Photon Upconversion in Crystalline Rubrene: Resonant Enhancement

23, 25. In addition, these processes govern the unusual excitation intensity dependence of ..... Polarizing the excitation light along the b axis (0°...
2 downloads 0 Views 745KB Size
Subscriber access provided by University of Winnipeg Library

C: Energy Conversion and Storage; Energy and Charge Transport

Photon Upconversion in Crystalline Rubrene: Resonant Enhancement by an Interband State Chad D. Cruz, Hyun Ho Choi, Vitaly Podzorov, Eric L. Chronister, and Christopher J. Bardeen J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.8b02838 • Publication Date (Web): 13 Jul 2018 Downloaded from http://pubs.acs.org on July 16, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Photon Upconversion in Crystalline Rubrene: Resonant Enhancement by an Interband State Chad D. Cruz,1 Hyun Ho Choi,2 Vitaly Podzorov,2 Eric L. Chronister1, Christopher J. Bardeen*1

1

Department of Chemistry, University of California Riverside, Riverside, California 92521, USA

2

Department of Physics and Astronomy, Rutgers University, Piscataway, New Jersey 08854, USA

Abstract Triplet-triplet exciton annihilation after sensitization of the triplet states by a near-infrared (NIR) absorbing sensitizer enables rubrene to function as a photon upconversion material. In this paper, we demonstrate an alternate pathway to NIR upconversion in pristine rubrene crystals: resonantly enhanced two-photon absorption via a weakly allowed interband state. We find that all crystalline rubrene samples exhibit NIR-to-visible upconversion that can be easily observed by eye under low-intensity (20 W/cm2) continuous wave excitation. The amount of continuous wave photoluminescence is comparable to what is observed under femtosecond pulsed excitation with the same average intensity. A wide range of excitation intensities (I) for the photoluminescence power dependence are explored and careful fitting of the intensity dependence of the upconverted photoluminescence shows that it has an approximate I4I2 transition. Moreover, there is a pronounced dependence of the per-pulse upconverted photoluminescence signal on the laser repetition rate. A four-state kinetic model with a longlived (~1 µs) interband state that takes into account fission and fusion dynamics can reproduce both the I4I2 transition and the dependence of the photoluminescence on pulse repetition rate. The modeling suggests that this interband state arises from a low-concentration species, possibly a crystal defect or defective rubrene molecules. Several other polyacene crystals (tetracene, diphenylhexatriene and perylene) measured under the same conditions did not exhibit similar behavior. The observation of resonantly enhanced upconverted PL without the addition of chemically distinct sensitizers suggests that interband states in organic molecular crystals can generate new and possibly useful photophysical behavior.

1 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 30

Introduction Photophysical phenomena in organic molecular materials can potentially form the basis of high-efficiency solar energy conversion technologies. For instance, many conjugated organic molecules, including derivatives of tetracene (TET),1-3 pentacene,4-5 perylene6-7 and isobenzofuran8-9 can efficiently convert a high energy singlet exciton into two lower energy triplet excitons via the spin conserving process of singlet fission. The inverse of singlet fission is triplet-triplet annihilation, in which a pair of low energy triplet excitons fuse into one higher energy singlet exciton that can emit a high energy photon.10-11 Both phenomena can be used to downconvert (fission) or upconvert (fusion) light energy. These two physical processes thus provide strategies to eventually surpass the Shockley-Queisser limit for single junction photovoltaic devices.12-15 In a crystal composed of rubrene (RUB), the triplet pair state is approximately isoenergetic with the singlet state and both singlet fission and triplet fusion can be quite efficient.16-24 These two processes are reflected in time-resolved photoluminescence (PL) experiments where both prompt and delayed components are observed.23,

25

In addition, these

processes govern the unusual excitation intensity dependence of photoconductivity in RUB with fractional power exponents.26 For organic triplet-triplet annihilation upconversion (TTA-UC), RUB is a commonly used emitter molecule due its long-lived triplet states and high PL quantum yield.27-34 The low energy of RUB’s triplet state (1.15 eV) allows it to be sensitized by NIR absorbing chromophores such as porphyrins35-36 and inorganic nanocrystals.28 After sensitization, triplet excited states on separate RUB molecules can fuse into a higher energy singlet state which then emits a visible photon. Most reported upconversion schemes are performed in solutions, 2 ACS Paragon Plus Environment

Page 3 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

since facile diffusion permits the triplet sensitization and subsequent fusion events to occur in different spatial regions. This prevents energy transfer from the emitter back to the sensitizer. UC in solid-state systems is more challenging due to decreased diffusion and typically has a lower overall efficiency. Nevertheless, progress has been made on visible-to-UV solid-state systems.37-39 Furthermore, two recent results suggest that RUB could form the basis of a solidstate material that upconverts NIR photons. First, the Baldo and Bulovic groups demonstrated that a layer of inorganic nanocrystals could sensitize triplets in a RUB film, enabling upconversion.40 Even more surprising, an earlier paper by Liu et al. reported visible PL excited by a continuous wave (CW) 980 nm laser in crystalline powders of RUB in the absence of any sensitization.41 The authors of that work attributed the “upconversion” to nonresonant twophoton absorption (2PA). Although they did not attempt to extract a cross section, their data suggested it could be very high. Interestingly, there is significant disagreement over the magnitude of the two-photon cross section in crystalline RUB. Gurzadyan and coworkers performed an open aperture Z-scan to determine the cross section at 740 nm in crystalline RUB arriving at a value of 1000 GM (1 GM = 10-50 cm4/s).42 Biaggio and coworkers measured the nonlinear transmissivity of RUB between 750 and 920 nm and find a much lower two-photon cross section of 170 GM peaked at 850 nm,43 that is in close agreement to the theoretical study by Zhao et al. which determined a two-photon cross section of 185 GM.44 Gurzadyan and coworkers use an 80 MHz laser to perform their Z-scan, while Biaggio and coworkers use a 1 kHz repetition rate. This difference in repetition rates was cited as a possible reason for the different magnitudes of the two-photon cross section. Given the important role of RUB in solid-state NIR upconversion schemes, we decided to revisit the origin of the upconversion in pristine crystals. In agreement with Liu et al.,41 we find

3 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

that crystalline RUB exhibits surprisingly intense NIR-visible upconversion using CW excitation at intensities as low as 20 W/cm2. The PL generated by CW excitation is comparable to that generated by a femtosecond pulse train with the same average intensity but with peak intensities on the order of 105 times higher. The fact that the upconverted PL depends only on average intensity, not peak intensity, suggests that an intermediate electronic state is facilitating resonant two-photon absorption in pristine crystals of RUB. A large range of intensities for the PL power dependence are explored and a quartic regime is found at lower intensities. Moreover, there is a pronounced dependence of the per-pulse upconverted PL on the laser repetition rate. Our observations lead to a model in which a long-lived (~1 µs) intermediate state facilitates twophoton absorption to the singlet state, after which it undergoes fission and fusion dynamics. Other polyacenes, TET, diphenylhexatriene and perylene do not exhibit similar behaviors, suggesting that the crystal packing in RUB plays a pivotal role in the observed photophysics. The observation of resonantly enhanced upconverted PL in a pristine molecular crystal suggests that intermolecular interactions in molecular crystals can give rise to interband states that enable photon upconversion even in the absence of triplet sensitizers.

