Download Quantum effects in chemistry - Fritz Haber Center for Molecular

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Copenhagen interpretation wikipedia , lookup

Probability amplitude wikipedia , lookup

Molecular Hamiltonian wikipedia , lookup

Path integral formulation wikipedia , lookup

Scalar field theory wikipedia , lookup

Quantum field theory wikipedia , lookup

Atomic theory wikipedia , lookup

Measurement in quantum mechanics wikipedia , lookup

Relativistic quantum mechanics wikipedia , lookup

Quantum dot wikipedia , lookup

Quantum fiction wikipedia , lookup

Wave–particle duality wikipedia , lookup

Quantum entanglement wikipedia , lookup

Bohr model wikipedia , lookup

Coherent states wikipedia , lookup

Max Born wikipedia , lookup

Renormalization wikipedia , lookup

Quantum computing wikipedia , lookup

Many-worlds interpretation wikipedia , lookup

Particle in a box wikipedia , lookup

Renormalization group wikipedia , lookup

Quantum teleportation wikipedia , lookup

Symmetry in quantum mechanics wikipedia , lookup

Density matrix wikipedia , lookup

Quantum electrodynamics wikipedia , lookup

Theoretical and experimental justification for the Schrödinger equation wikipedia , lookup

Franck–Condon principle wikipedia , lookup

Quantum machine learning wikipedia , lookup

EPR paradox wikipedia , lookup

Interpretations of quantum mechanics wikipedia , lookup

Electron configuration wikipedia , lookup

Orchestrated objective reduction wikipedia , lookup

Hydrogen atom wikipedia , lookup

Quantum key distribution wikipedia , lookup

Canonical quantization wikipedia , lookup

Quantum group wikipedia , lookup

History of quantum field theory wikipedia , lookup

Quantum state wikipedia , lookup

Quantum decoherence wikipedia , lookup

Hidden variable theory wikipedia , lookup

T-symmetry wikipedia , lookup

Transcript
Available online at www.sciencedirect.com
Procedia Chemistry 3 (2011) 63–81
22nd Solvay Conference on Chemistry
Quantum effects in chemistry: seven sample situations
Mark A. Ratnera, Ronnie Kosloffb
a
Northwestern University, Department of Chemistry, 2145 Sheridan Road, Evanston, IL, 60208, USA
b
Fritz Haber Institute, Hebrew University, Jerusalem 91904, Israel
Abstract
Significant quantum effects in chemistry range from static structure (electronic and geometrical) through dynamical
behavior, including optical properties, conductance, relaxation, decoherence, and thermalization. We outline seven
situations in which molecular systems exhibit ineluctably quantum behavior. These range from situations in which
the community can understand the problem quantitatively and conceptually (for example for dilute sets of spins in
NMR) to femtosecond/attosecond situations, which the community understands only primitively. In condensed
phase, the dynamics will always evolve in a system/bath environment, and we discuss here how to pose, and to start
understanding, problems of that sort.
Keywords: coherence; relaxation; electron transfer; excitons; energy transfer; electron transport
1. Quantum Chemistry
Historically, the first applications of quantum mechanics in chemistry occurred within two years of the introduction
of quantum mechanics, with the formulation of the Born-Oppenheimer separation between nuclear and electronic
motion [1], and the treatment of the hydrogen molecule by Heitler and London [2]. In the eighty-three years since
that time, the issue of quantum effects in chemistry has taken on many meanings. This Solvay Conference examines
the challenges that occur due to quantum effects in chemistry, in the context of the experimental and conceptual
landscape of 2010. In this short article, I discuss six special cases, in each of which quantum effects are important,
and then conclude (Section 8) with some technical but significant issues that arise in the general description of
quantum dynamical molecular evolution.
This first section deals very briefly with the historically important case of quantum chemistry, which is the major
success story in the area of quantum effects in chemistry. Sections 2-5 outline specific conceptual and experimental
steps that have been taken by presenters at the Solvay Conference, within the particular focus session for which this
paper is the rapporteur. We address (Sections 6-8) some aspects of quantum dynamical evolution, including
discussion of a non-equilibrium, open system problem as represented by molecular transport junctions.
Quantum chemistry is the 800-pound gorilla in the conference room. Starting from the work in 1927, it has
advanced steadily, and must now be counted as among the greatest successes in the theoretical chemistry of the 20th
century. Partly this success is due to the obvious importance of the field, since chemists always have cared about
the structure and stability of molecules. Partly it is because of the advent of electronic computers that permitted the
appropriate calculations to be made. Partly it is because of the advances in crystallography, electron scattering, and
1876–6196 © 2011 Published by Elsevier Ltd. Selection and/or peer-review under responsibility of G.R. Fleming, G.D. Scholes and A. De Wit
doi:10.1016/j.proche.2011.08.013
64
Mark A. Ratner and Ronnie Kosloff / Procedia Chemistry 3 (2011) 63–81
spectroscopy that permitted structures of small and medium sized molecules to be analyzed very accurately. Finally,
it succeeded because of very dedicated work by a number of brilliant individuals over eight decades.
It is also possible to ascribe the success that quantum chemistry has achieved to the fact that the problem was very
clearly defined. In terms of fundamental constants, the problem of quantum chemical electronic structure
0).
calculations is to solve the stationary state Schrödinger equation in a limit (kT0, cinfinity, me/m0, Here, k, T, c, me, m, and peed of light, electron mass,
mass of characteristic atom, and the density of the sample. These conditions correspond to zero temperature,
nonrelativistic, frozen nucleus, infinitely dilute samples. Under these conditions, one can define computational
“model chemistries” for particular categories of molecules. Work in the last three decades of the twentieth century
brought about the concept of “model chemistries” [3] and great success in so doing: if one now chooses to calculate
the geometry of a newly synthesized molecule containing (say) fourteen carbons, twenty hydrogens, and a couple of
oxygens and nitrogens, the “model chemistry” corresponding to MP/2 solution to the static Schrödinger equation
using a 6-31G++-** basis set will give the geometry accurately to 0.02 Å in bond length, and with comparable
accuracy in bond angles.
The successes in solving these “model chemistries” are huge, and the influence in the chemical sciences
appropriately large. Most organic chemists (and people well beyond chemistry including geologists, biologists,
physicists, and engineers) now utilize electronic structure calculations to find geometries and suggest properties [4].
1.1. Outline: Quantum Dynamical Effects in Dense Systems
The rapporteur function of this paper involves discussion of four elegant contributions, that focus on quantum
effects beyond the stationary ground state of isolated molecules. These are largely short-time phenomena, because
for electronic systems interacting with a dense-states bath, decoherence will always cause the loss of phase
dependent behavior on long enough time scales. In most of these situations, the assumptions made above about
isolated molecules are no longer true (kT e/m in its earliest days, when people were interested in liquids and solids. This attention to condensed phase matter has
become crucial to contemporary chemistry and to its applications in collaborative ventures with other sciences
including physics, materials science, chemical engineering, and biology [5]. Since coherent quantum evolution is
largely a short time phenomenon, our understanding of its processes and behaviors has advanced with the
availability of accurate short time measurement. From the nanosecond timescale of flash photolysis through the
picosecond timescales that were called “ultrafast” in the 1970s (using etalons, and the beginnings of fast laser
spectroscopy) to the femtosecond timescale that was made available and relatively straightforward by the advent of
titanium sapphire lasers through to the attosecond timescale that will challenge our understanding of the dynamics
surrounding the Born-Oppenheimer principle, and is already beginning to be pursued [6].
The protocols applied to make measurements using these short-time techniques involve either multi-pulse
