Adiabatic electron transfer

Last updated

Adiabatic electron-transfer is a type of oxidation-reduction processes. The mechanism is ubiquitous in nature in both the inorganic and biological spheres. Adiabatic electron-transfers proceed without making or breaking chemical bonds. Adiabatic electron-transfer can occur by either optical or thermal mechanisms. [1] [2] Electron transfer during a collision between an oxidant and a reductant occurs adiabatically on a continuous potential-energy surface.

History

Noel Hush is often credited with formulation of the theory of adiabatic electron-transfer. [3] [4]

Fig. 1. Electron transfer occurs between donor (D) and acceptor (A) species separated by distance R that may be found in many forms in both condensed phases and the gas phase. Internal structure, external structure, or chance collisions provide interconnection between the species. Upon electron transfer, the structure of the local chemical environments involving D and A change, as does the polarization these species induce on any surrounding media. Hush wiki Fig 1.png
Fig. 1. Electron transfer occurs between donor (D) and acceptor (A) species separated by distance R that may be found in many forms in both condensed phases and the gas phase. Internal structure, external structure, or chance collisions provide interconnection between the species. Upon electron transfer, the structure of the local chemical environments involving D and A change, as does the polarization these species induce on any surrounding media.

Figure 1 sketches the basic elements of adiabatic electron-transfer theory. Two chemical species (ions, molecules, polymers, protein cofactors, etc.) labelled D (for “donor”) and A (for “acceptor”) become a distance R apart, either through collisions, covalent bonding, location in a material, protein or polymer structure, etc. A and D have different chemical environments. Each polarizes their surrounding condensed media. Electron-transfer theories describe the influence of a variety of parameters on the rate of electron-transfer. All electrochemical reactions occur by this mechanism. Adiabatic electron-transfer theory stresses that intricately coupled to such charge transfer is the ability of any D-A system to absorb or emit light. Hence fundamental understanding of any electrochemical process demands simultaneous understanding of the optical processes that the system can undergo.

Fig. 2. When the donor species absorbs light energy, it goes into a high-energy excited state, generating significant changes to its local chemical environment and the polarization of its external environment. These environments facilitate coupling
V
D
A
{\displaystyle V_{DA}}
between the donor and acceptor, which drives photochemical charge separation with a rate given by Eqn. (3) in the weak-coupling limit. This rate is also dependent on the energy
l
{\displaystyle \lambda }
required to rearrange the atoms to the preferred local geometry and environment polarization of the charge-separated state D -A and the energy change
D
G
0
{\displaystyle \Delta G_{0}}
associated with charge separation. Hush wiki Fig 2.png
Fig. 2. When the donor species absorbs light energy, it goes into a high-energy excited state, generating significant changes to its local chemical environment and the polarization of its external environment. These environments facilitate coupling between the donor and acceptor, which drives photochemical charge separation with a rate given by Eqn. (3) in the weak-coupling limit. This rate is also dependent on the energy required to rearrange the atoms to the preferred local geometry and environment polarization of the charge-separated state D -A and the energy change associated with charge separation.

Figure 2 sketches what happens if light is absorbed by just one of the chemical species, taken to be the charge donor. This produces an excited state of the donor. As the donor and acceptor are close to each other and surrounding matter, they experience a coupling . If the free energy change is favorable, this coupling facilitates primary charge separation to produce D+-A , producing charged species. In this way, solar energy is captured and converted to electrical energy. This process is typical of natural photosynthesis as well as modern organic photovoltaic and artificial photosynthesis solar-energy capture devices. [5] The inverse of this process is also used to make organic light-emitting diodes (OLEDs).

Fig. 3. Light energy is absorbed by the donor and acceptor, initiating intervalence charge transfer to directly convert solar energy into electrical energy as D -A . In the weak-coupling limit, the coupling
V
D
A
{\displaystyle V_{DA}}
, reorganization energy
l
{\displaystyle \lambda }
, and the free energy change
D
G
0
{\displaystyle \Delta G_{0}}
control the rate of light absorption (and hence charge separation) via Eqn. (1). Hush wiki Fig 3.png
Fig. 3. Light energy is absorbed by the donor and acceptor, initiating intervalence charge transfer to directly convert solar energy into electrical energy as D -A . In the weak-coupling limit, the coupling , reorganization energy , and the free energy change control the rate of light absorption (and hence charge separation) via Eqn. (1).

