Summary
Singlet fission is known to improve solar energy utilization by circumventing the Shockley-Queisser limit. The two essential steps of singlet fission are the formation of a correlated triplet pair and its subsequent quantum decoherence. However, the mechanisms of the triplet pair formation and decoherence still remain elusive. Here we examined both essential steps in single crystalline hexacene and discovered remarkable anisotropy of the overall singlet fission rate along different crystal axes. Since the triplet pair formation emerges on the same timescale along both crystal axes, the quantum decoherence is likely responsible for the directional anisotropy. The distinct quantum decoherence rates are ascribed to the notable difference on their associated energy loss according to the Redfield quantum dissipation theory. Our hybrid experimental/theoretical framework will not only further our understanding of singlet fission, but also shed light on the systematic design of new materials for the third-generation solar cells.
Subject Areas: Spectroscopy, Theoretical Photophysics, Quantum Phenomena
Graphical Abstract
Highlights
-
•
Remarkable anisotropy of the overall singlet fission along different crystal axes
-
•
The correlated triplet pair emerges on the same timescale along both crystal axes
-
•
The quantum decoherence is predominantly driven by electron-phonon coupling
-
•
The anisotropic decoherence is due to the directional difference of its energy loss
Spectroscopy; Theoretical Photophysics; Quantum Phenomena
Introduction
Singlet fission has recently attracted extensive attention from both experimentalists (Busby et al., 2015, Chan et al., 2011, Chan et al., 2012, Congreve et al., 2013, Musser et al., 2015, Walker et al., 2013) and theoreticians (Berkelbach et al., 2013a, Berkelbach et al., 2013b, Berkelbach et al., 2014, Tiago et al., 2003, Yost et al., 2014, Zimmerman et al., 2010, Zirzlmeier et al., 2015), since it has the potential to circumvent the Shockley-Queisser limit for solar energy utilization. Many materials have exhibited singlet fission, and they range from small molecules to polymers, and from isolated molecules, thin films, to polycrystalline materials (Burdett et al., 2010, Busby et al., 2014, Busby et al., 2015, Cook et al., 2016, Johnson et al., 2010, Jundt et al., 1995, Katoh et al., 1997, Marciniak et al., 2009, Michl et al., 2007, Najafov et al., 2010, Pensack et al., 2016, Schwob and Williams, 1972, Takeda et al., 1996, Tayebjee et al., 2013, Walker et al., 2013, Watanabe et al., 2006, Wen et al., 2013, Wilson et al., 2011, Zenz et al., 1999). Singlet fission is a spin-conserving process, in which a photo-generated singlet exciton is converted into two individual triplet excitons (Smith and Michl, 2010). This unique photophysical phenomenon is usually considered to proceed in two steps, namely, the formation of a correlated triplet pair, 1(T1T1), and its subsequent quantum decoherence to engender a decoupled triplet dimer, T1+T1. A generic model often employed to describe the mechanism of singlet fission is given by (Merrifield et al., 1969):
(Equation 1) |
where S0 is a ground state, S1 is a singlet excited state, and kF and kQ denote the rate constants for the formation of 1(T1T1) and its quantum decoherence, respectively. Despite recent tremendous efforts, the mechanisms of these two constituent steps of singlet fission (Bakulin et al., 2016, Chan et al., 2011, Chan et al., 2012, Miyata et al., 2017, Musser et al., 2015, Stern et al., 2017) are still not well understood.
Previous studies were primarily focused on the first step (Berkelbach et al., 2013b, Chan et al., 2013, Monahan and Zhu, 2015, Schwerin et al., 2010, Smith and Michl, 2010, Yost et al., 2014, Zimmerman et al., 2010, Zimmerman et al., 2011) and have invoked three mechanisms so far to rationalize the formation of 1(T1T1). They are the direct mechanism (Zimmerman et al., 2010, Zimmerman et al., 2011, Zimmerman et al., 2013), the charge-transfer (CT)-mediated mechanism (Beljonne et al., 2013, Busby et al., 2015, Chan et al., 2013, Monahan and Zhu, 2015, Smith and Michl, 2010), and the sequential mechanism (Berkelbach et al., 2013b, Nakano et al., 2016). On the other hand, the quantum decoherence of 1(T1T1) is rarely explored, particularly from the theoretical perspective (Casanova, 2018). As a matter of fact, quantum decoherence plays a key role in singlet fission by decoupling 1(T1T1) through vibronic coupling. Therefore, a rapid decoherence of correlated 1(T1T1) is highly desired for an efficient production of decoupled T1+T1, the final product of singlet fission. Otherwise, the reverse of the singlet fission could be mediated by fusion of two triplets via triplet-triplet annihilation (Yong et al., 2017), thereby decreasing the photon-to-exciton conversion efficiency. The fundamental driving force for the quantum decoherence of 1(T1T1) is its interaction with its dissipative environment, which is typically described by a thermalized phonon path at finite temperature (Tao, 2014). This system-environment interaction varies greatly from amorphous, or polycrystalline, to ordered crystalline structures. Singlet fission materials in the form of single crystal are anticipated to exhibit pronounced collective characteristics in molecular vibrations because of their ordered crystalline structures.
In this work, we investigated the quantum decoherence of 1(T1T1) states in single crystalline hexacene by polarized transient absorption spectroscopy in conjunction with ab initio quantum mechanics simulations. Our results showed that both 1(T1T1) and T1+T1 states are of anisotropic origin owing to triplet Davydov splitting in hexacene, leading to distinctive quantum decoherence rates along crystal a and b axes. This notable anisotropy can be principally ascribed to the difference on the energy gap between 1(T1T1) and T1+T1 according to the Redfield quantum dissipation theory (Ishizaki and Fleming, 2009). Moreover, through our functional mode analysis, the anisotropic quantum decoherence is predominantly driven by some slow phonon modes with an effective frequency ωp < 50 cm−1.
Results
Singlet Davydov Splitting in Hexacene
Hexacene single crystals used in our experiments were grown with physical vapor transport by following a recently developed method (Laudise et al., 1998, Watanabe et al., 2012). Their typical lateral size was on the order of 20 μm × 40 μm (Figure 1A), whereas their thickness was estimated to be 0.5 μm as detailed in Supplemental Information. The triclinic unit cell of hexacene shown in Figure 1B is defined by experimental in-plane crystal lattice constants (Watanabe et al., 2012) of a = 7.673 Å (long axis) and b = 6.292 Å (short axis) (Yamagata et al., 2011). A home-built microscopy was coupled with a tungsten-halogen light source for polarization-dependent optical transmission measurements of hexacene single crystal (Figure S1), and its polarization-resolved linear absorption spectra for the ab plane is presented on Figure 1C. Interestingly, the polarized measurements exhibit different absorption features along these two lattice axes. The lowest resonant peak at 835 nm was assigned to be the transition along the short b axis and is called b1. On the other hand, the lowest transition at 761 nm was attributed to the transition along the long a axis and is named a1. Apparently, b1 is redshifted by 144 meV (1,160 cm−1) with respect to a1. This notable energy difference is due to the splitting of electronic excitations, the so-called singlet Davydov splitting (Davydov, 1948, Davydov, 2013) that stems from dipole-dipole interactions of two non-equivalent molecules in a hexacene unit cell (Chernikov et al., 2014, Hestand et al., 2015).
