Abstract

Optical interconnects transmit signals between different parts of integrated circuits using light and have lower power consumption and latency than the conventional metal interconnects. The modulation frequency of the on-chip light sources in optical interconnects determines the signal transmission speed. In principle, on-chip lasers relying on stimulated emission can have high modulation frequency but also have high fabrication complexity and energy consumption. Thus, it is desirable to use light-emitting devices that operate through spontaneous emission as the on-chip light source. However, the operation speed of these devices is limited by the radiative lifetime of the emitting materials. In this perspective, we envision that elaborately designed fluorescence molecules can give rise to materials that have radiative lifetime shorter by several orders of magnitude than conventional inorganic semiconductors through a phenomenon termed superradiance. Optical communication and interconnects utilize high-speed light sources for information transmission. The modulation frequency of light-emitting devices operating through spontaneous emission is fundamentally limited by the material intrinsic radiative lifetime. In this perspective, we examine the radiative lifetime of different materials and identify the superradiant molecular J aggregates as a promising class of materials for high-speed light-emitting devices. These molecular aggregates are relatively unexplored for electroluminescent devices and can have short radiative lifetime on the order of 10 ps while maintaining high photoluminescence quantum yield. The relation between intermolecular interactions, molecular packing geometry, and radiative lifetimes is presented theoretically in the context of Frenkel excitons and is corroborated with experimental examples. We further demonstrate the potential of designing superradiant materials through molecular engineering. We believe that these superradiant molecular materials will open up new opportunities in the fabrication of efficient and high-speed light-emitting devices. Optical communication and interconnects utilize high-speed light sources for information transmission. The modulation frequency of light-emitting devices operating through spontaneous emission is fundamentally limited by the material intrinsic radiative lifetime. In this perspective, we examine the radiative lifetime of different materials and identify the superradiant molecular J aggregates as a promising class of materials for high-speed light-emitting devices. These molecular aggregates are relatively unexplored for electroluminescent devices and can have short radiative lifetime on the order of 10 ps while maintaining high photoluminescence quantum yield. The relation between intermolecular interactions, molecular packing geometry, and radiative lifetimes is presented theoretically in the context of Frenkel excitons and is corroborated with experimental examples. We further demonstrate the potential of designing superradiant materials through molecular engineering. We believe that these superradiant molecular materials will open up new opportunities in the fabrication of efficient and high-speed light-emitting devices. Optical communications and interconnects rely on high-frequency modulation light sources for high-speed information transmission. The commonly used light sources include lasers and light-emitting diodes. Lasers can be modulated at high speed but require complex fabrication steps and are energy intensive. Light-emitting devices operating thorough spontaneous emission can have reduced fabrication complexity and energy consumption, and devices with modulation frequency above 1 GHz have been reported.1Koester R. Sager D. Quitsch W.-A. Pfingsten O. Poloczek A. Blumenthal S. Keller G. Prost W. Bacher G. Tegude F.-J. High-speed GaN/GaInN nanowire array light-emitting diode on silicon(111).Nano Lett. 2015; 15: 2318-2323Crossref PubMed Scopus (85) Google Scholar,2Shi J. Chi K. Wun J. Bowers J. Sheu J. GaN based cyan light-emitting diodes with GHz bandwidth.in: IEEE Photonics Conference (IPC 2016). IEEE, 2016: 623-624https://doi.org/10.1109/IPCon.2016.7831257Crossref Scopus (4) Google Scholar However, the speed of these devices is fundamentally limited by the spontaneous radiative lifetime of the light-emitting materials. Although the radiative lifetime can be reduced with well-designed cavities through the Purcell effect,3Noda S. Fujita M. Asano T. Spontaneous-emission control by photonic crystals and nanocavities.Nat. Photon. 