotta, Charles L.; Eckert, Charles A.Accounts of Chemical Research (2006), 39 (8), 531-538CODEN: ACHRE4; ISSN:0001-4842. (American Chemical Society)Gas-expanded liqs. (GXLs) are a new and benign class of liq. solvents, which may offer many advantages for sepns., reactions, and advanced materials. GXLs are intermediate in properties between normal liqs. and supercrit. fluids, both in solvating power and in transport properties. Other advantages include benign nature, low operating pressures, and highly tunable properties by simple pressure variations. The chem. community has only just begun to exploit the advantages of these GXLs for industrial applications. This account focuses on the synergism of exptl. techniques with theor. modeling resulting in a powerful combination for exploring chem. structure and transport in the cybotactic region of GXLs (at the nanometer lengthscale).
- 43Stockard, J.; Rohlfing, C. M.; Richmond, G. L. Equity for women and underrepresented minorities in STEM: Graduate experiences and career plans in chemistry. Proc. Natl. Acad. Sci. U. S. A. 2021, 118, e2020508118, DOI: 10.1073/pnas.2020508118Google Scholar43https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXitFaisL8%253D&md5=ada69c53528b1bf08e5d0de857fb553fEquity for women and underrepresented minorities in STEM: Graduate experiences and career plans in chemistryStockard, Jean; Rohlfing, Celeste M.; Richmond, Geraldine L.Proceedings of the National Academy of Sciences of the United States of America (2021), 118 (4), e2020508118CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)Recent events prompted scientists in the United States and throughout the world to consider how systematic racism affects the scientific enterprise. This paper provides evidence of inequities related to race-ethnicity and gender in graduate school experiences and career plans of PhD students in the top 100 ranked departments in one science, technol., engineering, and math (STEM) discipline, chem. Mixed-model regression analyses were used to examine factors that might moderate these differences. The results show that graduate students who identified as a member of a racial/ethnic group traditionally underrepresented in chem. (underrepresented minorities, URM) were significantly less likely than other students to report that their financial support was sufficient to meet their needs. They were also less likely to report having supportive relationships with peers and postdocs. Women, and esp. URM women, were significantly less likely to report supportive relationships with advisors. Despite their more neg. experiences in graduate school, students who identified as URM expressed greater commitment to finishing their degree and staying in the field. When there was at least one faculty member within their departments who also identified as URM they were also more likely than other students to aspire to a university professorship with an emphasis on research. Men were significantly more likely than women to express strong commitment to finishing the PhD and remaining in chem., but this difference was stronger in top-ranked departments. Men were also more likely than women to aspire to a professorship with an emphasis on research, and this difference remained when individual and departmental-level variables were controlled.
- 44Bowman, J. M. Autobiography of Joel M. Bowman. J. Phys. Chem. A 2013, 117, 6907– 6909, DOI: 10.1021/jp405529pGoogle ScholarThere is no corresponding record for this reference.
- 45Tucker, S. C. Solvent Density Inhomogeneities in Supercritical Fluids. Chem. Rev. 1999, 99, 391– 418, DOI: 10.1021/cr9700437Google Scholar45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXjt1Oquw%253D%253D&md5=6a00e344f9f2c3fd7c6d571935a79efaSolvent Density Inhomogeneities in Supercritical FluidsTucker, Susan C.Chemical Reviews (Washington, D. C.) (1999), 99 (2), 391-418CODEN: CHREAY; ISSN:0009-2665. (American Chemical Society)A review, with ∼184 refs., where fundamentals, evidence of local d. inhomogeneities, and addnl. characteristics of local d. inhomogeneities were discussed.
- 46Locker, C. R.; Hernandez, R. Folding behavior of model proteins with weak energetic frustration. J. Chem. Phys. 2004, 120, 11292– 11303, DOI: 10.1063/1.1751394Google Scholar46https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXksVCqurc%253D&md5=9f12cdcc8c3ef9f3115a3e21f1e224a6Folding behavior of model proteins with weak energetic frustrationLocker, C. Rebecca; Hernandez, RigobertoJournal of Chemical Physics (2004), 120 (23), 11292-11303CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The native structure of fast-folding proteins, albeit a deep local free-energy min., may involve a relatively small energetic penalty due to nonoptimal, though favorable, contacts between amino acid residues. The weak energetic frustration that such contacts represent varies among different proteins and may account for folding behavior not seen in unfrustrated models. Minimalist model proteins with heterogeneous contacts-as represented by lattice heteropolymers consisting of three types of monomers-also give rise to weak energetic frustration in their corresponding native structures, and the present study of their equil. and nonequil. properties reveals some of the breadth in their behavior. In order to capture this range within a detailed study of only a few proteins, four candidate protein structures (with their cognate sequences) have been selected according to a figure of merit called the winding index-a characteristic of the no. of turns the protein winds about an axis. The temp.-dependent heat capacities reveal a high-temp. collapse transition, and an infrequently obsd. low-temp. rearrangement transition that arises because of the presence of weak energetic frustration. Simulation results motivate the definition of a new measure of folding affinity as a sequence-dependent free energy-a function of both a reduced stability gap and high accessibility to non-native structures-that correlates strongly with folding rates.
- 47Leite, V. B. P.; Alonso, L. C. P.; Newton, M.; Wang, J. Single Molecule Electron Transfer Dynamics in Complex Environments. Phys. Rev. Lett. 2005, 95, 118301, DOI: 10.1103/PhysRevLett.95.118301Google Scholar47https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXpvFGktrc%253D&md5=b52f959f5825f207ddea56311f94e8e7Single Molecule Electron Transfer Dynamics in Complex EnvironmentsLeite, Vitor B. P.; Alonso, Luciana C. P.; Newton, Marshall; Wang, JinPhysical Review Letters (2005), 95 (11), 118301/1-118301/4CODEN: PRLTAO; ISSN:0031-9007. (American Physical Society)We propose a new theor. approach to study the kinetics of the electron transfer (ET) under the dynamical influence of the complex environments with the first passage times (FPT) of the reaction events. By measuring the mean and high order moments of FPT and their ratios, the full kinetics of ET, esp. the dynamical transitions across different temp. zones, is revealed. The potential applications of the current results to single mol. electron transfer are discussed.
- 48Sension, R.; Tokmakoff, A. Proceedings of “Optical Probes of Dynamics in Complex Environments”; OSTI, 2008; https://www.osti.gov/servlets/purl/1062182, accessed January 15, 2021.Google ScholarThere is no corresponding record for this reference.
