linker-dependent singlet fission in tetracene...

12
Linker-Dependent Singlet Fission in Tetracene Dimers Nadezhda V. Korovina, ,Jimmy Joy, Xintian Feng, § Cassidy Feltenberger, Anna I. Krylov, Stephen E. Bradforth,* and Mark E. Thompson* Department of Chemistry, University of Southern California, Los Angeles, California 90089, United States * S Supporting Information ABSTRACT: Separation of triplet excitons produced by singlet ssion is crucial for ecient application of singlet ssion materials. While earlier works explored the rst step of singlet ssion, the formation of the correlated triplet pair state, the focus of recent studies has been on understanding the second step of singlet ssion, the formation of independent triplets from the correlated pair state. We present the synthesis and excited-state dynamics of meta- and para-bis(ethynyltetracenyl)benzene dimers that are analogues to the ortho- bis(ethynyltetracenyl)benzene dimer reported by our groups previously. A comparison of the excited-state properties of these dimers allows us to investigate the eects of electronic conjugation and coupling on singlet ssion between the ethynyltetracene units within a dimer. In the para isomer, in which the two chromophores are conjugated, the singlet exciton yields the correlated triplet pair state, from which the triplet excitons can decouple via molecular rotations. In contrast, the meta isomer in which the two chromophores are cross-coupled predominantly relaxes via radiative decay. We also report the synthesis and excited-state dynamics of two para dimers with dierent bridging units joining the ethynyltetracenes. The rate of singlet ssion is found to be faster in the dimer with the bridging unit that has orbitals closer in energy to that of the ethynyltetracene chromophores. INTRODUCTION Singlet ssion is a process in which a singlet (S 1 ) excited state of a molecule shares energy with a neighboring chromophore, yielding a pair of triplet (T 1 ) excitons. 13 Incorporation of such materials in organic solar cells could improve the power conversion eciencies by doubling the number of excitons generated by high energy photons, thereby averting thermal losses. Hanna and Nozik have reported that the ShockleyQueisser limit, the theoretical power conversion eciency limit of 33% for a single junction device, could be increased up to 44.4% by combining an ecient singlet ssion material with a red absorber in a device, 4 while Tayebjee et al. take endothermic ssion into account to derive an eciency of 45.9%. 5 In the seminal work of Congreve et al., an internal quantum eciency of 160% was demonstrated in a pentacene/ poly(3-hexylthiophene) device; 6 however, the power con- version eciencies of most singlet ssion devices are suboptimal, which leaves room for improvement in regard to material design. An important criterion for singlet ssion materials for solar cell utilization is fast rate of ssion, to ensure that charge injection occurs from the triplet state, rather the singlet state; particularly, singlet ssion rates faster than 100 ps are desirable. 7 The rate of singlet ssion is proportional to the square of the coupling term: S 1 S 0 |H ̂ el | 1 (T 1 T 1 )between the [S 1 S 0 ] state and the biexciton state, often described as correlated triplet pair [ 1 (T 1 T 1 )], and also depends on the energy gap between these states. 8 Early models by Smith and Michl suggested that S 1 S 0 |H ̂ el | 1 (T 1 T 1 )is maximized in slip- stacked, π-overlapping arrangements of the chromophores; 1 however, more sophisticated theoretical treatments 3 show that coupling can also be signicant in sandwiched and even coplanar chromophores. 3,9,10 The eect of the electronic coupling on singlet ssion is best demonstrated using covalently coupled chromophores (such as dimers and larger arrays). 1117 Analysis of the excited-state dynamics of dimers in solution is advantageous because it allows one to separate the eects of coupling from the solid- state eects (such as intermolecular exciton delocalization) on the rates of singlet ssion. Substantial progress has been made toward the synthesis and photophysical characterization of acene dimers and larger arrays, as well as in covalently linked systems of other singlet ssion chromophores. 1836 Ecient singlet ssion has been reported in dimers wherein the constituent chromophores exhibit substantial π orbital overlap, or are linked together in a conjugated manner; conversely, dimers in which the chromophores are cross-conjugated exhibit slower singlet ssion. Guldi et al. have reported fast and ecient singlet ssion in ortho-, meta-, and para-TIPS- pentacene dimers, in which the favorable E(S 1 ) 2 × E(T 1 ) alignment provides the enthalpic driving force for ssion, such that even the weakly coupled meta pentacene dimers exhibited singlet ssion on picosecond time scales. 19,37 In dimers of tetracene and diphenylisobenzofuran, in which the chromo- phores are without any signicant π orbital overlap or conjugation, and are out of plane with respect to each other, singlet ssion was reported to be slow and inecient, likely due to weak coupling and the E(S 1 ) 2 × E(T 1 ) energy alignment. 3840 Recent work on singlet ssion with pentacene Received: April 25, 2018 Published: July 17, 2018 Article pubs.acs.org/JACS Cite This: J. Am. Chem. Soc. 2018, 140, 10179-10190 © 2018 American Chemical Society 10179 DOI: 10.1021/jacs.8b04401 J. Am. Chem. Soc. 2018, 140, 1017910190 Downloaded via UNIV OF SOUTHERN CALIFORNIA on August 23, 2018 at 12:42:17 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Upload: others

Post on 06-Sep-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Linker-Dependent Singlet Fission in Tetracene Dimersiopenshell.usc.edu/pubs/pdf/jacs-140-10179.pdfLinker-Dependent Singlet Fission in Tetracene Dimers ... Queisser limit, the theoretical

Linker-Dependent Singlet Fission in Tetracene DimersNadezhda V. Korovina,†,‡ Jimmy Joy,† Xintian Feng,§ Cassidy Feltenberger, Anna I. Krylov,Stephen E. Bradforth,* and Mark E. Thompson*

Department of Chemistry, University of Southern California, Los Angeles, California 90089, United States

*S Supporting Information

ABSTRACT: Separation of triplet excitons produced bysinglet fission is crucial for efficient application of singletfission materials. While earlier works explored the first step ofsinglet fission, the formation of the correlated triplet pair state,the focus of recent studies has been on understanding thesecond step of singlet fission, the formation of independenttriplets from the correlated pair state. We present the synthesisand excited-state dynamics of meta- and para-bis(ethynyltetracenyl)benzene dimers that are analogues to the ortho-bis(ethynyltetracenyl)benzene dimer reported by our groups previously. A comparison of the excited-state properties of thesedimers allows us to investigate the effects of electronic conjugation and coupling on singlet fission between the ethynyltetraceneunits within a dimer. In the para isomer, in which the two chromophores are conjugated, the singlet exciton yields the correlatedtriplet pair state, from which the triplet excitons can decouple via molecular rotations. In contrast, the meta isomer in which thetwo chromophores are cross-coupled predominantly relaxes via radiative decay. We also report the synthesis and excited-statedynamics of two para dimers with different bridging units joining the ethynyltetracenes. The rate of singlet fission is found to befaster in the dimer with the bridging unit that has orbitals closer in energy to that of the ethynyltetracene chromophores.

■ INTRODUCTIONSinglet fission is a process in which a singlet (S1) excited stateof a molecule shares energy with a neighboring chromophore,yielding a pair of triplet (T1) excitons.1−3 Incorporation ofsuch materials in organic solar cells could improve the powerconversion efficiencies by doubling the number of excitonsgenerated by high energy photons, thereby averting thermallosses. Hanna and Nozik have reported that the Shockley−Queisser limit, the theoretical power conversion efficiency limitof 33% for a single junction device, could be increased up to44.4% by combining an efficient singlet fission material with ared absorber in a device,4 while Tayebjee et al. takeendothermic fission into account to derive an efficiency of45.9%.5 In the seminal work of Congreve et al., an internalquantum efficiency of 160% was demonstrated in a pentacene/poly(3-hexylthiophene) device;6 however, the power con-version efficiencies of most singlet fission devices aresuboptimal, which leaves room for improvement in regard tomaterial design.An important criterion for singlet fission materials for solar

cell utilization is fast rate of fission, to ensure that chargeinjection occurs from the triplet state, rather the singlet state;particularly, singlet fission rates faster than 100 ps aredesirable.7 The rate of singlet fission is proportional to thesquare of the coupling term: ⟨S1S0|Hel|

1(T1T1)⟩ between the[S1S0] state and the biexciton state, often described ascorrelated triplet pair [1(T1T1)], and also depends on theenergy gap between these states.8 Early models by Smith andMichl suggested that ⟨S1S0|Hel|

1(T1T1)⟩ is maximized in slip-stacked, π-overlapping arrangements of the chromophores;1

however, more sophisticated theoretical treatments3 show that

coupling can also be significant in sandwiched and evencoplanar chromophores.3,9,10

The effect of the electronic coupling on singlet fission is bestdemonstrated using covalently coupled chromophores (such asdimers and larger arrays).11−17 Analysis of the excited-statedynamics of dimers in solution is advantageous because itallows one to separate the effects of coupling from the solid-state effects (such as intermolecular exciton delocalization) onthe rates of singlet fission. Substantial progress has been madetoward the synthesis and photophysical characterization ofacene dimers and larger arrays, as well as in covalently linkedsystems of other singlet fission chromophores.18−36 Efficientsinglet fission has been reported in dimers wherein theconstituent chromophores exhibit substantial π orbital overlap,or are linked together in a conjugated manner; conversely,dimers in which the chromophores are cross-conjugatedexhibit slower singlet fission. Guldi et al. have reported fastand efficient singlet fission in ortho-, meta-, and para-TIPS-pentacene dimers, in which the favorable E(S1) ≥ 2 × E(T1)alignment provides the enthalpic driving force for fission, suchthat even the weakly coupled meta pentacene dimers exhibitedsinglet fission on picosecond time scales.19,37 In dimers oftetracene and diphenylisobenzofuran, in which the chromo-phores are without any significant π orbital overlap orconjugation, and are out of plane with respect to each other,singlet fission was reported to be slow and inefficient, likelydue to weak coupling and the E(S1) ≤ 2 × E(T1) energyalignment.38−40 Recent work on singlet fission with pentacene

Received: April 25, 2018Published: July 17, 2018

Article

pubs.acs.org/JACSCite This: J. Am. Chem. Soc. 2018, 140, 10179−10190

© 2018 American Chemical Society 10179 DOI: 10.1021/jacs.8b04401J. Am. Chem. Soc. 2018, 140, 10179−10190

Dow

nloa

ded

via

UN

IV O

F SO

UT

HE

RN

CA

LIF

OR

NIA

on

Aug

ust 2

3, 2

018

at 1

2:42

:17

(UT

C).