Experimental Pristine crystals of RUB were grown via physical vapor transport as has been reported previously.45 The crystals were a mix of needle-like and platelet-like with the largest facet being the ab face. All measurements were taken with excitation incident upon the ab face. The crystals ranged in thickness with the platelet-like crystals being thinner on average (2 -100 µm) and the needle-like crystals being thicker (200 – 300 µm). The crystals were stored in the dark until ready for use. To avoid photooxidation of the crystals, measurements were performed in a Janis

4 ACS Paragon Plus Environment

Page 4 of 30

Page 5 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Research vacuum cryostat (10-5 torr) fitted with optical windows. Two different bottles of microcrystalline RUB powders were used (Sigma Aldrich, sublimed grade 99.99%; Acros Organics, 99%) when measuring the upconverted PL, and both gave the same results. Absorption measurements for RUB in CH2Cl2 solution were taken with a Varian Cary 500 UV-Vis/Near IR spectrophotometer. Absorption measurements for RUB single crystals were taken using a Craic Technologies 20/30 PV Microspectrophotometer. The dependence on average intensity of the PL signal was measured for both CW and pulsed NIR excitation. An 808 nm diode laser with an 810 nm interference filter (Thorlabs FBH 810-10) was used to perform the PL intensity dependence using front face detection (Figure S1). The 808 nm laser focused to a roughly square spot with a side length of 50 µm. A 980 nm diode laser with a 950 nm long wave pass filter was used to perform the PL intensity dependence using an angled configuration for detection (Figure S2), with a square focal spot with a side length of 300 µm. A Coherent Vitesse Ti:Sapphire oscillator was also used to measure the intensity dependence of the PL at 800 nm using front face detection (Figure S3). The Vitesse can be operated in either CW or pulsed modes (80 MHz, 100 fs) and the beam has a Gaussian profile with a measured waist of 30 µm (FWHM). For all excitation wavelengths, the excitation beam was modulated at 100 Hz and the PL was detected with a Hamamatsu mini photomultiplier tube connected to a lock-in amplifier. Two 750 nm short pass filters were placed before the detector to minimize the contribution of laser scatter to the signal. The laser power was varied with a set of neutral density filters and measured with a power meter. Multiple crystals were measured, and the intensity dependence is the same regardless of crystal thickness. The steady state PL spectra were measured with the Ocean Optics USB 4000 spectrophotometer using the angled setup shown in Figure S2 with excitation wavelengths of 532

5 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

nm, 808 nm and 980 nm. The same filters mentioned previously were used for the 808 nm and 980 nm measurements. An OG 550 nm long wave pass filter was used in front of the detector for the PL excited with the 532 nm laser. For all excitation wavelengths the incident polarization was aligned parallel to the b axis of the crystal. The IR emission was collected using an InGaAs detector which is part of the PTI QM-400 fluorimeter (Figure S4). The 808 nm laser with polarization aligned parallel to the b axis was used to excite RUB and an 850 nm long wave pass filter was placed over the detection window. The Coherent Vitesse oscillator described above was also used for the two-photon cross section estimation and the variable repetition rate experiments. In pulsed mode, the relative 2PA cross sections of RUB and TET were measured in degassed solutions of methylene chloride. A degassed solution of Coumarin 540A (C540A) in methylene chloride was used as the 2PA standard (δ = 47 x 10-50 cm4/s).46-47 The emission was probed with a Hamamatsu mini photomultiplier tube in a right-angle configuration (Figure S5). An interference filter was used to isolate a 20 nm spectral detection window (510 – 530 nm for TET, 540 – 560 nm for RUB and C540A). For the variable repetition rate experiments, the laser was operated in pulsed mode and a ConOptics 350 electro-optic modulator controlled by a ConOptics 25D driver was coupled to a ConOptics 300 synchronous countdown modulator. This pulse-picking set-up allowed the repetition rate to be varied over the range of 80 – 0.8 MHz. The emission was collected through two 750 nm short wave pass filters in a front face configuration using a Hamamatsu mini photomultiplier tube connected to a lock-in amplifier (Figure S6). For all the results reported in this paper, when we refer to the laser intensity it will correspond to the time-averaged intensity of the laser used. For the CW source, this also corresponds to the peak intensity. For the femtosecond laser, with a nominal pulse duration of

6 ACS Paragon Plus Environment

Page 6 of 30

Page 7 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

100 fs and a repetition rate of 80 MHz, this number must be multiplied by a factor of approximately 105 to estimate the peak intensity during the pulse.

Results and Discussion We first studied the 2PA properties of TET and RUB molecules in dilute solutions of methylene chloride. Surprisingly, we found only one prior report in the literature on the twophoton properties of RUB48 and no previous reports for TET. The laser dye C540A was used as a standard for the 2PA at 800 nm, with a measured 2PA cross section δ = 47 x 10-50 cm4/s.46-47 All three molecules showed measurable fluorescence only when excited using a femtosecond pulsed laser. When the same laser was used in CW mode, no fluorescence signal could be detected above the background. All three molecules also showed the quadratic dependence on intensity expected for nonresonant 2PA, and this dependence is shown in Figure 1. There is a slight saturation behavior for all three molecules at the highest intensities, as observed for other 2PA absorbers.49 Using the equation50-51

∙

 =   

∙ ∙ 



∙





(1)

where  represents the fraction of the total PL spectrum within the range of the interference filter, φ is the PL quantum yield and S is the measured signal, we can determine the relative values of δ for RUB and TET by comparing their signals to that measured for C540A under the same conditions. The experimental parameters used are summarized in Table 1, along with the calculated δ values. At 800 nm excitation, RUB and TET have very similar 2PA cross-sections, which is not surprising considering that they have similar conjugated cores. Both δ values are about 30× smaller than that of C540A as well as the previously reported value for RUB in a benzene solution (56 x 10-50 cm4/s).48 However, the open aperture Z-scan technique utilized in 7 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

reference 48 can yield an effective 2PA cross section that is enhanced by excited state absorption52 and their reported intensity dependence with a slope greater than 2 suggests that multiphoton processes other than 2PA are contributing to the cross section. Furthermore, our lower δ values for RUB and TET are consistent with experimentally determined values for other polycyclic aromatic hydrocarbons such as naphthalene, anthracene and perylene.47, 53-54 The crystal packing in RUB leads to a large separation of the conjugated cores due to the twisted phenyl rings, and Biaggio has suggested that the spectroscopic properties of this molecular crystal can be approximated by an oriented gas of RUB molecules, with only minor perturbations by Coulombic interactions.55 TET, on the other hand, forms a close-packed herringbone crystal where strong Coulomb and charge-transfer interactions lead to a large Davydov splitting56-57 and other excitonic effects such as superradiance.58 We expected crystalline RUB to behave similarly to the solution phase, while crystalline TET might show some significant deviations. Instead, it was crystalline RUB that exhibited qualitatively new behavior. The first surprising observation is illustrated in Figure 2. RUB microcrystalline powder displays bright yellow-orange emission (~600 nm) which is easily discerned by eye under both pulsed (Figure 2A) and CW excitation (Figure 2B) at 800 nm. The PL spectrum is similar for all three wavelengths (Figure 2C), but the high energy side of the emission is suppressed for longer wavelength excitation, as expected because these wavelengths penetrate deeper into the crystal and the PL undergoes more self-absorption as it escapes. All other molecular crystals tested (TET, perylene, diphenylhexatriene) show a clear dependence on the peak power of the laser, with at least an order of magnitude more upconverted PL under pulsed excitation compared to CW excitation, whereas upconverted PL in RUB is approximately the same under pulsed and CW excitation.

8 ACS Paragon Plus Environment

Page 8 of 30

Page 9 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

The dependence of the upconverted PL on the incident laser intensity highlights another difference between crystalline RUB and TET (Figure 3). The power law with exponent 2 observed in TET is a hallmark of a two-photon process, which is proportional to the incident laser intensity squared. Perylene and diphenylhexatriene crystals also showed the usual I2 dependence over the whole range of intensities for both pulsed and CW excitation (Figure S7). For these crystals, the dependence on peak power and the power law of 2 both suggest nonresonant 2PA as the dominant mechanism leading to upconverted PL in TET and the other crystals. However, direct absorption to the triplet state followed by TTA, can also generate an I2 dependence. This pathway has been previously observed in anthracene59 as well as TET.60 Even though the S0T1 transition is formally forbidden, in many organic crystals its weak absorption is sufficient to generate enough triplet density to observe TTA if intense NIR sources are used. The upconverted PL observed in RUB, on the other hand, has a different intensity dependence, with a log-log plot exhibiting a slope of 4 at lower intensities that turns to a slope of 2 and eventually begins to flatten out at the highest intensities. Similar intensity dependences are observed in RUB for excitation wavelengths of 800 nm, 808 nm and 980 nm (Figure S8). The results are quite robust amongst different samples of RUB. Pristine crystals, crystalline powder out of the supplier’s bottle and RUB crystals which are more than 4 years old, all show the same behavior (Figure S9). These results raise two questions: (1) why does CW excitation lead to efficient 2PA in RUB? (2) Why does the upconverted PL have a quartic dependence upon the incident intensity? To answer question (1), we consider the 2PA process in more detail. The mechanism of nonresonant 2PA can be described in terms of a three-state system in which the final state is populated via a virtual intermediate state between the initial and final real states.52,

