Robert Strich,
Shirin Faraji
and
Elisa Palacino-González
*
Institute of Theoretical and Computational Chemistry, Heinrich-Heine Universität Düsseldorf, Germany. E-mail: elisa.palacino.gonzalez@hhu.de
First published on 19th August 2025
Quantum coherences play a central role in a broad range of fields, including functional energy materials, biological systems, and molecular quantum information science. Coherences encode critical information about the phase and dynamics of a system, and their interaction with its environment. Particularly, the ultrafast charge transfer process between electron donor and acceptor species in functional energy materials is influenced by vibronic coherences. A key limitation arises from the dephasing of coherences due to dissipation, causing loss of information and limiting applications of molecular systems. Extending and controlling coherence lifetimes would enable the rational design of smarter materials with optimised properties. Here we introduce a novel idea using chirped excitations as a pathway to extend quantum coherence lifetimes, enhancing their robustness against dissipation. A detailed analysis of the light-induced molecular quantum dynamics and wave packet evolution from first-principles models constructed at the donor–acceptor heterojunction of an organic photovoltaic blend is discussed. We demonstrate that tuning the chirp of the excitation pulse, vibronic coherence lifetimes can be extended up to the picosecond timescale. Chirped excitations also enable tunable spatial localisation of the induced wave packet, with localisation controlled by the chirp intensity. These effects are observed consistently across different donor–acceptor adducts selected from the molecular dynamics structure of the blend. Our results introduce a new degree of freedom for coherent control in molecular systems, offering a promising pathway toward the development of advanced functional energy materials and applications in molecular quantum information science.
Achieving precise control over quantum processes in molecules remains a central barrier. Molecular systems feature fast, coupled electronic and nuclear motion that is highly sensitive to the environment. Designing laser pulses that can steer such dynamics with fidelity is non-trivial, especially when decoherence and dissipation play a role. Additionally, translating theoretical control schemes into experimentally feasible protocols often faces limitations in pulse shaping and system characterization. As a result, considerable effort has been devoted to developing approaches for coherent control of molecular systems with external pulses.5,7,8,11,22,25,29,42,44,51,59,62–64,66,67 In this context, ultrafast shaped laser pulses are tailored in amplitude, phase, or polarization to drive specific quantum pathways. Pulse shaping has proven to be a versatile tool for tailoring excitation conditions to extend coherence lifetimes and steer molecular dynamics with high precision. Early experimental27 and theoretical studies37 on control with tailored femtosecond pulses laid the groundwork for this field, revealing the potential of shaped femtosecond pulses to selectively control vibrational excitation and molecular dissociation pathways. Later theoretical studies14,50 have significantly advanced the theoretical framework of this field, demonstrating how shaped pulses can be used to manipulate nonadiabatic transitions, entangle electronic and vibrational degrees of freedom, and explore the coherent control of photo-induced processes in molecular systems. Their work emphasizes the sensitivity of vibronic coherence evolution to both the spectral content and phase structure of excitation pulses, underscoring the need for fine pulse design in coherent control protocols. Among pulse shaping techniques, linearly chirped pulses have attracted particular interest over the past decades. Several theoretical and experimental studies have shown that chirped excitations provide an additional degree of freedom to control the temporal ordering of excitation pathways and the population of vibronic states.2,3,9,13,24,31–34,38,53,55 For example, they can effectively enhance coherence excitations and be used to manipulate vibronic populations and coherence dynamics in molecular systems, revealing their potential to uncover rich dynamical information from complex molecular systems.
