Abstract
The photophysical processes of singlet fission and triplet fusion have numerous emerging applications. They involve the separation of a photo-generated singlet exciton into two dark triplet excitons and the fusion of two dark triplet excitons into an emissive singlet exciton, respectively. The role of the excimer state and the nature of the triplet-pair state in these processes have been a matter of contention. Here we analyse the room temperature time-resolved emission of a neat liquid singlet fission chromophore and show that it exhibits three spectral components: two that correspond to the bright singlet and excimer states and a third component that becomes more prominent during triplet fusion. This spectrum is enhanced by magnetic fields, confirming its origins in the recombination of weakly coupled triplet pairs. It is thus attributed to a strongly coupled triplet pair state. These observations unite the view that there is an emissive intermediate in singlet fission and triplet fusion, distinct from the broad, unstructured excimer emission.
Similar content being viewed by others
Main
Singlet fission (SF)1 and triplet fusion (TF)2 are related photophysical processes with promising applications spanning photovoltaics3,4,5, photocatalysis6, optical and magnetic resonance imaging7,8, multi-excitonic logic9,10 and advanced manufacturing11. In photovoltaics, these processes offer the possibility of much higher limiting efficiencies compared with single-threshold solar cells3,12,13,14. In SF, a photo-prepared excited singlet state couples to a ground-state chromophore and transfers about half its energy, ultimately bringing about two uncoupled chromophores in their lowest-energy triplet states1. The system conserves spin and is widely accepted to traverse a region of the potential energy landscape corresponding to a spin-0 (singlet) triplet-pair state, 1(TT), with exchange coupling much greater than the zero-field splitting of the individual triplet states15,16,17. As the triplets move apart, the exchange coupling diminishes and the spin-0 character of the bichromophoric state is mixed with other multiplicities18. These weakly coupled triplets may return to the singlet manifold by the process of TF2,18,19.
In TF, the reverse process occurs whereby weakly coupled triplets become more strongly coupled before crossing to the highly emissive excited singlet state2. In concentrated solutions, aggregates and films, the excited singlet state may form an excimer, corresponding to a deep well on the potential energy surface brought about by strong excitonic coupling20. The interplay between the excited singlet, S1, the excimer, 1Ex, and the 1(TT) state has been the subject of debate19,21,22,23,24,25,26,27,28. In concentrated solutions of 5,12-bis((triisopropylsilyl)ethynyl)tetracene (TIPS-Tc) and 6,13-bis((triisopropylsilyl)ethynyl)pentacene, it has been asserted that the excimer constitutes an intermediate in SF21,22. However, though excimers might be involved, Dvořák et al.23 showed, using total internal reflection excitation, that concentrated 6,13-bis((triisopropylsilyl)ethynyl)pentacene solutions do not exhibit an emissive excimer. Furthermore, through careful analysis of the time-resolved photoluminescence (TRPL) of concentrated TIPS-Tc solutions, Dover et al.19 demonstrated that the excimer, evidenced by a featureless and red-shifted emission spectrum, served as a trap, and that triplets could be generated from the S1 state without necessarily accessing the excimeric well. Bossanyi et al.24 showed that in pentacene single crystals and 2,8-difluoro-5,11-bis(triethylsilylethynyl)anthradithiophene films at 100 K, a red-shifted spectrum, distinct from the excimer, evinced an emissive intermediate state in TF, which they labelled as 1(TT). Such a weakly emissive state was previously observed in low-temperature tetracene thin films29,30. The 1(TT) intermediate has been the subject of a number of other reports in aggregates and films of acene derivatives15,31,32,33,34,35,36. This begs the question of can such a third, emissive state be observed in SF liquids or solutions? Such an observation would unite models of SF across different phases at device-relevant temperatures. Furthermore, such a spectrum should be enhanced by magnetic fields, which are known to attenuate the generation of free triplets37,38.
To elucidate this question, we synthesized a room temperature liquid SF material 5,12-bis(n-octyldiisopropylsilylethynyl)tetracene (NODIPS-Tc) (Fig. 1a). By adding solvent to the liquid, we bridge between a material composed of neat chromophore, which acts as a solid on experimental timescales, as revealed by molecular dynamics (MD) simulations and solutions of dynamic chromophores. Detailed analysis of the TRPL of this material, both neat and in concentrated solution, reveals three spectral components. The third component, attributed to 1(TT), becomes more prominent during TF. Furthermore, the spectrally resolved magnetic field effect (MFE) on the luminescence of this system reveals that luminescence enhancement is explained by enhanced 1(TT) emission. This study unites the observations of Dover et al.19 and Bossanyi et al.24, showing that there is a spectrally observable emissive intermediate in SF both in concentrated solutions and the solid state, but that this is distinct from the broad, red-shifted excimer emission.