Calculated Anisotropic Energy Levels in Hexacene Single Crystals
To explore the energy levels of all singlet-fission participating states in single crystalline hexacene, we performed computer simulations under periodic boundary conditions. Specifically, the single crystalline hexacene was modeled as a 9 × 9 × 3 supercell with a total of 20,412 atoms (Figure 2A). For a compromised balance between numerical efficiency and physical accuracy, the hybrid quantum mechanics/molecular mechanics approach (Warshel and Levitt, 1976) was adopted. A detailed description of the computational calculations can be found in Supplemental Information. As shown in Figure 2B, we denote the ground state, singlet excited state, charge-transfer state, correlated triplet pair, and decoupled triplet dimer as 1(S0S0), 1(S1S0), CT, 1(T1T1), and T1 + T1, respectively, and their constituent spin configurations were illustrated in our previous studies (Elenewski et al., 2017a, Elenewski et al., 2017b). Geometry optimization was carried out on all spin states before their relative energies are determined. The only exception is 1(S1S0)*, which assumes the same geometry as 1(S0S0), such that reflects the vertical optical gap of hexacene crystal. In particular, the linear-response time-dependent density functional theory(Casida and Huix-Rotllant, 2012) was utilized to delineate the 1(S1S0) and 1(S1S0)* states, whereas the constrained density functional theory (CDFT) (Kaduk et al., 2012) was used to construct the reference orbitals needed for the multi-configurational CT, 1(T1T1), and T1 + T1 states. As for the T1 + T1 dimer, a correction term of −kbTln(3 × 3) ≈ −56 meV was added to account for the electronic spin entropy. Interestingly, we found that the 1(T1T1) and T1 + T1 states along the short b axis are more thermodynamically stable than their counterparts along the long a axis, suggesting possible directional heterogeneity of quantum decoherence.
Spectroscopic Signatures of Triplet Davydov Splitting
Transient absorption spectroscopy is a proven tool to investigate singlet fission in solution and in thin film. We have designed and constructed a polarization-resolved transient absorption microscopy to examine anisotropic singlet fission in hexacene single crystals. A detailed description of our experimental setup is presented in Figure S1. In short, a regenerative amplifier Ti:Sapphire laser system operating at 795 nm and 1 kHz repetition rate was used for our measurements. A small portion of the laser pulse was taken as a pump light, and its fluence on samples was kept below 200 μJ/cm2 to avoid exciton-exciton annihilation (Figure S3) by using a half-wave plate and a polarizer. Another small part of the laser pulse with a pulse energy of ca. 1 μJ was focused onto a sapphire crystal to generate white light supercontinuum as a probe. The pump and probe beams were combined to become collinear with a pellicle beam splitter. They were then focused onto samples through a 10× microscope objective. The polarizations of the pump and probe beams were independently adjusted with two individual half-wave plates.
To explore anisotropic spectral kinetics in hexacene single crystals, polarization-resolved transient absorption measurements were performed along both axes using a pump along b axis under the near lowest photo-excitation of 795 nm. Their results were presented as pseudo-color plots of transient absorption spectra (ΔT/T) along b axis (Figures 3A and 3B) and a axis (Figures 3C and 3D). We chose two different color scales to highlight the difference on relaxation between the perturbed ground state and the newly populated excited states. The strong positive changes in ΔT/T between 729 and 620 nm were primarily due to the ground state bleaching and recovery. As expected, these positive responses ΔT/T in spectra are similar to those of the linear absorption spectra in hexacene as shown in Figure 1C. The negative spectral changes in ΔT/T between 496 and 620 nm were attributed to the photo-induced absorption of the excited states (Figures 3E and 3F). The transient spectra along both axes have broad, featureless, and short-lived bands from 496 to 577 nm (Figures 3E and 3F). The fast process dominated in this spectral region can be ascribed to the excited state absorption S1 to a higher singlet state Sn (S1→ Sn). By contrast, long-lived peaks appear in the range between 590 and 620 nm, as well as in the range between 508 and 544 nm. The latter even overlaps with the short-lived bands. We attributed these long-lived peaks to triplet transitions in hexacene. Again, a strong anisotropicity was observed on the triplet transitions. The peak along b axis is at 599 nm, whereas the one along a axis is redshifted by 9 nm. Since the redshift is an indicator of triplet Davydov splitting, we expect a coexistence of two kinds of triplet excitons with a splitting energy of 30 meV. Note that the correlated triplet pair, 1(T1T1), is a dark state in our case and did not show appreciable spectral feature in the transient spectra. Therefore, the 9-nm redshift must be induced by the energy difference between the decoupled triplet dimers, i.e., T1a + T1a and T1b + T1b. Interestingly, T1b + T1b was found to be energetically more stable than T1a + T1a probably due to the stronger Davydov splitting between two spatially closer triplet excitons. Nevertheless, the long-lived spectral features between 500 and 620 nm were assigned to the T1 → Tn transitions along both axes (Busby et al., 2014, Lee et al., 2013).
Moreover, we measured the anisotropic transient spectra using pumps along both axes to further verify the independence of triplet Davydov splitting on pump polarization. The independence was confirmed by Figure S4, which does not manifest appreciable change when the pump polarization is varied. Furthermore, we changed the excitation wavelengths to 400 nm and again observed distinctive triplet Davydov splitting (Figure S12). Therefore, the anisotropy of the triplet states and their derivatives is an intrinsic property of crystalline hexacene regardless of external electromagnetic perturbations.
Singlet Fission Kinetics along the Long and Short Crystal Axes
Figure 4 shows kinetic traces of hexacene in terms of ΔT/T after pumps are applied along b axis with the following wavelengths: (1) 525 ± 18.0 nm for S1→Sn; (2) 599 ± 8.7 nm for T1b →Tnb; (3) 608 ± 9.0 nm for T1a →Tna. A kinetic model (Johnson et al., 2010, Monahan et al., 2017, Wan et al., 2018) was used to derive the generation and dissociation rates of 1(T1T1) state from the experimental data shown in Figure 4. The detailed fitting procedure can be found in Supplemental Information. A global fitting of the kinetic traces affords the formation (kF) and decoherence (kQ) rates along b axis as kF, bb = 4.5 ± 0.2 ps−1 and kQ, bb = 6.7 ± 0.2 ps−1, respectively. Similarly, their counterparts along a axis are given by kF,aa = 5.0 ± 0.2 ps−1 and kQ, aa = 3.3 ± 0.2 ps−1. Apparently, kF remains nearly unchanged, whereas kQ has more than doubled after the probe polarization rotates from a axis to b axis. Once again, our experimental data suggest that the mechanism of singlet fission is of anisotropic nature in single crystalline hexacene.
