Abstract
The role of molecular vibrations for the persistence of quantum coherences, recently observed in photoinduced charge transfer reactions in both biological and artificial energy conversion systems at room temperature, is currently being intensely discussed. Experiments using two-dimensional electronic spectroscopy (2DES) suggest that vibrational motion – and its coupling to electronic degrees of freedom – may play a key role for such coherent dynamics and potentially even for device function. In organic photovoltaics materials, strong coupling of electronic and vibrational motion is predicted, especially for ubiquitous C=C stretching vibrations. The signatures of such strong vibronic couplings in 2DES are, however, debated. Here we analyse the effect of strong vibronic coupling in model simulations of 2DES spectra and dynamics for an electronic dimer coupled to a single high-frequency vibrational mode. This system represents the simplest conceivable model for a prototypical donor–acceptor interface in the active layer of organic solar cells. The vibrational mode is chosen to mimic C=C stretching vibrations with typical large vibronic couplings predicted in organic photovoltaics materials. Our results show that the decisive signatures of strong vibronic coupling mediating coherent charge transfer between donor and acceptor are not only temporally oscillating cross-peaks, but also most importantly characteristic peak splittings in the 2DES spectra. The 2DES pattern thus directly reflects the new eigenstates of the system that are formed by strong mixing of electronic states and vibrational mode.
1 Introduction
Photoinduced energy and charge transfer reactions in donor–acceptor molecular systems are fundamental for the energy conversion in natural and artificial light harvesting, in organic photovoltaics (OPV) and photocatalysis [1]. Conventionally, they have mostly been described using quasi-classical rate equation models, such as Förster theory for energy transfer [2] and Marcus theory for electron transfer [3]. In both models, the initial and final states (donor and acceptor) are assumed to be weakly coupled, and the transfer is expected to proceed as an incoherent, monotonic decay of the initial (donor) and a concurrent build-up on the final (acceptor) state population [1], [4], [5]. In the past decade, however, long-lived oscillations, persisting for hundreds of femtoseconds at room temperature, have been observed in spectroscopic experiments of photosynthetic light harvesting proteins [6], [7], [8], [9], [10], [11]. This suggested a coherent, i.e. wave-like, nature of energy migration at early times. Initially interpreted as persisting electronic coherence, these surprising findings have triggered tremendous interest in the potential role that quantum effects could have for the efficiency of energy conversion in photosynthesis and hence for biological functionality [12], [13], [14]. The persistence of such coherences may be a signature of spatial delocalisation of electronic wave functions and thus may give directionality to the transfer. In comparison to incoherent random walks, persistent quantum coherences would make energy transfer within the light harvesting complex faster and more efficient [15]. This led to an intense discussion of the experimental data in the community [16]. In fact, typical electronic dephasing times in molecules are of the order of only few tens of femtoseconds at room temperature [17] and thus decay much faster than the coherence signatures observed in experiments. This is because disorder and environmental fluctuations at room temperature, inherent in these systems, induce a rapid loss of coherent polarisations at optical frequencies between electronic ground and excited states [18]. Hence, it was thought to be unlikely for coherence, although it may be present right upon excitation, to be of any significant functional relevance [16]. To this aim, several experimental and theoretical efforts have been devoted to identifying the mechanisms behind such long-lived coherences [19], [20], [21], [22].
Probing coherent couplings experimentally has been possible by ultrafast multidimensional spectroscopy, in particular two-dimensional electronic spectroscopy (2DES) [23], [24]. The interpretation of 2DES data is, however, not always straightforward and unambiguous. Oscillatory signals in 2DES may arise not only from pure electronic coherence, but also from other quantum pathways [25]. In particular, the coupling between electronic and nuclear motion may also result in long-lived oscillations [26]. There is now growing evidence that the observed beatings in 2DES studies of light harvesting proteins actually reflect the interplay of electronic and vibrational degrees of freedom and hence are indicative of vibronic coherences [15], [26], [27], [28], [29], [30], [31]. During the past years, coherent dynamics of possible vibronic origin have been suggested in a variety of other systems, such as artificial light harvesting complexes [32], [33], molecular aggregates [34], and in some OPV materials [35], [36], [37], [38], [39], [40].
One of the critical and still open issues for the interpretation of 2DES is the isolation of ground and excited state coherences that may arise from excitation of vibrational wave packets [41]. In the experiments, excitation with femtosecond pulses inevitably launches vibrational wave packet motion on both excited and ground state potential energy surfaces. Thus, both these contributions are necessarily present in the 2DES signals and may both result in oscillatory features. While vibrational coherences in the excited state may be relevant for the transport of energy and charges, vibrational coherences in the ground state have no direct impact on these processes, but are mainly spectators. Nevertheless, vibrational wave packet motion in the ground state may result in persistent modulation of the measured 2DES signals and overlap with electronic features, thus complicating the interpretation of the experimental data. In systems with weak coupling to vibrations, such as biological light harvesting complexes, it has been recently shown theoretically that ground state vibrational coherences may dominate the 2DES spectra and enhance oscillatory features in the dynamics, thereby hiding the weaker electronic signatures [41], [42]. Therefore, the interpretation of 2DES data requires considerable care.
In conjugated oligomers and polymers, used, e.g. as active materials in OPV [43], vibronic couplings to high-frequency, underdamped vibrations play an important role for the stationary optical spectra [44], [45]. A measure of the strength of vibronic coupling between electronic states and vibrational modes is the Huang–Rhys factor, defining the relative displacement between the ground and excited state potential energy surfaces [46]. The characteristic rich optical spectra with pronounced vibrational progression in conjugated polymers, such as polythiophene (P3HT), reflect strong vibronic coupling to intramolecular C=C stretching modes oscillating at ∼1400–1500 cm−1 with Huang–Rhys factors of about unity [45], [47]. In many other materials relevant for OPV and organic electronics in general, similarly large Huang–Rhys factors are usually found for the coupling to these high-frequency molecular vibrations [45], [48]. Large vibronic couplings mainly result from the low dielectric function and comparatively high degree of structural disorder [49] that are typical in conjugated materials. This generally localises the electronic excitations to a few molecular units and thus increases electron-vibrational couplings [50]. A consequence of such vibronic couplings is the formation of quasi-particles such as excitons, intramolecular and intermolecular polaron excitons, and polarons [50]. These quasi-particles determine the optical, electronic, and transport properties in these materials. In solid-state organic thin films, the details of the molecular arrangement may also have a significant effect on vibronic couplings [50]. Hence, it appears that strong vibronic coupling may greatly influence the transport of energy and charge and thus play an important role for the initial steps of the energy conversion in OPV materials. Indeed, recent experimental and theoretical works point to the importance of delocalisation and vibronic coupling for the ultrafast coherent charge transfer dynamics of OPV materials [35], [36], [37], [38], [51], [52], [53], [54], [55], [56], [57], [58], [59]. As in the case of photosynthetic proteins, however, the assignment of spectral features, such as cross-peaks or persistent beatings in 2DES, to coherent dynamics underlying charge or energy transfer in OPV has to be treated very carefully. In OPV materials, the effects of such strong vibronic couplings on the shape and dynamics of ultrafast 2DES spectra have received comparatively little attention [60], [61], [62].
