Marcus Theory for Electron Transfer a short introduction

Similar documents
12.2 MARCUS THEORY 1 (12.22)

Theory of electron transfer. Winterschool for Theoretical Chemistry and Spectroscopy Han-sur-Lesse, Belgium, December 2011

Charge and Energy Transfer Dynamits in Molecular Systems

5.74 Introductory Quantum Mechanics II

11.1. FÖRSTER RESONANCE ENERGY TRANSFER

CHEM Core Chemistry 3. Inorganic Reaction Mechanisms

10 Theoretical considerations of electron-transfer reactions

Born-Oppenheimer Approximation

Of Electrons, Energy, and Excitons

Tutorial on rate constants and reorganization energies

Quantum Master Equations for the Electron Transfer Problem

Electron-proton transfer theory and electrocatalysis Part I

(Excerpt from S. Ji, Molecular Theory of the Living Cell: Concepts, Molecular Mechanisms, and Biomedical Applications, Springer, New York, 2012)

2m 2 Ze2. , where δ. ) 2 l,n is the quantum defect (of order one but larger

Laser Induced Control of Condensed Phase Electron Transfer

Lecture 12. Electron Transport in Molecular Wires Possible Mechanisms

CHEM3023: Spins, Atoms and Molecules

3.23 Electrical, Optical, and Magnetic Properties of Materials

Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Spring Semester 2006 Christopher J. Cramer. Lecture 9, February 8, 2006

An Introduction to Quantum Chemistry and Potential Energy Surfaces. Benjamin G. Levine

Schedule. Lecture 7: M-M bonds δ-bonds and bonding in metal clusters

I 2 Vapor Absorption Experiment and Determination of Bond Dissociation Energy.

Supporting Information for Electron transport in crystalline PCBM-like fullerene derivatives: a comparative computational study

Electron transfer theory and electrocatalysis. Marc Koper Leiden University (NL)

André Schleife Department of Materials Science and Engineering

eigenvalues eigenfunctions

Theoretical formulation for electron transfer coupled to multiple protons: Application to amidinium carboxylate interfaces

Energetics of Electron Transfer Reactions

Chem 442 Review of Spectroscopy

Advanced Analytical Chemistry

UV-vis (Electronic) Spectra Ch.13 Atkins, Ch.19 Engel

Theoretical Photochemistry WiSe 2017/18

Energy and Cells. Appendix 1. The two primary energy transformations in plants are photosynthesis and respiration.

What dictates the rate of radiative or nonradiative excited state decay?

I 2 Vapor Absorption Experiment and Determination of Bond Dissociation Energy.

(Ox) 6I - (aq) + BrO 3 - (aq) + 6H + (aq) 3I 2 (aq) + Br - (aq) + 3H 2 O(l)

NPTEL/IITM. Molecular Spectroscopy Lectures 1 & 2. Prof.K. Mangala Sunder Page 1 of 15. Topics. Part I : Introductory concepts Topics

Chem 344 Final Exam Tuesday, Dec. 11, 2007, 3-?? PM

Equilibrium stable isotope fractionation

Theoretical Photochemistry WiSe 2016/17

Quantum Chemistry. NC State University. Lecture 5. The electronic structure of molecules Absorption spectroscopy Fluorescence spectroscopy

Luminescence. Photoluminescence (PL) is luminescence that results from optically exciting a sample.

2 Reaction kinetics in gases

Spectroscopy in Inorganic Chemistry. Vibration and Rotation Spectroscopy

Energetics of protein charge transfer and photosynthesis

Molecular orbitals, potential energy surfaces and symmetry

Principles of Molecular Spectroscopy

Bimolecular processes

Polarons. University of Ljubljana Faculty of Mathematics and Physics. Seminar - Ib, 1. year II. cycle degrees

5.61 Physical Chemistry Final Exam 12/16/09. MASSACHUSETTS INSTITUTE OF TECHNOLOGY Department of Chemistry Chemistry Physical Chemistry

Energy Barriers and Rates - Transition State Theory for Physicists

Lecture 08 Born Oppenheimer Approximation

Ayan Chattopadhyay Mainak Mustafi 3 rd yr Undergraduates Integrated MSc Chemistry IIT Kharagpur

