
Lowering of the singlet-triplet energy gap via intramolecular exciton-exciton coupling
- Select a language for the TTS:
- UK English Female
- UK English Male
- US English Female
- US English Male
- Australian Female
- Australian Male
- Language selected: (auto detect) - EN
Play all audios:

ABSTRACT Organic dyes typically have electronically excited states of both singlet and triplet multiplicity. Controlling the energy difference between these states is a key factor for making
efficient organic light emitting diodes and triplet sensitizers, which fulfill essential functions in chemistry, physics, and medicine. Here, we propose a strategy to shift the singlet
excited state of a known sensitizer to lower energies without shifting the energy of the triplet state, thus without compromising the ability of the sensitizer to do work. We covalently
connect two to four sensitizers in such a way that their transition dipole moments are aligned in a head-to-tail fashion, but, through steric encumbrance, the delocalization is minimized
between each moiety. Exciton coupling between the singlet excited states considerably lowers the first excited singlet state energy. However, the energy of the lowest triplet excited state
is unperturbed because the exciton coupling strength depends on the magnitude of the transition dipole moments, which for triplets are very small. We expect that the presented strategy of
designed intramolecular exciton coupling will be a useful concept in the design of both photosensitizers and emitters for organic light emitting diodes as both benefits from a small
singlet-triplet energy gap. SIMILAR CONTENT BEING VIEWED BY OTHERS HIGHLY EFFICIENT LUMINESCENCE FROM SPACE-CONFINED CHARGE-TRANSFER EMITTERS Article 15 June 2020 HYBRIDIZATION OF
SHORT-RANGE AND LONG-RANGE CHARGE TRANSFER EXCITED STATES IN MULTIPLE RESONANCE EMITTER Article Open access 09 August 2023 TOWARDS EFFICIENT NEAR-INFRARED FLUORESCENT ORGANIC LIGHT-EMITTING
DIODES Article Open access 21 January 2021 INTRODUCTION Triplet photosensitizers (PS) are crucial for photocatalysis1,2,3,4, triplet-triplet annihilation upconversion5,6,7, photodynamic
therapy8,9,10,11, and various other fields and applications. New avenues of utilizing PS are frequently discovered and with the growing interest, the development of organic, highly efficient
PS has gained more attention. The crucial property of a triplet PS is its rate of intersystem crossing (ISC). ISC is a spin forbidden process which can be realized by magnetic
perturbations, like spin-orbit coupling12. To achieve a high rate of ISC, traditional PS make use of the heavy atom effect, known to enhance spin-orbit coupling13,14. Heavy elements like Pt,
Pd, Ru, Ir, Se, I and Br have shown to enhance ISC15,16. However, there are considerable disadvantages of using heavy atoms to increase the rate of ISC, such as their scarceness, toxicity,
and cost. Furthermore, the increased spin-orbit coupling also reduces the triplet state lifetime, which is a negative feature when used in PS applications17. To avoid some of these
drawbacks, development and improvement of organic, heavy atom free triplet PS is in high demand. Several strategies have been developed to realize an enhanced rate of ISC for compounds
without incorporating heavy atoms. One that has gained recognition is spin-orbit charge transfer intersystem crossing (SOCT-ISC)9,18,19. Systems exhibiting this phenomenon are usually
donor-acceptor dyads able to perform intramolecular excited state electron transfer reactions. Upon photoinduced electron transfer, the system forms a charge separated (CS) state. The donor
and acceptor moieties are designed to have an orthogonal geometry towards each other. Due to the orthogonality of the system, the change in molecular orbital angular momentum compensates for
the change in spin angular momentum, enhancing the rate of ISC20. Thus, from the CS state the molecule can undergo charge recombination to the triplet state in an efficient manner. This
approach does not rely on the heavy atom effect and is therefore highly appropriate for the design of heavy atom free photosensitizers. Recently, several triplet photosensitizers undergoing
SOCT-ISC based on the BODIPY scaffold have been reported5,8,21,22,23. Within the scope of these reports, a BODIPY-anthracene dyad stands out with a high yield of ISC24,25,26. These studies
also highlight the importance of the energy alignment between the S1 and CS states. The energy of the CS state is strongly dependent on the polarity of the solvent, and the S1-CS energy
alignment can therefore be tuned by the choice of solvent. In a non-polar solvent (toluene), the CS state was at a higher energy than the S1 state, resulting in a neglectable ISC but a high
yield of fluorescence (ФFluo = 0.81). However, with increasing solvent polarity, the CS state is stabilized, resulting in exergonic excited state electron transfer. The yield of ISC thus
increased (ФISC = 0.92) at the expense of the fluorescence (ФFluo = 0.002)24. Organic dyes such as BODIPY’s have a relatively large S1-T1 energy gap. In all applications relying on a triplet
sensitizer, this gap can be regarded as an energy loss. As small an energy gap as possible is thus preferable. In recent years, the field of organic electronics has been focusing on
reducing this energy gap by separating the HOMO and LUMO orbitals in space as doing so affects the electron exchange interaction12,27,28,29. However, there are other possible mechanisms that
theoretically can perturb the S1-T1 energy alignment30,31. One of these is exciton coupling, the effect seen in J-aggregates32. Exciton coupling influences the singlet state energy without
affecting either the triplet or CS state energies significantly. Even before the development of the theory of exciton coupling it was observed that aggregation affects the fluorescence and
phosphorescence differently33. Curiosity about this observation amongst other things, later lead to the finding that exciton coupling influences the singlet state energy without affecting
either the triplet or CS state energies significantly34. This is because the effect is based on a Columbic coupling of the transition dipole moments associated with the respective states,
and only singlet states have an appreciable magnitude of their transition dipole moments35. Thus, from a theoretical perspective, assembling an organic PS in the form of a J-aggregate, will
reduce energy losses by allowing a red-shifted excitation wavelength. Here, BODIPY-anthracene dyads were β-tethered into oligomers (2–4 units) in a non-conjugated fashion. The β-tethering
aligns the S0-S1 transition dipole moments of the BODIPYs in a ‘head-to-tail’ fashion. Intramolecular exciton coupling thus occurs, giving an allowed low energy transition that depends on
the number of coupled BODIPY-anthracene moieties. This is the same effect as seen in J-aggregates and has also previously been observed in ethylene-bridged BODIPY oligomers36,37. The
β-tethered oligomers have almost unperturbed energies of their T1 and CS states because direct electronic communication between the moieties was designed to be suppressed. However, they are
still able to perform ISC. The formation of covalently tethered J-aggregate mimics is thus a viable path to reduce energy losses in organic sensitizers. This by selectively lowering the S1
state without significantly influencing the energy of the triplet state. RESULTS To lower the S1 energy without affecting the CS and T1 energies, it is important to align the S0-S1
transition dipole moments without causing an extension of the aromatic system in oligomers. This can be achieved if steric hindrance forces the individual moieties into an orthogonal
geometry. The BODIPY anthracene dyad was synthesized following the literature procedure38. Dimerization of BODIPY dyes has previously been achieved in the meso, α, and β positions39,40,41.