2007; 1: 449-458Crossref Scopus (808) Google Scholar,4Fortuna S.A. Heidelberger C. Yablonovitch E. Fitzgerald E.A. Wu M.C. Nanoscale III-V light emitting diode with antenna-enhanced 250 picosecond spontaneous emission lifetime.in: IEEE International Semiconductor Laser Conference (ISLC). IEEE, 2018https://doi.org/10.1109/ISLC.2018.8516244Crossref Scopus (1) Google Scholar its implementation is material specific and introduces complications to device fabrication. Thus, developing materials with intrinsically short radiative lifetime is fundamental to increasing the speed of light-emitting devices. As the luminescent quantum yield is determined by the competition of radiative and non-radiative processes, materials with short radiative lifetime produce high modulation frequency while maintaining high electroluminescent quantum yields. For inorganic materials, their radiative lifetimes are generally determined by the crystal structures and can hardly be reduced below a few nanoseconds even with material engineering methods such as doping or downsizing. Molecular materials can have radiative lifetimes down to tens of picoseconds, which can be further reduced by molecular and crystal engineering. With favorable packing geometry, superradiance can be achieved in molecular aggregates where the radiative lifetime scales inversely with grain size. Thus, designing superradiant molecular crystals can lead to fundamental advancement in high-speed light-emitting devices. In light-emitting devices, the generation of light takes place when electrons and holes recombine radiatively at a rate that depends on the radiative lifetime. For some inorganic semiconductors, the high dielectric constant largely screens the Coulomb attraction between the electrons and holes, which can thus move independently in the conduction and valence band at room temperature. The radiative lifetime of these free carriers depends on the radiative recombination coefficient and the carrier concentration, which are related to the material band structure and doping level. Direct band-gap III–V semiconductors with moderate doping levels typically have radiative lifetimes of 10–100 ns.5Henry C. Levine B. Logan R. Bethea C. Minority carrier lifetime and luminescence efficiency of 1.3 μm InGaAsP-InP double heterostructure layers.IEEE J. Quan. Electron. 1983; 19: 905-912Crossref Scopus (51) Google Scholar, 6Im J.S. Moritz A. Steuber F. Härle V. Scholz F. Hangleiter A. Radiative carrier lifetime, momentum matrix element, and hole effective mass in GaN.Appl. Phys. Lett. 1997; 70: 631-633Crossref Scopus (212) Google Scholar, 7Juršėnas S. Miasojedovas S. Zukauskas A. Rate of radiative and nonradiative recombination in bulk GaN grown by various techniques.J. Cryst. Growth. 2005; 281: 161-167Crossref Scopus (10) Google Scholar, 8Lush G.B. MacMillan H.F. Keyes B.M. Levi D.H. Melloch M.R. Ahrenkiel R.K. Lundstrom M.S. A study of minority carrier lifetime versus doping concentration in n-type GaAs grown by metalorganic chemical vapor deposition.J. Appl. Phys. 1992; 72: 1436-1442Crossref Scopus (60) Google Scholar, 9Mickevičius J. Shur M.S. Fareed R.S.Q. Zhang J.P. Gaska R. Tamulaitis G. Time-resolved experimental study of carrier lifetime in GaN epilayers.Appl. Phys. Lett. 2005; 87: 241918Crossref Scopus (33) Google Scholar, 10Muth J.F. Lee J.H. Shmagin I.K. Kolbas R.M. Casey Jr., H.C. Keller B.P. Mishra U.K. DenBaars S.P. Absorption coefficient, energy gap, exciton binding energy, and recombination lifetime of GaN obtained from transmission measurements.Appl. Phys. Lett. 1997; 71: 2572-2574Crossref Scopus (627) Google Scholar, 11Semyonov O. Subashiev A. Chen Z. Luryi S. Radiation efficiency of heavily doped bulk n-InP semiconductor.J. Appl. Phys. 2010; 108: 013101Crossref Scopus (24) Google Scholar, 12Vilms J. Spicer W.E. Quantum efficiency and radiative lifetime in p-type gallium arsenide.J. Appl. Phys. 1965; 36: 2815-2821Crossref Scopus (105) Google Scholar Although the radiative lifetime