Adiabatic electron-transfer is also relevant to the area of solar energy harvesting. Here, light absorption directly leads to charge separation D+-A. Hush's theory for this process [2] considers the donor-acceptor coupling , the energy required to rearrange the atoms from their initial geometry to the preferred local geometry and environment polarization of the charge-separated state, and the energy change associated with charge separation. In the weak-coupling limit ( ), Hush showed [2] that the rate of light absorption (and hence charge separation) is given from the Einstein equation by

… (1)

This theory explained [2] how Prussian blue absorbes light, creating [6] [7] [8] [9] [10] the field of intervalence charge transfer spectroscopy.

Adiabatic electron transfer is also relevant to the Robin-Day classification system, which codifies types of mixed valence compounds. [11] [12] An iconic system for understanding Inner sphere electron transfer is the mixed-valence Creutz-Taube ion, wherein otherwise equivalent Ru(III) and Ru(II) are linked by a pyrazine. The coupling is not small: charge is not localized on just one chemical species but is shared quantum mechanically between two Ru centers, presenting classically forbidden half-integral valence states. [13] that the critical requirement for this phenomenon is

… (2)

Adiabatic electron-transfer theory stems from London's approach to charge-transfer and indeed general chemical reactions [14] applied by Hush using parabolic potential-energy surfaces. [15] [16] Hush himself has carried out many theoretical and experimental studies of mixed valence complexes and long range electron transfer in biological systems. Hush's quantum-electronic adiabatic approach to electron transfer was unique; directly connecting with the Quantum Chemistry concepts of Mulliken, it forms the basis of all modern computational approaches to modeling electron transfer. [17] [18] [19] [20] Its essential feature is that electron transfer can never be regarded as an “instantaneous transition”; instead, the electron is partially transferred at all molecular geometries, with the extent of the transfer being a critical quantum descriptor of all thermal, tunneling, and spectroscopic processes. It also leads seamlessly [21] to understanding electron-transfer transition-state spectroscopy pioneered by Zewail.

In adiabatic electron-transfer theory, the ratio is of central importance. In the very strong coupling limit when Eqn. (2) is satisfied, intrinsically quantum molecules like the Crautz-Taube ion result. Most intervalence spectroscopy occurs in the weak-coupling limit described by Eqn. (1), however. In both natural photosynthesis and in artificial solar-energy capture devices, is maximized by minimizing through use of large molecules like chlorophylls, pentacenes, and conjugated polymers. The coupling can be controlled by controlling the distance R at which charge transfer occurs- the coupling typically decreases exponentially with distance. When electron transfer occurs during collisions of the D and A species, the coupling is typically large and the “adiabatic” limit applies in which rate constants are given by transition state theory. [4] In biological applications, however, as well as some organic conductors and other device materials, R is externally constrained and so the coupling set at low or high values. In these situations, weak-coupling scenarios often become critical.

In the weak-coupling (“non-adiabatic”) limit, the activation energy for electron transfer is given by the expression derived independently by Kubo and Toyozawa [22] and by Hush. [16] Using adiabatic electron-transfer theory, [23] in this limit Levich and Dogonadze then determined the electron-tunneling probability to express the rate constant for thermal reactions as [24]

. … (3)

This approach is widely applicable to long-range ground-state intramolecular electron transfer, electron transfer in biology, and electron transfer in conducting materials. It also typically controls the rate of charge separation in the excited-state photochemical application described in Figure 2 and related problems.

Marcus showed that the activation energy in Eqn. (3) reduces to in the case of symmetric reactions with . In that work, [25] he also derived the standard expression for the solvent contribution to the reorganization energy, making the theory more applicable to practical problems. Use of this solvation description (instead [4] of the form that Hush originally proposed [16] ) in approaches spanning the adiabatic and non-adiabatic limits is often termed “Marcus-Hush Theory”. [18] [19] [26] [27] These and other contributions, including the widespread demonstration of the usefulness of Eqn. (3), [28] led to the award of the 1992 Nobel Prize in Chemistry to Marcus.