( r) = 1 Z. e Zr/a 0. + n +1δ n', n+1 ). dt ' e i ( ε n ε i )t'/! a n ( t) = n ψ t = 1 i! e iε n t/! n' x n = Physics 624, Quantum II -- Exam 1

Finite Ring Geometries and Role of Coupling in Molecular Dynamics and Chemistry

Beyond Born-Oppenheimer Design of Thermally Activated Delayed Fluorescence Emitters. Tom Penfold

van Quantum tot Molecuul

09 Intro to Mass Dependent Fractionation

Simulating charge transfer in (organic) solar cells

The electronic structure of materials 1

Chemistry 3502/4502. Final Exam Part I. May 14, 2005

16.1 Molecular Vibrations

Recommended Reading: 23, 29 (3rd edition); 22, 29 (4th edition) Ch 102 Problem Set 7 Due: Thursday, June 1 Before Class. Problem 1 (1 points) Part A

Fluorescence Spectroscopy

Ab-initio modeling of opto-electronic properties of molecules in solvents and in proximity to a semiconductor nanoparticle

Experiment 6: Vibronic Absorption Spectrum of Molecular Iodine

The linear electron-phonon coupling model for molecular nonadiabatic ET. Simple derivations of the electron transfer rate

Chem 3502/4502 Physical Chemistry II (Quantum Mechanics) 3 Credits Spring Semester 2006 Christopher J. Cramer. Lecture 10, February 10, / 4

ELECTRON TRANSFER AND SINGLE MOLECULAR EVENTS. W. Schmickler Department of Theoretical Chemistry, University of Ulm, D Ulm, Germany

Quantum mechanics can be used to calculate any property of a molecule. The energy E of a wavefunction Ψ evaluated for the Hamiltonian H is,

Chem120a : Exam 3 (Chem Bio) Solutions

CHAPTER 13 Molecular Spectroscopy 2: Electronic Transitions

Example questions for Molecular modelling (Level 4) Dr. Adrian Mulholland

Vibronic Coupling in Quantum Wires: Applications to Polydiacetylene

Optical Properties of Lattice Vibrations

Molecular spectroscopy Multispectral imaging (FAFF 020, FYST29) fall 2017

PHOTOCHEMISTRY NOTES - 1 -

Semiclassical molecular dynamics simulations of excited state double-proton transfer in 7-azaindole dimers

Coherence Vibrations and Electronic Excitation Dynamics in Molecular Aggregates and Photosynthetic Pigment-Proteins

Introduction to Solving the Time- Dependent Schrödinger Equation. Tom Penfold

Diatomic Molecules. 7th May Hydrogen Molecule: Born-Oppenheimer Approximation

Introduction to Franck-Condon Factors

1+e θvib/t +e 2θvib/T +

October Entrance Examination: Condensed Matter Multiple choice quizzes

Chemistry 3502/4502. Final Exam Part I. May 14, 2005

How Does the Solvent Control Electron Transfer? Experimental and Theoretical Studies of the Simplest Charge Transfer Reaction

Chapter 6 Chemical Reactivity and Mechanisms

Introduction to Vibrational Spectroscopy

Theoretical models for the solvent effect

Introduction to Relaxation Theory James Keeler

Physics 828 Problem Set 7 Due Wednesday 02/24/2010

Physical Chemistry Laboratory II (CHEM 337) EXPT 9 3: Vibronic Spectrum of Iodine (I2)

Physics 541: Condensed Matter Physics

Isotope-same element (same atomic #), different # of neutrons so mass is different

Study guide for AP test on TOPIC 1 Matter & Measurement

Structural Bioinformatics (C3210) Molecular Mechanics

Chem 232. Representation of Reaction Mechanisms. A Simple Guide to "Arrow Pushing"

Infrared Spectroscopy

Chapter 12 What Is Temperature? What Is Heat Capacity?

Part 1. Answer 7 of the following 8 questions. If you answer more than 7 cross out the one you wish not to be graded. 12 points each.