Results
MD
MD simulations were performed on neat NODIPS-Tc, and a 50% w/w mixture with toluene. The neat liquid is essentially immobile on experimental timescales, with a structural decorrelation time on the order of 100 μs. However, the structural decorrelation of the mixture was 104 times faster, with a time constant of about 10 ns (Extended Data Fig. 1).
The simulations predict that the closest nearest-neighbour interaction of both the neat liquid and the mixture are at a distance of 3.8 Å, with a π-stacking angle of about 60°. The second nearest configuration has the chromophores π-stacked in a perpendicular arrangement at a distance of 4.4 Å (Extended Data Fig. 2). These two arrangements occur as distinct peaks in the configurational probability distributions in the neat liquid, which merge into a single peak/ridge in the mixture (angular–radial distribution functions in Supplementary Figs. 9c and 10c). Configurations expected to be prone to excimer formation, with anti-parallel or parallel transition moments, are found at larger distances of around 5.6 and 6.5 Å, respectively, in both simulations.
TRPL
The room temperature TRPL of neat NODIPS-Tc liquid is shown in Fig. 1b. Each spectrum is normalized by its integral, accentuating the changes that occur as the spectrum evolves over time. At early times (0–8 ns, epoch 1), the emission is dominated by the photo-generated S1 state, rapidly red shifting into an 1Ex-dominated spectral shape in epoch 2 (8–80 ns). The S1 emission appears H-aggregated39, with the 0–0 band suppressed to a greater extent than in more dilute (yet optically thick) samples measured under identical conditions (Extended Data Fig. 3). This observation is consistent with the predominance of π-stacked nearest neighbours in the molecular dynamics simulations. For comparison, the steady-state absorption and photoluminescence spectra of dilute NODIPS-Tc solution are shown in Fig. 1a. The expected effect of self-absorption on the spectra is shown in Extended Data Fig. 4.
The early time spectral slices spanning epochs 1 and 2 are plotted in Fig. 1c. There is a rapid shift of the iso-emissive point (IEP) in the first few nanoseconds, which then settles down near 620 nm. As with TIPS-Tc, in NODIPS-Tc, both SF and TF pathways are thermally accessible19,40. After 100 ns, in epoch 3, the annihilation of SF-generated triplets dominates the spectrum. Spectral slices spanning all epochs are plotted in Fig. 1d.
The presence of a stable IEP after a few nanoseconds in the area-normalized spectra suggests that there are two major spectral components. The IEP deviates in the first few nanoseconds, indicating that there is probably a third component. To quantify the presence of a third component, we implemented principal component analysis (PCA).
PCA
The PCA of the dataset is shown in Fig. 2a. A scree plot41 showing the percentage variance explained by each principal component (PC) and the cumulative variance is plotted in Fig. 2b. It is important to note that PCA does not generate the spectra of species, and as the PCs are necessarily orthogonal, they can exhibit differently signed spectral regions. PC1 represents the average spectrum and PC2 accounts for the principal spectral changes, with the S1 component diminishing and the 1Ex component growing in time. The variance attributed to each PC drops steadily before an elbow appears at PC4, indicating the onset of insignificant factors41. The eigenvector of PC4 is very noisy. Since three spectra are required to reproduce the deviation from a single IEP in Fig. 1c, and there is no evidence of a significant fourth component, we proceed on the basis that there are three spectral components in the neat sample. The PCA and scree plots for the mixtures with toluene (75%, 50% and 30% w/w) are shown in Extended Data Fig. 5. The third component is not in evidence in these samples as shown by PCA, although the PC3 eigenvector seems to contain spectral information.
The number of PCs may be equated with the number of kinetically and spectrally distinct emissive species. As such, three PCs suggests at least three emissive species. Clearly, during the transition from epoch 1 to 2, the red-shifted 1Ex emission grows at the expense of the bluer, S1-like spectrum. Since the initially prepared state is S1 and the IEP deviates and then settles in the first few nanoseconds, there must be a rapid quasi-equilibrium established between the S1 state and another species. We propose that this third spectral component is due to emission by an excitonically coupled chromophore pair, which is spectrally distinct from 1Ex. We assign this species to the exchange-coupled triplet pair state, 1(TT), as reported in the solid state24,33,35,36,42,43.
Spectral decomposition
The kinetic scheme is illustrated in Fig. 3a. In epoch 1, there is a rapid onset of equilibrium between S1 and 1(TT), which then equilibrates with the 1Ex state in epoch 2. Dissociation of 1(TT) states (SF) generates a pool of free triplets, which then undergo mutual annihilation (TF) in epoch 3.
