Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Perspective

Cite This: J. Am. Chem. Soc. 2019, 141, 708−722 pubs.acs.org/JACS

From Fundamental Theories to Quantum Coherences in Electron


Transfer
Shahnawaz Rafiq and Gregory D. Scholes*
Frick Chemistry Laboratory, Princeton University, Princeton, New Jersey 08544, United States
transport leads to a competition between the delocalization
ABSTRACT: Photoinduced electron transfer (ET) is a and localization, and that is where coherences seem to matter
cornerstone of energy transduction from light to (Figure 1). The resulting dynamics are complex and could
chemistry. The past decade has seen tremendous advances sometimes translate to robust and disorder-resistant trans-
in the possible role of quantum coherent effects in the port.13,14 In this regime, due to competition of the time scales,
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

light-initiated energy and ET processes in chemical, the system exhibits coherent dynamics at short times, and at
biological, and materials systems. The prevalence of later times incoherent transport dominates. From the coupling
Downloaded via NATL TAIWAN UNIV on September 28, 2023 at 18:05:33 (UTC).

such coherence effects holds a promise to increase the perspective, the three regimes are analogous to the valence
efficiency and robustness of transport even in the face of electron localization or delocalization in the Creutz−Taube
energetic or structural disorder. A primary motive of this ions.15,16 Exploration of electron transfer (ET) systems capable
Perspective is to work out how to think about “coherence” of retaining phase information for a practical time scale could
in ET reactions. We will discuss how the interplay of basic help to direct or control charge separation and recombination.
parameters governing ET reactionslike electronic Electron transfer is a ubiquitous process whose mechanism
coupling, interactions with the environment, and intra- and dynamics are relevant to applications like solar cells,
molecular high-frequency quantum vibrationsimpact molecular electronic devices, biosensing techniques, and
coherences. This includes revisiting the insights from various photoactivated processes.17−24 Its importance is
the seminal work on the theory of ET and time-resolved emphasized every time an organic chemist draws a reaction
measurements on coherent dynamics to explore the role mechanism. Chemistry is about coercing electrons to move.
of coherences in ET reactions. We conclude by suggesting The advent of ultrafast broadband electronic spectroscopy has
that in addition to optical spectroscopies, validating the provided new insights into the microscopic mechanisms of
functional role of coherences would require simultaneous such a fundamental process.22,25−28 In this Perspective, we will
mapping of correlated electron motion and atomically start with a quantum/semiclassical description of ET reactions,
resolved nuclear structure. to lay a groundwork for introducing coherences in photo-
induced ET reactions.

■ INTRODUCTION
Recent work has suggested that effects collectively called
Coherences will be discussed in chemical, biological, and
material systems with examples cited in each section. We will
explore what is meant by coherences in ET. Specifically, we
will ask how to assess adiabaticity and friction (coupling of a
coherencelong thought to be irrelevant for robust function
reaction coordinate to solvent and/or nuclear modes) in
owing to their fleeting presencemight be harnessed
experimental studies of ET reactions. The role of high-
efficiently even when electronic coupling is not such a
frequency vibrations in ET reactions, while not a new subject,
dominant participant.1−7 Coherence refers to phase relations
has been highlighted by recent work using coherent spectros-
among the constituents of the superposition of quantum states
copies. We will examine the mechanistic role of high-frequency
that carries extra correlations beyond those predicted by the
vibrations using wavepacket spectroscopies. We present a
classical probability theory.8,9 When such phase relationships
perspective for new experimental strategies which could impart
among quantitieslike quantum amplitudeare retained long
functional validation to the coherences beyond the skepticism
enough to have a mechanistic and functional relevance to the emerging from pure optical spectroscopies. This is a review of
underlying process, the dynamics are rendered coherent. In early steps toward working out how to design systems that uses
relation to practical systems, when electronic states are coherence to assist in ET and inhibit back-reactions.


localized, transport of energy or charge is incoherentit
hops randomly from site to site. Similarly, stochastic QUANTUM AND SEMICLASSICAL DESCRIPTION
incoherent transport is prevalent when the electronic states OF ELECTRON TRANSFER
are strongly delocalized over multiple sites to form extended
eigenstatesthat is, population relaxes incoherently owing to The pioneering work by Marcus assumes the well-known
minimal interactions with the surroundings. These two thermally activated form of transition-state theory, wherein
extremes are prevalent when the electronic coupling between nuclear motions of the reactant molecules and the solvent are
the sites is either much smaller or much larger than the in thermal equilibrium before the reaction.29−31 Equilibrium
reorganization energy of the surroundings. In these limits, fluctuations of the solvent dielectric polarization enables the
optimization of the transport becomes very challenging.10−12
In the middle of the two extremeslocalization and Received: August 22, 2018
delocalizationthe interplay of energy fluctuations and Published: November 9, 2018

© 2018 American Chemical Society 708 DOI: 10.1021/jacs.8b09059


J. Am. Chem. Soc. 2019, 141, 708−722
Journal of the American Chemical Society Perspective

Figure 1. Limiting cases of transport as a function of electronic coupling. (a) Weak electronic coupling: incoherent transport from one curve to
other. (b) Strong electronic coupling: incoherent relaxation among highly delocalized states. (c) Intermediate electronic coupling: competition
between the energetic disorder and the coupling could give rise to a unique transport mechanism at short times of the dynamics.

reactant and product energies to occasionally become quasi- The overall ET rate is
degenerate, so that an ET event can occur, leaving the product 0→n
in an energetically rich state. The classical Marcus theory of ET kET = ∑ kNA
kET = Kelνn exp((λ + ΔG°)2/4λkBT) generally works well for n (3)
reactions where transition probability of ET is close to unity k0→n
NA is the non-adiabatic rate constant for each vibronic
most adiabatic ET reactionswhere Kel is the electronic channel.
transmission coefficient and νn is the frequency of passage
(nuclear motion) through the transition state. However, to ΔG°0 → n = ΔG° + nhν (4)
explain ET reactions where transition probability is much less (V 0 → n)2 = V 2 |⟨0|n⟩|2 (5)
than unitymost non-adiabatic ET reactionsa quantum
mechanical description to the nuclear motion is required. The |⟨0|n⟩|2 = (Sn/n! ) exp( −S) (6)
transition probability in such cases was formulated in a Golden
Rule description, originally by Levich and Dogonadze.32 The S, being the electron-vibrational coupling strength, or
concept of the electron tunneling from donor to acceptor as Huang−Rhys factor, is defined as a ratio of the reorganization
well as the nuclear tunneling from reactant to product were energy and frequency of the high-frequency mode (λvib/
introduced. The nuclear tunneling is required particularly in (hνvib)), n is the vibrational quantum number, and λs and λvib
the inverted regime of ET to move from the energy minimum are the solvent and vibrational reorganization energies.
of the donor surface to the acceptor surface due to the avoided In the high-temperature limit, this semiclassical Marcus
crossing between the two surfaces. Hopfield was the first to equation (eqs 2 and 3) reduces to a simpler and more familiar

ij −(λ + ΔG°)2 yz
study how nuclear tunneling explained low-temperature equation:

expjjj zz
biological ET kinetics by using a semiclassical treatment.33
j zz
2πV 2
k {
The Golden Rule treatment was also applied by Jortner and kET =
co-workers,34,35 as well as Marcus and co-workers,36 who ℏ 4πλkBT 4λkBT (7)
showed the role of high-frequency quantum vibrations of the
product as energy accepting modes. The Golden Rule where λ = λs + λvib is the total reorganization energy. Equation
expression for the ET rate constant is 7 is more like a classical Marcus equation. The 1950s Marcus
theory assumed a nuclear frequency factor characteristic of the
2π 2 potential energy surface curvature. Whereas, the later theories
kET = V FCWD
ℏ (1) of solvent-limited rates assumed overdamped or diffusive
where V is the matrix element for the electronic coupling motion along the reaction coordinate, leading to different pre-
between the initial and final states and FCWD is the Franck− factors of the classical and semiclassical Marcus rate equations.
Condon weighted density of states. In a full quantum The dependence of the ET rate on the energy gap leads to
treatment of ET, the FCWD contains a summation over all three distinct regimes based on the ratio of standard free
the vibronic states of the reactant, the product, and the solvent energy change between the reactant and product species and
vibrational modes. Knowing that solvent stochastic fluctuations the solvent reorganization energy: normal, barrierless, and
occur mainly at low frequencies and can therefore be treated inverted regimes.40 The classical and semiclassical Marcus
classically, the quantum treatment could only be applied to the equations both predict a symmetric quadratic logarithmic rate
high-frequency intramolecular vibrational modes. constant as a function of free energy gap on either side of the
An assumption was further made to reduce the number of barrierless regime because ET occurs only at the intersection
high-frequency vibrations to only one averaged mode of point of the two single free energy curves. However, in the
frequency ν by Jortner,37 Miller et al.,38 and Brunschwig and quantum description (eq 2), a more effective route is nuclear
Sutin,39 which resulted in the following semiclassical Marcus tunneling through vibrational overlaps, which can open extra
channels for ET. This condition is particularly important in the

ij −(λs + ΔG°0 → n)2 yz


equation for a non-adiabatic limit:

expjjj zz
inverted region, where vibrational wave functions of the

j zz
2π (V 0 → n)2 reactant and the product states are embedded, so Franck−

k {
0→n
kNA = Condon factors are much larger than the normal region. Siders
ℏ 4πλskBT 4λskBT (2) and Marcus36 confirmed that the inclusion of nuclear tunneling
709 DOI: 10.1021/jacs.8b09059
J. Am. Chem. Soc. 2019, 141, 708−722
Journal of the American Chemical Society Perspective

Figure 2. Semiclassical electron transfer models: (a) Sumi−Marcus model, (b) Jortner−Bixon model, and (c) Walker−Barbara model. These
models were built on the Fermi’s Golden Rule description of ET to accommodate the roles of solvation dynamics, adiabaticity, and high-frequency
quantum vibrations in the inverted region toward ET.