excitation/detection, multidimensional correlation spectroscopy, or single molecule spectroscopy.
In most of these situations, we want partial information, since in condensed matter there may be 1020 atoms or so in
the studied volume and its intimately coupled neighborhood. Since full theoretical quantum dynamics on this
number of particles is both unnecessary and impossible, the theoretical techniques used to interpret the data are
those appropriate to situations in which we are interested in the behavior of a subsystem coupled to an environment.
The most popular such approach is the density matrix technique [7-11], where the trace over the bath gives the
reduced density matrix corresponding to the system and its dynamics. The second approach is the QM/MM scheme,
in which one treats the quantum mechanics and dynamics of the system subject to an external bath that is treated
classically [12,13]. Finally, the self-energy formalism evolving from nonequilibrium Green’s functions can be used
to describe the quantum statistical mechanics of subsystems [14-18]. The second of these is the most popular in the
chemical community, the first is the most popular for specifically pursuing dynamical quantum effects in molecular
processes, and the third is the least common, but perhaps the most powerful. Remarks on these will be made in
Sections 6-8.
Mark A. Ratner and Ronnie Kosloff / Procedia Chemistry 3 (2011) 63–81
65
The next four sections overview the contributions of the other speakers in this session, and suggest some of the
challenges that arise in trying to understand the evolution of molecules in the quantum dynamics regime.
2. The contribution from the Kauffmann Group
The contribution from the Kauffmann Group deals with “the choreography of functionally important modes in
optical dynamics of complex molecules”. This contribution utilizes two-dimensional electron correlation
spectroscopy to access quantum kinetics, including the intramolecular vibrational relaxation regime. The systems
studied are large molecules, including phthalocyanines, carotenoid/purpurine dyads, and pinacyanol chloride.
Vibronic coupling pathways dictate both vibrational transfer and the redistribution, accessing the IVR regime.
This work uses multidimensional spectroscopy to investigate nuclear dynamics following an electronic excitation,
and the evolution of quantum wave packets and their dynamics over short time scales up to, say, one picosecond.
As part of this investigation, the Kauffmann Group works [19,20] in the sub ten- femtosecond regime to examine
energy transfer in these dyads, observing strongly damped, off diagonal, oscillatory signals as would be expected for
situations with this sort of intersubunit coupling.
3. The contribution from the Wolf Group
The contribution from the Wolf Group extends this molecular emphasis to examine interfacial dynamics. It uses the
technique of two-photon photoemission, which has been beautifully developed by the Berlin group [21], combined
with surface science techniques to study the dynamics following photoinjection of electrons from a metal into the
adsorbate. The subsequent solvation and localization of the electron within the adlayer is the issue to be examined
and illuminated.
This research is the successor to the work [22,23] on the time-dependent Stokes shifts carried out in the 1970’s and
1980’s, to determine dynamical relaxation of nuclear modes around a newly-introduced charged species. Here,
however, the photoinjection of the bare electron gives a cleaner system, because the counterion (a hole in this case)
is screened by the electrons of the metal. The full dynamics involving molecular dipole reorientation and inertial
motion can then be examined. The observation is that the behavior depends strongly on the phase ordering in the
adlayer. In crystalline structures, Wolf’s group sees relaxation processes over a time scale of 1012 times longer than
in amorphous materials! This huge difference is explained in terms of the relevant energy landscapes that the
electron would explore in fully crystalline versus fully amorphous materials.
These lovely experiments characterize nuclear relaxation phenomena around excess charge, that is introduced very
rapidly into a specific place in an adlayer. The characterization of the dynamics, and its understanding in terms both
of the motions and the static and dynamic disorder of the nuclei in the adlayer, is characteristic of the best work on
interfacial dynamics.
4. The contribution from the Scholes Lab
The contribution from the Scholes Lab in Toronto comes from one of the leaders in investigating very short-time
quantum coherences in real systems [24-26]. In this contribution, they focus on conjugated polymers, which are of
interest both fundamentally and because of their applications in bulk heterojunction molecular photovoltaic cells
[27-29]. It has been clearly established now (and follows from some arguments originally given by Flory half a
century ago [30]) that conformational subunits in conjugated polymers are linked by regions of disorder, and that the
energy moves among these conformational subunits by a Förster-like process [31], migrating on characteristic
picosecond time scales, and winding up in the energy sink (the site with the reddest adsorption structure). This can
66
Mark A. Ratner and Ronnie Kosloff / Procedia Chemistry 3 (2011) 63–81
be understood utilizing simple models, that extend the two-site Förster mechanism to multiple sites along the
disordered structure of the conjugated polymer [32].
At shorter time scales, the Scholes group and others [33] have shown that the dynamics among the photoexcited
conformational subunits exhibit quantum coherence in the transport of excitation between nearby subunits.
Probably most interestingly, vibrational motions along the conjugated chains themselves seem to delocalize the
excitation energy. Scholes presents both the observations and a mechanistic argument about how the chemical
bonds introduce quantum effects in the dynamics of energy transfer.
Energy transfer has become the area in which some of the most striking and elegant spectroscopic studies have been
done, studies that indicate the appearance of quantum coherence in situations in which it would not be expected
[34]. While some of these observations can be understood using density matrix theory and some special assumptions
about the nature of the relaxation processes [35], their observation in several different systems, in several different
laboratories, suggests that the standard arguments from complete and careful simulations of systems like the
solvated electron [36,37] or isolated chromophore dynamics in solution [38] are not necessarily relevant to large
molecular species. (If they were, the conclusion that follows from the work of many authors that in polar solvents
decoherence is essentially complete on the time scale of a few tens of femtoseconds [39] might preclude some of the
long-time coherences observed in polymers, and in biological and biomimetic systems).
5. The contribution by the Waldeck Lab
The contribution by the Waldeck Lab deals with some fundamental conceptual issues involving pathways for
coherent transfer between molecular sites. Effectively, the authors argue [40-42] that if one chooses to measure and
describe electronic tunneling from a given site to another within a molecule, many different quantum mechanical
pathways might be accessed. Nuclear reorganization can evolve, and eventually an optimal tunneling geometry
(corresponding classically to the activated complex picture of Eyring [43]) can be found. Then, conceptually, the
electron is imagined to tunnel through a barrier presented by the nuclear geometry. The tunneling can occur by a
single pathway, but might occur by several different pathways, and quantum phase interferences among those
pathways can occur in the propagation scheme [44-50]. This quantum interference among coupling pathways can
be important in electron transfer kinetics. Decoherence among these pathways can arise from inelastic effects in the
molecules, or from motions that describe either pure dephasing (T2*) or energy relaxation (T1). As examples, the
authors discuss biological and biomimetic systems, where nuclear fluctuations, as might be expected, help to
determine the nature of the bridges and which ones are utilized.
This work deals not so much with coherence along a single quantum trajectory, but rather with the phenomenon of
phase interference, familiar from the two-slit problem. Such phase interferences are responsible for some of the