Building on prior work by some of the present authors,53 on a minimal model with charge-transfer signatures to investigate the effect of chirped excitation on double-pump femtosecond fluorescence spectroscopic signals, here an advancement is presented. This prior work revealed that chirped pulses can reduce long-term charge-transfer population and enable selective tuning of short-time dynamics via modulation of wavepacket motion. Built on this, double-pump fluorescence signals revealed that chirped excitation enhanced vibrational signatures in fluorescence spectra, offering better interpretability under low-signal-to-noise ratio. Tested on a physical model, these findings lay the foundation for the work introduced here, where the model is improved to include a first-principles description of a charge-transfer dimer from a prototypical organic photovoltaic blend. Besides, a detailed analysis of quantum coherences lifetime, in the presence of dissipation, is described alongside its connection with the properties of the chirped excitation. A first-principles Hamiltonian is constructed using time-dependent density functional theory (TDDFT) calculations in combination with molecular dynamics (MD) simulations for describing the intermolecular charge-transfer process in a prototypical organic photovoltaic blend. A detailed analysis of the role of chirped excitations for manipulating the system quantum dynamics is presented. Particularly, we demonstrate that the lifetime of quantum coherences can be significantly extended by using chirped excitations. Our results indicate that chirped excitation induces a vibronic wave packet with a substantially longer coherence lifetime compared to that generated by a transform-limited (TL) pulse. Importantly, this extension is tunable, with longer lifetimes observed at higher chirp rates. In addition, chirped pulses lead to greater localization of the wave packet, enhancing the vibronic character of the system. These insights not only advance our understanding of quantum coherence evolution under shaped laser fields but also suggest promising strategies for designing coherence-preserving quantum control protocols in dissipative environments, with potential applications in quantum information science45,47,60 and advanced functional energy materials.18
The remainder of this paper is organized as follows. Section 2 outlines the theoretical methodology employed throughout this work. We begin by presenting the formalism used to describe the system and its interaction with a dissipative environment, focusing on ultrafast intermolecular charge transfer in a prototypical organic photovoltaic (OPV) blend, the poly(3-hexylthiophene):phenyl-C61-butyric acid methyl ester (P3HT:PCBM).1,12,17,20,21,23,26,36,43,52,56,58,65 We detail the first-principles approach developed to model the dimer adduct at the heterojunction, followed by a discussion of the dissipative environment arising from the blend. Subsequently, we introduce the concept of linearly chirped pulses and describe the system–field Hamiltonian utilized in our simulations. The section concludes with a presentation of the quantum master equation governing the time evolution of the system in the presence of the dissipative environment. Section 3 presents the results, including an analysis of the population dynamics and wave packet evolution (Section 3.1), Wigner function representations and quantum coherences (Section 3.2), and additional results for an alternative model system based on a different dimer configuration within the blend (Section 3.3). Finally, Section 4 provides concluding remarks and perspectives for future work. The SI is provided in a separate file and contains a detailed account of the numerical methods employed, including TDDFT and MD simulations, molecular orbital analyses, and additional data from the quantum dynamical simulations.
H = HS + HB + HSB. | (1) |
Here, HS denotes the Hamiltonian of the system of interest, HB the bath Hamiltonian, and HSB their interaction. The system dynamical information is contained in the reduced density matrix ρ(t), defined by taking the trace of the overall system ρtot(t) over the bath degrees of freedom.
To show the purpose of the present work on the effect of a chirped excitation on extending the lifetime of vibronic coherences, we focus on a well-known photophysical process, i.e. the ultrafast intramolecular charge-transfer. In this work we focus on the P3HT:PCBM organic photovoltaic blend. Upon photoexcitation, an exciton delocalization is created along the polymer chain (the local excitation), and within a sub-100 fs timescale, a electron-transfer process takes place to a charge-transfer state, leaving a charge-separation between the P3HT (donor) and the PCBM (acceptor). An illustration of the blend is shown in Fig. 2, with a close-up of the dimer adduct at the heterojunction.
HS represents a P3HT:PCBM dimer adduct from the blend, such as the highlighted in the close-up on Fig. 1. A detailed description of how HS has been modelled as first principles calculations is contained in Section 2.1.1. The bath Hamiltonian HB describing the collective effect of the blend environment surrounding a specific dimer adduct, and the system–bath interaction HSB Hamiltonian are further discussed in the following Section 2.1.2.
HS = Hg + He, | (2) |
Hg = |g〉hg〈g|, | (3) |
![]() | (4) |
![]() | (5) |
![]() | (6) |
Here, the |LE〉 state is optically bright, whereas the lower-lying |CT〉 state is optically dark. Hence, the raising and lowering components of the transition dipole moment operator are defined as
X = ê![]() ![]() | (7) |
In eqn (7), ê denotes the unit vector of the polarization of the pulse, and μg,LE the electronic transition dipole vector, μg,LE = |μg,LE|, g,LE = μg,LE/μg,LE. In this study, we set for convenience ê
g,XT = 1.