predicts an asymmetric fall off in the logarithmic rate constant the non-adiabatic and the adiabatic limits through a parameter
in the inverted region. for adiabaticity.56
The existence of the inverted region and the asymmetric kNA
shape of the ln(kET) vs ΔG° curve were experimentally verified kET =
by Miller et al. in intermolecular charge transfer (CT) between 1 + HA (8)
the biphenyl radical anion and various acceptors in a rigid low- HA is the adiabaticity parameter to entertain the transition
temperature glass as well as in biphenyl radical anion from the non-adiabatic to the solvent-controlled adiabatic ET
connected to various acceptors by a hydrocarbon bridge.38,40 and implicates the delocalization of the vibronic wave function
Later, researchers in the groups of Farid41 and Gray42,43 in the intersection region, given by
verified a Marcus inverted region with a symmetric ln(kET) vs
ΔG° curve, which was an indicative of nuclear tunneling 4πV 2τs
HA =
through vibrational overlaps of low-frequency modes. More ℏλs (9)
recent advances in photoinduced ET are highlighted in the
work by Vauthey and co-workers.22,44−47 where τs is the longitudinal dielectric relaxation time in the
The semiclassical treatment works well in the weak-coupling reaction coordinate. Equation 8 represents a continuous
non-adiabatic limit, where the ET process is treated within the change from the non-adiabatic limit (HA ≪ 1) to the adiabatic
equilibrium transition-state framework. However, in the limit (HA ≫ 1). In the solvent-controlled adiabatic limit, the
strong-coupling limit, the ET reaction proceeds adiabatically rate constant can be recast into the Kramer’s-like expression

ij −(λs + ΔG°)2 yz
on the lower energy surface, so that the probability of ET does for adiabatic ET rate:

expjjj zz
zz
j
not directly depend on the electronic coupling, but rather on 1 λs
k {
the probability of the reaction coordinate to reach the barrier kA =
crossing. Therefore, instead of equilibrium contributions, the τs 16πkBT 4 λ s k BT (10)
response of the solvent polarization contributes to the reaction The above approaches connect the non-adiabatic limit,
coordinate, and ET dynamics is discussed in the framework of where the probability of ET is determined by the electronic
Kramer’s theory.48−51 Kramer’s theory treats a reaction as a tunneling and tunneling fluctuations, and the adiabatic limit,
potential barrier crossing by stochastic motion on the potential where it is determined by the reaction coordinate dynamics. In
energy surface, where in the reaction rate is inversely other words, the two limits differ in the nature of nuclear
proportional to the dielectric relaxation time. The reaction motion through the curve crossing region, which is considered
can thus be identified as a solvent-controlled adiabatic to be overdamped for non-adiabatic and ballistic or diffusive
reaction.51 for adiabatic transfer.
The transition from non-adiabatic ET to solvent-controlled In the adiabatic case of strong coupling, an electron is
ET, was originally studied by Zusman using stochastic motion transferred ballistically by nuclear motion through the
along the reaction coordinate with transition probability transition state, while in the solvent-controlled adiabatic
proportional to V2.52 Friedman and Newton53 and Sumi and limit, the electronic coupling may still be weak; however, the
Marcus54 used the first-passage approach, while Rips and rate is affected by frictional couplingcoupling of the nuclear
Jortner adopted the real-time path integral formalism.55 Garg, reaction coordinate to other nuclear coordinates and the
Onuchic, and Ambegaokar introduced a block-summation solvent relaxation coordinate. Frictional coupling ensures that
technique to derive an expression for the rate, which bridges the time spent in the curve crossing region is long enough that
710 DOI: 10.1021/jacs.8b09059
J. Am. Chem. Soc. 2019, 141, 708−722
Journal of the American Chemical Society Perspective

every time an electron approaches the crossing region, the and coumarins.59 The Sumi−Marcus model predicts a slow
donor−acceptor encounter is effective, even though electronic rate of ET reaction in these systemsall these ET reactions
coupling is weak. Hence, the reaction appears adiabatic owing fall in Marcus inverted regime. The Bixon−Jortner treatment
to the nuclear control rather than electronic tunneling. reproduces these experimental ET rates in fast relaxing solvents
In the solvent-controlled adiabatic limit or in the adiabatic only, but overestimates the rate in slow-relaxing solvents,
limit of strong coupling, the nuclear motions and the solvent where it does not scale with solvation dynamics. Walker et al.
relaxation rates thus decide the ET probability. Therefore, discussed how the Jortner−Bixon treatment was unable to
many theories were put forward to address the role of reproduce the switch from the solvation dynamical control to
vibrational motion and solvation dynamics in ET. the vibrational control.57,61,64,65
Sumi and Marcus, in 1986, used a low-frequency (classical) Walker and Barbara, in 1992, thus came up with a hybrid
vibrational degree of freedom and a classical solvent degree of model by extending the Sumi−Marcus model to incorporate
freedom to address the role of vibrational motion and solvation the effect of high-frequency vibrations in a manner similar to
dynamics in ET (Figure 2a).54 In this two-dimensional model, that developed in the Jortner−Bixon model (Figure 2c).57 The
the ET occurs along the classical vibrational coordinate (q), solvent dependent rate constant was expressed as
and the free energy barrier in the vibrational direction is 0→n
circumvented by energy fluctuations via reorganization along k(X ) = ∑ kNA (X )
the solvent coordinate (X). This model, however, predicts a n (13)

ij −(λi + ΔG°0 → n(X ))2 yz


expjjj zz
vanishing ET rate constant as temperature approaches zero

j zz
and also propounds that the barrier must be crossed for ET to 4π 2(V 0 → n)2
k {
0→n
occur, implying that there is no scope for possible quantum kNA (X ) =
h 4πλikBT 4λ i k BT
tunneling through the barrier.57
Jortner and Bixon, in 1988, while retaining the classical (14)
solvent degree of freedom to account for dielectric relaxation,
ΔG°0 → n(X ) = ΔG° + λs − 2λsX + nhv (15)
incorporated high-frequency quantum vibrations of the donor
and acceptor centers as the intramolecular vibrational The classical reorganization was split into two parts: one
coordinate coupled to the ET process (Figure 2b).58 High- corresponding to the classical vibration (λi) and second to the
frequency vibrations were shown to influence ET dynamics in solvent reorganization (λs). ΔG°0→n(X) is the solvent-
three ways: providing a multitude of parallel kinetic channels. coordinate dependent effective free energy gap between the
A “polaron dressing” effect (dressing electronic coupling by the ground state of the reactant and the nth vibrational level of the
nuclear Franck−Condon factor) reduces both the non- product state. The issues of the time scale for solvent
adiabatic rate constant (eq 1) and the adiabaticity parameter polarization that contributes to the ET reorganization energy
(eq 8). Vibrational excitation reduces the effective energy gap, have been much discussed by Matyushov and others.66−69
lowering the activation barrier in the inverted region. The The ET models described above established the role of
overall ET rate is a summation over all the high-frequency vibrations in ET reactions, which can be understood more
vibronic channels, comprehensibly in the context of the proton-coupled electron
∞ 0→n transfer (PCET) reactions. In PCET reactions, while the ET
kNA
kET = ∑ can be treated as an event occurring between a localized initial
n=0
1 + HAn (11) and a localized final state, the proton transfer occurs between
manifold of localized proton states on the initial state and a
where k0→nNA is given by eq 6, and the Franck−Condon corresponding manifold of vibrational states on the final
dressed adiabaticity parameter is HnA = HA|⟨0|n⟩|2. The solvent- state.70 The initial state of the proton may not be thermally
controlled adiabatic limit can be realized for some of the n = 0 accessible (as the proton potential well may have a frequency
→ n channels; i.e., the polaron dressed adiabaticity parameter of ∼2000 cm−1), but with a simultaneous ET event, the solvent
should be greater than unity for some values of n, which are reorganization can provide near-degenerate multiple proton
denoted by {n̅}. The ET rate is written as accepting levels on the final statenested energy surfaces. For
∞ 0→n ∞ 0→n more details, Hammes-Schiffer group has developed a general
kNA kNA theory for PCET reactions.71−77
kET = ∑ + ∑
n ∈{n ̅ }
HAn n ≠{n ̅ }
1 + HAn (12) Now it is established that ET reactions in the intermediate
or strong coupling limits depend significantly on the vibronic
This model thus introduces nuclear tunneling, and the interactions in addition to the solvent bath polarization.
extent of tunneling is decided by the vibrational mixing Therefore, frictional coupling as explained above, plays an
between the reactant and product states. Only those essential role in ET reactions. When friction is strong
intramolecular modes which involve large displacements compared to other competing energy-scales, the reactant
from the reactant to the product potentials need to be executes a random walk along the reaction coordinate to the
considered. barrier crossing (non-adiabatic ET).56 This is the incoherent
Photoinduced ET reactions much faster than solvation limit. A contrary situation where there is no frictiona
dynamics have been observed experimentally.59−61 Walker et complete separation of the nuclear reaction coordinate from
al. studied ultrafast photoinduced back-electron transfer in other nuclear and/or solvent degreesshould manifest as a
betaines.57 Tominaga et al. found intervalence CT in mixed- coherent transfer (ballistic) with fastest rates of ET. However,
valence compounds faster than the solvation dynamics.62 the findings of Garg et al. showed that fastest rate does not
Kobayashi et al. studied ultrafast intermolecular ET in the come from a frictionless limit.56 Instead, the rate increases as
electron-donating solvents as fast as 100 fs though fluorescence fiction is introduced. This is resonant with Wolynes’s work,
quenching.63 Similar observations were found with oxazines which suggested that quantum mechanical effects, for example
711 DOI: 10.1021/jacs.8b09059
J. Am. Chem. Soc. 2019, 141, 708−722
Journal of the American Chemical Society Perspective

Figure 3. Wavepacket dynamics and its spectroscopy. (a) Generation of a coherent wavepacket on electronic potentials by a broadband laser pulse.
(b) Motion of the wavepacket in the phase space of the electronic potential modulates the signal amplitude of Cresyl Violet (in methanol), and that
modulation appears as ripples in the time-domain spectroscopic signals. (c) Fourier transform of the Cresyl Violet residual ripples produces
frequency of the wavepacket generated along Franck−Condon active nuclear vibrations. Panels b and c are reprinted with permission from ref 103.
Copyright 2016 American Chemical Society.