- 49Virshup, A. M.; Punwong, C.; Pogorelov, T. V.; Lindquist, B. A.; Ko, C.; Martínez, T. J. Photodynamics in Complex Environments: Ab Initio Multiple Spawning Quantum Mechanical/Molecular Mechanical Dynamics. J. Phys. Chem. B 2009, 113, 3280– 3291, DOI: 10.1021/jp8073464Google Scholar49https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXhsFSqsLvL&md5=e0f0c87e8ccf752578878b4256ee89e3Photodynamics in Complex Environments: Ab Initio Multiple Spawning Quantum Mechanical/Molecular Mechanical DynamicsVirshup, Aaron M.; Punwong, Chutintorn; Pogorelov, Taras V.; Lindquist, Beth A.; Ko, Chaehyuk; Martinez, Todd J.Journal of Physical Chemistry B (2009), 113 (11), 3280-3291CODEN: JPCBFK; ISSN:1520-6106. (American Chemical Society)Our picture of reactions on electronically excited states has evolved considerably in recent years, due to advances in our understanding of points of degeneracy between different electronic states, termed "conical intersections" (CIs). CIs serve as funnels for population transfer between different electronic states, and play a central role in ultrafast photochem. Because most practical photochem. occurs in soln. and protein environments, it is important to understand the role complex environments play in directing excited-state dynamics generally, as well as specific environmental effects on CI geometries and energies. In order to model such effects, we employ the full multiple spawning (FMS) method for multistate quantum dynamics, together with hybrid quantum mech./mol. mech. (QM/MM) potential energy surfaces using both semiempirical and ab initio QM methods. In this article, we present an overview of these methods, and a comparison of the excited-state dynamics of several biol. chromophores in solvent and protein environments. Aq. solvation increases the rate of quenching to the ground state for both the photoactive yellow protein (PYP) and green fluorescent protein (GFP) chromophores, apparently by energetic stabilization of their resp. CIs. In contrast, solvation in methanol retards the quenching process of the retinal protonated Schiff base (RPSB), the rhodopsin chromophore. Protein environments serve to direct the excited-state dynamics, leading to higher quantum yields and enhanced reaction selectivity.
- 50Hernandez, R.; Popov, A. Molecular dynamics out of equilibrium: Mechanics and measurables. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2014, 4, 541– 561, DOI: 10.1002/wcms.1190Google Scholar50https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhvVelu7jK&md5=83f3cee550944e4259712b398279c52eMolecular dynamics out of equilibrium: mechanics and measurablesHernandez, Rigoberto; Popov, Alexander V.Wiley Interdisciplinary Reviews: Computational Molecular Science (2014), 4 (6), 541-561CODEN: WIRCAH; ISSN:1759-0884. (Wiley-Blackwell)Mol. dynamics is fundamentally the integration of the equations of motion over a representation of an at. and mol. system. The most rigorous choice for performing mol. dynamics entails the use of quantum-mech. equations of motion and a representation of the mol. system through all of its electrons and atoms. For most mol. problems involving at least hundreds of atoms, but generally many more, this is simply computationally prohibitive. Thus the art of mol. dynamics lies in choosing the representation and the appropriate equations of motion capable of addressing the requisite measurables. When used adroitly, it can provide both equil. (averaged) and time-dependent properties of a mol. system. Many computational packages now exist that perform mol. dynamics simulations. They generally include force fields to represent the interactions between atoms and mols. (smoothing out electrons through the Born-Oppenheimer approxn.) and integrate the remaining particles classically. Despite these simplifications, all-atom mol. dynamics remains computationally inaccessible if one includes the no. of atoms required to simulate mesoscopic solvents. Here we use anal. models to demonstrate how mol. dynamics can be used to limit the solvent size in systems experiencing either equil. or nonequil. conditions. It is equally important to address the measurables (such as reaction rates) that are to be obtained prior to the generation of the data-intensive trajectories. WIREs Comput Mol Sci 2014, 4:541-561. doi: 10.1002/wcms.1190 For further resources related to this article, please visit the . Conflict of interest: The authors have declared no conflicts of interest for this article.
- 51Press release: Nobel Prize in Chemistry 2013. https://www.nobelprize.org/prizes/chemistry/2013/press-release/, accessed January 15, 2021.Google ScholarThere is no corresponding record for this reference.
- 52MacKerell, A. D., Jr. All-atom empirical potential for molecular modeling and dynamics studies of proteins. J. Phys. Chem. B 1998, 102, 3586– 3616, DOI: 10.1021/jp973084fGoogle Scholar52https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXivVOlsb4%253D&md5=ebb5100dafd0daeee60ca2fa66c1324aAll-Atom Empirical Potential for Molecular Modeling and Dynamics Studies of ProteinsMacKerell, A. D., Jr.; Bashford, D.; Bellott, M.; Dunbrack, R. L.; Evanseck, J. D.; Field, M. J.; Fischer, S.; Gao, J.; Guo, H.; Ha, S.; Joseph-McCarthy, D.; Kuchnir, L.; Kuczera, K.; Lau, F. T. K.; Mattos, C.; Michnick, S.; Ngo, T.; Nguyen, D. T.; Prodhom, B.; Reiher, W. E., III; Roux, B.; Schlenkrich, M.; Smith, J. C.; Stote, R.; Straub, J.; Watanabe, M.; Wiorkiewicz-Kuczera, J.; Yin, D.; Karplus, M.Journal of Physical Chemistry B (1998), 102 (18), 3586-3616CODEN: JPCBFK; ISSN:1089-5647. (American Chemical Society)New protein parameters are reported for the all-atom empirical energy function in the CHARMM program. The parameter evaluation was based on a self-consistent approach designed to achieve a balance between the internal (bonding) and interaction (nonbonding) terms of the force field and among the solvent-solvent, solvent-solute, and solute-solute interactions. Optimization of the internal parameters used exptl. gas-phase geometries, vibrational spectra, and torsional energy surfaces supplemented with ab initio results. The peptide backbone bonding parameters were optimized with respect to data for N-methylacetamide and the alanine dipeptide. The interaction parameters, particularly the at. charges, were detd. by fitting ab initio interaction energies and geometries of complexes between water and model compds. that represented the backbone and the various side chains. In addn., dipole moments, exptl. heats and free energies of vaporization, solvation and sublimation, mol. vols., and crystal pressures and structures were used in the optimization. The resulting protein parameters were tested by applying them to noncyclic tripeptide crystals, cyclic peptide crystals, and the proteins crambin, bovine pancreatic trypsin inhibitor, and carbonmonoxy myoglobin in vacuo and in a crystal. A detailed anal. of the relationship between the alanine dipeptide potential energy surface and calcd. protein φ, χ angles was made and used in optimizing the peptide group torsional parameters. The results demonstrate that use of ab initio structural and energetic data by themselves are not sufficient to obtain an adequate backbone representation for peptides and proteins in soln. and in crystals. Extensive comparisons between mol. dynamics simulation and exptl. data for polypeptides and proteins were performed for both structural and dynamic properties. Calcd. data from energy minimization and dynamics simulations for crystals demonstrate that the latter are needed to obtain meaningful comparisons with exptl. crystal structures. The presented parameters, in combination with the previously published CHARMM all-atom parameters for nucleic acids and lipids, provide a consistent set for condensed-phase simulations of a wide variety of mols. of biol. interest.