A parametrization of the P3HT:PCBM heterojunction followed TDDFT calculations using the computational software Q-Chem 6.216 on dimer structures obtained from a MD trajectory on the P3HT:PCBM blend from previous work by some of the present authors.41 The MD simulation was simulated for a total duration of 100 ns at a reference temperature of 298.15 K. Additional details of the simulation and trajectory parameters are provided in the SI. To represent the realistic conditions of the blend, the following protocol was adopted: several P3HT:PCBM dimer adduct structures were sampled from different regions of the blend. To reflect the configurational heterogeneity within the MD blend, we focused on a single MD snapshot containing dimers with markedly different relative orientations and intermolecular distances. The relative orientation between the two species and the intermolecular distance was different for the dimers sampled. Only dimers with intermolecular distance between 3–5 Å were considered. The model parameters obtained from the TDDFT calculations include adiabatic excitation energies for the locally-excited and charge-transfer states, frequencies of the system vibrational modes, dimensionless displacements Δe and value of the electronic coupling. Single-point TDDFT calculations were performed on the selected geometry of the dimer adduct using the CAMB3LYP functional and 6-31+G* basis set, yielding the adiabatic excitation energies. The reason of performing the single-point excited state calculations on the geometries directly obtained from the blend, without performing a prior geometry optimization is the following: such an optimization would yield a geometry that deviates significantly from the true structure of the blend system. Which would lose the essential blend configurational information that we address in this study. It should be noted that the vibrational mode accounted for in the quantum dynamical simulations is a relatively high-frequency mode, corresponding to a stretching mode of the CC bond in the P3HT backbone, therefore, its frequency is not expected to change significantly upon geometry optimization. The selection of the functional was made after carefully benchmarking several functionals including B3LYP and ωB97X-D. B3LYP, although widely used, is well known to underestimate excitation energies and fail to correctly describe CT states due to the lack of long-range correction. In our tests, ωB97X-D produced CT states with inconsistent orbital character (non-physical charge-transfer from the PCBM to the P3HT), complicating the reliable assignment of excitonic and CT transitions. In contrast, CAM-B3LYP yielded orbital characters and excitation energies in line with the expected physics of our system, offering a consistent and minimal long-range correction necessary for the accurate description of CT states. Following the orbital character and excited states analysis, described in the SI, CAM-B3LYP was chosen as the most appropriate functional for this system. To reduce the computational times, the hexyl groups of the P3HT have been substituted by methyl groups. This approach has been previously adopted in on theoretical studies on modelling the excitonic delocalization of P3HT30,35,41 and it is shown that it does not affect the electronic and optical properties of the P3HT and similar conjugated polymers. Following a detailed analysis of the molecular orbitals for the first 20 electronic states in the dimer, the characterization of the excited states reveals the presence of a strong optically “bright” electronic state with locally-excited character on the P3HT (ELE = 3.0402 eV). The lower-lying electronic states have a very low transition dipole, and can be viewed as optically “dark” states. Particularly, only one of them is associated with a charge-transfer character from the P3HT (donor) to the PCBM (acceptor), being the HOMO orbital in the P3HT and the LUMO orbital in the PCBM. Besides, these orbitals are also shared with the LE state, suggesting a strong charge-transfer coupling between both. The rest of electronic states below the locally excited state reveal a character of local-excitation on the PCBM unit. In a subsequent step, vibrational frequency and gradient and electronic coupling calculations associated to the LE and CT electronic states were performed on the selected dimer adduct. All the (mass weighted) normal modes of the system on the ground electronic state were calculated from TDDFT, and rescaled by
to make them dimensionless. The excited-state potential energy surfaces are modeled as harmonic, with the same vibrational frequencies as the ground state. The displacements along the normal modes were reconstructed using excited-state gradient calculations at the MD snapshot geometry. To achieve this, the gradient obtained from TDDFT was transformed from Cartesian coordinates into dimensionless normal mode coordinates. The displacement for each mode was then determined by taking the ratio of the transformed gradient to the corresponding mode frequency. In this way, the gradient provides a direct estimate of the normal mode displacements upon excitation.10 The diabatic electronic coupling was evaluated using the Boys diabatization method,54 which has been previously demonstrated to be a good method for calculating intermolecular charge-transfer couplings. Further computational details of the TDDFT calculations can be found in the SI.