The spectra of the three epochs, σ1−3, are shown in Fig. 3c, and are linear combinations of the species-associated spectra, σS, σTT and σEx. Epoch 1, σ1, is clearly σS dominated, and epoch 2 is σEx dominated. Epoch 3 closely resembles the steady-state emission spectrum, indicating that TF principally populates the S1 state after transiting the 1(TT) state. Subsequent emission is equivalent to the steady-state spectrum. Since σ3 is generated by 1(TT) states, it will naturally have a higher proportion of 1(TT) emission than σ1 or σ2.
The spectral evolution that occurs in the first nanoseconds suggests that the 1(TT) spectrum, σTT, is slightly red shifted compared with the photo-generated S1 state spectrum, σS, but does not have the same intensity in the deep-red region as the 1Ex state spectrum, σEx. As such, we may isolate σTT by assuming that the bluest emission is due to the S1 state and the reddest emission is due to the 1Ex state. Subtracting contributions due to σS, and the excimer-dominated spectrum σ2, from σ3 results in the σTT spectrum displayed in Fig. 3d. The other species-associated emission spectra, σS and σEx are also displayed in Fig. 3d.
Magnetophotoluminescence
To shed further light on the spectral characteristics of the 1(TT) state, we performed magnetic photoluminescence experiments. Though not directly observable through photoluminescence, there must also be a bichromophoric state in which the exchange coupling is weaker than the zero-field splitting, 1(T…T). This weakly coupled regime is included in the kinetic scheme illustrated in Fig. 3a. For aligned chromophores, at zero-field there are three out of nine triplet pair sublevels with singlet character (\(\left\vert xx\right\rangle ,\left\vert\, yy\right\rangle\) and \(\left\vert zz\right\rangle\), where the triplet basis is \(\{\left\vert x\right\rangle ,\left\vert\, y\right\rangle ,\left\vert z\right\rangle \}\)). As a magnetic field is applied, these states mix with other sublevels and the number of levels with singlet character increases, essentially increasing the entropy of the 1(T…T) state. At higher fields where the basis transforms into \(\{\left\vert -\right\rangle ,\left\vert 0\right\rangle ,\left\vert +\right\rangle \}\), only two of the nine weakly coupled substates have singlet character, \(\left\vert 00\right\rangle\) and \((\left\vert +-\right\rangle +\left\vert -+\right\rangle )/\sqrt{2}\), and the entropy of the state is diminished37,44,45,46. Low entropy of an intermediate is accompanied by a higher free energy of activation, and thus rate constants are effectively manipulated by the magnetic field (Supplementary Information). This effect of the magnetic field on the free energy of the 1(T…T) state is illustrated schematically in Fig. 3a.
The effect of increasing the free energy of the 1(T…T) state is to attenuate SF, and thus enhance photoluminescence. In Fig. 4a, we see that at fields as high as 210 mT, for the neat sample, the photoluminescence is enhanced by about 3.6% (ΔPL/PL). Furthermore, we see that at low fields, the photoluminescence is diminished, which is characteristic of chromophores that exhibit hindered rotational diffusion on the timescales of SF and TF47. This is less evident but persists in the 75% w/w sample. In the 50% w/w sample, both the effect of hindered rotation and the SF rate are diminished, giving rise to a smaller positive MFE at all fields. This is consistent with the results of the MD simulations, which showed that the mobilities were 104 times higher in the 50% w/w mixture than the neat liquid.
The ΔPL/PL spectrum is shown in Fig. 4b. All samples exhibit peaks around 560 nm and 595 nm. Raising the free energy of the 1(T…T) state with a magnetic field should enhance the emission non-uniformly by initially returning a population to the 1(TT) state. Indeed, we would expect, as a first approximation, that the magnetic enhancement would be identical to the spectrum observed during epoch 3, where TF dominates. The ratio of the TF spectrum in epoch 3 to the steady-state spectrum is plotted in Fig. 4b (in blue). It exhibits the same shape as the ΔPL/PL spectrum. The population that returned to the S1 state by the action of the magnetic field can be expected to have an identical fate to photo-generated S1 chromophores, and thus the enhanced emission will be identical to the steady-state emission. Contributions to the ΔPL/PL spectrum from TF-generated S1 states will thus result in a featureless ΔPL/PL spectrum. Any spectral features can thus be attributed to excess 1Ex or 1(TT) states.