Here, we try to bridge this gap by analysing the signatures of strong vibronic coupling on the 2DES pattern for a specific model system, an electronic dimer coupled to a single high-frequency vibrational mode. We start by briefly introducing ultrafast 2DES explaining the signals giving rise to the spectra. We then discuss the general spectral signatures of coupling in 2DES for an electronic dimer in the absence of vibrations. Including vibronic coupling to a high-frequency mode, we discuss the 2DES pattern and the dynamics in two limiting cases. In the first case, vibronic coupling results in large amplitude vibrational wave packet motion in the single moieties without affecting charge transfer. In the second case, we analyse how strong vibronic coupling influences the 2DES spectra and charge transfer dynamics.
Our results suggest that strong vibronic coupling involved in the excited state dynamics results in a splitting of the 2DES peak pattern, in agreement with recent experimental observations [37]. The analysis of such peak splittings provides important information about the structure and dynamics of the excited state wave packets and thus may help in probing coherent charge transfer dynamics in complex OPV materials.
2 Model Simulations
In this section, we introduce the dimer model that we use to discuss the effects of strong vibronic coupling on the charge transfer dynamics in a donor–acceptor system and illustrate the procedure for the calculation of the 2DES signals that are discussed in this article.
2.1 Model Dimer of a Donor–Acceptor Interface
We consider an electronic dimer system as the conceptually simplest conceivable model for a donor–acceptor interface. We define
where

Scheme of the pulse sequence used in a 2DES experiment. Three ultrashort pulses (red) interact with the sample inducing a coherent nonlinear polarisation P(3). The corresponding field Es reemitted by the sample (blue) after interaction with the third pulse is measured as a function of the coherence and waiting times τ and T.
where the transition dipole moment matrix element
To model vibronic coupling, we assume that the electronic system is coupled to a single vibrational mode with vibrational energy Evib along the nuclear coordinate Q for all electronic states. The equilibrium geometry of the ground state is chosen as the origin of the coordinate system Q = 0, and each potential energy surface is described by the same normal mode coordinate. The energy of the ground state then becomes that of a one-dimensional harmonic oscillator, i.e.
For the numerical calculations, we choose to work in the basis of displaced harmonic oscillator states [1]. This basis transformation is obtained introducing the unitary displacement operator for the vibrational mode,
The optical excitation of the vibronic system is described by a similar light–matter interaction Hamiltonian as introduced above (2). Here we make use of the Condon approximation [1], in which the expectation value of the dipole operator between ground state with vibrational level n and the excited state with vibrational level mj, j = D, A, is given by
2.2 Master Equation and Numerical Calculation of the 2DES Signals
For sufficiently weak optical pulses, 2DES is a third-order nonlinear optical spectroscopy technique. The measured signal in 2DES is proportional to the third-order nonlinear polarisation,
The total Hamiltonian
For our model system, we define creation and annihilation operators describing electronic dephasing as
The master equation (3) is solved numerically by a nonperturbative approach, which yields the total polarisation, including all allowed phase-matched signals as well as all linear and third-order nonlinear contributions arising from the interaction with the pulse sequence [66]. In the nonperturbative scheme, the optical pulses are included in the light–matter interaction part of the Hamiltonian, and no assumptions are done on the timings between the three pulses [66]. This has the advantage that the calculated signals include directly the effect of pulse overlaps, which may result, e.g. in perturbed free induction decays or other coherent interferences during pulse overlap [67]. The nonlinear signal in a specific phase-matching condition is subsequently isolated from the total polarisation. Experiments probing absorptive signals are usually most interesting for 2DES of molecular systems. Therefore, in our simulations, we will calculate absorptive pathways by applying a phase-cycling algorithm. To this aim, at each time delay τ, T, we calculate the total polarisation for four different phases
2.3 Wave Packet Dynamics
To discuss the effect of possible ground state contributions to spectra, we calculate wave packet dynamics in the ground and excited donor state of such a dimer model coupled to a high-frequency harmonic oscillator mode for different time profiles of the excitation pulses. Such wave packet dynamics are calculated by projecting the density matrix
where the wave functions are that defined above, and
3 Results and Discussion
A detailed treatment of the signals involved in 2DES has previously been reported in the literature [18], [23], [71], [72], [73]. Here we briefly introduce the main signals that are relevant for the modelling and interpretation of the 2DES spectra presented in the next sections. In a 2DES experiment, the sample is excited by a sequence of three ultrashort optical pulses
In the most general implementation, the three pulses impinge on the sample with different wavevectors
In this report, we therefore focus our attention on the analysis of absorptive 2DES spectra. Such spectra are usually recorded in a partially collinear implementation of 2DES [78], which can be obtained as a straightforward extension of a conventional pump-probe geometry. In this scheme, the first two (pump) pulses

(a) Phase-locked pulse pair used for the excitation in 2DES and scheme of a dimer of noninteracting donor and acceptor with electronic transition dipole moments μD and
Repeating the experiment for many time delays τ at a fixed waiting time, e.g. T = 0 fs, the nonlinear polarisation signal results in a map in the time domain as a function of the excitation pulse delay and the detection time. In Figure 2d, we show the nonlinear polarisation field, reemitted by the system in response to the interaction with the three pulses for a waiting time T = 0 fs, in the phase-matching condition corresponding to absorptive pathways, as introduced above. After a two-dimensional Fourier transform of the reemitted field
3.1 Electronic Coherence and Charge Transfer
To discuss the signatures of strong couplings and coherent charge transfer in 2DES, we first consider the case of electronic coherence without contribution of vibrations. For this we use a dimer model mimicking a donor–acceptor interface and assume that the effect of vibrations merely results in decoherence of the optically driven coherent polarisations, accounted for by a finite decoherence time of the optical excitations. This is equivalent to the case of weak coupling to all molecular vibrations in the system, which is, however, quite unlikely in functional organic materials [50], but helps us to introduce the typical spectral signatures of coherent coupling in 2DES.