curve crossing, conspire with dissipation in an interesting way. environment, then quantum coherences should always be
He argued that friction generally increases the adiabaticity of a subjected to strong dephasing.
barrier crossing because of the increased time spent by the Some recent research suggests that quantum coherence
electron in the crossing region and multiple chances to move could be facilitated by resonance with vibrational coher-
back and forth.78 In the regime where friction and adiabaticity encesidentified as vibronic coherences.6,9,88−91 Reimers and
corroborate, the dynamics becomes complex, and coherent Hush92,93 have studied ET involving a single vibrational mode,
interactions enabled by the reaction coordinate may matter. providing insights into vibronic coupling. In the adiabatic limit,


the reactant and product are coupled together to make Born−
COHERENCE IN PHOTOINDUCED ELECTRON Oppenheimer adiabatic ground and excited states. Theoretical
TRANSFER treatments by Reimers and Hush suggest that on the reactant
and product sides of the reaction coordinate, the electrons are
The Non-adiabatic Limit. In the non-adiabatic limit of
localized and vibrations are localized, but in the vicinity of the
Marcus’s classical theory,29 the jump of electron from the
transition state the electronic wave function and vibrations are
reactant free energy curve to the product curve follows a
delocalized.94,95 Recent work has shown that ET in the
Franck−Condon principle in the same way as it is obeyed in
photoexcitation. In the normal regime of ET, this instanta- inverted regime in a photosynthetic reaction center (photo-
neous transfer of the electron requires the reactant and product system II) is expedited when vibrational levels bridge
energies to be degeneratedriven by solvent energy reactant−product electronic gaps.96−99 This work highlights
fluctuations. In the inverted regime, that instantaneous transfer the importance of elucidating how molecular vibrations aid ET
is achieved by invoking nuclear tunneling to accomplish the reactions. That issue is salient because these vibrational
energy matching, and the phase coherence is equivalently frequencies and the nature of those modes can be tuned by
maintained during the ET reaction. However, it is a very chemical structurethus giving the chemist control over this
simplistic aspect of coherent transfer, as the coherence is apparently powerful and exotic design ingredient.
maintained only for a few femtosecondsthe time scale Vibrational Coherences. The interest in the role of
during which the actual ET occurs. In this non-adiabatic limit, coherences in ET reactions was triggered by the short-pulse
the solvent coordinate and the nuclear coordinate are pump−probe experiments of Vos and co-workers in the early
completely orthogonal to each other, and the reaction is 1990s, in the bacterial photosynthetic reaction centers.100−102
promoted by the solvent dielectric relaxation. An implicit manifestation of using short pump pulses is the
Elegant examples of such quantum coherence can be found generation of superpositions of vibrational wave functions
abundantly at the molecule scale, when the relevant energy wavepackets comprising Franck−Condon active modes
scales are very large compared to thermal energy. For example, (Figure 3).103 These wavepackets modulate the signal
in superexchange-mediated ET in donor−bridge−acceptor amplitude of the electronic surfaces. The modulations when
complexes, the different electronic coupling pathways across probed by another laser pulse, appear as oscillations on top of a
the bridge interfere quantum mechanically at the amplitude smooth population decay in the time-domain signal.
level, leading to an utterly different outcome than predicted by Oscillations are Fourier transformed to show the vibrational
the classical probability laws.79−86 frequency of the Franck−Condon active modes accessed by
The Adiabatic Limit. Quantum coherence involving the bandwidth of the pump pulse (Figure 3). In the
electronic coupling trajectoriesas described in the non- photosynthetic reaction center, the vibrational coherences
adiabatic limitcannot be considered as a powerful tuning along the low frequency nuclear modes surprisingly survived
element, even though it is omnipresent, because it is strongly the primary ET event.100−102 This work implicated a possible
subjected to dephasing.87 One may imagine that since ET is functional role of vibrational coherences for ET. An outcome
strongly slaved to stochastic polarization fluctuations of the of that period was a better understanding of vibrational
712 DOI: 10.1021/jacs.8b09059
J. Am. Chem. Soc. 2019, 141, 708−722
Journal of the American Chemical Society Perspective

coherences in ET, succinctly pre-empted by Jortner and Bixon have a set of values which define the contribution of each
(discussed above).58 vibrational level to the superposition (i.e., wavepacket).
The Bixon−Jortner description of the role of vibrations in Multiple crossings due to the nesting of the free energy
ET formed the basis of interpreting vibrational coherences in surfaces will modify the value of the coefficients of each
ET.104 More direct questions pertain to the role of vibrations: vibrational level as the reactant passes into the product, leading
Are there particular vibrations that promote or attenuate the to dispersion of the wavepacket (Figure 4b). Thus, the
ET? Is the bath polarization coordinate orthogonal to the dephasing time of the reactive mode wavepacket should
vibrational coordinate, or is there a collective additive effect of correspond to the mean first passage time to the transition
bath polarization and intramolecular vibrations that affects the state, which is defined as the time needed to reach the
ET reaction? To answer these questions, many ET systems transition state or the mixed region by a statistical ensemble of
have been studied using ultrafast coherent pump−probe-type initial states. Mean first passage time is expected to be shorter
spectroscopies in parallel with theoretical treatments to than the actual ET time, as the overall ET time includes the
speculate the role of vibrational coherences toward the ET time during which the transition state evolves into the
reactions. equilibrated final state.
Coherent vibrational wavepackets generated on the reactant A Note on Basis. We wish to point out that understanding
state are uniquely suited to report on the coupling of nuclear the experimental signatures of coherence demands a general
motions to the ET reactions, especially outside the Franck− understanding of basis, because without knowing the basis of
Condon region. On one hand, vibrational coherence that is the initial state, it is challenging to identify coherence.
weakly coupled to the reaction coordinate does not affect the Experiments such as two-dimensional electronic spectroscopy
reaction and will have no bearing on the ET dynamicshence (2DES) detect coherence in the adiabatic basis in the Franck−
superpositions (wavepackets) involving such modes will not be Condon region. For example, in CdSe nanoplatelets, where the
affected by ET (Figure 4a). The free energy curves in such basis is clearly delocalized, the oscillating off-diagonal cross-
peaks were assigned to an excitonic coherence (see Figure 7d−
f, below).87 That oscillating cross-peak indicates that the two
excitonic states are coupled and share excitation. The
modulation of the electronic wavepacket between the two
excitonic states lives for the time scale of the optical coherence
between the ground and the excitonic states. So, experiments
measure coherence in the adiabatic basis, because it probes
coupling between the electronic states through wave function
delocalization. In other words, oscillations in the cross peak of
2D maps can definitively assess state coherences. However,
when the dephasing is strong/rapid, the experiment will detect
localized states, more like the theorist exploring a localized
basis.
In the case of ET reactions, the reaction coordinate follows a
basis rotation, as epitomized by the method Larrson used to
calculate ET matrix elements.105 Calculations find that the
electron is localized on the donor in the Franck−Condon
region, but an external electric field can be changed, mimicking
Figure 4. Free energy curves and intramolecular modes. The effect of solvent reorganization, until the calculation predicts a perfectly
coupling a vibrational mode to the ET on the wavepacket dynamics. delocalized state where the electron is shared between the
Due to nested free energy crossings enabled by coupling a nuclear donor and acceptor. Similarly, when we think about most ET
mode to the ET reaction, the wavepacket dispersion is enhanced, and reactions, an initial state is prepared in a localized reactant
that translates into faster dephasing of the reactive mode wavepacket. basis, which evolves into the transition state that is a quantum-
This dephasing time could correspond to the mean first passage time
to the transition state of the ET reaction.
mechanical superposition of the reactant and product statesa
state best described in the delocalized basis. The transition
state finally collapses into the product localized basis. In such a
cases cross like pure diabats. On the other hand, ET aided by a situation, 2DES cannot assess coherence definitively at the off-
nuclear mode modifies that picture. The coupled vibrational diagonal cross peaks because basis changes during ET.
mode “nests” the free energy curves much like the way proton However, if the reactant and product states are coupled in
vibrations nest the energy curves of a proton-coupled ET the Franck−Condon region, then 2DES can provide definitive
reaction. Thereby, the vibration increases the number of ET insights in the coherent ET dynamics.
channels (Figure 4b) as well as the range of the activation Coherences in Chemical Systems. The first and foremost
barriers and effective reorganization energies by the constraint to experimentally exploring the role of coherences in
permutation of offsets (Bixon−Jortner theory). This peculiar ET is to focus on systems with ET rate is faster than vibrational
case is common to non-adiabatic and adiabatic ET reactions, dephasing. Barbara and co-workers used ca. 20 fs time-
though the picture gets more complicated in the adiabatic resolution pump−probe spectroscopy to measure an ca. 85 fs
regime. ET time-constant corresponding to the photoinduced back-ET
The nesting of the free energy curves dephases the reaction in the mixed-valence metal dimers.106−108 Analysis of
vibrational wavepackets during the ET reaction. To elaborate, the oscillations demonstrated that the average vibrational
when a superposition is generated along a coupled vibration on dephasing for the observed modes is ca. 300 fs. The longer
the reactant state, coefficients of individual vibrational levels dephasing time of the vibrational coherences relative to the ET
713 DOI: 10.1021/jacs.8b09059
J. Am. Chem. Soc. 2019, 141, 708−722
Journal of the American Chemical Society Perspective

Figure 5. Wavepacket dynamics during electron transfer. (a) Wavepacket branching at the branching space of the conical intersection along the
tuning and the coupling modes that define the CI between CT states and neutral ground state in the π-stacked dimer. (b) Integrated Fourier
transform of Betaine-30. Multiple frequency components are separated by using super-Gaussian windows of different widths followed by inverse
Fourier transformation of the frequency domain signal into the time domain to calculate dephasing times. Point to note is that certain high-
frequency modes dephase much faster in acetonitrile solvent than methanol, which is in correspondence with the solvent-dependent ET rate. (c)
Reactant−product coupling enabled by displaced high-frequency quantum vibration. The effective coupling is shown to be reduced by dressing of
Franck−Condon factors to the electronic coupling. Panel a is adapted with permission from ref 116. Copyright 2016 American Chemical Society.
Panels b and c are reprinted with permission from ref 121. Copyright 2017 Elsevier.