The 1(TT)-associated spectrum σTT (Fig. 3d) is ratioed to the steady-state spectrum and plotted in Fig. 4b (in green). The ratio of the excimer spectrum to the steady-state spectrum is shown in black. Clearly, the peaks at 560 nm and 595 nm are reproduced by the 1(TT) spectrum. However, σTT cannot account for the excess emission in the red region, which must be due to excess excimer formation (compared with the steady-state spectrum). The TF spectrum exhibits a greater excimer contribution than observed in the ΔPL/PL spectrum. The differences suggest that the generation and recombination of geminate 1(T…T) pairs does not produce the same spectrum as TF from uncorrelated triplets. It may be hypothesized that the sites that are conducive to SF are those that are less susceptible to excimer formation than a randomly chosen chromophore pair. For instance, the nearest-neighbour sites identified in the MD simulations may act as SF sites that do not easily form excimers due to the hindered mutual rotation of the tetracene subunits. However, the parallel and anti-parallel pairs at longer distance may become engaged in TF through random hopping of triplet excitons and are geometrically predisposed to the formation of an excimer. These dimer geometries are depicted in Extended Data Fig. 2.
MD simulations find that similar dimer geometries occur at 50% w/w in toluene. Figure 4b shows remnants of the spectral features attributed to the 1(TT) state, showing that the 1(TT) state is in evidence in concentrated solutions, but is displayed to a lesser extent. At 30%, there is no longer any evidence of this emission (Extended Data Fig. 6). Interestingly, at low magnetic fields corresponding to the dip in Fig. 4a, a spectrum showing negative 1(TT) features is observed (Extended Data Fig. 7).
The shape of σTT is essentially that of a broadened and red-shifted S1 spectrum. The broadening can be attributed to a distribution of chromophore pair geometries, as predicted by the MD simulations (Supplementary Figs. 9–11). The energy of the 1(TT) state is estimated to be about 2.30 eV. Notwithstanding the triplet binding energy, this places the T1 state of the NODIPS-Tc chromophore at >1.15 eV, consistent with energy transfer experiments48. The energy of the excimer state was estimated from a van’t Hoff plot, shown in Extended Data Fig. 8. The S1 and 1(TT) emission is activated by 0.313(16) eV, placing the 1Ex state at 1.99 eV. The relative energies of these states are shown in Fig. 3b.
The results in this work underscore the complexity of the excited states in the most studied SF chromophore (tetracene). In addition to unifying the debate regarding the fundamental photophysics of SF and TF in solids and solutions, our results have many implications for technologies that exploit SF and TF. First, multi-exciton logic and magnetic resonance imaging are underpinned by an understanding of spin evolution and is affected by the equilibrium in Fig. 3a. Second, the development of TF-based light-emitting diodes necessitates an understanding of all the emissive states in the system. Finally, magnetic field dependent measurements are often used to characterize SF-based solar cells5,49,50. These analyses usually assume that the magnetic field perturbs the equilibrium between (S0S1 and T1 + T1), mediated by the singlet character of the 1(T…T) pair. However, as shown in Fig. 4, this is an oversimplication, as it does not consider the interplay between these states and the coupled-pair 1(TT) and 1Ex states.
Conclusions
In this study, we analysed the time-resolved emission of NODIPS-tetracene, a liquid-state SF and TF material. As seen in other systems, the photo-generated bright state generates excimers that emit with an unstructured, red-shifted spectrum. In the first few nanoseconds, there is a shift of the IEP, which settles down near 620 nm and remains in place at longer times, during the TF phase. This initial shift in the IEP indicates the presence of a third spectral component, which is confirmed by PCA.
Further insight into the spectral components generated by a triplet-pair is gained from magnetic field experiments. Photoluminescence enhancement is observed at particular wavelengths, as well as in the excimer spectrum, compared with the steady-state spectrum. The peaks in the ΔPL/PL spectrum are assigned to the 1(TT) state. MD simulations show that the neat chromophore behaves essentially as an amorphous solid on experimental timescales, but that the 50% mixture with toluene is dynamic. They both exhibit the same features in the magnetic luminescence attributed to the 1(TT) state, demonstrating a common SF mechanism in two very different states of matter.
These results highlight several features of SF and TF systems in general. They further show that the excimer state is not a necessary intermediate in SF or TF, but serves as a trap. Furthermore, it is shown that emission from the 1(TT) state can be observed at room temperature in non-solid state systems, and that this emission is enhanced by a magnetic field more so than the steady-state emission. In more dilute solutions of chromophore, these effects are less evident.
Methods
MD
MD simulations were carried out using the Large-scale Atomic/Molecular Massively Parallel Simulator51,52,53 (2 August 2023 version with graphic processing unit acceleration). Short-ranged non-bonded interactions were cut off at a distance of 11 Å, while the long-ranged part of electrostatic interactions were computed with the particle–particle particle–mesh method53,54. Bond lengths involving hydrogen atoms were constrained using the SHAKE algorithm55. Constant temperature and pressure were maintained using a Nosé–Hoover56,57 style thermostat and barostat58 in a cubic simulation box with periodic boundary conditions enforced in all dimensions.