For simplicity, we consider only the electronic ground and first (singly) excited state of each moiety and neglect the effect of possible higher-lying electronic states. Specifically, we do not include the possible excited state absorption (ESA) signals that may arise from transitions to these states. This simplification is motivated by the fact that for P3HT ESA of singlet excitons is strongly red-shifted with respect to the excitonic resonances centred around 600 nm [35], [80], [81], [82]. The ESA peaks around 1200 nm, and therefore, such transitions are not excited even when using extremely short optical pulses in the 2DES experiments. Quite generally, such ESA transitions will lead to negative peaks in absorptive 2DES spectra. These negative peaks may easily mask the 2DES signatures of the transitions between the ground and singly excited electronic excited state, which can lead to both ground state bleaching (GSB) or SE. In contrast to ESA, both GSB and SE enhance the transmission through the sample and thus result in positive peaks in absorptive 2DES maps. It will be important for the later discussion that, in OPV materials, negative peaks in 2DES associated with ESA may also arise from other types of quasiparticle excitation, e.g. from polaronic excitations, charge transfer excitons, or also from triplet excitons [81]. Signatures of polaronic excitations in experimental 2DES studies of P3HT thin films and of its blend with fullerene acceptors have indeed been observed in recent experiments [37], [39]. In the case of P3HT, these ESA features reflect photoinduced absorption of delocalised polaron excitons in the polymer aggregate and overlap with the positive SE cross-peaks expected from singlet excitons [37]. This complicates the interpretation of the spectra. In our dimer model, we neglect for simplicity higher excited states that could give rise to ESA transitions, even though they may be relevant for the interpretation of certain experiments.
In this electronic dimer model, we describe the interaction between donor and acceptor by an electronic coupling strength J connecting their excited states. This electronic coupling controls the transfer of energy from one moiety to the other. When this coupling exceeds the damping, i.e. for
In a typical configuration in OPV materials, the donor is the main absorber with a broad absorption covering much of the visible spectral range [43]. The acceptor, on the other hand, is only weakly absorbing and essentially transparent in this range. A specific example for such donor–acceptor systems is a conjugated polymer mixed with a high concentration of fullerene derivatives as acceptors [43]. In the following, we therefore assume
To illustrate the effect of electronic coupling on 2DES spectra, we assume a resonant case with

Electronic dimer in the strong coupling limit. (a) Scheme depicting the excited states of the electronically coupled dimer in the absence (dashed) and presence (solid) of electronic coupling. We have assumed that the (uncoupled) donor and acceptor states are initially resonant and the coupling strength J = 50 meV exceeds the damping
Upon excitation at T = 0 fs, the corresponding absorptive 2DES map shows diagonal peaks, at excitation and detection energies corresponding to the hybrid eigenstates, and their cross-peaks suggesting their interaction (Fig. 3b). The peaks are arranged in a regular square pattern with the splittings defined by the coupling strength and thus reflecting the structure of the excited states (Fig. 3a). For the resonant case in Figure 3, the splitting is 2J = 100 meV. The waiting time dynamics (Fig. 3c–i) show amplitude oscillations of the peaks with a period defined by the energetic splitting and decaying with a time constant determined by T2 (Fig. 3d, g). These dynamics probe the coherent transfer of population between the coupled states of an electronic donor–acceptor dimer. The oscillatory peaks are signature of electronic coherence during the transfer. The appearance of these oscillations implies that electronic charge is transferred from the donor to acceptor, i.e. over a distance of approximately 0.5–1.0 nm, on a timescale of
For coupling strengths smaller than the damping,
3.2 Vibronic Coupling and Charge Transfer
To discuss the effect of vibrations on the 2DES spectra and dynamics, we include coupling of the electronic states to a high-frequency, underdamped vibrational mode in the electronic dimer model. In OPV materials, C=C stretching and ring breathing vibrations in the range of 1400–1500 cm−1 (∼170–180 meV) are usually strongly coupled to electronic excitations with large Huang–Rhys factors of about unity [45], [84] and result in pronounced vibrational progression in the linear optical spectra. In nonlinear experiments of donor conjugated polymers, long-lived vibrational oscillations up to about 1 ps, arising from intramolecular C=C stretching, have been observed [35], [37], [65], [85]. This suggests long vibrational relaxation times of the order of 500–600 fs. For fullerene acceptors widely used in OPV, the dominant breathing mode is found in the same frequency range (1470 cm−1) [84]. Therefore, in our model, we assume that donor and acceptor strongly couple to the same high-frequency vibrational mode with vibrational energy
In light of experimental observations [35], [36], [37], these assumptions seem a reasonable approximation to explain the signatures of vibronic coupling in 2DES in a dimer model mimicking a prototypical donor–acceptor system with parameters relevant for OPV materials. As for the case of the electronic dimer (Section 3.1), we neglect the effect of possible higher excited states that could result in ESA transitions. In the next two sections, we discuss the signatures of vibronic coupling in this dimer model in 2DES for the limiting cases of weak and strong vibronic coupling.
3.2.1 Weak Vibronic Coupling Between Donor and Acceptor
We first consider the case that vibronic coupling does not influence the charge transfer dynamics between an optically bright donor and an optically dark acceptor. For that, we assume an electronic coupling J = 50 meV and the same vibronic coupling strength
As the vibrational mode frequency exceeds by far thermal fluctuations at room temperature, the population is initially in the vibronic ground state

Wave packet dynamics in the (a–c) ground and (d–f) excited state potentials of the donor moiety along the nuclear coordinate Q for increasing values (left to right) of the excitation pulse duration of (a, d) 1 fs, (b, e) 5 fs, and (c, f) 10 fs. Note that, for the ground state wave packets (d–f), the amplitude square of the ground state vibrational wave function
Impulsive stimulated Raman scattering [86], [87] also drives coherent vibrational wave packet motion in the ground state. Here, the shape and dynamics of the wave packet launched in the ground state depend critically on the duration of the excitation pulses.
This can be seen in Figure 4a–c, showing ground state vibrational wave packets for different values of the pulse duration tF. To display the oscillatory part of the ground state wave packet more clearly, we have subtracted the amplitude square of the first vibrational wave function
For the shortest pulses (tF = 1 fs) in our model, the wave packet displays a characteristic oscillation with a period of ∼11.5 fs, half the period of vibrational mode (∼23 fs). For such short pulses, the excited state wave packet will not evolve substantially before being driven back to the ground state. In this case, two weak, counter-propagating wave packets, phase shifted by 180°, are launched in the ground state (Fig. 4a). For longer pulse durations (Fig. 4b, c), the excited state wave packet can evolve during the duration of the pulse. Hence, for the chosen displacement ΔD, the ground state wave packet is preferentially launched near the outer turning point (Q > 0), and the wave packet motion becomes more similar, except for a 180° phase shift, to that in the excited state.