Se

e ht

tps:

//pub

s.ac

s.or

g/sh

arin

ggui

delin

es f

or o

ptio

ns o

n ho

w to

legi

timat

ely

shar

e pu

blis

hed

artic

les.

Page 2: Linker-Dependent Singlet Fission in Tetracene Dimersiopenshell.usc.edu/pubs/pdf/jacs-140-10179.pdfLinker-Dependent Singlet Fission in Tetracene Dimers ... Queisser limit, the theoretical

has also explored the role of conformational dynamics infacilitating excited-state processes,41 observing the correlatedtriplet pair state with mid-IR spectroscopy42 and polarizationresolved transient absorption microscopy,43 and using electronparamagnetic resonance44 and femtosecond stimulated Ramanspectroscopy45 as probes for singlet fission. Ito et al. havetheoretically studied how the next-nearest-neighbor interac-tions affect singlet fission in ortho-, meta-, and para-pentacenedimers,46 while Zeng has used quantum dynamical simulationsto explore the role of through-bond coupling in singlet fissiondimers.47

The currently accepted mechanistic picture of singlet fissioninvolves an intermediate biexciton state,48 which can bedescribed as a correlated triplet pair [1(T1T1)], connecting thesinglet excited state [S0S1] with the two independent tripletstates [3T1].

1 Previously, we reported fast and efficientformat ion of the biexc i ton sta te in or tho -b i s -(ethynyltetracenyl)benzene (o-BETB), in which the twotetracene units are oriented in a twisted configuration withπ-overlap and appear to be strongly electronically coupled; thatis, there is good coupling between the [S0S1] and the[1(T1T1)] states. Although the π orbital overlap between thetetracenes provides significant through-space coupling, abinitio calculations showed that the conjugating, ortho-diethynylbenzene linker also contributes to the coupling, thatis, via through-bond mechanism.49 To disentangle the role ofthrough-bond versus through-space coupling, we synthesizedthe meta and the para analogues of ortho -b is-(ethynyltetracenyl)benzene (m-BETB, p-BETB-ehex, and p-BETB-ohex, Figure 1). The position of the alkyne substitutionon the benzene ring heavily influences the ground-stateelectronic structure of molecules containing these diethynyl-benzene units.50 In the ortho and the para isomers, the π-systems of the two chromophores are conjugated, which leadsto through-bond contributions to the coupling, whereas in themeta isomer the linker does not facilitate conjugation betweenthe tetracene units that are pendant on the alkynes, resulting inonly small through-space coupling. The conjugation throughthe linkers can be understood in terms of the cumulene-likeresonance form of the ortho- and para-diethynylbenzenes.51,52

Importantly, the conjugation manifests itself in stronginteractions between the chromophores both in the groundand in the excited state. The para-diethynylbenzene partic-ularly facilitates efficient electronic communication betweenthe chromophores it bridges;53−55 as a result, compoundsbased on this chemical motif have been incorporated intoultrasensitive sensors,56 energy harvesting materials,57 and

molecular electronic devices.58,59 The meta-diethynylbenzeneis cross-conjugating and has been demonstrated to provideonly weak coupling between the chromophores tethered toit.60,61 Calculations of tetracene-based dimers and modelstructures have shown that the linker contributes significantlyto the coupling, via through-bond interaction, which affects thecharacter of the states involved in singlet fission.62−64 A recenttheoretical work by Abraham and Mayhall provides a model fordetermining the effect of the covalent linker on theboundedness of the 1(T1T1) state based on the Ovchinnikovrule. They concluded that the ortho- and para-benzene dimersof ethynylpentacene and ethynyltetracene have bound 1(T1T1)states, whereas meta coupled dimers should have unbound1(T1T1) states.

65 The boundedness of the 1(T1T1) state is animportant factor when considering the applications of singletfission materials. The materials in which the 1(T1T1) state isbound will not readily produce free triplet excitons, which areessential for reaping the benefits of singlet fission in devices.Therefore, it is valuable to understand the mechanism bywhich the triplets can decouple from the initially formedbiexciton state.Although the favorable E(S1) ≥ 2 × E(T1) energy aligment

in the dimers reported by Guldi et al. ensured fast singletfission, it was unresolvable with the instrument response oftheir transient absorption spectrometer in ortho and paradimers, and rapid annihilation to the ground state wasobserved. The isoenergetic E(S1) ≈ 2 × E(T1) conditions inour bis(ethynyltetracenyl)benzene dimers result in singletfission rates that are fast but still resolvable with our instrumentresponse, and the dynamics of the population of each of thethree states involved in singlet fission could be observed in ourpara dimers. In comparison to the tetracene dimers reportedby Bardeen et al., in which the acenes were directly attached tothe phenyl linker and exhibited slow singlet fission, thetetracene units in our dimers are connected to the phenyllinker via ethynyl groups. This ensures conjugation betweenthe tetracenes (via the cumulene resonance structure) and lesssteric hindrance within a dimer, which could result in fastersinglet fission.40 The presence of ethynyl groups in ourethynyltetracene dimers also perturbs the S1 and T1 stateenergies in a favorable way, as evidenced by faster fissionrates.49 Using ultrafast spectroscopy, we observe that the para-bis(ethynyltetracenyl)benzene dimers (p-BETB-ohex, R =OC6H13; and p-BETB-ehex, R = 2-(ethyl)hexyl in Figure 1)undergo singlet fission from independent triplet states, whereasno significant singlet fission in the meta-bis(ethynyltetracenyl)-benzene dimer (m-BETB) is observed. To understand thedifferences in the excited-state behavior of the ortho versus thepara dimers, we present the transient spectroscopy of p-BETB-ehex in THF and p-BETB-ehex suspended in a rigid polymermatrix (poly(methyl methacrylate), PMMA) and model thekinetics using two populations that exhibit distinct excited-state dynamics. On the basis of modeling and calculatingelectronic coupling as a function of the angle between theethynyltetracenes, we suggest that the rotation of ethynylte-tracenes to a perpendicular orientation facilitates the formationof independent triplets (i.e., productive singlet fission). Tostudy the role of the linker in facilitating singlet fission betweenthe tetracenes, we have also synthesized the para-bis-(ethynyltetracenyl)benzene dimers p-BETB-ehex and p-BETB-ohex (Scheme 1) with different linker units and observethat singlet fission is faster in p-BETB-ohex.

Figure 1. Chemical (top row) and space-filling structures (bottomrow) of o-BETB, m-BETB, and p-BETB-ehex and p-BETB-ohex. Twopara dimers with different substituents on the linker were synthesized(p-BETB-ohex, R = OC6H13; and p-BETB-ehex, R = 2-(ethyl)hexyl).

Journal of the American Chemical Society Article

DOI: 10.1021/jacs.8b04401J. Am. Chem. Soc. 2018, 140, 10179−10190

10180

Page 3: Linker-Dependent Singlet Fission in Tetracene Dimersiopenshell.usc.edu/pubs/pdf/jacs-140-10179.pdfLinker-Dependent Singlet Fission in Tetracene Dimers ... Queisser limit, the theoretical

■ SYNTHETIC AND SPECTROSCOPIC METHODSThe meta and the para dimers were synthesized as shown in Scheme1. The alkynyl motif in these dimers allows them to be constructed viathe highly efficient Sonogashira reaction.66 m-BETB was constructedfrom 5-bromotetracene and the commercially available meta-diethynylbenzene. The para dimers were synthesized by Sonogashirareaction between the appropriate diethynylbenzene linker unit and 5-bromotetracene. The synthetic details and characterization data areavailable in the Supporting Information. Broadband femtosecond andnanosecond transient absorption spectra and time-correlated singlephoton counting traces were collected from solutions of theethynyltetracene dimers made in 1 mm path length quartz cuvettesusing freshly distilled, deaerated THF. Briefly, the broadbandfemtosecond TA setup consists of a Coherent Legend amplifierpumping a Spectra Physics OPA to provide the pump beam and aCaF2 crystal to provide the white light super continuum probe beam.Nanosecond transient absorption spectroscopy was performed usingthe same probe beam and pump beam from a Alphalas Pulselasnanosecond laser. More details on steady state and transientabsorption spectroscopy, fluorescence lifetime measurements, andelectrochemistry are provided in the Supporting Information.