Of these attachment positions, β-tethering also leads to oligomerization of BODIPY dyes42,43,44. Both conjugated oligomers42 as well as non-conjugated oligomers, where methyl groups on
adjacent carbons result in a large dihedral angle between moieties due to steric hindrance43,44, have been accessed. The conjugated BODIPY oligomers were obtained via Suzuki-Miyaura
coupling. The benefit of this reaction is the controllability of the degree of oligomerization. However, a fully conjugated system would not just influence the singlet energy, but also the
triplet energy to a similar extent. The non-conjugated systems were obtained by treating the monomer with either the hypervalent iodine reagent phenyliodine(III)bistrifluoroacetate (PIFA),
or anhydrous iron(III)trichloride (Fig. 1). In both cases, trimers were the highest order of oligomers isolated43,44. However, polymerization was achieved using FeCl3, when leaving the
reaction for a longer time. Both conditions were tested here, resulting in the dimer regardless of conditions used (Supplementary Note 1, 2, Supplementary Figs. 1–19, Supplementary Tables 1,
2). The reaction using FeCl3 yielded the dimer to a larger extent, 17% as compared to 8% when using PIFA. Furthermore, the reaction using FeCl3 also yielded higher oligomers, the trimer,
and the tetramer. Atropisomers of the trimer could be isolated. These are hypothesized to correspond to the cis and trans isomers with regards to the anthracene units on the end BODIPYs and
show discernible 1H NMR spectra (Supplementary Figs. 11, 14; unfortunately, the distance in between the anthracene units is too large to obtain NOESY signals, which would allow for
identification of the two different isomers). After purification, only trace amounts, 1–2% of the two atropisomers of the trimer, and >1% of the tetramer were obtained. However, the small
amounts collected were enough to confirm their structure through 1H, 13C (Supplementary Dataset 1), diffusion NMR and MALDI, and to probe the excited state energetics and dynamics. THE
ENERGIES OF THE SINGLET AND TRIPLET STATES In the presentation of the excited state energetics, we start by discussing the S1 state. We then continue with the T1 and the CS states, so that
we in the conclusion can discuss the complete energetic picture of our systems. To obtain information on the S1 state energy, and to see if it could be decreased successfully, the absorption
spectra of the oligomers were analyzed. The two conformers of the trimer exhibit identical photophysical properties. In this discussion, the results are therefore combined as the trimer.
The absorption envelope (Fig. 2a) of all compounds can be deconvoluted into an anthracene part (370–400 nm) and a BODIPY part (500–600 nm). The wavelength (i.e., energy) of the anthracene
absorption is not affected by oligomerization. This means that the anthracene moieties are located far enough apart, and/or the orientation between them is disadvantageous, for Columbic
interactions between the transition dipole moments to become significant. The situation for the BODIPY part, where the transition dipole moments are designed to be in a head-to-tail fashion,
is quite different. The absorption of the oligomers is shifted to longer wavelengths, with absorption maxima at 506, 542, 569, and 583 nm for the monomer, dimer, trimer, and tetramer,
respectively. As expected, the absorption energy converges as the number of moieties increases. The excitation energy thus decreases asymptotically, reducing the gain in making even longer
oligomers (Fig. 2b). Furthermore, the magnitude of the transition dipole moments were calculated from the molar absorptivity spectra (Supplementary Note 3, Supplementary Fig. 20)45. As
expected for an exciton coupled system, the transition dipole moments scale well with the square root of the number of moieties (Fig. 2c)35, although limited purity of the tetramer
(Supplementary Fig. 17) is exhibited as a less than expected transition dipole moment for that molecule. Previously reported conjugated BODIPY dimers and trimers, show a different trend than
we observe here. Much larger bathochromic shifts have been reported for conjugated oligomers42. The different observations indicate different main communication pathways for conjugated and
non-conjugated oligomers. In the latter, the lack of conjugation in between the monomer units minimizes electron delocalization. Exciton coupling could however explain the observed spectral
changes, as it does not involve conjugation between moieties46. Using a point dipole approximation, the Coulombic interaction energies between moieties in the oligomers were calculated
(Supplementary Note 4). Based on this, the spectral shift solely due to Coulombic interactions could be extracted. The vast majority of the experimentally observed spectral shift can be
reproduced by taking Coulombic interactions into account. Thus, the observed spectral changes are mainly a consequence of exciton coupling rather than electronic delocalization. Quantum
mechanical calculations were performed to validate the degree to which the triplet energy remains constant across the series. The geometry of singlet and triplet ground states were optimized
in vacuum at the ωB97XD/6-31 g(d) level of theory (Supplementary Fig. 21, Supplementary Dataset 2) and singlet excitation energies were obtained through TD-DFT using the singlet ground
state geometry using Gaussian 16, revision B.0147,48. No imaginary frequencies for the optimized ground state geometry verifies that a minimized structure was obtained. Although the overall
trend is captured well, these calculations overestimate the reduction of the S1 energy as well as the increase of the transition dipole moment with oligomer length (Fig. 2b, c, Supplementary
Table 3). But, more importantly, the calculations predict that the energy of the T1 states only marginally decrease across the monomer-to-tetramer series (Fig. 2b). EMISSION EFFICIENCY AND
DYNAMICS We will now turn our attention to the energy relaxation pathways towards the charge separated state after exciting the oligomers. Figure 3 shows the steady state emission spectra of
the monomer and oligomers in toluene (for other solvents and oligomers, see Supplementary Figs. 22–26). Clear BODIPY like emission envelopes that follow the same energetic trend as the
absorption spectra can be seen. However, when increasing the polarity of the solvent, a distinguishable second band is observed. This has, for the monomer, previously been ascribed to
emission from the charge separated state24. The emission from the CS state is intense and spectrally well resolved for the monomer, but it is also present to a minor degree in the oligomers
as a shoulder at lower energies in the high polar solvents (Fig. 3). It is difficult to draw any conclusions regarding the efficiency of reaching the CS state based on the magnitude of the
CS emission. It is better to compare how well the presence of an energetically accessible CS state quenches the fluorescence (Table 1). The fluorescence quantum yield of the monomer shows a
drop from 0.80 in toluene to 0.02 in DCM and 0.01 in ACN. This effect can be explained by the polarity and thus the dielectric constants of the solvents. The Rehm-Weller equation predicts
that the energy of the CS state lowers as the dielectric constant increases15,49. ACN has a larger dielectric constant than DCM and toluene. Therefore, the CS state in ACN is lower than in
DCM, which is lower than in toluene. The driving force for populating the CS state is thus larger in ACN and DCM, explaining the rise of the second emission band in these solvents. The drop
in ФEm is less intense for the dimer, trimer and tetramer in DCM, but it is still considerable in ACN. From a driving force perspective, this difference indicates that the energy of the
singlet state decreases more than the CS state in the oligomers. Thus, the energy difference between the S1 and CS states only remains large enough in ACN for charge separation to outcompete