Adiabatic electron-transfer theory is also widely applied in Molecular Electronics. [29] In particular, this reconnects adiabatic electron-transfer theory with its roots in proton-transfer theory [30] and hydrogen-atom transfer, [15] leading back to London's theory of general chemical reactions. [14]

Related Research Articles

<span class="mw-page-title-main">Electrode</span> Electrical conductor used to make contact with nonmetallic parts of a circuit

An electrode is an electrical conductor used to make contact with a nonmetallic part of a circuit. Electrodes are essential parts of batteries that can consist of a variety of materials depending on the type of battery.

<span class="mw-page-title-main">Quantum field theory</span> Theoretical framework combining classical field theory, special relativity, and quantum mechanics

In theoretical physics, quantum field theory (QFT) is a theoretical framework that combines classical field theory, special relativity, and quantum mechanics. QFT is used in particle physics to construct physical models of subatomic particles and in condensed matter physics to construct models of quasiparticles.

<span class="mw-page-title-main">Ionization</span> Process by which atoms or molecules acquire charge by gaining or losing electrons

Ionization is the process by which an atom or a molecule acquires a negative or positive charge by gaining or losing electrons, often in conjunction with other chemical changes. The resulting electrically charged atom or molecule is called an ion. Ionization can result from the loss of an electron after collisions with subatomic particles, collisions with other atoms, molecules and ions, or through the interaction with electromagnetic radiation. Heterolytic bond cleavage and heterolytic substitution reactions can result in the formation of ion pairs. Ionization can occur through radioactive decay by the internal conversion process, in which an excited nucleus transfers its energy to one of the inner-shell electrons causing it to be ejected.

<span class="mw-page-title-main">Ionization energy</span> Energy needed to remove an electron

In physics and chemistry, ionization energy (IE) (American English spelling), ionisation energy (British English spelling) is the minimum energy required to remove the most loosely bound electron of an isolated gaseous atom, positive ion, or molecule. The first ionization energy is quantitatively expressed as

Matter waves are a central part of the theory of quantum mechanics, being half of wave–particle duality. All matter exhibits wave-like behavior. For example, a beam of electrons can be diffracted just like a beam of light or a water wave.

<span class="mw-page-title-main">Rydberg formula</span> Formula for spectral line wavelengths in alkali metals

In atomic physics, the Rydberg formula calculates the wavelengths of a spectral line in many chemical elements. The formula was primarily presented as a generalization of the Balmer series for all atomic electron transitions of hydrogen. It was first empirically stated in 1888 by the Swedish physicist Johannes Rydberg, then theoretically by Niels Bohr in 1913, who used a primitive form of quantum mechanics. The formula directly generalizes the equations used to calculate the wavelengths of the hydrogen spectral series.

<span class="mw-page-title-main">Renormalization</span> Method in physics used to deal with infinities

Renormalization is a collection of techniques in quantum field theory, the statistical mechanics of fields, and the theory of self-similar geometric structures, that are used to treat infinities arising in calculated quantities by altering values of these quantities to compensate for effects of their self-interactions. But even if no infinities arose in loop diagrams in quantum field theory, it could be shown that it would be necessary to renormalize the mass and fields appearing in the original Lagrangian.

In theoretical physics, the term renormalization group (RG) refers to a formal apparatus that allows systematic investigation of the changes of a physical system as viewed at different scales. In particle physics, it reflects the changes in the underlying force laws as the energy scale at which physical processes occur varies, energy/momentum and resolution distance scales being effectively conjugate under the uncertainty principle.

<span class="mw-page-title-main">Coupling constant</span> Parameter describing the strength of a force

In physics, a coupling constant or gauge coupling parameter, is a number that determines the strength of the force exerted in an interaction. Originally, the coupling constant related the force acting between two static bodies to the "charges" of the bodies divided by the distance squared, , between the bodies; thus: in for Newtonian gravity and in for electrostatic. This description remains valid in modern physics for linear theories with static bodies and massless force carriers.