- 53Mackerell, A. D., Jr.; Feig, M.; Brooks, C. L., III Extending the treatment of backbone energetics in protein force fields: limitations of gas-phase quantum mechanics in reproducing protein conformational distributions in molecular dynamics simulations. J. Comput. Chem. 2004, 25, 1400– 1415, DOI: 10.1002/jcc.20065Google Scholar53https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXlsVOgt7c%253D&md5=b2451bb5df548447f8b172a211bc1848Extending the treatment of backbone energetics in protein force fields: Limitations of gas-phase quantum mechanics in reproducing protein conformational distributions in molecular dynamics simulationsMacKerell, Alexander D., Jr.; Feig, Michael; Brooks, Charles L., IIIJournal of Computational Chemistry (2004), 25 (11), 1400-1415CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)Computational studies of proteins based on empirical force fields represent a powerful tool to obtain structure-function relationships at an at. level, and are central in current efforts to solve the protein folding problem. The results from studies applying these tools are, however, dependent on the quality of the force fields used. In particular, accurate treatment of the peptide backbone is crucial to achieve representative conformational distributions in simulation studies. To improve the treatment of the peptide backbone, quantum mech. (QM) and mol. mech. (MM) calcns. were undertaken on the alanine, glycine, and proline dipeptides, and the results from these calcns. were combined with mol. dynamics (MD) simulations of proteins in crystal and aq. environments. QM potential energy maps of the alanine and glycine dipeptides at the LMP2/cc-pVxZ/MP2/6-31G* levels, where x = D, T, and Q, were detd., and are compared to available QM studies on these mols. The LMP2/cc pVQZ//MP2/6-31G* energy surfaces for all three dipeptides were then used to improve the MM treatment of the dipeptides. These improvements included addnl. parameter optimization via Monte Carlo simulated annealing and extension of the potential energy function to contain peptide backbone .vphi., ψ dihedral crossterms or a .vphi., ψ grid-based energy correction term. Simultaneously, MD simulations of up to seven proteins in their cryst. environments were used to validate the force field enhancements. Comparison with QM and crystallog. data showed that an addnl. optimization of the .vphi., ψ dihedral parameters along with the grid-based energy correction were required to yield significant improvements over the CHARMM22 force field. However, systematic deviations in the treatment of .vphi. and ψ in the helical and sheet regions were evident. Accordingly, empirical adjustments were made to the grid-based energy correction for alanine and glycine to account for these systematic differences. These adjustments lead to greater deviations from QM data for the two dipeptides but also yielded improved agreement with exptl. crystallog. data. These improvements enhance the quality of the CHARMM force field in treating proteins. This extension of the potential energy function is anticipated to facilitate improved treatment of biol. macromols. via MM approaches in general.
- 54Kalé, L.; Skeel, R.; Bhandarkar, M.; Brunner, R.; Gursoy, A.; Krawetz, N.; Phillips, J.; Shinozaki, A.; Varadarajan, K.; Schulten, K. NAMD2: Greater scalability for parallel molecular dynamics. J. Comput. Phys. 1999, 151, 283– 312, DOI: 10.1006/jcph.1999.6201Google Scholar54https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXivFejt7Y%253D&md5=f40c0fc219c6fef216fae5f0dc8c9003NAMD2: Greater Scalability for Parallel Molecular DynamicsKale, Laxmikant; Skeel, Robert; Bhandarkar, Milind; Brunner, Robert; Gursoy, Attila; Krawetz, Neal; Phillips, James; Shinozaki, Aritomo; Varadarajan, Krishnan; Schulten, KlausJournal of Computational Physics (1999), 151 (1), 283-312CODEN: JCTPAH; ISSN:0021-9991. (Academic Press)Mol. dynamics programs simulate the behavior of biomol. systems, leading to understanding of their functions. However, the computational complexity of such simulations is enormous. Parallel machines provide the potential to meet this computational challenge. To harness this potential, it is necessary to develop a scalable program. It is also necessary that the program be easily modified by application-domain programmers. The NAMD2 program presented in this paper seeks to provide these desirable features. It uses spatial decompn. combined with force decompn. to enhance scalability. It uses intelligent periodic load balancing, so as to maximally utilize the available compute power. It is modularly organized, and implemented using Charm++, a parallel C++ dialect, so as to enhance its modifiability. It uses a combination of numerical techniques and algorithms to ensure that energy drifts are minimized, ensuring accuracy in long running calcns. NAMD2 uses a portable run-time framework called Converse that also supports interoperability among multiple parallel paradigms. As a result, different components of applications can be written in the most appropriate parallel paradigms. NAMD2 runs on most parallel machines including workstation clusters and has yielded speedups in excess of 180 on 220 processors. This paper also describes the performance obtained on some benchmark applications. (c) 1999 Academic Press.