mechanistic regularities of organic chemistry, such as the very weak coupling between meta-substituted sites in
benzene, compared to the much stronger interaction between sites that are para or ortho to one another. Indeed, a
significant part of physical organic chemistry is based on arguments concerning coherence and phase cancellation
[44,51].
The understanding of pathways is highly intuitive to most chemists; this contribution focuses on the concepts
involved, and the situations under which coherence effects can be observed clearly, or molecular motion can lead to
decoherence. Such concepts have been recently investigated experimentally in specially-designed two-path systems
[52], and their explanation in terms of the shaky pathways scheme [53] seems quite powerful.
These four contributions suggest four of the most important areas in which coherences and quantum effects are now
seen in condensed phase chemistry (vibrational interactions with electronic states, injection at surfaces and structural
relaxations around introduced charges, coherence effects in energy transfers in disordered and model systems, and
interference among different pathways in electron tunneling processes).
Mark A. Ratner and Ronnie Kosloff / Procedia Chemistry 3 (2011) 63–81
67
6. Density Matrices, Partial Information, and Hot Injection
The introductory article by Rice covered many of the issues confronting our current understanding and formulation
of quantum effects in dynamical systems. In Sections 6-8, we suggest a few situations, in which such difficulties
can occur. In this chapter, we will deal with situations in which we wish only partial information on a subsystem,
interacting with a larger heat bath. These correspond to closed systems in the sense that the number of electrons, nel,
within the system, is fixed. Energy can flow in and out of the system, and in this sense it resembles the canonical
ensemble of classical statistical mechanics.
Under these conditions, we can represent the density matrix operator by Eq. 1, where i and j label arbitrary basis
functions. The full density matrix describes the full system, and its dynamics are fixed by the quantum Liouville
equation, Eq. 2. We are generally interested only in the dynamics of a subsystem (e.g. a molecule dissolved in
fluid). We then [7-11] define the reduced density matrix by Eq. 3, where is the reduced density matrix
(depending on system parameters only) and the trace is taken over all the states of the bath that enter into Eq. 1.
(1)
i, j i i, j j
(2)
i d / dt [ H , ]
(3)
Tr B ( )
H total H system H bath H int eraction
(4)
(5)
i d / dt [ H system , ] i d / d t ) d
It then follows that, if we represent the total Hamiltonian of the physical sample by Eq. 4, the evolution of the
reduced density matrix can be written as Eq. 5. Here the d subscript in the last term implies dissipative behavior,
behavior corresponding to decoherence and relaxation within the quantum subsystem, caused by the last two terms
in Eq. 4.
There are many formulations of the dissipative term, some of which are Markovian and quite general (usually called
Lindblad models [54-58]), whereas others are approximated at second order, in the expansion of the density matrix
(Redfield theory [59]). More recent treatments have been developed. One can indeed expand to higher orders in
the density matrix, resulting in diagrammatic expansions using double sided Green’s functions. This has been
elegantly introduced and utilized by Mukamel and his group [10].
Sophisticated analyses using these schemes are common in descriptions of condensed phase chemical dynamics.
But as was pointed out in a significant paper [55] by Tannor and collaborators, each of these schemes suffers from
one or another fundamental inadequacy (Redfield
equations can give negative populations, Lindblad
approximations are only accurate up to an unknown intensity parameter, and are not necessarily translationally
invariant). Nevertheless, sophisticated analyses using these schemes are indeed common in descriptions of
condensed phase chemical dynamics.
To introduce multilevel, not-necessarily Markov relaxation process, Kosloff has introduced the so-called stochastic
surrogate model Hamiltonian [60,61]. This is represented by Eq. 6, where B and B’ represent the primary bath (here
taken as two-level systems rather than harmonic oscillators) and the secondary bath (also consisting of two-level
systems).
H
total
H S H B H SB H B ' H
BB '
(6)
68
Mark A. Ratner and Ronnie Kosloff / Procedia Chemistry 3 (2011) 63–81
The secondary bath B’ couples to the primary bath, and effectively gives the elements of the primary bath a lifetime
(a similar scheme was utilized by Subotnik and collaborators [62] in constructing a model for molecular transport
junctions utilizing a density matrix approach – see Section 7). In this formulation, the HBB’ effectively gives the
states of the bath represented by B a lifetime. Kosloff has developed a convenient swapping scheme that describes
the full time evolution of the system subject to these two baths. This avoids any Markovian assumption, and by
appropriate selection of the system’s matrix element in the HSB term, one can propagate computationally the system
Hamiltonian, and deduce the effects of bath relaxation and decoherence. This is a promising scheme, because it is
computationally relatively straightforward, and yet is non-Markov, and readily interpreted in a physical way.
As an artificial example, we consider [63] the photodynamics in a three-state problem. Fig.1 gives the three lowest
potential energy surfaces of an anharmonic diatomic molecule. Upon photoexcitation from the ground state to the
acceptor (bright) state labeled b, the wave packet starts to evolve on the excited state, and eventually crosses the
curve into the dark state, or acceptor state, a. When the wave packet hits the outer wall of the potential
corresponding to the dark state, it is reflected. If, now, it can decohere before it gets back to the crossing between
the bright and dark states, then (following the initially nonintuitive but quantum mechanically correct) argument
given in the classic review [64], the system simply cannot cross. Therefore, if the bottom of the dark state potential
is higher than the bottom of the bright state potential, the excess energy denoted EB in Fig.1 can be captured by the
dark state, and remain there. The energy EB is thus prevented from thermalizing, and if injection of the electron
from the dark state into the electrode then follows, the thermal excitation within the initial Franck-Condon vertical
excitation from the ground state to the bright state can be partially trapped; EB represents this trapped energy. It is,
however, no longer in the vibrational manifold, but rather in the electronic energy of the dark state. Subsequent
injection into the electrode thus captures not only the energy corresponding to the bright state minimum, but also the
excess vibrational energy which is accessed by the Franck-Condon, vertical excitation.
This sort of capture of energy before it can transform from electronic to vibrational (heat bath) was first suggested
by Nozik, who called it hot injection [65]. Nozik was mostly interested in semiconductor quantum dot systems, but
while the example presented here is artificial, nevertheless if the decoherence process is fast enough (a few
femtoseconds in this case), then indeed energy can be captured using this hot injection formulation.
Details of the argument here are given in Ref. [63]. Unpublished work from the Kosloff group suggests that in twodimensional systems, far better capture of the excess vibrational energy into the electronic manifold can occur, and
this could cause hot injection to be useful in organic photovoltaic systems. One interesting aspect of this dark state
capture of vibrational excitation is that, in this case, it is the decoherence rather than the coherence that permits
manipulation of energy capture.
The utilization of decoherence effects to improve evolution behavior has also been seen [66,67] in new work from a
very different part of the chemical dynamics interest spectrum. Plenio and coworkers, [67] and also Aspuru-Guzik
and co-workers [66], pursued the FMO complex that was observed [34] to exhibit long-lasting quantum coherences.
These investigators saw an unanticipated sort of resonance, when the energy gap frequency between states was
comparable to the decoherence rate. Under these conditions, the energy transfer efficiency can actually be enhanced
by the decoherence [64] ; in this sense, it resembles the “hot injection” situation, in that the decoherence actually