Out of the sample of dimers, two structures (dimers A and B) differing in intermolecular distance and relative orientation were selected to illustrate the message on this paper on the effect of chirped excitations. Table 1 summarises the system parameters in the diabatic representation for a one-dimensional model constructed on dimer A.
εCT [eV] | εLE [eV] | ΔCT [adim.] | ΔLE [adim.] | Ω [eV] | V [eV] |
---|---|---|---|---|---|
2.684 | 2.822 | 0.538 | 0.0037 | 0.190 | −0.012 |
Table 1 highlights the vibrational frequency selected to model the vibronic dynamics of the system Hamiltonian HS. This mode is the vibration with the largest displacement in the charge-transfer state and is the reaction mode that couples to the bath. The mode corresponds to a CC stretch of the P3HT backbone.
Fig. 2 shows a sketch of the harmonic potential energy functions in the diabatic representation for the ground, first, and second electronic states. The validity of the displaced harmonic oscillator model for this type of systems has been previously validated in extensive theoretical studies by Burghardt et al.57
The model parameters of dimer B and additional dimers analysed from the blend are contained in the SI.
![]() | (8) |
![]() | (9) |
![]() | (10) |
J(ω) = ζω![]() | (11) |
Htot(t) = H − HF(t), | (12) |
In the dipole approximation and the rotating-wave approximation (RWA), the Hamiltonian describing the system–field interaction HF(t) is given by
HF(t) = −μg,LE·E(t), | (13) |
![]() | (14) |
In the present work we leverage the potential behind the properties of a linearly chirped pulse. The particularity of chirped pulses, which set them apart from their TL analogous is, that the carrier frequency evolves linearly in time. This results in a systematic temporal separation of frequency components, either increasing (negative chirp) or decreasing (positive chirp) across the pulse duration. As a consequence, the temporal evolution of the phase of a chirped pulse becomes quadratic, in contrast with the linear temporal evolution of a TL pulse phase. Despite this, chirping a TL pulse does not alter the energy information of the pulse, that is, the Fourier transform of the pulse temporal evolution does not change when chirping it. Here, we highlight that these properties within a chirped pulse can be seen as an additional tool to steer the dynamical evolution of a molecular system.
The temporal evolution of a linearly chirped Gaussian pulse is described by
![]() | (15) |
![]() | (16) |
The temporal envelope of a chirped pulse, the maximum pulse amplitude and the effective pulse duration are described, respectively, by
![]() | (17) |
![]() | (18) |
![]() | (19) |
Fig. 3 shows a sketch of a positively linearly chirped pulse (blue) and its TL analogous (black). Following the equations above, the sketch shows that when chirping a 15 fs TL pulse centred at t = 0 fs, the pulse amplitude becomes smaller at the centre of the pulse while its duration extends over a longer time (here ∼ 400 fs). As previously discussed, both the TL and chirped pulse configurations contain the same total energy. The fundamental distinction lies in the temporal behaviour of their spectral components: in a TL pulse, all frequency components arrive simultaneously, whereas in a chirped pulse, the carrier frequencies, defined by the pulse's spectral resolution, arrive in a time-ordered sequence. This temporal evolution of the frequency within a chirped pulse is governed by the chirp parameter η.
For the simulations contained in Section 3, a base TL pulse of 15 fs pulse duration, carrier frequency ω0 = 3.00 eV and amplitude E0 = 0.001 eV was used. Different chirped pulses were modelled for various values of |η|. The excitation energy region of the TL pulse and from its chirped versions is shadowed in yellow in Fig. 2.
