Thomas P.
Fay
*a and
David T.
Limmer
*abcd
aDepartment of Chemistry, University of California, Berkeley, CA 94720, USA. E-mail: tom.patrick.fay@gmail.com; dlimmer@berkeley.edu
bKavli Energy Nanoscience Institute, Berkeley, CA 94720, USA
cChemical Science Division Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA
dMaterial Science Division Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA
First published on 2nd April 2024
Triplet excited state generation plays a pivotal role in photosensitizers, however the reliance on transition metals and heavy atoms can limit the utility of these systems. In this study, we demonstrate that an interplay of competing quantum effects controls the high triplet quantum yield in a prototypical boron dipyrromethene-anthracene (BD-An) donor–acceptor dyad photosensitizer, which is only captured by an accurate treatment of both inner and outer sphere reorganization energies. Our ab initio-derived model provides excellent agreement with experimentally measured spectra, triplet yields and excited state kinetic data, including the triplet lifetime. We find that rapid triplet state formation occurs primarily via high-energy triplet states through both spin–orbit coupled charge transfer and El-Sayed's rule breaking intersystem crossing, rather than direct spin–orbit coupled charge transfer to the lowest lying triplet state. Our calculations also reveal that competing effects of nuclear tunneling, electronic state recrossing, and electronic polarizability dictate the rate of non-productive ground state recombination. This study sheds light on the quantum effects driving efficient triplet formation in the BD-An system, and offers a promising simulation methodology for diverse photochemical systems.
In this work we focus on a prototypical heavy-atom-free photocatalyst, the boron-dipyrromethene-anthracene (BD-An) dyad (chemical structure in Fig. 1B).5,17–19 BD-An has recently found applications in synthetic chemistry20–22 and its derivatives have been investigated for phototheraputic applications.23 The competing photophysical processes and the electronic excited states involved are summarized in Fig. 1. BD-anuses excited state charge transfer from an anthracenyl (An) group to the photoexcited SBD* forming an SCT state, to enable rapid triplet TBD* formation with a high experimental yield, ΦT = 0.93–0.96.18,24 Naively one might expect excited state charge transfer to reduce the triplet quantum yield, since the charge transfer state provides a charge recombination pathway for relaxation to the singlet ground state. However experiments indicate that charge recombination is suppressed by the large charge recombination free energy change, pushing this reverse electron transfer deep into the Marcus inverted regime, where increasing the free energy change increases the activation energy.18 This effect is captured qualitatively by Marcus' theory for the reaction rate constant25,26
(1) |
We aim to investigate the efficiency of BD-An triplet state generation in solution, going beyond the Marcus picture through first principles computational and theoretical methods, in order to explain how spin-crossover competes with charge recombination and fluorescence in solution. To this end, we interrogate each of the photophysical processes outlined in Fig. 1A by combining electronic structure calculations, molecular dynamics simulations and non-adiabatic rate theories.26,27 Our aim is to develop models that quantitatively predict experimental observables and give physical insight into mechanisms of triplet formation. We find that effects not captured by Marcus theory, including nuclear tunneling and zero-point energy, have a large effect on the non-adiabatic reaction rate constants, and must be accounted for in our description of these systems.27–30 Furthermore, Marcus theory relies on weak coupling between charge transfer states that does not hold for some of the important processes in BD-An, which we investigate with numerically exact open-system quantum dynamics calculations.31–34
The importance of solvent effects poses a particular challenge in developing a first principle understanding of triplet state formation, because this necessitates the use of explicit solvent models and molecular dynamics.27 However common general force fields for organic molecules are only applicable to describe the ground electronic state of these systems. Previous studies have primarily used gas phase electronic structure calculations to rationalize observed behavior,18,19 but these have not attempted to quantitatively predict rate constants from first principles. To address these challenges, we have developed a protocol for excited state force field parameterization, enabling us to accurately describe solvent fluctuations that control charge transfer processes in ground and excited states. With these tools, we show that the photophysics of BD-An can be quantitatively predicted and mechanisms of triplet formation can be understood in detail. We start by providing a brief description of the computational methods used in this study. We then show our results for predicted spectra, free energy changes and rate constants, followed by a discussion of how these can be used to understand efficient triplet formation in BD-An.