would decrease with doping, heavy doping can result in non-radiative recombination by introducing trap states and promoting Auger processes (Figure 1A). When the thermal energy of the electrons and holes are not enough to overcome their Coulomb attraction, they become bound electron-hole pairs and behave as neutral quasiparticles termed excitons. The excitons can be categorized by their radius, which is the distance between the bound electron and hole. In inorganic semiconductors, due to the screened Coulomb interaction and large orbital overlap between atoms, the excitons typically have radius larger than the size of a lattice unit cell and are categorized as Wannier-Mott excitons. The bound electrons and holes have lower energy than the free carriers (Figure 1B) in the conduction and valence band corresponding to their binding energy, which is on the order of 10 meV and can be increased by quantum confinement.13Elward J.M. Chakraborty A. Effect of dot size on exciton binding energy and electron-hole recombination probability in CdSe quantum dots.J. Chem. Theor. Comput. 2013; 9: 4351-4359Crossref PubMed Scopus (61) Google Scholar,14Meulenberg R.W. Lee J.R.I. Wolcott A. Zhang J.Z. Terminello L.J. van Buuren T. Determination of the exciton binding energy in CdSe quantum dots.ACS Nano. 2009; 3: 325-330Crossref PubMed Scopus (140) Google Scholar II–VI quantum dots are a class of materials that host Wannier-Mott excitons, and their radiative lifetime is on the order of 10 ns. Although their radiative lifetime can be reduced by decreasing the particle size,13Elward J.M. Chakraborty A. Effect of dot size on exciton binding energy and electron-hole recombination probability in CdSe quantum dots.J. Chem. Theor. Comput. 2013; 9: 4351-4359Crossref PubMed Scopus (61) Google Scholar,15Gong K. Martin J.E. Shea-Rohwer L.E. Lu P. Kelley D.F. Radiative lifetimes of zincblende CdSe/CdS quantum dots.J. Phys. Chem. C. 2015; 119: 2231-2238Crossref Scopus (49) Google Scholar below a certain size (∼2 nm) the quantum dots become chemically unstable and defective. The excitons can also form trions (with charged carriers) or biexcitons (with another exciton), which can have different radiative and non-radiative lifetimes.16Sahin M. Koç F. A model for the recombination and radiative lifetime of trions and biexcitons in spherically shaped semiconductor nanocrystals.Appl. Phys. Lett. 2013; 102: 183103Crossref Scopus (19) Google Scholar,17Jha P.P. Guyot-Sionnest P. Trion decay in colloidal quantum dots.ACS Nano. 2009; 3: 1011-1015Crossref PubMed Scopus (237) Google Scholar For II–VI quantum dots, the radiative lifetime for trions and biexcitons can be twice as short as for the excitons, but still on the order of several nanoseconds. The photophysics of excitons in molecular materials can be understood in the context of molecular orbitals (Figures 1C and 1D).18Hestand N.J. Spano F.C. Expanded theory of H- and J-molecular aggregates: the effects of vibronic coupling and intermolecular charge transfer.Chem. Rev. 2018; 118: 7069-7163Crossref PubMed Scopus (590) Google Scholar, 19Bardeen C.J. The structure and dynamics of molecular excitons.Annu. Rev. Phys. Chem. 2014; 65: 127-148Crossref PubMed Scopus (167) Google Scholar, 20Frenkel J. On the transformation of light into heat in solids.Phys. Rev. 1931; 37: 17-44Crossref Scopus (491) Google Scholar An isolated molecule can be conceptually simplified as a two-level system with the two energy levels being the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO, Figure 1D). A molecule at its ground state typically has a pair of electrons with opposite spin occupying the HOMO, which is a singlet state denoted S0. With optical excitation, one electron in the HOMO can be excited into the LUMO without changing its spin. The resulting singlet excited state (S1) is conceptually equivalent to having one electron in LUMO and one hole in HOMO. In molecular solids, these excitations correspond to Frenkel excitons where the electron and hole reside in the same molecule. The radiative lifetime of a molecule is determined by the transition rate between the two states, which is proportional to the transition dipole moment μ (Equation 1). Here ϕg and ϕe denote the molecular ground and excited state wavefunction, q is