In physics, mathematics and statistics, scale invariance is a feature of objects or laws that do not change if scales of length, energy, or other variables, are multiplied by a common factor, and thus represent a universality.

<span class="mw-page-title-main">Rydberg atom</span> Excited atomic quantum state with high principal quantum number (n)

A Rydberg atom is an excited atom with one or more electrons that have a very high principal quantum number, n. The higher the value of n, the farther the electron is from the nucleus, on average. Rydberg atoms have a number of peculiar properties including an exaggerated response to electric and magnetic fields, long decay periods and electron wavefunctions that approximate, under some conditions, classical orbits of electrons about the nuclei. The core electrons shield the outer electron from the electric field of the nucleus such that, from a distance, the electric potential looks identical to that experienced by the electron in a hydrogen atom.

<span class="mw-page-title-main">Aufbau principle</span> Principle of atomic physics

The aufbau principle, also called the aufbau rule, states that in the ground state of an atom or ion, electrons fill subshells of the lowest available energy, then they fill subshells of higher energy. For example, the 1s subshell is filled before the 2s subshell is occupied. In this way, the electrons of an atom or ion form the most stable electron configuration possible. An example is the configuration 1s2 2s2 2p6 3s2 3p3 for the phosphorus atom, meaning that the 1s subshell has 2 electrons, and so on.

Møller–Plesset perturbation theory (MP) is one of several quantum chemistry post–Hartree–Fock ab initio methods in the field of computational chemistry. It improves on the Hartree–Fock method by adding electron correlation effects by means of Rayleigh–Schrödinger perturbation theory (RS-PT), usually to second (MP2), third (MP3) or fourth (MP4) order. Its main idea was published as early as 1934 by Christian Møller and Milton S. Plesset.

In quantum field theory, and specifically quantum electrodynamics, vacuum polarization describes a process in which a background electromagnetic field produces virtual electron–positron pairs that change the distribution of charges and currents that generated the original electromagnetic field. It is also sometimes referred to as the self-energy of the gauge boson (photon).

In quantum mechanics, the Hellmann–Feynman theorem relates the derivative of the total energy with respect to a parameter to the expectation value of the derivative of the Hamiltonian with respect to that same parameter. According to the theorem, once the spatial distribution of the electrons has been determined by solving the Schrödinger equation, all the forces in the system can be calculated using classical electrostatics.

In theoretical chemistry, Marcus theory is a theory originally developed by Rudolph A. Marcus, starting in 1956, to explain the rates of electron transfer reactions – the rate at which an electron can move or jump from one chemical species (called the electron donor) to another (called the electron acceptor). It was originally formulated to address outer sphere electron transfer reactions, in which the two chemical species only change in their charge with an electron jumping (e.g. the oxidation of an ion like Fe2+/Fe3+), but do not undergo large structural changes. It was extended to include inner sphere electron transfer contributions, in which a change of distances or geometry in the solvation or coordination shells of the two chemical species is taken into account (the Fe-O distances in Fe(H2O)2+ and Fe(H2O)3+ are different).

The conductance quantum, denoted by the symbol G0, is the quantized unit of electrical conductance. It is defined by the elementary charge e and Planck constant h as:

Car–Parrinello molecular dynamics or CPMD refers to either a method used in molecular dynamics or the computational chemistry software package used to implement this method.

Surface hopping is a mixed quantum-classical technique that incorporates quantum mechanical effects into molecular dynamics simulations. 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.

Electric dipole spin resonance (EDSR) is a method to control the magnetic moments inside a material using quantum mechanical effects like the spin–orbit interaction. Mainly, EDSR allows to flip the orientation of the magnetic moments through the use of electromagnetic radiation at resonant frequencies. EDSR was first proposed by Emmanuel Rashba.