Two systems were simulated: a pure NODIPS-Tc system containing 216 molecules and an approximately 50% w/w NODIPS-Tc/toluene mixture containing 108 NODIPS-Tc molecules and 855 toluene molecules. For each system, the initial configuration was built using Moltemplate59 (version 2.20.19), with molecules placed on a cubic grid at very low density to avoid overlapping molecules. The system energy was minimized and then a short (50–75 ps) simulation was carried out at constant volume and temperature at a a high temperature of 600 K, during which the simulation timestep was progressively increased to its final value. The simulation was then continued at constant pressure and temperature at 600 K and 100 atm for at least 1 ns to shrink the system to liquid density and to randomize the molecular positions and orientations. For the NODIPS-Tc/toluene mixture, the system was then simulated in the ensemble at 298 K and 1 atm. Owing to the low molecular mobility of pure NODIPS-Tc at room temperature, equilibration of this system at 1 atm was carried out in stages, with the temperature progressively decreased from 500 K to 450 K to 400 K to 350 K to 298 K, with sufficient time spent at each intermediate temperature to reach equilibrium. In addition, between the 350 K and 298 K simulations, the temperature was decreased at a constant rate over a period of 150 ns. Details of the simulations that were performed are given in Supplementary Table 6.
Chemical synthesis
NODIPS acetylene was prepared as previously reported60. See Supplementary Information for full details of the synthetic procedures60,61,62.
Sample preparation
All samples used for TRPL spectroscopy were measured under an inert atmosphere using a Teflon-stopcock sealed quartz cuvette with a 1 mm path length. Toluene was added for diluted samples.
Optical spectroscopy
TRPL was carried out using a SpitLight Picolo, ND:YVO4 laser system with pulse duration of 800 ps at a repetition rate of 1 kHz and wavelength of 532 nm. Photoluminescence was detected using a spectrometer (Princeton Instruments 2300i) equipped with a 300 groove-per-millimetre grating blazed at 500 nm, and an intensified time-gated camera (Princeton Instruments Pi Max-4).
Steady-state optical absorption spectroscopy was performed with a Cary 60 ultraviolet–visible–near infrared spectrometer covering a wavelength range of 190–1,100 nm.
For steady-state photoluminescence measurements, samples were excited by a 532 nm continuous-wave laser. The photoluminescence was collected and imaged onto a multi-mode optical fibre by two off-axis parabolic mirrors. An Ocean Optics Flame spectrometer was used to collect the photoluminescence spectrum. The magnetic field was provided by a benchtop electromagnet (Magnetech MFG-6-24) controlled by a d.c. power supply (Keithley 2230G-30-1).
PCA
PCA was used to extract the emissive species (PCs) contributing to the TRPL. These PCs form a new basis set and can be used to reconstruct the experimental data. First, the experimental TRPL dataset was organized as an m × n matrix X, where m is the number of measurement types, that is, dimensions (in this case the number of wavelengths) and n is the number of observations (in this case the number of time series). Second, the covariance matrix C = XXT was computed, where C is a square symmetric m × m matrix. The diagonal terms of C are the variance of each measurement type and the off-diagonal terms are the covariance between measurement types. The C matrix contains the correlations between all possible pairs of measurements. Last, eigendecomposition of the covariance matrix C was performed, which yields the eigenvectors of PCs and eigenvalues, that is, the variance of each PC.
Data availability
Data supporting Figs 1–4 are publicly available at https://doi.org/10.5061/dryad.05qfttf8r.
Code availability
Not applicable.
References
Smith, M. B. & Michl, J. Recent advances in singlet fission. Annu. Rev. Phys. Chem. 64, 361–386 (2013).
Feng, J., Alves, J., de Clercq, D. M. & Schmidt, T. W. Photochemical upconversion. Annu. Rev. Phys. Chem. 74, 145–168 (2023).
Schulze, T. F. & Schmidt, T. W. Photochemical upconversion: present status and prospects for its application to solar energy conversion. Energy Environ. Sci. 8, 103–125 (2015).
MacQueen, R. W. et al. Crystalline silicon solar cells with tetracene interlayers: the path to silicon-singlet fission heterojunction devices. Mater. Horiz. 5, 1065–1075 (2018).
Einzinger, M. et al. Sensitization of silicon by singlet exciton fission in tetracene. Nature 571, 90–94 (2019).
Ravetz, B. D. et al. Photoredox catalysis using infrared light via triplet fusion upconversion. Nature 565, 343–346 (2019).
Zhu, X., Su, Q., Feng, W. & Li, F. Anti-stokes shift luminescent materials for bio-applications. Chem. Soc. Rev. 46, 1025–1039 (2017).
Kawashima, Y. et al. Singlet fission as a polarized spin generator for dynamic nuclear polarization. Nat. Commun. 14, 1056 (2023).