favors the process that is wished. Such interactions between spacings in the system and frequencies of decoherence
are reminiscent, also, of nmr lineshape behavior.
Mark A. Ratner and Ronnie Kosloff / Procedia Chemistry 3 (2011) 63–81
69
b
a
g
Fig.1. The simplest scheme for molecular “hot injection.” The bright state in red (b) holds vibrational energy according to the Franck-Condon
principle. Excess energy EB is captured for injection into the electrode, rather than becoming thermal waste.
7. Self Energies, Time Scales, and Some Molecular Transport Issues
Density matrix methods were originally developed for canonical, closed systems. They are, generally, the method
of choice for the analysis of quantum decoherence problems, because they so beautifully permit an understanding of
partial information. In open quantum systems, where the number of electrons is neither an integer nor a known
quantity, density matrix schemes are substantially more difficult to formulate [68]. Under these conditions, the most
attractive way to deal with understanding transport is to use the nonequilibrium Green’s function formulation. The
formalism here is a bit complex [14-18, 69], and in this section we will simply note some of the understandings that
have been gained, and mention issues such as time scales and interferences.
Experimentally, the molecular junction transport system has been extensively explored over the last twenty years
[70-73]. A large number of experimental test beds, many devolving from more traditional mesoscopic physics,
some from electrochemistry, have been used [74-80].
Fig.2. Schematic of electrochemical break junction device for conductance measurement of single molecules.
70
Mark A. Ratner and Ronnie Kosloff / Procedia Chemistry 3 (2011) 63–81
Fig.2 shows the electrochemical break junction scheme [78] that has the advantage of measuring current through a
molecular transport junction, but also of providing enough statistics to estimate significant physical issues including
(in particular) geometric variation. The simplest model for a molecular transport junction is shown in Fig. 3.
Fig.3. Highly schematic picture of transport junctions for measurement of molecular (a) conductance and quantum dot (b) conductance
Voltage is applied between two electrodes, with the molecule bound to both electrodes through individual binding
sites. This is the theoretical model that is nearly always used, and in many cases it is believed that the experiment
measures something very much like this.
The experimental and theoretical analysis of molecular transport junctions has become quite active, and with more
than a thousand papers published on the subject, this is not an appropriate place for review. These systems are
appropriate here, because they offer examples of open quantum systems interacting with baths, but the baths are not
the harmonic oscillator or two-level system baths that are appropriate for the modeling of quantum coherence,
decoherence, and dynamics in isolated (or solvated) molecular systems.
The fundamental understanding of what happens [81] in molecular transport was provided by Landauer [82]
Buttiker, and Imry [83]. They were concerned not with molecular transport, but with a similar problem in
mesoscopic systems. A simple picture of such systems is a quantum dot behavior, as sketched in Fig.3b: here the
two electrodes can cause charge to flow through the system consisting of the macroscopic electrodes and the
nanoscopic quantum dot. The Landauer/Imry (L/I) approach is based on the understanding that the decoherence and
loss of quantum effects are caused by a bath, but that bath comprises the electronic states of the leads – it is a Fermi
bath, not a Bose bath or classical bath. But it effectively relaxes the system, can cause full quantum decoherence,
and permits irreversible flow of current from the higher to the lower potential.
The Landauer argument is a lovely one, because it is so simple. The resulting Landauer/Imry (L/I) equation for the
zero-voltage current is
g g0
T
i ,i
channels , i
(7)
Here g0 = 2e2/h, where e is the electronic charge, h is Planck’s constant, and g0 is the fundamental quantum of
conductance (77.48 microsiemens). The transmission coefficients Tii describe behavior in channel i. Fundamentally,
Eq.7 says that “conductance is scattering” - as the electrons pass from the macroscopic lead into the nanoscopic/
mesoscopic system (be it a quantum dot, or an ultra-thin heterostructure layer, or a molecular entity), it can scatter
elastically, either into the product bath or back to the reactant bath. If in each channel only forward scattering
happens (so that its Tii is unity), summing over all transport channels that are transverse to the electrode pair gives
the number of channels times the quantum of conductance.
This is a remarkable construct: note that it has little to do with Ohm’s law, because here conductance is quantized,
and the only considered inelastic behavior is from the electrons in the leads. The most striking examples are seen
experimentally for metallic wires, as shown in Fig.4; the typical conductance quantization is seen, exactly at the
values suggested by the Landauer formula.
Mark A. Ratner and Ronnie Kosloff / Procedia Chemistry 3 (2011) 63–81
71
Fig.4. Observed conductance quantization in units of 2e2/h. From A. Yanson, GR Bollinger, H, van den Brom, N. Agrait, JM van Ruitenbeek,
Nature 395, 783 (1998).
While Eq.7 is useful for certain metallic wire systems, it is less effective for molecular systems. There are three
reasons for this: first, it is difficult to define what one means by a transverse channel in the molecular situation.
Second, molecules do have vibronic coupling, and it is difficult to include that simply in the Landauer formulation.
Third, the electronic structure of molecules is not normally thought of in channels – to be consistent with the
advanced computational methods available for molecular entities, it is more convenient to formulate the problem
using a fundamental nonequilibrium approach, that can deal with interelectronic interactions, vibrational and
classical baths, photonic excitation, and so forth. This is offered by the nonequilibrium Green’s function (NEGF)
method [14-18,69,84].
The NEGF method was originally developed for problems of nonequilibrium statistical mechanics and for quantum
transport. It bears a close relationship to the two-sided Green’s function approach of Mukamel [10], but generalizes
the density matrix scheme, because there are two different times, with different time orderings, associated with
NEGF. Because of the time ordering, four different Green’s functions are generally used: the retarded, advanced,
greater, or lesser Green’s functions. Each has its own associated self-energy, and they differ simply because of the
time orderings. The equations of motion for these Green’s functions are typical many-body equations, in that no
closed- form solution exists except for very special cases. There are many approaches to solving the NEGF
equations, utilizing perturbation theory, or decoupling methods, or scattering approximations. It is then possible to
develop self-energy terms that describe interactions of the molecular system (as chosen) with the environment.
For the example of transport between two metallic electrodes, ignoring any corrections to the Born-Oppenheimer
approximation (no vibronic coupling) and no photonic coupling, the Caroli [85] equations 8 will hold. Here Gr is
the retarded Green’s
2e 2
(8a)
I (V ) T ( E ,V )[ f ( E ) f ( E )]dE
h
i
f
T ( E ,V ) Tr{i ( E ,V )G R ( E ,V ) f ( E ,V )G A ( E ,V )}
(8b)
G R ( E ,V ) [ E H ext TOT ]1
(8c)
= i,f
(8d)
i[ ]
72
Mark A. Ratner and Ronnie Kosloff / Procedia Chemistry 3 (2011) 63–81
function, and is the so-called spectral density. The understanding of these equations is straightforward: the
transmission T given by Eq. 8b is equivalent to the sum over channels in the Landauer expression. The spectral
density refers to the coupling between the “extended molecule” (the molecule plus a small number of atoms
representing part of the electrode; these are included to describe properly the coordination of the molecular entity to
the electrode) and the two electrodes.
The Green’s function G is given by eq.8c, in terms of the full self-energies. These self-energies contain real and
imaginary parts – the real part corresponds to the energy shifting of the system levels by the environment. The
imaginary part (in the simplest situation that the Caroli equations describe, again neglecting all interactions with any