9 ACS Paragon Plus Environment

61-62

The

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

virtual state cannot support a real population, and the two photons must arrive within a very short time interval ∆ = 10−15 − 10−16 $% to reach the excited singlet state. If the intermediate state is a real state that can support population, this lifts the requirement of the simultaneous arrival of two photons.62 To access the excited singlet through the resonant intermediate, the second photon must only arrive before the decay of this state. A resonantly enhanced intermediate state would explain the fact that the upconverted PL observed in RUB depends only on average intensity, rather than the peak intensity within the femtosecond pulse train. If the 2PA in crystalline RUB proceeds through a real intermediate state, then this state should have some spectroscopic signature. The most logical candidate for a transition in the NIR would be the S0T1 transition in molecular RUB. A concentrated (~ 2 mM) solution of RUB in methylene chloride was prepared and placed in a 10 cm cuvette for the NIR absorption measurement, however, no measurable absorption is observed beyond 650 nm (Figure 4A). Given that this measurement is sensitive to an absorbance of 0.005 or less, we can place an upper limit of ε ≤ 0.005 M-1cm-1 for any NIR transition in the isolated molecule. Of course, the CW 2PA is only seen in crystalline samples, so the lack of any absorptive features in the solution spectrum is not surprising. We attempted to directly measure the absorption of a single RUB crystal out to 1700 nm using a microspectrometer with unpolarized light (Figure 4B). The oscillatory feature present over the NIR range is an etalon effect which allows us to estimate a crystal thickness of 5 µm.63 There was no detectable NIR absorption above the level of 0.04 for a 5 µm thick crystal. This suggests that the crystal has an absorption coefficient of less than 40 cm-1. To compare with solution studies, we also calculate an effective concentration of 2.4 M in the crystal, which places an upper limit on the molar absorption coefficient ε ≤ 35 M-1cm-1. It should be noted that the absorption in a crystal is highly anisotropic, and it is possible that

10 ACS Paragon Plus Environment

Page 10 of 30

Page 11 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

correctly polarized light would be able to pick up a low amplitude absorption in this wavelength region. The light from the microspectrometer used in this study was unpolarized, and the absorption coefficients here should be regarded as a rough estimate. A full polarization study is necessary to accurately assess the absorption properties of the crystal in the NIR. Even though we could not detect the presence of a low-energy intermediate state in the absorption, we did find evidence for such states by looking at the NIR emission after CW 808 nm excitation. In Figure 5 we show the steady-state PL spectrum for this excitation wavelength. The emission extends from 950 nm past 1200 nm, with two prominent peaks at 980 nm and 1060 nm. The peak at 1060 nm is likely a Raman peak with a separation of 3000 cm-1 from the excitation wavelength. The 980 nm peak of this PL is at a higher energy than that expected for RUB phosphorescence, which should be around 1070 nm based on a triplet energy of 1.15 eV. We were unable to find any experimental phosphorescence spectra for RUB in the literature. It is also not clear that this emitting state is solely responsible for the enhanced 2PA. Given that its emission peaks at 980 nm, it is unlikely to absorb strongly at this wavelength, so it may not be able to facilitate the experimentally observed 2PA at 980 nm. Nevertheless, the presence of PL in this wavelength range proves the existence of low-energy states in RUB with appreciable oscillator strength. Since the upconverted PL is the most easily detectable signature of the intermediate state, we have used it to investigate the nature of this state in more detail. The upconverted PL in crystalline RUB depends upon the polarization of the incident photons (Figure 6). The vapor )* transport grown RUB crystals used for this study are orthorhombic in the &'( point group and

the crystallographic axes are defined in the space group Acam (a = 14.4 Å, b = 7.18 Å, c = 26.9 Å), consistent with previous studies.55, 64 We observe a distinct anisotropy in the PL intensity

11 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

induced on the ab surface with the 980 nm diode laser when the excitation is polarized parallel to different crystallographic axes. Polarizing the excitation light along the b axis (0°) yields the brightest PL while polarizing along the a axis (90°) minimizes the PL. However, a similar anisotropy for single crystals was reported by Biaggio and coworkers using a pulsed 772 nm laser, with a factor of 4 enhancement in the upconverted PL when the polarization is aligned parallel to the b axis.43 The anisotropy depicted in Figure 6 is more pronounced than what has been previously reported, with 3 orders of magnitude more intense upconverted PL from exciting with polarization parallel with the b axis relative to the a axis. The highly oriented nature of the two-photon transition at 980 nm suggests that it does not originate from a randomly oriented defect or impurity. Additional evidence for the role of a long-live intermediate state is obtained by measuring the upconverted PL intensity as a function of the laser repetition rate. In the limit of negligible ground state depletion, the PL signal of a material which solely undergoes nonresonant 2PA should show no dependence on the repetition rate of the laser. Each pulse transfers an equal amount of population to the excited state, independent of how quickly the pulses follow each other. If a resonant intermediate state participates in the 2PA process, however, a dependence upon the pulse spacing of the laser will be observed when the lifetime of the intermediate state is comparable to or longer than the pulse spacing.65-66 The initial pulse populates the excited singlet state through a resonant 2PA processes, but some population will be left in the intermediate state. Between pulses, the intermediate state population will partially decay back to the ground state, but if a pulse arrives before it fully decays, it can transfer population from both S0 and the intermediate state, increasing the total PL signal. For a resonant 2PA mechanism, the PL signal per pulse should be greater for shorter pulse spacings and

12 ACS Paragon Plus Environment

Page 12 of 30

Page 13 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

gradually decrease as the pulse spacing increases. Indeed, this is observed for single crystals of RUB (Figure 7). For comparison, we also measured the repetition rate dependent PL of TET powder and the laser dye Rhodamine B (Rhod B) in a PMMA matrix. Both of these samples show no dependence on the laser repetition rate to within the measurement error, confirming that nonresonant 2PA is the dominant mechanism leading to the PL. In order to quantitatively model the dependence of the PL on laser intensity and repetition rate, we use the 4-state model outlined in Figure 8. This model leads to the following system of kinetic equations: +,- +.

+,> +.

= −/ 0 12 + 40 20 + 4' − 4 055 %26 + 4.708 296 + (4 99; + 4 99 )296 '

= / 0 12 − (/0' 1 + 40 )20













(3)











(4)

= /0' 120 − 4' + 4 055 %26 + 499 296 '

+,A6

= −4.708 296 + 24 055 26 − (2499 + 499; )296 '

+.

(2)



+,-6 +.