Discussion
Our studies of single crystalline hexacene have three main traits: (1) a large triplet Davydov splitting of 30 meV; (2) a homogeneous 1(T1T1) formation rate regardless of molecular directionality; (3) a notable anisotropy of 1(T1T1) quantum decoherence rate. We have also proposed a unified vibronic model for singlet fission by treating the formation of a correlated triplet pair and its quantum decoherence on the same footing using our non-adiabatic functional mode approach (Chen, 2014, Elenewski et al., 2017b). As clearly depicted in Scheme 1, the optical excitation results in anisotropic structural and dynamical properties for both the 1(T1T1) and T1 + T1 states, allowing us to gain more insights into the intrinsic directional heterogeneity of triplet excitons in single crystalline hexacene.
Formation of 1(T1T1) Directly from 1(S1S0)∗
The formation of 1(T1T1) is generally considered as the rate-determining step of singlet fission under the assumption of ultrafast quantum decoherence of 1(T1T1). Surprisingly, our results show that the formation of 1(T1T1) outpaces its quantum decoherence in hexacene, making the overall singlet fission rate predominantly decided by the latter. More interestingly, the formation rates of 1(T1T1) were found to be similar along b axis (kF, bb = 4.5 ± 0.2 ps−1) and a axis (kF,aa = 5.0 ± 0.2 ps−1). As previously discussed, the formation of 1(T1T1) could proceed through the direct, the CT-mediated, and the sequential mechanisms. To reconcile which mechanism is responsible for the fast and isotropic formation of 1(T1T1) in hexacene, we employed our functional mode vibronic theory that has been successfully applied to investigate singlet fission in tetracene and pentacene (Elenewski et al., 2017a, Elenewski et al., 2017b). If the direct mechanism is followed, 1(T1T1) is formed without the aid of any CT intermediate. By contrast, the sequential mechanism demands a thermodynamically stable CT state with an energy lower than the hot 1(S1S0)∗. Similarly, the mediated mechanism entails a virtual CT state, whose energy should be close to 1(S1S0)∗ for an ideal vibrational wave-function mixing between them. Since our calculated energy diagram in Figure 2 manifests a relatively large energy gap of over 0.25 eV between the CT and 1(S1S0)∗ states, both the sequential and mediated mechanisms should be effectively blocked. In this regard, we assume that the 1(T1T1) pair production in single crystalline hexacene only proceeds through the direct mechanism (Elenewski et al., 2016, Elenewski et al., 2017b):
(Equation 2) |
where J is the electronic strength between 1(S1S0)∗ and 1(T1T1), is the driving force of singlet fission, λ is the associated reorganization energy, Ω is the angular frequency of the effective singlet-fission driving vibrational mode, Γ is its thermal relaxation rate, and is the optically populated initial vibrational quanta where Einc is the incident light energy, which is 795 nm in the present study. We can then apply Equation 2 to calculate the 1(T1T1) formation rate kF under non-thermalized condition. It was found that kF,aa of 0.93 ps−1 is nearly the same as kF,bb of 0.83 ps−1, in spite of a modest disparity in energetic parameters such as ΔG0 and λ. The diminished difference on kS along a and b axes primarily arises from their fast Ωs, whose zero-point energies become much more overwhelming for vibrational quantum tunneling. As suggested by Busby et al., 2014, almost all singlet excitons in hexacene crystal undergo singlet fission, resulting in a triplet quantum yield of nearly 200% that also affords a negligible fluorescence quantum yield. If so, the overall decay rate, kD, of 1(S1S0)∗ is simply kF + Γ, leading to kD,aa of 1.86 ps−1 and kD,bb of 2.08 ps−1. These computed rates are in good agreement with those obtained from our experimental measurements, indicating a direct formation of 1(T1T1) from the vibrationally “hot” 1(S1S0)∗ in hexacene. In fact, previous studies of hexacene thin films attributed a slower singlet fission to multi-phonon relaxations, as compared with pentacene and tetracene (Busby et al., 2014). For example, ultrafast vibronic spectroscopies of pentacene and its derivatives showed that the singlet fission process is a single-molecular internal conversion via a conical intersection due to strong coupling between nuclear and electronic motions (Musser et al., 2015). As a result, the thermal relaxation of the 1(S1S0)∗ state should be taken into account when its decay rate is probed by transient spectroscopy techniques.
Anisotropic Nature of T1 + T1 and 1(T1T1)
Triplet Davydov splitting exhibits a difference in energy as high as 30 meV along the a and b axes in hexacene single crystals, indicating that the T1 + T1 states remain strong dipole-dipole interactions in the anisotropic environment. The energy level of T1b + T1b is lower than that of T1a + T1a in the case of hexacene. This energy alignment is opposite to that of pentacene (Figure S5) and tetracene (Schwoerer and Birech, 2014), both of which have T1b + T1b higher than T1a + T1a as reported. Since the anisotropic lowest spin-correlated triplet T1 + T1 states arises from the decoherence of 1(T1T1), it is legitimate to surmise that the 1(T1T1) states should be anisotropic in hexacene single crystals. We further hypothesized that the energy level for 1(T1bT1b) is lower than that for 1(T1aT1a) in hexacene single crystals from the measured triplet Davydov splitting.
To reveal the nature of the anisotropic 1(T1T1) and T1 + T1 states, we need to address the anti-ferromagnetic electron correlation in acenes as revealed by a complete active space simulation using the density matrix renormalization group algorithm (Hachmann et al., 2007). This anti-ferromagnetism was later corroborated by a spin-polarized density functional theory study (Jiang and Dai, 2008), which discovered spatially separated magnetizations in acenes. Similarly, electron spins in hexacene favor anti-ferromagnetic correlation, resulting in the notable presence of diradical covalence resonance structures. For example, a pair of adjacent hexacene molecules along a axis had been extracted from the crystal structure before a complete active space self-consistent field calculation (Aquilante et al., 2016) with an active space of CAS(6,6) was conducted to reveal an enthalpy difference, , of 16.7 meV. In fact, a similar value of 13.5 meV was also achieved by our employed CDFT method that had been demonstrated to achieve a comparable accuracy against the multireference configuration interaction MRCI + Q level of theory on polyacenes (Kubas et al., 2014). Since a T1 + T1 dimer has less anti-ferromagnetic characteristics than its parental 1(T1T1) pair in spite of more available spin microstates, its relative energetic stability is determined by a direct competition between enthalpy and entropy, i.e., EHS→LS versus kBTln9. It turns out that entropy is a more dominant factor than enthalpy for crystalline hexacene at room temperature, making T1 + T1 thermodynamically more stable than 1(T1T1) as shown in Figure 2. By the same token, the 1(T1aT1a) pair is expected to have a higher energy than its counterpart along axis b owing to diminished anti-ferromagnetic coupling strength over distance. Nevertheless, the difference on the spin-flip energy between the two axes is not strong enough to override their gap on the 1(T1T1) pair, thus sustaining the relative stability of T1b + T1b over T1a + T1a.