bath except for the electronic one in the leads) describes the coupling between the electronic states of the extended
molecule and the electronic states in the leads. At a frequency corresponding to an excitation on the extended
molecule, the first two terms in the eq.8c will give zero, and the self-energy terms will cancel the spectral density
terms in eq.8b, so that effectively the transmission is g0, and the L/I limit holds.
The integration in eq.8a says that for any given voltage, the Fermi functions fi and ff are necessary for the correct
statistical counting: the electron must go from a full electrode level to an empty one, for the transition to be
allowed. Since these factors contain the Fermi energy EF , we need to be careful in choosing EF for the real
situation. This is usually taken to be EF of the bare metal, but this is almost certainly incorrect, because the metal in
a molecular transport junction will have molecules attached to it, and its geometric structure will certainly not be
that of the perfect crystal. These modifications should shift the relevant EF, and this issue, called the “band lineup
problem” makes calculations of the actual current or conductance in molecular transport junctions challenging
(though it can be overcome by actually calculating EF for the full system).
This formulation of molecular conductance is purely quantal; the quantum of conductance g0 contains Planck’s
constant. The problem of “quantized conductance” was effectively discussed by Landauer in the 1950s, and so this
quantum effect is neither new nor unanticipated. But for molecular systems, the area has only been discussed since
the 1990s, and very interesting progress has been made. We will limit ourselves here to two significant issues: first,
interference effects for multiple quantum channels; second, decoherence effects and the transition from tunneling to
hopping motion.
7.1. Interference Effects in Transport Junctions
Pi systems contain delocalized electrons, that can be described, qualitatively, by very simple tight-binding or Huckel
models. Since they are delocalized, we might expect interference between pathways. Indeed, this is the mechanism
by which we understand that (for example) in benzene rings the interaction between substituents at meta sites is very
much smaller than that at ortho or para sites. Other coherence-driven molecular effects include ring currents in
NMR [86], cross conjugation [87,88], and transfer through squaraine structures [89].
Recent analyses of current patterns through organic molecules strung between metallic electrodes, using either
extended Huckel theory or tight binding density functional calculations, have been completed by Solomon and coworkers [90-94]. These illustrate some aspects of molecular interference. The formalism to develop the tunneling
currents [95,96] follows directly from the Landauer transmission, and Eqs.8. It is then necessary to determine the
surfaces through which the charge will move, and to recall that currents can run both forward and backward. The
result is not only a calculation of the conductance, but an interpretive mechanism for understanding the conductance.
In these calculations, there is no direct treatment of decoherence, because the structures start with the Landauer
(elastic) formulation, and no dephasing-type terms are permitted, other than the spectral density itself (this is at the
frontier, and should be explored soon; some initial considerations using the Buttiker probe approach have appeared
[96]).
Mark A. Ratner and Ronnie Kosloff / Procedia Chemistry 3 (2011) 63–81
73
Fig.5. A simple cross coupling situation. Transport through the cross-coupled link (bottom) shows interference patterns absent in the linear link
(above)
Fig.6. A three-ring reversal in the current passing through a molecule. The arrow diameter is proportional to the current between leads (at lower
right and left). Blue arrows point against the voltage gradient, and are characteristic of interference behavior.
Figs.5, 6 illustrate the current flow pictures that one calculates, for different systems. Fig.5 illustrates the simplest
example of cross conjugation: cross conjugation is defined [97] as “three unsaturated groups, two of which although
conjugated to a third unsaturated center are not conjugated to each other”. The simplest example is a T junction
(stub resonator) situation. The transmission from one side to the other of cross conjugated versus normally
conjugated systems are shown: the deep minimum in the center of the curve arises because the pi system transport
is destructively interfered with by the cross conjugation, leading to the transmittance dip, with conduction being
dominated (in this energetic region) by the sigma system. Then the cross conjugated system conducts less well than
the fully-conjugated system, and indeed comparably to the sigma-only (presence of methylene groups in the
backbone) structure. Unpublished work (M. Wasielewski et al.) has investigated these effects using electron transfer,
rather than conductance, measurements. The results are in substantial agreement with these predictions, with the
cross-conjugated molecules showing anomalously low rates.
74
Mark A. Ratner and Ronnie Kosloff / Procedia Chemistry 3 (2011) 63–81
Earlier measurements also bear this out. Mayor and collaborators measured [98] conductance in two different
molecules, one characterized by para linkages and the other by meta linkages: the doubly meta-linked species
conduct essentially an order of magnitude less well than the para-linked ones, again because of interferences in the
pi system.
Fig.6 indicates the kinds of current branching situations that occur near interferences (interferences occur at given
values of the voltage, since the transmission is voltage dependent, and so are the currents and the conductance). The
three-ring structures correspond to currents flowing forwards and backwards (blue and red in the picture). They are
often found near interference minima in the pi system. In unpublished work, Solomon has also found that some of
the gating structures (in which photoisomerization is used to switch between a highly conductive and a less
conductive entity) can be described well in terms of these current flow pictures.
While these current calculations are done using a relatively primitive basis set, and do not deal with correlation or
vibronic coupling or decoherence problems in any way, nevertheless they are instructive. One purpose for doing
theory is to interpret and predict experiment, and these local current maps are helpful in that regard.
7.2. Decoherence and Turnovers
Quantum mechanics can describe systems that are generated when an electron donor (for electron transfer) or an
electrode (for electron transport) injects an electron onto a molecular bridge (consisting of one or a series of local
orbitals that can be occupied by the transferring electron) and then onto a different large entity, (an acceptor in
donor/bridge/acceptor electron transfer, or an electrode in voltage-dependent current spectroscopy).
Fig. 7. Schematic energy level picture for either a donor/bridge/acceptor molecule or a molecular transport junction (in the latter case, the red
sites are the highest occupied electrode band states).
When the gap between the levels on the bridge (Fig.7) and the donor/acceptor states is large, one expects a
“superexchange” kind of process, in which the rate of electron transfer (or the conductance) falls off exponentially
with distance. This can be rationalized in many ways, the simplest of which is to use one- electron perturbation
theory, with the smallness parameter being the ratio of the intersite tunneling integral over the donor-bridge energy
level difference.
Extensive modeling has demonstrated [99-103], and observations have shown [104-107], that as the bridge gets
longer, coherence can be destroyed by either T1-type or pure dephasing processes. In either situation, one eventually
reaches a case where the quantum behavior is no longer coherent over the whole bridge, and a transition (sometimes
gradual, sometimes abrupt) to hopping type incoherent transport occurs. This was predicted by a simple density
matrix treatment; the prediction, and some early experimental results, for DNA species are given in Fig.8.
Mark A. Ratner and Ronnie Kosloff / Procedia Chemistry 3 (2011) 63–81
75
Fig. 8. Transport turnovers seen in DNA oligomers.
These turnovers are fascinating, because they occur in many systems. We can rationalize these behaviors utilizing a
time scale argument. The tunneling matrix element can be compared with the polaron stabilization energy