In the 2DES, such coherent vibrational wave packet motion in the excited donor state results in a series of diagonal and cross-peaks. These peaks reflect optical transitions between vibronic ground states

Vibronic dimer in the weak vibronic coupling limit. (a) Scheme of the harmonic oscillator potential energy surfaces of donor and acceptor. (b–c, e–f, h–i) Absorptive 2DES maps show diagonal and cross-peaks arranged in a checkerboard pattern suggesting excitation of a vibrational wave packet mainly in the donor moiety. Peaks for excitation and detection energies between 2.0 and 2.2 eV reflect bleaching transitions of the donor. Cross-peaks at detection energies ED ∼ 1.8 eV arise from SE of the excited donor to the ground state. The peak splittings along detection and excitation energy axes are determined by the frequency of the coupled vibrational mode
The coherent motion of the wave packets in the donor potentials modulates the amplitude of the peaks along the waiting time (Fig. 5b–i), resulting in coherent vibrational beatings with a period of
In this vibronic dimer model, population transfer between donor and acceptor is not affected by the vibronic coupling but it is determined only by the coupling strength between electronic states J and by the energetic detuning between them
This is partly due to the relative displacement between donor and acceptor
With the exception of the SE peaks in Figure 5, the pattern generated by the excitation of a vibrational wave packet in the donor is very similar to the one resulting from strong coherent electronic transfer between donor and acceptor in Figure 3. Importantly, despite the fact that the physics of the underlying system is quite different, both models result in oscillatory dynamics. While in Figure 3 they reflect electronic coherence between donor and acceptor, the beatings in Figure 5 arise from vibrational coherence in only one of the moiety. It is now evident that the interpretation of persistent beatings of the 2DES peaks is not straightforward and may easily lead to misinterpretations of the underlying physics [6], [7]. Hence, the presence of cross-peaks and long-lived beatings in experimental 2DES data cannot be taken as an unambiguous signature of electronic coherence [21], [25], [61], [62]. Simulations of 2DES spectra and their dynamics based on model Hamiltonians, as presented in the present work, or more advanced ab initio methods are therefore indispensable for a substantiated analysis of experimental data.
3.2.2 Strong Vibronic Coupling Mediating Coherent Transfer
We now consider different strengths of the vibronic couplings, i.e. different displacements, for the excited donor and acceptor states in the dimer (Fig. 6a). When the vibronic couplings of donor and acceptor are significantly different, they may modulate the electronic coupling between the states and hence may take an active role in the charge transfer dynamics. In this case, both the 2DES spectral pattern and dynamics may differ significantly from those seen in the case of pure electronic or vibrational coherence. In the dimer model coupled to a high-frequency vibrational mode introduced in Section 3.2, the excited states are electronically coupled to each other and mutually strongly coupled to a single vibrational mode. For a sufficiently strong electronic coupling, this results in the formation of new eigenstates of the system that are mixed states between the electronic donor and acceptor states and the vibrational mode ladder. The coupled potential energy surfaces show an avoided crossing along the vibrational coordinate Q where the uncoupled potentials of each component would instead cross (Fig. 6a, dashed black). The strong mixing of electronic and vibrational states induces delocalisation of the wave function between donor and acceptor. Excitation by the pulse sequence in 2DES results in a coherent superposition of these mixed donor–acceptor vibronic states. The optically excited wave packet is initially launched on the donor side of the coupled potential energy surface. This wave packet is now no more a vibrational wave packet oscillating only within the donor moiety, as in the case discussed in Section 3.2.1, but it is a coherent superposition of vibronic wave functions delocalised across the donor and acceptor. Hence, a complex and completely new energy level structure forms in the excited state. In the 2DES pattern, this results in a splitting of the vibrational wave packet structure defined by the vibrational mode into a subpeak structure, with many peaks (Fig. 6a) corresponding to optical transitions in the ground and the complex mixed excited state.

Strong vibronic coupling mediating coherent charge transfer dynamics between donor and acceptor. (a) Scheme showing the coupled donor–acceptor excited state potential energy surface. Strong mixing of electronic and vibrational degrees of freedom results in vibronic states (solid) delocalised across the donor and acceptor. The dashed lines show the position of vibrational states of donor and acceptor in the absence of coupling. (b–c, e–f, h–i) Absorptive 2DES maps show a clear substructure of the checkerboard pattern on top of the one observed for excitation of a vibrational wave packet in the donor (c.f. Fig. 5). Multiple peak splittings reflect the more complex structure of the excited state and probe the strongly coupled vibronic eigenstates. (d, g) The waiting time dynamics of selected (d) diagonal and (g) cross-peaks of this subpeak structure show clear and persistent beatings, suggesting multiple oscillation frequencies, which reflect the different splittings. Note that, although the value of electronic dephasing time and electronic coupling are the same as in Figure 3, the coherent vibronic oscillations are not damped within the electronic dephasing time T2, but persist on a longer timescale set by the vibrational relaxation time
The waiting time dynamics of diagonal and cross-peaks (Fig. 6d, g) show beatings suggesting multiple oscillation frequencies that reflect both the coherent vibrational motion and also the coherent population oscillations between donor and acceptor induced by the electronic coupling. Such beatings persist on a timescale that is much longer than the electronic dephasing time. In fact, although T2 and J are the same as in case discussed in Sections 3.1 and 3.2.1 (c.f. Figs. 3 and 5d, g), here the coherent oscillations do not decay within T2, but persist on a much longer timescale, which is set by the vibrational relaxation time
Due to the strong vibronic coupling, the optical excitation generates a coherent wave packet of strongly mixed donor–acceptor electronic and vibrational states. As long as vibronic coherence survives, the transfer can be regarded as essentially barrierless across the interface. As such, energetic detuning
We note that, while vibronic coupling strengths govern the timescale for equilibration of the population between donor and acceptor, the relaxation time of the coupled vibrational mode controls the coherent population beatings between the donor and acceptor during transfer. Hence, for shorter
Our results show that the observation of oscillating cross-peaks in 2DES does not unequivocally indicate coherent transfer. Instead, additional features in the 2DES spectra, in particular spectral peak splittings, seem to be more reliable signatures of coherent transfer mediated by a strongly coupled vibration (Fig. 6). Recently, we have indeed observed such peak splittings in ultrafast 2DES experiments of the prototypical polymer P3HT, reflecting strong vibronic couplings between localised excitons and more delocalised polaron excitons within the polymer moiety [37]. In P3HT, spectral subpeaks are detected for both excitonic GSB transitions and for ESA cross-peaks probing transitions within the polaron manifold [37]. An even more complex peak pattern is seen at early waiting times in experimental 2DES maps of P3HT blended with the fullerene acceptor PCBM [39]. These experimental features could reasonably well be explained by strong mixing of the electronic states (excitons and polaron excitons) to the underdamped intramolecular C=C stretching mode, giving rise to delocalised vibronic states [37]. This explanation, based on a relatively simple model Hamiltonian similar to the one used here, has been recently fully confirmed by advanced first-principles quantum theory [88]. The resulting peak structures and dynamics in 2DES are similar to the ones reported in Figure 6. In other 2DES experiments on P3HT, such peak splittings have not been directly resolved [65]. The dynamics show clear peak oscillations with the period of the high-frequency vibrational mode but little other dynamic signatures of strong vibronic coupling. These 2DES spectra can be also understood within the vibronically coupled dimer model presented here, but they reflect vibrational wave packet motion in a weakly coupled dimer, as discussed in Section 3.2.1 in this article. It appears that, despite its simplicity, this dimer model can explain surprisingly well the essential experimental features and thus aid in the interpretation of 2DES experiments of complex OPV materials.