the elemental charge, r→ is the position operator, and kr is the rate of radiative recombination and the inverse of radiative lifetime. For molecules in singlet excited states, the recombination between the electron and hole gives fluorescence, which has radiative lifetime on the order of 1–100 ns. Besides radiative decay, the singlet excited state can also relax to the triplet state through intersystem crossing, whereby the electron excited into the LUMO would change its spin so that its spin is the same as the electron remaining in the HOMO. Due to the spin-pairing energy, the triplet state would have a lower energy compared with the singlet state. The triplet state can decay to ground state via phosphorescence, which is spin forbidden and has radiative lifetime beyond hundreds of nanoseconds.μ→=⟨ϕg|qr→|ϕe⟩kr=1τrad∝43ν3ℏc3|μ→|2.(Equation 1) The optoelectronic properties of molecular aggregates and crystals are different from isolated molecules due to the many-body Coulomb interactions between the electrons and nuclei of different molecules. When the intermolecular distance is significantly greater than the separation of charges within a molecule, Coulomb interaction can be approximated by the electrostatic interaction of electric dipoles. These materials host Frenkel excitons, which are molecular excitations delocalized in the material with each molecule having a certain possibility to be in the excited state. When the molecules are placed closer to each other, their molecular orbitals start to overlap. In these materials, the intermolecular interactions depend on the shape, phase, and orientation of the molecular orbitals of adjacent molecules and require quantum chemical treatment. The orbital overlap between nearby molecules facilitates charge transfer and delocalization. As a result, an electron from the HOMO of one molecule can be excited into the LUMO of other molecules in the material. This type of excitation gives charge-transfer excitons, which have the electron and hole bound to two different molecules. In actual molecular aggregates, although the intermolecular distance is comparable with the molecular size, the far field dipole approximation of Coulomb interaction is largely valid in materials without significant charge-transfer characteristics. For aromatic molecules with π conjugation, π-π stacking could lead to substantial orbital overlap between adjacent molecules; thus, charge-transfer excitons need to be considered when modeling their optical properties. As the molecules have singlet and triplet excited states, Frenkel excitons in solids have singlet and triplet bands. Phosphorescence from the triplet band, similar to the phosphorescence from molecules, has long lifetime beyond 100 ns. Efficient phosphorescence typically involves heavy atoms that promote spin-orbit coupling.21Sommer J.R. Shelton A.H. Parthasarathy A. Ghiviriga I. Reynolds J.R. Schanze K.S. Photophysical properties of near-infrared phosphorescent π-extended platinum porphyrins.Chem. Mater. 2011; 23: 5296-5304Crossref Scopus (114) Google Scholar,22Sarma M. Tsai W.-L. Lee W.-K. Chi Y. Wu C.-C. Liu S.-H. Chou P.-T. Wong K.-T. Anomalously long-lasting blue PhOLED featuring phenyl-pyrimidine cyclometalated iridium emitter.Chem. 2017; 3: 461-476Abstract Full Text Full Text PDF Scopus (68) Google Scholar The electron excited into the singlet band would first decay to the lowest energy state in the exciton band (internal conversion) before transition to the ground state via fluorescence,23Kasha M. Characterization of electronic transitions in complex molecules.Faraday Discuss. 1950; 9: 14-19Crossref Google Scholar which can have radiative lifetime down to nanoseconds. With increased wavefunction overlaps between the molecules, the charge-transfer exciton states can have energy low enough to mix with the Frenkel exciton and lower the energy of the first exciton band. Thus, these materials have red-shifted optical spectra, which are also broad due to the coupling to intermolecular vibrational modes.18Hestand N.J. Spano F.C. Expanded theory of H- and J-molecular aggregates: the effects of vibronic coupling and intermolecular charge transfer.Chem. Rev. 2018; 118: 7069-7163Crossref PubMed Scopus (590) Google Scholar,19Bardeen C.J. The structure and dynamics of molecular excitons.Annu. Rev. Phys. Chem. 2014; 65: 127-148Crossref PubMed Scopus (167) Google Scholar Charge-transfer excitons tend to have longer radiative lifetime than Frenkel excitons due to larger separations of the electrons and holes.24Katoh R. Suzuki K. Furube A. Kotani M. Tokumaru K. Fluorescence quantum yield of aromatic hydrocarbon crystals.J. Phys. Chem. C. 2009; 113: 2961-2965Crossref Scopus (192) Google Scholar|k⟩=1N∑neikn|n⟩k=0,±2πN,…,π|n⟩=ϕen∏i≠nϕgi,(Equation 2) J=(3cos2θ−1)|μ→|24πε0R3Ek=2Jcos(k).