In polycrystalline singlet fission materials, the anisotropy of the 1(T1T1) states may preserve within individual crystalline grains. The random orientation of those grain boundaries, however, is bound to suppress the global expression of the anisotropy. Very uniquely, our studies in the form of single crystals allow us to unravel the anisotropic nature of the 1(T1T1) state without ambiguity. Previous studies of polycrystalline acenes and other singlet fission materials showed that the energy level of a triplet pair 1(T1T1) is higher than that of a decoupled triplet dimer(Basel et al., 2019, Yong et al., 2017). Recently, a triplet pair 1(T1T1) was reported to be bound relative to free triplets with an energy of ca. 30 meV, which is largely material independent (Yong et al., 2017). Our results not only demonstrated a higher energy of 1(T1T1) than T1 + T1, but also exhibit their anisotropic features in hexacene.
Anisotropic Quantum Decoherence of 1(T1T1)
The transition dipole strength for the triplet pair 1(T1T1) state is generally too weak to be detected so that its quantum decoherence is usually assumed to be much faster than its formation (Berkelbach et al., 2014, Chan et al., 2013, Coto et al., 2015, Mirjani et al., 2014, Tamura et al., 2015, Yost et al., 2014). Nevertheless, some studies implied that the quantum decoherence rate of the 1(T1T1) state could be as slow as its formation rate (Feng and Krylov, 2016, Kolomeisky et al., 2014, Monahan et al., 2017, Pensack et al., 2016). In the present study, we found that kQ is even slower than kF along the long a axis, whereas kQ along the short b axis is comparable with kF. Our results indicate that the relatively slow quantum decoherence is as important as triplet pair formation for the overall singlet fission rate in hexacene. According to the Redfield quantum dissipation theory (Ishizaki and Fleming, 2009), the quantum decoherence rate, kQ, due to system-bath coupling is given by:
(Equation 3) |
where ΔE is the energy gap between the initial and final states of quantum decoherence, λ is the corresponding reorganization energy, and ωp is the system-bath coupling frequency. If the quantum decoherence of 1(T1T1) pair in hexacene crystal is assumed to be driven by vibronic coupling only, ωp can be ascertained by the functional mode analysis (Hub and de Groot, 2009), which projects the diabatic energy gaps between 1(T1T1) and T1 + T1 onto the dimer's vibrational normal modes for 5,000 snapshots extracted from our MD trajectory. It was found that the quantum decoherence is predominantly driven by a few slow vibrational modes with ωp < 50 cm−1 (Figure S11). Thus, the quantum decoherence is mediated by the electron-phonon interaction in crystalline hexacene. Even if we account for the small contributions from other vibrational modes that are usually much faster, the calculated effective driving mode, , is still as slow as ∼40 cm−1 (Table S3). Therefore, ωp is an important factor for the quantum decoherence of 1(T1T1) pair in hexacene because of its notable mismatch with ΔE, i.e., ћωp≪ΔE→kQ∝ωpλΔE−1. Interestingly, the along a axis is nearly two times as large as its counterpart along b axis, resulting in a remarkable directional heterogeneity of kD, i.e., . Besides the experimentally consistent heterogeneity factor, our calculated values of kQ along the two crystal axes are also well in line with the results from our ultrafast spectroscopy studies (Table S2).
Previous studies of singlet fission in pentacene and hexacene were performed on their polycrystalline films (Bakulin et al., 2016, Busby et al., 2014, Chan et al., 2011, Kandada et al., 2014, Rao et al., 2011, Wilson et al., 2011, Wilson et al., 2013). Fission rates of the films are, in general, the averaged result over different orientations of the polycrystalline structures, leading to a loss of intrinsically anisotropic features in these materials. Transient absorption microscopy experiments found that long-lived correlated triplet pairs exist and triplet interaction and binding could be induced by a π-staked geometry in crystalline pentacene derivative (Folie et al., 2018). Polarization measurements of tetracene demonstrated that there was no difference in kinetics along the long and short axes (Schwoerer and Birech, 2014, Wan et al., 2015, Zhang et al., 2014). By contrast, our results showed that the anisotropic quantum decoherence of 1(T1T1) occurs in single crystalline hexacene. This decoherence process is driven by the vibronic coupling between the 1(T1T1) state and some low-frequency phonon modes of hexacene crystal. Owing to the directional asymmetry of the crystal, this quantum decoherence process might be intrinsically anisotropic as governed by the energy gap between 1(T1T1) and T1 + T1.
In summary, we discovered anisotropic singlet fission in single crystalline hexacene and have revealed its mechanism by deciphering its two constituent steps. For the first step that gives rise to correlated triplet pairs through the so-called direct mechanism, no anisotropy was found as they were formed on the same timescale of ∼200 fs along the two crystal axes. In the second step that involves quantum decoherence of the correlated triplet pairs, the rate along the short b axis is nearly two times as fast as that along the long a axis. The difference on the rate might be ascribed to the directional heterogeneity of the hexacene crystal that results in distinct thermodynamic stabilities between the initial and final states of the quantum decoherence, which is driven by system-bath coupling with a slow effective frequency of ωp < 50 cm−1. Our findings not only corroborated the importance of quantum decoherence in fission process, but will also guide future experimental and theoretical endeavors of exploring anisotropic photophysical properties in novel materials for multiple exciton generation.
Limitations of the Study
From a fundamental point of view, one of the most challenging tasks for the singlet fission community is the characterization and monitoring of the optically dark 1(T1T1) state due to its small absorption cross section and short lifetime upon photo-excitation. Although the transient absorption spectra of the 1(T1T1) state were presented in recent experiments (Burdett and Bardeen, 2012, Korovina et al., 2016, Lukman et al., 2015, Lukman et al., 2016, Stern et al., 2015, Stern et al., 2017), those results are limited to intramolecular singlet fission or intermolecular one for dimers in solution (Rao and Friend, 2017) with unusually long 1(T1T1) lifetime. Therefore, it is not surprising that the optical identification of the short-lived 1(T1T1) state in hexacene has not been reported yet with the transient absorption spectroscopy method. More recently, the two-dimensional optical spectroscopy was employed to demonstrate the existence of 1(T1T1) state in thin film pentacene and its derivatives (Bakulin et al., 2016). Nevertheless, this technique has to be coupled with a microscopy approach for small-size samples such as single crystalline hexacene, our system of interest.