References

  1. Piechota, Eric J.; Meyer, Gerald J. (2019). "Introduction to Electron Transfer: Theoretical Foundations and Pedagogical Examples". Journal of Chemical Education. 96 (11): 2450–2466. Bibcode:2019JChEd..96.2450P. doi:10.1021/acs.jchemed.9b00489. S2CID   208754569.
  2. 1 2 3 4 Hush, N. S. (1967). Intervalence-transfer absorption. II. Theoretical considerations and spectroscopic data. Progress in Inorganic Chemistry. Vol. 8. pp. 391–444. doi:10.1002/9780470166093.ch7. ISBN   9780470166093.
  3. "Fellows Details". Royal Society. Retrieved 18 September 2015.
  4. 1 2 3 Hush, N. S. (1961). "Adiabatic theory of outer sphere electron-transfer reactions in solution". Transactions of the Faraday Society. 57: 577. doi:10.1039/TF9615700557.
  5. Warman, J. M.; Haas, M. P. d.; Paddon-Row, M. N.; Cotsaris, E.; Hush, N. S.; Oevering, H.; Verhoeven, J. W. (1986). "Light-induced giant dipoles in simple model compounds for photosynthesis". Nature. 320 (6063): 615–616. Bibcode:1986Natur.320..615W. doi:10.1038/320615a0. S2CID   4346663.
  6. Nelsen, S. F.; Ismagilov, R. F.; Trieber, D. A. (1997). "Adiabatic Electron Transfer: Comparison of Modified Theory with Experiment" (PDF). Science. 278 (5339): 846–849. Bibcode:1997Sci...278..846N. doi:10.1126/science.278.5339.846. PMID   9346480.
  7. German, E. D. (1979). "Intramolecular intervalence charge transfer in bimolecular mixed-valence complexes of metals". Chemical Physics Letters. 64 (2): 295–298. Bibcode:1979CPL....64..295G. doi:10.1016/0009-2614(79)80516-3.
  8. Sun, D. L.; Rosokha, S. V.; Lindeman, S. V.; Kochi, J. K. (2003). "Intervalence (Charge-Resonance) Transitions in Organic Mixed-Valence Systems. Through-Space versus Through-Bond Electron Transfer between Bridged Aromatic (Redox) Centers". Journal of the American Chemical Society. 125 (51): 15950–15963. doi:10.1021/ja037867s. PMID   14677987.
  9. Nelsen, S. F.; Weaver, M. N.; Luo, Y.; Lockard, J. V.; Zink, J. I. (2006). "Use of the neighboring orbital model for analysis of electronic coupling in Class III intervalence compounds". Chemical Physics. 324 (1): 195–201. Bibcode:2006CP....324..195N. doi:10.1016/j.chemphys.2006.01.023.
  10. Rosokha, S. V.; Kochi, J. K. (2008). "Fresh Look at Electron-Transfer Mechanisms via the Donor/Acceptor Bindings in the Critical Encounter Complex". Accounts of Chemical Research. 41 (5): 641–653. doi:10.1021/ar700256a. PMID   18380446.
  11. Robin, M. B.; Day, P. (1967). Mixed Valence Chemistry-A Survey and Classification. Advances in Inorganic Chemistry and Radiochemistry. Vol. 10. pp. 247–422. doi:10.1016/S0065-2792(08)60179-X. ISBN   9780120236107.
  12. Day, P.; Hush, N. S.; Clark, R. J. H. (2008). "Mixed valence: origins and developments". Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences. 366 (1862): 5–14. Bibcode:2008RSPTA.366....5D. doi:10.1098/rsta.2007.2135. PMID   17827130. S2CID   5912503.
  13. Hush, N. S. (1975). "Inequivalent XPS [x-ray photoelectron spectroscopy] binding energies in symmetrical delocalized mixed-valence complexes". Chemical Physics. 10 (2–3): 361–366. Bibcode:1975CP.....10..361H. doi:10.1016/0301-0104(75)87049-2.
  14. 1 2 London, F. (1932). "On the theory of non-adiabatic chemical reactions". Quantum Chemistry. World Scientific Series in 20th Century Chemistry. Vol. 74. pp. 32–60. doi:10.1142/9789812795762_0003. ISBN   978-981-02-2771-5.