- 55Pearlman, D. A.; Case, D. A.; Caldwell, J.; Ross, W. R.; Cheatham, T. E., III; DeBolt, S.; Ferguson, D.; Seibel, G.; Kollman, P. AMBER, a computer program for applying molecular mechanics, normal mode analysis, molecular dynamics and free energy calculations to elucidate the structures and energies of molecules. Comput. Phys. Commun. 1995, 91, 1– 41, DOI: 10.1016/0010-4655(95)00041-DGoogle Scholar55https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXps1Wrtrw%253D&md5=8dc71939a46bbc17d5da08782d4e6ec8"AMBER", a package of computer programs for applying molecular mechanics, normal mode analysis, molecular dynamics and free energy calculations to stimulate the structural and energetic properties of moleculesPearlman, David A.; Case, David A.; Caldwell, James W.; Ross, Wilson S.; Cheatham, Thomas E. III; DeBolt, Steve; Ferguson, David; Seibel, George; Kollman, PeterComputer Physics Communications (1995), 91 (1-3), 1-42CODEN: CPHCBZ; ISSN:0010-4655. (Elsevier)We describe the development, current features, and some directions for future development of the AMBER package of computer programs. This package has evolved from a program that was constructed to do Assisted Model Building and Energy Refinement to a group of programs embodying a no. of the powerful tools of modern computational chem.-mol. dynamics and free energy calcns.
- 56Brown, W. M.; Petersen, M. K.; Plimpton, S. J.; Grest, G. S. Liquid crystal nanodroplets in solution. J. Chem. Phys. 2009, 130, 044901, DOI: 10.1063/1.3058435Google Scholar56https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhtVWntbg%253D&md5=fd0b4c29e8994e5da4aa8eec53178918Liquid crystal nanodroplets in solutionBrown, W. Michael; Petersen, Matt K.; Plimpton, Steven J.; Grest, Gary S.Journal of Chemical Physics (2009), 130 (4), 044901/1-044901/7CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The aggregation of liq. crystal nanodroplets from a homogeneous soln. is studied by mol. dynamics simulations. The liq. crystal particles are modeled as elongated ellipsoidal Gay-Berne particles while the solvent is modeled as spherical Lennard-Jones particles. Extending previous studies of , we find that liq. crystal nanodroplets are not stable and that after sufficiently long times the nanodroplets always aggregate into a single large droplet. Results describing the droplet shape and orientation for different temps. and shear rates are presented. The implementation of the Gay-Berne potential for biaxial ellipsoidal particles in a parallel mol. dynamics code is also briefly discussed. (c) 2009 American Institute of Physics.
- 57Durrant, J. D.; Kochanek, S. E.; Casalino, L.; Ieong, P. U.; Dommer, A. C.; Amaro, R. E. Mesoscale All-Atom Influenza Virus Simulations Suggest New Substrate Binding Mechanism. ACS Cent. Sci. 2020, 6, 189– 196, DOI: 10.1021/acscentsci.9b01071Google Scholar57https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXjtFKhsL0%253D&md5=84c3498ba325642e2389c06d4c118de5Mesoscale All-Atom Influenza Virus Simulations Suggest New Substrate Binding MechanismDurrant, Jacob D.; Kochanek, Sarah E.; Casalino, Lorenzo; Ieong, Pek U.; Dommer, Abigail C.; Amaro, Rommie E.ACS Central Science (2020), 6 (2), 189-196CODEN: ACSCII; ISSN:2374-7951. (American Chemical Society)Influenza virus circulates in human, avian, and swine hosts, causing seasonal epidemic and occasional pandemic outbreaks. Influenza neuraminidase, a viral surface glycoprotein, has two sialic acid binding sites. The catalytic (primary) site, which also binds inhibitors such as oseltamivir carboxylate, is responsible for cleaving the sialic acid linkages that bind viral progeny to the host cell. In contrast, the functional annotation of the secondary site remains unclear. Here, we better characterize these two sites through the development of an all-atom, explicitly solvated, exptl. based integrative model of the pandemic influenza A H1N1 2009 viral envelope, contg. ∼160 million atoms and spanning ∼115 nm in diam. Mol. dynamics simulations of this crowded subcellular environment, coupled with Markov state model theory, provide a novel framework for studying realistic mol. systems at the mesoscale and allow us to quantify the kinetics of the 150-loop transition between the open and closed states. An anal. of chloride ion occupancy along the neuraminidase surface implies a potential new role for the neuraminidase secondary site, wherein the terminal sialic acid residues of the linkages may bind before transfer to the primary site where enzymic cleavage occurs. Altogether, our work breaks new ground for mol. simulation in terms of the size, complexity, and methodol. analyses of the simulated components, as well as provides fundamental insights into the understanding of substrate recognition processes for this vital influenza drug target, suggesting a new strategy for the development of anti-influenza therapeutics.
- 58Hagy, M. C.; Hernandez, R. Dynamical simulation of dipolar Janus colloids: Dynamical properties. J. Chem. Phys. 2013, 138, 184903, DOI: 10.1063/1.4803864Google Scholar58https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXnsVyrsLw%253D&md5=f4d2faca35479eab47b17f30643d245dDynamical simulation of dipolar Janus colloids: Dynamical propertiesHagy, Matthew C.; Hernandez, RigobertoJournal of Chemical Physics (2013), 138 (18), 184903/1-184903/10CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The dynamical properties of dipolar Janus particles are studied through simulation using our previously-developed detailed pointwise (PW) model and an isotropically coarse-grained (CG) model. The CG model is found to have accelerated dynamics relative to the PW model over a range of conditions for which both models have near identical static equil. properties. Phys., this suggests dipolar Janus particles have slower transport properties (such as diffusion) in comparison to isotropically attractive particles. Time rescaling and damping with Langevin friction are explored to map the dynamics of the CG model to that of the PW model. Both methods map the diffusion const. successfully and improve the velocity autocorrelation function and the mean squared displacement of the CG model. Neither method improves the distribution of reversible bond durations f(tb) obsd. in the CG model, which is found to lack the longer duration reversible bonds obsd. in the PW model. We attribute these differences in f(tb) to changes in the energetics of multiple rearrangement mechanisms. This suggests a need for new methods that map the coarse-grained dynamics of such systems to the true time scale. (c) 2013 American Institute of Physics.