Fig. 2 (A) Absorption and (B) emission spectra of BD-An comparing calculated and experimental spectra with and without shifts in the excited state energies. The simulated line-shapes are obtained from the spin-boson mapping described in the main-text with bespoke force-fields for the excited states. The energy differences between excited states were obtained from DLPNO-STEOM-CCSD/def2-TZVP(-f) calculations combined with solvation energies from molecular dynamics. Experimental spectra obtained from ref. 18. |
In the absence of solvation effects, the SBD* state is lower in energy than the SCT state by about 0.5 eV (see ESI† for list of energies), which is inconsistent with the fluorescence spectrum, which shows a clear peak from the CT state at lower energies than the SBD* peak. Thus in order to predict solvation effects and spectral line-shapes, we constructed bespoke force-fields for the ground and excited states of BD-An, which enabled us to perform molecular dynamics simulations to efficiently compute spectra with the spin-boson mapping.45 Geometries and Hessians from TDA-TDDFT to were used to parameterize intramolecular force-fields46,47 based on the OPLS-AA force-field.48,49 Electronic polarizability was accounted for using the Drude oscillator model.50 We used the same procedure to parameterize both polarizable50 and non-polarizable force-fields for the acetonitrile (ACN) solvent, with further non-bonded parameter refinement targeting the dielectric properties of the solvent. The BD-An molecule was solvated in a box of 512 ACN molecules, and energy gap correlation functions were calculated from NVE trajectories, initial after NPT and NVT equilibration (full details are given in the ESI†).
From the molecular dynamics (MD) trajectories, the spin-boson mapping was constructed, from which spectra were then calculated.45,51,52 In this approach the full anharmonic potential energy surfaces VJ are mapped onto effective harmonic potential energy surfaces. Observables of this harmonic model are fully determined by the spectral density We fit the spectral distribution from the energy gap correlation function obtained from molecular dynamics,29
(2) |
(3) |
The absorption, AJ(ω), and fluorescence, FJ(ω), spectra (with unit area) are then given by
(4) |
(5) |
Further details of force-field development and the spin-boson mapping are provided in the ESI.†
The unshifted spectra calculated from the spin-boson mapping using DLPNO-STEOM-CCSD/def2-TZVP(-f) gas phase energy gaps are shown in Fig. 2 as dashed lines. Our calculated spectra show good overall agreement in the spectral line shapes, without any additional fitting, capturing the narrow SBD* peak in the absorption and fluorescence spectra, including a small vibronic side band at about 1500 cm−1 from the main peak, as well as the broad SCT fluorescence band. The agreement in the vibronic structure in the SBD* peaks suggests the fitted force fields capture the reorganization energies between excited states relatively well. However we see that the unshifted absorption spectrum calculations underestimates the SBD* energy, which we attribute to the fact that the triple zeta def2-TZVP(-f) basis set is likely still not sufficient for this system. As a result, we shifted all excited states by 805 cm−1 in order to fit the experimental absorption spectrum. This simple shift is justified by the fact that all excited states shift by ∼0.15 eV on increasing the basis set size from def2-SVP to def2-TZVP(-f), but differences between excited state energies change by much less (see ESI† for details). Furthermore it has been found the EOM-CCSD has typical errors of around 0.3 eV ≈ 2400 cm−1 for charge transfer states, so introducing a shift of 805 cm−1 seems justifiable. This shift is also used later in the free energy and rate calculations.
Using the shift from the absorption spectrum, the fluorescence spectrum (Fig. 2B) was calculated as a weighted sum of the SCT and SBD* emission spectra, with weights given by the transition dipole moments from DLPNO-STEOM-CCSD, μSBD*,S02 = 7.59 a.u. and μSCT,S02 = 0.54 a.u., and equilibrium populations of the two states given by the free energy change of charge separation ΔACS, i.e.