(5)

In Equations (2)-(5), I is the laser intensity (photons⋅s-1⋅cm-2), and the absorption cross sections for transitions from S0 → i and i → S1 are given by σ0i and σi2, respectively. The rate ki describes nonradiative relaxation from i → S0, while k2 represents the radiative relaxation rate from S1 → S0. The rate of singlet fission is given by kfiss, triplet fusion by kTTS and nonradiative relaxation of the triplet by ktrip. A triplet quenching term is included, kTTQ describing an interaction between triplets which returns one of them to the ground state, T1 + T1 → T1 + S0. We know k2 = 6.25x107 s-1 from time-resolved PL measurements,28 and we estimate ktrip to be 1.67x104 s-1 based on the decay of the delayed fluorescence signal (Figure S10) as well as previous measurements by Biaggio and coworkers.25 We obtain kfiss = 5x1010 s-1 from previous work which used pump-probe spectroscopy where the correlation between the decay of singlet induced

13 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

absorption and the growth of the triplet induced absorption is assigned to singlet fission.21-22, 24, 42 Exponential fits to the time traces of the singlet and triplet features yield two fission times, we have taken the longer time, which is reported to be approximately 20 ps, to represent kfiss here. The triplet recombination rate, kTTS = 10-12 cm3/s, is determined from previous work where a lower limit (10-13 cm3/s) was estimated from a fit to the PL decay25 and a more precise value (5.4±1.0 x 10-12 cm3/s) was found from a transient grating measurement.67 We assume that kTTQ < kTTS. This leaves σ0i, σi2, and ki as the only completely unknown parameters. We can use this kinetic model to describe both the intensity dependence of the PL shown in Figure 3 and the repetition rate dependence shown in Figure 7. Both measurements depend nonlinearly on the model parameters, so our goal was to find a set of reasonable parameters that could consistently describe both sets of data. We cannot rule out that there may exist a different set of parameters that reproduce these data equally well. From numerical simulations, we find that the detailed shape of the PL signal versus pulse period curve is sensitive to the σ0i and σi2 values as well as ki. In the low S0 depletion limit, when less than 1% of the population is excited by each pulse, the decrease of the per-pulse PL signal versus pulse period is not a simple exponential decay. By assuming the ratio σ0i/σi2 = 0.01, we can numerically solve Equations (2)(5) for a train of δ-function pulses. The calculated signals are overlaid with the data in Figure 9A. A value for the intermediate lifetime τi = 1/ki = 1.5 µs yields a decent match to the data. Representative curves for other values of τi deviated from the data points at long times. This lifetime is long compared to the S1 state but much shorter than that of the T1 state as detected via the delayed fluorescence decay (60 µs). We now turn to the physical origin of the intensity dependence of the PL signal. At high intensities, the slope of 2 which turns to a slope of 1 can be understood as PL induced by 2PA

14 ACS Paragon Plus Environment

Page 14 of 30

Page 15 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

which begins to saturate at high intensities and has been observed before in other materials.68 At lower intensities, there is an asymptotic quartic (power exponent 4) dependence on laser intensity that is more surprising. This type of intensity dependence has been observed previously in anthracene, however.11,69 It arises because the PL due to TTA is proportional to the square of the triplet population, which in turn is proportional to the initial singlet population. Since the initial singlet population generated via 2PA is also proportional to the square of the incident intensity, the delayed PL is expected to have a quartic dependence on the laser intensity, as observed. It is important to emphasize that this quartic intensity dependence is a general phenomenon that should be observed any time a triplet population that undergoes TTA is created by 2PA. To analyze the intensity dependence quantitatively, we first fix ki = 0.67 µs-1 to be consistent with the repetition rate data. Assuming a CW laser excitation at 800 nm with constant steady-state densities of all the involved species, we can set dNx/dt=0 and solve Equations (2)-(5) for the equilibrium populations for a fixed laser intensity I. We first assume that NS0(0) = 1.434x1021 molecules/cm3, the density of molecules in a pure RUB crystal. In this case, we obtain the best match to the data when σ0i = 2x10-20 cm2 and σi2 = 2x10-16 cm2. But we found that these parameters could not reproduce the saturation behavior at the highest intensities (Figure 9). The calculated signal did not turn over as quickly as the experimental signal at the highest intensities. In view of the low absorbance of the samples in this wavelength range, increasing the σ0i value did not seem reasonable. Instead, we lowered the NS0(0) value. We found an improved fit to the data when NS0(0) = 1.44x1019 molecules/cm3 with σ0i = 2x10-18 cm2 and σi2 = 2x10-16 cm2. The lower NS0(0) value corresponds to a 1% concentration of sites with an enhanced 0i absorption. The linear absorbance, proportional to NS0(0)×σ0i, remains the same. But the nonlinear intensity curve is quite sensitive to the change: as the number of participating ground

15 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

state molecules is decreased, it becomes easier to saturate the system by putting all of them in an excited state. The PL curve now rolls over at lower intensities and more closely matches the data. Our calculations also confirm that the I4 dependence of the PL arises from the combination of 2PA and TTA, both of which contribute a factor of I2 to the emissive S1 population. When kfiss = 0 and no T1 population is produced, we observe only an I2 dependence over the whole intensity range (Figure S11). The existence of singlet fission in RUB serves to suppress the contribution of the initially excited S1 population to the overall PL signal, making PL generated by TTA the dominant component, as surmised by previous workers.16, 70 The I4 dependence for a 2PA process is to be expected for any system in which TTA is the dominant contributor to the PL. Finally, we emphasize that the parameters in Table 2 should be interpreted with caution. For example, in most cases changing the parameter by a factor of 2 did not generate a large deviation from the data. In Figure S11 in the Supporting Information we provide graphs that illustrate the sensitivity of the calculated intensity dependence on different model parameters. The identity of the low energy interband state in RUB is the most interesting question raised by this work. One obvious candidate is the unperturbed molecular T1 state. However, this state is unlikely for several reasons. First, the anisotropy in Figure 6 suggests that the S0i transition is not oriented along the short axis of the molecule, which lies along the direction of light propagation, perpendicular to the ab crystal face. If we assume that the S0T1 transition in RUB is similar to that of anthracene and lies along the short molecular axis,71 it would not couple to the light incident perpendicular to the ab face and would not be expected to show a strong anisotropy. Second, the S0T1 cross section for molecular RUB has not been measured, but it can safely be assumed to be less than 10-21 cm2 based on our measurements, much smaller than

16 ACS Paragon Plus Environment

Page 16 of 30

Page 17 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

the values of σ0i needed to fit our data. Third, from our measurements of the delayed fluorescence lifetime, the bulk crystal T1 lifetime is too long to be consistent with the τi value obtained for the interband state. Finally, in dilute solution, molecular RUB acts as a normal chromophore that undergoes nonresonant 2PA. Furthermore, if there was significant absorption directly to the T1 state, TTA could only generate an I2 dependence at low intensities, not the observed I4 dependence (Supporting Information). Although the unperturbed S0T1 transition has negligible oscillator strength at the molecular level, there is evidence that intermolecular interactions can enhance this oscillator strength. One example is the noncovalent association of O2 to a conjugated molecule, which can enhance its S0T1 transition via a charge-transfer interaction,72-73 resulting in absorption coefficients on the order of 100 M-1cm-1 in pressurized solutions.74 We do not think that O2 plays a role in the current experiments, since exposing the crystals to a vacuum of 10-5 Torr for 48 hours did not reduce the upconverted PL signal. Furthermore, O2 has a low energy singlet state at ~8000 cm-1 that is known to act as an energy acceptor for RUB triplet states, so we would expect an O2 defect to quench the TTA.75 But O2 is not the only molecule that can lead to this effect. Adding a small amount of 2-methylanthracene to an anthracene crystal has been shown to enhance its phosphorescence yield by an order of magnitude.76 This result suggests that interactions between conjugated molecules can also affect the S0T1 transition, although the mechanism has not been investigated in any detail. Given that the upconversion effect is observed for samples of varying purity and under different atmospheres, it is unlikely that it results from an extrinsic impurity molecule. Instead, we believe the most likely explanation is that the interband state arises from some defect that is present in the RUB itself. One possibility is that intermolecular interactions involving only RUB

17 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 30

molecules could be responsible for the enhanced 2PA. RUB can crystallize in a variety of polymorphs77 and its PL spectrum is highly sensitive to different preparation conditions.55,

78

These facts suggest that RUB single crystals can possess a variety of internal molecular packing configurations, and it is possible that one or more of these configurations leads to an enhanced S0T1 transition. A second possibility is that chemically modified RUB molecules give rise to interband states. Recent theoretical work has shown that atomic defects generated by O or Hatom insertion into pentacene and RUB molecules can induce interband states, although the spin multiplicity of these states is not specified.79-80 Irkhin et al. used H bombardment to introduce a high concentration of such defects into a RUB crystal and found they lowered the triplet lifetime to ~1 µs, close to what we obtain for the intermediate state.81 In our samples, such defective molecules would be present in much lower concentrations and not affect the bulk triplet lifetime. Chemical defects can also be introduced during an experiment. When organic crystals are exposed to high-vacuum gauges in vacuum chambers, certain defects (traps) can be created at the surface of the crystals.82 However, we do not think these vacuum-induced surface traps are the origin of the interband absorption, since it is present in air, under vacuum, and also in inert gas atmospheres. Finally, we note that these mechanisms are not mutually exclusive. For example, a chemically modified RUB molecule could also give rise to defective packing at specific sites within the crystal. Such sites would still have a well-defined orientation that leads to highly anisotropic absorption, as observed. A triplet on such a site would also be expected to have a different energy than that of the bulk triplet excitons. This site would act as a 2PA “hotspot,” producing singlets that rapidly fission and produce at least one mobile triplet that can diffuse and undergo TTA to generate the delayed fluorescence signal. It is possible that the different

18 ACS Paragon Plus Environment

Page 19 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

molecular configuration at this site would also affect the rates of other photophysical processes. This would explain why our τi = 1.5 µs is shorter than that measured for bulk triplets. Characterizing the structure of such defect sites will be a challenge for future investigations.