Transcription:

Marcus Theory for Electron Transfer a short introduction Minoia Andrea MPIP - Journal Club -Mainz - January 29, 2008 1

Contents 1 Intro 1 2 History and Concepts 2 2.1 Frank-Condon principle applied to Electron Transfer..... 3 2.2 The Marcus Correction...................... 4 2.3 Potential energy surfaces: adiabatic and non-adiabatic process 5 2.4 Classical and semi-classical ET................. 6 3 Classical electron transfer theory 6 4 The semi-classical formulation 9 5 Applications 12 6 References 12 1 Intro Electron transfer (ET) is one of the most important chemical process in nature and it plays a central role in many biological, physical and chemical (both organic and inorganic) systems. In nature ET occurs in photosynthetic reaction center where transfer of electrons is used to create charge imbalance across a membrane, originating a proton pumping mechanism to produce ATP. In chemical system, ET at the metal surface with oxygen is responsible of the corrosion. Solid state electronics depends on the control of the ET in semiconductors and the new area of molecular electronics depends critically on the understanding and the control of the transfer of electrons in and between molecules. Figure 1 shows some fields related to the ET. 2 History and Concepts The main contribute to the ET theory comes from Marcus, Nobel for the chemistry in the 1992, which has given a first formulation of the theory in 1956[1]. Early experiments in the electron transfer field were on isotopic exchange reactions (self-exchange reactions) and later cross reactions. Equation 2.1 shows a typical self exchange process, where the asterisk denote a radioactive isotope. Fe 2+ + Fe 3+ Fe 3+ + Fe 2+ (2.1) 2

Figure 1.1: Examples of topics in the electron transfer field. These experiments originate by the huge amount of radioactive isotopes available after the Second World War and by the introduction of new instrumentation which permitted to study the rates of rapid chemical reactions such the electron transfer. Before 1940s only slow electrochemical reactions, such as the discharge of an H 3 O + ion at en electrode to form H 2, had been investigate extensively. For Marcus the inspiration to develop his theory was a paper of Bill Libby in which Franck-Condon principle was used to explain why reactions, as in eq. 2.1, between pairs of small cations in aqueous solution are relatively slow with respect the same electron transfer reactions involving larger ions, such as Fe(CN) 4 6. In contrast with the classical chemical reaction (eq. 2) in simple electron transfer reactions no chemical bonds are broken or formed, so a different picture is needed for the reaction-rate for electron transfer. AB + C A + BC (2.2) 2.1 Frank-Condon principle applied to Electron Transfer In his paper in 1952 Libby noted that when an electron transfer from one reacting ion or molecule to another, the two new ions or molecules formed 3

Figure 2.1: Typical nuclear configurations for reactants, products and surrounding solvent molecules in reaction 2.1. The longer M-OH 2 bond length in the state +2 is indicated schematically by the larger ionic radius. are in the wrong environment of the solvent molecules since the nuclei do not have time to move during the fast electron jump. Looking at the electron transfer reaction 2.1, after the electron jump the Fe 2 + ion would be formed in some configuration of the many nearby dipolar solvent molecules that was appropriate to the original Fe 3 + ion. This introduce a solvatation energy barrier for the process (Figure 2.1). The solvatation energy barrier is not the only mechanism playing an important role in the electron transfer reaction rate: the self exchange reaction 2.3 involve big ions but the experimental rate for the reaction is really slow, in contrast to the picture of a ET governed by the solvatation energy barrier. [Co(NH 3 ) 6 ] 2+ + [Co(NH 3 ) 6 ] 3+ [Co(NH 3 ) 6 ] 3+ + [Co(NH 3 ) 6 ] 2+ (2.3) The missing ingredient to explain the slow rate of the reaction 2.3 is the fact that there is a dramatic difference in the equilibrium Co-N bond length in the 3+ and 2+ ions so that each ions would be formed in a very foreign configuration of the vibrational coordinates. It is clear then that electron transfer implies changing in the chemical structure of the reactants. To understand how the Franck-Condon principle is used, is better to recall its classical definition in spectroscopy: an electronic transition is most likely to occur without changes in the positions of the nuclei in the molecular entity and its environment. The resulting state is called a Franck-Condon state, and the transition involved, a vertical transition. As electronic transitions, electron transfers are instantaneous compared to the motion of the nuclei of the molecules or ions involved in the process and of the orientation of the molecules in the medium (e.g. solvent molecules). The foreign environment for the new electronic state after the electronic jump can be seen as an energetic barrier for the ET process. 4