4 Conclusion
In this report, we have discussed how the coupling between electronic and vibrational degrees of freedom affects the photoinduced charge transfer dynamics in donor–acceptor systems and, most importantly, how 2DES can be used to reveal the signatures of such vibronic coupling. For that we have used a simple model dimer system that mimics a typical donor–acceptor interface and takes into account the vibronic coupling to a single high-frequency, underdamped vibrational mode. Our simulations show that strong vibronic coupling can drive coherent charge transfer between donor and acceptor on a timescale that is essentially determined by half the period of the vibrational mode. This is a direct consequence of strong mixing of electronic and vibrational states resulting in delocalisation of the (vibronic) wave functions across the interface. This is in very good agreement with recent results of more sophisticated theoretical studies on the role of vibronic couplings for nonadiabatic excited state dynamics in molecular systems [51], [53], [54]. Our results show that the decisive signatures of such coherent couplings in 2DES experiments are not only oscillating cross-peaks, but also most importantly multiple spectral peak splittings reflecting strong mixing of the donor and acceptor potentials.
The selectivity of vibronic coupling to a high-frequency underdamped mode, as described here, may be indeed expected for many OPV materials due to their strongly pronounced coupling to C=C stretching modes [50]. In well-designed systems, this could promote directionality of the coherent wave packet along specific pathways in the donor moiety and across the donor–acceptor interface. Making use of this selectivity of the vibronic coupling could be the key to exploit vibronic coherence in applications. The interesting perspective of using coherence to optimise specific device functions [38] has been very recently suggested also by other groups [89]. In OPV, it could not only facilitate charge separation at the donor–acceptor interface, but also improve transport of excitons towards the interface with the acceptor. We have previously observed long-lived vibronic coherence on a timescale of ∼600 fs in 2DES experiments of P3HT thin films without any particular morphology optimisation [37]. Considering typical distances of donor and acceptor centres of mass of ∼1 nm and strong coupling to the C=C stretching mode (1450 cm−1, i.e. vibrational period of ∼22 fs) with a vibrational relaxation of ∼600 fs, coherence lengths for ballistic wave packet propagation of ∼55 nm can be estimated. As the molecular structure may greatly influence electronic and vibronic couplings, this suggests that longer coherence lengths might be achieved in optimised structures. In highly ordered molecular wires, coherent exciton transport has been suggested in [90], and more recently, long-range exciton diffusion of ∼200 nm has been estimated in polymer fibres [91]. In disordered OPV materials, long coherence lengths, however, have not been directly measured yet. Recently, efforts to move from the observation of long-lived coherence in some materials to controlling such coherent dynamics are being discussed in the literature [40], [89], [92]. A promising avenue seems to be the strong coupling of molecular excitations in currently available materials to electromagnetic fields in, e.g. cavities or plasmonic nanostructures [93], [94], [95], [96]. Enhancement of the charge mobility could indeed be achieved by strong coupling to surface plasmons in thin film transistors [93]. Theoretical studies also suggest the feasibility of this approach and suggest that it not only can increase coherent transport lengths [96], but also can control the pathways and dynamics of chemical reactions [97]. Hence, harnessing coherence to enhance transport and charge separation in donor–acceptor materials seems a promising approach not only for OPV. An important challenge is the experimental verification and coherent control of such coherences and, most importantly, the rational design of organic materials with optimised coherent transport properties. Our results suggest that advanced methods such as 2DES, in particular when being combined with high-resolution optical or electron microscopy, will be a particularly valuable and powerful tool for probing and controlling coherent charge transport in organic nanomaterials. Our results also emphasise the particular importance of model simulations providing direct comparison with experimental observables. Direct comparison between experiment and theory is indispensable for unravelling the large amount of information contained in experimental 2DES spectra.
Funding source: Deutsche Forschungsgemeinschaft
Award Identifier / Grant number: SPP1839
Award Identifier / Grant number: SPP1840
Award Identifier / Grant number: K20815000003
Funding source: German-Israeli Foundation
Award Identifier / Grant number: 1256
Funding statement: Financial support by the Deutsche Forschungsgemeinschaft (Funder Id: http://dx.doi.org/10.13039/501100001659, grant no. SPP1839 and SPP1840), the Korea Foundation for International Cooperation of Science and Technology (Global Research Laboratory project, K20815000003), and the German-Israeli Foundation (Funder Id: http://dx.doi.org/10.13039/501100001736, grant no. 1256) is gratefully acknowledged. We wish to thank Ephraim Sommer, Filippo Troiani, Carlo Andrea Rozzi, Elisa Molinari, Giulio Cerullo, Jaemin Lim, Susana Huelga, and Martin Plenio for many helpful discussions and valuable contributions to this project.