Jacobberger, R. M., Qiu, Y., Williams, M. L., Krzyaniak, M. D. & Wasielewski, M. R. Using molecular design to enhance the coherence time of quintet multiexcitons generated by singlet fission in single crystals. J. Am. Chem. Soc. 144, 2276–2283 (2022).
Hudson, R. J. et al. A framework for multiexcitonic logic. Nat. Rev. Chem. 8, 136–151 (2024).
Sanders, S. N. et al. Triplet fusion upconversion nanocapsules for volumetric 3D printing. Nature 604, 474–478 (2022).
Tayebjee, M. J. Y., McCamey, D. R. & Schmidt, T. W. Beyond Shockley–Queisser: molecular approaches to high-efficiency photovoltaics. J. Phys. Chem. Lett. 6, 2367–2378 (2015).
Tayebjee, M. J. Y., Gray-Weale, A. A. & Schmidt, T. W. Thermodynamic limit of exciton fission solar cell efficiency. J. Phys. Chem. Lett. 3, 2749–2754 (2012).
Rao, A. & Friend, R. H. Harnessing singlet exciton fission to break the Shockley–Queisser limit. Nat. Rev. Mat. 2, 1–12 (2017).
Pensack, R. D. et al. Observation of two triplet-pair intermediates in singlet exciton fission. J. Phys. Chem. Lett. 7, 2370–2375 (2016).
Bakulin, A. A. et al. Real-time observation of multiexcitonic states in ultrafast singlet fission using coherent 2D electronic spectroscopy. Nat. Chem. 8, 16–23 (2016).
Monahan, N. R. et al. Dynamics of the triplet-pair state reveals the likely coexistence of coherent and incoherent singlet fission in crystalline hexacene. Nat. Chem. 9, 341–346 (2017).
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. 134, 8597–8607 (2012).
Dover, C. B. et al. Endothermic singlet fission is hindered by excimer formation. Nat. Chem. 10, 305–310 (2018).
Schmidt, T. W. A Marcus–Hush perspective on adiabatic singlet fission. J. Chem. Phys. 151, 054305 (2019).
Walker, B. J., Musser, A. J., Beljonne, D. & Friend, R. H. Singlet exciton fission in solution. Nat. Chem. 5, 1019–1024 (2013).
Stern, H. L. et al. Identification of a triplet pair intermediate in singlet exciton fission in solution. Proc. Natl Acad. Sci. USA 112, 7656–7661 (2015).
Dvořák, M. et al. Singlet fission in concentrated tips-pentacene solutions: the role of excimers and aggregates. J. Am. Chem. Soc. 143, 13749–13758 (2021).
Bossanyi, D. G. et al. Emissive spin-0 triplet-pairs are a direct product of triplet–triplet annihilation in pentacene single crystals and anthradithiophene films. Nat. Chem. 13, 163–171 (2021).
Ryerson, J. L. et al. Two thin film polymorphs of the singlet fission compound 1,3-diphenylisobenzofuran. J. Phys. Chem. C 118, 12121–12132 (2014).
Schrauben, J. N., Ryerson, J. L., Michl, J. & Johnson, J. C. Mechanism of singlet fission in thin films of 1,3-diphenylisobenzofuran. J. Am. Chem. Soc. 136, 7363–7373 (2014).
Feng, X. & Krylov, A. I. On couplings and excimers: lessons from studies of singlet fission in covalently linked tetracene dimers. Phys. Chem. Chem. Phys. 18, 7751–7761 (2016).
Mauck, C. M. et al. Singlet fission via an excimer-like intermediate in 3,6-bis(thiophen-2-yl)diketopyrrolopyrrole derivatives. J. Am. Chem. Soc. 138, 11749–11761 (2016).
Burdett, J. J., Gosztola, D. & Bardeen, C. J. The dependence of singlet exciton relaxation on excitation density and temperature in polycrystalline tetracene thin films: kinetic evidence for a dark intermediate state and implications for singlet fission. J. Chem. Phys. 135, 214508 (2011).
Tayebjee, M. J., Clady, R. G. & Schmidt, T. W. The exciton dynamics in tetracene thin films. Phys. Chem. Chem. Phys. 15, 14797–14805 (2013).
Grieco, C. et al. Direct observation of correlated triplet pair dynamics during singlet fission using ultrafast mid-IR spectroscopy. J. Phys. Chem. C 122, 2012–2022 (2018).
Sanders, S. N. et al. Understanding the bound triplet-pair state in singlet fission. Chem 5, 1988–2005 (2019).
Hausch, J. et al. Distinguishing between triplet-pair state and excimer emission in singlet fission chromophores using mixed thin films. J. Phys. Chem. C 126, 6686–6693 (2022).
Miyata, K., Conrad-Burton, F. S., Geyer, F. L. & Zhu, X.-Y. Triplet pair states in singlet fission. Chem. Rev. 119, 4261–4292 (2019).