■ RESULTS AND DISCUSSIONElectrochemistry. The extent to which the ethynylte-

tracene units are conjugated in ortho-, meta-, and para-bis(ethynyltetracenyl)benzene dimers can be estimated fromcyclic voltametry (CV). In the dimer the first oxidation shouldbe localized on one of the tetracenes, and the second oxidationshould take place on the second tetracene. Conjugationbetween the ethynyltetracenes in a dimer should give rise to asignificant difference between the first and second oxidationpotentials, because the two ethynyltetracene units are notacting independently. Conversely, if the ethynyltetracene unitsdo not interact, both units in a dimer would oxidize at thesame potential, and the difference between the first and secondoxidation potentials would be close to zero. The cyclicvoltamograms of the ortho-, meta-, and para-bis-(ethynyltetracenyl)benzene dimers, as well as phenylethynylte-

tracene (PET, monomer), were collected in dichloromethane,with ferrocene as the internal standard (additional exper-imental details in the Supporting Information). We observedtwo oxidation peaks in all dimers; however, the peaks werepoorly resolved in m-BETB, due to their proximity (see theSupporting Information). The oxidation potentials and thevalues of the splitting between the oxidation peaks are given inTable 1. On the basis of the magnitude of the splitting of theoxidation potentials, the chromophores that are the moststrongly conjugated are o-BETB and p-BETB-ohex, withΔE(Ox2−Ox1) values of 0.22 V, m-BETB at 0.07 V is theleast conjugated, and p-BETB-ehex is intermediate with adifference of 0.13 V between the two oxidations. Theelectrochemical properties of the dimers also provideinformation about the energetics of intramolecular chargeresonance (CR) states, which are important in the singletfission mechanism.49,67 The approximate energy of theintramolecular CR state (Eredox, Table 1) can be roughlyestimated from the difference between the oxidation (Eox) andthe reduction (Ered) potentials and a Coulombic term.68 Thevalues of Eox, Ered, and the electrochemical energy gap for thedimers and the monomer obtained by CV are given in TableS1. The values of the first excited state (Eoptical), estimated fromthe absorption spectra of these dimers, are also supplied inTable 1 for comparison. In all molecules, the electrochemicalgap is energetically comparable to the optical gap, suggestingthat CR configurations lie close enough in energy to the locallyexcited ones to potentially influence singlet fission kinetics, assuggested in the case of the series of TIPS-pentacenedimers.37,69

Steady-State Spectra. The intensities and line shapes ofthe steady-state extinction spectra are representative of themagnitude of the transition dipole moments in the moleculesand can reflect the extent of the S1S0−S0S1 coupling betweenthe chromophores within the dimers. In our previous work, weshowed that the ethynyltetracene moieties in o-BETB arealigned such that the transition dipole moments of the lowerenergy transition (along the short axis of the acene) arecoupled in an H-aggregate-like manner, as evidenced by theenhanced ν01 and diminished ν00 absorption as compared toPET (Figure 2).49 The absorption of the ortho dimer isredistributed to longer wavelengths, but the integratedintensity in the visible region is less than twice that of themonomer. The absorption spectrum of m-BETB is nearlyidentical to that of PET, but with a small red-shift of 4 nm(0.02 eV). The relative intensities of the ν00 and the ν01vibronic features of the PET and the m-BETB are identical.The most notable difference between PET and m-BETBabsorption properties is in their molar absorptivity spectra: themolar absorptivity of m-BETB is twice that of PET in thevisible part of the spectrum (the ratio of integrated areas is

Scheme 1. Synthetic Scheme for m-,p-BETB-ohex and p-BETB-ehexa

a(i) NBS, DMF, CH2Cl2, 40 °C; (ii) Pd(PPh3)4, CuI, THF, Et3N, 60°C.

Table 1. Electrochemical Properties of the Four Dimers As Compared To the Monomer, and the Optical Gaps of TheseMoleculesa

molecule first oxidation (V) second oxidation (V) ΔE(Ox2−Ox1) (V) Eredox (eV) Eoptical (eV) Eoptical − Eredox (eV)

PET 0.46 2.38 2.42 0.04o-BETB 0.37 0.59 0.22 2.32 2.37 0.05m-BETB 0.41 0.48 0.07 2.30 2.40 0.10p-BETB-ehex 0.46 0.59 0.13 2.38 2.34 −0.04p-BETB-ohex 0.33 0.55 0.22 2.21 2.28 0.07

aThe ferrocenium/ferrocene redox couple in dichloromethane was used as an internal standard for electrochemical measurements.

Journal of the American Chemical Society Article

DOI: 10.1021/jacs.8b04401J. Am. Chem. Soc. 2018, 140, 10179−10190

10181

Page 4: Linker-Dependent Singlet Fission in Tetracene Dimersiopenshell.usc.edu/pubs/pdf/jacs-140-10179.pdfLinker-Dependent Singlet Fission in Tetracene Dimers ... Queisser limit, the theoretical

2.0). This suggests that tetracenes in m-BETB are effectivelyindependent of each other and the electronic coupling betweenthem is weak.The absorption properties of the p-BETB-ohex and p-BETB-

ehex are different from those of o-BETB, m-BETB, and themonomer PET. The molar absorptivity of p-BETB-ehex in thevisible region is more than twice that of PET, with the areaunder the curve greater than twice that of PET by a factor of2.6. The red shift and significant difference between theintegrated absorptivity of the dimer versus twice that of themonomer suggest strong exciton coupling between the twoethynyltetracene units in p-BETB-ehex.70,71 The increase inmolar absorptivity of p-BETB-ohex over PET is even greater,with p-BETB-ohex having an absorptivity more than 3 timesgreater than that of PET in the visible region (the ratio of theintegrated areas of the absorption bands is 3.3), and isaccompanied by a substantial red-shift in the visibleabsorption. This further demonstrates that the electron-donating alkoxy substituents on the bridging benzene ringsignificantly alter the electronic structure of the dimers and inthis case appear to enhance exciton coupling. The intensity ofthe emission (see the Supporting Information) gives insightinto the rates of the excited-state decay channels. Analogouslyto o-BETB, the para dimers are very weakly emissive (Φfl <1%), suggesting that the process, which quenches thefluorescence in these systems, is orders of magnitude fasterthan ∼10 ns (viz., the excited-state lifetime of the monomerPET; see the Supporting Information). The analogous para-bis(ethynylanthracenyl)benzene dimers emit with high quan-tum efficiencies: 97% for para-bis(ethynylanthracenyl)-benzene,52 and 60% for the analogous hexyloxy-substitutedpara-bis(ethynylanthracenyl)benzene.72 Similarly to the com-parison of o-BETB to its anthracene analogue in our previouswork, the high quantum efficiency of emission of theanthracene dimers as compared to the tetracene dimerssuggests that a rapid excited-state decay channel is available inthe para-tetracene dimers, which is not available in theanthracene analogues. Conversely, m-BETB is stronglyemissive in solution (Figure 3), with a quantum yield of62%, suggesting that the rate and yield of nonradiative excited-state decay processes are not significant in this system. Theemission decay, shown in Figure 3b, has a small delayedcomponent (∼190 ns), analogous to those observed for weaklycoupled tetracene dimers by Bardeen73 and Damrauer,74 andcould be suggestive of triplet−triplet recombination. Presum-

ing that the delayed fluorescence is a result of recombination oftwo geminate triplets generated by singlet fission, we used akinetic model analogous to that reported by Muller et al. andCook et al.40,73,74 to estimate the yield of singlet fission to beapproximately 1% in m-BETB, and the rate of the decay of thedelayed fluorescence to be solvent dependent (details in theSupporting Information). The photoluminescence lifetimes ofall solutions of m-BETB in THF with concentrations varyingfrom 1.5 to 21 μM (see the Supporting Information for details)are the same (∼170 ns), indicating the absence of anyintermolecular reactions between two dimers in solution. Thetriplets on m-BETB therefore recombine via a geminatepathway.

Transient Absorption of m-BETB. The femtosecondtransient absorption spectra of m-BETB in THF followingexcitation at 515 nm are shown in Figure 4a. The characteristicacene Sn ← S1 absorption peak is observed at 400 nm, alongwith ground-state bleach at 516 and 480 nm; similar featureswere observed in the case of the ethynyltetracene monomer(ETTMS) in our earlier studies.49 Between 10 and 975 ps,there is no significant evolution of these spectral features. TheSn ← S1 absorption decays after 5 ns when excited at 532 nmin a nanosecond pump broadband transient absorptionexperiment (Figure 4b), with a concomitant rise of a positivefeature at ∼510 nm. A peak in this spectral region has beenpreviously observed for Tn ← T1 absorption in acenes like5,12-diphenyltetracene (DPT) and ET-TMS.37,49,75 Moreimportantly, the spectral shape at 135 ns is identical to thesensitized Tn ← T1 absorption of m-BETB (inset to Figure4b). The rise of the absorption feature at 510 nm is slow,suggesting that the triplets could be generated in this system by

Figure 2. Extinction spectra of m-BETB, p-BETB-ehex, and p-BETB-ohex as compared to o-BETB and PET in THF solution.

Figure 3. (a) Steady-state absorption and emission of m-BETB inTHF. (b) Emission decay of m-BETB in THF (excitation at 405 nm,emission at 570 nm), with noticeable delayed fluorescence on the∼200 ns time scale.

