Integrated Photon Upconversion Dye-Sensitized ... - ACS Publications

Jun 28, 2019 - light source, such as sunlight.1,2 In the solar cell field, it is anticipated that the .... useful for triplet energy transfer from a p...
0 downloads 0 Views 973KB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article Cite This: ACS Omega 2019, 4, 11271−11275

http://pubs.acs.org/journal/acsodf

Integrated Photon Upconversion Dye-Sensitized Solar Cell by Coadsorption with Derivative of Pt−Porphyrin and Anthracene on Mesoporous TiO2 Tatsuro Morifuji, Yuya Takekuma, and Morio Nagata* Department of Industrial Chemistry, Graduate School of Engineering, Tokyo University of Science, 12-1 Ichigaya-funagawara, Shinjuku, Tokyo 162-0826, Japan Downloaded via 37.44.253.52 on July 29, 2019 at 05:24:56 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

S Supporting Information *

ABSTRACT: Photon upconversion via triplet−triplet annihilation (TTA-UC) is a process that converts two lower-energy photons to a higher-energy photon, which is expected to increase the maximum solar cell efficiency beyond the Shockley−Queisser limit. To incorporate TTA-UC into a dye-sensitized solar cell (DSSC), we used a coadsorption approach, in which both a TTA-UC donor with an alkyl carboxylic acid chain and a TTA-UC acceptor dye were adsorbed onto mesoporous TiO2. Incident photon-to-current conversion efficiency spectra and excitation intensity dependence indicated that a photocurrent was generated under irradiation at wavelengths above 490 nm by the TTA-UC mechanism. The power-conversion efficiency of the DSSC was increased to 0.72%, and the photocurrent contributed by TTA-UC was 0.036 mA cm−2 under 1 sun irradiation.

1. INTRODUCTION Photon upconversion via triplet−triplet annihilation (TTAUC) is a process whereby two lower-energy photons are converted to a higher-energy photon. Upconversion is an excellent strategy for solar applications because higher-energy photons can be generated with a low-intensity noncoherent light source, such as sunlight.1,2 In the solar cell field, it is anticipated that the maximum solar cell efficiency will exceed the Shockley−Queisser limit.3−5 Among them, the application of TTA-UC technology is expected to greatly contribute to the improvement of the efficiency of dye-sensitized solar cells (DSSCs).6 The DSSCs require the energy of the lowest unoccupied molecular orbital (LUMO) of the dye to be higher than the energy of the conduction band edge of TiO2, while the highest occupied molecular orbital (HOMO) energy of the dye must be lower than the redox potential of the redox agent. The maximum voltage is the difference between the conduction band edge of TiO2 and the redox potential of the redox agent. The trade-off of increasing the maximum voltage is that the spectral response corresponding to the difference between the HOMO and the LUMO remains unchanged. One strategy for integrating TTA-UC into DSSCs is to use TTA-UC as a filter or reflector in existing solar cells.7,8 In such a system, however, there is a significant loss of light before it reaches the dyes. This is due to self-absorption, reflection, and scattering of the used glass substrate. The strategy of directly incorporating TTA-UC into photocurrent generation in solar cells has been studied as a means of bypassing the parasitic absorption problem.9−14 In a report by Simpson et al., the © 2019 American Chemical Society

triplet acceptor molecule in a DSSC was adsorbed on TiO2 and used as the sensitizing dye, while the triplet donor molecule was dissolved in the electrolyte.9 In addition, Ahmad et al. assembled a photoelectrochemical device that incorporated a metal−organic framework, a triplet acceptor molecule adsorbed onto TiO2, and a triplet donor molecule dissolved in an electrolyte.10 In the power-conversion mechanism of the homogeneous device, the energy from low-energy photons is transferred from the triplet donor molecule in the electrolyte to the triplet acceptor molecule on TiO2, and a high-energy photon is generated by triplet−triplet annihilation (TTA). In contrast, DSSCs in which both the triplet acceptor and the donor are present as self-assembled multilayers on TiO2 have been devised.11−14 In these DSSCs, sensitizer molecules are coordinated to surface-bound acceptor molecules via metal-ion linkages. These devices exhibit higher photocurrents and better overall efficiency than heterogeneous devices. This approach circumvents the problem of limited triplet-donor solubility in the electrolyte. In this work, we introduce a co-adsorption approach based on previous reports of Morandeira and co-workers. The TTAUC donor, Pt(II)−porphyrin dye with an attached carboxylic alkyl chain spacer (PtTPO), and the acceptor, diphenylanthracene dye (ADDA), are both adsorbed on mesoporous TiO2. In this co-adsorption approach, it is expected to improve the efficiency of the triplet energy transfer and triplet−triplet Received: April 27, 2019 Accepted: June 18, 2019 Published: June 28, 2019 11271