Yong, C. K. et al. The entangled triplet pair state in acene and heteroacene materials. Nat. Commun. 8, 15953 (2017).
Lukman, S. et al. Efficient singlet fission and triplet-pair emission in a family of zethrene diradicaloids. J. Am. Chem. Soc. 139, 18376–18385 (2017).
Merrifield, R., Avakian, P. & Groff, R. Fission of singlet excitons into pairs of triplet excitons in tetracene crystals. Chem. Phys. Lett. 3, 155–157 (1969).
Wakasa, M. et al. What can be learned from magnetic field effects on singlet fission: role of exchange interaction in excited triplet pairs. J. Phys. Chem. C 119, 25840–25844 (2015).
Hestand, N. J. & Spano, F. C. Expanded theory of H- and J-molecular aggregates: the effects of vibronic coupling and intermolecular charge transfer. Chem. Rev. 118, 7069–7163 (2018).
Bayliss, S. L. et al. Geminate and nongeminate recombination of triplet excitons formed by singlet fission. Phys. Rev. Lett. 112, 238701 (2014).
Cattell, R. B. The scree test for the number of factors. Multivar. Behav. Res. 1, 245–276 (1966).
Musser, A. J. & Clark, J. Triplet-pair states in organic semiconductors. Ann. Rev. Phys. Chem. 70, 323–351 (2019).
Stern, H. L. et al. Vibronically coherent ultrafast triplet-pair formation and subsequent thermally activated dissociation control efficient endothermic singlet fission. Nat. Chem. 9, 1205–1212 (2017).
Johnson, R. C. & Merrifield, R. E. Effects of magnetic fields on the mutual annihilation of triplet excitons in anthracene crystals. Phys. Rev. B 1, 896–902 (1970).
Johnson, R. C., Merrifield, R. E., Avakian, P. & Flippen, R. B. Effects of magnetic fields on the mutual annihilation of triplet excitons in molecular crystals. Phys. Rev. Lett. 19, 285–287 (1967).
Merrifield, R. E. Theory of magnetic field effects on the mutual annihilation of triplet excitons. J. Chem. Phys. 48, 4318–4319 (1968).
Tapping, P. C. & Huang, D. M. Comment on ‘magnetic field effects on singlet fission and fluorescence decay dynamics in amorphous rubrene’. J. Phys. Chem. C 120, 25151–25157 (2016).
Gray, V. et al. Thiol-anchored TIPS-tetracene ligands with quantitative triplet energy transfer to pbs quantum dots and improved thermal stability. J. Phys. Chem. Lett. 11, 7239–7244 (2020).
Daiber, B. et al. Change in tetracene polymorphism facilitates triplet transfer in singlet fission-sensitized silicon solar cells. J. Phys. Chem. Lett. 11, 8703–8709 (2020).
Congreve, D. N. et al. External quantum efficiency above 100% in a singlet-exciton-fission-based organic photovoltaic cell. Science 340, 334–337 (2013).
Plimpton, S. J. Fast parallel algorithms for short-range molecular dynamics. J. Comput. Phys. 117, 1–19 (1995).
Brown, W. M., Wang, P., Plimpton, S. J. & Tharrington, A. N. Implementing molecular dynamics on hybrid high performance computers—short range forces. Comput. Phys. Commun. 182, 898–911 (2011).
Brown, W. M., Kohlmeyer, A., Plimpton, S. J. & Tharrington, A. N. Implementing molecular dynamics on hybrid high performance computers—particle–particle particle-mesh. Comput. Phys. Commun. 183, 449–459 (2012).
Hockney, R. W. & Eastwood, J. W. Computer Simulation Using Particles (Taylor & Francis, 1988).
Ryckaert, J. P., Ciccotti, G. & Berendsen, H. J. C. Numerical integration of the cartesian equations of motion of a system with constraints: molecular dynamics of n-alkanes. J. Comput. Phys. 23, 327–341 (1977).
Nosé, S. A molecular dynamics method for simulations in the canonical ensemble. Mol. Phys. 52, 255–268 (1984).
Hoover, W. G. Canonical dynamics—equilibrium phase-space distributions. Phys. Rev. A 31, 1695–1697 (1985).
Shinoda, W., Shiga, M. & Mikami, M. Rapid estimation of elastic constants by molecular dynamics simulation under constant stress. Phys. Rev. B 69, 134103 (2004).
Jewett, A. I. et al. Moltemplate: a tool for coarse-grained modeling of complex biological matter and soft condensed matter physics. J. Mol. Biol. 433, 166841 (2021).
Kumarasamy, E. et al. Properties of poly- and oligopentacenes synthesized from modular building blocks. Macromol. 49, 1279–1285 (2016).