![]() | (20) |
As introduced above, the Redfield superoperator is modelled here from Ohmic general spectral density functions. The dephasing superoperator , however, is modelled from previous work by some of the authors of the current paper, where a TDDFT/MD approach was developed to model electronic dephasing rates on the P3HT:PCBM blend.41 This worked revealed typical electronic dephasing times of ∼100 fs induced by the environment of P3HT and PCBM molecules in a thin-film structure.
The (super) operator accounts for pure electronic dephasing and is given by
![]() | (21) |
Before the arrival of the laser pulse, the system is in thermal equilibrium in the electronic ground state
ρ(−∞) = ρB|G〉〈G|. | (22) |
Here
ρB = ZB−1e−Hg/(kBT) | (23) |
To show the action of the Redfield operator R on the system dynamics explicitly, we write the eigenvalue solution of the system Hamiltonian HS
HS|μ〉 = Eμ|μ〉 | (24) |
![]() | (25) |
![]() | (26) |
The elements of the Redfield tensor can be written as
![]() | (27) |
![]() | (28) |
![]() | (29) |
P = Tr{|e〉〈e|ρe(t)}, | (30) |
Fig. 4 illustrates the population probabilities of the (a) optically bright |LE〉 state and (b) and (c) dark |CT〉 state following excitation with a single pulse. Panel (b) shows the data for different positive chirps scenarios, while the lower panel displays data for negative values of the chirp intensity η. Before the arrival of the pulse, the system is in a thermal equilibrium, following a Boltzmann distribution in the ground electronic state. Upon the pulse arrival, centred at t = 0, |LE〉 gets slowly populated. Panel (a) shows that for a chirp intensity of η = 5, chirping the excitation (blue, red) results in a threefold increase in the population of |LE〉 at long times, once the pulse has gone, in contrast to the excitation with a TL case (black line). Besides, due to the effectively longer duration of a chirped pulse, the excitation of the ground electronic state happens earlier than for the TL, approximately by 150 fs. As a consequence of the longer pulse duration, the population of |LE〉 extends over a longer period, proportional to the pulse duration. Particularly, for the case of negative chirp, periodic beatings are observed to be more pronounced, reflecting a stronger coherent superposition of the different vibrational states excited. An interpretation for this is that chirping the excitation allows the pulse frequency to match the resonance condition with the vibrational energy levels of the system, promoting the creation of constructive interferences, which enhances population transfer to |LE〉. In contrast, a TL pulse lacks this time-dependent resonance, leading to less efficient excitation and a reduced population in |LE〉. As a consequence of the favoured excitation with a chirped pulse, the population of the dark charge-transfer state also increases when the excitation is chirped, as can depicted in panels (b) and (c), in contrast to the unchirped excitation case (black line). The simulations reveal two key phenomena. First, due to the effectively longer duration of a chirped pulse, with a phase evolving quadratically with time, the sign of the chirp can delay (positive chirp) or advance (negative chirp) the population of |CT〉. Our simulations show that increasing the intensity of the chirp parameter enhances this effect. This provides a potential method for indirectly controlling the population of an optically dark state over time. Second, the periodic oscillations of ∼90 fs period observed in panels (b) and (c), reflects that the vibronic structure of the system are enhanced when chirping the excitation. An interesting effect observed is the phase-shift of the population dynamics when increasing the intensity of the chirp.