(6) |
The assumption of equilibrium between the SBD* and SCT states is justified by the fact the time-scale of equilibration of these states is ∼103 times shorter than the lifetime of these states (as we will discuss shortly). We have also computed the fluorescence spectrum assuming the populations of the SBD* and SCT states are given by the experimental estimate, ΔACS,exp, based on the approximate Weller equation, which is about 0.2 eV larger than our estimate.18 Because ΔACS,exp > 0, the SCT state is significantly less populated relative to the SBD* state and the SCT fluorescence peak is almost completely suppressed, which does not agree with the experimental spectrum. This suggests that the Weller equation cannot be used reliably when free energy changes are close to zero. As an interesting aside, the strongest SCT–Sn coupling (see Table 1) is to the S0 state, by over a factor of 10, which indicates that the intensity borrowing effect responsible for the SCT emission arises primarily from mixing between SCT and S0 states, rather than SCT and SBD* states, as has previous been assumed.24
A | B | Calc. ΔAA→B (eV) | Exp. ΔAA→Bb (eV) | λ (eV) | |HAB|d (cm−1) | k A→B (s−1) |
---|---|---|---|---|---|---|
a Free energy changes calculated with non-polarizable ACN, from thermodynamic integration/MBAR. b Estimated free energy changes from ref. 18 calculated with the Rehm–Weller equation ΔA ≈ ΔG = e(ED − EA) − ΔE* − e2/(4πε0εrrDA). c Reorganization energies from equating the pGaussianJ(ε = 0) with pJ(ε = 0) (see ESI for details). d Couplings averaged over gas-phase equilibrium geometries of A and B, |HAB|2 = (|HAB,A|2 + |HAB,B|2)/2. Details of calculations given in ESI. Ref. 18, estimated from spectroscopic measurements. e Rate constants from spin boson mapping. f Linear response value: λ = (〈ΔV〉B − 〈ΔV〉A)/2. g Radiative rate constant (eqn (9)). h With recrossing correction and. i Without recrossing correction. j Estimated from spectroscopic measurements.18 k Using reorganization energy from non-polarizable umbrella sampling calculations (see ESI). | ||||||
SBD* | SCT | −0.057 ± 0.005 | +0.13 | 0.550 ± 0.002 | 99 | (1.46 ± 0.04) × 1011 |
SBD* | S0 | −2.4542 ± 0.0004 | −2.460 | (8.77 ± 0.08) × 10−2f | — | (1.0747 ± 0.0006) × 108g |
SCT | S0 | −2.397 ± 0.003 | −2.59 | 0.483 ± 0.003 | 1904 | (3.4 ± 0.5) × 107h/(3.6 ± 0.6) × 107i |
SCT | TBD* | −0.826 ± 0.005 | −0.97 | 0.584 ± 0.001 | 0.79 | (7.9 ± 0.2) × 107 |
SCT | TAN* | −0.524 ± 0.004 | — | −0.477 ± 0.002 | 0.63 | (9.7 ± 0.1) × 107 |
SCT | TCT | −0.112 ± 0.001 | — | −0.119 ± 0.002 | 0.21 | (2.86 ± 0.02) × 107 |
TAN* | TBD* | −0.302 ± 0.003 | — | 0.565 ± 0.002 | 2.57 | (1.09 ± 0.01) × 109 |
TBD* | S0 | −1.638 ± 0.001a | −1.62j | 0.512 ± 0.002k | 0.19 | (1.045 ± 0.006) × 104k |
We have also calculated the rates of these processes from the same MD simulations, by calculating the probability of two states being at resonance. This probability controls the classical Fermi's Golden rule (FGR) rate for the transition between A and B.55 The free energy along the energy gap coordinate, ΔV = VB − VA, is related to the energy gap distribution pJ(ε) = 〈δ(ΔV − ε)〉J by27
AJ(ε) = −kBTln(pJ(ε)) + (AB − AJ) | (7) |
Fig. 3 (A–F) Free energy curves for the six A → B processes considered with the reaction A → B labeled on each figure. Points correspond to free energy curves calculated with MBAR and lines correspond to polynomials fitted to the MBAR cumulative distribution functions (see ESI† for details). (G) A snapshot for molecular dynamics simulations on the S0 potential energy surface. (H) A scheme highlighting the processes in (A–F). |
The free energy curves for charge separation and charge recombination are shown in Fig. 3A and B, where we see charge separation lies in the Marcus normal regime, whereas charge recombination is deep in the Marcus inverted regime, with a much larger free energy barrier. Using diabatic state couplings calculated from the generalized Mulliken–Hush method57 with DLPNO-STEOM-CCSD calculations, we can directly calculate the classical FGR rates for these processes (couplings |HAB| are shown in Table 1). The classical FGR charge separation rate is 4.8 × 1010 s−1, about a factor of 10 smaller than the experimentally observed rate of 5.4 × 1011 s−1, however the charge recombination rate is predicted to be 1.1 × 10−17 s−1, which is more than 1024 times too small compared to the experimental estimate of 2.3 × 107 s−1.18 This enormous discrepancy can be attributed to nuclear quantum effects, in particular the important role of nuclear tunneling in the inverted regime.