Journal of the American Chemical Society Article

DOI: 10.1021/jacs.8b04401J. Am. Chem. Soc. 2018, 140, 10179−10190

10182

Page 5: Linker-Dependent Singlet Fission in Tetracene Dimersiopenshell.usc.edu/pubs/pdf/jacs-140-10179.pdfLinker-Dependent Singlet Fission in Tetracene Dimers ... Queisser limit, the theoretical

intersystem crossing, with rates similar to those of othertetracene derivatives.76 The transient absorption data were fitwith a two-state sequential model (see the SupportingInformation) with a time constant of 15 ns for the rise ofthe triplet feature and a triplet lifetime of 4.6 μs. The tripletscould also arise from a channel in which very slow (andtherefore inefficient) singlet fission results in the delayedfluorescence from m-BETB in solution (∼190 ns, Figure 3b).It is reasonable to attribute the negligible singlet fission and

slow formation of triplets in m-BETB in solution to the weaknonadiabatic coupling, similar to the meta dimer in a series oftetracene dimers reported by Muller et al.40 Ab initiocalculations further support this notion. The couplingparameter, ||γ||2, between the S1 and the 1(T1T1) states in m-BETB is calculated to be 4.6 × 10−5, a value that is about 3orders of magnitude smaller than for o-BETB (0.04).64

Additionally, the computed binding energy (Eb) of the triplets(calculated as the difference between the energy of thebiexciton stateand the two separated triplets) indicates how

strongly the T1 excitons are coupled in the 1(T1T1) statewithin the dimer and is reflective of the intramolecularchromophore coupling. The value of Eb in m-BETB is <0.001eV (as compared to 0.15 eV for o-BETB), suggesting that thetetracenes are weakly coupled in this dimer. Although singletfission was observed to proceed on a 200 ps time scale in theTIPS-pentacene analogue of m-BETB, the rate of singlet fissionreported in the meta dimer in the series of the TIPS-pentacenedimers was slower by at least 3 orders of magnitude ascompared to the ortho and para dimers.37,77

Transient Absorption of p-BETB-ohex and p-BETB-ehex in THF. The TA spectra of p-BETB-ohex and p-BETB-ehex in THF, excited at 545 and 530 nm, respectively, areshown in Figures 5a and 6a. At early times (<25 ps), bothcompounds display the characteristic excited-state absorptionband at 425 nm and ground-state bleach at 520−540 nm. Thevibronic structure associated with ground-state absorption andexcited-state emission (through contributions from ground-state bleaching and stimulated emission) are superimposed onthe broader excited-state absorption bands in the overalltransient absorption spectra as seen in Figures 5b and 6b. Thebroad absorption feature at 425 nm decays within 100 ps,giving rise to an induced absorption peak at 555 nm in p-BETB-ohex and 540 nm in p-BETB-ehex. This feature of thepara dimers decays much more rapidly than that of m-BETB,and most importantly the spectral shape of the inducedabsorption of p-BETB-ehex and p-BETB-ohex at ∼100 ps isalso different from that of o-BETB reported previously.49

To get an idea of the excited-state dynamics in the p-BETBdimers, we subjected the TA data to target analysis.78 Weobtained the best fit using a three-state sequential model([state 1] → [state 2] → [state 3]). The species associateddifference spectra (SADS) and corresponding concentrationcurves obtained from target analysis are shown in Figures 5and 6, respectively. The shapes of the SADS of p-BETB-ohexand p-BETB-ehex in THF are qualitatively similar (Figures 5band 6b). The spectral signature of state 1 in both systemsresembles the typical acene Sn ← S1 absorption,

49 with a sharppeak in the 375−470 nm region with a larger (negative)amplitude from ground-state bleach and stimulated emission.In both systems, state 1 rapidly decays to state 2 with a broaderband in the 375−470 nm region, a net absorption from 575−650 nm, and ground-state bleach features similar to those ofstate 1. The spectral shape of state 2 of p-BETB-ohex and p-BETB-ehex in THF resembles that of the 1(T1T1) stateabsorption in o-BETB. The difference in the SADScorresponding to state 2 in p-BETB-ohex and p-BETB-ehex

Figure 4. (a) Femtosecond and (b) nanosecond transient absorptionof m-BETB, with the sensitized Tn ← T1 absorption shown in black.Inset to (b) shows an overlay of the transient spectrum of m-BETB at135 ns and the sensitized Tn ← T1 spectrum.

Figure 5. Femtosecond transient absorption (a), the species associated difference spectra (b), and concentration curves (c) from the target analysisof the TA data of p-BETB-ohex in THF.

Journal of the American Chemical Society Article

DOI: 10.1021/jacs.8b04401J. Am. Chem. Soc. 2018, 140, 10179−10190

10183

Page 6: Linker-Dependent Singlet Fission in Tetracene Dimersiopenshell.usc.edu/pubs/pdf/jacs-140-10179.pdfLinker-Dependent Singlet Fission in Tetracene Dimers ... Queisser limit, the theoretical

suggests that the spectrum originating from this state isdependent upon the structure of the bridging linker, andtherefore the coupling between acenes. Recall that, on the basisof the splitting in the oxidation potential peaks, theconjugation between the tetracenes is the strongest in o-BETB and p-BETB-ohex (0.22 V), and significantly weaker(0.13 V) in p-BETB-ehex. The rate of formation of state 2 inthe para dimers, however, is much faster than the formation of1(T1T1) in o-BETB. The target analysis gives values of 0.1 ps(essentially on the time scale of the instrument response) for p-BETB-ohex and 0.4 ps for p-BETB-ehex (Table 2), while in o-BETB, 1(T1T1) was formed in 2 ps. The faster rate offormation in the para dimers than in the ortho dimer highlightsthe importance of through-bond coupling. The result suggeststhat through-bond coupling in the planar p-BETB-ehex isstronger than the coupling in o-BETB, in which thechromophores are predominantly coupled through-space, butalso exhibit some through-bond coupling via the linker, albeitto a lesser extent because the chromophores in o-BETB are notin plane with the linker. Unlike in o-BETB, the 1(T1T1) state inp-BETB-ohex and p-BETB-ehex decays to a third state with anidentifiably different spectral signature, with prominentinduced absorption peaks at 555 and 540 nm in p-BETB-ohex and p-BETB-ehex, respectively. The spectral shape ofstate 3 in p-BETB-ohex and p-BETB-ehex in solution bears astrong resemblance to that of the sensitized Tn ← T1 inducedabsorption of the monomer, ET-TMS.49

The absorption of state 3 could correspond to either one ortwo triplets per dimer. There are two factors, which suggestthat state 3 corresponds to absorption of two triplets on thesame dimer: (i) The rate of formation of state 3 is fast, 19 ps inp-BETB-ohex and 48 ps in p-BETB-ehex (Table 2). Theserates of T1 formation are too fast for intersystem crossing inacenes,76 but are in the range of the measured singlet fissionrates in tetracene.8,37,79,80 (ii) The rate of the decay of state 3 isvery fast, 20 ps in p-BETB-ohex and 76 ps in p-BETB-ehex.

Typically, a single triplet exciton localized on an acene dimerdecays on the time scale of microseconds.76 However, therecent studies of pentacene dimers showed that when twotriplet excitons are located on the same dimer, the Tn ← T1absorption decays on the 10−100 ps time scale, due to thefavorable T1−T1 fusion pathway to the ground state.77,81

Therefore, the SADS for state 3 in the para dimers in THF isthe spectral signature of two triplets in close proximity to eachother. The most likely origin of these two triplets is via singletfission.In Figure 7, an overlay of the SADS for state 3 of p-BETB-

ehex is made with Tn ← T1 absorption spectra for a singletriplet sensitized onto p-BETB-ehex and triplet sensitized ontothe model monomer compound, ETTMS. The detailsregarding sensitization experiments have been previouslyreported.49 For both of the sensitized spectra, a spectral shiftis required to line up the peaks with the state 3 SADS. Despitesmall discrepancies in the intensities in the 400−500 nmregion, the relative peak positions and relative intensitiesoverlay well, strongly suggesting that state 3 corresponds to Tn← T1 induced absorption in p-BETB-ehex. It is worth notingthat state 3 resembles the Tn ← T1 induced absorption of themonomer more closely than that of the dimer. We surmise thatin the case of the sensitized triplet on the dimer, the Tn ← T1excitation may be delocalized over more than just the one

Figure 6. Femtosecond transient absorption (a), the species associated difference spectra (b), and concentration curves (c) from the target analysisof the TA data of p-BETB-ehex in THF.

Table 2. Time Constants (in ps) of the Decays Used To Fitthe TA Data of the Dimers in THF

o-BETBa m-BETBb

p-BETB-ohex

p-BETB-ehex

τ [S1 →1(T1T1)] 2.0 1.5 × 104 0.1 0.4

τ [1(T1T1) →3T1] 19 48

τ [3T1] 4.6 × 106 20 76

aTaken from previously published work, ref 17. bTime constantscorresponding to the intersystem crossing rate and the decay of thattriplet.

Figure 7. Plot shows the spectral signature of state 3 from targetanalysis of p-BETB-ehex in THF as compared to the normalizedsensitized Tn ← T1 absorption of ET-TMS (red-shifted, 20 nm) andsensitized Tn ← T1 absorption of p-BETB-ehex in THF (blue-shifted,9 nm).

Journal of the American Chemical Society Article

DOI: 10.1021/jacs.8b04401J. Am. Chem. Soc. 2018, 140, 10179−10190

10184

Page 7: Linker-Dependent Singlet Fission in Tetracene Dimersiopenshell.usc.edu/pubs/pdf/jacs-140-10179.pdfLinker-Dependent Singlet Fission in Tetracene Dimers ... Queisser limit, the theoretical

ethynyltetracene unit onto the phenyl linker or even thesecond ethynyltetracene. Therefore, the sensitized tripletabsorption of the dimer differs from that of the monomerETTMS: when singlet fission produces two triplet excitons inthe dimer they likely localize to their respective acenes, andwhen the acenes rotate out of plane from each other thespectrum of the triplet excitons effectively resembles thesensitized triplet spectrum of monomer ETTMS.49 Weestimated the triplet yield in these compounds by scaling thetarget analysis populations by the extinction coefficients of thesinglet and the triplet absorptions (see the SupportingInformation for details) and obtained 94% for p-BETB-ehexand 70% for p-BETB-ohex. The triplet yields are low in thesepara dimers because the rate of recombination of the triplets isfast relative to the rate of their formation.Surprisingly, the observed behavior of our ortho-,49 meta-,