Conclusion We have shown that crystalline RUB supports a NIR 2PA process that generates upconverted PL in the visible range even using low power CW laser excitation. This upconverted luminescence is readily observed by eye. The origin of this enhanced 2PA process is a resonant interband state with a lifetime of 1.5 µs. After absorption, a model that assumes the singlets undergo fission and then triplet-triplet annihilation can reproduce the intensity dependence of the PL signal specifically the I4I2 which has been seen before in anthracene crystals. Our results suggest that the interband state that facilitates the 2PA arises from a low concentration species. The generality of this effect suggests that this species is not an extrinsic impurity molecule like O2, but possibly a chemically modified RUB molecule or packing defect. We hypothesize that this species can enhance oscillator strength to the S0T1 optical transition. However, we should stress that the precise origin of this interband state and its oscillator strength has not been conclusively determined, and further experiments, for example a careful measurement of the PL excitation spectrum, will be necessary to establish its exact nature. Such investigations are warranted, since the results in this paper suggest that it may be possible to design interband states in purely organic molecular solids with useful photophysical properties.

Supporting Information Diagrams of experimental setups; Intensity dependent PL measurements of other acenes; Timeresolved PL trace of RUB; Sensitivity of model to fit parameters.

19 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Author Information Corresponding Author *

Email: [email protected]

Acknowledgements This research was supported by the National Science Foundation grant DMR-1508099 (CJB). HHC and VP are grateful to Rutgers Energy Institute (REI), and Center for Advanced SoftElectronics funded by the Ministry of Science, ICT and Future Planning as Global Frontier Project (CASE-2011-0031628), for partial support. We thank Dr. Jon Burdett and Craic Technologies for the use of a microscope spectrophotometer. References 1. Burdett, J. J.; Bardeen, C. J., Quantum Beats in Crystalline Tetracene Delayed Fluorescence Due to Triplet Pair Coherences Produced by Direct Singlet Fission. J. Am. Chem. Soc. 2012, 134, 8597-8607. 2. Roberts, S. T.; McAnally, R. E.; Mastron, J. N.; Webber, D. H.; Whited, M. T.; Brutchey, R. L.; Thompson, M. E.; Bradforth, S. E., Efficient Singlet Fission Discovered in a Disordered Acene Film. J. Am. Chem. Soc. 2012, 134, 6388-6400. 3. Burdett, J. J.; Bardeen, C. J., The Dynamics of Singlet Fission in Crystalline Tetracene and Covalent Analogs. Acc. Chem. Res. 2013, 46, 1312-1320. 4. Rao, A.; Wilson, M. W. B.; Hodgkiss, J. M.; Albert-Seifried, S.; Bassler, H.; Friend, R. H., Exciton Fission and Charge Generation via Triplet Excitons in Pentacene/C60 Bilayers. J. Am. Chem. Soc. 2010, 132, 12698-12703. 5. Wilson, M. W. B.; Rao, A.; Clark, J.; Kumar, R. S. S.; Brida, D.; Cerullo, G.; Friend, R. H., Ultrafast Dynamics of Exciton Fission in Polycrystalline Pentacene. J. Am. Chem. Soc. 2011, 133, 11830-11833. 6. Albrecht, W. G.; Michel-Beyerle, M. E.; Yakhot, V., Exciton Fission in Excimer Forming Crystal. Dynamics of an Excimer Build-Up in α-Perylene. Chem. Phys. 1978, 35, 193-200. 7. Eaton, S. W.; Shoer, L. E.; Karlen, S. D.; Dyar, S. M.; Margulies, E. A.; Veldkamp, B. S.; Ramanan, C.; Hartzler, D. A.; Savikhin, S.; Marks, T. J.; Wasielewski, M. R., Singlet Exciton Fission in Polycrystalline Thin Films of a Slip-Stacked Perylenediimide. J. Am. Chem. Soc. 2013, 135, 14701-14712. 8. Schwerin, A. F.; Johnson, J. C.; Smith, M. B.; Sreearunothai, P.; Popović, D.; Černý, J.; Havlas, Z.; Paci, I.; Akdag, A.; MacLeod, M. K. et al., Toward Designed Singlet Fission: Electronic States and Photophysics of 1,3-Diphenylisobenzofuran. J. Phys. Chem. A 2010, 114, 1457-1473. 9. Schrauben, J. N.; Ryerson, J. L.; Michl, J.; Johnson, J. C., Mechanism of Singlet Fission in Thin Films of 1,3-Diphenylisobenzofuran. J. Am. Chem. Soc. 2014, 136, 7363-7373. 10. Kepler, R. G.; Caris, J. C.; Avakian, P.; Abramson, E., Triplet Excitons and Delayed Fluorescence in Anthracene Crystals. Phys. Rev. Lett. 1963, 10, 400-402. 11. Singh, S.; Jones, W. J.; Siebrand, W.; Stoicheff, B. P.; Schneider, W. G., Laser Generation of Excitons and Fluorescence in Anthracene Crystals. J. Chem. Phys. 1965, 42, 330-342. 12. Shockley, W.; Queisser, H. J., Detailed Balance Limit of Efficiency of p-n Junction Solar Cells. J. Appl. Phys. 1961, 32, 510-519. 13. Shpaisman, H.; Niitsoo, O.; Lubomirsky, I.; Cahen, D., Can Up- and Down-Conversion and MultiExciton Generation Improve Photovoltaics? Sol. Energy Mater. Sol. Cells 2008, 92, 1541-1546. 14. Congreve, D. N.; Lee, J.; Thompson, N. J.; Hontz, E.; Yost, S. R.; Reusswig, P. D.; Bahlke, M. E.; Reineke, S.; Van Voorhis, T.; Baldo, M. A., External Quantum Efficiency Above 100% in a Singlet-ExcitonFission–Based Organic Photovoltaic Cell. Science 2013, 340, 334-337. 20 ACS Paragon Plus Environment