Methods
All methods can be found in the accompanying Transparent Methods supplemental file.
Acknowledgments
This work was supported by the Honda Research Institute (Ohio) before June 1, 2015, by Air Force Office of Scientific Research (AFOSR) (FA9550-15-1-0213) between June 1, 2015, and July 31, 2017, and by Utah State University after August 1, 2017. D.S. was supported by the W. M. Keck Foundation. Computational resources were provided by the Argonne Leadership Computing Facilities at Argonne National Laboratory under Department of Energy contract DE-AC-06CH11357 and by the Extreme Science and Engineering Discovery Environment at Texas Advanced Computing Center under National Science Foundation contract TG-CHE130008. Y.R. thank Drs. Tony F. Heinz, Omer Yaffe, Xiaoyang Zhu, and Francis Spano for their beneficial discussion.
Author Contributions
D.S., G.-H.D., B.X., Y.W., and Y.Q. performed the transient absorption measurements and characterizations. E.X. performed the theoretical calculation. X.L. and Y.Z. prepared the samples. D.S., G.-H.D., B.X., C.N., A.R.H., H.-L.D., G.C., H.C., and Y.R. analyzed the data. Y.R., G.C., A.R.H., H.C., C.N., and H.-L.D. designed the project. Y.R., G.C., C.N., A.R.H., and H.-L.D. supervised the experiments. H.C. supervised the theoretical calculation. D.S, G.-H.D, B.X., and E.X. wrote the original draft. Y.R., H.C., and G.C. revised the manuscript.
Declaration of Interests
The authors declare no competing interests.
Published: September 27, 2019
Footnotes
Supplemental Information can be found online at https://doi.org/10.1016/j.isci.2019.08.053.
Contributor Information
Gugang Chen, Email: gchen@honda-ri.com.
Hanning Chen, Email: chenhanning@email.gwu.edu.
Yi Rao, Email: yi.rao@usu.edu.
Supplemental Information
References
- Aquilante F., Autschbach J., Carlson R.K., Chibotaru L.F., Delcey M.G., De Vico L., Fdez Galván I., Ferré N., Frutos L.M., Gagliardi L. Molcas 8: new capabilities for multiconfigurational quantum chemical calculations across the periodic table. J. Comput. Chem. 2016;37:506–541. doi: 10.1002/jcc.24221. [DOI] [PubMed] [Google Scholar]
- Bakulin A.A., Morgan S.E., Kehoe T.B., Wilson M.W.B., Chin A.W., Zigmantas D., Egorova D., Rao A. Real-time observation of multiexcitonic states in ultrafast singlet fission using coherent 2D electronic spectroscopy. Nat. Chem. 2016;8:16–23. doi: 10.1038/nchem.2371. [DOI] [PubMed] [Google Scholar]
- Basel B.S., Papadopoulos I., Thiel D., Casillas R., Zirzlmeier J., Clark T., Guldi D.M., Tykwinski R.R. Pentacenes: a molecular Ruler for singlet fission. Trends Chem. 2019;1:11–21. [Google Scholar]
- Beljonne D., Yamagata H., Bredas J.L., Spano F.C., Olivier Y. Charge-transfer excitations steer the Davydov splitting and mediate singlet exciton fission in pentacene. Phys. Rev. Lett. 2013;110:226402. doi: 10.1103/PhysRevLett.110.226402. [DOI] [PubMed] [Google Scholar]
- Berkelbach T.C., Hybertsen M.S., Reichman D.R. Microscopic theory of singlet exciton fission. I. General formulation. J. Chem. Phys. 2013;138:114102. doi: 10.1063/1.4794425. [DOI] [PubMed] [Google Scholar]
- Berkelbach T.C., Hybertsen M.S., Reichman D.R. Microscopic theory of singlet exciton fission. II. Application to pentacene dimers and the role of superexchange. J. Chem. Phys. 2013;138:114103. doi: 10.1063/1.4794427. [DOI] [PubMed] [Google Scholar]
- Berkelbach T.C., Hybertsen M.S., Reichman D.R. Microscopic theory of singlet exciton fission. III. Crystalline pentacene. J. Chem. Phys. 2014;141:074705. doi: 10.1063/1.4892793. [DOI] [PubMed] [Google Scholar]
- Burdett J.J., Bardeen C.J. Quantum beats in crystalline tetracene delayed fluorescence due to triplet pair coherences produced by direct singlet fission. J. Am. Chem. Soc. 2012;134:8597–8607. doi: 10.1021/ja301683w. [DOI] [PubMed] [Google Scholar]
- Burdett J.J., Muller A.M., Gosztola D., Bardeen C.J. Excited state dynamics in solid and monomeric tetracene: the roles of superradiance and exciton fission. J. Chem. Phys. 2010;133:144506. doi: 10.1063/1.3495764. [DOI] [PubMed] [Google Scholar]
- Busby E., Berkelbach T.C., Kumar B., Chernikov A., Zhong Y., Hlaing H., Zhu X.Y., Heinz T.F., Hybertsen M.S., Sfeir M.Y. Multiphonon relaxation slows singlet fission in crystalline hexacene. J. Am. Chem. Soc. 2014;136:10654–10660. doi: 10.1021/ja503980c. [DOI] [PubMed] [Google Scholar]
- Busby E., Xia J., Wu Q., Low J.Z., Song R., Miller J.R., Zhu X.Y., Campos L.M., Sfeir M.Y. A design strategy for intramolecular singlet fission mediated by charge-transfer states in donor-acceptor organic materials. Nat. Mater. 2015;14:426–433. doi: 10.1038/nmat4175. [DOI] [PubMed] [Google Scholar]
- Casanova D. Theoretical modeling of singlet fission. Chem. Rev. 2018;118:7164–7207. doi: 10.1021/acs.chemrev.7b00601. [DOI] [PubMed] [Google Scholar]
- Casida M.E., Huix-Rotllant M. Progress in time-dependent density-functional theory. Annu. Rev. Phys. Chem. 2012;63:287–323. doi: 10.1146/annurev-physchem-032511-143803. [DOI] [PubMed] [Google Scholar]
- Chan W.L., Berkelbach T.C., Provorse M.R., Monahan N.R., Tritsch J.R., Hybertsen M.S., Reichman D.R., Gao J.L., Zhu X.Y. The quantum coherent mechanism for singlet fission: experiment and theory. Acc. Chem. Res. 2013;46:1321–1329. doi: 10.1021/ar300286s. [DOI] [PubMed] [Google Scholar]