(8) |
The calculated spectral distributions ρJ(ω) can be decomposed into inner sphere, outer sphere and cross-correlated contributions, by decomposing the energy gap into molecular and the remaining environment contributions ΔV = ΔVmol + ΔVenv. We find that the cross-correlated contribution is generally negligible for all processes in BD-An, so the reorganization energy is well-described by a simple sum of inner and outer sphere contributions. The inner and outer sphere spectral distributions are calculated with the non-polarizable ACN/solute model, with the outer sphere contribution scaled down to match the polarizable model outer sphere contributions. As can be seen in Fig. 4A, the low frequency proportion of the spectral distribution for the SCT → S0 transition is dominated by the outer sphere contribution arising from solvent molecule fluctuations, making up ∼50% of the reorganization energy, which is well approximated by the Debye model.32 In contrast, the high frequency region of the spectral density is dominated by the inner sphere contribution from changes in equilibrium bond lengths in the BD-An molecule on charge transfer. The inner sphere spectral distribution has contributions over a range of frequencies from around 500 to 1600 cm−1, all of which contribute to tunneling enhancement of the SCT → S0 rate, although the dominant mode at ∼1400 cm−1 likely corresponds to a CC stretching motion within the aromatic rings. Qualitatively similar spectral distributions were found for the other charge transfer processes. For processes which do not involve charge transfer the spectral distribution is dominated by the inner sphere contribution, as can be seen for the TAN* → TBD* process in Fig. 4B.
When accounting for nuclear quantum effects, the SBD* → SCT rate goes up by a factor of ∼3 to 1.46 × 1011 s−1, and the SCT → S0 rate goes up by over 1024 to 1.0 × 108 s−1, and both calculated rates are now much closer to the experimentally measured values, agreeing much better with the experimental value. Application of Marcus–Levich–Jortner theory with the same inner and outer sphere reorganization energies and a characteristic inner-sphere frequency of 1500 cm−1 also predicts about a 1024-fold increase in the rate constant, compared to Marcus theory. This suggests that the large increase is robust to the details of the spectral density. Electronic polarizability is essential to account for in calculating the charge recombination rate. When a non-polarizable model is used instead, the free energy change of the reaction is effectively unchanged but the reorganization energy goes up by nearly 0.1 eV. This lowers the activation energy and accelerates the rate by around a factor of three.