Jennings, W. B. Chemical shift nonequivalence in prochiral groups. Chem. Rev. 75, 307–322 (1975).
McFarlane, W. Chemical shift and coupling constant nonequivalence of isopropyl methyl protons in a tertiary phosphine. Chem. Commun. 229–230 (1968).
Acknowledgements
This work was supported by the Australian Research Council (Centre of Excellence in Exciton Science CE170100026 (J.H.C. and T.W.S.) and FT230100002 (M.J.Y.T.)) and used computational resources provided by the University of Adelaide’s Phoenix High-Performance Computing service (D.M.H.).
Funding
Open access funding provided through UNSW Library
Author information
Authors and Affiliations
Contributions
J.F., P.H., D.M.d.C., B.P.C. and S.K.K.P. performed the measurements. T.W.S., J.F., P.H., M.J.Y.T., J.H.C., N.J.E.-D., M.P.N., B.P.C. and D.M.d.C. analysed the experimental results. M.W.B., A.A.D.F., S.N.S., J.E.B., H.L.L. and P.T. synthesized and analysed the NODIPS-Tc material. D.M.H. performed and analysed MDs simulations. J.F. and D.M.d.C. designed the figures. T.W.S. wrote the manuscript.
Corresponding author
Ethics declarations
Competing interests
The authors declare no competing interests.
Peer review
Peer review information
Nature Chemistry thanks Yasuhiro Kobori and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Extended data
Extended Data Fig. 1 Diffusion coefficients from molecular dynamics.
(a) Translational diffusion coefficient of the center-of-mass and (b) rotational diffusion coefficient of each principal axis (short, intermediate, and long) of the tetracene carbon backbone versus temperature for neat NODIPS-Tc (filled symbols) and 50% w/w NODIPS-Tc/toluene (empty symbols).
Extended Data Fig. 2 Typical π-stacked dimer geometries sampled from the MD simulations corresponding to peaks in the ADRFs.
Parameters correspond to centre-of-mass displacement and angle between long-axes. (a) 3.8 Å (60°) (b) 4.4 Å (90°) (c) 5.6 Å (180°) (d) 6.5 Å (0°).
Extended Data Fig. 3 Steady-state photoluminescence spectra of NODIPS-Tc, neat and in toluene at various % w/w concentrations.
Samples measured with 532 nm notch filter.
Extended Data Fig. 4 Effect of self-absorption.
Steady-state photoluminescence spectrum (black line) of NODIPS-Tc in a solution of toluene at a concentration of 0.01 mg/mL (14 μM) and calculated self-absorbed spectra at excitation wavelengths of 532 nm (red) and 480 nm (blue).
Extended Data Fig. 5 Principal component analysis of the TRPL of dilute solutions.
a PCs for the 75% w/w sample with PC3 showing signs of being diminished. b Scree plot of the 75% w/w sample. c PCs of the 50% w/w sample accompanied with d a scree plot. e PCs of the 30% w/w sample. At this dilution the PC3 is no longer relevant. This is further seen in f, the scree plot, which shows a clear elbow at PC3.
Extended Data Fig. 6 Plots of various concentrations of spectrally resolved MPL at a magnetic field of 210 mT.
a Shows the neat chromophore, which exhibits peaks at 560 nm and 595 nm. Theses peaks become less pronounced for b 75% w/w and c 50% w/w. In d, for the 30% w/w sample, these peaks are no longer present. The error bars in the plots represent the standard error of the mean after 20 repeat measurements. Error arises due to small fluctuations in the laser diode power.
Extended Data Fig. 7 Wavelength dependence of the magnetic-field effect on the PL of NODIPS-Tc at 210 mT and 20 mT magnetic field.
The vertical axis is expressed in the percentage PL change (ΔPL/PL). For comparison, we include a best-fit combination of σTT and σEx ratioed to the steady-state spectrum from the main text (red). The peaks at 560 nm and 595 nm are dips at 20 mT, where the integrated MPL is negative.
Extended Data Fig. 8 Plot of the logarithm of the ratio of the sum of S1 and 1(TT) components to the 1Ex component against inverse temperature for the spectra of Epoch 2.
The linear fit indicates that the excimer is 0.31 eV below the other singlet states. The intercept is not meaningful without information on the relative radiative rates of these states.
Supplementary information
Supplementary Information
Supplementary Text Sections 1–7, Figs. 1–19, Tables 1–6 and references.
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
About this article
Cite this article
Feng, J., Hosseinabadi, P., de Clercq, D.M. et al. Magnetic fields reveal signatures of triplet-pair multi-exciton photoluminescence in singlet fission. Nat. Chem. (2024). https://doi.org/10.1038/s41557-024-01591-0
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41557-024-01591-0