fluorescence in the longer oligomers. To evaluate the efficiency of internal conversion from higher excited states towards the S1 state, different excitation energies were used to excite
the monomer. Excitation of the BODIPY moiety is described above. Upon excitation at higher energies, at the anthracene absorption, the emission spectrum looks almost identical (Supplementary
Fig. 22). The only difference is seen in the relative emission intensity between the S1 and the CS states emission bands. When the anthracene unit is excited, the CS state emission is
relatively increased. This indicates that charge separation competes with Sn to S1 internal conversion in these systems. The fluorescence quantum yield of the monomer is greatly reduced in
ACN compared to toluene. However, the observed emission lifetime is not reduced to the same extent (Table 1, see Supplementary Figs. 27–33 for emission decays). This surprising feature has
been previously observed24, and we will try to rationalize it. When recording the emission lifetime for the monomer, a long-lived transient as well as an instrument response function limited
signal is present when monitoring the S1 but not the CS state emission (Supplementary Figs. 27, 29). The rate of charge separation is thus faster than the time resolution of our
instrumentation. The long-lived emission must therefore not be due to the directly excited BODIPY unit. Furthermore, when observing the steady state emission as a function of decreasing
temperature (Supplementary Fig. 34), we see a reduction of the S1 emission until a plateau value is reached. Based on these two observations, we suggest that the plateau value represent
prompt emission from the S1 state, and the temperature dependent emission represents endothermic charge recombination from the CS state back to the S1 state. To strengthen this hypothesis,
the system was modeled using rate equations (Supplementary Note 5). In the modeling, it was assumed that the S1 and CS states are connected by a microscopic reversibility, implying that the
ratio of the rate constants between them being equal to the Boltzmann factor. The modeled system thus strongly resembles that used when modeling molecules exhibiting E-type delayed
fluorescence (with the triplet state then being replaced by a CS state). Using the energy difference between the two states as a fitting parameter, both the steady state and time resolved
temperature dependent emission could be qualitatively reproduced (Supplementary Fig. 35). Furthermore, the fitted energy difference between the S1 and CS states was 12 kJ/mol (this
corresponds to a 28 nm shift compared to the S1 energy), which is well in line with the emission data (Fig. 3). It is thus very plausible that the observed long lifetime represents the
lifetime of the CS state, illustrating the complex dynamics in these systems. ENERGY OF THE CS STATE To investigate whether the energy of the CS state is influenced by oligomerization, the
redox potentials of the monomer and the dimer were determined in DCM (Supplementary Note 6). Limited solubility in ACN prevented experimental determination of the CS state energy, and we
therefore relied on the Rehm-Weller equation for this value. The monomer shows a reversible one-electron oxidation and reduction at +0.71 V and −1.73 V vs. Fc/Fc+, respectively in DCM
(Supplementary Fig. 36a, Supplementary Table 4). These values correlate very well with the previously reported redox potentials of the monomer24, where the reduction was assigned to the
BODIPY and oxidation to the anthracene moieties. The dimer in DCM shows two reversible one-electron waves for reduction (Supplementary Fig. 36b, Supplementary Table 4). The first reduction
signal at −1.71 V is almost identical to the one observed for the monomer. This agreement indicates that the reduction of the monomeric unit of the dimer is almost identical to that in the
monomer. In contrast to the reduction, only one oxidation potential for the dimer was observed close to the end of the solvent window. A small shift of 80 meV for the oxidation potential was
seen. It is unlikely that this is caused by anthracene-anthracene interactions, because of their large distance to each other. However, the oxidation peak is not well resolved, and the
error margins are therefore large. Nevertheless, the change in S1 energy is notably larger compared to the change in CS energy (about 160 meV vs 100 meV), and the CS energy change due to
oligomerization is small compared to the effect of solvent polarity. It should be noted here that limited solubility of the trimer and tetramer prevented cyclic voltammetry to be conducted.
It is interesting to compare the energy of the CS state measured electrochemically and spectroscopically. The energy of the electrochemically determined CS state is summarized in
Supplementary Table 5 and shown schematically in Fig. 3 as a vertical line for the monomer and dimer. It is located at a higher energy as compared to the CS state emission maximum, which is
reasonable. Thus, the electrochemically and spectroscopically measured energy of the CS state corroborate each other. In summary, the energy of the CS state is less affected in comparison to
the changes seen for the S1 state by oligomerization. Furthermore, it can be tuned separately by the solvent polarity. FORMATION OF THE TRIPLET STATE The ability of the monomer, dimer,
trimer, and tetramer to relax from the CS state to the T1 state was assessed by nanosecond transient absorption. The monomer is known to form the triplet state in high yield in both DCM and
ACN. Indeed, after excitation to the S1 state, all compounds show both a ground state bleach, as well as an excited state absorption in degassed DCM as well as ACN solutions (Fig. 4a, b).
Furthermore, as oxygen is added, all transient absorption features disappear (Supplementary Figs. 37–41). This indicates that all the oligomers populate the triplet state in these solvents.
The kinetics of the recovery to the ground state was fitted well assuming first order kinetics, resulting in lifetimes of around 150 μs for the monomer and generally longer for the oligomers
(Table 1, Fig. 4c, d). The triplet lifetime is on the same order of magnitude as iodinated BODIPY50, and several metal porphyrins. However, what stand out is that the triplet lifetime does
not decrease with the number of moieties in the series, despite that the S1 energy decreases. If the reduction of the singlet energy also reduced the triplet energy, a reduction of the
triplet lifetime would have been expected from the energy gap law51,52,53. Thus, this feature can be explained by the change in singlet energy being decoupled from the triplet energy. The
yield of intersystem crossing for the monomer is known, and it was therefore used as a reference when determining the yield of intersystem crossing for the oligomers. Figures 4c, d show
decays in DCM and ACN, respectively, using absorption and excitation fluence matched conditions (low solubility of the trimer prevented such experiment in ACN). Under such conditions, the
yield of intersystem crossing can be determined by comparing the ground state bleach shortly after the excitation pulse. In DCM, the yield of intersystem crossing reduces through the series
(Table 1). This is an expected result as the driving force for forming the CS state is continuously reduced through the series. Still, it is rewarding to simulate the yields of emission and
intersystem crossing from known measurables in order to gain a thorough understanding of the observed photophysics in the series. A rate equation model was therefore constructed
(Supplementary Note 7). It is based on a three-level system, the S1, CS, and T1 states, of which the S1 and CS states are in dynamic equilibrium with each other (Fig. 5a), where the
individual rate constants are shown in Supplementary Note 7 and Supplementary Tables 6, 7. Figure 5b shows the simulated and measured quantum yields of fluorescence and intersystem crossing
in DCM. The trends in oligomer lengths are captured very well, with a first increasing and then decreasing yield of fluorescence, and a continuously decreasing yield of intersystem crossing.