Figure 2.2: Potential energy surface for the reactants and the products 2.2 The Marcus Correction The idea of Libby turns out to be incorrect, or incomplete, because for reactions occurring in the dark the energy is not conserved: the ions would be formed in the wrong high-energy environment, but the only way such a non conserving event could happen would be by absorption of light (vertical transition) and not in the dark. To clarify let s consider the energy transfer reaction 2.4. [Fe(H 2 O) 6 ] 2+ + [Fe(H 2 O) 6 ] 3+ [Fe(H 2 O) 6 ] 3+ + [Fe(H 2 O) 6 ] 2+ (2.4) Here the Fe-O equilibrium distance in the ion Fe(II) is 2.21Åwhile is 2.05Åfor the ion Fe(III). Scheme 2.2 shows the potential energy surfaces for the reactants and the products: If the electron transfer takes place at their equilibrium positions, then we would have compressed Fe(II) ion and a stretched Fe(III) ion. These are in vibrational excited states and would release energy: if the ET occurs in dark, then the energy is not conserved. Thus the reagents must match their energies before electron transfer can occur. In the limit of weak electron interactions between the reactants both the Franck-Condon principle (vertical transition) and energy conservation must be satisfied. Fluctuations had to occur in the various nuclear coordinations as well as in the orientation coordinates of the solvent molecules and in any other coordi- 5

Figure 2.3: Potential energy surface for the reactants and the products nates whose most probable distribution for the products differs from that of the reactants. With such fluctuations, values of the coordinates (i.e. reaction coordinates) could be reached which satisfy both the Franck-Condon and energy conservation conditions so to permit ET to occur in the dark. 2.3 Potential energy surfaces: adiabatic and non-adiabatic process The potential energy of the system is a function of the translational, rotational and vibrational coordinates of the reacting species and of the molecules of the surrounding medium. A profile of the potential energy surface is given in figure 2.3. The abscissa, a line draw in the many-dimensional coordinates space introduced before, represents any concerted motion (fluctuation) leading from any spatial configuration (on all atoms) that is suited to the electronic structure of the reactants to one suited to that for the products. Surface R denotes the potential energy profile when the reacting species have the electronic structure of the reactants, while surface P corresponds to their having the electronic structure of the products. If the distance between the reactants is sufficiently small there is the usual splitting of the two surfaces at their intersection. If the electronic interaction is very weak the splitting at the intersection of the two surfaces is negligible. In the first case the system will always remain on the lowest surface as it moves from left to right in figure 2.3. Thus the system has moved from the R to the P surface adiabatically. On the other hand, if the splitting is negligible, a system initially on 6

Figure 3.1: Potential energy surface for the reactants and the products the surface R will tend to remain on R as it passes to the right across the intersection. The probability that as a result of this nuclear motion the system ends up on surface P is calculated by treating this motion non-adiabatically. Please note that the potential curves can be approximate to a parabola. 2.4 Classical and semi-classical ET Considering figure 2.3, it should be noted that the system can undergo this electron transfer either by surmounting the barrier if it has enough energy (classical) or by tunneling (semi-classical). If nuclear tunneling is unimportant, the rate of the ET reaction is simply the rate of crossing the barrier top times its relative population. The top of the barrier is then the activation energy for the electron transfer. 3 Classical electron transfer theory The free energy barrier, G (figure 3) is the difference between the crossing point (for diabatic process) and the bottom of the reactant free energy parabola. If the entropy changes are ignored, the free energy becomes energy or potential energy. Referring to figure 3 the barrier top occurs at the crossing (position q c ) and is given by the conditions 3.1. V R (q c ) = V P (q c ) (3.1) In the approximation of parabolic potential energy, the condition 3.1 can be rewritten in terms of the position of the minima for the potential energy of 7