The effect of chirping the excitation can be seen by exploring the photoinduced wave packet under different pulse excitation conditions. The vibrational density of an electronic state, given by the electronic state density matrix in the space representation defined by the reaction coordinate in the system, Q, can give information about the wave packet. Fig. 5 depicts the time evolution of the wave packet in the dark electronic state, for different pulse excitation conditions. The simulations with a TL pulse excitation (a) reveal a delocalized wave packet along the coordinate space Q, that slowly relaxes due to the effect of the dissipative environment after an almost instantaneous creation with the pulse. The wavepacket is quasi-instantaneously created at Q = 0, revealing the transfer to the dark electronic |CT〉 state from the bright electronic |LE〉 state, is almost immediate after excitation. Panel (a) shows the motion of the wavepacket along the |CT〉 potential, displaying the vibronic beatings as a consequence of the electronic coupling between the two electronic states. The wave packet oscillates along the potential displaying a localized behaviour around Q ∼ 1.5, and broadening along the rest of the coordinate space. A node is observed at Q ∼ 0.5 in the wavepacket evolution before dissipation takes place at long times. This feature originates from the interference phenomena originated during the formation of the wave packet, and it is determined by the system–pulse interaction. Along time, the wave packet slowly dissipates to the minima of the |CT〉 state at Q = 0.5. When the excitation is chirped, two key things happen to the wave packet, which manifest differently depending on the sign of the chirp parameter, η. First is a localization of the wave packet along the potential. Panels (c), (e), (g) and (d), (f), (h) show that the chirped excitation localizes the wave packet along the potential, in contrast to the unchirped excitation case displayed in (a). In particular, the beatings appears to be more localized in two return points at the dark-state potential, reflecting a more localized scenario, revealing enhanced vibrational quantum coherence signatures. This occurs independently on the sign of the chirp parameter, however, for the negative chirp case, vibrational features are even more enhanced with the chirp. Second, when increasing the value of the chirp, the final creation time of the wave packet is shifted along time, and increases with the intensity of the chirp. For a positive chirp (c), (e) and (g), the wave packet is delayed in time, whereas for the case of a negative chirp (d), (f) and (h), a similar effect is observed, where the wave packet appears to be more localized along the potential and the vibrational quantum coherence features are enhanced. A significative difference with respect to the positive chirp scenario depicted is, that the creation of the wave packet is shifted to earlier times when increasing the value of the chirp. This time shift in wave packet creation (whether to earlier or later times) arises from the temporal evolution of the pulse's carrier frequency. In the case of a positive chirp, the higher vibrational levels of the bright state are initially excited, followed by the excitation of lower-lying vibrational states as the frequency decreases over time. Conversely, for a negative chirp, the excitation process occurs in the opposite order, with lower vibrational levels being excited first and higher levels excited later. These different excitation pathways leads to distinct interference patterns, resulting in a delayed wave packet formation for the positive chirp case and an earlier wave packet for the negative one. The nodal signature at Q ∼ 0.5 is still observed when the pulse is chirped. An important aspect to consider is, that the observed effects are not a mere consequence of a longer interaction with the pulse. To check the effect of the pulse interaction time one the shift experienced on the wave packet, an interaction with a TL pulse of similar effective duration to the one in panels (c) and (d), ∼200 fs is simulated, shown in panel (b). We observe that even though the interaction with the pulse is present for a longer time, the creation time of the wave packet is not shifted from the central arrival time of the pulse (t = 0 fs). This indicates that the shift described on the creation time of the wave packet in the dark state potential is not due to the longer interaction time, but to the phase interference between the oscillatory phase of the chirp and the one associated to the originating wave packet.
![]() | (31) |
The analysis of the WDF provides information about the quantum phenomena in a system, such as the the presence of quantum coherence and interference effects. Fig. 6 shows the WDF for the vibrational density in the |CT〉 state. The vertical axis is the vibrational coordinate Q in the system and the horizontal axis represents the momentum P. Exciting the system with a pulse of constant carrier frequency leads to a wave packet that oscillates harmonically along the potential defined by Q. The negative region present in the WDF, that reflect the quantum coherence character of the wave packet, originates with the arrival of the pulse at ∼20 fs. These vibrational coherences survive up to ∼500 fs, slowly dissipating due to the weak interaction with the bath. As previously describes in the population dynamics and wave packet analysis section, chirping the excitation introduces key differences in the photoinduced dynamics. This is reflected in the WDF as two main signatures. First, the maxima intensity of the WDF is shifted in time with respect to the central arrival time of the pulse, at t = 0 fs, depending on the direction of the chirp, in consonance with the wave packet simulations. Second, the quantum coherence signatures (blue features) survive for a longer time when the pulse is chirped (b) and (c). An interesting effect observed is the revival of the negative features in the positive chirped WDF for >340 fs, for the pulse conditions used. This hints to a weak reversibility of the wave packet within the dissipative bath, which brings back the non-classical behaviour of a dissipating wave packet.