(equivalent to the Marcus reorganization energy). When the tunneling matrix element is larger, the tunneling is
coherent, and then there isn’t time for the electron to trap itself locally.
Another time of interest is the so-called Landauer/Buttiker contact time, which is supposed to be the time that the
electron actually spends in the forbidden region between donor and acceptor (or between electrodes). In a
semiclassical approximation [108], this time
tcontact N / EG
(9)
is simply given by eq.9, where N is the length of the chain in repeating units, and EG is the gap energy – the energy
between donor (or acceptor) levels and bridge levels. As the gap becomes zero, the system becomes fully
delocalized. It can then transfer directly (the transmittance goes to one, and one would expect quantum conductance
as in Fig.4) or, since the time formally becomes very large, trapping might occur on the bridge.
The issue of bridge trapping by polarons has been examined extensively for conducting polymers, both theoretically
and experimentally. One interesting question issue is the time scale of charge transport versus the time scale of
polaron formation. The polaron coupling is proportional to the density of electrons on the bridge; if EG is large, the
populations are small, and one expects that the golden rule formula, with the perturbation proportional to the square
of the population, will indicate that the vibronic coupling is very weak. Whether or not the polaron formation time
is the same for electrons moving with kinetic energy and for electrons localized at one site seems unclear.
If one takes an empirical approach to the decoherence, simply using the representation with relaxation terms added
to the off-diagonal elements of the density matrix, one can actually show that para-, ortho-, and meta-type effects
completely disappear, when the dephasing gets strong enough. An example is shown for the Davidene molecule in
Fig.9 – this is not surprising, but it is an interesting prediction if these measurements are ever made. Similar
considerations have been suggested [96] for actual molecular devices, based on interference.
76
Mark A. Ratner and Ronnie Kosloff / Procedia Chemistry 3 (2011) 63–81
Fig.9. As the dephasing parameter g increases, the o, m and p linkages all agree, since their interference patterns have been lost. Modeling from
R. Goldsmith, M.A. Ratner and M.R. Wasielewski J. Phys. Chem. B, 20258–20262 (2006).
8. Theoretical Worries, Technical but Significant
Dealing with the quantum dynamics of molecular systems interacting with a condensed phase environment has
made great progress experimentally, theoretically, and computationally. Density matrix, QM/MM and
nonequilibrium Green’s function approaches are useful for formulating and understanding the quantum/classical
dynamics. Moreover, computational advances (particularly the use of Feynman path integral methods for model
problems [110,111], density functional theory and time-dependent density functional theory for electronic structure
and response, and extensive simulation methodologies, all coupled with substantial advances in the technology of
computation (including most recently GPU cards)) have allowed meaningful calculations to be done. Nevertheless,
several challenges need to be overcome before analysis of real chemical systems, with the electronic structure
details and the dynamics, can be completed. Here we simply note some of these, and suggest possible solutions.
8.1. Basis Sets, and Pointer Basis
When density matrices are used (as they broadly are) to describe quantum dynamical evolution and relaxation in
molecular systems, we need basis sets to represent the systems (interacting with the bath). Quantum chemistry
provides an extensive set of basis sets for use in static structures, but a question like how energy transfer occurs
between a phthalocyanine and a nearby quinone poses the problem of basis sets in a different way. In perturbation
theories, one needs to calculate the matrix elements of a perturbation between initial and final states -should they be
stationary states? Usually not, because one is generally experimentally interested in initial states that have been
photoexcited, excited by collision or ionization – that is, these are not equilibrium eigenstates. These states can be
generated computationally, by computing the time evolution under the exciting field from an initial ground state
[112]. This excited wave function should represent the initial state in the chemical dynamics.
If one is interested in the evolution of this initially-excited molecular state, it needs to be represented somehow in
terms of the basis. For model systems as simple as a two- or three-level model, this is not a problem. But in
applications to real systems, we need to represent this initial state such that its evolution can be computed. Again, as
long as we operate within a full basis (we study the complete problem quantum mechanically), the propagation is
straightforward although so time consuming that as of 2011it is impractical. But if the calculation is of density
matrix or NEGF type, to describe the effects of the bath interaction with the system, the problem becomes more
complex. Several problems with the basis set then follow. For a two-level system, in which the field enters through
Bloch-type relaxation of the diagonal and off-diagonal elements of the density matrix, the dynamics can be
completely different depending on the representation in which the T1 and T2 relaxations are applied. To pick an
extreme situation, for a two-site orthogonal system, one could choose either the local states or the fully delocalized
Mark A. Ratner and Ronnie Kosloff / Procedia Chemistry 3 (2011) 63–81
77
states as the basis functions for the density matrix. Limiting ourselves to pure dephasing, it is clear that using the
local basis can result in the system being trapped in either the left or the right localized state, when the dephasing
rate overwhelms the tunneling. But if the eigenstate basis is used, the system remains in the fully delocalized,
nodeless, “binding” structure, if (again) the dephasing is fast enough.
These are trivial examples of the so-called pointer state problem. The pointer basis is supposed to be the most
reasonable in which to carry out quantum dynamical relaxation calculations. The difficulty is agreeing on what the
pointer basis might be. There has been discussion of this in the literature [113,114], but it seems that this situation is
not yet resolved. The usual approach is to revert to a model, in which one assumes an initial state that is fully
localized, assumes a coupling Hamiltonian (often of simple Huckel type), and carries out the dynamics assuming
some sort of bath. This could be some form of the so-called spin-boson system [64] (a two level molecular system
coupled to a set of harmonic oscillators corresponding to the bath). Even that is complicated, though it completely
avoids the problem of the choice of basis sets.
8.2. Initial optical preparation
If a full molecular representation in terms of a reasonable electronic structure basis is used, it is possible [112] to
generate a photoexcited state by propagating the initial static ground state under the perturbation V as exp(. But the optical interaction with the molecule is proportional to the molecular dipole operator, and this
can distribute density throughout the molecule. That initially prepared state will generally not resemble our intuitive
notion of, say, an optically excited donor state in an electron transfer reaction, or an initial exciton state for energy
transfer. Intuitively, we expect that over a very short time (a few tens of femtoseconds) a sort of electronic
decoherence occurs, causing effective localization of the charge into something that closely resembles a localized
excited state. These modifications are not the same sorts of shifts that are measured at much longer times by the
time-dependent Stokes shift spectrum, because that is often dominated by relatively slow motions (for example
rotations) in the environment. Intuitively, the sorts of decoherence involved in making an initial photoexcited state
correspond to “chopping off the tails” of the photoexcited, largely delocalized, state to form a reasonably localized
photoexcited state. This has been addressed by the (interesting but still far from complete) idea of einselected
(environment - induced superselection) states, that are the appropriate initial states for the quantum dynamics. The
einselection process [113,114] arises from the system/bath coupling, and occurs very much faster than other