time (85 fs) indicated that coherence is maintained for some A new question arises then: Does a coherent excitation of a
degrees of freedom during the back-ET reaction. Scherer and wavepacket modify the ET dynamics? A coherent excitation
co-workers demonstrated that the wavepacket motion in instantaneously prepares a non-Boltzmann distribution of an
Prussian blue intervalence CT complex originated from the initial reactant state as superpositions of vibrational states of
ground state and the excited CT state.109 Further, it was the Franck−Condon active nuclear modes. In the case of weak
proposed that several low-frequency vibrational modes are electronic coupling, due to the separation of time scales
coupled to the CT coordinate.109 The longer lifetime of the between the electronic and nuclear motion, ET occurs only
vibrational coherences than ET could implicate that wave- after the relaxation of nuclei to the Boltzmann distribution.103
packet survives the ET event and is transferred to the product Therefore, the temporal coherences created by wavepacket
surface. motion dephase in the reactant state. On the other hand, in the
Wynne et al. in 1996 used a 40 fs, 810 nm pump pulse to intermediate regime of coupling, an ET reaction may be faster
create vibrational coherences in the CT state of an electron than the nuclear relaxation that dephases the wavepacket
donor−acceptor π-stacked complex between tetracyanoethy- (often similar to the time scale of intramolecular vibrational
lene (TCNE) and pyrene.110 Clear oscillations were observed relaxation); thus, an ET reaction occurs from “hot” vibrational
in the ground-state bleach and the stimulated emission signals states. This will be interesting to study, but the challenge is
during the back-ET process in 250 fs−1.5 ps. It was found that how to analyze the effect of a non-Boltzmann distribution of
the oscillations due to 160 cm−1 mode in the stimulated initially prepared states on the ET dynamics.
emission signal survived for the time scale of the ET and Mathies and co-workers studied a π-stacked CT complex
modulated the rate of the back-ET from the CT to the ground tetramethylbenzene−tetracyanoquinodimethane, using two-
state. This vibrational wavepacket was implicated to result in dimensional Raman spectroscopy and proposed that intra-
coherent formation of the product state via stepwise molecular CT is equivalent to a conical intersection between
modulations of the ground-state signal.110,111 However, excited and ground states.115,116 The branching space of the
identifying a temporal modulation on the basis of visual conical intersection is defined by the anharmonically coupled
distinction between vibrationally modulated recovery of totally symmetric 323 cm−1 CCN bend (tuning mode) and
population and sinusoidal wavepacket oscillations on the non-totally symmetric 1271 cm−1 CC rocking mode
ground state is almost impossible, unless we are dealing with a (coupling mode) (Figure 5a). The coherences produced
diatomic moleculefor example, the familiar dissociation of along the tuning and coupling modes branch at the conical
sodium iodide studied by Zewail and co-workers.112−114 intersection, and a portion of it crosses over to the ground state
Bixon and Jortner in 1997 modeled the temporal vibrational during the back-ET process.117−119
coherence effects in non-adiabatic ET dynamics in the π- Betaine-30 is another example, in which photoexcitation
stacked TCNE-pyrene ET complex.104 The calculations accesses a CT state, followed by a solvent-dependent ultrafast
predicted that for a nonradiative decay, the amplitude of back-ET1.2 ps in acetonitrile and 3.1 ps in methanol.120 The
coherences is determined by the off-diagonal matrix element possible role of coherent vibrational motion was explored by
(vibronic or electronic coupling between the states) and the Scholes and co-workers using a ca. 12 fs broadband pump−
reactant−product correlation (a function of FC vibrational probe spectroscopy.121 It was found that dephasing times of
overlaps between the reactant and the product states). It was certain high-frequency vibrations (1356, 1581, 1602 cm−1)
proposed that in the TCNE−pyrene ET complex, the most correlate with the solvent-dependent ET rate, which suggests
pronounced modulations originate in the vibrational wave- they are pertinent for the ET (Figure 5b). For example, the
packet motion of the initial state and are not due to the dephasing time constants of 1581 and 1602 cm−1 modes are
vibrationally coherent product formation. 0.27 and 0.29 ps in acetonitrile, where the rate of ET is 1.2 ps,
714 DOI: 10.1021/jacs.8b09059
J. Am. Chem. Soc. 2019, 141, 708−722
Journal of the American Chemical Society Perspective

and the corresponding dephasing time constants in methanol presence of coherent oscillations in the product state indicated
are 0.57 and 0.84 ps, where the rate of ET is 3.1 ps. Silva and wavepacket transfer during the ET reaction. It was proposed
co-workers reported that these high-frequency modes show that vibrational motion along the 600 cm−1 mode plays an
frequency shifts and are anharmonically coupled to the low- active role in the ultrafast ET from DMA to oxazine. Scholes
frequency modesprompting them to suggest that these and co-workers revisited the same ET systems and concluded
modes are relevant to the back-ET.122 Back-ET in Betaine-30 that the wavepacket motion on the product state is not an
represents a radiationless crossing of the excited-state and the outcome of coherence transfer.126 It is merely a spectator that
ground-state potentials in the adiabatic limit.121 As per Bixon− passes from the reactant to the product state without affecting
Jortner theory, to bridge the energy gap between the two the ET reaction. This is consistent with the Born−
potentials, high-frequency quantum vibrations couple to the Oppenheimer separation of electronic and nuclear degrees of
ET process. In Betaine-30, the free energy change for the freedom. The ultrafast rate of the ET may indeed derive from
product formation in acetonitrile is ca. 11 000 cm−1, so for the the coupling of high-frequency quantum vibrations to the ET
high-frequency mode (say 1600 cm−1) to participate in the reaction as represented in the Bixon−Jortner theory.104
back-ET process, it must provide a fourth or fifth vibrational The examples cited above highlight the use of vibrational
quantum on the product surface (Figure 5c) to make the wavepacket dynamics as internal clocks in probing the role of
process barrierless (considering a total reorganization energy of high-frequency quantum vibrations in the ET reactions. To
ca. 3500 cm−1 for acetonitrile).121 Note that this large energy summarize, the main challenge is to identify the reaction-
gap relative to the reorganization energy in Betaine-30 makes promoting vibrations from a sea of spectator vibrations. The
this a case of Marcus inverted region. reaction promoting vibrations are part of the ET reaction
A vibrational quantum (fourth or fifth) that is high in the coordinate, while the spectator modes, as the name suggests,
potential will have a very small Franck−Condon factor do not participate in the ET reaction. Spectator modes are
associated with it. Thus, the enhancement in the rate due to orthogonal to the reaction coordinate, so their corresponding
the vibronic interactions is an outcome of near resonance wavepackets dephase on a longer time scales than wavepackets
condition between the reactant and product states but not due of the reactive modes. In other words, a wavepacket
to the electronic coupling (ca. 2500 cm−1). The effective comprising spectator vibrations can be transmitted through
coupling (<5 cm−1) is significantly smaller due to the dressing the crossing region of ET without undergoing any significant
by the Franck−Condon factor. With that small effective additional dephasing. Contrarily, a wavepacket comprising
coupling, the correlation between the generated superposition reaction-promoting vibrations can undergo rapid dephasing
on the reactant surface and the acceptor vibrational states on due to the nesting of free energy curves, as discussed above.
the product will be random due to the strong amplitude of Vibronic coupling may have varying effects on the ET
energy gap fluctuations produced by solvent polarization. dynamics depending on the extent of vibronic wave function
Therefore, vibrational coherences generated on the reactant delocalization, which strongly depends on the reactant−-
state thus largely dephase on the reactant surface, with a very product energy gap. In the presence of a large reactant−-
negligible transfer to the product surface. Thereby any product energy gap, the bottom acceptor levels are offset from
coherence effect is lost due to strong energy gap disorder the donor level leading to highly localized wave functions on
induced by the surrounding fluctuations, and hence the back- the product potential. The ET along localized channels occur
ET will be predominantly incoherent.121 only when energy fluctuations induce transient degeneracy of
The correlation between the back-ET rate and vibrational the reactant and product. However, the higher accepting level
wavepacket dephasing of certain high-frequency modes, of the product state may be in accidental quasi-degeneracy
mentioned above, can be put in perspective by invoking the with the donor level, therefore resulting in vibrational wave
nested free energy crossings, discussed above (Figure 4). Along function delocalization between the reactant and product wells.
all the vibronic channels furnished by the coupled high- Such wave function delocalization is a result of mixing
frequency modes, multiple crossings between reactant and electronic wave functions of the reactant and product states
product free energies induce dispersion of the vibrational in the presence of a resonant vibration as well as electronic
coherence involving the coupled modes. That dispersion coupling. Due to these vibronic mixing effects, ET may be fast
causes faster dephasing of the wavepacket in acetonitrile along certain intramolecular coordinates.
solvent (1.2 ps) relative to the methanol solvent (3.1 ps). We In a strong coupling regime, where vibronic mixing is
therefore suggest that the mean first passage time of the ET to prominent, photoexcitation will directly access a vibronic CT
the transition state is shorter in acetonitrile compared to state, so the concept of photoexcited ET from reactant to
methanol solvent, and that these values scale proportionally to product is superfluous. However, in an intermediate coupling
the reaction rates. regime, the ET dynamics will generally involve interplay of
In the case of intermolecular or outer-sphere ET systems solvent reorganization and the vibrational reaction coordinate.
with inherently large heterogeneities, the correlations between Certain solvent fluctuations might, for instance, produce
reactant and product states are subject to stronger energetic conditions enabling vibronic resonance, which may provide
disorder originating from solvent reorganization; hence, coherent contributions to the transfer. In the non-adiabatic
vibrational coherence dynamics may present differently. limit, this scenario is precisely that indicated by the nested free
Zinth and co-workers, while studying ET from an electron- energy curve model.
donating solvent (dimethylaniline) to a photoexcited electron- In the past few years, researchers in the groups of
accepting dye (oxazine)with an ET time constant of ca. 80 Rubtsov127,128 and Weinstein129−131 have used infrared
fs, identified high-frequency wavepacket motion of 600 cm−1 perturbations and demonstrated the effect of vibrational
using a 20 fs pump pulse.123−125 While earlier experiments excitation of bridging modes on the ET reactions. The
probed wavepacket motion in the reactant state only, their measurements on donor−bridge−acceptor CT complexes
work probed oscillations in the product state also. The revealed that the rate of ET is highly dependent on the
715 DOI: 10.1021/jacs.8b09059
J. Am. Chem. Soc. 2019, 141, 708−722
Journal of the American Chemical Society Perspective