References
[1] V. May and O. Kühn, Charge and Energy Transfer Dynamics in Molecular Systems, Wiley-VCH, Weinheim, Germany 2011.10.1002/9783527633791Suche in Google Scholar
[2] T. Förster, Ann. Phys. 437, 55 (1948).10.1002/andp.19484370105Suche in Google Scholar
[3] R. A. Marcus, Rev. Mod. Phys. 65, 599 (1993).10.1103/RevModPhys.65.599Suche in Google Scholar
[4] S. Trotzky, T. Hoyer, W. Tuszynski, C. Lienau, and J. Parisi, J. Phys. D: Appl. Phys. 42, 055105 (2009).10.1088/0022-3727/42/5/055105Suche in Google Scholar
[5] M. D. Heinemann, K. von Maydell, F. Zutz, J. Kolny-Olesiak, H. Borchert, et al., Adv. Funct. Mater. 19, 3788 (2009).10.1002/adfm.200900852Suche in Google Scholar
[6] G. S. Engel, T. R. Calhoun, E. L. Read, T.-K. Ahn, T. Mancal, et al., Nature 446, 782 (2007).10.1038/nature05678Suche in Google Scholar PubMed
[7] E. Collini, C. Y. Wong, K. E. Wilk, P. M. G. Curmi, P. Brumer, et al., Nature 463, 644 (2010).10.1038/nature08811Suche in Google Scholar PubMed
[8] T. Brixner, J. Stenger, H. M. Vaswani, M. Cho, R. E. Blankenship, et al., Nature 434, 625 (2005).10.1038/nature03429Suche in Google Scholar PubMed
[9] G. Panitchayangkoon, D. Hayes, K. A. Fransted, J. R. Caram, E. Harel, et al., P. Natl. Acad. Sci. 107, 12766 (2010).10.1073/pnas.1005484107Suche in Google Scholar PubMed PubMed Central
[10] R. Hildner, D. Brinks, J. B. Nieder, R. J. Cogdell, and N. F. van Hulst, Science 340, 1448 (2013).10.1126/science.1235820Suche in Google Scholar PubMed
[11] H. Lee, Y.-C. Cheng, and G. R. Fleming, Science 316, 1462 (2007).10.1126/science.1142188Suche in Google Scholar PubMed
[12] M. Mohseni, Y. Omar, G. S. Engel, and M. B. Plenio, Quantum Effects in Biology, Cambridge University Press, Cambridge, United Kingdom 2014.10.1017/CBO9780511863189Suche in Google Scholar
[13] A. D. L. Lande, N. S. Babcock, J. Rezac, B. Levy, B. C. Sanders, et al., Phys. Chem. Chem. Phys. 14, 5902 (2012).10.1039/c2cp21823bSuche in Google Scholar PubMed
[14] F. Levi, S. Mostarda, F. Rao, and F. Mintert, Rep. Prog. Phys. 78, 082001 (2015).10.1088/0034-4885/78/8/082001Suche in Google Scholar PubMed
[15] S. F. Huelga and M. B. Plenio, Contemp. Phys. 54, 181 (2013).10.1080/00405000.2013.829687Suche in Google Scholar
[16] H.-G. Duan, V. I. Prokhorenko, R. J. Cogdell, K. Ashraf, A. L. Stevens, et al., P. Natl. Acad. Sci. 114, 8493 (2017).10.1073/pnas.1702261114Suche in Google Scholar PubMed PubMed Central
[17] O. V. Prezhdo and P. J. Rossky, Phys. Rev. Lett. 81, 5294 (1998).10.1103/PhysRevLett.81.5294Suche in Google Scholar
[18] S. Mukamel, Principles of Nonlinear Optical Spectroscopy, Oxford University Press, New York 1995.Suche in Google Scholar
[19] E. L. Read, G. S. Engel, T. R. Calhoun, T. Mančal, T. K. Ahn, et al., P. Natl. Acad. Sci. 104, 14203 (2007).10.1073/pnas.0701201104Suche in Google Scholar PubMed PubMed Central
[20] N. Christensson, H. F. Kauffmann, T. Pullerits, and T. Mančal, J. Phys. Chem. B 116, 7449 (2012).10.1021/jp304649cSuche in Google Scholar PubMed PubMed Central
[21] M. B. Plenio, J. Almeida, and S. F. Huelga, J. Chem. Phys. 139, 235102 (2013).10.1063/1.4846275Suche in Google Scholar PubMed
[22] A. Chenu and G. D. Scholes, Annu. Rev. Phys. Chem. 66, 69 (2015).10.1146/annurev-physchem-040214-121713Suche in Google Scholar PubMed
[23] D. M. Jonas, Annu. Rev. Phys. Chem. 54, 425 (2003).10.1146/annurev.physchem.54.011002.103907Suche in Google Scholar PubMed
[24] S. T. Cundiff, and S. Mukamel, Phys. Today 66, 44 (2013).10.1063/PT.3.2047Suche in Google Scholar
[25] V. Butkus, D. Zigmantas, L. Valkunas, and D. Abramavicius, Chem. Phys. Lett. 545, 40 (2012).10.1016/j.cplett.2012.07.014Suche in Google Scholar
[26] V. Tiwari, W. K. Peters, and D. M. Jonas, P. Natl. Acad. Sci. 110, 1203 (2013).10.1073/pnas.1211157110Suche in Google Scholar PubMed PubMed Central
[27] T. Pullerits, D. Zigmantas, and V. Sundström, P. Natl. Acad. Sci. 110, 1148 (2013).10.1073/pnas.1221058110Suche in Google Scholar PubMed PubMed Central
[28] F. D. Fuller, J. Pan, A. Gelzinis, V. Butkus, S. S. Senlik, et al., Nat. Chem. 6, 706 (2014).10.1038/nchem.2005Suche in Google Scholar PubMed
[29] E. Romero, R. Augulis, V. I. Novoderezhkin, M. Ferretti, J. Thieme, et al., Nat. Phys. 10, 676 (2014).10.1038/nphys3017Suche in Google Scholar PubMed PubMed Central
[30] A. W. Chin, J. Prior, R. Rosenbach, F. Caycedo-Soler, S. F. Huelga, et al., Nat. Phys. 9, 113 (2013).10.1038/nphys2515Suche in Google Scholar
[31] V. Tiwari, W. K. Peters, and D. M. Jonas, Nat. Chem. 6, 173 (2014).10.1038/nchem.1881Suche in Google Scholar PubMed
[32] C. A. Rozzi, S. M. Falke, N. Spallanzani, A. Rubio, E. Molinari, et al., Nat. Commun. 4, 1602 (2013).10.1038/ncomms2603Suche in Google Scholar PubMed PubMed Central
[33] J. Lim, D. Paleček, F. Caycedo-Soler, C. N. Lincoln, J. Prior, et al., Nat. Commun. 6, 7755 (2015).10.1038/ncomms8755Suche in Google Scholar PubMed PubMed Central
[34] A. Halpin, J. M. JohnsonPhilip, R. Tempelaar, R. S. Murphy, J. Knoester, et al., Nat. Chem. 6, 196 (2014).10.1038/nchem.1834Suche in Google Scholar PubMed
[35] S. M. Falke, C. A. Rozzi, D. Brida, M. Maiuri, M. Amato, et al., Science 344, 1001 (2014).10.1126/science.1249771Suche in Google Scholar PubMed