Page 20 of 30

Page 21 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

15. Schulze, T. F.; Schmidt, T. W., Photochemical Upconversion: Present Status and Prospects for its Application to Solar Energy Conversion. Energy Environ. Sci. 2015, 8, 103-125. 16. Biaggio, I.; Irkhin, P., Extremely Efficient Exciton Fission and Fusion and its Dominant Contribution to the Photoluminescence Yield in Rubrene Single Crystals. Appl. Phys. Lett. 2013, 103, 263301. 17. Jankus, V.; Snedden, E. W.; Bright, D. W.; Arac, E.; Dai, D.; Monkman, A. P., Competition Between Polaron Pair Formation and Singlet Fission Observed in Amorphous Rubrene Films. Phys. Rev. B 2013, 87, 224202. 18. Bera, K.; Douglas, C. J.; Frontiera, R. R., Femtosecond Raman Microscopy Reveals Structural Dynamics Leading to Triplet Separation in Rubrene Singlet Fission. J. Phys. Chem. Lett. 2017, 8, 59295934. 19. Sutton, C.; Tummala, N. R.; Beljonne, D.; Brédas, J.-L., Singlet Fission in Rubrene Derivatives: Impact of Molecular Packing. Chem. Mater. 2017, 29, 2777-2787. 20. Breen, I.; Tempelaar, R.; Bizimana, L. A.; Kloss, B.; Reichman, D. R.; Turner, D. B., Triplet Separation Drives Singlet Fission after Femtosecond Correlated Triplet Pair Production in Rubrene. J. Am. Chem. Soc. 2017, 139, 11745-11751. 21. Ishibashi, Y.; Inoue, Y.; Asahi, T., The Excitation Intensity Dependence of Singlet Fission Dynamics of a Rubrene Microcrystal Studied by Femtosecond Transient Microspectroscopy. Photochem. Photobiol. Sci. 2016, 15, 1304-1309. 22. Ma, L.; Zhang, K.; Kloc, C.; Sun, H.; Michel-Beyerle, M. E.; Gurzadyan, G. G., Singlet Fission in Rubrene Single Crystal: Direct Observation by Femtosecond Pump-Probe Spectroscopy. Phys. Chem. Chem. Phys. 2012, 14, 8307-8312. 23. Piland, G. B.; Burdett, J. J.; Kurunthu, D.; Bardeen, C. J., Magnetic Field Effects on Singlet Fission and Fluorescence Decay Dynamics in Amorphous Rubrene. J. Phys. Chem. C 2013, 117, 1224-1236. 24. Miyata, K.; Kurashige, Y.; Watanabe, K.; Sugimoto, T.; Takahashi, S.; Tanaka, S.; Takeya, J.; Yanai, T.; Matsumoto, Y., Coherent Singlet Fission Activated by Symmetry Breaking. Nat. Chem. 2017, 9, 983. 25. Ryasnyanskiy, A.; Biaggio, I., Triplet Exciton Dynamics in Rubrene Single Crystals. Phys. Rev. B 2011, 84, 193203. 26. Irkhin, P.; Najafov, H.; Podzorov, V., Steady-State Photoconductivity and Multi-Particle Interactions in High-Mobility Organic Semiconductors. Sci. Rep. 2015, 5, 15323. 27. Singh-Rachford, T. N.; Castellano, F. N., Photon Upconversion Based on Sensitized Triplet–Triplet Annihilation. Coord. Chem. Rev. 2010, 254, 2560-2573. 28. Huang, Z.; Li, X.; Mahboub, M.; Hanson, K. M.; Nichols, V. M.; Le, H.; Tang, M. L.; Bardeen, C. J., Hybrid Molecule–Nanocrystal Photon Upconversion Across the Visible and Near-Infrared. Nano Lett. 2015, 15, 5552-5557. 29. Nienhaus, L.; Wu, M.; Geva, N.; Shepherd, J. J.; Wilson, M. W. B.; Bulović, V.; Van Voorhis, T.; Baldo, M. A.; Bawendi, M. G., Speed Limit for Triplet-Exciton Transfer in Solid-State PbS NanocrystalSensitized Photon Upconversion. ACS Nano 2017, 11, 7848-7857. 30. Mahboub, M.; Huang, Z.; Tang, M. L., Efficient Infrared-to-Visible Upconversion with Subsolar Irradiance. Nano Lett. 2016, 16, 7169-7175. 31. Amemori, S.; Sasaki, Y.; Yanai, N.; Kimizuka, N., Near-Infrared-to-Visible Photon Upconversion Sensitized by a Metal Complex with Spin-Forbidden yet Strong S0–T1 Absorption. J. Am. Chem. Soc. 2016, 138, 8702-8705. 32. Duan, P.; Yanai, N.; Nagatomi, H.; Kimizuka, N., Photon Upconversion in Supramolecular Gel Matrixes: Spontaneous Accumulation of Light-Harvesting Donor–Acceptor Arrays in Nanofibers and Acquired Air Stability. J. Am. Chem. Soc. 2015, 137, 1887-1894.

21 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

33. Mahboub, M.; Maghsoudiganjeh, H.; Pham, A. M.; Huang, Z.; Tang, M. L., Triplet Energy Transfer from PbS(Se) Nanocrystals to Rubrene: the Relationship Between the Upconversion Quantum Yield and Size. Adv. Funct. Mater. 2016, 26, 6091-6097. 34. Deng, F.; Sun, W.; Castellano, F. N., Texaphyrin Sensitized Near-IR-to-Visible Photon Upconversion. Photochem. Photobiol. Sci. 2014, 13, 813-819. 35. Deng, F.; Sommer, J. R.; Myahkostupov, M.; Schanze, K. S.; Castellano, F. N., Near-IR Phosphorescent Metalloporphyrin as a Photochemical Upconversion Sensitizer. Chem. Comm. 2013, 49, 7406-7408. 36. Cheng, Y. Y.; Khoury, T.; Clady, R. G. C. R.; Tayebjee, M. J. Y.; Ekins-Daukes, N. J.; Crossley, M. J.; Schmidt, T. W., On the Efficiency Limit of Triplet-Triplet Annihilation for Photochemical Upconversion. Phys. Chem. Chem. Phys. 2010, 12, 66-71. 37. Wu, T. C.; Congreve, D. N.; Baldo, M. A., Solid State Photon Upconversion Utilizing Thermally Activated Delayed Fluorescence Molecules as Triplet Sensitizer. Appl. Phys. Lett. 2015, 107, 031103. 38. Sripathy, K.; MacQueen, R. W.; Peterson, J. R.; Cheng, Y. Y.; Dvorak, M.; McCamey, D. R.; Treat, N. D.; Stingelin, N.; Schmidt, T. W., Highly Efficient Photochemical Upconversion in a Quasi-Solid Organogel. J. Mater. Chem. C 2015, 3, 616-622. 39. Singh-Rachford, T. N.; Lott, J.; Weder, C.; Castellano, F. N., Influence of Temperature on LowPower Upconversion in Rubbery Polymer Blends. J. Am. Chem. Soc. 2009, 131, 12007-12014. 40. Wu, M.; Congreve, D. N.; Wilson, M. W. B.; Jean, J.; Geva, N.; Welborn, M.; Van Voorhis, T.; Bulović, V.; Bawendi, M. G.; Baldo, M. A., Solid-State Infrared-to-Visible Upconversion Sensitized by Colloidal Nanocrystals. Nat. Photon. 2016, 10, 31-34. 41. Liu, H.; Yan, F.; Li, W.; Lee, C.-S.; Chu, B.; Chen, Y.; Li, X.; Han, L.; Su, Z.; Zhu, J.; Kong, X.; Zhang, L.; Luo, Y., Up-Conversion Luminescence of Crystalline Rubrene Without any Sensitizers. Org. Electron. 2010, 11, 946-950. 42. Ma, L.; Galstyan, G.; Zhang, K.; Kloc, C.; Sun, H.; Soci, C.; Michel-Beyerle, M. E.; Gurzadyan, G. G., Two-Photon-Induced Singlet Fission in Rubrene Single Crystal. J. Chem. Phys. 2013, 138, 184508. 43. Irkhin, P.; Biaggio, I., Two-Photon Absorption Spectroscopy of Rubrene Single Crystals. Phys. Rev. B 2014, 89, 201202. 44. Zhao, L.; Yang, G.; Su, Z.; Qin, C.; Yang, S., Theoretical Studies on One- and Two-Photon Absorption Properties of Rubrene and its Derivatives. Synth. Met. 2006, 156, 1218-1224. 45. de Boer, R. W. I.; Gershenson, M. E.; Morpurgo, A. F.; Podzorov, V., Organic Single-Crystal FieldEffect Transistors. Phys. Status Solidi A 2004, 201, 1302-1331. 46. Wolleschensky, R.; Feurer, T.; Sauerbrey, R.; Simon, U., Characterization and Optimization of a Laser-Scanning Microscope in the Femtosecond Regime. Appl. Phys. B 1998, 67, 87-94. 47. Makarov, N. S.; Drobizhev, M.; Rebane, A., Two-Photon Absorption Standards in the 550–1600 nm Excitation Wavelength Range. Opt. Express 2008, 16, 4029-4047. 48. Sailaja, R.; Bisht, P. B.; Singh, C. P.; Bindra, K. S.; Oak, S. M., Influence of Multiphoton Events in Measurement of Two-Photon Absorption Cross-Sections and Optical Nonlinear Parameters Under Femtosecond Pumping. Opt. Commun. 2007, 277, 433-439. 49. Fischer, A.; Cremer, C.; Stelzer, E. H. K., Fluorescence of Coumarins and Xanthenes after TwoPhoton Absorption with a Pulsed Titanium–Sapphire Laser. Appl. Opt. 1995, 34, 1989-2003. 50. Xu, C.; Webb, W. W., Measurement of Two-Photon Excitation Cross Sections of Molecular Fluorophores with Data from 690 to 1050 nm. J. Opt. Soc. Am. B 1996, 13, 481-491. 51. Lee, S.; Thomas, K. R. J.; Thayumanavan, S.; Bardeen, C. J., Dependence of the Two-Photon Absorption Cross Section on the Conjugation of the Phenylacetylene Linker in Dipolar Donor−Bridge−Acceptor Chromophores. J. Phys. Chem. A 2005, 109, 9767-9774. 52. Rumi, M.; Perry, J. W., Two-Photon Absorption: an Overview of Measurements and Principles. Adv. Opt. Photon. 2010, 2, 451-518. 22 ACS Paragon Plus Environment