amount of excess energy localized on the bridge. By applying


an infrared excitation pulse resonant with a bridge vibration, an
ET pathway was selectively turned off or slowed down. One of
the reasons for such a control has recently been proposed to
originate from level matching or mismatching between the
reactant and product states upon vibrational excitation.132
These studies have used vibrations as an active control for the
ET reactions and are potentially promising to shine light on
the vibronic mixing effects, coherent contributions enabled by
vibrations, etc.
Coherences in Biological Systems. Vibronic mixing enables
wave function delocalization in the presence of a disorder
comparable to the electronic coupling and has been
demonstrated for photoinduced ET in photosystem-II reaction
center.98,99 The charge separation process in photosynthesis
converts excitation energy into a stable charge-separated
state.133 In isolated reaction centers that lack quinones, the
absorbed energy relaxes to the lowest state in a few hundred
femtoseconds.134 This lowest energy state evolves into a CT
state characterized by localization of the electron and the hole
on adjacent chlorophyll molecules.135,136
Romero et al.98 and Fuller et al.99 applied 2DES137−139 and
simulations to study the role of coherences during photo-
induced charge separation event in the photosystem II reaction
center. Their measurements revealed coherent oscillations in
the time domain signals like those observed during the energy
transfer in photosynthetic systems.140−143 To determine the
origin of the coherent oscillationselectronic, vibrational, or Figure 6. Effect of resonant vibration on dynamics. The effect of the
mixed electronic−vibrational (vibronic)a comparison was absence or presence of a vibration resonant with the energy gap
made between the experimentally observed pattern of between the exciton and charge-transfer state on electron-vibrational
oscillations and those from Redfield theory-predicted simu- mixing (top panel) and charge-separation dynamics (bottom panel) is
lations. This comparison revealed that the temporal oscillations shown. In the presence of finite electronic coupling (150 cm−1 in this
are predominantly of vibronic origin (mixed electronic− example), under resonant conditions, zero vibrational quanta of the
vibrational), and the coherence is maintained by resonant exciton mixes with higher vibrational quanta of the CT state. This
nonequilibrium vibrational modes between the excitonic states mixing induces delocalization of the vibronic wave functions.
Whereas, under nonresonant conditions, the vibrational wave
and the CT state. The coherences were found to persist on the functions remain localized, so mixing or delocalization is prevented.
time scale of charge separation. These mixed electronic− The mixing of the vibronic wave functions from resonant interaction
vibrational (vibronic) coherences were proposed to enhance results in 3.3-fold increases in the ET rate. This mixing essentially
the rate of charge separation due to vibrational level bridging prevents back-reactions and thus results in directionality and
the electronic gap resulting in wave function delocalization. efficiency of the forward process. Adapted with permission from ref
Such electronic vibrational resonances have been suggested to 5. Copyright 2017 Springer Nature.
be potentially important for enhancing photosynthetic energy
transport.6,88,90,143−147 These types of studies have lead individual wells, but also between the wells over the curve
researchers to propound a quantum design of photosynthesis crossing region. It was further shown that the presence of a
for bioinspired solar-energy conversion.96,148 resonant vibration enables 3.3-fold increase in the population
The effect of the absence and presence of a vibration transfer rate from the exciton state to the CT state.148
resonant with the exciton-CT energy gap on electron- Such interplay between electronic states and coupled
vibrational mixing and the CT dynamics was shown by van vibrational modes has also been predicted in organic
Grondelle and co-workers145,148 and also by Scholes and co- photovoltaic (OPV) material systems. Coercing ET events
workers (on energy transfer dynamics)6 for a two-level one through vibrations make it a particularly attractive design
mode system (Figure 6). In the absence of the resonant principle, which has a potential for preparing more efficient
vibration (vibrational frequency detuned from the energy gap), OPV materials.
the vibrational wave functions of the exciton and the CT state Coherences in Material Systems. A largely accepted
do not overlap; therefore, they are predominantly localized in viewpoint regarding OPVs is that light absorption creates a
their individual potential wells. However, in the presence of the localized exciton that hops randomly and incoherently before
resonant vibration, the vibronic coupling leads to mixing of the dissociating into free charge carriers.149 Achieving efficient
excitonic and CT state potential wells. From the perspective of dissociation of exciton into free carriers has long been
the wavepacket dynamics, under non-resonant conditions the recognized as a vital challenge for molecular-based solar
wavepacket oscillates in the localized potential wells of the cells.149−152 Recent work has indicated that coherence effects
exciton and CT state. However, under resonant conditions, the can be prevalent in OPVs also; therefore, coherence may have
wavepacket oscillates between the potential wells of the exciton an active role in ultrafast charge separation.10 Heeger and co-
and the CT state. Which means that phase coherence of the workers performed transient absorption measurements on
vibronic wave functions is not only maintained in the polymer−fullerene bulk heterojunction solar cells and found
716 DOI: 10.1021/jacs.8b09059
J. Am. Chem. Soc. 2019, 141, 708−722
Journal of the American Chemical Society Perspective

Figure 7. Coherent dynamics in material systems. (a,b) Two-dimensional electronic spectra of a P3HT/PCBM blend at 35 and 300 fs after
photoexcitation. (c) Time dependence of the amplitude of signal at the hole PIA cross-peak and P3HT bleach. (d) Absorption spectrum of
cadmium selenide (CdSe) nanoplatelets showing HX and LX transitions. Shaded curve is the laser pulse spectrum used to measure 2DES data. (e)
2D spectrum at 52 fs showing clear diagonal and anti-diagonal peaks. (f) Kinetic trace extracted at the lower cross peak showing amplitude
modulations as a function of waiting time. The frequency of the modulation corresponds to the energy difference between the two excitonic
statesreferred to as electronic coherence. Panels a−c are adapted with permission from ref 160. Copyright 2014 Springer Nature. Panels d−f are
adapted with permission from ref 87. Copyright 2015 Springer Nature.

that majority of charge carriers are generated within ca. 100 quantum coherent motion of electrons governing the primary
fs.153 It was proposed that the initial photoexcited state is a charge flow, which is coupled to the delocalized nuclear
coherently delocalized state instead of a localized exciton, and motion across the porphyrin−fullerene region. The CT was
excitonic delocalization is responsible for ultrafast charge shown to occur in ca. 70 fs, and on this time scale the launched
separation over long distance.153 Friend and co-workers electronic wavepacket of 30 fs period oscillates between the
studied the time dependence of the separation of photo- carotenoid−porphyrin group and the fullerene.158 They further
generated electron−hole pairs across the donor−acceptor extended the study to a conjugated polymer−fullerene blend
heterojunction.154 They found that the Coulombic barrier and observed coherent vibrational motion of the fullerene
between electron and hole is surmounted in ca. 40 fs, which acceptor after impulsive excitation of the P3HT polymer
would be possible only in the presence of rapid charge motion donor.159 An electronic oscillation was coherently produced on
away from the interface through delocalized band states. The the fullerene acceptor after photoexcitation of the polymer
electronic motion on the acceptor site was modeled to be blend, and the period of this oscillation (25 fs) matches with
coherently coupled to the nearest neighbors, which gave rise to the vibrational frequency of the blend. It was proposed that
a band of delocalized eigenstates, and this coherent coupling upon optical excitationwhich creates an electron−hole pair
causes charge separation over a distance of 3−5 nm within a on the P3HT polymerthe strong electronic−vibrational
few hundred femtoseconds.154 coupling induces a delocalization of the electronic wave
Burghardt and co-workers applied a quantum mechanical function across the interface. The correlated motion of the
treatment of the combined electronic and nuclear evolution to electron density and the nuclei on the same time scale was
study the ultrafast charge separation in the conjugated proposed to be responsible for the early time ultrafast charge
oligomer−fullerene complex of P3HT and C60.155 The initial separation in this model OPV system.159
CT dynamics was shown to be guided by vibronic coherence, Once this coherently driven delocalization of electronic wave
which determines the observed oscillatory features of the function across the donor−acceptor interface catches up with
transient dynamics. It was shown that coupled phonon the phonon cloudmolecular relaxation that stabilizes the
modesparticularly high-frequency modes of the acceptor chargethe localization of electron density occurs. This
are essential to dissipate energy and induce dephasing, thus localization, however, occurs at some distance from the
leading to immediate charge separation in ca. 100 fs. The work interface resulting in reducing the electron−hole attraction.
predicts that it is the vibronic coupling which directs charge Further, due to the associated phonon interaction, the electron
separation, instead of the thermally induced barrier crossing as must now be transported forward via hopping, thereby
considered previously.156,157 reducing the chance of back-reactions.10
Lienau and co-workers investigated the primary CT process To illustrate the experimental detection of coherent
in a prototypical artificial reaction centera supramolecular oscillations in an OPV active layer, 2DES measurements of
triad of carotene−porphyrin−fullerene.158 Combined exper- the P3HT−PCBM blend are shown in Figure 7.160 Excitation
imental and theoretical data provided evidence for a correlated, of a P3HT donor, when blended with fullerene acceptor, the
717 DOI: 10.1021/jacs.8b09059
J. Am. Chem. Soc. 2019, 141, 708−722
Journal of the American Chemical Society Perspective

Figure 8. Mapping electron and nuclear dynamics. (a) Direct measurement of ultrafast charge transfer in amino acid phenylalanine after a sub-300-
attosecond XUV pump excitation. The charge transfer manifested itself as a sub-4.5 fs-oscillation in the formation of the doubly charged immonium
ion suggestive of periodic electronic wavepacket motion. (b) Theoretically simulated variation of hole density after XUV excitation as a function of
time. Snapshots at time intervals of 4.4 and 3.4 fs show same hole density at the highlighted amino group suggestive of the periodic wavepacket
dynamics. (c) Schematic drawing of the ultrafast electron diffraction experiment which uses an optical pump pulse and an electron probe pulse. The
subsequent real-space analysis of diffraction patterns of CF3I molecule are shown. (d) Detailed wavepacket dynamics over multiple potential energy
surfaces of CF3 I deduced from UED experiments. Panels a and b are reprinted with permission from ref 162. Copyright 2014 AAAS. Panels c and
d are reprinted with permission from ref 170. Copyright 2018 AAAS.

C60 derivative PCBM, results in rapid ET from the P3HT to energy disorder.161 Indeed, spectral line widths of the
the fullerene domain. This is confirmed by a growth of the electronic transitions would imply the dephasing time scale
P3HT hole transient signalindicated by a photoinduced of <100 fs. A shift in the perspective has, however, been offered
absorption that is clearly resolved as a cross peak in the 2DES by the developments in the area of photosynthetic energy
maps (Figure 7a−c). Vibrational wavepackets generated in transfer where coherence effects are more robust to dephasing
both the ground and excited states of the P3HT exciton when electronic energy gaps are matched by vibrational energy
spectral region are later detected in the product photoinduced gaps. We anticipate that such vibronically enabled coherence is
absorption signal of the P3HT hole. Such transfer of a functionally relevant for a dynamical process like ET at least in
wavepacket is an indicative of either the spectator nature of the phase space representing a quantum mixture of donor and
modes or a more obscured reactive significance. This example
acceptor states. Putting it in perspective, the donor and
shows how vibrational wavepackets can label excited states and
provide information through amplitude and phase of the acceptor potentials may be local at their equilibrium positions,
oscillations as the reactant state evolves into the product state but the crossing region can be adiabatic, and hence the system
during ET event. may repeatedly pass from a localized to delocalized basis (or
In addition to observing vibrational wavepackets, the incoherent to coherent) as a function of vibronically coupled
prowess of 2DES in detecting electronic coherences exclusively nuclear coordinate. Therefore, the functional role of
was shown by Cassette et al. between two excitonic states, coherences should not rely on the fragile and rapidly dephasing
heavy-hole exciton (HX) and light-hole exciton (LX), of a superposition of states. Instead, it represents the trajectories in
semiconductor “nanoplatelet” colloid dispersed in a solution at that small phase space through which donor states transport
ambient temperature (Figure 7d−f).87 The amplitude of the electrons into the acceptor states.
cross peak, HX-LX and LX-HX, in the 2D signal map oscillate We can ask, What is needed to shift these pioneering studies
as a function of waiting time, showing that the quantum from being notional to incisive? A key question relates to how
amplitude of HX and LX bands are correlated until the we think of “coherences” in ET reactions beyond what we
excitonic superposition dephases. The dephasing time constant know. What are the experimental variables, which can provide
of this ensemble was found to be of 13 fs, while the typical decisive information about the coherent or incoherent nature
electronic decoherence time constant is ≤100 fs. of the process? How do we establish the role of coherence for