- 59Downey-Mavromatis, A.; Widener, A. Chemistry faculty’s diversity has changed little since 2011. Chem. Eng. News 2020, 98, 22– 25, DOI: 10.1021/cen-09843-feature2Google Scholar59https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXit12ju7fF&md5=c06a7f4abf3ff20ec34c69dcb7cc7575Chemistry faculty's diversity has changed little since 2011Downey-Mavromatis, ArmindaChemical & Engineering News (2020), 98 (43), 22-25CODEN: CENEAR; ISSN:1520-605X. (American Chemical Society)For the past decade, C&EN has worked with the Open Chem. Collaborative in Diversity Equity (OXIDE) to publish statistics on gender and racial and ethnic diversity of US chem. faculty. OXIDE collects the data, a task that's not as easy as it might sound, its leaders say. Nevertheless, it's important work, esp. with the increased focus from universities on campus diversity efforts after protests about racism and police brutality against Black people this year. "This is definitely a moment of reflection for our nation-and one cannot effectively reflect without knowing the facts," says Dontarie Stallings, assoc. director of OXIDE and a teaching professor of chem. at the University of California San Diego. The current facts: the racial and ethnic diversity of US chem. faculty has changed little since 2011. OXIDE began documenting diversity among chem. faculty because although some universities survey their faculty and students to understand their diversity and OXIDE hopes data will help departments improve faculty's diversity, inclusion, and equity 10.47287/cen-09843-feature2-meeting-gr1Rigoberto Hernandez (left) and Dontarie Stallings at an American Chem. Society meeting in 2016 (Credit: Courtesy of Rigoberto Hernandez) oxidesurveyRigoberto HernandezDontarie StallingsA photo of Rigoberto Hernandez and Dontarie Stallings.
- 60Hernandez, R. The private sector’s role in chemistry’s future. Chem. Eng. News 2015, 93, 33, DOI: 10.1021/cen-09337-commentGoogle ScholarThere is no corresponding record for this reference.
- 61Hernandez, R. OneChemistry in the marketplace of ideas. Chem. Eng. News 2017, 95, 41, DOI: 10.1021/cen-09517-commentGoogle ScholarThere is no corresponding record for this reference.
- 62Craven, G. T.; Junginger, A.; Hernandez, R. Lagrangian descriptors of driven chemical reaction manifolds. Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top. 2017, 96, 022222, DOI: 10.1103/PhysRevE.96.022222Google ScholarThere is no corresponding record for this reference.
- 63Schraft, P.; Junginger, A.; Feldmaier, M.; Bardakcioglu, R.; Main, J.; Wunner, G.; Hernandez, R. Neural network approach to time-dependent dividing surfaces in classical reaction dynamics. Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top. 2018, 97, 042309, DOI: 10.1103/PhysRevE.97.042309Google ScholarThere is no corresponding record for this reference.
- 64Feldmaier, M.; Reiff, J.; Benito, R. M.; Borondo, F.; Main, J.; Hernandez, R. Influence of external driving on decays in the geometry of the LiCN isomerization. J. Chem. Phys. 2020, 153, 084115, DOI: 10.1063/5.0015509Google Scholar64https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXhslWqtr7F&md5=8e4c634fd35a0574038dee134be46e23Influence of external driving on decays in the geometry of the LiCN isomerizationFeldmaier, Matthias; Reiff, Johannes; Benito, Rosa M.; Borondo, Florentino; Main, Joerg; Hernandez, RigobertoJournal of Chemical Physics (2020), 153 (8), 084115CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The framework of transition state theory relies on the detn. of a geometric structure identifying reactivity. It replaces the laborious exercise of following many trajectories for a long time to provide chem. reaction rates and pathways. In this paper, recent advances in constructing this geometry even in time-dependent systems are applied to the LiCN .dblharw. LiNC isomerization reaction driven by an external field. We obtain decay rates of the reactant population close to the transition state by exploiting local properties of the dynamics of trajectories in and close to it. We find that the external driving has a large influence on these decay rates when compared to the non-driven isomerization reaction. This, in turn, provides renewed evidence for the possibility of controlling chem. reactions, like this one, through external time-dependent fields. (c) 2020 American Institute of Physics.
- 65Bureau, H.; Quirk, S.; Hernandez, R. The relative stability of trpzip1 and its mutants determined by computation and experiment. RSC Adv. 2020, 10, 6520, DOI: 10.1039/D0RA00920BGoogle Scholar65https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXjtVeisbw%253D&md5=47af55f3a988711d077a6af9263cc27aThe relative stability of trpzip1 and its mutants determined by computation and experimentBureau, Hailey R.; Quirk, Stephen; Hernandez, RigobertoRSC Advances (2020), 10 (11), 6520-6535CODEN: RSCACL; ISSN:2046-2069. (Royal Society of Chemistry)Six mutants of the tryptophan zipper peptide trpzip1 have been computationally and exptl. characterized. We det. the varying roles in secondary structure stability of specific residues through a mutation assay. Four of the mutations directly effect the Trp-Trp interactions and two of the mutations target the salt bridge between Glu5 and Lys8. CD spectra and thermal unfolding are used to det. the secondary structure and stability of the mutants compared to the wildtype peptide. Adaptive steered mol. dynamics has been used to obtain the energetics of the unfolding pathways of the mutations. The hydrogen bonding patterns and side-chain interactions over the course of unfolding have also been calcd. and compared to wildtype trpzip1. The key finding from this work is the importance of a stabilizing non-native salt bridge pair present in the K8L mutation.
- 66Zhuang, Y.; Bureau, H.; Quirk, S.; Hernandez, R. Adaptive Steered Molecular Dynamics of Biomolecules. Mol. Simul. 2021, 47, 408, DOI: 10.1080/08927022.2020.1807542Google Scholar66https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXhs1Kks7nN&md5=c51977e6a907041213010ccd153ccfb8Adaptive steered molecular dynamics of biomoleculesZhuang, Yi; Bureau, Hailey R.; Quirk, Stephen; Hernandez, RigobertoMolecular Simulation (2021), 47 (5), 408-419CODEN: MOSIEA; ISSN:0892-7022. (Taylor & Francis Ltd.)Adaptive steered mol. dynamics (ASMD) is a variant of steered mol. dynamics (SMD) in which the driving of the auxiliary - viz steered - particle is performed in stages. In SMD, many nonequil. trajectories are generated allowing for fast sampling of the ensemble space and access to rare events. Equil. observables, such as the potential of mean force along the pathway, result from averaging over these trajectories using Jarzynski's Equality (JE). Unfortunately, in SMD, a large no. of trajectories are needed to cover all possible configurations in order to obtain converged quantities in the exponential av. of the JE, and this is computationally expensive. ASMD reduces the no. of trajectories that must be sampled by discarding those trajectories that have deviated far from the equil. path in stages. At the end of a stage, one chooses - or contracts - one (in naive ASMD) or some (in multi-branched ASMD or MB-ASMD) of the configurations produced from the previous stage to initiate the trajectories in the next stage. Alternatively, in full-relaxation ASMD (FR-ASMD), all generated structures are relaxed under the constraint of a fixed auxiliary particle exerting no net work on the system. We provide a direct comparison of the energetics and other observables obtained from these approaches. We find that FR-ASMD is preferred when the unfolding pathways follow up along a single funnel and the system is sufficiently small that computational resources are not a limiting concern. It gives the highest accuracy in such cases while avoiding the inefficiencies of SMD. However, for complex energy landscapes typical of most multi-domain proteins, MB-ASMD is preferred because it provides a mechanism to sample alternative pathways while suffering only a modest loss in accuracy compared to FR-ASMD.