- Chan W.L., Ligges M., Jailaubekov A., Kaake L., Miaja-Avila L., Zhu X.Y. Observing the multiexciton state in singlet fission and ensuing ultrafast multielectron transfer. Science. 2011;334:1541–1545. doi: 10.1126/science.1213986. [DOI] [PubMed] [Google Scholar]
- Chan W.L., Ligges M., Zhu X.Y. The energy barrier in singlet fission can be overcome through coherent coupling and entropic gain. Nat. Chem. 2012;4:840–845. doi: 10.1038/nchem.1436. [DOI] [PubMed] [Google Scholar]
- Chen H. Functional mode electron-transfer theory. J. Phys. Chem. C. 2014;118:7586–7593. doi: 10.1021/jp504418c. [DOI] [PubMed] [Google Scholar]
- Chernikov A., Yaffe O., Kumar B., Zhong Y., Nuckolls C., Heinz T.F. Spectroscopic study of anisotropic excitons in single crystal hexacene. J. Phys. Chem. Lett. 2014;5:3632–3635. doi: 10.1021/jz501693g. [DOI] [PubMed] [Google Scholar]
- Congreve D.N., Lee J., Thompson N.J., Hontz E., Yost S.R., Reusswig P.D., Bahlke M.E., Reineke S., Van Voorhis T., Baldo M.A. External quantum efficiency above 100% in a singlet-exciton-fission–based organic photovoltaic cell. Science. 2013;340:334–337. doi: 10.1126/science.1232994. [DOI] [PubMed] [Google Scholar]
- Cook J.D., Carey T.J., Damrauer N.H. Solution-phase singlet fission in a structurally well-defined norbornyl-bridged tetracene dimer. J. Phys. Chem. C. 2016;120:4473–4481. doi: 10.1021/acs.jpca.6b04367. [DOI] [PubMed] [Google Scholar]
- Coto P.B., Sharifzadeh S., Neaton J.B., Thoss M. Low-lying electronic excited states of pentacene oligomers: a comparative electronic structure study in the context of singlet fission. J. Chem. Theory Comput. 2015;11:147–156. doi: 10.1021/ct500510k. [DOI] [PubMed] [Google Scholar]
- Davydov A. Theory of absorption spectra of molecular crystals. Zh. Eksp. Teor. Fiz. 1948;18:210–218. [Google Scholar]
- Davydov A. Springer; 2013. Theory of Molecular Excitons. [Google Scholar]
- Elenewski J.E., Cai J.Y., Jiang W., Chen H. Functional mode hot electron transfer theory. J. Phys. Chem. C. 2016;120:20579–20587. [Google Scholar]
- Elenewski J.E., Cubeta U.S., Ko E., Chen H. Computer simulation of singlet fission in single crystalline pentacene by functional mode vibronic theory. J. Phys. Chem. C. 2017;121:11159–11165. [Google Scholar]
- Elenewski J.E., Cubeta U.S., Ko E., Chen H. Functional mode singlet fission theory. J. Phys. Chem. C. 2017;121:4130–4138. [Google Scholar]
- Feng X.T., Krylov A.I. On couplings and excimers: lessons from studies of singlet fission in covalently linked tetracene dimers. Phys. Chem. Chem. Phys. 2016;18:7751–7761. doi: 10.1039/c6cp00177g. [DOI] [PubMed] [Google Scholar]
- Folie B.D., Haber J.B., Refaely-Abramson S., Neaton J.B., Ginsberg N.S. Long-lived correlated triplet pairs in a π-stacked crystalline pentacene derivative. J. Am. Chem. Soc. 2018;140:2326–2335. doi: 10.1021/jacs.7b12662. [DOI] [PubMed] [Google Scholar]
- Hachmann J., Dorando J.J., Avilés M., Chan G.K.-L. The radical character of the acenes: a density matrix renormalization group study. J. Chem. Phys. 2007;127:134309. doi: 10.1063/1.2768362. [DOI] [PubMed] [Google Scholar]
- Hestand N.J., Yamagata H., Xu B., Sun D., Zhong Y., Harutyunyan A.R., Chen G., Dai H.-L., Rao Y., Spano F.C. Polarized absorption in crystalline pentacene: theory vs experiment. J. Phys. Chem. C. 2015;119:22137–22147. [Google Scholar]
- Hub J.S., de Groot B.L. Detection of functional modes in protein dynamics. PLoS Comput. Biol. 2009;5:e1000480. doi: 10.1371/journal.pcbi.1000480. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ishizaki A., Fleming G.R. On the adequacy of the Redfield equation and related approaches to the study of quantum dynamics in electronic energy transfer. J. Chem. Phys. 2009;130:234110. doi: 10.1063/1.3155214. [DOI] [PubMed] [Google Scholar]
- Jiang D.-E., Dai S. Electronic ground state of higher acenes. J. Phys. Chem. C. 2008;112:332–335. doi: 10.1021/jp0765087. [DOI] [PubMed] [Google Scholar]
- Johnson J.C., Nozik A.J., Michl J. High triplet yield from singlet fission in a thin film of 1,3-diphenylisobenzofuran. J. Am. Chem. Soc. 2010;132:16302–16303. doi: 10.1021/ja104123r. [DOI] [PubMed] [Google Scholar]
- Jundt C., Klein G., Sipp B., Lemoigne J., Joucla M., Villaeys A.A. Exciton dynamics in pentacene thin films studied by pump-probe spectroscopy. Chem. Phys. Lett. 1995;241:84–88. [Google Scholar]
- Kaduk B., Kowalczyk T., Van Voorhis T. Constrained density functional theory. Chem. Rev. 2012;112:321–370. doi: 10.1021/cr200148b. [DOI] [PubMed] [Google Scholar]
- Kandada A.R.S., Petrozza A., Lanzani G. Ultrafast dissociation of triplets in pentacene induced by an electric field. Phys. Rev. B. 2014;90:075310. [Google Scholar]
- Katoh R., Kotani M., Hirata Y., Okada T. Triplet exciton formation in a benzophenone single crystal studied by picosecond time-resolved absorption spectroscopy. Chem. Phys. Lett. 1997;264:631–635. [Google Scholar]
- Kolomeisky A.B., Feng X., Krylov A.I. A simple kinetic model for singlet fission: a role of electronic and entropic contributions to macroscopic rates. J. Phys. Chem. C. 2014;118:5188–5195. [Google Scholar]
- Korovina N.V., Das S., Nett Z., Feng X., Joy J., Haiges R., Krylov A.I., Bradforth S.E., Thompson M.E. Singlet fission in a covalently linked cofacial alkynyltetracene dimer. J. Am. Chem. Soc. 2016;138:617–627. doi: 10.1021/jacs.5b10550. [DOI] [PubMed] [Google Scholar]
- Kubas A., Hoffmann F., Heck A., Oberhofer H., Elstner M., Blumberger J. Electronic couplings for molecular charge transfer: benchmarking CDFT, FODFT, and FODFTB against high-level ab initio calculations. J. Chem. Phys. 2014;140:104105. doi: 10.1063/1.4867077. [DOI] [PubMed] [Google Scholar]