Care should however be taken when using FGR to calculated the charge recombination rate. This is because the diabatic coupling for charge recombination process, HAB = 1904 cm−1, is about 20 times larger than kBT, and thus higher-order diabatic coupling effects beyond FGR, may be important (although large nuclear quantum effects in the FGR rate have been observed to reduce the importance of higher order effects).34 The large difference in couplings arises from the BD π orbitals involved in the transitions. The SBD* → SCT coupling involves an interaction between πAn and πBD (Fig. 5A) orbitals, whereas SCT → S0 coupling involves the πAn and (Fig. 5B) orbitals. As can be seen in Fig. 5 the πBD has minimal density on the carbon atom bonded to the An, group, whereas the orbital does. In order to investigate the potential role of higher-order diabatic coupling effects in the SCT → S0 transition, we have performed Hierarchical Equations of Motion (HEOM) calculations a simple model for this transition. The spectral density for the transition is coarse-grained down to a low-frequency outer-sphere portion described with a Debye spectral density and the inner sphere portion is described with a single under-damped Brownian oscillator spectral density, with a characteristic frequency of 1400 cm−1. The coarse-grained spectral density is shown in Fig. 6A. For this simplified model the exact open quantum system dynamics can be obtained using the HEOM method, and from this the rate constant as a function of HAB can be obtained. These rates are shown in Fig. 6B. We see that the rate constant is still fortuitously very well described by Fermi's Golden rule for this model, with only a factor of ∼0.9 reduction in the rate constant at the calculated value of HAB. We include this as a correction to the Fermi's Golden rule kSCS0 that we calculated with the full spectral density.
Fig. 5 BD orbitals involved in charge separation and recombination (A) πBD and (B) calculated with ωB97X-D3/def2-TZVPP/CPCM(ACN) at the S0 equilibrium geometry. |
Fig. 6 (A) The coarse-grained model spectral distribution for the SCT → S0 transition, consisting of a low frequency Debye contribution ρD(ω) = (1/2π)/(1 + (ω/ωD)2), with βωD = 0.1831, and an under-damped Brownian oscillator contribution ρBO(ω) = (1/2π)γΩ2/((ω2 − Ω2)2 + γ2ω2) with βγ = 4 and βΩ = 6.76. The reorganization energy for the Brownian oscillator portion is βλ = 8.6780 and for the Debye portion is βλ = 10.1459. (B) The rate constant from HEOM calculations for the coarse-grained spectral density as a function of HAB together with the FGR predictions. The value of HAB for the SCT → S0 transition is also indicated. Calculations were performed using the heom-lab code58 using the HEOM truncation scheme from ref. 59. |
Radiative recombination from the SCT state can also occur in BD-An, either through thermally activated delayed fluorescence via the SBD* state, or directly. The radiative rates can be calculated from the fluorescence spectra obtained from the spin-boson mapping as26,60
(9) |
We have also calculated the SOC couplings between the different SCT and triplet states using TDDFT (ωB97X-D3/def2-TZVPP/CPCM(ACN)) and the spin–orbit mean-field (SOMF) treatment of spin–orbit coupling.61,62 The two spin–orbit coupled charge transfer (SOCT) pathways have the largest SOC couplings, at 0.79 cm−1 and 0.63 cm−1 for the TBD* and TAN* whilst the formally El-Sayed's rule forbidden pathway has a smaller coupling at 0.21 cm−1. Using these couplings and the spin-boson mapping, we find that two El-Sayed's rule allowed transitions, viaTAN* and TBD*, occur at very similar rates, with SCT → TBD* occurring only about 20% faster than the SCT → TAN* formation. The triplet–triplet TAN* → TBD* energy transfer is also activation-less (Fig. 3C), and has a coupling from fragment energy/charge density (FED/FCD) calculations47,63,64 of 2.57 cm−1, and so occurs about 10 times faster than the triplet formation rate, accelerated by a factor of 1.6 by nuclear quantum effects, so the steady state population of TAN* would be difficult to observe spectroscopically at room temperature. The El-Sayed's rule forbidden transition to the TCT state also contributes to triplet formation, although it occurs about 4.5 times slower than TBD* formation. The TCT state very rapidly recombines to the TAN* or TBD* states, with these spin allowed transitions occurring at least ∼104 times faster than the corresponding spin-forbidden transitions, so the TCT state would be very difficult to observe directly at room temperature. Overall the TCT TAN*, and TBD* pathways contribute 14%, 47%, and 39% respectively to the overall triplet formation. Surprisingly the most significant pathway is the TAN* pathway and not the direct TBD* pathway, which can be rationalized by the lower activation barrier for the TAN* spin–orbit coupled charge recombination. The observation is consistent with TREPR experiments in which all three triplet states were observed, although at much lower temperatures (80 K) in a very different medium (a dichloromethane/isopropanol solid matrix). This work shows that multiple triplet formation pathways, including those forbidden by El-Sayed's rule, can contribute at room temperature in polar solvents. The presence of multiple triplet recombination pathways may also explain the large spread of effective spin–orbit coupled charge transfer rates observed in the family of BD-Aryl molecules studied in ref. 18.