(Equation 3) The radiative lifetime of materials hosting Frenkel excitons can be shortened by engineering the molecular interactions, which can give rise to the superradiance phenomenon in molecular aggregates.25Kasha M. Rawls H.R. El-Bayoumi M.A. The exciton model in molecular spectroscopy.Pure Appl. Chem. 1965; 11: 371-392Crossref Scopus (3455) Google Scholar,26Spano F.C. Mukamel S. Superradiance in molecular aggregates.J. Chem. Phys. 1989; 91: 683-700Crossref Scopus (250) Google Scholar In this phenomenon, fluorescence from a certain exciton state can lead to radiative lifetime inversely proportional to the size of the aggregate. Molecular aggregates can be categorized as J and H aggregate depending on the sign of intermolecular Coulomb interactions. As a simple example, we describe the photophysics of molecular aggregates and the origin of superradiance with a linear chain of N molecules where the molecular transition dipole moments are parallel. With a far field dipole approximation, the effect of intermolecular interaction can be treated analytically and can lead to a simple relationship between molecular orientations and material radiative lifetimes. The ground state of the molecular chain is not affected by the intermolecular interactions and can be expressed as the multiplication of the molecular ground state wavefunctions. The excited states of the chain are linear combinations of the degenerate single-molecule excited states (Equation 2) in which the electrons and holes are bound within the same molecule.18Hestand N.J. Spano F.C. Expanded theory of H- and J-molecular aggregates: the effects of vibronic coupling and intermolecular charge transfer.Chem. Rev. 2018; 118: 7069-7163Crossref PubMed Scopus (590) Google Scholar Here, |n> denotes the single-molecule excited states wavefunction where the nth molecule is in the excited state while all other molecules are in the ground state, and |k> denotes the wavefunctions of the Frenkel excitons that form an exciton band. The radiative lifetime of the exciton can be related to the transition dipole moment after substituting the molecular electronic states with the exciton states. If we only consider the interaction between adjacent molecules, the energy of exciton states depends on the relative orientation between the molecular transition dipole and displacement vector connecting the molecular mass centers (Equations 2 and 3; Figure 2).18Hestand N.J. Spano F.C. Expanded theory of H- and J-molecular aggregates: the effects of vibronic coupling and intermolecular charge transfer.Chem. Rev. 2018; 118: 7069-7163Crossref PubMed Scopus (590) Google Scholar With negative Coulomb interaction energy, the all-symmetrical (k = 0) state has lowest energy in the exciton band, which is the state from which the radiative decay will occur. The transition dipole moment of this state is √N times that of the molecular state, which gives the molecular aggregate N times shorter radiative lifetime as well as brighter fluorescence. In the classical picture in which the fluorescent molecules are treated as oscillating dipoles (the transition dipole), these dipoles in the molecular chain are aligned in the same direction into a giant oscillating dipole that has N times the oscillator strength of an isolated molecule. In actual molecular aggregates, due to the presence of disorder and the effect of finite temperature, the number of molecules that “coherently” interact with each other is often smaller than the size of the aggregates. In the absence of disorder, the coherence length is proportional to the square root of the Coulomb energy J divided by thermal energy kT (where k is Boltzmann’s constant and T is temperature).27Yamagata H. Spano F.C. Strong photophysical similarities between conjugated polymers and J-aggregates.J. Phys. Chem. Lett. 2014; 5: 622-632Crossref PubMed Scopus (61) Google Scholar As the all-symmetrical state has lower energy than the molecular excited states, the optical spectra of the aggregate are red-shifted with respect to the isolated molecules. This type of molecular aggregate was discovered by Jelley and named as J aggregate.28Jelley E.E. Spectral absorption and fluorescence of dyes in the molecular state.Nature. 