- Laudise R.A., Kloc C., Simpkins P.G., Siegrist T. Physical vapor growth of organic semiconductors. J. Cryst. Growth. 1998;187:449–454. [Google Scholar]
- Lee J., Bruzek M.J., Thompson N.J., Sfeir M.Y., Anthony J.E., Baldo M.A. Singlet exciton fission in a hexacene derivative. Adv. Mater. 2013;25:1445–1448. doi: 10.1002/adma.201203982. [DOI] [PubMed] [Google Scholar]
- Lukman S., Chen K., Hodgkiss J.M., Turban D.H.P., Hine N.D.M., Dong S., Wu J., Greenham N.C., Musser A.J. Tuning the role of charge-transfer states in intramolecular singlet exciton fission through side-group engineering. Nat. Commun. 2016;7:13622. doi: 10.1038/ncomms13622. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lukman S., Musser A.J., Chen K., Athanasopoulos S., Yong C.K., Zeng Z., Ye Q., Chi C., Hodgkiss J.M., Wu J. Tuneable singlet exciton fission and triplet–triplet annihilation in an orthogonal pentacene dimer. Adv. Funct. Mater. 2015;25:5452–5461. [Google Scholar]
- Marciniak H., Pugliesi I., Nickel B., Lochbrunner S. Ultrafast singlet and triplet dynamics in microcrystalline pentacene films. Phys. Rev. B. 2009;79:235318. doi: 10.1103/PhysRevLett.99.176402. [DOI] [PubMed] [Google Scholar]
- Merrifield R.E., Avakian P., Groff R.P. Fission of singlet excitons into pairs of triplet excitons in tetracene crystals. Chem. Phys. Lett. 1969;3:155–157. [Google Scholar]
- Michl J., Nozik A.J., Chen X., Johnson J.C., Rana G., Akdag A., Schwerin A.F. Organic Photovoltaics VIII, Proc. SPIE. Vol. 6656. International Society for Optics and Photonics; 2007. Toward Singlet Fission for Excitonic Solar Cells; p. 66560E. [Google Scholar]
- Mirjani F., Renaud N., Gorczak N., Grozema F.C. Theoretical investigation of singlet fission in molecular dimers: the role of charge transfer states and quantum interference. J. Phys. Chem. C. 2014;118:14192–14199. [Google Scholar]
- Miyata K., Kurashige Y., Watanabe K., Sugimoto T., Takahashi S., Tanaka S., Takeya J., Yanai T., Matsumoto Y. Coherent singlet fission activated by symmetry breaking. Nat. Chem. 2017;9:983–989. doi: 10.1038/nchem.2784. [DOI] [PubMed] [Google Scholar]
- Monahan N., Zhu X.-Y. Charge transfer–mediated singlet fission. Annu. Rev. Phys. Chem. 2015;66:601–618. doi: 10.1146/annurev-physchem-040214-121235. [DOI] [PubMed] [Google Scholar]
- Monahan N.R., Sun D., Tamura H., Williams K.W., Xu B., Zhong Y., Kumar B., Nuckolls C., Harutyunyan A.R., Chen G. Dynamics of the triplet-pair state reveals the likely coexistence of coherent and incoherent singlet fission in crystalline hexacene. Nat. Chem. 2017;9:341–346. doi: 10.1038/nchem.2665. [DOI] [PubMed] [Google Scholar]
- Musser A.J., Liebel M., Schnedermann C., Wende T., Kehoe T.B., Rao A., Kukura P. Evidence for conical intersection dynamics mediating ultrafast singlet exciton fission. Nat. Phys. 2015;11:352–357. [Google Scholar]
- Najafov H., Lee B., Zhou Q., Feldman L.C., Podzorov V. Observation of long-range exciton diffusion in highly ordered organic semiconductors. Nat. Mater. 2010;9:938–943. doi: 10.1038/nmat2872. [DOI] [PubMed] [Google Scholar]
- Nakano M., Ito S., Nagami T., Kitagawa Y., Kubo T. Quantum master equation approach to singlet fission dynamics of realistic/artificial pentacene dimer models: relative relaxation factor analysis. J. Phys. Chem. C. 2016;120:22803–22815. [Google Scholar]
- Pensack R.D., Ostroumov E.E., Tilley A.J., Mazza S., Grieco C., Thorley K.J., Asbury J.B., Seferos D.S., Anthony J.E., Scholes G.D. Observation of two triplet-pair intermediates in singlet exciton fission. J. Phys. Chem. Lett. 2016;7:2370–2375. doi: 10.1021/acs.jpclett.6b00947. [DOI] [PubMed] [Google Scholar]
- Rao A., Friend R.H. Harnessing singlet exciton fission to break the Shockley–Queisser limit. Nat. Rev. Mater. 2017;2:17063. [Google Scholar]
- Rao A., Wilson M.W.B., Albert-Seifried S., Di Pietro R., Friend R.H. Photophysics of pentacene thin films: the role of exciton fission and heating effects. Phys. Rev. B. 2011;84:195411. [Google Scholar]
- Schwerin A.F., Johnson J.C., Smith M.B., Sreearunothai P., Popović D., Černý J., Havlas Z., Paci I., Akdag A., MacLeod M.K. Toward designed singlet fission: electronic states and photophysics of 1,3-diphenylisobenzofuran. J. Phys. Chem. C. 2010;114:1457–1473. doi: 10.1021/jp907401t. [DOI] [PubMed] [Google Scholar]
- Schwob H.P., Williams D.F. Charge transfer exciton fission in anthracene crystals. Chem. Phys. Lett. 1972;13:581–584. [Google Scholar]
- Schwoerer H., Birech Z. Frontiers in Optics 2014. Optical Society of America; 2014. Davydov splitting in triplet excitons of tetracene single crystals; p. FTu1G.8. [Google Scholar]
- Smith M.B., Michl J. Singlet fission. Chem. Rev. 2010;110:6891–6936. doi: 10.1021/cr1002613. [DOI] [PubMed] [Google Scholar]
- Stern H.L., Cheminal A., Yost S.R., Broch K., Bayliss S.L., Chen K., Tabachnyk M., Thorley K., Greenham N., Hodgkiss J.M. Vibronically coherent ultrafast triplet-pair formation and subsequent thermally activated dissociation control efficient endothermic singlet fission. Nat. Chem. 2017;9:1205–1212. doi: 10.1038/nchem.2856. [DOI] [PubMed] [Google Scholar]