■ PERSPECTIVE ON COHERENCES IN ELECTRON


TRANSFER
function? To this end, we must categorize more definitely the
different ways that coherence contributes to ET reactions. Is
adiabaticity significant in this regard, or obvious? Is the role of
The argument against a functional role of coherence in high-frequency intramolecular modes the key? Do we need to
complex systems is that the phase-locking of the wave develop new experiments which can distinguish the effects of
functions is rapidly lost by dephasing in systems having strong local environment and complex electronic and vibrational
718 DOI: 10.1021/jacs.8b09059
J. Am. Chem. Soc. 2019, 141, 708−722
Journal of the American Chemical Society Perspective

structure of the systems more adequately than current the wavepacket become out of step owing to dispersion of the
femtosecond experiments? transition states for the different pathways. Therefore, the
Pure optical spectroscopies have so far been pivotal in reactive wavepackets dephase rapidly as the system reaches the
revealing time-domain coherent oscillations. We believe that transition state.
the functional role of coherences might emerge by exploring The direct participation of coherence in controlling function
real-time atomic-scale spatial imaging of charge and nuclear is a fascinating idea that might possibly have revolutionary
dynamics. That would require time resolution into the prospects, as discussed recently.6 To date, the most direct
attosecond regimecapable of revealing electron correlations implication has been envisioned in OPVs, in which coherent
and non-adiabatic dynamics.162 Emerging X-ray free electron generation of mobile charge carriers has been hypothesized. It
laser (XFEL) sources capable of producing ultrashort X-ray is actually the subsequent decoherence which is important for
pulses at high repetition rates hold the key to capturing the inhibiting recombination of the electrons and holes.10
atomically resolved structures.163−165 X-rays have enough Irrespective of the amount of work done in the field of
momentum to provide a complete map of the electronic and coherences, much more is needed to get a better under-
vibrational states. A suite of X-ray-based diffractive, spectro- standing of coherences in ET. Often researchers are tempted to
scopic, and imaging techniques can be applied to capture the focus solely on measurement of time-domain wavepacket
motion of electron density excited locally and subsequently oscillations. The challenge, though, is to disentangle oscillatory
moving rapidly across to the acceptor. signals from the actual role played by vibrations or coherence.
A demonstration of the impact of attosecond spectroscopy In our studies we have found that in most systems, oscillations
was shown by Calegari et al. for the CT dynamics in amino result from spectator modes; therefore, they have little
acid phenylalanine where CT manifested itself as a sub-4.5-fs relevance to the process. One must fully explore the large
oscillation (Figure 8a,b).162 The initial creation of electronic state-space, identify the reactive vibrations, and then assess
coherences and coherently driven electronic rearrangement coherences along those reactive modes. Coherence in ET is
can be probed by mapping the charge as a function of time. firmly connected to adiabaticity, environment fluctuations and
Electronic coherence would mean that charge oscillates from their time scales, and intramolecular vibrations, so exploring
reactant to the product potentials before localization, which the effect of these parameters on reactive wavepacket dynamics
can be captured by free electron lasers.162,166,167 Because of the could help us in first identifying the presence of coherent
spatial resolution, this technique can further probe the position effects and then assessing their potential functional relevance.
of nuclei as the electronic wave function evolves, and hence the The key question is whether coherence in ET can provide new
vibronic effects, which may give rise to coherent dynamics, can function or whether it simply modifies details of the
also be probed. mechanism? Exploring these issues in ET may provide new
Ultrafast electron diffraction (UED) is another direct probe connections between chemical design at the molecular level
of structural dynamics exhibiting high spatial and temporal and energy transduction dynamics and efficacy.


resolution.168,169 UED uses electron pulses to probe structural
dynamics from the diffraction pattern created by the electron AUTHOR INFORMATION
pulses, after optical pulses access high-energy states. Recent
Corresponding Author
work by Yang et al. imaged conical intersection and
photodissociation dynamics of CF3I using UED (Figure *gscholes@princeton.edu
8c,d).170,171 Conical intersections, as described above, play ORCID
an important role in controlling excited-state dynamics and Shahnawaz Rafiq: 0000-0002-9660-4495
have been proposed to be functional in the coherent back-ET Gregory D. Scholes: 0000-0003-3336-7960
of π-stacked donor−acceptor complexes, etc. UED has been Notes
shown to image real-space trajectories of coherent nuclear
The authors declare no competing financial interest.


wavepacket bifurcating at the branching space of the conical
intersection. The beauty of UED like XFELs is that it probes
both electronic and nuclear motions simultaneously, so if a ACKNOWLEDGMENTS
correlation exists between the electronic and the nuclear S.R. and G.D.S. acknowledge support from the Division of
motion, these techniques can resolve such coherent Chemical Sciences, Geosciences, and Biosciences, Office of
interactions in both time and space. Therefore, we believe Basic Energy Sciences, of the U.S. Department of Energy
that validating the functional role of coherences requires to go through Grant No. DE-SC0015429.


beyond the pure optical spectroscopies and move toward time-
resolved dynamical probes. REFERENCES

■ CONCLUSION
In this Perspective we discussed how vibrational coherences
(1) Engel, G. S.; Calhoun, T. R.; Read, E. L.; Ahn, T.-K.; Mancal, T.;
Cheng, Y.-C.; Blankenship, R. E.; Fleming, G. R. Nature 2007, 446
(7137), 782−786.
can help characterize some elements of reactant-product (2) Cheng, Y.-C.; Fleming, G. R. Annu. Rev. Phys. Chem. 2009, 60,
interactions in electron transfer reactions. That character- 241−262.
ization hinges on watching the evolution of coherences (3) Collini, E.; Wong, C. Y.; Wilk, K. E.; Curmi, P. M. G.; Brumer,
P.; Scholes, G. D. Nature 2010, 463 (7281), 644−U69.
involving the spectator and reactive vibrational modes. (4) Scholes, G. D.; Fleming, G. R.; Olaya-Castro, A.; van Grondelle,
Coherences comprising spectator modes do not undergo R. Nat. Chem. 2011, 3 (10), 763−774.
phase dispersion during ET; therefore, no additional dephasing (5) Romero, E.; Novoderezhkin, V. I.; van Grondelle, R. Nature
of the wavepacket is induced. Contrarily, in the ET reactions 2017, 543, 355−365.
where nested free energy curves provide multiple kinetic (6) Scholes, G. D.; et al. Nature (London, U. K.) 2017, 543, 647−
pathways from reactant to product, the phases that make up 656.