The dependence of the fluorescence yield on oligomer length can be explained by the balance between a decreased driving force for charge separation and an increased non-radiative decay to
the ground state. The energetics of the charge separation goes from exothermic for the monomer to endothermic for the oligomers. The result being that the equilibrium concentrations are
shifted towards the S1 state with an increased number of moieties in the oligomer. The energetics also explain why a very fast decay component in the TCSPC experiments only could be seen for
the monomer (vide supra). Counteracting the shift in the dynamic equilibrium concentration is the non-radiative decay to the ground-state, which increases throughout the series. That
internal conversion increases with a reduced excitation energy is commonly observed and is usually explained by the energy gap law51,52. We note, however, that for J-aggregates there are
examples where a reduced reorganization energy counteracts the effect of the energy gap law32,54,55. The simulations also capture the yields of fluorescence and intersystem crossing in ACN
solution very well. Here, a high driving force for charge separation results in a low yield of fluorescence throughout the series. A slight increase in emission yield with oligomer length is
observed, which can be explained from a reduction in the driving force with oligomer length. Overall, the trends of the fluorescence and intersystem crossing yields with oligomer length is
quite straightforwardly explained by the driving force for charge separation. Thus, highlighting the importance of the delicate balance of the energetics between the S1, CS and T1 states for
the SOCT-ISC process. Optimization of these energetics could for instance be done by changing the oxidation potential of the anthracene moiety through chemical modifications or by changing
the solvent. To demonstrate that solvent polarity can be used as an optimization tool to increase the yield of triplet formation, measurements were conducted on the dimer in benzonitrile, a
solvent with a polarity (ε = 25.9) between that of DCM and ACN. As expected, the dimer exhibited an emission quantum yield (29%) that was intermediate between those observed in DCM and ACN.
Additionally, the triplet quantum yield (24%; Supplementary Fig. 43), increased by 41% and 71% compared to DCM and ACN, respectively. Also, although our simulations capture the overall trend
well, the absolute values are less well explained. We here note that it is possible to perform excellent fits with our model, by for instance using the energy of the CS state, the rate of
charge recombination to the T1 state, or the rate of charge separation as fitting parameters. However, as we do not know which of these parameters changes most, we stop at the point where
approximated rates fit the overall trends in yields well. The yields of intersystem crossing and fluorescence are relatively low for the oligomers in ACN. This is attributed to an increase
in the non-radiative rate constant in ACN for the CS state, which is well explained by Marcus theory for electron transfer. To investigate the rate changes throughout the series, we
conducted femtosecond transient absorption experiments (which corroborated the time-resolved emission experiments) and analyzed the photophysics using Marcus theory (Supplementary Note 8,
Supplementary Figs. 44–46). The initial formation of the CS state as well as its relaxation to the triplet and ground states are electron transfer processes. The rate of electron transfer
reactions normally increase steeply with the driving force until the driving force equals the reorganization energy (so-called Marcus normal region). With further increase in the driving
force, the rate decreases sharply (so-called Marcus inverted region). Redox potentials for the monomer and dimer could be obtained (vide infra), and these were used to analyse the rate for
charge recombination to the triplet and ground states. The obtained rates align with the predictions from Marcus theory (Supplementary Fig. 46). The rate of recombination to the triplet
state increases with increasing driving force, with the monomer exhibiting a larger driving force and consequently a higher rate. Conversely, the rate of recombination to the ground state
decreases with increasing driving force, meaning the monomer, having a larger driving force, shows a slower recombination rate to the ground state. This combination gives the monomer the
prerequisites for forming triplets more efficiently than the oligomers, explaining the short lifetime observed for the CS state of the oligomers. Thus, the increase in the non-radiative rate
with oligomer length is not a consequence of oligomerization as such, but rather of the mechanism through which the triplet state is reached. Furthermore, the recombination from the CS to
the ground state is likely to be in the inverted Marcus regime for any molecular system. High efficiencies for the SOCT-ISC mechanism are therefore only feasible at considerable CS state
energies because the unwanted recombination to the ground state becomes slower the higher the CS state energy. Thus, limiting this mechanism to energies roughly equaling the visible part of
the electromagnetic spectrum. It is important to note that the low yields, attributed to fast ground state recombination in the oligomers, could have other explanations beyond the one
suggested by the analysis based on Marcus theory. However, investigating these alternative explanations is beyond the scope of this work. DISCUSSIONS Exciton coupling is shown to selectively
stabilize the first excited singlet state energy, without significantly perturbing the energy of the first excited triplet state. This approach thus enables a route to significantly reduce
energy losses for triplet sensitizers. We show this in a so-called SOCT-ISC sensitizer by synthesizing a set of new oligomeric BODIPY-anthracene dyads. The design principles in the coupling
of dyad moieties focused on the transition dipole moments being aligned in a head-to-tail fashion, but steric hindrance breaking the conjugated system. The oligomers show a gradual lowering
of the S1 energy with length (Fig. 6). The dimer shows a shift of 36 nm, the trimer 63 nm and the tetramer 77 nm compared to the monomer. This, as well as the increasing molar absorptivity
with oligomer length indicates exciton coupling in the system. Thus, the shift in absorption confirms a successful shift of the S1 state energy to longer wavelengths. The energy of the
charge separated state, where the BODIPY and anthracene units are reduced and oxidized, respectively, is highly dependent on solvent polarity. The choice of solvent therefore modulates the
driving force for charge separation from the S1 state. Because of the lower energy of the S1 state in the oligomers compared to the monomer, a solvent with higher polarity was needed to
quench the fluorescence for the oligomers. The dynamics of the S1-CS state interaction was examined in detail for the monomer. It was found that a model that included a dynamic equilibrium
between these two states were needed to explain the photophysical properties. Recombination from the CS to the ground state was shown to be in the inverted Marcus regime, practically
limiting the SOCT-ISC mechanism to relatively high energies. Electrochemical analysis was used to complement spectroscopical investigation and showed that the reduction and oxidation