and para-ethynyltetracene dimers is different from that ofisomorphic dimers of TIPS-pentacene reported by Tykwinski,Guldi, and co-workers.37 All TIPS-pentacene dimers undergosinglet fission, with only the meta-TIPS-pentacene dimerundergoing fission slowly enough to be resolved by theinstrument.45,82,83 In our previous work on o-BETB, theformation of only the biexciton state was observed, whereashere we find evidence of predominantly radiative decay in m-BETB. We observe that the para-ethynyltetracene dimer formstwo independent triplet states via the biexciton state. UsingTCSPC (with more dynamic range), we observe that a smallfraction (1%, see the Supporting Information) of m-BETBmolecules gives rise to delayed fluorescence, which were notcaptured in transient absorption. The delayed fluorescence inm-BETB is solvent dependent, which is similar to the meta-TIPS-pentacene dimer. Solvent dependence suggests involve-ment of charge resonance configurations. However, in contrastto meta-TIPS-pentacene,84−89 we do not observe the spectralsignatures of the charge resonance states in the transientabsorption spectra of m-BETB. We explain this difference bynoting that the optical and electrochemical gaps are closer inmeta-TIPS-pentacene than in m-BETB (1.88 vs 1.87 eV in theformer and 2.30 versus 2.40 eV in the latter), which suggeststhat there is a smaller contribution of the charge resonancestates in m-BETB. The analysis of the S1 wave function shows0.2% of charge resonance configurations in m-BETB. In theTIPS-pentacene dimers, the energy of two triplets at 1.54 eV isfar below the energy of the S1 state at 1.88 eV. This energeticdriving force for singlet fission in the TIPS−pentacene systemsresults in ultrafast kinetics from which mechanistic informationis not easily obtained. In the case of ethynyltetracene dimerspresented here, the E(S1) ≈ 2E(T1) isoenergetic conditionsmake singlet fission only slightly favorable, as a result of whichthe subtle differences in the effect of through-bond andthrough-space coupling can be investigated.

The ortho- and the para-bis-ethynylbenzene dimers exhibitdistinctly different excited-state dynamics. The S1 state in thepara dimers yields the 1(T1T1) state, which decays to formindependent triplets, whereas the 1(T1T1) state in o-BETBrelaxes to the ground state without producing independenttriplet excitons. According to the model proposed by Abrahamand Mayhall (AM model), both ortho and para dimers shouldproduce bound 1(T1T1) states that are unlikely to formnoninteracting triplet excitons.65 Although the AM modelcorrectly captures the important aspect of the nature of thebiexciton states in these systems, as compared against therestricted active space configuration interaction with doublespin-flip method (RAS- 2SF-CI), and explains the observationsfor o-BETB, it does not explain the observed singlet fissiondynamics in our para dimers, which we attribute to the staticnature of the model. The difference in the excited-stateprocesses in the ortho and para dimers and the deviation fromthe AM model can be rationalized by considering rotationalmotions of the chromophores in these dimers. In o-BETB thetetracenes can rotate, and they always retain some degree of πoverlap, which results in nonzero biexciton binding energy. Inthe para dimers, on the other hand, the chromophoreinteraction is limited to through-bond only. Upon twisting ofone of the tetracenes perpendicular to the plane of the rest ofthe molecule, the electronic coupling between the chromo-phores is broken and Eb becomes zero (Table 3).Consequently, the triplets from the correlated triplet pairstate become independent of each other. Thus, to explainsinglet-fission trends in flexible dimers, the AM model needs tobe extended to account for changes in electronic structureupon twisting. Figure 8 shows the possible molecularorientations corresponding to the correlated and independenttriplet pair states in p-BETB-ehex. Further insight into theeffect of rotation on the excited-state energies of p-BETB-ehexcan be obtained from theory. The energies of the S1,

1(T1T1),and 5(T1T1) states were computed as a function of the dihedralangle between the tetracenes in p-BETB-ehex using RAS-2SF-CI (Table 3). The minimum of the S1 potential energy surfacecorresponds to a planar geometry, and the 1(T1T1) and5(T1T1) states are energetically downhill from the S1 state byabout 0.4 eV at all rotational angles. The coupling between theS1 and the 1(T1T1) states (||γ||

2) was calculated to be largest ata planar geometry (0.020), and is 0.004 at 60° and drops tozero at 90°. The 1(T1T1) surface has a minimum at 60°, whilethe 5(T1T1) surface has a minimum at 90°, at which the1(T1T1) and the 5(T1T1) states become degenerate (see Table3). These values indicate that the triplet excitons from thecorrelated triplet pair state in p-BETB-ehex can becomenoninteracting in the 5(T1T1) state when the tetracenes rotateperpendicular to each other within the dimer.

Table 3. RAS-2SF-CI Computed Excited-State Energies (ineV), and Binding Energy (in eV) and Coupling for p-BETB-ehex, the para Dimer, As a Function of the Dihedral Anglebetween the Tetracenes

angle (deg) E(S1) E(1(T1T1)) Eb ||γ||2

0 3.84 3.38 0.04 0.02030 3.91 3.45 0.04 0.02060 3.86 3.37 0.01 0.00490 3.87 3.39 0.00 0.000

Figure 8. Proposed decoupling of the correlated triplet state in p-BETB-ehex (alkyl chains have been removed for clarity) to give rise totwo independent triplets.

Journal of the American Chemical Society Article

DOI: 10.1021/jacs.8b04401J. Am. Chem. Soc. 2018, 140, 10179−10190

10185

Page 8: Linker-Dependent Singlet Fission in Tetracene Dimersiopenshell.usc.edu/pubs/pdf/jacs-140-10179.pdfLinker-Dependent Singlet Fission in Tetracene Dimers ... Queisser limit, the theoretical

Transient Absorption of p-BETB-ehex in PMMA. Ifrotational motion of the tetracene in para dimer in solutionprovides the mechanism for generating free triplets, thenrestricting that motion should reduce the rate and the yield ofthe formation of free triplets. To test this hypothesis, a solutionof p-BETB-ehex in a rigid polymer matrix, poly(methylmethacrylate) (PMMA) was prepared. p-BETB-ehex inPMMA produces a substantial amount of emission at roomtemperature, with a quantum efficiency of ∼10% (see theSupporting Information), which means that a significantfraction of the molecules are confined to a geometry inwhich the radiative decay is the predominant relaxationpathway from the singlet excited state. From a comparisonof the transient absorption spectra of p-BETB-ehex in THFand in PMMA (Figure 9), it is evident that in PMMA theexcited-state absorption corresponding to S1 → Sn absorptioncentered at 420 nm decays much more slowly. Thecharacteristic Tn ← T1 absorption peak at 550 nm still formsin this media; however, the rate of its formation is much slowerthan for p-BETB-ehex in THF. Unlike p-BETB-ehex in THF,the transient signals observed from p-BETB-ehex in PMMAcannot be fit using a 3-state sequential model. PMMA, whichbeing a more rigid medium than THF, raises the barrier torotation for the tetracene units in p-BETB-ehex along thediethynylbenzene linker. This results in p-BETB-ehex dimersthat are conformationally locked in the polymer matrix in adistribution of rotational angles. Upon excitation in this rigidmedium, not all of the p-BETB-ehex molecules that end up inthe 1(T1T1) state will be able to rotate and form independenttriplets. The TA spectra from p-BETB-ehex in PMMA with thetetracene units locked in all dihedral angles can be modeled asa sum of signals from two model populations with extremeconformations of the para dimer: population 1 with thetetracenes coplanar with each other and population 2 withtetracenes perpendicular (Figure 10). Population 1 representsthe dimer molecules in PMMA, which cannot undergo singletfission completely, and the 1(T1T1) remains in equilibriumwith the S1 state, giving rise to the fluorescence signal from thePMMA sample. Population 2 is representative of dimermolecules that can fully undergo singlet fission to twoindependent triplets, and gives rise to the triplet band at 550nm in the TA spectra at 957 ns. Using compartmental modelsto represent the excited-state processes in the distinctpopulations, it is possible to obtain concentration curves that

describe the dynamics of the excited states (details in theSupporting Information).78 The transient absorption spectracorresponding to both model populations 1 and 2 can bederived by multiplying the concentration curves with thespecies associated difference spectra (from Figure 6b)corresponding to the S1,

1(T1T1), and T1−T1 states (see theSupporting Information). The fractions of the modelpopulation 1 and population 2 that best represent p-BETB-ehex in PMMA are obtained by comparing the time slice of thetransient absorption data at 425 nm with the time slices at thesame wavelength of the modeled transient absorption data; thelatter is obtained by combining the TA spectra of thepopulation 1 and population 2 in different fractions andoptimizing the time constants for the various excited-stateprocesses by repeated iterations in MATLAB. As seen inFigure 11, the best representation of the excited-state dynamicsis obtained by combining the dynamics of populations 1 and 2in equal fractions and reproduces the 10% quantum yieldmeasured from the PMMA film. This suggests that in a morerealistic model, the tetracene units in p-BETB-ehex in thepolymer matrix would be locked at all dihedral angles withequal probability. This is quite likely the case because themonomer units in PMMA are small, and would not hold thetetracenes at any one angle preferentially by noncovalent

Figure 9. Femtosecond transient absorption of p-BETB-ehex inPMMA.

Figure 10. State diagrams showing (a) population 1 and (b)population 2 used to model the TA data from p-BETB-ehex inPMMA.

Figure 11. Dynamics of excited-state absorption at 425 nm of p-BETB-ehex in PMMA (black trace) overlaid with the time slices at425 nm of transient absorption spectra obtained by combiningmodeled transient absorption of population 1 and population 2 indifferent fractions. Numbers in the legend show the fraction ofpopulation 1 in that model. The rate constants (s−1) that gave the bestfit are kr = 5.5 × 107, knr = 2.9 × 106, k12 = 5 × 1011, k21 = 4 × 1012, k22= 2 × 109, k12′ = 1.6 × 1012, k23 = 1.2 × 1010, and k33 = 1.2 × 109.