- 67Zhuang, Y.; Bureau, H. R.; Lopez, C.; Bucher, R.; Quirk, S.; Hernandez, R. Energetics and structure of alanine-rich _-helices via Adaptive Steered Molecular Dynamics (ASMD). Biophys. J. 2021, 120, 2009, DOI: 10.1016/j.bpj.2021.03.017Google Scholar67https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXoslGkt74%253D&md5=b8e34535a458f3191440b0477bb2cc6cEnergetics and structure of alanine-rich α-helices via adaptive steered molecular dynamicsZhuang, Yi; Bureau, Hailey R.; Lopez, Christine; Bucher, Ryan; Quirk, Stephen; Hernandez, RigobertoBiophysical Journal (2021), 120 (10), 2009-2018CODEN: BIOJAU; ISSN:0006-3495. (Cell Press)The energetics and hydrogen bonding profiles of the helix-to-coil transition were found to be an additive property and to increase linearly with chain length, resp., in alanine-rich α-helical peptides. A model system of polyalanine repeats was used to establish this hypothesis for the energetic trends and hydrogen bonding profiles. Numerical measurements of a synthesized polypeptide Ac-Y(AEAAKA)kF-NH2 and a natural α-helical peptide a2N (1-17) provide evidence of the hypothesis's generality. Adaptive steered mol. dynamics was employed to investigate the mech. unfolding of all of these alanine-rich polypeptides. We found that the helix-to-coil transition is primarily dependent on the breaking of the intramol. backbone hydrogen bonds and independent of specific side-chain interactions and chain length. The mech. unfolding of the α-helical peptides results in a turnover mechanism in which a 310-helical structure forms during the unfolding, remaining at a near const. population and thereby maintaining additivity in the free energy. The intermediate partially unfolded structures exhibited polyproline II helical structure as previously seen by others. In summary, we found that the av. force required to pull alanine-rich α-helical peptides in between the endpoints-namely the native structure and free coil-is nearly independent of the length or the specific primary structure.
- 68Mahala, B.; Hernandez, R. Solvent softness effects on unimolecular chemical reaction rate constants. Chem. Phys. Lett. 2020, 744, 137182, DOI: 10.1016/j.cplett.2020.137182Google Scholar68https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXjtF2iu70%253D&md5=f44619c6871d6d5799fda42ba532e37cSolvent softness effects on unimolecular chemical reaction rate constantsMahala, Benjamin D.; Hernandez, RigobertoChemical Physics Letters (2020), 744 (), 137182CODEN: CHPLBC; ISSN:0009-2614. (Elsevier B.V.)The stochastic hard collision (SHC) model for a coarse-grained solvent has been used to solvate a class of model chem. reactions. Although there is precedent from our prior work that the use of nondeterministic interaction potentials can lead to accurate dynamics at long-enough time scales, this work provides evidence that this model has dynamical consistency specifically in the response to the reactant motion. That is, the agreement between the theor. rate consts. and those obtained from the stochastic simulations provides support for the SHC solvent without sacrificing the correlations in the dynamics of the reacting solutes.
- 69Murphy, C. J.; Vartanian, A. M.; Geiger, F. M.; Hamers, R. J.; Pedersen, J.; Cui, Q.; Haynes, C. L.; Carlson, E. E.; Hernandez, R.; Klaper, R. D.; Orr, G.; Rosenzweig, Z. Biological responses to engineered nanomaterials: Needs for the next decade. ACS Cent. Sci. 2015, 1, 117, DOI: 10.1021/acscentsci.5b00182Google Scholar69https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhtVaku7vM&md5=35006c55ffaa627f090c5f23d8ca8d8aBiological Responses to Engineered Nanomaterials: Needs for the Next DecadeMurphy, Catherine J.; Vartanian, Ariane M.; Geiger, Franz M.; Hamers, Robert J.; Pedersen, Joel; Cui, Qiang; Haynes, Christy L.; Carlson, Erin E.; Hernandez, Rigoberto; Klaper, Rebecca D.; Orr, Galya; Rosenzweig, Ze'evACS Central Science (2015), 1 (3), 117-123CODEN: ACSCII; ISSN:2374-7951. (American Chemical Society)A review. The interaction of nanomaterials with biomols., cells, and organisms is an enormously vital area of current research, with applications in nanoenabled diagnostics, imaging agents, therapeutics, and contaminant removal technologies. Yet the potential for adverse biol. and environmental impacts of nanomaterial exposure is considerable and needs to be addressed to ensure sustainable development of nanomaterials. In this Outlook four research needs for the next decade are outlined: (i) measurement of the chem. nature of nanomaterials in dynamic, complex aq. environments; (ii) real-time measurements of nanomaterial-biol. interactions with chem. specificity; (iii) delineation of mol. modes of action for nanomaterial effects on living systems as functions of nanomaterial properties; and (iv) an integrated systems approach that includes computation and simulation across orders of magnitude in time and space.