Page 22 of 30

Page 23 of 30 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

53. Webman, I.; Jortner, J., Energy Dependence of Two-Photon-Absorption Cross Sections in Anthracene. J. Chem. Phys. 1969, 50, 2706-2716. 54. Birks, J. B., Photophysics of Aromatic Molecules. Wiley - Interscience: New York, 1970. 55. Irkhin, P.; Ryasnyanskiy, A.; Koehler, M.; Biaggio, I., Absorption and Photoluminescence Spectroscopy of Rubrene Single Crystals. Phys. Rev. B 2012, 86, 085143. 56. Nakashima, T. T.; Offen, H. W., Crystal Spectra of Tetracene and Rubrene Under Pressure. J. Chem. Phys. 1968, 48, 4817-4821. 57. Yamagata, H.; Norton, J.; Hontz, E.; Olivier, Y.; Beljonne, D.; Brédas, J. L.; Silbey, R. J.; Spano, F. C., The Nature of Singlet Excitons in Oligoacene Molecular Crystals. J. Chem. Phys. 2011, 134, 204703. 58. Lim, S.-H.; Bjorklund, T. G.; Spano, F. C.; Bardeen, C. J., Exciton Delocalization and Superradiance in Tetracene Thin Films and Nanoaggregates. Phys. Rev. Lett. 2004, 92, 107402. 59. Port, H.; Rund, D., Temperature Dependent Triplet Absorption Spectra of Anthracene Crystals. Photoexcitation Spectroscopy with a Dye-Laser. Chem. Phys. Lett. 1978, 54, 474-478. 60. Vaubel, G.; Baessler, H., Delayed Fluorescence and Triplet Lifetime in Tetracene Crystals. Phys. Status Solidi B 1970, 37, K31-K34. 61. Peticolas, W. L., Multiphoton Spectroscopy. Annu. Rev. Phys. Chem. 1967, 18, 233-260. 62. Wu, C. W., W. W., Multiphoton Excitation of Molecular Fluorophores and Nonlinear Laser Microscopy. In Top. Fluoresc. Spectrosc., Lakowicz, J. R., Ed. Plenum Press: New York, 1997; Vol. 5: Nonlinear and Two-Photon-Induced Fluorescence, pp 471-540. 63. Manifacier, J. C.; Gasiot, J.; Fillard, J. P., A Simple Method for the Determination of the Optical Constants n, k and the Thickness of a Weakly Absorbing Thin Film. J. Phys. E: Sci. Instrum. 1976, 9, 1002. 64. Podzorov, V.; Menard, E.; Borissov, A.; Kiryukhin, V.; Rogers, J. A.; Gershenson, M. E., Intrinsic Charge Transport on the Surface of Organic Semiconductors. Phys. Rev. Lett. 2004, 93, 086602. 65. Jiang, X.-F.; Pan, Y.; Jiang, C.; Zhao, T.; Yuan, P.; Venkatesan, T.; Xu, Q.-H., Excitation Nature of Two-Photon Photoluminescence of Gold Nanorods and Coupled Gold Nanoparticles Studied by TwoPulse Emission Modulation Spectroscopy. J. Phys. Chem. Lett. 2013, 4, 1634-1638. 66. Ippen, E. P.; Shank, C. V.; Woerner, R. L., Picosecond Dynamics of Azulene. Chem. Phys. Lett. 1977, 46, 20-23. 67. Ward, K. A.; Richman, B. R.; Biaggio, I., Nanosecond Pump and Probe Observation of Bimolecular Exciton Effects in Rubrene Single Crystals. Appl. Phys. Lett. 2015, 106, 223302. 68. Wang, C. H.; Tai, O. Y.-H.; Wang, Y.; Tsai, T.-H.; Chang, N.-C., Non-Quadratic-Intensity Dependence of Two-Photon Absorption Induced Fluorescence of Organic Chromophores in Solution. J. Chem. Phys. 2005, 122, 084509. 69. Silver, M.; Zahlan, A. B., Intensity Dependence of Laser-Induced Delayed Fluorescence. J. Chem. Phys. 1964, 40, 1458-1458. 70. Ma, L.; Zhang, K.; Kloc, C.; Sun, H.; Soci, C.; Michel-Beyerle, M. E.; Gurzadyan, G. G., Fluorescence from Rubrene Single Crystals: Interplay of Singlet Fission and Energy Trapping. Phys. Rev. B 2013, 87, 201203. 71. Padhye, M. R.; McGlynn, S. P.; Kasha, M., Lowest Triplet State of Anthracene. J. Chem. Phys. 1956, 24, 588-594. 72. Tsubomura, H.; Mulliken, R. S., Molecular Complexes and their Spectra. XII. Ultraviolet Absorption Spectra Caused by the Interaction of Oxygen with Organic Molecules. J. Am. Chem. Soc. 1960, 82, 5966-5974. 73. Kristiansen, M.; Scurlock, R. D.; Iu, K. K.; Ogilby, P. R., Charge-Transfer State and Singlet Oxygen (1.DELTA.g O2) Production in Photoexcited Organic Molecule-Molecular Oxygen Complexes. J. Phys. Chem. 1991, 95, 5190-5197. 74. Evans, D. F., 257. Perturbation of Singlet-Triplet Transitions of Aromatic Molecules by Oxygen Under Pressure. J. Chem. Soc. 1957, 0, 1351-1357. 23 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

75. Darmanyan, A. P., Laser Photolysis Study of the Mechanism of Rubrene Quenching by Molecular Oxygen. Chem. Phys. Lett. 1982, 86, 405-410. 76. Brillante, A.; Craig, D. P.; Mau, A. W. H.; Rajikan, J., Impurity Induced Phosphorescence in Anthracene Crystal. Chem. Phys. Lett. 1975, 30, 5-10. 77. Jurchescu, O. D.; Meetsma, A.; Palstra, T. T. M., Low-Temperature Structure of Rubrene Single Crystals Grown by Vapor Transport. Acta Crystallogr. B 2006, 62, 330-334. 78. Chen, Y.; Lee, B.; Fu, D.; Podzorov, V., The Origin of a 650 nm Photoluminescence Band in Rubrene. Adv. Mater. 2011, 23, 5370-5375. 79. Northrup, J. E.; Chabinyc, M. L., Gap States in Organic Semiconductors: Hydrogen- and OxygenInduced States in Pentacene. Phys. Rev. B 2003, 68, 041202. 80. Tsetseris, L.; Pantelides, S. T., Large Impurity Effects in Rubrene Crystals: First-Principles Calculations. Phys. Rev. B 2008, 78, 115205. 81. Irkhin, P.; Biaggio, I.; Zimmerling, T.; Dobeli, M.; Batlogg, B., Defect Density Dependent Photoluminescence Yield and Triplet Diffusion Length in Rubrene. Appl. Phys. Lett. 2016, 108, 063302. 82. Podzorov, V.; Menard, E.; Pereversev, S.; Yakshinsky, B.; Madey, T.; Rogers, J. A.; Gershenson, M. E., Interaction of Organic Surfaces with Active Species in the High-Vacuum Environment. Appl. Phys. Lett. 2005, 87, 093505. 83. Grabolle, M.; Spieles, M.; Lesnyak, V.; Gaponik, N.; Eychmüller, A.; Resch-Genger, U., Determination of the Fluorescence Quantum Yield of Quantum Dots: Suitable Procedures and Achievable Uncertainties. Anal. Chem. 2009, 81, 6285-6294. 84. Murov, S. L. C., Ian; Hug, Gordon L., Handbook of Photochemistry. 2nd ed.; Marcel Dekker, Inc.: New York, 1993.