Journal of the American Chemical Society Article

DOI: 10.1021/jacs.8b04401J. Am. Chem. Soc. 2018, 140, 10179−10190

10186

Page 9: Linker-Dependent Singlet Fission in Tetracene Dimersiopenshell.usc.edu/pubs/pdf/jacs-140-10179.pdfLinker-Dependent Singlet Fission in Tetracene Dimers ... Queisser limit, the theoretical

interactions. The TA spectra of p-BETB-ehex in PMMA andits modeling using conformationally restricted populationsemphasize the importance of the rotation of the tetracene unitsin allowing the 1(T1T1) state to separate to independenttriplets in p-BETB-ehex and not in o-BETB, despite thesimilarities in electronic coupling. A similar rotationalmechanism of formation of triplets from the correlated tripletpair state has been reported in bipentacene dimers by Tayebjeeet al.90 The phenylene bridges in these dimers resulted in along-lived 5(T1T1) state that had a lifetime >200 ns andallowed the authors to characterize the quintet state usingtime-resolved EPR. According to the authors, the coupling inthe bipentacene BP2 with the pentacenes separated by twophenylene units is greater than that in BP3 with the pentacenesseparated by three phenylene units. This difference is reflectedin the fact that in BP3 the quintet state evolves into dissociatedtriplets, while in BP2 the quintet state gives rise to triplets thatare still weakly interacting. Although an analogous analysis forour p-BETB systems would be insightful, it is unlikely to befruitful in the case of our dimers. On the basis of the fast timescale (<100 ps) of the conversion from 1(T1T1) to electroni-cally independent triplets (2 × T1) in p-BETB dimers, followedby rapid return to the ground state, we can expect that thetriplet excitons do not have enough time to undergo spindephasing, and only undergo electronic decoupling. ESRexperiments analogous to those carried out by Tayebjee et al.would be hampered by the short excited-state lifetime observedin p-BETB dimers.Finally, of additional interest are the differences in the

excited-state kinetics of p-BETB-ohex and p-BETB-ehex. Allthree of the processes, relaxation of S1 into

1(T1T1), conversionof 1(T1T1) to independent triplets, and the decay of theresultant triplets, are faster in p-BETB-ohex as compared to p-BETB-ehex (Table 2). The difference in the coupling of thetetracene units in p-BETB-ohex and p-BETB-ehex can beunderstood in terms of the relative orbital energies of thetetracenes and the two bis(ethynyl)benzene linkers. Aqualitative diagram showing the relative orbital energies ofthese molecules is provided in Figure 12. In the alkyl-substituted benzene linker, the orbitals are expected to lieenergetically far from those of the tetracenes, and the couplingis not expected to be very large. However, introducing theelectron-donating hexyloxy substituents onto the benzenelinker raises the energy of the HOMO, and therefore brings the

orbital energies of the linker closer to those of the tetracenes,resulting in stronger coupling between the acenes and fastersinglet fission in p-BETB-ohex.

■ CONCLUSIONSWe have synthesized the meta and para analogues of thepreviously reported ortho-bis(ethynyltetracenyl)benzene tostudy the role of through-bond versus through-space couplingin facilitating singlet fission between the tetracene units. Theortho- and para- ethynylbenzene linkers provide conjugationbetween the tetracene units, while the meta-ethynylbenzenelinker is cross-conjugating in the ground state. Usingbroadband femtosecond transient absorption spectroscopy,we have established that the meta-(ethynyltetracenyl)benzenepredominantly undergoes radiative decay in the excited state.Surprisingly, the para-bis(ethynyltetracenyl)benzene dimerundergoes complete singlet fission, forming independent tripletstates via rotation of the acenes out of plane from each other.As reported previously, the ortho-bis(ethynyltetracenyl)-benzene dimer only forms the intermediate correlated tripletpair state.Comparing the behavior of ortho-, meta- and para-

bis(ethynyltetracenyl)benzene dimers provides crucial insightinto the role of conjugation and through-bond coupling insinglet fission in dimers. The ortho and para dimers in whichthe tetracene units are conjugated through the ethynylbenzenelinker exhibit singlet fission, while the cross-linked meta dimerpredominantly undergoes radiative decay. Furthermore, the[S1] → [1(T1T1)] process is faster in p-BETB-ohex and p-BETB-ehex than in o-BETB, suggesting that the through-bondcoupling operating in the para dimers is stronger than thecombined through-space and through-bond coupling operatingin the ortho dimer. Unlike in the case of a series of pentacenedimers, where the energetics are so favorable for singlet fissionthat the ortho-, meta-, and para-TIPS-pentacene dimers alldisplayed triplet absorption signatures in picosecond timescales, in our case, only the para dimer forms independenttriplets on similar time scales. This provides an insightfulcomparison with the meta dimer, which does not form asignificant number of triplets, and the ortho dimer, which onlyforms the intermediate correlated triplet pair state. Weattribute the difference in the excited-state dynamics of theortho and para dimers to the rotational flexibility of the acenesin the para dimer, which breaks the coupling between thechromophores and allows the triplets to separate from thecorrelated triplet pair state. Such rotation of the tetracene unitsin the ortho dimer is hindered because they are confined to acofacial orientation. The role of tetracenes rotating to formseparated triplets is clear when comparing the transientabsorption of the para dimer in PMMA to that in THF.PMMA being a more rigid medium prevents the rotation andsignificantly hinders the formation of the two noninteractingtriplets.A comparison of the two para-bis(ethynyltetracenyl)-

benzene dimers (p-BETB-ohex vs p-BETB-ehex) providesinsight into the role of the energetics of the linker in singletfission in covalent dimers. The hexyloxy-bis(ethynyl)benzenelinker in p-BETB-ohex has molecular orbitals that are closer inenergy to orbitals of tetracenes than the 2-ethylhexyl-bis(ethynyl)benzene linker in p-BETB-ehex. This results inbetter mixing of the linker orbitals with those of the tetracenesand thus better coupling in p-BETB-ohex than in p-BETB-ehex. Better coupling between the tetracene units results in

Figure 12. A diagram showing the closer lying orbitals of the linkerand the tetracenes in (a) p-BETB-ehex as compared to (b) p-BETB-ohex. la/b and ha/b are the frontier molecular orbitals of tetracene.

Journal of the American Chemical Society Article

DOI: 10.1021/jacs.8b04401J. Am. Chem. Soc. 2018, 140, 10179−10190

10187

Page 10: Linker-Dependent Singlet Fission in Tetracene Dimersiopenshell.usc.edu/pubs/pdf/jacs-140-10179.pdfLinker-Dependent Singlet Fission in Tetracene Dimers ... Queisser limit, the theoretical

faster formation of the correlated triplet pair state and theseparated triplets in p-BETB-ohex than in p-BETB-ehex,highlighting the role of linker energetics in singlet fission.In separate experiments reported elsewhere, a comparison of

the excited-state dynamics of the meta and para dimers in neatthin films91 has allowed us to separately probe intra- and inter-dimer singlet fission between tetracene units. While the metadimer does not fission in THF solution, in film interdimersinglet fission is turned on facilitated by through-spaceinteractions between the ethynyltetracene units on adjacentdimers. In films of the para dimer, singlet fission was facilitatedby both intra-dimer through-bond coupling, and inter-dimerthrough-space coupling, such that the depopulation of thesinglet state happened at least twice (within the instrumentresponse) as fast as compared to in THF solution.

■ ASSOCIATED CONTENT*S Supporting InformationThe Supporting Information is available free of charge on theACS Publications website at DOI: 10.1021/jacs.8b04401.

Synthetic details, structural characterization of thematerials, electrochemical data, details of samplepreparation for spectroscopy, and excited-state kineticmodels (PDF)

■ AUTHOR INFORMATIONCorresponding Authors*[email protected]*[email protected] I. Krylov: 0000-0001-6788-5016Stephen E. Bradforth: 0000-0002-6164-3347Mark E. Thompson: 0000-0002-7764-4096Present Addresses‡National Renewable Energy Laboratory, 15013 Denver WestParkway, Golden, Colorado 80401, United States§Department of Chemistry, University of California, Berkeley,California 94720, United States; and Q-Chem Inc., 6601Owens Drive, Suite 105, Pleasanton, California 94588, UnitedStatesAuthor Contributions†N.V.K. and J.J. contributed equally.NotesThe authors declare no competing financial interest.

■ ACKNOWLEDGMENTSM.E.T., N.V.K., S.E.B., and J.J. acknowledge support from theCenter for Energy Nanoscience, an Energy Frontier ResearchCenter funded by the U.S. Office of Science, Office of BasicEnergy Sciences (DE SC0001013). M.E.T. and N.V.K.additionally were supported by the Nanoflex Power Corp.A.I.K. acknowledges support through the DE FG0205ER15685 grant (Department of Energy).

■ REFERENCES(1) Smith, M. B.; Michl, J. Chem. Rev. 2010, 110 (11), 6891−6936.(2) Piland, G. B.; Burdett, J. J.; Dillon, R. J.; Bardeen, C. J. J. Phys.Chem. Lett. 2014, 5 (13), 2312−2319.(3) Casanova, D. Chem. Rev. 2018, DOI: 10.1021/acs.chem-rev.7b00601.(4) Hanna, M. C.; Nozik, A. J. J. Appl. Phys. 2006, 100 (7), 074510.