- 70Wu, M.; Vartanian, A. M.; Chong, G.; Pandiakumar, A. K.; Hamers, R. J.; Hernandez, R.; Murphy, C. J. Solution NMR analysis of ligand environment in quaternary ammonium-terminated self-assembled monolayers on gold nanoparticles: The effect of surface curvature and ligand structure. J. Am. Chem. Soc. 2019, 141, 4316, DOI: 10.1021/jacs.8b11445Google Scholar70https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXjtFCrt74%253D&md5=533e812bc895c8d2b559484e647f1d15Solution NMR Analysis of Ligand Environment in Quaternary Ammonium-Terminated Self-Assembled Monolayers on Gold Nanoparticles: The Effect of Surface Curvature and Ligand StructureWu, Meng; Vartanian, Ariane M.; Chong, Gene; Pandiakumar, Arun Kumar; Hamers, Robert J.; Hernandez, Rigoberto; Murphy, Catherine J.Journal of the American Chemical Society (2019), 141 (10), 4316-4327CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)We report a soln. NMR-based anal. of (16-mercaptohexadecyl)trimethylammonium bromide (MTAB) self-assembled monolayers on colloidal gold nanospheres (AuNSs) with diams. from 1.2 to 25 nm and gold nanorods (AuNRs) with aspect ratios from 1.4 to 3.9. The chem. shift anal. of the proton signals from the solvent-exposed headgroups of bound ligands suggests that the headgroups are satd. on the ligand shell as the sizes of the nanoparticles increase beyond ∼10 nm. Quant. NMR shows that the ligand d. of MTAB-AuNSs is size-dependent. Ligand d. ranges from ∼3 mols. per nm2 for 25 nm particles to up to 5-6 mols. per nm2 in ∼10 nm and smaller particles for in situ measurements of bound ligands; after I2/I- treatment to etch away the gold cores, ligand d. ranges from ∼2 mols. per nm2 for 25 nm particles to up to 4-5 mols. per nm2 in ∼10 nm and smaller particles. T2 relaxation anal. shows greater hydrocarbon chain ordering and less headgroup motion as the diam. of the particles increases from 1.2 nm to ∼13 nm. Mol. dynamics simulations of 4, 6, and 8 nm (11-mercaptoundecyl)trimethylammonium bromide-capped AuNSs confirm greater hydrophobic chain packing order and satn. of charged headgroups within the same spherical ligand shell at larger nanoparticle sizes and higher ligand densities. Combining the NMR studies and MD simulations, we suggest that the headgroup packing limits the ligand d., rather than the sulfur packing on the nanoparticle surface, for ∼10 nm and larger particles. For MTAB-AuNRs, no chem. shift data nor ligand d. data suggest that two populations of ligands that might correspond to side-ligands and end-ligands exist; yet T2 relaxation dynamics data suggest that headgroup mobility depends on aspect ratio and abs. nanoparticle dimensions.
- 71Daly, C. A., Jr.; Allen, C. R.; Rozanov, N. D.; Chong, G.; Melby, E. S.; Kuech, T. R.; Lohse, S. E.; Murphy, C. J.; Pedersen, J. A.; Hernandez, R. Surface Coating Structure and Its Interaction with Cytochrome c in EG6-Coated Nanoparticles Varies with Surface Curvature. Langmuir 2020, 36, 5030– 5039, DOI: 10.1021/acs.langmuir.0c00681Google Scholar71https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXnsVCns7g%253D&md5=9e71f49e51279feec665b49fd71e8e55Surface Coating Structure and Its Interaction with Cytochrome c in EG6-Coated Nanoparticles Varies with Surface CurvatureDaly, Clyde A.; Allen, Caley; Rozanov, Nikita; Chong, Gene; Melby, Eric S.; Kuech, Thomas R.; Lohse, Samuel E.; Murphy, Catherine J.; Pedersen, Joel A.; Hernandez, RigobertoLangmuir (2020), 36 (18), 5030-5039CODEN: LANGD5; ISSN:0743-7463. (American Chemical Society)The compn., orientation, and conformation of proteins in biomol. coronas acquired by nanoparticles in biol. media contribute to how they are identified by a cell. While numerous studies have investigated protein compn. in biomol. coronas, relatively little detail is known about how the nanoparticle surface influences the orientation and conformation of the proteins assocd. with them. The authors previously showed that the peripheral membrane protein cytochrome c adopts preferred poses relative to neg. charged 3-mercaptopropionic acid (MPA)-gold nanoparticles (AuNPs). Here, the authors employ mol. dynamics simulations and complementary expts. to establish that cytochrome c also assumes preferred poses upon assocn. with nanoparticles functionalized with an uncharged ligand, specifically ω-(1-mercaptounde-11-cyl)hexa(ethylene glycol) (EG6). The authors find that the display of the EG6 ligands is sensitive to the curvature of the surface-and, consequently, the effective diam. of the nearly spherical nanoparticle core-which in turn affects the preferred poses of cytochrome c.
- 72Bathe, M.; Hernandez, R.; Komiyama, T.; Machiraju, R.; Neogi, S. Autonomous computing materals. ACS Nano 2021, 15, 3586– 3592, DOI: 10.1021/acsnano.0c09556Google Scholar72https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXlt1Shs7k%253D&md5=f02169e68d5a3171fcfaf8b3f25cfa1cAutonomous Computing MaterialsBathe, Mark; Hernandez, Rigoberto; Komiyama, Takaki; Machiraju, Raghu; Neogi, SanghamitraACS Nano (2021), 15 (3), 3586-3592CODEN: ANCAC3; ISSN:1936-0851. (American Chemical Society)A review. Conventional materials are reaching their limits in computation, sensing, and data storage capabilities, ushered in by the end of Moore's law, myriad sensing applications, and the continuing exponential rise in worldwide data storage demand. Conventional materials are also limited by the controlled environments in which they must operate, their high energy consumption, and their limited capacity to perform simultaneous, integrated sensing, computation, and data storage and retrieval. In contrast, the human brain is capable of multimodal sensing, complex computation, and both short- and long-term data storage simultaneously, with near instantaneous rate of recall, seamless integration, and minimal energy consumption. Motivated by the brain and the need for revolutionary new computing materials, we recently proposed the data-driven materials discovery framework, autonomous computing materials. This framework aims to mimic the brain's capabilities for integrated sensing, computation, and data storage by programming excitonic, phononic, photonic, and dynamic structural nanoscale materials, without attempting to mimic the unknown implementational details of the brain. If realized, such materials would offer transformative opportunities for distributed, multimodal sensing, computation, and data storage in an integrated manner in biol. and other nonconventional environments, including interfacing with biol. sensors and computers such as the brain itself.
- 73Webb, D. R. National Research Council (Us) Chemical Sciences Roundtable. Minorities in the Chemical Workforce: Diversity Models That Work: A Workshop Report to the Chemical Sciences Roundtable; NRC, 2003; https://www.ncbi.nlm.nih.gov/books/NBK36322/, accessed March 31, 2018.Google ScholarThere is no corresponding record for this reference.