potentials of the dimer do not change to a great extent as compared to the monomer. The ability of all oligomers to populate the triplet state was shown by nanosecond transient absorption
spectroscopy, and the experimentally obtained yields could successfully be simulated by rate equations. Importantly, the corresponding kinetic traces showed no decrease in the triplet
lifetime when increasing the oligomer length. That the lifetime of a molecular series does not reduce with decreasing excitation energy is unusual but can be explained by the triplet energy
being constant in the series, which was confirmed by quantum mechanical calculations. The conserved triplet lifetime highlights a secondary beneficial feature of using exciton coupling for
decreasing the excitation energy of a sensitizer. This is important since long triplet lifetimes are an essential property of photo sensitizers. We suggest the following design concepts to
optimize the photophysical performances beyond those presented herein. Electron delocalization is here restricted through steric bulk. This method is easily synthetically implemented but
allows for a distribution of dihedral angles between monomers (thus a small but variable amount of electron delocalization), resulting in broader than expected absorption and emission
spectral envelopes. By locking the conformation between monomers in an orthogonal orientation, for instance through the introduction of spiro-carbons, electron delocalization is minimized,
and a significant narrowing of the absorbance and emission spectra is expected. Spiro-carbons have previously been introduced in for instance fluorene to form spirobifluorene, which crystal
structure shows perpendicular directions of the two fluorene moieties in the dimer56. The benefits of the SOCT-ISC mechanism are that the lifetime of the acquired triplet state is long, and
that the method works in absence of heavy atoms. The drawback is that it depends on an intermediate CT state, which energy is highly solvent polarity dependent18. Furthermore, in exciton
coupled systems, the driving force for reaching the triplet state is reduced at the same time as the rate of charge recombination to the ground state is increased, lowering the yield of
triplet formation. However, the yields of emission are high and lifetimes relatively long for all oligomers in toluene solution, indicating that exciton coupling itself does not
significantly deactivate the excited state. Thus, limitations for the SOCT-ISC process can most likely be overcome by the utilization of other methods of increasing ISC, for instance through
the incorporation of heavy atoms. In summary, exciton coupling was used to selectively lower the energy of the S1 state without affecting the triplet state energy to a great extent. This
intramolecular exciton-exciton coupling strategy can be a useful addition in the design of heavy atom free photosensitizers and dyes for light emitting diodes with a small singlet-triplet
energy gap. METHODS Synthesis: All reactions were carried out under ambient conditions unless stated differently, for example performed under N2 atmosphere. Glassware were oven dried prior
to use, unless indicated otherwise. Common reagents, solvents, or materials were obtained from Sigma-Aldrich Chemical Co. and used without further purification. Dry solvents for reactions
sensitive to moisture and/or oxygen were obtained through a solvent purifying system (inert PureSolv-MD-5). Column chromatography was performed using silica gel (VWR 40 to 63 μm) unless
stated otherwise. Preparative thin layer chromatography was performed on silica gel plates (UniplateTM, UV254 indicator, 20×20 cm, 2000 micron, Miles Scientific, purchased from Sigma-Aldrich
Chemical Co). Size exclusion column chromatography (SEC) was performed using Bio-Beads S-X3 Support as a stationary phase and DCM as eluent. NMR spectra (1H, 13C, 19F) were recorded on a
BRUKER Avance III spectrometers (600/700/800 MHz 1H; 151/176/201 MHz 13C; 564/659/753 MHz 19F). The 700 MHz spectrometer is equipped with an AVIII console and a 5 mm QCI H/F–C/N–D–05 Z
cryo–probe. The 600 MHz spectrometer is equipped with a NEO console and a QCI H&F/P/C–N–D–05 Z XT cryo–probe. Spectra were taken in CDCl3 (containing tetramethylsilane with 0.00 ppm as
an internal reference) or CD2Cl2 as solvent. Coupling constants (_J_ values) are given in Hertz (Hz) and chemical shifts are reported in parts per million (ppm). 13C spectra are decoupled
from 1H. High-resolution MS was obtained from an Agilent 1290 infinity LC system equipped with an auto sampler in tandem with an Agilent 6520 Accurate Mass Q-TOF LC/MS. LDI MS data were
collected with a Rapiflex MALDI ToF/ToF mass spectrometer (Bruker Daltonics, Bremen, Germany). The instrument is equipped with a Smartbeam 3D UV laser (355 nm) and operated in
reflector-positive ionization mode. A number of 200 shots were acquired at 1 kHz with a mass range set to m/z 400–3200. Melting points were measured using a BÜCHI Melting Point B-545
instrument. IR spectra were recorded using an INVENIO R FT-IR from BRUKER. Analytical HPLC was performed on a Thermo scientific, Dionex Ultimate 3000 system equipped with a Dionex Ultimate
3000 Variable Wavelength Detector and a Reverse Phase Dionex Acclaim PolarAdvantage II column. The solvent used was a mixture of MilliQwater and acetonitrile, both containing 0.1% formic
acid. The oligomers (Monomer, Dimer, Trimer1, Trimer2 and Tetramer) were run using the λmax of the regarding compound as the detection wavelength. The compounds were run using a gradient
ramping from 5% to 95% MeCN over a time period of 15 min with afterward holding of the final ratio for 10 min in the case of Monomer, dimer, trimer1 and trimer2. The tetramer was run using a
gradient ramping from 5% to 95% MeCN over a time period of 10 min with afterward holding of the final ratio for 20 min. Purities are given in the caption of Supplementary Figs. 2, 6, 11,
14, and 17. Optical spectroscopy: Absorption spectra were measured using a spectrophotometer (LAMBDA 950, PerkinElmer). Steady-state emission spectra, excitation spectra and emission
lifetimes were measured with a spectrofluorometer (FLS1000, Edinburgh Instrument) and are corrected using the emission correction files provided by the manufacturer. The emission quantum
yields were calculated using the relative method, using a standard (Fluorescein) with known emission quantum yield57. Fluorescein in 0.1 M NaOH (Φf = 0.91) was used as a reference compound
for ΦEm determination57. (Excitation at 480 nm, Refractive index: 1.33). The samples were measured in solution at 22 °C, refractive indices: toluene = 1.497, DCM = 1.4125, ACN = 1.3405. For
the emission lifetime measurements, the samples were excited by a 375 nm or 510 nm picosecond pulsed diode laser (Edinburgh Instruments). Nanosecond transient absorption measurements were
performed on an Edinburgh Instrument LP 980 spectrometer equipped with an ICCD (Andor). A Spectra-Physics Nd:YAG 532 nm laser (pulse width ~7 ns) coupled to a Spectra-Physics primoscan
optical parametric oscillator (OPO) was used as pump source. The samples used for transient absorption measurements were prepared in a glovebox using anhydrous and degassed solvents. The
samples used for steady-state absorption and emission spectra were prepared using anhydrous solvents (including molar absorptivities, fluorescence decays and fluorescence quantum yield).