1936; 138: 1009-1010Crossref Scopus (325) Google Scholar,29Bricks J.L. Slominskii Y.L. Panas I.D. Demchenko A.P. Fluorescent J-aggregates of cyanine dyes: basic research and applications review.Methods Appl. Fluoresc. 2017; 6: 012001Crossref PubMed Scopus (168) Google Scholar Conversely, with positive Coulomb interaction energy, the all-symmetrical state is the highest energy state in the exciton band. The electron excited into this state would decay through internal conversion to the lowest energy state in the exciton band, which has zero transition dipole moment and thus a long radiative lifetime. In this case the aggregate is named H aggregate, as the absorption spectra have hypsochromic shift compared with the molecular monomer. With the existence of faster non-radiative decays, the H aggregate typically has weak fluorescence. Thus, the photophysics of molecular materials depends highly on the geometry of molecular assembly, which dictates the molecular interactions that determine material properties important to high-speed light-emitting devices. The lifetimes of a various fluorescent materials are summarized in Figure 3. The UV-emitting GaN has one of the shortest radiative lifetimes in materials hosting free carriers approaching 1 ns,6Im J.S. Moritz A. Steuber F. Härle V. Scholz F. Hangleiter A. Radiative carrier lifetime, momentum matrix element, and hole effective mass in GaN.Appl. Phys. Lett. 1997; 70: 631-633Crossref Scopus (212) Google Scholar,10Muth J.F. Lee J.H. Shmagin I.K. Kolbas R.M. Casey Jr., H.C. Keller B.P. Mishra U.K. DenBaars S.P. Absorption coefficient, energy gap, exciton binding energy, and recombination lifetime of GaN obtained from transmission measurements.Appl. Phys. Lett. 1997; 71: 2572-2574Crossref Scopus (627) Google Scholar with the most widely used GaAs and InGaAsP having lifetime beyond 10 ns.5Henry C. Levine B. Logan R. Bethea C. Minority carrier lifetime and luminescence efficiency of 1.3 μm InGaAsP-InP double heterostructure layers.IEEE J. Quan. Electron. 1983; 19: 905-912Crossref Scopus (51) Google Scholar,8Lush G.B. MacMillan H.F. Keyes B.M. Levi D.H. Melloch M.R. Ahrenkiel R.K. Lundstrom M.S. A study of minority carrier lifetime versus doping concentration in n-type GaAs grown by metalorganic chemical vapor deposition.J. Appl. Phys. 1992; 72: 1436-1442Crossref Scopus (60) Google Scholar For lead iodide perovskite, although the direct band-gap material has radiative recombination coefficient similar to that of GaAs, its lower carrier concentration leads to the long radiative lifetime up to tens of microseconds.30Alarousu E. El-Zohry A.M. Yin J. Zhumekenov A.A. Yang C. Alhabshi E. Gereige I. AlSaggaf A. Malko A.V. Bakr O.M. et al.Ultralong radiative states in hybrid perovskite crystals: compositions for submillimeter diffusion lengths.J. Phys. Chem. Lett. 2017; 8: 4386-4390Crossref PubMed Scopus (66) Google Scholar,31Yamada Y. Nakamura T. Endo M. Wakamiya A. Kanemitsu Y. Photocarrier recombination dynamics in perovskite CH3NH3PbI3 for solar cell applications.J. Am. Chem. Soc. 2014; 136: 11610-11613Crossref PubMed Scopus (617) Google Scholar The range of radiative lifetime of these free carrier materials corresponds to materials with different doping levels. Low-dimensional semiconductors hosting Wannier-Mott excitons such as quantum dots and transition metal dichalcogenides have radiative lifetime on the order of 10 ns.15Gong K. Martin J.E. Shea-Rohwer L.E. Lu P. Kelley D.F. Radiative lifetimes of zincblende CdSe/CdS quantum dots.J. Phys. Chem. C. 2015; 119: 2231-2238Crossref Scopus (49) Google Scholar,32Amani M. Lien D.H. Kiriya D. Xiao J. Azcatl A. Noh J. Madhvapathy S.R. Addou R. Santosh K.C. Dubey M. et al.Near-unity photoluminescence quantum yield in MoS2.Science. 