the reactants and the products, q R and q P respectively. 1 2 f (q c q R ) 2 = G 0 + 1 2 (q c q P ) 2 (3.2) We can solve for the crossing the previous equation obtaining q c = G0 f 1 (q c q P ) + (q P + q R ) 2 (3.3) A fundamental physical quantity is the reorganization energy, λ which can be defined as λ = 1 2 f (q R q P ) 2 (3.4) This energy increases with increasing the separation between the minima corresponding to the initial and final state and with increasing the stiffness or force constant f. The potential energy barrier is thus given by: G = G R (q c ) G R (q R ) = V R (q c ) V R (q r ) = frac12f (q c q R ) 2 (3.5) Substituting for q c the expression in eq 3.3 G = λ ] 2 [1 + G0 (3.6) 4 λ This express the barrier height or the free energy activation for crossing two adiabatic curves in terms of the overall free energy of reaction G 0 and the reorganization energy λ. The expression for the rate constant k et of the reaction is given from the standard Arrhenius reletionship in eq 3.7. [ ] G k et = A exp (3.7) K B T The prefactor A in equation 3.7 depends on the nature of the electron transfer reaction (e.g., bimolecular or intramolecular). G 0 is the standard free energy for the reactions (and equals to zero for a self-exchange reaction) and λ is a reorganization term composed of solvational (λ 0 ) and vibrational (λ i ) components: λ = λ 0 + λ i (3.8) The standard estimation for solvational energy in eq 3.9 was obtained by Marcus by using a model in which reactants and products were modeled as spheres and the solvent as a dielectric continuum. λ 0 = ( e) 2 [ 1 2a 1 + 1 2a 2 1 R 8 ][ 1 ǫ 1 ǫ 0 ] (3.9)

Here a 1, a2, R, ǫ and ǫ 0 are respectively the radii of the donor and acceptor in the Marcus two spheres model, the distance between the centers of the donor and acceptor spheres and the optical and static dielectric constants. e is the amount of charge transferred. The difference in the inverse dielectric constants relates to the fact that nuclear degrees of freedom cannot readjust instantaneously to the motion of the electrons and thus contribute to the barrier. This is a manifestation of the Born-Oppenheimer separation. The vibrational contribute to the reorganization energy λ i is given by the equation 3.10. λ i = λ i,l = 1 f l ( q e,l ) 2 (3.10) 2 l l Here, the summation is over the coupled intramolecular vibrations. The contribution of the lth normal mode to the reorganization energy is given in terms of its constant force f l and the change in equilibrium positions between the reactants and the products, q e,l = q P,e q R,e. As mentioned above, in case of self-exchange reactions the driving force G 0 vanishes and equation 3.7 becomes: [ ] λ k et ( G 0 = 0) = A exp K B T (3.11) and the activation free energy for self-exchange ET is simply one-fourth of the reorganization energy. A more extensive mathematical treatment can be found in the paper of Newton.[3] 4 The semi-classical formulation The equations 3.7 and 3.11 completely characterize the electron transfer reaction in terms of three quantities: the prefactor A, the free energy G 0 and the reorganization energy λ. On the other way, the theory as presented before shows some limits, and the most important is probably the temperature dependence. Equation 3.7, in fact, predicts a vanishing electron transfer rate constant at zero temperature, in contrast with experiments (Figure 4). This and other problems arise from the classical assumption that the energy barrier must be crossed (overcome) while, in reality, it exists the possibility to have tunneling through the barrier. A semi-classical approach has been addressed by a number of workers to include quantum effects into the classical Marcus theory. Experimental results in figure 4 shows that the predictions of eq. 3.7 are verified at high temperature, where the electron transfer is an activated process. Relation 4.1 provides a guide to which modes l coupled 9

Figure 4.1: Comparison of theories with data on C. vinosum cytochrome oxidation. Data are points and various theoretical models yield the curves. Note the nonvanishing rate constant at low temperature. to electron transfer must be treated as quantum mechanical and which can be assumed classical. hω l K b T 1 quantum hω l 1 K b T classical (4.1) The necessity to introduce quantum effects become clear introducing the harmonic oscillator vibrational energy levels within the potential curves of figure 2.3, as shown in figure 4. As shown there, tunneling can occur between the ground level in the reactant potential and the ninth vibrational level in the product curve. The extent to which tunneling plays a role depends on the extent of the vibrational overlap between the initial and the final states. The rate constant for electron transfer, from the Fermi s golden rule of perturbation theory with the full Hamiltonian of the system, gives the ET rate constant K E T as in eq. 4.2. k ET = 2π h < R H P > 2 < r vib p vib > 2 δ (E R E P ) (4.2) = 2π h < R H P > 2 < r vib p vib > ρ(e P ) (4.3) = 2π h < R H P > 2 (DFWC) (4.4) = 2π h H RP (DFWC) (4.5) 10