Anhydrous solvents were necessary since the CS state was quenched by the presence of even small amounts of water. There are literature reports showing that the charge separated state
lifetime can be heavily influenced by hydrogen bonding58. However, such interactions are not possible in our system. The quenching mechanism involving water was not investigated further
since it is beyond the scope of this report. In intense laser light the concentration of excited state species might become high enough to allow bimolecular reactions. For triplet states,
triplet-triplet annihilation might occur if two excited molecules collide leading to increased deactivation rate of the excited state population59. The time dependence of the triplet excited
state concentration (if the triplet formation occurs within the instrument response function of the instrument) is given by:
$$\frac{d\left[{{\scriptstyle{3}\atop}A}{*}\right]}{{dt}}=-2{k}_{{TTA}}{\left[{{\scriptstyle{3}\atop}A}{*}\right]}^{2}-{k}_{T}\left[{{\scriptstyle{3}\atop}A}{*}\right]$$ (1) where _k__T_ and
_k__TTA_ are the first order intrinsic and second order triplet-triplet annihilation rate constants, respectively. Equation 1 has an analytical solution giving the time dependent triplet
excited state concentration. $$\left[{{\scriptstyle{3}\atop}A}{*}\right]={\left[{{\scriptstyle{3}\atop}A}{*}\right]}_{0}\frac{1-\beta }{{e}^{{k}_{T}t}-\beta },\,{{{\rm{where}}}} \,
\beta=\frac{2{k}_{{TTA}}{\left[{{\scriptstyle{3}\atop}A}{*}\right]}_{0}}{{k}_{T}+2{k}_{{TTA}}{\left[{{\scriptstyle{3}\atop}A}{*}\right]}_{0}}$$ (2) The parameter _β_ describes the ratio of
second order decay over total decay rate of the triplet state. Equation 2 was used to fit the nsTA data of the monomer measured at a series of different intensities in a global fit to obtain
the triplet lifetime. Femtosecond transient absorption (fsTA) measurements were performed with a Ti:sapphire oscillator (Mai-Tai, Spectra Physics) which was used as seed to a regenerative
amplifier (Solstice Ace, Spectra Physics) pumped by a frequency-doubled diode-pumped Nd:YLF laser (Ascend, Spectra Physics). This produced pulses of around 60 fs duration (FWHM) at a 1 kHz
repetition rate. The 800 nm output from the amplifier was split, and the two beams were used as pump and probe light. To achieve the desired excitation wavelength of 390 nm (with an energy
of 1μJ per pulse at the sample), we employed an optical parametric amplifier (TOPAS PRIME, Light Conversion Ltd.) The probe light was directed onto a translating CaF2 plate to generate a
supercontinuum, while the pump beam’s timing was adjusted relative to the probe beam using an optical delay stage (with a range from 0 to 10 ns). The supercontinuum was split into a probe
and reference beam, and the probe beam overlapped with the pump at the sample. The transmitted probe and reference beam were directed to optical fibers and detected by a CCD camera
(iXon-Andor) operating synchronously with the 1 kHz laser. The transient spectra were obtained from the difference of the probe light divided by the reference with and without excitation of
the sample by the pump beam; 2000 spectra were averaged per delay time using a custom LabVIEW program controlling the setup. The cyclic voltammetry measurements were performed with a
CHI-potentiostat controlled using CHI650A software (version 11.15). Platinum electrodes were used as the working and counter electrodes and Ag/AgCl in saturated KCl was used as the reference
electrode. The measurements were performed in thoroughly degassed (using Argon) dichloromethane (DCM) with 0.1 M tetra-_n_-butylammonium perchlorate TBAPF6 (Sigma–Aldrich) at a scan rate of
0.1 V/s. Ferrocene/Ferrocenium (Fc/Fc+) was used as an external standard with E1/2 at approximately 0.61 V vs. Ag/AgCl in DCM. Quantum mechanical calculations were performed using Gaussian
16, revision B.0147,48. The geometry of singlet and triplet ground states were optimized in vacuum at the ωB97XD/6-31 g(d) level of theory and singlet excitation energies were obtained
through TD-DFT using the singlet ground state geometry. The energy difference between the ground singlet and triplet states therefore represents adiabatic transitions, whereas the energy
difference between the ground and excited singlet states represents vertical transitions. DATA AVAILABILITY The data that support the findings of this work are available within the Article
and its Supplementary Information files. Source data are provided with this paper. REFERENCES * Romero, N. A. & Nicewicz, D. A. Organic photoredox catalysis. _Chem. Rev._ 116,
10075–10166 (2016). Article CAS PubMed Google Scholar * Zhao, X., Hou, Y., Liu, L. & Zhao, J. Triplet photosensitizers showing strong absorption of visible light and long-lived
triplet excited states and application in photocatalysis: a mini review. _Energy Fuels_ 35, 18942–18956 (2021). Article CAS Google Scholar * De Bonfils, P., Péault, L., Nun, P. &
Coeffard, V. State of the art of bodipy-based photocatalysts in organic synthesis. _Eur. J. Org. Chem._ 2021, 1809–1824 (2021). Article Google Scholar * García-Santos, W. H. et al.
Dibromo-BODIPY as an organic photocatalyst for radical–ionic sequences. _J. Org. Chem._ 86, 16315–16326 (2021). Article PubMed Google Scholar * Wu, W., Guo, H., Wu, W., Ji, S. & Zhao,
J. Organic triplet sensitizer library derived from a single chromophore (BODIPY) with long-lived triplet excited state for triplet–triplet annihilation based upconversion. _J. Org. Chem._
76, 7056–7064 (2011). Article CAS PubMed Google Scholar * Peng, J., Guo, X., Jiang, X., Zhao, D. & Ma, Y. Developing efficient heavy-atom-free photosensitizers applicable to TTA
upconversion in polymer films. _Chem. Sci._ 7, 1233–1237 (2016). Article CAS PubMed Google Scholar * Carrod, A. J., Gray, V. & Börjesson, K. Recent advances in triplet–triplet
annihilation upconversion and singlet fission, towards solar energy applications. _Energy Environ. Sci._ 15, 4982–5016 (2022). Article CAS Google Scholar * Prieto-Montero, R. et al.
Exploring BODIPY derivatives as singlet oxygen photosensitizers for PDT. _Photochem Photobio._ 96, 458–477 (2020). Article CAS Google Scholar * Nguyen, V.-N., Yan, Y., Zhao, J. &
Yoon, J. Heavy-atom-free photosensitizers: from molecular design to applications in the photodynamic therapy of cancer. _Acc. Chem. Res._ 54, 207–220 (2021). Article CAS PubMed Google
Scholar * Xiao, X., Ye, K., Imran, M. & Zhao, J. Recent development of heavy atom-free triplet photosensitizers for photodynamic therapy. _Appl. Sci._ 12, 9933 (2022). Article CAS
Google Scholar * Xu, K., Zhao, J., Escudero, D., Mahmood, Z. & Jacquemin, D. Controlling triplet–triplet annihilation upconversion by tuning the PET in aminomethyleneanthracene
derivatives. _J. Phys. Chem. C._ 119, 23801–23812 (2015). Article CAS Google Scholar * Xu, S. et al. Tuning the singlet-triplet energy gap: a unique approach to efficient photosensitizers
with aggregation-induced emission (AIE) characteristics. _Chem. Sci._ 6, 5824–5830 (2015). Article CAS PubMed PubMed Central Google Scholar * Balzani V., Ceroni P., Juris A.
_Photochemistry and photophysics: concepts, research, applications_. (John Wiley & Sons, 2014). * Turro N. J., Ramamurthy V., Scaiano J. C. _Modern molecular photochemistry of organic
molecules_, 188. (University Science Books Sausalito, CA, 2010). * Turro N. J., Ramamurthy V., Ramamurthy V., Scaiano J. C. _Principles of molecular photochemistry: an introduction_.