The appearance of negative values in the WDF is a crucial hallmark of quantum phenomena, being an indicator of non-classicality. This offers a powerful tool for quantifying quantum coherence and revealing non-classical effects such as superposition and entanglement. A convenient measure of this has been previously introduced as defined by the negative volume fraction,19 given by
![]() | (32) |
This theoretical observation offers a potential approach for estimating the position of a crossing between coupled states, which can be of experimental relevance. The comparison between the two models indicates that achieving wave packet localization via interaction with chirped pulses requires the excitation energy to be above the crossing point. This situation can be achieved by setting a detuning between the TL pulse central frequency and the system resonance energy, where the pulse energy is higher.
Last, the δ−(t) was calculated for the WDF of the charge-transfer electronic state, along the propagation time, and it is shown in panel (e) of Fig. 8. The simulations reveal that the coherences lifetime extends from 40 fs to 500 fs and 900 fs when using a negative and positive chirped pulses, respectively. This suggests that the enhanced robustness of coherences due to chirping is an intrinsic property of a chirped pulse. The interference between the oscillatory phase of the chirp and that of the originating wave packet sustains these coherences for an extended duration, counteracting the effects of dissipation.
The simulations show that chirping the excitation pulse significantly enhances the transfer of population to optically dark states and the vibronic features, in comparison to an excitation with a TL pulse. The analysis of the wave packets in the charge-transfer state for different chirp conditions shows that it is possible to exert some control on the formation time of the wave packet, advancing or delaying it over the TL excitation case. Besides, the chirp can serve as a tool to localise the wave packet in the charge-transfer state, affecting its overall lifetime before irreversible relaxation due to the interaction with the bath. The extent of the localization of the wave packet in the dark state is also sensitive to the relative displacements of the potentials. For crossing points located at small displacements at the excited states, the localization of the wave packet is strong, when using a chirped excitation. For higher crossings at higher energies, creating such localized wave packet requires a chirped excitation with a higher detuning of the pulse central frequency, being the resonance condition the energy difference between the minima of the ground state and the minima of the bright electronic state. A key observation, reported in this work, is the substantial impact of the chirp on the wave packet evolution, where the chirped pulse significantly extends the lifetime of vibrational coherences, supported by analysis of WDF for different excitonic conditions. This underscores the non-trivial interplay between the phase evolution of the excitation pulse and the intrinsic system dynamics. This insight is crucial for understanding how the temporal control in the excitation can determine the robustness of quantum systems over extended timescales.
The analysis presented in this work was extended to different sets of system parameters, derived from first-principles parametrisation of various dimer adducts within the blend. It was observed that these chirp-induced effects are generally present across adducts located in different regions of the P3HT:PCBM blend, indicating a broader behaviour characteristic of this type of donor–acceptor system. The two models used to illustrate this effect differed significantly in the displacements of the potential energy functions characterising the local excitation and charge transfer states, as well as in the strength of the electronic coupling. It was shown that the relative position of the crossing between the coupled electronic states plays a crucial role in determining the influence of the chirp on wave packet localisation and on extending the lifetime of light-induced vibronic coherences. In addition, it was demonstrated that by tuning the properties of the chirped pulse, such as the chirp intensity, the evolution of the field-induced dynamics can be substantially manipulated to achieve a desired target. These results provide evidence for a potential experimental route to indirectly estimate the position of crossings between coupled electronic states. At the same time, they highlight the role of chirped pulses in enhancing the properties of light-induced systems, establishing the chirp as a promising tool to extend quantum coherence lifetimes. This holds relevance for the development of more efficient functional energy materials and for applications in the growing field of molecular quantum information science.
Overall, this work demonstrates the ability of chirped pulses to control both electronic and nuclear degrees of freedom in nonadiabatic systems modelled from first-principles, offering a promising strategy for the manipulation of quantum dynamics. Ongoing work is focused on extending these findings by modelling spectral bath densities from TDDFT/MD data, including a characterization of the dependence of the charge-transfer coupling along the fluctuating MD trajectory.
For additional information, please contact the corresponding author.
Footnote |
† Authors dedicate this paper to Prof. Christel Marian on the occasion of her 70th birthday. |
This journal is © the Owner Societies 2025 |