timescales in the system. The states formed by the einselection process may be pointer states, and in the simplest
view they are minimally influenced by the bath. Issues of generality and uniqueness still surround this notion, which
has been used only for a few selected model systems. Still, these are significant issues, particularly as the bath
changes from aqueous solution (where coherences last only very short times, usually less than 100 femtoseconds) to
slower-relaxing hosts such as less polar liquids, bio-environments, polymers, solids or glasses
One way around this problem, investigated in the context of energy and of electron transfer, is to choose a state that
is maximally localized, in the Edmiston-Ruedenberg sense. In an initial study, Subotnik and collaborators [115]
were able to match very well classic energy transfer studies, utilizing these localized initial electronic states.
8.3
Overlap issues
It is convenient to assume (almost always assumed in formal derivations), that the initial and final states are
orthogonal to one another. But conventional orbital basis sets, as used in electronic structure calculations, are hardly
ever orthogonal. They can be made orthogonal by Lowdin or Schmidt or other procedures, but this leaves
difficulties in understanding what happens at short times after photoexcitation.
Suppose we imagine a very fast photoexcitation of an electron in site A, where the localized orbitals A and B are
not orthogonal. The prepared state is then just c A 0 , with
population on B at this initial time:
c A the creation operator on A. Then we find for the
78
Mark A. Ratner and Ronnie Kosloff / Procedia Chemistry 3 (2011) 63–81
nB (0) 0 c AcB cB c A 0
(10)
2
,
S AB
this follows from the anticommutation relations between the creation and destruction operators, and SAB is the
overlap integral. This suggests that the expectation value of the population on site B, even at time zero, is the square
of this overlap integral. This raises the question of whether or not we can actually separate donor and acceptor states
using this sort of molecular basis. The interpretations can become quite difficult; it appears that when the electron is
created on site A, there is a small part of it that is on site B even at zero time. Perhaps the same sort of rapid
relaxation (einselection) referred to in Section 8.1 could occur here too, but formally this is a nettle.
8.4 Other molecular systems
Coherence became an important theme in modern condensed phase science with the development of magnetic
resonance. Issues of relaxation phenomena and loss of phase were brilliantly handled by the pioneers of magnetic
resonance spectroscopy. Magnetic resonance is the first situation in which coherence and its loss have been
significant issues. It is also a fortunate place for doing this kind of work, because the eigenvalue spectrum of the
spins is generally very sparse – for a proton, there are only two levels. This means that relatively simple
representations like the spin-boson one are appropriate, and it also means that there is enough time to do all sorts of
wonderful pulse sequencing and pulse shaping, both to measure and to drive the system.
The systems described in this manuscript (and indeed in this volume) are different. They involve coherences that are
electronic in origin, and coherence in energy transfer and charge transfer becomes a dominant issue. Because these
systems are not sparse, and because in many cases (in particular polar materials and polar environments) the
interactions with the environment are very strong, the existence of long time coherence is unusual, and
understanding it requires a new approach to this second situation for the decoherence problem, involving quantum
effects in chemistry.
There is a third area that has been explored only slightly. Vibrational coherences can also occur, and indeed one
might argue that they are partway between magnetic resonance and electronic behavior. Unlike magnetic resonance,
their eigenvalue spectrum is not sparse, and Fermi resonances and near-Fermi resonances are common. But unlike
electronic phenomena, their coupling to the bath is often quite weak. This allows coherences of quite high orders to
be observed, even in relatively large systems.
A wonderful example is found in the recent work of Wright [116,117], who examined, in particular, a rhodium
dicarbonyl chelate organometallic compound and was able to observe very high-level coherences, to understand
them both energetically and dynamically, and to suggest (in collaboration with Sibert) that they could be described
very well in terms of only two degrees of freedom. This vibrational coherence spectroscopy is another example of
coherent quantum effects in chemistry, and here the understanding and mastery of coherences is probably halfway
between its elegant state in spin systems and its challenging state in electronic systems. One task for the conceptual
mastery of dynamical quantum effects in chemistry is to transfer some of the learnings that occurred with spins and
vibrations to the challenging and significant area of electronic coherence behaviour.
8.5 Variational Principles
When the system is both open and nonequilibrium (for example, electron transport in junctions under voltage), a
particular problem occurs. Because the number of electrons is not known, and because the momentum of electrons
is finite since current is flowing, the standard derivation of the Rayleigh-Ritz variational principle fails. While there
are time- dependent variational principles that may well be valid, it is difficult to use these to optimize geometry.
The usual approach is to optimize the geometry in the absence of the field, and then to assume that the geometry
remains unchanged in the presence of the nonequilibrium potential and electronic momentum. This is generally not
correct, and the issue of optimized geometry under current flow conditions is one of the challenges remaining in the
field of molecular transport junctions.
Mark A. Ratner and Ronnie Kosloff / Procedia Chemistry 3 (2011) 63–81
79
These four issues represent some of the next chapter in our understanding of how to understand nonequilibrium
quantum dynamical systems. Very recent work by Martinez has focused on another issue: what coordinates should
be used to conceptualize the dynamics? Using GPU-enabled very fast computation, he was able to generalize the
description of a reaction coordinate, by permitting the overall computed dynamics (in his case QM/MM dynamics)
to be expressed in terms of the motions of the molecule itself and the surrounding solvent [118]. This idea of
allowing the computation to suggest its own coordinates is attractive, and may well provide a different approach to
understanding the molecular pathways, molecular interferences, and molecular decoherence that characterize
quantum dynamics of electronic energy and charge in condensed phase.
Acknowledgements
We thank Abe Nitzan, Gil Katz, Joe Subotnik, Troy van Voorhis, Stuart Rice, Josef Michl, George Schatz, Shaul
Mukamel, Greg Scholes and Graham Fleming for helpful discussions. This material is based upon work supported
as part of the NERC, an Energy Frontier Research Center funded by the U.S. Department of Energy, Office of
Science, Office of Basic Energy Sciences under Award Number DE-SC0001059.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
M. Born and J.R. Oppenheimer, Ann. Phys. (Leipzig) 84 (1927) 457.
W. Heitler and F. London, Z. Phys. 44 (1927) 455.
W. J. Hehre, L. Radom, P.V. Schleyer, J. Pople Ab-Initio Molecular Orbital Theory, Mc Graw-Hill, 1986.
Y. Shao, L.F. Molnar, Y. Jung, J. Kussmann et al., Phys. Chem. Chem. Phys. 8 (2006) 3172.
G.R.Fleming and M.A. Ratner, Eds., Directing Matter and Energy:
Five Challenges for Science and the Imagination, U. S. Department of Energy, 2008.
J.H. Worner, J.B. Bertrand, D.V. Kartashov, P.B. Corkum et al., Nature 466 (2010) 604.
J. von Neumann (translated by R.T. Beyer) Mathematical Foundations of Quantum Mechanics, Princeton, 1955.
E.R. Bittner, Quantum Dynamics : Applications in Biological and Materials Systems, CRC Press Boca Raton, 2010.
G.C. Schatz and M.A. Ratner, Quantum Mechanics in Chemistry, Prentice Hall, 1993.
K. Blum, Density Matrix Theory and Applications , Plenum, 1996; D. Abramavicius, S. Mukamel,
J. Chem. Phys. 133 (2010) 064510; S. Mukamel, Principles of Nonlinear Optical Spectroscopy Oxford, 1999.
H.-P. Breuer, F. Petruccione, The Theory of Open Quantum Systems, Oxford, 2007.
A. Osted, J. Kongsted, K.V. Mikkelsen, O. Christiansen, Chem. Phys. Lett. 429 (2006) 430.