(University Science Books, 2009). * Lee, D. R. et al. Heavy atom effect of selenium for metal-free phosphorescent light-emitting diodes. _Chem. Mater._ 32, 2583–2592 (2020). Article ADS
CAS Google Scholar * Zhao, J., Wu, W., Sun, J. & Guo, S. Triplet photosensitizers: from molecular design to applications. _Chem. Soc. Rev._ 42, 5323–5351 (2013). Article CAS PubMed
Google Scholar * Zhang, X. et al. Recent development of heavy-atom-free triplet photosensitizers: molecular structure design, photophysics and application. _J. Mater. Chem. C._ 9,
11944–11973 (2021). Article ADS CAS Google Scholar * Hussain, M. et al. Spin–orbit charge-transfer intersystem crossing of compact naphthalenediimide-carbazole electron-donor–acceptor
triads. _J. Phys. Chem. B_ 125, 10813–10831 (2021). Article CAS PubMed Google Scholar * Dong, Y. et al. Spin–orbit charge-transfer intersystem crossing (SOCT-ISC) in bodipy-phenoxazine
dyads: effect of chromophore orientation and conformation restriction on the photophysical properties. _J. Phys. Chem. C._ 123, 22793–22811 (2019). Article CAS Google Scholar * Chen, K.
et al. Bodipy derivatives as triplet photosensitizers and the related intersystem crossing mechanisms. _Front. Chem._ 7, 821 (2019). Article ADS CAS PubMed Google Scholar * Filatov, M.
A. et al. BODIPY-pyrene and perylene dyads as heavy-atom-free singlet oxygen sensitizers. _ChemPhotoChem_ 2, 606–615 (2018). Article CAS Google Scholar * Buglak, A. A. et al. Quantitative
structure-property relationship modelling for the prediction of singlet oxygen generation by heavy-atom-free BODIPY photosensitizers. _Chem. Eur. J._ 27, 9934–9947 (2021). Article CAS
PubMed Google Scholar * Wang, Z. & Zhao, J. Bodipy–anthracene dyads as triplet photosensitizers: effect of chromophore orientation on triplet-state formation efficiency and application
in triplet–triplet annihilation upconversion. _Org. Lett._ 19, 4492–4495 (2017). Article CAS PubMed Google Scholar * Filatov, M. A. et al. Control of triplet state generation in heavy
atom-free BODIPY–anthracene dyads by media polarity and structural factors. _Phys. Chem. Chem. Phys._ 20, 8016–8031 (2018). Article CAS PubMed Google Scholar * Filatov, M. A. et al.
Generation of triplet excited states via photoinduced electron transfer in meso-anthra-BODIPY: fluorogenic response toward singlet oxygen in solution and in vitro. _J. Am. Chem. Soc._ 139,
6282–6285 (2017). Article CAS PubMed Google Scholar * Freeman, D. M. E. et al. Synthesis and exciton dynamics of donor-orthogonal acceptor conjugated polymers: reducing the
singlet–triplet energy gap. _J. Am. Chem. Soc._ 139, 11073–11080 (2017). Article CAS PubMed Google Scholar * Han, G., Hu, T. & Yi, Y. Reducing the singlet−triplet energy gap by
end-group π−π stacking toward high-efficiency organic photovoltaics. _Adv. Mater._ 32, 2000975 (2020). Article CAS Google Scholar * Liu, Y., Li, C., Ren, Z., Yan, S. & Bryce, M. R.
All-organic thermally activated delayed fluorescence materials for organic light-emitting diodes. _Nat. Rev. Mater._ 3, 18020 (2018). Article ADS CAS Google Scholar * Yu, Y., Mallick,
S., Wang, M. & Börjesson, K. Barrier-free reverse-intersystem crossing in organic molecules by strong light-matter coupling. _Nat. Commun._ 12, 3255 (2021). Article ADS CAS PubMed
PubMed Central Google Scholar * Stranius, K., Hertzog, M. & Börjesson, K. Selective manipulation of electronically excited states through strong light–matter interactions. _Nat.
Commun._ 9, 2273 (2018). Article ADS PubMed PubMed Central Google Scholar * Cravcenco, A. et al. Exciton delocalization counteracts the energy gap: a new pathway toward NIR-emissive
dyes. _J. Am. Chem. Soc._ 143, 19232–19239 (2021). Article CAS PubMed PubMed Central Google Scholar * Kautsky, H. & Merkel, H. Phosphoreszenz, Selbstauslöschung und
Sensibillisatorwirkung organischer Stoffe. _Naturwissenschaften_ 27, 195–196 (1939). Article ADS CAS Google Scholar * Kasha, M., Rawls, H. R. & El-Bayoumi, M. A. The exciton model in
molecular spectroscopy. _Pure Appl. Chem._ 11, 371–392 (1965). Article CAS Google Scholar * 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). Article CAS PubMed Google Scholar * Patalag, L. J., Ho, L. P., Jones, P. G. &
Werz, D. B. Ethylene-bridged Oligo-BODIPYs: access to intramolecular J-aggregates and superfluorophores. _J. Am. Chem. Soc._ 139, 15104–15113 (2017). Article CAS PubMed Google Scholar *
Patalag, L. J., Hoche, J., Mitric, R., Werz, D. B. & Feringa, B. L. Transforming dyes into fluorophores: exciton-induced emission with chain-like Oligo-BODIPY superstructures. _Angew.
Chem. Int Ed._ 61, e202116834 (2022). Article ADS CAS Google Scholar * Peng, Y.-Z. et al. Charge transfer from donor to acceptor in conjugated microporous polymer for enhanced
photosensitization. _Angew. Chem. Int. Ed._ 60, 22062–22069 (2021). Article CAS Google Scholar * Bröring, M. et al. Bis(BF2)-2,2′-Bidipyrrins (BisBODIPYs): highly fluorescent BODIPY
dimers with large stokes shifts. _Chem. Eur. J._ 14, 2976–2983 (2008). Article PubMed Google Scholar * Cakmak, Y. et al. Designing excited states: theory-guided access to efficient
photosensitizers for photodynamic action. _Angew. Chem. Int. Ed._ 50, 11937–11941 (2011). Article CAS Google Scholar * Pang, W. et al. Modulating the singlet oxygen generation property of
meso–β directly linked BODIPY dimers. _Chem. Commun._ 48, 5437–5439 (2012). Article CAS Google Scholar * Hayashi, Y., Yamaguchi, S., Cha, W. Y., Kim, D. & Shinokubo, H. Synthesis of
directly connected BODIPY oligomers through Suzuki–Miyaura coupling. _Org. Lett._ 13, 2992–2995 (2011). Article CAS PubMed Google Scholar * Nepomnyashchii, A. B., Bröring, M., Ahrens, J.
& Bard, A. J. Synthesis, photophysical, electrochemical, and electrogenerated chemiluminescence studies. multiple sequential electron transfers in BODIPY monomers, dimers, trimers, and
polymer. _J. Am. Chem. Soc._ 133, 8633–8645 (2011). Article CAS PubMed Google Scholar * Rihn, S., Erdem, M., De Nicola, A., Retailleau, P. & Ziessel, R. Phenyliodine(III)
Bis(trifluoroacetate) (PIFA)-promoted synthesis of bodipy dimers displaying unusual redox properties. _Org. Lett._ 13, 1916–1919 (2011). Article CAS PubMed Google Scholar * Lewis, J.
& Maroncelli, M. On the (uninteresting) dependence of the absorption and emission transition moments of coumarin 153 on solvent. _Chem. Phys. Lett._ 282, 197–203 (1998). Article ADS
CAS Google Scholar * Kasha, M. Energy transfer mechanisms and the molecular exciton model for molecular aggregates. _Radiat. Res._ 20, 55–70 (1963). Article ADS CAS PubMed Google
Scholar * Chai, J.-D. & Head-Gordon, M. Long-range corrected hybrid density functionals with damped atom–atom dispersion corrections. _Phys. Chem. Chem. Phys._ 10, 6615–6620 (2008).