719 DOI: 10.1021/jacs.8b09059


J. Am. Chem. Soc. 2019, 141, 708−722
Journal of the American Chemical Society Perspective

(7) Dijkstra, A. G.; Wang, C.; Cao, J. S.; Fleming, G. R. J. Phys. (46) Koch, M.; Rosspeintner, A.; Adamczyk, K.; Lang, B.; Dreyer, J.;
Chem. Lett. 2015, 6 (4), 627−632. Nibbering, E. T. J.; Vauthey, E. J. Am. Chem. Soc. 2013, 135 (26),
(8) Caruso, F.; Chin, A. W.; Datta, A.; Huelga, S. F.; Plenio, M. B. 9843−9848.
Phys. Rev. A: At., Mol., Opt. Phys. 2010, 81 (6), 062346. (47) Rosspeintner, A.; Koch, M.; Angulo, G.; Vauthey, E. J. Am.
(9) O’Reilly, E. J.; Olaya-Castro, A. Nat. Commun. 2014, 5, 3012. Chem. Soc. 2012, 134 (28), 11396−11399.
(10) Bredas, J. L.; Sargent, E. H.; Scholes, G. D. Nat. Mater. 2017, (48) Kosower, E. M.; Dodiuk, H.; Tanizawa, K.; Ottolenghi, M.;
16 (1), 35−44. Orbach, N. J. Am. Chem. Soc. 1975, 97 (8), 2167−2178.
(11) Walschaers, M.; Schlawin, F.; Wellens, T.; Buchleitner, A. Annu. (49) Huppert, D.; Kanety, H.; Kosower, E. M. Faraday Discuss.
Rev. Condens. Matter Phys. 2016, 7, 223−248. Chem. Soc. 1982, 74, 161−175.
(12) Chenu, A.; Scholes, G. D. Annu. Rev. Phys. Chem. 2015, 66, 69− (50) Kosower, E. M.; Huppert, D. Chem. Phys. Lett. 1983, 96 (4),
96. 433−435.
(13) Ke, Y. L.; Liu, Y. X.; Zhao, Y. J. Phys. Chem. Lett. 2015, 6 (9), (51) Kramers, H. A. Physica 1940, 7, 284−304.
1741−1747. (52) Zusman, L. D. Chem. Phys. 1980, 49 (2), 295−304.
(14) Park, H.; et al. Nat. Mater. 2016, 15 (2), 211−216. (53) Friedman, H. L.; Newton, M. D. Faraday Discuss. Chem. Soc.
(15) Demadis, K. D.; Hartshorn, C. M.; Meyer, T. J. Chem. Rev. 1982, 74, 73−81.
2001, 101 (9), 2655−2686. (54) Sumi, H.; Marcus, R. A. J. Chem. Phys. 1986, 84, 4894−4914.
(16) Reimers, J. R.; Wallace, B. B.; Hush, N. S. Philos. Trans. R. Soc., (55) Rips, I.; Jortner, J. J. Chem. Phys. 1987, 87 (4), 2090−2104.
A 2008, 366 (1862), 15−31. (56) Garg, A.; Onuchic, J. N.; Ambegaokar, V. J. Chem. Phys. 1985,
(17) Wasielewski, M. R. Chem. Rev. 1992, 92 (3), 435−461. 83 (9), 4491−4503.
(18) Ponseca, C. S.; Chábera, P.; Uhlig, J.; Persson, P.; Sundström, (57) Walker, G. C.; Aakesson, E.; Johnson, A. E.; Levinger, N. E.;
V. Chem. Rev. 2017, 117 (16), 10940−11024. Barbara, P. F. J. Phys. Chem. 1992, 96 (9), 3728−3736.
(19) Xiang, D.; Wang, X.; Jia, C.; Lee, T.; Guo, X. Chem. Rev. 2016, (58) Jortner, J.; Bixon, M. J. Chem. Phys. 1988, 88, 167−170.
116 (7), 4318−4440. (59) Yoshihara, K.; Tominaga, K.; Nagasawa, Y. Bull. Chem. Soc. Jpn.
(20) Grabowski, Z. R.; Rotkiewicz, K.; Rettig, W. Chem. Rev. 2003, 1995, 68 (3), 696−712.
103 (10), 3899−4032. (60) Barbara, P. F.; Meyer, T. J.; Ratner, M. A. J. Phys. Chem. 1996,
(21) Liu, Y.; Wang, M. K.; Zhao, F.; Xu, Z. A.; Dong, S. J. Biosens. 100 (31), 13148−13168.
Bioelectron. 2005, 21 (6), 984−988. (61) Barbara, P. F.; et al. In Dynamics and Mechanisms of
(22) Kumpulainen, T.; Lang, B.; Rosspeintner, A.; Vauthey, E. Chem. Photoinduced Electron Transfer and Related Phenomena, 1st ed.;
Rev. 2017, 117, 10826−10939. Mataga, N., Okada, T., Masuhara, H., Eds.; North Holland, 1992;
(23) Rafiq, S.; Bezdek, M. J.; Koch, M.; Chirik, P. J.; Scholes, G. D. J. pp 21−38.
Am. Chem. Soc. 2018, 140 (20), 6298−6307. (62) Tominaga, K.; Kliner, D. A. V.; Johnson, A. E.; Levinger, N. E.;
(24) Rafiq, S.; Yadav, R.; Sen, P. J. Phys. Chem. A 2011, 115 (30), Barbara, P. F. J. Chem. Phys. 1993, 98 (2), 1228−1243.
8335−8343. (63) Kobayashi, T.; Takagi, Y.; Kandori, H.; Kemnitz, K.; Yoshihara,
(25) Elsaesser, T. Chem. Rev. 2017, 117 (16), 10621−10622. K. Chem. Phys. Lett. 1991, 180 (5), 416−422.
(26) Gruebele, M.; Zewail, A. H. J. Chem. Phys. 1993, 98 (2), 883− (64) Walker, G. C.; Barbara, P. F.; Doorn, S. K.; Dong, Y. H.; Hupp,
902. J. T. J. Phys. Chem. 1991, 95 (15), 5712−5715.
(27) Zewail, A. H. Nature 2001, 412 (6844), 279−279. (65) Barbara, P. F.; Walker, G. C.; Smith, T. P. Science 1992, 256
(28) Baskin, J. S.; Zewail, A. H. J. Chem. Educ. 2001, 78 (6), 737− (5059), 975−981.
751. (66) Matyushov, D. V. Acc. Chem. Res. 2007, 40 (4), 294−301.
(29) Marcus, R. A. J. Chem. Phys. 1956, 24 (5), 966−978. (67) Ghorai, P. K.; Matyushov, D. V. J. Chem. Phys. 2006, 124 (14),
(30) Marcus, R. A. Can. J. Chem. 1959, 37 (1), 155−163. 144510.
(31) Marcus, R. A. J. Chem. Phys. 1956, 24 (5), 979−989. (68) Ghorai, P. K.; Matyushov, D. V. J. Am. Chem. Soc. 2005, 127
(32) Levich, V. G.; Dogonadze, R. R. Dokl. Akad. Nauk SSSR 1959, (47), 16390−16391.
124 (1), 123−126. (69) LeBard, D. N.; Lilichenko, M.; Matyushov, D. V.; Berlin, Y. A.;
(33) Hopfield, J. J. Proc. Natl. Acad. Sci. U. S. A. 1974, 71 (9), 3640− Ratner, M. A. J. Phys. Chem. B 2003, 107 (51), 14509−14520.
3644. (70) Cukier, R. I.; Nocera, D. G. Annu. Rev. Phys. Chem. 1998, 49,
(34) Kestner, N. R.; Logan, J.; Jortner, J. J. Phys. Chem. 1974, 78 337−369.
(21), 2148−2166. (71) Soudackov, A.; Hammes-Schiffer, S. J. Chem. Phys. 1999, 111
(35) Ulstrup, J.; Jortner, J. J. Chem. Phys. 1975, 63 (10), 4358−4368. (10), 4672−4687.
(36) Siders, P.; Marcus, R. A. J. Am. Chem. Soc. 1981, 103 (4), 748− (72) Soudackov, A.; Hammes-Schiffer, S. J. Chem. Phys. 2000, 113
752. (6), 2385−2396.
(37) Jortner, J. J. Chem. Phys. 1976, 64 (12), 4860−4867. (73) Soudackov, A.; Hatcher, E.; Hammes-Schiffer, S. J. Chem. Phys.
(38) Miller, J. R.; Beitz, J. V.; Huddleston, R. K. J. Am. Chem. Soc. 2005, 122 (1), 014505.
1984, 106 (18), 5057−5068. (74) Hammes-Schiffer, S.; Soudackov, A. V. J. Phys. Chem. B 2008,
(39) Brunschwig, B. S.; Ehrenson, S.; Sutin, N. J. Phys. Chem. 1987, 112 (45), 14108−14123.
91 (18), 4714−4723. (75) Hammes-Schiffer, S. J. Am. Chem. Soc. 2015, 137 (28), 8860−
(40) Miller, J. R.; Calcaterra, L. T.; Closs, G. L. J. Am. Chem. Soc. 8871.
1984, 106 (10), 3047−3049. (76) Skone, J. H.; Soudackov, A. V.; Hammes-Schiffer, S. J. Am.
(41) Gould, I. R.; Ege, D.; Mattes, S. L.; Farid, S. J. Am. Chem. Soc. Chem. Soc. 2006, 128 (51), 16655−16663.
1987, 109 (12), 3794−3796. (77) Sirjoosingh, A.; Hammes-Schiffer, S. J. Phys. Chem. A 2011, 115
(42) McCleskey, T. M.; Winkler, J. R.; Gray, H. B. J. Am. Chem. Soc. (11), 2367−2377.
1992, 114 (17), 6935−6937. (78) Wolynes, P. G. J. Chem. Phys. 1987, 86 (4), 1957−1966.
(43) Fox, L. S.; Kozik, M.; Winkler, J. R.; Gray, H. B. Science 1990, (79) Beratan, D. N.; Hopfield, J. J. J. Am. Chem. Soc. 1984, 106 (6),
247 (4946), 1069−1071. 1584−1594.
(44) Rosspeintner, A.; Lang, B.; Vauthey, E. Annu. Rev. Phys. Chem. (80) Onuchic, J. N.; Beratan, D. N. J. Am. Chem. Soc. 1987, 109
2013, 64, 247−271. (22), 6771−6778.
(45) Dereka, B.; Koch, M.; Vauthey, E. Acc. Chem. Res. 2017, 50 (2), (81) Beratan, D. N.; Onuchic, J. N.; Hopfield, J. J. J. Chem. Phys.
426−434. 1987, 86 (8), 4488−4498.