T. Vreven, K. Morokuma, O. Farkas, H.B. Schlegel et al., J. Comp. Chem. 24 (2003) 760.
S. Datta, Quantum transport : Atom to Transistor, Cambridge, 2005.
L.P. Kadanoff, G. Baym, Quantum Statistical Mechanics; Green’s Function Methods in Equilibrium and Nonequilibrium Problems,
Benjamin, 1962.
H.J.W. Haug, A.-P. Jauho, Quantum Kinetics in Transport and Optics of Semiconductors, Springer, 2009.
G.D. Mahan, Many-Particle Physics, Kluwer/Plenum Publishers, 2000.
L.V. Keldysh, Soviet Physics JETP-USSR 20 (1965) 1018.
A. Nemeth, F. Milota, T. Mancal, V. Lukes et al., J. Chem. Phys. 132 (2010) 11.
N. Christensson, F. Milota, A. Nemeth, J. Sperling et al., J. Phys. Chem. B 113 (2009) 16409.
M. Bertin, M. Mever, J. Stahler, C. Gahl et al., Faraday Disc. 141 (2009) 293.
B. Bagchi, B. Jana, Chem. Soc. Rev. 39 (2010) 1936.
J. Lobaugh, P.J. Rossky , J. Phys. Chem. A 103 (1999) 9432.
G.D. Scholes, J. Phys. Chem. Lett. 1 (2010) 2.
E. Collini, C.Y. Wong, K.E. Wilk, P.M.G. Curmi et al., Nature 463 (2010) 644.
H. Hossein-Nejad, G.D. Scholes, New J. Phys. 12 (2010) 20.
D. Jarzab, F. Cordella, M. Lenes, F. B. Kooistra et al., J. Phys. Chem. B 113 (2009) 16513.
C.J. Brabec, J.R. Durrant, MRS Bulletin 33 (2008) 670.
A.J. Heeger, Chem. Soc. Rev. 39 (2010) 2354.
P.J. Flory, Statistical Mechanics of Chain Molecules, Interscience, 1969.
T. Förster, Ann. Physik 437, 55 (1948).
A. Huijser, B. Suijkerbuijk, R. Gebbink, T.J. Savenije et al., J. Am. Chem. Soc. 130 (2008) 2485.
L. Fradkin, R.E. Palacios, J.C. Bolinger, K.J. Lee et al., J. Phys. Chem. A 113 (2009) 4739.
G.S. Engel, T.R. Calhoun, E.L. Read, T.K. Ahn et al., Nature 446 (2007) 782.
A. Ishizaki, G.R. Fleming, Proc. Nat. Acad. Sci. 106 (2009) 17255.
D. Borgis, P.J. Rossky, L. Turi, J. Chem. Phys. 125 (2006) 13.
80
Mark A. Ratner and Ronnie Kosloff / Procedia Chemistry 3 (2011) 63–81
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]
[51]
[52]
[53]
[54]
[55]
[56]
[57]
[58]
[59]
[60]
[61]
[62]
[63]
[64]
[65]
[66]
[67]
[68]
[69]
[70]
[71]
[72]
[73]
[74]
[75]
[76]
[77]
[78]
[79]
[80]
[81]
[82]
[83]
[84]
[85]
[86]
[87]
[88]
[89]
[90]
[91]
[92]
[93]
[94]
[95]
[96]
[97]
A. E. Bragg, J.R.R. Verlet, A. Kammrath, O. Cheshnovsky et al., J. Am. Chem. Soc. 127 (2005) 15283.
L. Turi, G. Hantal, P.J. Rossky, D. Borgis, J. Chem. Phys. 131 (2009) 9.
I.J. Finkelstein, A. Goj, B.L. McClain, A.M. Massari et al., J. Phys. Chem. B 109 (2005) 16959.
S. S Skourtis, D. H Waldeck, D. N.Beratan, Fluctuations in Biological and Bioinspired Electron-Transfer Reactions. In Annual
Review of Physical Chemistry, Annual Reviews: Palo Alto, (201) 461.
S.Chakrabarti, M. Liu, D. H.Waldeck, A. M. Oliver et al., J. Am. Chem. Soc. 129 (2007) 3247.
S. S. Skourtis, D. N. Beratan, R. Naaman, A. Nitzan et al., Phys. Rev. Lett. 101 (2008) 238103.
S. Glasstone, K.J. Laidler and H. Eyring, The Theory of Rate Processes, McGraw-Hill, 1941.
L. Salem, The Molecular Orbital Theory of Conjugated Systems. Benjamin, 1966.
A.A. Kocherzhenko, F.C. Grozema, L.D.A. Siebbeles, J. Phys. Chem. C 114 (2010) 7973.
D. Walter, D. Neuhauser, R. Baer, Chem. Phys. 299 (2004) 139.
G.C. Solomon, C. Herrmann, T. Hansen, V. Mujica et al., Nat. Chem. 2 (2010) 223.
V. Ben-Moshe, A. Nitzan, S.S. Skourtis, D.N. Beratan, J. Phys. Chem. C 114 (2010) 8005.
S. Gerlich, M. Gring, H. Ulbricht, K. Hornberger, J. Tüxen et. al., Angew. Chem. Int. Ed. 47 (2008), 6195.
M.H. van der Veen, M.T. Rispens, H.T. Jonkman, J.C. Hummelen, Adv. Funct. Mater. 14 (2004) 215.
E.V. Anslyn, D.Daugherty, Modern Physical Organic Chemistry. University Science, (2006).
Z.W. Lin, C.M. Lawrence, D.Q. Xiao, V.V. Kireev et al., J. Am. Chem. Soc. 131 (2009) 18060.
D.Q. Xiao, S.S. Skourtis, I.V. Rubtsov, D.N. Beratan, Nano Lett. 9 (2009) 1818.
G. Lindblad, Comm. Math. Phys. 48 (1976) 119.
D. Kohen, C.C. Marston, D.J. Tannor, J. Chem. Phys. 107 (1997) 5236.
R. Alicki, K. Lendl, Quantum Dynamical Semigroups and Applications. Springer, 1987.
R. Kosloff, M.A. Ratner, W.B. Davis, J. Chem. Phys. 106 (1997) 7036.
D.A. Lidar, I. L. Chuang, K.B. Whaley, Phys. Rev. Lett. 81 (1998) 2594.
A.G. Redfield, IBM J. Res. Dev. 1 (1957) 19.
G. Katz, D. Gelman, M.A. Ratner, R. Kosloff, J. Chem. Phys. 129 (2008) 6.
G. Katz, M.A. Ratner, R. Kosloff, New J. Phys. 12 (2010) 13.
J. E. Subotnik, T. Hansen, M.A. Ratner, A. Nitzan, J. Chem. Phys. 130 (2009) 12.
G. Katz, R. Kosloff and M. A. Ratner, J. Phys. Chem. 114 (2010).
A.J. Leggett, S. Chakravarty, A.T. Dorsey, M.P.A. Fisher et al., Rev. Mod. Phys. 59 (1987) 1.
G. Cooper, J.A. Turner, B.A. Parkinson, A.J. Nozik, J. Appl. Phys. 54 (1983) 6463.
P. Rebentrost, M. Mohseni, A. Aspuru-Guzik, J. Phys. Chem. B 113 (2009) 9942.
J. Prior, A.W. Chin, S.F. Huelga, M.B. Plenio, Phys. Rev. Lett. 105 (2010) 4.
B. Muralidharan, A.W. Ghosh, S. Datta, Phys. Rev. B 73 (2006) 5.
M. Galperin, M.A. Ratner, A. Nitzan, J. Phys.-Cond. Matt. 19 (2007) 81.
W.Y. Wang, T.H. Lee, M.A. Reed, Proc. IEEE 93 (2005) 1815.
X. Chen, Y.M. Jeon, J.W. Jang, L. Qin, et al., J. Am. Chem. Soc. 130 (2008) 8166.
H.B. Akkerman, B. de Boer, J. Phys.-Cond. Matt. 20 (2008) 20.
G. Cuniberti, G. Fagas, K. Richter, Introducing Molecular Electronics Springer (2005).
Y. Selzer, D.L. Allara, Ann. Rev. Phys. Chem. 57 (2006) 593.
A.R. Champagne, A.N. Pasupathy, D.C. Ralph, Nano Lett. 5 (2005) 305.
E.A. Osorio, K. Moth-Poulsen, H.S.J. van der Zant, J. Paaske et al., Nano Lett. 10 (2010) 105.
O. Tal, M. Krieger, B. Leerink, J. M. van Ruitenbeek, Phys. Rev. Lett. 100 (2008) 4.
B.Q. Xu, N.J. Tao, Science 301 (2003) 1221; J. He, O. Sankey, M. Lee, N.J. Tao et al., Faraday Disc. 131 (2006) 145.
L. Venkataraman, J.E. Klare, I.W. Tam, et al. Nano Lett. 6 (2006) 458-62.
E. Lortscher, H.B. Weber, H. Riel, Phys. Rev. Lett. 98 (2007).
M. Kamenetska, M. Koentopp, A.C.Whalley, Y.S. Park, et al., Phys. Rev. Lett. 102 (2009) 4.
R. Landauer, IBM J. Res. Develop. 1, 233 (1957).
M. Buttiker, Y. Imry, R. Landauer, S. Pinhas, Phys. Rev. B 31 (1985) 6207.
J. Taylor, H. Guo, J. Wang, Phys. Rev. B 63 (2001) 245407.
C. Caroli, R. Combesco, P. Nozieres, D. Saintjam, J. Phys. Pt. C.- Sol. St. Phys. 4 (1971) 916.
J.A. Pople, K.G. Untch, J. Am. Chem. Soc. 88 (1966) 4811.
R.R. Tykwinski, F. Diederich, Liebigs Annalen-Recueil (1997) 649.
R. Stadler, Phys. Rev. B 80 (2009) 11.
C.T. Chen, S.R. Marder, L.T. Cheng, J. Am. Chem. Soc. 116 (1994) 3117.
G.C. Solomon, C. Herrmann, J. Vura-Weis, M.R. Wasielewski et al., J. Am. Chem. Soc. 132 (2010) 7887.
C. Herrmann, G.C. Solomon, M.A. Ratner, J. Am. Chem. Soc. 132 3682.
G.C. Solomon, C. Herrmann, T. Hansen, V. Mujica et al., Nat. Chem. 2 (2010) 223.
C. Herrmann, G. C. Solomon and M.A. Ratner J. Phys. Chem. C 114 (2010) 20813-20820
G. C. Solomon , J. Vura-Weis , C. Herrmann et al, J. Phys. Chem. B 114, 14735-14744 (2010) 14735-14744
M. Ernzerhof, M. Zhuang, P. Rocheleau, J. Chem. Phys. 123 (2005) 5.
T.N. Todorov, J. Phys.-Cond. Matt. 14 (2002) 3049.
N.F. Phelan, M. Orchin, J. Chem. Ed. 45 (1968) 633.
Mark A. Ratner and Ronnie Kosloff / Procedia Chemistry 3 (2011) 63–81
[98] M. Mayor, H.B. Weber, J. Reichert, M. Elbing et al., Ang. Chem.-Int. Ed. 42 (2003) 5834.
[99] E. Sim, J. Phys. Chem. B 108 (2004) 19093.
[100] H. Kim, E. Sim, J. Phys. Chem. B 112 (2008) 2557.
[101] G. Ashkenazi, R. Kosloff, M.A. Ratner, J. Am. Chem. Soc. 121 (1999) 3386.
[102] D. Segal, A. Nitzan, W.B. Davis, Wasielewski, M. R. et al., J. Phys. Chem. B 104 (2000) 3817.
[103] Y. A. Berlin, F.C. Grozema, L.D.A. Siebbeles, M.A. Ratner, J. Phys. Chem. C 112 (2008) 10988.
[104] T. Hines, I. Diez-Perez, J. Hihath, H.M. Liu et al., J. Am. Chem. Soc. 132 (2010) 11658.
[105] L. Luo, C.D. Frisbie, J. Am. Chem. Soc. 132 (2010) 8854.
[106] D.N. Beratan, S.S. Skourtis, I.A. Balabin, A. Balaeff et al., Accounts Chem. Res. 42 (2009) 1669.
[107] S.H. Choi, C. Risko, M.C.R. Delgado, B. Kim et al., J. Am. Chem. Soc. 132 (2010) 4358.
[108] A. Nitzan, J. Jortner, J. Wilkie, A.L. Burin et al., J. Phys. Chem. B 104 (2000) 5661.
[109] W.Y. Wang, T. Lee, M.A. Reed, Reports on Progress in Physics 68 (2005) 523.
[110] N. Makri, Ann. Rev. Phys. Chem. 50 (1999) 167.
[111] N. Makri, A. Nakayama, N.J. Wright, J. Theo. Comp. Chem. 3 (2004) 391.
[112] R. Kosloff, Ann. Rev. Phys. Chem. 45 (1994) 145.
[113] H. Ollivier, D. Poulin, W.H. Zurek, Phys. Rev. Lett. 93 (2004) 4.
[114] W.H. Zurek, Rev. Mod. Phys. 75 (2003) 715.
[115] J.E.Subotnik, T. Hansen, M.A. Ratner, A. Nitzan, J. Chem. Phys. 130 (2009) 12.
[116] A.V. Pakoulev, S.B. Block, L.A. Yurs, N.A. Mathew, K.M. Kornau, J.C. Wright, J. Phys. Chem. Lett. 1 (2010) 822.
[117] N.A. Mathew, L.A. Yurs, S.B. Block et al., J. Phys.Chem. A 113 (2009) 9261-65.
[118] I.S. Ufimtsev, T.J. Martinez, J. Chem. Theo. Comp. 5 (2009) 3138.
81