[36] Y. Song, S. N. Clafton, R. D. Pensack, T. W. Kee, and G. D. Scholes, Nat. Commun. 5, 4933 (2014).10.1038/ncomms5933Suche in Google Scholar PubMed
[37] A. De Sio, F. Troiani, M. Maiuri, J. Réhault, E. Sommer, et al., Nat. Commun. 7, 13742 (2016).10.1038/ncomms13742Suche in Google Scholar PubMed PubMed Central
[38] A. De Sio and C. Lienau, Phys. Chem. Chem. Phys. 19, 18813 (2017).10.1039/C7CP03007JSuche in Google Scholar PubMed
[39] A. De Sio, F. V. d. A. Camargo, K. Winte, E. Sommer, F. Branchi, et al., Eur. Phys. J. B 91, 236 (2018).10.1140/epjb/e2018-90216-4Suche in Google Scholar
[40] J.-L. Bredas, E. H. Sargent, and G. D. Scholes, Nat. Mater. 16, 35 (2017).10.1038/nmat4767Suche in Google Scholar PubMed
[41] R. Tempelaar, T. L. C. Jansen, and J. Knoester, J. Phys. Chem. B 118, 12865 (2014).10.1021/jp510074qSuche in Google Scholar PubMed
[42] A. Chenu, N. Christensson, H. F. Kauffmann, and T. Mančal, Sci. Rep. 3, 2029 (2013).10.1038/srep02029Suche in Google Scholar PubMed PubMed Central
[43] C. J. Brabec, V. Dyakonov, J. Parisi, and N. S. Sariciftici (Eds.) Organic Photovoltaics, Springer, Berlin, Heidelberg 2003.10.1007/978-3-662-05187-0Suche in Google Scholar
[44] J.-L. Brédas, D. Beljonne, V. Coropceanu, and J. Cornil, Chem. Rev. 104, 4971 (2004).10.1021/cr040084kSuche in Google Scholar PubMed
[45] F. C. Spano, Annu. Rev. Phys. Chem. 57, 217 (2006).10.1146/annurev.physchem.57.032905.104557Suche in Google Scholar PubMed
[46] A. P. Shreve, E. H. Haroz, S. M. Bachilo, R. B. Weisman, S. Tretiak, et al., Phys. Rev. Lett. 98, 037405 (2007).10.1103/PhysRevLett.98.037405Suche in Google Scholar PubMed
[47] J. Clark, C. Silva, R. H. Friend, and F. C. Spano, Phys. Rev. Lett. 98, 206406 (2007).10.1103/PhysRevLett.98.206406Suche in Google Scholar PubMed
[48] C. M. Pochas, and F. C. Spano, J. Chem. Phys. 140, 244902 (2014).10.1063/1.4882696Suche in Google Scholar PubMed
[49] D. Chirvase, J. Parisi, J. C. Hummelen, and V. Dyakonov, Nanotechnology 15, 1317 (2004).10.1088/0957-4484/15/9/035Suche in Google Scholar
[50] S. Kilina, D. Kilin, and S. Tretiak, Chem. Rev. 115, 5929 (2015).10.1021/acs.chemrev.5b00012Suche in Google Scholar PubMed
[51] H. Tamura, J. G. S. Ramon, E. R. Bittner, and I. Burghardt, J. Phys. Chem. B 112, 495 (2008).10.1021/jp077270pSuche in Google Scholar PubMed
[52] R. Binder, D. Lauvergnat, and I. Burghardt, Phys. Rev. Lett. 120, 227401 (2018).10.1103/PhysRevLett.120.227401Suche in Google Scholar PubMed
[53] T. R. Nelson, D. Ondarse-Alvarez, N. Oldani, B. Rodriguez-Hernandez, L. Alfonso-Hernandez, et al., Nat. Commun. 9, 2316 (2018).10.1038/s41467-018-04694-8Suche in Google Scholar PubMed PubMed Central
[54] M. Polkehn, H. Tamura, and I. Burghardt, J. Phys. B-At. Mol. Opt. Phys. 51, 014003 (2017).10.1088/1361-6455/aa93d0Suche in Google Scholar
[55] T. Nelson, S. Fernandez-Alberti, A. E. Roitberg, and S. Tretiak, J. Phys. Chem. Lett. 8, 3020 (2017).10.1021/acs.jpclett.7b00790Suche in Google Scholar PubMed
[56] S. Gélinas, A. Rao, A. Kumar, S. L. Smith, A. W. Chin, et al., Science 343, 512 (2014).10.1126/science.1246249Suche in Google Scholar PubMed
[57] H. Tamura and I. Burghardt, J. Am. Chem. Soc. 135, 16364 (2013).10.1021/ja4093874Suche in Google Scholar PubMed
[58] M. Huix-Rotllant, H. Tamura, and I. Burghardt, J. Phys. Chem. Lett. 6, 1702 (2015).10.1021/acs.jpclett.5b00336Suche in Google Scholar PubMed
[59] S. Joseph, M. K. Ravva, and J.-L. Bredas, J. Phys. Chem. Lett. 8, 5171 (2017).10.1021/acs.jpclett.7b02049Suche in Google Scholar PubMed
[60] A. G. Dijkstra, C. Wang, J. Cao, and G. R. Fleming, J. Phys. Chem. Lett. 6, 627 (2015).10.1021/jz502701uSuche in Google Scholar PubMed
[61] D. Egorova, J. Chem. Phys. 140, 034314 (2014).10.1063/1.4861634Suche in Google Scholar PubMed
[62] D. Egorova, M. F. Gelin, and W. Domcke, J. Chem. Phys. 126, 074314 (2007).10.1063/1.2435353Suche in Google Scholar
[63] M. O. Scully and M. S. Zubairy, Quantum Optics, Cambridge University Press, Cambridge, NY 1997.10.1017/CBO9780511813993Suche in Google Scholar
[64] B. Palmieri, D. Abramavicius, and S. Mukamel, J. Chem. Phys. 130, 204512 (2009).10.1063/1.3142485Suche in Google Scholar
[65] Y. Song, C. Hellmann, N. Stingelin, and G. D. Scholes, J. Chem. Phys. 142, 212410 (2015).10.1063/1.4916325Suche in Google Scholar
[66] M. F. Gelin, D. Egorova, and W. Domcke, Accounts Chem. Res. 42, 1290 (2009).10.1021/ar900045dSuche in Google Scholar
[67] P. Hamm, Chem. Phys. 200, 415 (1995).10.1016/0301-0104(95)00262-6Suche in Google Scholar
[68] L. Seidner, G. Stock, and W. Domcke, J. Chem. Phys. 103, 3998 (1995).10.1063/1.469586Suche in Google Scholar
[69] S. Meyer and V. Engel, Appl. Phys. B 71, 293 (2000).10.1007/s003400000342Suche in Google Scholar
[70] T. Mančal, A. V. Pisliakov, and G. R. Fleming, J. Chem. Phys. 124, 234504 (2006).10.1063/1.2200704Suche in Google Scholar PubMed
[71] S. Mukamel, Annu. Rev. Phys. Chem. 51, 691 (2000).10.1146/annurev.physchem.51.1.691Suche in Google Scholar PubMed
[72] M. Cho, Chem. Rev. 108, 1331 (2008).10.1021/cr078377bSuche in Google Scholar PubMed
[73] P. Hamm and M. Zanni, Concepts and Methods of 2D Infrared Spectroscopy, Cambridge University Press, Cambridge, United Kingdom 2011.10.1017/CBO9780511675935Suche in Google Scholar