720 DOI: 10.1021/jacs.8b09059


J. Am. Chem. Soc. 2019, 141, 708−722
Journal of the American Chemical Society Perspective

(82) Jordan, K. D.; Paddonrow, M. N. Chem. Rev. 1992, 92 (3), (118) Levine, B. G.; Martinez, T. J. Annu. Rev. Phys. Chem. 2007, 58,
395−410. 613−634.
(83) Curtiss, L. A.; Naleway, C. A.; Miller, J. R. J. Phys. Chem. 1995, (119) Farag, M. H.; Jansen, T. L. C.; Knoester, J. J. Phys. Chem. Lett.
99 (4), 1182−1193. 2016, 7 (17), 3328−3334.
(84) Skourtis, S. S.; Waldeck, D. H.; Beratan, D. N. J. Phys. Chem. B (120) Kovalenko, S. A.; Eilers-Konig, N.; Senyushkina, T. A.;
2004, 108 (40), 15511−15518. Ernsting, N. P. J. Phys. Chem. A 2001, 105, 4834−4843.
(85) Prytkova, T. R.; Kurnikov, I. V.; Beratan, D. N. Science 2007, (121) Rafiq, S.; Scholes, G. D. Chem. Phys. Lett. 2017, 683, 500−
315 (5812), 622−625. 506.
(86) Beratan, D. N.; Skourtis, S. S.; Balabin, I. A.; Balaeff, A.; Keinan, (122) Silva, W. R.; Frontiera, R. R. Phys. Chem. Chem. Phys. 2016,
S.; Venkatramani, R.; Xiao, D. Q. Acc. Chem. Res. 2009, 42 (10), 18, 20290−20297.
1669−1678. (123) Seel, M.; Engleitner, S.; Zinth, W. Chem. Phys. Lett. 1997, 275
(87) Cassette, E.; Pensack, R. D.; Mahler, B.; Scholes, G. D. Nat. (3−4), 363−369.
Commun. 2015, 6, 6086. (124) Wolfseder, B.; Seidner, L.; Domcke, W.; Stock, G.; Seel, M.;
(88) Chin, A. W.; Prior, J.; Rosenbach, R.; Caycedo-Soler, F.; Engleitner, S.; Zinth, W. Chem. Phys. 1998, 233 (2−3), 323−334.
Huelga, S. F.; Plenio, M. B. Nat. Phys. 2013, 9 (2), 113−118. (125) Engleitner, S.; Seel, M.; Zinth, W. J. Phys. Chem. A 1999, 103
(89) Christensson, N.; Kauffmann, H. F.; Pullerits, T.; Mancal, T. J. (16), 3013−3019.
Phys. Chem. B 2012, 116 (25), 7449−7454. (126) Rafiq, S.; Dean, J. C.; Scholes, G. D. J. Phys. Chem. A 2015,
(90) Dean, J. C.; Mirkovic, T.; Toa, Z. S. D.; Oblinsky, D. G.; 119 (49), 11837−11846.
Scholes, G. D. Chem. 2016, 1 (6), 858−872. (127) Lin, Z.; Lawrence, C. M.; Xiao, D.; Kireev, V. V.; Skourtis, S.
(91) Pullerits, T.; Zigmantas, D.; Sundstrom, V. Proc. Natl. Acad. Sci. S.; Sessler, J. L.; Beratan, D. N.; Rubtsov, I. V. J. Am. Chem. Soc. 2009,
U. S. A. 2013, 110 (4), 1148−1149. 131 (50), 18060−18062.
(92) Hush, N. S. Trans. Faraday Soc. 1961, 57 (4), 557−580. (128) Yue, Y. K.; Grusenmeyer, T.; Ma, Z.; Zhang, P.; Schmehl, R.
(93) Hush, N. S. J. Chem. Phys. 1958, 28 (5), 962−972. H.; Beratan, D. N.; Rubtsov, I. V. Dalton Trans. 2015, 44 (18), 8609−
(94) Reimers, J. R.; Hush, N. S. Chem. Phys. 2004, 299 (1), 79−82. 8616.
(95) Reimers, J. R.; Hush, N. S. Chem. Phys. 1996, 208 (2), 177− (129) Delor, M.; Archer, S. A.; Keane, T.; Meijer, A. J. H. M.;
193. Sazanovich, I. V.; Greetham, G. M.; Towrie, M.; Weinstein, J. A. Nat.
(96) Ma, F.; Romero, E.; Jones, M. R.; Novoderezhkin, V. I.; van Chem. 2017, 9 (11), 1099−1104.
Grondelle, R. J. Phys. Chem. Lett. 2018, 9 (8), 1827−1832. (130) Delor, M.; Keane, T.; Scattergood, P. A.; Sazanovich, I. V.;
(97) Novoderezhkin, V. I.; Romero, E.; Prior, J.; van Grondelle, R. Greetham, G. M.; Towrie, M.; Meijer, A. J. H. M.; Weinstein, J. A.
Phys. Chem. Chem. Phys. 2017, 19 (7), 5195−5208.
Nat. Chem. 2015, 7 (9), 689−695.
(98) Romero, E.; Augulis, R.; Novoderezhkin, V. I.; Ferretti, M.;
(131) Delor, M.; Scattergood, P. A.; Sazanovich, I. V.; Parker, A. W.;
Thieme, J.; Zigmantas, D.; van Grondelle, R. Nat. Phys. 2014, 10 (9),
Greetham, G. M.; Meijer, A. J. H. M.; Towrie, M.; Weinstein, J. A.
676−682.
Science 2014, 346 (6216), 1492−1495.
(99) Fuller, F. D.; et al. Nat. Chem. 2014, 6 (8), 706−711.
(132) Ma, Z.; Lin, Z. W.; Lawrence, C.; Rubtsov, I. V.; Antoniou, P.;
(100) Vos, M. H.; Rappaport, F.; Lambry, J. C.; Breton, J.; Martin, J.
Skourtis, S. S.; Zhang, P.; Beratan, D. N. Chem. Sci. 2018, 9 (30),
L. Nature 1993, 363 (6427), 320−325.
(101) Vos, M. H.; Jones, M. R.; Hunter, C. N.; Breton, J.; Lambry, J. 6395−6405.
C.; Martin, J. L. Biochemistry 1994, 33 (22), 6750−6757. (133) Blankenship, R. E. Molecular mechanisms of photosynthesis;
(102) Vos, M. H.; Jones, M. R.; Hunter, C. N.; Breton, J.; Martin, J. Blackwell Science: Oxford/Malden, MA, 2002.
L. Proc. Natl. Acad. Sci. U. S. A. 1994, 91 (26), 12701−12705. (134) Durrant, J. R.; Hastings, G.; Joseph, D. M.; Barber, J.; Porter,
(103) Rafiq, S.; Scholes, G. D. J. Phys. Chem. A 2016, 120 (34), G.; Klug, D. R. Proc. Natl. Acad. Sci. U. S. A. 1992, 89 (23), 11632−
6792−6799. 11636.
(104) Bixon, M.; Jortner, J. J. Chem. Phys. 1997, 107 (5), 1470− (135) Steffen, M. A.; Lao, K. Q.; Boxer, S. G. Science 1994, 264
1482. (5160), 810−816.
(105) Larsson, S. J. Am. Chem. Soc. 1981, 103 (14), 4034−4040. (136) Rappaport, F.; Diner, B. A. Coord. Chem. Rev. 2008, 252 (3−
(106) Reid, P. J.; Silva, C.; Barbara, P. F.; Karki, L.; Hupp, J. T. J. 4), 259−272.
Phys. Chem. 1995, 99 (9), 2609−2616. (137) Jonas, D. M. Annu. Rev. Phys. Chem. 2003, 54, 425−463.
(107) Barbara, P. F.; Walker, G. C. Abstracts of Papers Am. Chem. (138) Brixner, T.; Mancal, T.; Stiopkin, I. V.; Fleming, G. R. J. Chem.
Soc. 1992, 203, 49-Phys. Phys. 2004, 121 (9), 4221−4236.
(108) Kambhampati, P.; Son, D. H.; Kee, T. W.; Barbara, P. F. J. (139) Turner, D. B.; Dinshaw, R.; Lee, K.-K.; Belsley, M. S.; Wilk, K.
Phys. Chem. A 2000, 104 (46), 10637−10644. E.; Curmi, P. M. G.; Scholes, G. D. Phys. Chem. Chem. Phys. 2012, 14
(109) Arnett, D. C.; Voehringer, P.; Scherer, N. F. J. Am. Chem. Soc. (14), 4857−4874.
1995, 117 (49), 12262−12272. (140) Jun, S.; Yang, C.; Isaji, M.; Tamiaki, H.; Kim, J.; Ihee, H. J.
(110) Wynne, K.; Reid, G. D.; Hochstrasser, R. M. J. Chem. Phys. Phys. Chem. Lett. 2014, 5 (8), 1386−1392.
1996, 105 (6), 2287−2297. (141) Roscioli, J. D.; Ghosh, S.; LaFountain, A. M.; Frank, H. A.;
(111) Wynne, K.; Hochstrasser, R. M. Electron Transfer-from Isolated Beck, W. F. J. Phys. Chem. Lett. 2017, 8 (20), 5141−5147.
Molecules to Biomolecules, Pt 2 2007, 107, 263−309. (142) Ma, F.; Yu, L.-J.; Hendrikx, R.; Wang-Otomo, Z.-Y.; van
(112) Potter, E. D.; Herek, J. L.; Pedersen, S.; Liu, Q.; Zewail, A. H. Grondelle, R. J. Phys. Chem. Lett. 2017, 8 (12), 2751−2756.
Nature 1992, 355 (6355), 66−68. (143) Richards, G. H.; Wilk, K. E.; Curmi, P. M. G.; Davis, J. A. J.
(113) Mokhtari, A.; Cong, P.; Herek, J. L.; Zewail, A. H. Nature Phys. Chem. Lett. 2014, 5 (1), 43−49.
1990, 348 (6298), 225−227. (144) Tiwari, V.; Peters, W. K.; Jonas, D. M. Proc. Natl. Acad. Sci. U.
(114) Rose, T. S.; Rosker, M. J.; Zewail, A. H. J. Chem. Phys. 1988, S. A. 2013, 110 (4), 1203−1208.
88 (10), 6672−6673. (145) Novoderezhkin, V. I.; Romero, E.; van Grondelle, R. Phys.
(115) Hoffman, D. P.; Ellis, S. R.; Mathies, R. A. J. Phys. Chem. A Chem. Chem. Phys. 2015, 17 (46), 30828−30841.
2014, 118 (27), 4955−4965. (146) Jumper, C. C.; Rafiq, S.; Wang, S.; Scholes, G. D. Curr. Opin.
(116) Hoffman, D. P.; Mathies, R. A. Acc. Chem. Res. 2016, 49 (4), Chem. Biol. 2018, 47, 39−46.
616−625. (147) Novelli, F.; Nazir, A.; Richards, G. H.; Roozbeh, A.; Wilk, K.
(117) Manthe, U.; Koppel, H. J. Chem. Phys. 1990, 93 (3), 1658− E.; Curmi, P. M. G.; Davis, J. A. J. Phys. Chem. Lett. 2015, 6 (22),
1669. 4573−4580.

721 DOI: 10.1021/jacs.8b09059


J. Am. Chem. Soc. 2019, 141, 708−722
Journal of the American Chemical Society Perspective

(148) Romero, E.; Novoderezhkin, V. I.; van Grondelle, R. Nature


2017, 543 (7645), 355−365.
(149) Clarke, T. M.; Durrant, J. R. Chem. Rev. 2010, 110 (11),
6736−6767.
(150) Sariciftci, N. S.; Smilowitz, L.; Heeger, A. J.; Wudl, F. Science
1992, 258 (5087), 1474−1476.
(151) Peet, J.; Kim, J. Y.; Coates, N. E.; Ma, W. L.; Moses, D.;
Heeger, A. J.; Bazan, G. C. Nat. Mater. 2007, 6 (7), 497−500.
(152) Ke, Y.; Liu, Y.; Zhao, Y. J. Phys. Chem. Lett. 2015, 6 (9),
1741−1747.
(153) Kaake, L. G.; Moses, D.; Heeger, A. J. J. Phys. Chem. Lett.
2013, 4 (14), 2264−2268.
(154) Gelinas, S.; Rao, A.; Kumar, A.; Smith, S. L.; Chin, A. W.;
Clark, J.; van der Poll, T. S.; Bazan, G. C.; Friend, R. H. Science 2014,
343 (6170), 512−516.
(155) Tamura, H.; Martinazzo, R.; Ruckenbauer, M.; Burghardt, I. J.
Chem. Phys. 2012, 137 (22), 22A540.
(156) Tamura, H.; Ramon, J. G. S.; Bittner, E. R.; Burghardt, I. Phys.
Rev. Lett. 2008, 100 (10), 107402.
(157) Tamura, H.; Ramon, J. G. S.; Bittner, E. R.; Burghardt, I. J.
Phys. Chem. B 2008, 112 (2), 495−506.
(158) Rozzi, C. A.; et al. Nat. Commun. 2013, 4, 1602.
(159) Falke, S. M.; et al. Science 2014, 344 (6187), 1001−1005.
(160) Song, Y.; Clafton, S. N.; Pensack, R. D.; Kee, T. W.; Scholes,
G. D. Nat. Commun. 2014, 5, 1−7.
(161) Scholes, G. D. J. Phys. Chem. Lett. 2018, 9 (7), 1568−1572.
(162) Calegari, F.; et al. Science 2014, 346 (6207), 336−339.
(163) Kupper, J.; et al. Phys. Rev. Lett. 2014, 112 (8), 083002.
(164) Minitti, M. P.; et al. Phys. Rev. Lett. 2015, 114 (25), 255501.
(165) Heinz, T.; Shpyrko, O. Opportunities for Basic Research at the
Frontiers of XFEL Ultrafast Science, Basic Energy Sciences Roundtable;
U.S. Department of Energy, 2017.
(166) Remacle, F.; Levine, R. D. Proc. Natl. Acad. Sci. U. S. A. 2006,
103 (18), 6793−6798.
(167) Kuleff, A. I.; Cederbaum, L. S. J. Phys. B: At., Mol. Opt. Phys.
2014, 47 (12), 124002.
(168) Ischenko, A. A.; Weber, P. M.; Miller, R. J. D. Chem. Rev.
2017, 117 (16), 11066−11124.
(169) Ihee, H.; Lobastov, V. A.; Gomez, U. M.; Goodson, B. M.;
Srinivasan, R.; Ruan, C. Y.; Zewail, A. H. Science 2001, 291 (5503),
458−462.
(170) Yang, J.; et al. Science 2018, 361 (6397), 64−67.
(171) Fielding, H. H. Science 2018, 361 (6397), 30−31.

722 DOI: 10.1021/jacs.8b09059


J. Am. Chem. Soc. 2019, 141, 708−722

You might also like