2015; 350: 1065Crossref PubMed Scopus (806) Google Scholar, 33Du H. Chen C. Krishnan R. Krauss T.D. Harbold J.M. Wise F.W. Thomas M.G. Silcox J. Optical properties of colloidal PbSe nanocrystals.Nano Lett. 2002; 2: 1321-1324Crossref Scopus (456) Google Scholar, 34Liu H. Guyot-Sionnest P. Photoluminescence lifetime of lead selenide colloidal quantum dots.J. Phys. Chem. C. 2010; 114: 14860-14863Crossref Scopus (67) Google Scholar, 35Kim H. Ahn G.H. Cho J. Amani M. Mastandrea J.P. Groschner C.K. Lien D.-H. Zhao Y. Ager J.W. Scott M.C. et al.Synthetic WSe2 monolayers with high photoluminescence quantum yield.Sci. Adv. 2019; 5: eaau4728Crossref PubMed Scopus (53) Google Scholar As previously discussed, phosphorescent materials have long radiative lifetimes on the order of microseconds.21Sommer J.R. Shelton A.H. Parthasarathy A. Ghiviriga I. Reynolds J.R. Schanze K.S. Photophysical properties of near-infrared phosphorescent π-extended platinum porphyrins.Chem. Mater. 2011; 23: 5296-5304Crossref Scopus (114) Google Scholar,22Sarma M. Tsai W.-L. Lee W.-K. Chi Y. Wu C.-C. Liu S.-H. Chou P.-T. Wong K.-T. Anomalously long-lasting blue PhOLED featuring phenyl-pyrimidine cyclometalated iridium emitter.Chem. 2017; 3: 461-476Abstract Full Text Full Text PDF Scopus (68) Google Scholar Among aromatic molecules, smaller molecules have larger optical band gap and weaker π stacking, and thus can host Frenkel excitons with radiative lifetime of a few nanoseconds.36Horiguchi R. Iwasaki N. Maruyama Y. Time-resolved and temperature-dependent fluorescence spectra of anthracene and pyrene in crystalline and liquid states.J. Phys. Chem. 1987; 91: 5135-5139Crossref Scopus (48) Google Scholar,37Selvakumar S. Sivaji K. Arulchakkaravarthi A. Sankar S. Enhanced fluorescence and time resolved fluorescence properties of p-terphenyl crystal grown by selective self seeded vertical Bridgman technique.Mater. Lett. 2007; 61: 4718-4721Crossref Scopus (23) Google Scholar With increasing conjugation size and without steric hindrance from side chains or functional groups, the tendency of π stacking between molecules increases, leading to a substantial portion of charge-transfer excitons and longer radiative lifetime of around 100 ns.24Katoh R. Suzuki K. Furube A. Kotani M. Tokumaru K. Fluorescence quantum yield of aromatic hydrocarbon crystals.J. Phys. Chem. C. 2009; 113: 2961-2965Crossref Scopus (192) Google Scholar In general, extensive π stacking often leads to side-by-side packing geometry behaving as H aggregates, which is also undesirable for shortening the radiative lifetime. The pseudoisocyanine (PIC) molecular aggregates, where the J-aggregate behavior was first discovered, have radiative lifetime of 1.5 ns.38Ermolaeva G.M. Maslov V.G. Orlova A.O. Panfutova A.S. Rosanov N.N. Fainberg B.D. Shakhverdov T.A. Shilov V.B. Dynamics of optical response of solutions of pseudoisocyanine J aggregates upon pico- and subnanosecond excitation.Opt. Spectrosc. 2011; 110: 871-879Crossref Scopus (9) Google Scholar J-aggregate behavior can also be achieved in crystals of aromatic molecules, where some small molecules that host Frenkel excitons can directly crystallize into J aggregates (e.g., 2,6-diphenylanthracene).39Li J. Zhou K. Liu J. Zhen Y. Liu L. Zhang J. Dong H. Zhang X. Jiang L. Hu W. Aromatic extension at 2,6-positions of anthracene toward an elegant strategy for organic semiconductors with efficient charge transport and strong solid state emission.J. Am. Chem. Soc. 2017; 139: 17261-17264Crossref PubMed Scopus (110) Google Scholar For large conjugated molecules, bulky side chains can be used to reduce the charge-transfer exciton and facilitate the desirable molecular packing geometry.40Chan J.M.W. Tischler J.R. Kooi S.E. Bulović V. Swager T.M. Synthesis of J-aggregating dibenz[a,j]anthracene-based macrocycles.J. Am. Chem. Soc. 2009; 131: 5659-5666Crossref PubMed Scopus (68) Google Scholar,41Zhang B. Soleimaninejad H. Jones D.J. White J.M. Ghiggino K.P. Smith T.A. Wong W.W.H. Highly fluorescent molecularly insulated perylene diimides: effect of concentration on photophysical properties.Chem. Mater. 2017; 29: 8395-8403Crossref Scopus (81) Google S

Full Text
Published version (Free)

Talk to us

Join us for a 30 min session where you can share your feedback and ask us any queries you have

Schedule a call