Article CAS PubMed Google Scholar * Frisch M. J., et al. _Gaussian 16 Rev. B.01_. (Wallingford, CT; 2016). * Hou, Y. et al. Charge separation, charge recombination, long-lived charge
transfer state formation and intersystem crossing in organic electron donor/acceptor dyads. _J. Mater. Chem. C._ 7, 12048–12074 (2019). Article CAS Google Scholar * Wang, Z. et al.
Insight into the drastically different triplet lifetimes of BODIPY obtained by optical/magnetic spectroscopy and theoretical computations. _Chem. Sci._ 12, 2829–2840 (2021). Article CAS
Google Scholar * Englman, R. & Jortner, J. The energy gap law for non-radiative decay in large molecules. _J. Lumin_ 1-2, 134–142 (1970). Article Google Scholar * Englman, R. &
Jortner, J. The energy gap law for radiationless transitions in large molecules. _Mol. Phys._ 18, 145–164 (1970). Article ADS CAS Google Scholar * Siebrand, W. Radiationless transitions
in polyatomic molecules. II. triplet‐ground‐state transitions in aromatic hydrocarbons. _J. Chem. Phys._ 47, 2411–2422 (2004). Article ADS Google Scholar * Scharf, B. & Dinur, U.
Striking dependence of the rate of electronic radiationless transitions on the size of the molecular system. _Chem. Phys. Lett._ 105, 78–82 (1984). Article ADS CAS Google Scholar *
Humeniuk, A., Mitrić, R. & Bonačić-Koutecký, V. Size dependence of non-radiative decay rates in J-aggregates. _J. Phys. Chem. A_ 124, 10143–10151 (2020). Article CAS PubMed Google
Scholar * Poriel, C. et al. Dispirofluorene–indenofluorene derivatives as new building blocks for blue organic electroluminescent devices and electroactive polymers. _Chem. Eur. J._ 13,
10055–10069 (2007). Article CAS PubMed Google Scholar * Brouwer, A. M. Standards for photoluminescence quantum yield measurements in solution (IUPAC Technical Report). _Pure Appl. Chem._
83, 2213–2228 (2011). Article CAS Google Scholar * Hankache, J., Niemi, M., Lemmetyinen, H. & Wenger, O. S. Hydrogen-bonding effects on the formation and lifetimes of
charge-separated states in molecular triads. _J. Phys. Chem. A_ 116, 8159–8168 (2012). Article CAS PubMed Google Scholar * Haefele, A., Blumhoff, J., Khnayzer, R. S. & Castellano, F.
N. Getting to the (square) root of the problem: how to make noncoherent pumped upconversion linear. _J. Phys. Chem. Lett._ 3, 299–303 (2012). Article CAS Google Scholar Download
references ACKNOWLEDGEMENTS The Swedish NMR center is acknowledged for access to high-field NMRs. The authors especially acknowledge Zoltan Takacs for performing diffusion NMR spectroscopy
and analysis thereof. K.B. gratefully acknowledge financial support from the Knut and Alice Wallenberg Foundation (KAW 2017,0192). B.A. acknowledges the Swedish research council (VR) for
support. Our research relied on computational resources provided by the National Academic Infrastructure for Supercomputing in Sweden (NAISS) at C3SE and NSC partially funded by the Swedish
research council through grant agreement no. 2022-06725. FUNDING Open access funding provided by University of Gothenburg. AUTHOR INFORMATION Author notes * These authors contributed
equally: Clara Schäfer, Rasmus Ringström. AUTHORS AND AFFILIATIONS * Department of Chemistry and Molecular Biology, University of Gothenburg, Box 462, 405 30, Gothenburg, Sweden Clara
Schäfer & Karl Börjesson * Department of Chemistry and Chemical Engineering, Chalmers University of Technology, Kemivägen 10, 412 96, Gothenburg, Sweden Rasmus Ringström, Martin Rahm
& Bo Albinsson * Department of Psychiatry and Neurochemistry, Institute of Neuroscience and Physiology, Sahlgrenska Academy at the University of Gothenburg, Mölndal Hospital, House V3,
431 80, Mölndal, Sweden Jörg Hanrieder * Department of Neurodegenerative Disease, Queen Square Institute of Neurology, University College London, London, WC1N 3BG, London, UK Jörg Hanrieder
Authors * Clara Schäfer View author publications You can also search for this author inPubMed Google Scholar * Rasmus Ringström View author publications You can also search for this author
inPubMed Google Scholar * Jörg Hanrieder View author publications You can also search for this author inPubMed Google Scholar * Martin Rahm View author publications You can also search for
this author inPubMed Google Scholar * Bo Albinsson View author publications You can also search for this author inPubMed Google Scholar * Karl Börjesson View author publications You can also
search for this author inPubMed Google Scholar CONTRIBUTIONS K.B. together with C.S. designed the project idea and the molecules. C.S. synthesized and analyzed the molecules, and performed
absorption, emission and initial transient absorption experiments and analysis thereof. R.R. performed electrochemical analysis of the compounds as well as transient absorption measurements.
M.R. performed DFT calculations. B.A., K.B., R.R., and C.S. analyzed all collected results. J.H. performed MALDI analysis of the synthesized BODIPY oligomers. All authors contributed to
writing the manuscript. All authors have given approval to the final version of the manuscript. CORRESPONDING AUTHOR Correspondence to Karl Börjesson. ETHICS DECLARATIONS COMPETING INTERESTS
The authors declare no competing interest. PEER REVIEW PEER REVIEW INFORMATION _Nature Communications_ thanks the anonymous reviewer(s) for their contribution to the peer review of this
work. A peer review file is available. ADDITIONAL INFORMATION PUBLISHER’S NOTE Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional
affiliations. SUPPLEMENTARY INFORMATION SUPPLEMENTARY INFORMATION PEER REVIEW FILE DESCRIPTION OF ADDITIONAL SUPPLEMENTARY FILES SUPPLEMENTARY DATASET 1 SUPPLEMENTARY DATASET 2 SOURCE DATA
SOURCE DATA 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/. Reprints and permissions ABOUT THIS ARTICLE CITE THIS
ARTICLE Schäfer, C., Ringström, R., Hanrieder, J. _et al._ Lowering of the singlet-triplet energy gap via intramolecular exciton-exciton coupling. _Nat Commun_ 15, 8705 (2024).
https://doi.org/10.1038/s41467-024-53122-7 Download citation * Received: 24 November 2023 * Accepted: 27 September 2024 * Published: 08 October 2024 * DOI:
https://doi.org/10.1038/s41467-024-53122-7 SHARE THIS ARTICLE Anyone you share the following link with will be able to read this content: Get shareable link Sorry, a shareable link is not
currently available for this article. Copy to clipboard Provided by the Springer Nature SharedIt content-sharing initiative