24 ACS Paragon Plus Environment

Page 24 of 30

Page 25 of 30

Figures and Tables

δ Sample

QY

fcoll

(× CDEFD GHI ∗ K ∗ LMNONPKEC)

C540A

0.5383

0.19

4746-47

RUB

0.9884

0.61

1.6

TET

0.1784

0.16

1.4

Table 1. The parameters used to determine the two-photon cross sections of the chromophores used. The dilute solutions (300 uM) in methylene chloride were all degassed prior to measurement. QY is the quantum yield of the chromophore and fcoll is the integrated area under the PL spectrum excited with the 800 nm pulsed laser over the range of the interference filter.

ki (s-1) 6.67 x 105

k2 (s-1) 6.25 x 107

kfiss (s-1) 5 x 1010

ktrip (s-1) 1.67 x 104

kTTS (cm3 /s) 10-12

kTTQ (cm3 /s) 10-14

σ0i (cm2) 2 x 10-18

σi2 (cm2) 2 x 10-16

Table 2. The parameters used for the kinetic model described in the text that reproduce data in Figure 9. We emphasize that these parameters are not the result of a linear-least squares fit but are derived from literature sources and constraints described in the text.

1000

Corrected Signal (mV)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

100 10 1

C540A RUB TET Slope 2

0.1 0.01 10

100

Power (mW)

Figure 1. Log-log plots of the relative intensity of the integrated PL for samples of coumarin, rubrene and tetracene in dilute solutions of dichloromethane excited via nonresonant 2PA.

25 ACS Paragon Plus Environment

The Journal of Physical Chemistry

Normalized Intensity

1.0

C 532 nm 808 nm 980 nm

0.5

0.0 550

600

650

700

750

Wavelength (nm)

Figure 2. Images of the upconverted visible (~ 600 nm) emission observed from crystalline rubrene powder with femtosecond pulsed (A) and continuous wave (B) NIR excitation at 800 nm. An average intensity of 4 kW/cm2 was used to excite the powder which corresponds to a peak intensity of 0.4 GW/cm2 for the femtosecond pulse train. The PL intensity appears essentially unchanged using either continuous wave or pulsed excitation. (C) The steady-state PL obtained from pristine single crystals of RUB using CW excitation. For each wavelength the laser was incident upon the ab facet of the crystal and polarized parallel to the b axis. 3

3

A 1 0

B 2

TET Log10[Signal]

2

Log10[Signal]

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 30

Pulsed Continuous wave Slope 2

-1 -2 1.5

1 0

RUB Pulsed Continuous wave Slope 2 Slope 4

-1 -2

2.0

2.5

3.0

3.5

4.0

1.5

2.0

2.5

3.0

3.5

4.0

Log10[Intensity]

Log10[Intensity]

Figure 3. Comparison of the excitation intensity dependence (W/cm2) of the integrated upconverted PL signal for pulsed and continuous wave laser excitation at 800 nm for TET (A) and RUB (B) crystals. Lines with a slope of 2 and 4 are included to guide the eye. Linear fits of the intensity dependence yield slopes of 2.03 ± 0.03 and 3.75 ± 0.16 for the middle and low range intensities, respectively.

26 ACS Paragon Plus Environment

Page 27 of 30

0.5

A

Absorbance

Absorbance

0.08

B RUB Etalon Model

RUB DCM

0.4

0.3

0.2

0.06

0.04

0.02

0.1

0.0 400

0.00

500

600

700

800

900

400

1000

600

800

1000

1200

1400

1600

Wavelength (nm)

Wavelength (nm)

Figure 4. The NIR absorption of rubrene solution (black line) in a 10 cm cuvette compared to the baseline absorption of the solvent dichloromethane (red line) (A) and crystalline rubrene absorption spectrum (black line) overlaid with the calculated background (red line) using the method described in ref. 63 (B). 1.0

Normalized Intensity

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

RUB BaSO4

0.8 0.6 0.4 0.2 0.0 900

950

1000

1050

1100

1150

1200

1250

Wavelength (nm)

Figure 5. NIR PL of RUB crystal (red line) and BaSO4 powder (blue line) to show background sensitivity. The PL is excited along the b axis of the RUB crystal with a vertically polarized 808 nm laser diode and collected through an 850 nm long wave pass filter. The peak at 1060 nm is likely a Raman peak with a separation of 3000 cm-1 from the excitation wavelength.

27 ACS Paragon Plus Environment

The Journal of Physical Chemistry

90 120

60

150

30

180

0

210

330

240

300 270

Figure 6. Polarization dependence of upconverted (visible) PL in crystalline rubrene excited along the ab facet using 980 nm NIR laser excitation. An angle of zero degrees corresponds to excitation linearly polarized along the b axis and an angle of 90 degrees corresponds to the crystallographic a axis. Crystal packing on the ab facet of rubrene is shown for clarity.

Signal / Pulse (mV / MHz)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

25

RUB TET Rhod B / PMMA

20 15 10 5 0 0

200

400

600

800

1000

1200

Period (ns) Figure 7. Dependence of the per-pulse upconverted (visible) PL intensity on the pulse spacing of the exciting NIR laser (800 nm). The error bars represent the standard deviation of 9 separate scans taken on five different spots of the same pristine RUB crystal which was ~100 µm thick (red squares). RUB was excited along the crystallographic b axis. For comparison, the same data for crystalline TET and the dye Rhodamine B in a polymer show no dependence on pulse period. All measurements were performed under vacuum.

28 ACS Paragon Plus Environment

Page 28 of 30

Page 29 of 30

Figure 8. A Jablonski diagram showing the parameters used in the kinetic model (Equations 2-5). The parameters are described in the text. Fission and fusion produce or annihilate triplets on separate RUB molecules, respectively. Here, the T1 state represents these separate, degenerate triplets for simplicity. The triplet quenching mechanism described by rate kTTQ is omitted for clarity.

3

1.00

A

B 2

0.37

Log10[Signal]

Ln[Signal / Pulse]

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

0.14

RUB τi = 60 µs

0.05

1

0

RUB NS0 = 1.434×1021

-1

τi = 10 µs

NS0 = 1.434×1019

τi = 1.5 µs

NS0 = 1.434×1017

0.02 0

200

400

600

800

1000

1200

-2 3.5

4.0

4.5

5.0

Log10[Intensity]

Period (ns)

Figure 9. (A) Modelling of the dependence of the per pulse signal on the period of the laser pulses. The red squares are the data points. The solid lines are calculated with the δ pulse simulation described in the text assuming an intermediate state lifetime (τi) of 60 µs (orange), 10 µs (blue), 1.5 µs (black). (B) Modelling of the dependence of the upconverted PL observed in single crystals on average 800 nm laser intensity. The red squares are the data points. The solid lines represent a four-state model described by rate Equations 2-5 in a steady state (dNX/dt = 0) for different interband state densities. NS0=1.434×1021 cm-3 assumes that the density of interband states equals the density of RUB molecules in the crystal (orange line). Reducing the interband density by a factor of 102 (blue line) leads to a good overlap with the experimental points, while reducing it by a factor of 104 (black line) leads to poor overlap. The intensity units for this plot are (photons/ns*µm2) for computational convenience and multiplying this number by 2.48×10-2 gives intensity in W/cm2 at 800 nm.

29 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

TOC Image

30 ACS Paragon Plus Environment

Page 30 of 30