- 74Hernandez, R.; Stallings, D.; Iyer, S. In Diversity in the Scientific Community Vol. 1: Quantifying Diversity and Formulating Success; Cheng, H. N., Nelson, D., Eds.; ACS Symposium Series; American Chemical Society; Oxford University Press: Washington DC, 2017; Vol. 1255, pp 101– 112.Google ScholarThere is no corresponding record for this reference.
- 75Berryman, S. E. Who will Do Science? Minority and Female Attainment of Science and Mathematics Degrees: Trends and Causes; Rockefeller Foundation, 1983.Google ScholarThere is no corresponding record for this reference.
- 76Steele, C. M. A Threat in the Air: How Stereotypes Shape Intellectual Identity and Performance. Am. Psychol. 1997, 52, 613, DOI: 10.1037/0003-066X.52.6.613Google Scholar76https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADyaK2szhsleksw%253D%253D&md5=82b49b03cac3553c9fb53d20dbfadebaA threat in the air. How stereotypes shape intellectual identity and performanceSteele C MThe American psychologist (1997), 52 (6), 613-29 ISSN:0003-066X.A general theory of domain identification is used to describe achievement barriers still faced by women in advanced quantitative areas and by African Americans in school. The theory assumes that sustained school success requires identification with school and its subdomains; that societal pressures on these groups (e.g., economic disadvantage, gender roles) can frustrate this identification; and that in school domains where these groups are negatively stereotyped, those who have become domain identified face the further barrier of stereotype threat, the threat that others' judgments or their own actions will negatively stereotype them in the domain. Research shows that this threat dramatically depresses the standardized test performance of women and African Americans who are in the academic vanguard of their groups (offering a new interpretation of group differences in standardized test performance), that it causes disidentification with school, and that practices that reduce this threat can reduce these negative effects.
- 77Raber, L. R. Georgia Section Celebrates 75th Herty Medal. Chem. Eng. News 2009, 87, 43– 44, DOI: 10.1021/cen-v087n042.p043Google Scholar77https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhtlWktLvJ&md5=95fb2b73b17aafc1d0cb75ed247fefc8Georgia Section celebrates 75th Herty MedalRaber, Linda R.Chemical & Engineering News (2009), 87 (42), 43-44CODEN: CENEAR; ISSN:0009-2347. (American Chemical Society)The ACS Georgia Section celebrates the 75th awarding of the Charles Holmes Herty Medal for the year 2009. Craig L. Hill, a professor at Emory University, received the award for his contributions in catalytic oxidn. In particular, he is being recognized for his development of the most reactive known catalysts for removal of environmental pollutants, odors, and toxics from air; prepn. of terminal noble-metal oxo compds.; and creation of the only stable and sol. catalysts for oxidn. of water.
- 78Friend, C. M.; Houk, K. N. Workshop on Building Strong Academic Chemistry Departments through Gender Equity; American Chemical Society, 2006; https://www.acs.org/content/dam/acsorg/funding/awards/national/gender-equity-report-cover.pdf, accessed January 16, 2021.Google ScholarThere is no corresponding record for this reference.
- 79Ali, H. B. Workshop on Excellence Empowered by a Diverse Academic Workforce: Achieving Racial & Ethnic Equity in Chemistry; OSTI, 2008; https://www.osti.gov/biblio/952471, accessed March 31, 2018.Google ScholarThere is no corresponding record for this reference.
- 80Hernandez, R.; Watt, S. In Career Challenges and Opportunities in the Global Chemistry Enterprise; Cheng, H. N., Shah, S., Wu, M. L., Eds.; ACS Symposium Series; American Chemical Society; Oxford University Press: Washington DC, 2014; Vol. 1169, pp 207– 224.Google ScholarThere is no corresponding record for this reference.
- 81Stallings, D.; Iyer, S.; Hernandez, R. In National Diversity Equity Workshops in Chemical Sciences (2011–2017); Hernandez, R., Stallings, D., Iyer, S. K., Eds.; ACS Symposium Series; American Chemical Society; Oxford University Press: Washington DC, 2018; Vol. 1277, pp 1– 19.Google ScholarThere is no corresponding record for this reference.
- 82Iyer, S.; Stallings, D.; Hernandez, R. In National Diversity Equity Workshops in Chemical Sciences (2011–2017); Hernandez, R., Stallings, D., Iyer, S. K., Eds.; ACS Symposium Series; American Chemical Society; Oxford University Press: Washington DC, 2018; Vol. 1277, pp 21– 49.Google ScholarThere is no corresponding record for this reference.
- 83Stallings, D.; Iyer, S.; Hernandez, R. In National Diversity Equity Workshops in Chemical Sciences (2011–2017); Hernandez, R., Stallings, D., Iyer, S. K., Eds.; ACS Symposium Series; American Chemical Society; Oxford University Press: Washington DC, 2018; Vol. 1277, pp 51– 77.Google ScholarThere is no corresponding record for this reference.
- 84Iyer, S.; Stallings, D.; Hernandez, R. In National Diversity Equity Workshops in Chemical Sciences (2011–2017); Hernandez, R., Stallings, D., Iyer, S. K., Eds.; ACS Symposium Series; American Chemical Society; Oxford University Press: Washington DC, 2018; Vol. 1277, pp 79– 107.Google ScholarThere is no corresponding record for this reference.
- 85Stallings, D.; Iyer, S.; Hernandez, R. In National Diversity Equity Workshops in Chemical Sciences (2011–2017); Hernandez, R., Stallings, D., Iyer, S. K., Eds.; ACS Symposium Series; American Chemical Society; Oxford University Press: Washington DC, 2018; Vol. 1277, pp 109– 140.Google ScholarThere is no corresponding record for this reference.
- 86Jacobs, M. Reasons Sought for Lack of Diversity. Chem. Eng. News 2001, 79, 100– 103, DOI: 10.1021/cen-v079n040.p100Google ScholarThere is no corresponding record for this reference.
- 87Hernandez, R. Advancing the chemical sciences through diversity. Chem. Eng. News 2014, 92, 45Google ScholarThere is no corresponding record for this reference.
- 88Kotov, A. Think Like a Grandmaster; Batsford, 1970.Google ScholarThere is no corresponding record for this reference.
- 89Clark, A. E.; Adams, H.; Hernandez, R.; Krylov, A. I.; Niklasson, A. M. N.; Sarupria, S.; Wang, Y.; Wild, S. M.; Yang, Q. Y. ACS Cent. Sci. 2021, DOI: 10.1021/acscentsci.1c00685 .Google ScholarThere is no corresponding record for this reference.