[74] S. T. Cundiff, Opt. Express 16, 4639 (2008).10.1364/OE.16.004639Suche in Google Scholar PubMed
[75] M. Khalil, N. Demirdöven, and A. Tokmakoff, Phys. Rev. Lett. 90, 047401 (2003).10.1103/PhysRevLett.90.047401Suche in Google Scholar PubMed
[76] D. Brida, C. Manzoni, and G. Cerullo, Opt. Lett. 37, 3027 (2012).10.1364/OL.37.003027Suche in Google Scholar PubMed
[77] W. D. Zhu, R. Wang, C. F. Zhang, G. D. Wang, Y. L. Liu, et al., Opt. Express 25, 21115 (2017).10.1364/OE.25.021115Suche in Google Scholar PubMed
[78] L. P. DeFlores, R. A. Nicodemus, and A. Tokmakoff, Opt. Lett. 32, 2966 (2007).10.1364/OL.32.002966Suche in Google Scholar
[79] A. P. Heberle, J. J. Baumberg, and K. Köhler, Phys. Rev. Lett. 75, 2598 (1995).10.1103/PhysRevLett.75.2598Suche in Google Scholar PubMed
[80] J. M. Guo, H. Ohkita, S. Yokoya, H. Benten, and S. Ito, J. Am. Chem. Soc. 132, 9631 (2010).10.1021/ja9108787Suche in Google Scholar PubMed
[81] J. M. Guo, H. Ohkita, H. Benten, and S. Ito, J. Am. Chem. Soc. 131, 16869 (2009).10.1021/ja906621aSuche in Google Scholar PubMed
[82] J. Piris, T. E. Dykstra, A. A. Bakulin, P. H. M. van Loosdrecht, W. Knulst, et al., J. Phys. Chem. C 113, 14500 (2009).10.1021/jp904229qSuche in Google Scholar
[83] C. Cohen-Tannoudji, F. Laloèe, and B. Diu, Quantum Mechanics, Wiley, New York 1997.Suche in Google Scholar
[84] S. Falke, P. Eravuchira, A. Materny, and C. Lienau, J Raman Spectrosc 42, 1897 (2011).10.1002/jrs.2966Suche in Google Scholar
[85] C. J. Brabec, G. Zerza, G. Cerullo, S. De Silvestri, S. Luzzati, et al., Chem Phys Lett 340, 232 (2001).10.1016/S0009-2614(01)00431-6Suche in Google Scholar
[86] L. Dhar, J. A. Rogers, and K. A. Nelson, Chem. Rev. 94, 157 (1994).10.1021/cr00025a006Suche in Google Scholar
[87] U. Banin, A. Bartana, S. Ruhman, and R. Kosloff, J. Chem. Phys. 101, 8461 (1994).10.1063/1.468108Suche in Google Scholar
[88] W. Popp, M. Polkehn, R. Binder, and I. Burghardt, J. Phys. Chem. Lett., 3326 (2019).10.1021/acs.jpclett.9b01105Suche in Google Scholar PubMed
[89] G. D. Scholes, G. R. Fleming, L. X. Chen, A. Aspuru-Guzik, A. Buchleitner, et al., Nature 543, 647 (2017).10.1038/nature21425Suche in Google Scholar PubMed
[90] A. T. Haedler, K. Kreger, A. Issac, B. Wittmann, M. Kivala, et al., Nature 523, 196 (2015).10.1038/nature14570Suche in Google Scholar PubMed
[91] X.-H. Jin, M. B. Price, J. R. Finnegan, C. E. Boott, J. M. Richter, et al., Science 360, 897 (2018).10.1126/science.aar8104Suche in Google Scholar PubMed
[92] G. D. Scholes, J. Phys. Chem. Lett. 9, 1568 (2018).10.1021/acs.jpclett.8b00734Suche in Google Scholar PubMed
[93] E. Orgiu, J. George, J. A. Hutchison, E. Devaux, J. F. Dayen, et al., Nat. Mater. 14, 1123 (2015).10.1038/nmat4392Suche in Google Scholar PubMed
[94] F. Herrera and F. C. Spano, Phys. Rev. Lett. 116, 238301 (2016).10.1103/PhysRevLett.116.238301Suche in Google Scholar PubMed
[95] J. Galego, F. J. Garcia-Vidal, and J. Feist, Phys. Rev. X 5, 041022 (2015).10.1103/PhysRevX.5.041022Suche in Google Scholar
[96] J. Feist and F. J. Garcia-Vidal, Phys. Rev. Lett. 114, 196402 (2015).10.1103/PhysRevLett.114.196402Suche in Google Scholar PubMed
[97] J. Galego, F. J. Garcia-Vidal, and J. Feist, Nat. Commun. 7, 13841 (2016).10.1038/ncomms13841Suche in Google Scholar PubMed PubMed Central
© 2019 Walter de Gruyter GmbH, Berlin/Boston
Artikel in diesem Heft
- Frontmatter
- Trends and Perspectives in Energy Research
- Quantitative Assessment of the Influence of Camera and Parameter Choice for Outdoor Electroluminescence Investigations of Silicon Photovoltaic Panels
- Sequentially Deposited Compact and Pinhole-Free Perovskite Layers via Adjusting the Permittivity of the Conversion Solution
- Efficient Solution Processed CH3NH3PbI3 Perovskite Solar Cells with PolyTPD Hole Transport Layer
- Potential of CZTSe Solar Cells Fabricated by an Alloy-Based Processing Strategy
- Two-Dimensional Absorbers for Solar Windows: A Simulation
- Harvesting the Electromagnetic Energy Confined Close to a Hot Body
- Activation of Small Organic Molecules on Ti2+-Rich TiO2 Surfaces: Deoxygenation vs. C–C Coupling
- Recent Advances in the Colloidal Synthesis of Ternary Transition Metal Phosphides
- Signatures of Strong Vibronic Coupling Mediating Coherent Charge Transfer in Two-Dimensional Electronic Spectroscopy
Artikel in diesem Heft
- Frontmatter
- Trends and Perspectives in Energy Research
- Quantitative Assessment of the Influence of Camera and Parameter Choice for Outdoor Electroluminescence Investigations of Silicon Photovoltaic Panels
- Sequentially Deposited Compact and Pinhole-Free Perovskite Layers via Adjusting the Permittivity of the Conversion Solution
- Efficient Solution Processed CH3NH3PbI3 Perovskite Solar Cells with PolyTPD Hole Transport Layer
- Potential of CZTSe Solar Cells Fabricated by an Alloy-Based Processing Strategy
- Two-Dimensional Absorbers for Solar Windows: A Simulation
- Harvesting the Electromagnetic Energy Confined Close to a Hot Body
- Activation of Small Organic Molecules on Ti2+-Rich TiO2 Surfaces: Deoxygenation vs. C–C Coupling
- Recent Advances in the Colloidal Synthesis of Ternary Transition Metal Phosphides
- Signatures of Strong Vibronic Coupling Mediating Coherent Charge Transfer in Two-Dimensional Electronic Spectroscopy