Using all of the computed rates, we have estimated the observed charge separation and charge recombination rates, as well as the triplet yield. The effective charge separation rate corresponds to the observed equilibration rate between SBD* and SCT states i.e. kCS,eff = kSBD*→SCT + kSCT→SBD*. Likewise the effective charge recombination rate corresponds to the observed decay rate of the SCT state, which under a pre-equilirbium approximation for the SBD* ⇌ SCT interconversion is given by
kCR,eff = pSCT(kCR + kF,SCT→S0) + pSBD*kF,SBD*→S0 | (10) |
kCR = kSCT→S0 + kSCT→TCT + kSCTAN* + kSCTT→BD*. | (11) |
The triplet quantum yield ΦT is calculated as ΦT = pSCT(kSCT→TCT + kSCT→TAN* + kSCT→TBD*)τCR, with τCR = 1/kCR,eff, and the fluorescence yield ΦF is ΦF = pSBD*kF,SBD*τCR. We also computed the fraction of non-radiative transitions which produce a triplet state, ϕCRT = ΦT/(1 − ΦF), as measured in ref. 18.
The calculated and experimental values of the rates and yields are summarized in Table 2. Overall we see excellent agreement between the calculated rates/yields and the experimental measurements from ref. 18 and 24, with less than a factor of 4 error in the charge separation rate and only a factor of ∼1.6 error in the charge recombination rate. Similar we only slightly underestimate the triplet yield, with our calculations yielding 0.80, compared to the experimental measurements between 0.93 and 0.96. If we only included the dominant SCT → TBD* triplet formation pathway, the triplet quantum yield would only be ∼0.6, and the error in the rate would be over a factor of 3. We also find that suppression of the charge recombination also plays a large role in efficient triplet formation, which is facilitated by polarizability and recrossing effects. Without including electronic polarizability, the charge recombination rate would be enhanced to ∼1.0 × 108 s−1, which would reduce the triplet quantum yield to ∼0.63. This corroborates the conclusions drawn in ref. 18, although we find that multiple triplet pathways also enable the triplet formation to compete with charge recombination, which is suppressed by several effects. The net fluorescence quantum yield from SBD* that we calculate, 0.045, is also in good agreement with the experimental values, between 0.01 and 0.018. These results suggest that the intersystem crossing rates are being slightly underestimated by our models, possibly due to errors in the reorganization energies or the spin–orbit couplings obtained from TD-DFT, which are all less than 1 cm−1.
The triplet lifetime τT = 1/kTBD*→S0 plays an important role in determining the utility of a triplet sensitizer or photocatalyst, with longer-lived triplet states allowing more time for diffusive encounters with other molecules enabling more efficient energy transfer. We have also calculated the triplet lifetime for BD-Anusing the methods described above, and we also find good agreement between our calculated value for τT and experimental measurements (Table 2), with an error of only ∼20%. From a simulation perspective, this requires an accurate calculation of the free-energy barrier, which requires enhanced sampling since the transition is very deep in the Marcus inverted regime, since it displays a highly non-quadratic free energy curve. This was achieved using the non-polarizable model with umbrella sampling65 on the energy gap coordinate ΔV sampled with the Fast-Forward Langevin algorithm.66 Use of the non-polarizable model is justified because over 99% of the reorganization energy is inner sphere for both ACN models, and solvent polarizability has less than a 1 meV effect on the free energy of the TBD* state. As with the spin-conserving charge recombination, because the transition is deep in the inverted regime and the spectral distribution is dominated by high frequency inner sphere contributions, there is a very large nuclear quantum effect of over 107 in the rate constant. One significant source of uncertainty in this is the validity of the spin-boson mapping, where rates calculated from the spectral distribution obtained from TBD* and S0 dynamics vary by about 50%. This means that methods that more rigorously account for asymmetry and anharmonicity in the potential energy surfaces, while also accounting for nuclear quantum effects, may be needed to more accurately compute triplet lifetimes for this system and other related systems.55,67,68 However given the simplicity of the spin-boson mapping and its accuracy in this case, it is clearly still useful in prediction of non-adiabatic rates.