(5) Tayebjee, M. J. Y.; Gray-Weale, A. A.; Schmidt, T. W. J. Phys.Chem. Lett. 2012, 3 (19), 2749−2754.(6) Congreve, D. N.; Lee, J.; Thompson, N. J.; Hontz, E.; Yost, S. R.;Reusswig, P. D.; Bahlke, M. E.; Reineke, S.; Van Voorhis, T.; Baldo,M. A. Science 2013, 340 (6130), 334−337.(7) Schrauben, J. N.; Zhao, Y.; Mercado, C.; Dron, P. I.; Ryerson, J.L.; Michl, J.; Zhu, K.; Johnson, J. C. ACS Appl. Mater. Interfaces 2015,7 (4), 2286−2293.(8) Kolomeisky, A. B.; Feng, X.; Krylov, A. I. J. Phys. Chem. C 2014,118 (10), 5188−5195.(9) Feng, X.; Kolomeisky, A. B.; Krylov, A. I. J. Phys. Chem. C 2014,118 (34), 19608−19617.(10) Matsika, S.; Feng, X.; Luzanov, A. V.; Krylov, A. I. J. Phys.Chem. A 2014, 118 (51), 11943−11955.(11) Margulies, E. A.; Miller, C. E.; Wu, Y.; Ma, L.; Schatz, G. C.;Young, R. M.; Wasielewski, M. R. Nat. Chem. 2016, 8, 1120.(12) Sung, J.; Nowak-Krol, A.; Schlosser, F.; Fimmel, B.; Kim, W.;Kim, D.; Wurthner, F. J. Am. Chem. Soc. 2016, 138 (29), 9029−9032.(13) Schrauben, J. N.; Akdag, A.; Wen, J.; Havlas, Z.; Ryerson, J. L.;Smith, M. B.; Michl, J.; Johnson, J. C. J. Phys. Chem. A 2016, 120(20), 3473−3483.(14) Freeman, D. M. E.; Musser, A. J.; Frost, J. M.; Stern, H. L.;Forster, A. K.; Fallon, K. J.; Rapidis, A. G.; Cacialli, F.; McCulloch, I.;Clarke, T. M.; Friend, R. H.; Bronstein, H. J. Am. Chem. Soc. 2017,139 (32), 11073−11080.(15) Wu, Y.; Wang, Y.; Chen, J.; Zhang, G.; Yao, J.; Zhang, D.; Fu,H. Angew. Chem., Int. Ed. 2017, 56 (32), 9400−9404.(16) Mauck Catherine, M.; Bae Youn, J.; Chen, M.; Powers-Riggs,N.; Wu, Y. L.; Wasielewski Michael, R. ChemPhotoChem. 2018, 2 (3),223−233.(17) Pace, N. A.; Zhang, W.; Arias, D. H.; McCulloch, I.; Rumbles,G.; Johnson, J. C. J. Phys. Chem. Lett. 2017, 8 (24), 6086−6091.(18) Sakuma, T.; Sakai, H.; Araki, Y.; Mori, T.; Wada, T.;Tkachenko, N. V.; Hasobe, T. J. Phys. Chem. A 2016, 120 (11),1867−1875.(19) Zirzlmeier, J.; Casillas, R.; Reddy, S. R.; Coto, P. B.; Lehnherr,D.; Chernick, E. T.; Papadopoulos, I.; Thoss, M.; Tykwinski, R. R.;Guldi, D. M. Nanoscale 2016, 8 (19), 10113−10123.(20) Sanders, Samuel N.; Kumarasamy, E.; Pun, Andrew B.;Steigerwald, Michael L.; Sfeir, Matthew Y.; Campos, Luis M. Chem.2016, 1 (3), 505−511.(21) Cook, J. D.; Carey, T. J.; Damrauer, N. H. J. Phys. Chem. A2016, 120 (26), 4473−4481.(22) Carey, T. J.; Snyder, J. L.; Miller, E. G.; Sammakia, T.;Damrauer, N. H. J. Org. Chem. 2017, 82 (9), 4866−4874.(23) Trinh, M. T.; Pinkard, A.; Pun, A. B.; Sanders, S. N.;Kumarasamy, E.; Sfeir, M. Y.; Campos, L. M.; Roy, X.; Zhu, X. Y.Science Advances 2017, 3, 7.(24) Lehnherr, D.; Alzola, J. M.; Mulzer, C. R.; Hein, S. J.; Dichtel,W. R. J. Org. Chem. 2017, 82 (4), 2004−2010.(25) Liu, H.; Wang, R.; Shen, L.; Xu, Y.; Xiao, M.; Zhang, C.; Li, X.Org. Lett. 2017, 19 (3), 580−583.(26) Kumarasamy, E.; Sanders, S. N.; Tayebjee, M. J. Y.;Asadpoordarvish, A.; Hele, T. J. H.; Fuemmeler, E. G.; Pun, A. B.;Yablon, L. M.; Low, J. Z.; Paley, D. W.; Dean, J. C.; Choi, B.; Scholes,G. D.; Steigerwald, M. L.; Ananth, N.; McCamey, D. R.; Sfeir, M. Y.;Campos, L. M. J. Am. Chem. Soc. 2017, 139 (36), 12488−12494.(27) Pun Andrew, B.; Sanders Samuel, N.; Kumarasamy, E.; SfeirMatthew, Y.; Congreve Daniel, N.; Campos Luis, M. Adv. Mater.2017, 29 (41), 1701416.(28) Dean, J. C.; Zhang, R.; Hallani, R. K.; Pensack, R. D.; Sanders,S. N.; Oblinsky, D. G.; Parkin, S. R.; Campos, L. M.; Anthony, J. E.;Scholes, G. D. Phys. Chem. Chem. Phys. 2017, 19 (34), 23162−23175.(29) Cook, J. D.; Carey, T. J.; Arias, D. H.; Johnson, J. C.; Damrauer,N. H. J. Phys. Chem. A 2017, 121 (48), 9229−9242.(30) Bettinger, H. F.; Einholz, R.; Gottler, A.; Junge, M.; Sattele, M.-S.; Schnepf, A.; Schrenk, C.; Schundelmeier, S.; Speiser, B. Org. Chem.Front. 2017, 4 (5), 853−860.

Journal of the American Chemical Society Article

DOI: 10.1021/jacs.8b04401J. Am. Chem. Soc. 2018, 140, 10179−10190

10188

Page 11: Linker-Dependent Singlet Fission in Tetracene Dimersiopenshell.usc.edu/pubs/pdf/jacs-140-10179.pdfLinker-Dependent Singlet Fission in Tetracene Dimers ... Queisser limit, the theoretical

(31) Wang, X.; Wang, R.; Shen, L.; Tang, Z.; Wen, C.; Dong, B.; Liu,H.; Zhang, C.; Li, X. Phys. Chem. Chem. Phys. 2018, 20 (9), 6330−6336.(32) Hetzer, C.; Guldi Dirk, M.; Tykwinski Rik, R. Chem. - Eur. J.2018, 24, 1−14.(33) Carey, T. J.; Miller, E. G.; Gilligan, A. T.; Sammakia, T.;Damrauer, N. H. Org. Lett. 2018, 20 (2), 457−460.(34) Wang, L.; Wu, Y.; Chen, J.; Wang, L.; Liu, Y.; Yu, Z.; Yao, J.;Fu, H. J. Phys. Chem. Lett. 2017, 8 (22), 5609−5615.(35) Liu, H.; Wang, X.; Pan, L.; Shen, L.; Wang, X.; Chen, Q.; Li, X.J. Photochem. Photobiol., A 2017, 340, 21−28.(36) Yamakado, T.; Takahashi, S.; Watanabe, K.; Matsumoto, Y.;Osuka, A.; Saito, S. Angew. Chem., Int. Ed. 2018, 57, 5438−5448.(37) Zirzlmeier, J.; Lehnherr, D.; Coto, P. B.; Chernick, E. T.;Casillas, R.; Basel, B. S.; Thoss, M.; Tykwinski, R. R.; Guldi, D. M.Proc. Natl. Acad. Sci. U. S. A. 2015, 112 (17), 5325−5330.(38) Muller, A. M.; Avlasevich, Y. S.; Mullen, K.; Bardeen, C. J.Chem. Phys. Lett. 2006, 421 (4), 518−522.(39) Johnson, J. C.; Akdag, A.; Zamadar, M.; Chen, X.; Schwerin, A.F.; Paci, I.; Smith, M. B.; Havlas, Z.; Miller, J. R.; Ratner, M. A.;Nozik, A. J.; Michl, J. J. Phys. Chem. B 2013, 117 (16), 4680−4695.(40) Muller, A. M.; Avlasevich, Y. S.; Schoeller, W. W.; Mullen, K.;Bardeen, C. J. J. Am. Chem. Soc. 2007, 129 (46), 14240−14250.(41) Yamakado, T.; Takahashi, S.; Watanabe, K.; Matsumoto, Y.;Osuka, A.; Saito, S. Angew. Chem., Int. Ed. 2018, 57 (19), 5438−5443.(42) Grieco, C.; Kennehan, E. R.; Kim, H.; Pensack, R. D.;Brigeman, A. N.; Rimshaw, A.; Payne, M. M.; Anthony, J. E.; Giebink,N. C.; Scholes, G. D.; Asbury, J. B. J. Phys. Chem. C 2018, 122 (4),2012−2022.(43) Folie, B. D.; Haber, J. B.; Refaely-Abramson, S.; Neaton, J. B.;Ginsberg, N. S. J. Am. Chem. Soc. 2018, 140 (6), 2326−2335.(44) Basel, B. S.; Zirzlmeier, J.; Hetzer, C.; Reddy, S. R.; Phelan, B.T.; Krzyaniak, M. D.; Volland, M. K.; Coto, P. B.; Young, R. M.;Clark, T.; Thoss, M.; Tykwinski, R. R.; Wasielewski, M. R.; Guldi, D.M. Chem. 2018, 4 (5), 1092−1111.(45) Hart, S.; Silva, W. R.; Frontiera, R. R. Chemical Science 2018, 9(5), 1242−1250.(46) Ito, S.; Nagami, T.; Nakano, M. J. Phys. Chem. A 2016, 120(31), 6236−6241.(47) Zeng, T. J. Phys. Chem. Lett. 2016, 7 (21), 4405−4412.(48) Feng, X.; Luzanov, A. V.; Krylov, A. I. J. Phys. Chem. Lett. 2013,4 (22), 3845−3852.(49) Korovina, N. V.; Das, S.; Nett, Z.; Feng, X.; Joy, J.; Haiges, R.;Krylov, A. I.; Bradforth, S. E.; Thompson, M. E. J. Am. Chem. Soc.2016, 138 (2), 617−627.(50) Perkampus, H.; Sandeman, I.; Timmons, C. DMS UV Atlas ofOrganic Compounds; Verlag Chemie: Weinheim, Butterworths,London, England, 1966.(51) Pensack, R. D.; Tilley, A. J.; Parkin, S. R.; Lee, T. S.; Payne, M.M.; Gao, D.; Jahnke, A. A.; Oblinsky, D.; Li, P.-F.; Anthony, J. E. J.Am. Chem. Soc. 2015, 137, 6790−6803.(52) Murov, S. L.; Carmichael, I.; Hug, G. L. Handbook ofPhotochemistry; CRC Press: New York, 1993.(53) Sanders, S. N.; Kumarasamy, E.; Pun, A. B.; Trinh, M. T.; Choi,B.; Xia, J.; Taffet, E. J.; Low, J. Z.; Miller, J. R.; Roy, X.; Zhu, X. Y.;Steigerwald, M. L.; Sfeir, M. Y.; Campos, L. M. J. Am. Chem. Soc.2015, 137 (28), 8965−8972.(54) Roberts, S. T.; McAnally, R. E.; Mastron, J. N.; Webber, D. H.;Whited, M. T.; Brutchey, R. L.; Thompson, M. E.; Bradforth, S. E. J.Am. Chem. Soc. 2012, 134 (14), 6388−6400.(55) Reusswig, P. D.; Congreve, D. N.; Thompson, N. J.; Baldo, M.A. Appl. Phys. Lett. 2012, 101 (11), 113304.(56) Rolczynski, B. S.; Szarko, J. M.; Son, H. J.; Liang, Y.; Yu, L.;Chen, L. X. J. Am. Chem. Soc. 2012, 134 (9), 4142−4152.(57) Roberts, S. T.; Schlenker, C. W.; Barlier, V.; McAnally, R. E.;Zhang, Y.; Mastron, J. N.; Thompson, M. E.; Bradforth, S. E. J. Phys.Chem. Lett. 2011, 2 (2), 48−54.

