Surface hopping

Last updated

Surface hopping is a mixed quantum-classical technique that incorporates quantum mechanical effects into molecular dynamics simulations. [1] [2] [3] [4] Traditional molecular dynamics assume the Born-Oppenheimer approximation, where the lighter electrons adjust instantaneously to the motion of the nuclei. Though the Born-Oppenheimer approximation is applicable to a wide range of problems, there are several applications, such as photoexcited dynamics, electron transfer, and surface chemistry where this approximation falls apart. Surface hopping partially incorporates the non-adiabatic effects by including excited adiabatic surfaces in the calculations, and allowing for 'hops' between these surfaces, subject to certain criteria.

Contents

Motivation

Molecular dynamics simulations numerically solve the classical equations of motion. These simulations, though, assume that the forces on the electrons are derived solely by the ground adiabatic surface. Solving the time-dependent Schrödinger equation numerically incorporates all these effects, but is computationally unfeasible when the system has many degrees of freedom. To tackle this issue, one approach is the mean field or Ehrenfest method, where the molecular dynamics is run on the average potential energy surface given by a linear combination of the adiabatic states. This was applied successfully for some applications, but has some important limitations. When the difference between the adiabatic states is large, then the dynamics must be primarily driven by only one surface, and not an average potential. In addition, this method also violates the principle of microscopic reversibility. [3]

Surface hopping accounts for these limitations by propagating an ensemble of trajectories, each one of them on a single adiabatic surface at any given time. The trajectories are allowed to 'hop' between various adiabatic states at certain times such that the quantum amplitudes for the adiabatic states follow the time dependent Schrödinger equation. The probability of these hops are dependent on the coupling between the states, and is generally significant only in the regions where the difference between adiabatic energies is small.

Theory behind the method

The formulation described here is in the adiabatic representation for simplicity. [5] It can easily be generalized to a different representation. The coordinates of the system are divided into two categories: quantum () and classical (). The Hamiltonian of the quantum degrees of freedom with mass is defined as:

,

where describes the potential for the whole system. The eigenvalues of as a function of are called the adiabatic surfaces :. Typically, corresponds to the electronic degree of freedom, light atoms such as hydrogen, or high frequency vibrations such as O-H stretch. The forces in the molecular dynamics simulations are derived only from one adiabatic surface, and are given by:

where represents the chosen adiabatic surface. The last equation is derived using the Hellmann-Feynman theorem. The brackets show that the integral is done only over the quantum degrees of freedom. Choosing only one adiabatic surface is an excellent approximation if the difference between the adiabatic surfaces is large for energetically accessible regions of . When this is not the case, the effect of the other states become important. This effect is incorporated in the surface hopping algorithm by considering the wavefunction of the quantum degrees of freedom at time t as an expansion in the adiabatic basis:

,

where are the expansion coefficients. Substituting the above equation into the time dependent Schrödinger equation gives

,

where and the nonadiabatic coupling vector are given by

The adiabatic surface can switch at any given time t based on how the quantum probabilities are changing with time. The rate of change of is given by:

,

where . For a small time interval dt, the fractional change in is given by

.

This gives the net change in flux of population from state . Based on this, the probability of hopping from state j to n is proposed to be

.

This criterion is known as the "fewest switching" algorithm, as it minimizes the number of hops required to maintain the population in various adiabatic states.

Whenever a hop takes place, the velocity is adjusted to maintain conservation of energy. To compute the direction of the change in velocity, the nuclear forces in the transition is

where is the eigen value. For the last equality, is used. This shows that the nuclear forces acting during the hop are in the direction of the nonadiabatic coupling vector . Hence is a reasonable choice for the direction along which velocity should be changed.

Frustrated hops

If the velocity reduction required to conserve energy while making a hop is greater than the component of the velocity to be adjusted, then the hop is known as frustrated. In other words, a hop is frustrated if the system does not have enough energy to make the hop. Several approaches have been suggested to deal with these frustrated hops. The simplest of these is to ignore these hops. [2] Another suggestion is not to change the adiabatic state, but reverse the direction of the component of the velocity along the nonadiabatic coupling vector. [5] Yet another approach is to allow the hop to happen if an allowed hopping point is reachable within uncertainty time , where is the extra energy that the system needed to make the hop possible. [6] Ignoring forbidden hops without any form of velocity reversal does not recover the correct scaling for Marcus theory in the nonadiabatic limit, but a velocity reversal can usually correct the errors [7]

Decoherence time

Surface hopping can develop nonphysical coherences between the quantum coefficients over large time which can degrade the quality of the calculations, at times leading the incorrect scaling for Marcus theory. [8] To eliminate these errors, the quantum coefficients for the inactive state can be damped or set to zero after a predefined time has elapsed after the trajectory crosses the region where hopping has high probabilities. [5]

Outline of the algorithm

The state of the system at any time is given by the phase space of all the classical particles, the quantum amplitudes, and the adiabatic state. The simulation broadly consists of the following steps:

Step 1. Initialize the state of the system. The classical positions and velocities are chosen based on the ensemble required.