  15. 1 2 Hush, N. S. (1953). "Quantum-mechanical discussion of the gas phase formation of quinonedimethide monomers". Journal of Polymer Science. 11 (4): 289–298. Bibcode:1953JPoSc..11..289H. doi:10.1002/pol.1953.120110401.
  16. 1 2 3 Hush, N. S. (1958). "Adiabatic rate processes at electrodes". The Journal of Chemical Physics. 28 (5): 962–972. Bibcode:1958JChPh..28..962H. doi:10.1063/1.1744305.
  17. Kornyshev, A. A.; Tosi, M.; Ulstrup, J. (1997). Electron and Ion Transfer in Condensed Media. Singapore: World Scientific. ISBN   978-9810229290.
  18. 1 2 Kuznetsov, A.; Ulstrup, J. (1998). Electron Transfer in Chemistry and Biology: An introduction to the theory. Chichester: Wiley. ISBN   978-0-471-96749-1.
  19. 1 2 Devault, D. (1980). "Quantum mechanical tunnelling in biological systems". Quarterly Reviews of Biophysics. 13 (4): 387–564. doi:10.1017/S003358350000175X. PMID   7015406. S2CID   26771752.
  20. Cave, R. J.; Newton, M. D. (1996). "Generalization of the Mulliken-Hush treatment for the calculation of electron transfer matrix elements". Chemical Physics Letters. 249 (1–2): 15–19. Bibcode:1996CPL...249...15C. doi:10.1016/0009-2614(95)01310-5.
  21. Reimers, J. R.; Hush, N. S. (2017). "Relating transition-state spectroscopy to standard chemical spectroscopic processes". Chemical Physics Letters. 683: 467–477. Bibcode:2017CPL...683..467R. doi:10.1016/j.cplett.2017.04.070. hdl: 10453/125251 .
  22. Kubo, R.; Toyozawa, Y. (1955). "Application of the Method of Generating Function to Radiative and Non-Radiative Transitions of a Trapped Electron in a Crystal". Progress of Theoretical Physics. 13 (2): 160–182. Bibcode:1955PThPh..13..160K. doi: 10.1143/PTP.13.160 .
  23. Levich, V. G.; Dogonadze, R. R. (1960). "Adiabatic theory for electron-transfer processes in solution". Proc. Akad. Naukl. SSSR. 133: 591.
  24. Levich, V. G.; Dogonadze, R. R. (1959). "Theory of rediationless electron transitions between ions in solution". Proc. Akad. Naukl. SSSR. 29: 9.
  25. Marcus, R. A. (1956). "On the Theory of Oxidation-Reduction Reactions Involving Electron Transfer. 1" (PDF). The Journal of Chemical Physics. 24 (5): 966–978. Bibcode:1956JChPh..24..966M. doi:10.1063/1.1742723. S2CID   16579694.
  26. Schmickler, W. Electron transfer and single molecule events (PDF). Paris: Eolss Publishers.
  27. Efrima, S.; Bixon, M. (1976). "Vibrational effects in outer-sphere electron-transfer reactions in polar media". Chemical Physics. 13 (4): 447–460. Bibcode:1976CP.....13..447E. doi:10.1016/0301-0104(76)87014-0.
  28. Marcus, R. A.; Sutin, N. (1985). "Electron transfers in chemistry and biology". Biochimica et Biophysica Acta (BBA) - Reviews on Bioenergetics. 811 (3): 265–322. doi:10.1016/0304-4173(85)90014-x.
  29. Reimers, J. R.; McKemmish, L.; McKenzie, R. H.; Hush, N. S. (2015). "A unified diabatic description for electron transfer reactions, isomerization reactions, proton transfer reactions, and aromaticity". Physical Chemistry Chemical Physics. 17 (38): 24598–24617. Bibcode:2015PCCP...1724598R. doi:10.1039/C5CP02236C. PMID   26193994.
  30. Horiuti, J.; Polanyi, M. (2003). "Outlines of a theory of proton transfer". Journal of Molecular Catalysis A: Chemical. 199 (1–2): 185–197. doi:10.1016/s1381-1169(03)00034-7.