(58) Ryerson, J. L.; Schrauben, J. N.; Ferguson, A. J.; Sahoo, S. C.;Naumov, P.; Havlas, Z.; Michl, J.; Nozik, A. J.; Johnson, J. C. J. Phys.Chem. C 2014, 118 (23), 12121−12132.(59) Ryerson, J. L.; Schrauben, J. N.; Ferguson, A. J.; Sahoo, S. C.;Naumov, P. e.; Havlas, Z. k.; Michl, J.; Nozik, A. J.; Johnson, J. C. J.Phys. Chem. C 2014, 118 (23), 12121−12132.(60) Thompson, N. J.; Wilson, M. W.; Congreve, D. N.; Brown, P.R.; Scherer, J. M.; Bischof, T. S.; Wu, M.; Geva, N.; Welborn, M.; VanVoorhis, T. Nat. Mater. 2014, 13, 1039−1043.(61) Thorsmølle, V. K.; Averitt, R. D.; Demsar, J.; Smith, D.;Tretiak, S.; Martin, R.; Chi, X.; Crone, B.; Ramirez, A.; Taylor, A.Phys. Rev. Lett. 2009, 102 (1), 017401.(62) Muller, A. M.; Avlasevich, Y. S.; Schoeller, W. W.; Mullen, K.;Bardeen, C. J. J. Am. Chem. Soc. 2007, 129 (46), 14240−14250.(63) Feng, X.; Krylov, A. I. Phys. Chem. Chem. Phys. 2016, 18 (11),7751−7761.(64) Feng, X.; Casanova, D.; Krylov, A. I. J. Phys. Chem. C 2016, 120(34), 19070−19077.(65) Abraham, V.; Mayhall, N. J. J. Phys. Chem. Lett. 2017, 8 (22),5472−5478.(66) Sonogashira, K. J. Organomet. Chem. 2002, 653 (1), 46−49.(67) Yost, S. R.; Lee, J.; Wilson, M. W. B.; Wu, T.; McMahon, D. P.;Parkhurst, R. R.; Thompson, N. J.; Congreve, D. N.; Rao, A.; Johnson,K.; Sfeir, M. Y.; Bawendi, M.; Swager, T. M.; Friend, R. H.; A, M.; vanVoorhis, T. Nat. Chem. 2014, 6, 492.(68) Grabowski, Z. R.; Rotkiewicz, K.; Rettig, W. Chem. Rev. 2003,103 (10), 3899−4032.(69) Berkelbach, T. C.; Hybertsen, M. S.; Reichman, D. R. J. Chem.Phys. 2013, 138 (11), 114103.(70) Lukman, S.; Musser, A. J.; Chen, K.; Athanasopoulos, S.; Yong,C. K.; Zeng, Z.; Ye, Q.; Chi, C.; Hodgkiss, J. M.; Wu, J.; Friend, R. H.;Greenham, N. C. Adv. Funct. Mater. 2015, 25 (34), 5452−5461.(71) Lippitz, M.; Hubner, C. G.; Christ, T.; Eichner, H.; Bordat, P.;Herrmann, A.; Mullen, K.; Basche, T. Phys. Rev. Lett. 2004, 92 (10),103001.(72) Nguyen, M.-H.; Nguyen, V. H.; Yip, J. H. Organometallics 2013,32 (24), 7283−7291.(73) Burdett, J. J.; Bardeen, C. J. Acc. Chem. Res. 2013, 46 (6),1312−1320.(74) Arias, D. H.; Ryerson, J. L.; Cook, J. D.; Damrauer, N. H.;Johnson, J. C. Chemical Science 2016, 7 (2), 1185−1191.(75) Marciniak, H.; Pugliesi, I.; Nickel, B.; Lochbrunner, S. Phys.Rev. B: Condens. Matter Mater. Phys. 2009, 79 (23), 235318.(76) Montalti, M.; Credi, A.; Prodi, L.; Gandolfi, M. T. Handbook ofPhotochemistry, 3rd ed.; CRC Press: Boca Raton, FL, 2006.(77) McQuade, D. T.; Pullen, A. E.; Swager, T. M. Chem. Rev. 2000,100, 2537.(78) van Stokkum, I. H. M.; Larsen, D. S.; van Grondelle, R.Biochim. Biophys. Acta, Bioenerg. 2004, 1657 (2), 82−104.(79) McQuade, D. T.; Kim, J.; Swager, T. M. J. Am. Chem. Soc. 2000,122, 5885.(80) Van Dyke, D. A.; Pryor, B. A.; Smith, P. G.; Topp, M. R. J.Chem. Educ. 1998, 75 (5), 615.(81) Giaimo, J. M.; Lockard, J. V.; Sinks, L. E.; Scott, A. M.; Wilson,T. M.; Wasielewski, M. R. J. Phys. Chem. A 2008, 112 (11), 2322−2330.(82) Miller, C. E.; Wasielewski, M. R.; Schatz, G. C. J. Phys. Chem. C2017, 121 (19), 10345−10350.(83) Breen, I.; Tempelaar, R.; Bizimana, L. A.; Kloss, B.; Reichman,D. R.; Turner, D. B. J. Am. Chem. Soc. 2017, 139 (34), 11745−11751.(84) Hetzer, C.; Guldi Dirk, M.; Tykwinski Rik, R. Chem. - Eur. J.2018, 0 (0), 8245.(85) Monahan, N.; Zhu, X. Y. Annu. Rev. Phys. Chem. 2015, 66 (1),601−618.(86) Zimmerman, P. M.; Bell, F.; Casanova, D.; Head-Gordon, M. J.Am. Chem. Soc. 2011, 133 (49), 19944−19952.(87) Aryanpour, K.; Shukla, A.; Mazumdar, S. J. Phys. Chem. C 2015,119 (13), 6966−6979.

Journal of the American Chemical Society Article

DOI: 10.1021/jacs.8b04401J. Am. Chem. Soc. 2018, 140, 10179−10190

10189

Page 12: Linker-Dependent Singlet Fission in Tetracene Dimersiopenshell.usc.edu/pubs/pdf/jacs-140-10179.pdfLinker-Dependent Singlet Fission in Tetracene Dimers ... Queisser limit, the theoretical

(88) Beljonne, D.; Yamagata, H.; Bredas, J. L.; Spano, F. C.; Olivier,Y. Phys. Rev. Lett. 2013, 110 (22), 226402.(89) Liu, H.; Wang, Z.; Wang, X.; Shen, L.; Zhang, C.; Xiao, M.; Li,X. J. Mater. Chem. C 2018, 6 (13), 3245−3253.(90) Tayebjee, M. J. Y.; Sanders, S. N.; Kumarasamy, E.; Campos, L.M.; Sfeir, M. Y.; McCamey, D. R. Nat. Phys. 2017, 13, 182.(91) Joy, J.; Korovina, N. V.; Thompson, M. E.; Bradforth, S. E. Proc.SPIE 10348, Physical Chemistry of Semiconductor Materials andInterfaces XVI 2017, 24.

Journal of the American Chemical Society Article

DOI: 10.1021/jacs.8b04401J. Am. Chem. Soc. 2018, 140, 10179−10190

10190