Step 2. Compute forces using Hellmann-Feynman theorem, and integrate the equations of motion by time step to obtain the classical phase space at time .

Step 3. Integrate the Schrödinger equation to evolve quantum amplitudes from time to in increments of . This time step is typically much smaller than .

Step 4. Compute probability of hopping from current state to all other states. Generate a random number, and determine whether a switch should take place. If a switch does occur, change velocities to conserve energy. Go back to step 2, till trajectories have been evolved for the desired time.

Applications

The method has been applied successfully to understand dynamics of systems that include tunneling, conical intersections and electronic excitation. [9] [10] [11] [12]

Limitations and foundations

In practice, surface hopping is computationally feasible only for a limited number of quantum degrees of freedom. In addition, the trajectories must have enough energy to be able to reach the regions where probability of hopping is large.

Most of the formal critique of the surface hopping method comes from the unnatural separation of classical and quantum degrees of freedom. Recent work has shown, however, that the surface hopping algorithm can be partially justified by comparison with the Quantum Classical Liouville Equation. [13] It has further been demonstrated that spectroscopic observables can be calculated in close agreement with the formally exact hierarchical equations of motion. [14]

See also

Related Research Articles

Bra–ket notation, also called Dirac notation, is a notation for linear algebra and linear operators on complex vector spaces together with their dual space both in the finite-dimensional and infinite-dimensional case. It is specifically designed to ease the types of calculations that frequently come up in quantum mechanics. Its use in quantum mechanics is quite widespread.

In quantum mechanics, the Hamiltonian of a system is an operator corresponding to the total energy of that system, including both kinetic energy and potential energy. Its spectrum, the system's energy spectrum or its set of energy eigenvalues, is the set of possible outcomes obtainable from a measurement of the system's total energy. Due to its close relation to the energy spectrum and time-evolution of a system, it is of fundamental importance in most formulations of quantum theory.

In quantum chemistry and molecular physics, the Born–Oppenheimer (BO) approximation is the best-known mathematical approximation in molecular dynamics. Specifically, it is the assumption that the wave functions of atomic nuclei and electrons in a molecule can be treated separately, based on the fact that the nuclei are much heavier than the electrons. Due to the larger relative mass of a nucleus compared to an electron, the coordinates of the nuclei in a system are approximated as fixed, while the coordinates of the electrons are dynamic. The approach is named after Max Born and his 23-year-old graduate student J. Robert Oppenheimer, the latter of whom proposed it in 1927 during a period of intense ferment in the development of quantum mechanics.

In physics, an operator is a function over a space of physical states onto another space of physical states. The simplest example of the utility of operators is the study of symmetry. Because of this, they are useful tools in classical mechanics. Operators are even more important in quantum mechanics, where they form an intrinsic part of the formulation of the theory.

The adiabatic theorem is a concept in quantum mechanics. Its original form, due to Max Born and Vladimir Fock (1928), was stated as follows:

In physics, a free particle is a particle that, in some sense, is not bound by an external force, or equivalently not in a region where its potential energy varies. In classical physics, this means the particle is present in a "field-free" space. In quantum mechanics, it means the particle is in a region of uniform potential, usually set to zero in the region of interest since the potential can be arbitrarily set to zero at any point in space.

<span class="mw-page-title-main">Two-state quantum system</span> Simple quantum mechanical system

In quantum mechanics, a two-state system is a quantum system that can exist in any quantum superposition of two independent quantum states. The Hilbert space describing such a system is two-dimensional. Therefore, a complete basis spanning the space will consist of two independent states. Any two-state system can also be seen as a qubit.

Vibronic coupling in a molecule involves the interaction between electronic and nuclear vibrational motion. The term "vibronic" originates from the combination of the terms "vibrational" and "electronic", denoting the idea that in a molecule, vibrational and electronic interactions are interrelated and influence each other. The magnitude of vibronic coupling reflects the degree of such interrelation.