The simulation techniques and bespoke force-field parametrization approach developed here paves the way for a quantitative modeling of other triplet photosensitizers and related systems,69 possibly even enabling straightforward computational screening for properties such as the triplet lifetime. Comparison between simulated and experimental optical spectra indicates that a major source of error is in gas phase energies of excited states. Even the popular wave-function-based DLPNO-STEOM-CCSD method appears to significantly underestimate transition energies, although the ground-state DLPNO-CCSD(T) method which can be used to calculate the T1–S0 gap seems robust. We also note that whilst the approximate spin-boson mapping seems fairly reliable for these systems, its application to deep inverted regime processes requires scrutiny. Thus BD-Ancould provide an interesting test-bed for recently developed approaches to calculating non-adiabatic transition rates applicable to high-dimensional anharmonic systems.34,55,67,68,70–79 The SCT → S0 transition poses a particular challenge, since it is deep in the inverted regime, nuclear quantum effects are very large and strong diabatic coupling means there may be some effects missed by FGR, which we have estimated using open quantum dynamics simulations. Furthermore in this study we have neglected non-Condon effects80 and potential spin-vibronic effects,81 which could also play a role in determine the rates of conversion between excited states in this system. Future investigations into these potential effects could provide further insight into triplet formation in BD-An.
Overall, we believe the mechanistic insights gained from this study, which would be difficult to probe directly with experiment alone, could help light the path towards the development of novel and interesting photochemistry in related systems. The observation that high-energy triplet pathways dominate at room temperature opens the door to the intriguing possibility of engineering triplet anti-Kasha's rule systems,82 in which higher energy triplet states could be used to drive photochemistry. This could be particularly promising since triplet-triplet energy transfer is strongly distance dependent,83 so spatial separation of chromophore units could be used to extend the lifetime of high-lying triplet states. In summary, our comprehensive study highlights the intricate balance of factors influencing triplet formation, including the significance of charge separation efficiency, multiple recombination pathways, and nuclear quantum effects. Moving forward, this mechanistic understanding could steer the development of novel photochemical systems, with a wide range of potential applications.
ACN | Acetonitrile |
An | Anthracene |
BD | BODIPY, boron dipyrromethane |
CPCM | Conductor-like polarizable continuum |
CR | Charge recombination |
CS | Charge separation |
CT | Charge transfer |
DLPNO-STEOM-CCSD | Domain local pair natural orbital similarity transformed equation of motion coupled cluster singles and doubles |
DLPNO-CCSD(T) | Domain local pair natural orbital coupled cluster singles and doubles with perturbative triples |
EOM-CCSD | Equation of motion coupled cluster singles and doubles |
FGR | Fermi's golden rule |
HEOM | Hierarchical equations of motion |
MBAR | Multi-state Bennett acceptance ratio |
MD | Molecular dynamics |
NPT | Constant particle number/pressure/temperature molecular dynamics |
NVE | Constant particle number/volume/energy molecular dynamics |
NVT | Constant particle number/volume/temperature molecular dynamics |
WHAM | Weighted histogram analysis |
SOCT | Spin–orbit coupled charge transfer |
TDA | Tamm-Dancoff approximation |
TDDFT | Time dependent density functional theory |
TREPR | Time resolved electron paramagnetic resonance |
Footnote |
† Electronic supplementary information (ESI) available: Details on electronic structure calculations and bespoke force-field parameterization. Details of rate constant calculations. MD simulation details. Supplemental tables of solvent model properties, gas phase state energies, and reorganization energies. See DOI: https://doi.org/10.1039/d4sc01369g |
This journal is © The Royal Society of Chemistry 2024 |