- Stern H.L., Musser A.J., Gelinas S., Parkinson P., Herz L.M., Bruzek M.J., Anthony J., Friend R.H., Walker B.J. Identification of a triplet pair intermediate in singlet exciton fission in solution. Proc. Natl. Acad. Sci. U S A. 2015;112:7656–7661. doi: 10.1073/pnas.1503471112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Takeda Y., Katoh R., Kobayashi H., Kotani M. Fission and fusion of excitons in perylene crystal studied with VUV and x-ray excitation. J. Electron. Spectrosc. Relat. Phenom. 1996;78:423–425. [Google Scholar]
- Tamura H., Huix-Rotllant M., Burghardt I., Olivier Y., Beljonne D. First-principles quantum dynamics of singlet fission: coherent versus thermally activated mechanisms governed by molecular π stacking. Phys. Rev. Lett. 2015;115:107401. doi: 10.1103/PhysRevLett.115.107401. [DOI] [PubMed] [Google Scholar]
- Tao G. Bath effect in singlet fission dynamics. J. Phys. Chem. C. 2014;118:27258–27264. [Google Scholar]
- Tayebjee M.J.Y., Clady R., Schmidt T.W. The exciton dynamics in tetracene thin films. Phys. Chem. Chem. Phys. 2013;15:14797–14805. doi: 10.1039/c3cp52609g. [DOI] [PubMed] [Google Scholar]
- Tiago M.L., Northrup J.E., Louie S.G. Ab initio calculation of the electronic and optical properties of solid pentacene. Phys. Rev. B. 2003;67:115212. [Google Scholar]
- Walker B.J., Musser A.J., Beljonne D., Friend R.H. Singlet exciton fission in solution. Nat. Chem. 2013;5:1019–1024. doi: 10.1038/nchem.1801. [DOI] [PubMed] [Google Scholar]
- Wan Y., Guo Z., Zhu T., Yan S., Johnson J., Huang L. Cooperative singlet and triplet exciton transport in tetracene crystals visualized by ultrafast microscopy. Nat. Chem. 2015;7:785–792. doi: 10.1038/nchem.2348. [DOI] [PubMed] [Google Scholar]
- Wan Y., Wiederrecht G.P., Schaller R.D., Johnson J.C., Huang L. Transport of spin-entangled triplet excitons generated by singlet fission. J. Phys. Chem. Lett. 2018;9:6731–6738. doi: 10.1021/acs.jpclett.8b02944. [DOI] [PubMed] [Google Scholar]
- Warshel A., Levitt M. Theoretical studies of enzymic reactions: dielectric, electrostatic and steric stabilization of the carbonium ion in the reaction of lysozyme. J. Mol. Biol. 1976;103:227–249. doi: 10.1016/0022-2836(76)90311-9. [DOI] [PubMed] [Google Scholar]
- Watanabe M., Chang Y.J., Liu S.-W., Chao T.-H., Goto K., Islam M.M., Yuan C.-H., Tao Y.-T., Shinmyozu T., Chow T.J. The synthesis, crystal structure and charge-transport properties of hexacene. Nat. Chem. 2012;4:574–578. doi: 10.1038/nchem.1381. [DOI] [PubMed] [Google Scholar]
- Watanabe S., Furube A., Katoh R. Generation and decay dynamics of triplet excitons in Alq3 thin films under high-density excitation conditions. J. Phys. Chem. C. 2006;110:10173–10178. doi: 10.1021/jp0618227. [DOI] [PubMed] [Google Scholar]
- Wen X.M., Yu P., Yuan C.T., Ma X.Q., Tang J. Singlet and triplet carrier dynamics in rubrene single crystal. J. Phys. Chem. C. 2013;117:17741–17747. [Google Scholar]
- Wilson M.W.B., Rao A., Clark J., Kumar R.S.S., Brida D., Cerullo G., Friend R.H. Ultrafast dynamics of exciton fission in polycrystalline pentacene. J. Am. Chem. Soc. 2011;133:11830–11833. doi: 10.1021/ja201688h. [DOI] [PubMed] [Google Scholar]
- Wilson M.W.B., Rao A., Ehrler B., Friends R.H. Singlet exciton fission in polycrystalline pentacene: from photophysics toward devices. Acc. Chem. Res. 2013;46:1330–1338. doi: 10.1021/ar300345h. [DOI] [PubMed] [Google Scholar]
- Yamagata H., Norton J., Hontz E., Olivier Y., Beljonne D., Bredas J.L., Silbey R.J., Spano F.C. The nature of singlet excitons in oligoacene molecular crystals. J. Chem. Phys. 2011;134:204703. doi: 10.1063/1.3590871. [DOI] [PubMed] [Google Scholar]
- Yong C.K., Musser A.J., Bayliss S.L., Lukman S., Tamura H., Bubnova O., Hallani R.K., Meneau A., Resel R., Maruyama M. The entangled triplet pair state in acene and heteroacene materials. Nat. Commun. 2017;8:15953. doi: 10.1038/ncomms15953. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yost S.R., Lee J., Wilson M.W.B., Wu T., McMahon D.P., Parkhurst R.R., Thompson N.J., Congreve D.N., Rao A., Johnson K. A transferable model for singlet-fission kinetics. Nat. Chem. 2014;6:492–497. doi: 10.1038/nchem.1945. [DOI] [PubMed] [Google Scholar]
- Zenz C., Cerullo G., Lanzani G., Graupner W., Meghdadi F., Leising G., De Silvestri S. Ultrafast photogeneration mechanisms of triplet states in para-hexaphenyl. Phys. Rev. B. 1999;59:14336–14341. [Google Scholar]
- Zhang B., Zhang C., Xu Y., Wang R., He B., Liu Y., Zhang S., Wang X., Xiao M. Polarization-dependent exciton dynamics in tetracene single crystals. J. Chem. Phys. 2014;141:244303. doi: 10.1063/1.4904385. [DOI] [PubMed] [Google Scholar]
- Zimmerman P.M., Bell F., Casanova D., Head-Gordon M. Mechanism for singlet fission in pentacene and tetracene: from single exciton to two triplets. J. Am. Chem. Soc. 2011;133:19944–19952. doi: 10.1021/ja208431r. [DOI] [PubMed] [Google Scholar]
- Zimmerman P.M., Musgrave C.B., Head-Gordon M. A correlated electron view of singlet fission. Acc. Chem. Res. 2013;46:1339–1347. doi: 10.1021/ar3001734. [DOI] [PubMed] [Google Scholar]
- Zimmerman P.M., Zhang Z.Y., Musgrave C.B. Singlet fission in pentacene through multi-exciton quantum states. Nat. Chem. 2010;2:648–652. doi: 10.1038/nchem.694. [DOI] [PubMed] [Google Scholar]
- Zirzlmeier J., Lehnherr D., Coto P.B., Chernick E.T., Casillas R., Basel B.S., Thoss M., Tykwinski R.R., Guldi D.M. Singlet fission in pentacene dimers. Proc. Nat. Acad. Sci. U S A. 2015;112:5325–5330. doi: 10.1073/pnas.1422436112. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.