One of the guiding principles in modern chemical dynamics and spectroscopy is that the motion of the nuclei in a molecule is slow compared to that of its electrons. This is justified by the large disparity between the mass of an electron, and the typical mass of a nucleus and leads to the Born–Oppenheimer approximation and the idea that the structure and dynamics of a chemical species are largely determined by nuclear motion on potential energy surfaces.

The Ehrenfest theorem, named after Austrian theoretical physicist Paul Ehrenfest, relates the time derivative of the expectation values of the position and momentum operators x and p to the expectation value of the force on a massive particle moving in a scalar potential ,

The Born–Huang approximation is an approximation closely related to the Born–Oppenheimer approximation. It takes into account diagonal nonadiabatic effects in the electronic Hamiltonian better than the Born–Oppenheimer approximation. Despite the addition of correction terms, the electronic states remain uncoupled under the Born–Huang approximation, making it an adiabatic approximation.

<span class="mw-page-title-main">Jaynes–Cummings model</span> Model in quantum optics

The Jaynes–Cummings model is a theoretical model in quantum optics. It describes the system of a two-level atom interacting with a quantized mode of an optical cavity, with or without the presence of light. It was originally developed to study the interaction of atoms with the quantized electromagnetic field in order to investigate the phenomena of spontaneous emission and absorption of photons in a cavity.

The theoretical and experimental justification for the Schrödinger equation motivates the discovery of the Schrödinger equation, the equation that describes the dynamics of nonrelativistic particles. The motivation uses photons, which are relativistic particles with dynamics described by Maxwell's equations, as an analogue for all types of particles.

<span class="mw-page-title-main">Quantum vortex</span> Quantized flux circulation of some physical quantity

In physics, a quantum vortex represents a quantized flux circulation of some physical quantity. In most cases, quantum vortices are a type of topological defect exhibited in superfluids and superconductors. The existence of quantum vortices was first predicted by Lars Onsager in 1949 in connection with superfluid helium. Onsager reasoned that quantisation of vorticity is a direct consequence of the existence of a superfluid order parameter as a spatially continuous wavefunction. Onsager also pointed out that quantum vortices describe the circulation of superfluid and conjectured that their excitations are responsible for superfluid phase transitions. These ideas of Onsager were further developed by Richard Feynman in 1955 and in 1957 were applied to describe the magnetic phase diagram of type-II superconductors by Alexei Alexeyevich Abrikosov. In 1935 Fritz London published a very closely related work on magnetic flux quantization in superconductors. London's fluxoid can also be viewed as a quantum vortex.

In 1927, a year after the publication of the Schrödinger equation, Hartree formulated what are now known as the Hartree equations for atoms, using the concept of self-consistency that Lindsay had introduced in his study of many electron systems in the context of Bohr theory. Hartree assumed that the nucleus together with the electrons formed a spherically symmetric field. The charge distribution of each electron was the solution of the Schrödinger equation for an electron in a potential , derived from the field. Self-consistency required that the final field, computed from the solutions, was self-consistent with the initial field, and he thus called his method the self-consistent field method.

A flavor of the k·p perturbation theory used for calculating the structure of multiple, degenerate electronic bands in bulk and quantum well semiconductors. The method is a generalization of the single band k·p theory.

An electric dipole transition is the dominant effect of an interaction of an electron in an atom with the electromagnetic field.

In physics, Berry connection and Berry curvature are related concepts which can be viewed, respectively, as a local gauge potential and gauge field associated with the Berry phase or geometric phase. The concept was first introduced by S. Pancharatnam as geometric phase and later elaborately explained and popularized by Michael Berry in a paper published in 1984 emphasizing how geometric phases provide a powerful unifying concept in several branches of classical and quantum physics.

In pure and applied mathematics, quantum mechanics and computer graphics, a tensor operator generalizes the notion of operators which are scalars and vectors. A special class of these are spherical tensor operators which apply the notion of the spherical basis and spherical harmonics. The spherical basis closely relates to the description of angular momentum in quantum mechanics and spherical harmonic functions. The coordinate-free generalization of a tensor operator is known as a representation operator.

<span class="mw-page-title-main">Newton-X</span> Molecular dynamics simulation software

Newton-X is a general program for molecular dynamics simulations beyond the Born-Oppenheimer approximation. It has been primarily used for simulations of ultrafast processes in photoexcited molecules. It has also been used for simulation of band envelops of absorption and emission spectra.