Figure 4.2: Schematic one-dimensional representation of the intersection between reactant and product vibrational levels. Here the splitting at the crossing is not shown and the optimal overlap of the ν = 0 initial vibrational level is with the ν = 9 product level. 11

Here the electron transfer rate constant is given first in terms of isolated levels for the reactant and product, then in terms of density of states,ρ(e P ) of the product, and finally in terms of density of states weighted Franck-Condon factor (DWFC). H RP is the electronic matrix element that mixes the donor and acceptor states. The states r vib > and p vib > are vibrational states of reactant and product respectively. H is the Hamiltonian for the entire system, and the matrix elements in eq. 4.2 and 4.3 are averaged over the full Hamiltonian of the system. In the case of non-adabatic ET, with the Condon approximation this formulation is correct, while for adiabatic ET a more general prefactor is required. We focus now on the case of non-adiabatic ET. Using the polaron model, each electronic state is coupled with a number of vibration, treated as harmonic oscillators with separation of nuclear and electronic coordinates assumed (i.e. in Born-Oppenheimer approximation regime). Under these conditions, the Hamiltonian for the system can be written as: [ H = R >< R ER 0 + 1 f l (q l q R,l ) ]+ 2 2 l [ P >< P EP 0 + 1 f l (q l q P,l ) ]+ 2 2 l ( R >< P + P >< R )H RP (4.6) Here, the first term in braces is the energy represented by the left parabola of figure 4, and the second set of braces is the same energy term for the right potential. ER 0 and E0 P are respectively the energy origin for these states and are characterized by the force constant f l for the lth mode, whose displacement origin is q R,L and q P,l in the reactants and products respectively. This Hamiltonian is called polaron models and consists in two electronic states coupled with a number of vibrational levels, which can be either classical or quantum mechanical, according to relation 4.1. In addition to the classical solvent motion, one can specifies one coupled vibration characterized by frequency ω and equilibrium displacement q e, then is possible to define the intramolecular reorganization energy for this mode as λ i = ( f 2 )( q e) 2 (4.7) This can be related to a related adimensional quantity: the electron vibrational coupling constant, or Huang-Rhys factor as define in eq. 4.8. S = λ i hω (4.8) 12

The generalization of equations 3.7 for non-adiabatic ET for one coupled mode in the quantum mechanical limits is then: k ET = 2π h ( ) 1/2 1 H2 RP (FC) (4.9) 4πλ 0 K B T where FC is the Franck-Condon factor (FC) = [ exp( S) Sν ν! exp (λ0 + ν hω ] + G 0 ) 2 4λ ν 0 K B T (4.10) The term in front the FC factor is the frequency of electron transfer in the absence of barrier and contains H RP and the classical density of states. The FC factor consists of the sum over all possible vibrational overlap integrals between the initial vibrational level ν and the final level. ν. Each individual ν represents a separate ν = 0 ν reaction channel. Each separated exponential term in the sum is the population of molecules having the required energy to undergo electron transfer with energy conservation through channel ν = 0 ν. The sum is dominated by those channels for which G 0 λ 0 + ν hω, so there is a close energy match between the energy released ( G 0 ) and the sum of the reorganization energy and the initial product vibrational energy (ν hω). A more complete treatment of the semi-classical ET theory can be found in papers [2] and [4]. 5 Applications 6 References 1. Marcus, R.A., Rev. Modern Physics, 65 (1993), 3, 599-610 2. Barbara P.F., J. Phys. Chem. 1996, 100,13148-13161 3. Newton, M.D., Chem. Ren, 1991, 91, 767-792 4. Jortner, J. and Freed, K.F. J. Chem. Phys. 1970,52, 6272-6291 13