References

  1. Herman, Michael F. (1984). "Nonadiabatic semiclassical scattering. I. Analysis of generalized surface hopping procedures". The Journal of Chemical Physics. 81 (2): 754–763. Bibcode:1984JChPh..81..754H. doi:10.1063/1.447708.
  2. 1 2 Tully, John C. (1990). "Molecular dynamics with electronic transitions". The Journal of Chemical Physics. 93 (2): 1061–1071. Bibcode:1990JChPh..93.1061T. doi:10.1063/1.459170. S2CID   15191625.
  3. 1 2 Quantum simulations of complex many-body systems: from theory to algorithms : winter school, 25 February - 1 March 2002, Rolduc Conference Centre, Kerkrade, the Netherlands ; lecture notes. Grotendorst, Johannes., Winter school (2002.02.25-03.01 : Kerkrade). Jülich: NIC-Secretariat. 2002. ISBN   3000090576. OCLC   248502198.{{cite book}}: CS1 maint: others (link)
  4. Barbatti, Mario (2011). "Nonadiabatic dynamics with trajectory surface hopping method". Wiley Interdisciplinary Reviews: Computational Molecular Science. 1 (4): 620–633. doi:10.1002/wcms.64. S2CID   123626773.
  5. 1 2 3 Hammes-Schiffer, Sharon; Tully, John C. (1994). "Proton transfer in solution: Molecular dynamics with quantum transitions". The Journal of Chemical Physics. 101 (6): 4657. Bibcode:1994JChPh.101.4657H. doi:10.1063/1.467455.
  6. Jasper, Ahren W.; Stechmann, Samuel N.; Truhlar, Donald G. (2002). "Fewest-switches with time uncertainty: A modified trajectory surface-hopping algorithm with better accuracy for classically forbidden electronic transitions". The Journal of Chemical Physics. 116 (13): 5424. Bibcode:2002JChPh.116.5424J. doi:10.1063/1.1453404.
  7. Jain, Amber; Subotnik, Joseph (2015). "Surface hopping, transition state theory, and decoherence. II. Thermal rate constants and detailed balance". The Journal of Chemical Physics. 143 (13): 134107. Bibcode:2015JChPh.143m4107J. doi:10.1063/1.4930549. PMID   26450292. S2CID   205207864.
  8. Landry, Brian R.; Subotnik, Joseph (2015). "Standard surface hopping predicts incorrect scaling for Marcus' golden-rule rate: The decoherence problem cannot be ignored". The Journal of Chemical Physics. 135 (19): 191101. Bibcode:2011JChPh.135s1101L. doi: 10.1063/1.3663870 . PMID   22112058.
  9. Tapavicza, Enrico; Tavernelli, Ivano; Rothlisberger, Ursula (2007). "Trajectory surface hopping within linear response time-dependent density-functional theory". Physical Review Letters. 98 (2): 023001. Bibcode:2007PhRvL..98b3001T. doi:10.1103/PhysRevLett.98.023001. PMID   17358601.
  10. Jiang, Ruomu; Sibert, Edwin L. (2012). "Surface hopping simulation of vibrational predissociation of methanol dimer". The Journal of Chemical Physics. 136 (22): 224104. Bibcode:2012JChPh.136v4104J. doi:10.1063/1.4724219. PMID   22713033.
  11. Müller, Uwe; Stock, Gerhard (22 October 1997). "Surface-hopping modeling of photoinduced relaxation dynamics on coupled potential-energy surfaces". The Journal of Chemical Physics. 107 (16): 6230–6245. Bibcode:1997JChPh.107.6230M. doi:10.1063/1.474288.
  12. Martens, Craig C. (2016-07-07). "Surface Hopping by Consensus". The Journal of Physical Chemistry Letters. 7 (13): 2610–2615. doi:10.1021/acs.jpclett.6b01186. ISSN   1948-7185. PMID   27345103.
  13. Subotnik, Joseph E.; Wenjun Ouyang; Brian R. Landry (2013). "Can we derive Tully's surface-hopping algorithm from the semiclassical quantum Liouville equation? Almost, but only with decoherence". The Journal of Chemical Physics. 139 (21): 214107. Bibcode:2013JChPh.139u4107S. doi:10.1063/1.4829856. PMID   24320364.
  14. Tempelaar, Roel; van de Vegte, Cornelis; Knoester, Jasper; Jansen, Thomas L. C. (2013). "Surface hopping modeling of two-dimensional spectra" (PDF). The Journal of Chemical Physics. 138 (16): 164106. Bibcode:2013JChPh.138p4106T. doi:10.1063/1.4801519. PMID   23635110.