DOI: 10.1021/acsomega.9b01210 ACS Omega 2019, 4, 11271−11275

ACS Omega

Article

the ADDA−TiO2 electrode was similar to the absorption spectrum of ADDA in solution (Figure S1). The absorption spectra of the TiO2 electrode with co-adsorbed ADDA and PtTPO contained peaks that overlapped with those of both ADDA and PtTPO, indicating the two dyes co-adsorbed on the TiO2 electrode. DSSCs were assembled with stained TiO2 photoanodes, Pt counter electrodes held together with Surlyn thermoforming film, and acetonitrile containing 0.6 M 1,2-dimethyl-3propylimidazolium iodide (DMPII), 0.1 M Lil, 0.05 M I2, and 0.5 M 4-tert-butylpyridine (TBP) as the electrolyte solution. The electrolyte solution was injected through a hole drilled into the counter electrode. The photocurrent density−voltage (J−V) curves obtained under AM 1.5G (100 mW cm−2) simulated solar irradiation are shown in Figure 2.

annihilation (TTA) due to shorter distances between the coadsorbed TTA-UC donor and acceptor in addition to solving the solubility problem. The co-adsorbed TiO2 photoanode exhibits efficient photon-to-electron conversion integrated TTA-UC. A schematic of the TTA-UC molecules co-adsorbed on TiO2 is shown in Figure 1a. The proposed power-

Figure 1. (a) Schematic representation of TiO2 electrode with coadsorbed TTA-UC molecules. (b) Absorption spectra of ADDA, PtTPO, and ADDA/PtTPO (10:1) adsorbed on TiO2 film.

generation mechanism begins with excitation of the donor molecule. Following the intersystem crossing to the triplet excited state, triplet energy is transferred from the donor molecule to the acceptor molecule via Dexter energy transfer.15,16 The triplet excited state then migrates through the acceptor until two triplet acceptors in close enough proximity are annihilated. One acceptor returns to the ground state and the other is promoted to the singlet excited state. In fact, TTA-UC on the metal-oxide ZrO2, which has surface properties that are comparable to those of TiO2, has been reported.17−19 Its conduction band is approximately 1.3 eV higher than that of TiO2. Thus, TTA-UC fluorescence can be achieved by eliminating the possibility of electron injection from the singlet excited state of the acceptor into the semiconductor.17 In this study, an alkyl chain molecular spacer was introduced between the dye and the TiO2 anchor to prevent the quenching of PtTPO by electron injection and energy transfer into TiO2. When an alkyl chain spacer was employed for a DSSC in which a light-harvesting dye was applied to TiO2, electron injection from the donor molecules into TiO2 was prevented, and Förster resonant energy transfer from the light-harvesting dyes to the sensitizing dyes was allowed to occur.20,21 It has been suggested that a molecular spacer might also be useful for triplet energy transfer from a porphyrin derivative to an anthracene derivative.22 The acceptor molecule in the singlet excited state would then inject electrons into the conduction band of TiO2 to generate a photocurrent. In this co-adsorbed configuration, ADDA acts as a TiO2-sensitizing dye. Photons are converted to electrons, which are then injected by ADDA in the excited state into the conduction band of TiO2.

Figure 2. Photocurrent density−voltage (J−V) curves of DSSCs based on TiO2 (black), ADDA−TiO2 (blue), PtTPO−TiO2 (red), and ADDA/PtTPO−TiO2 (green) under simulated AM 1.5G reference solar irradiation.

The incident photon-to-current conversion efficiency (IPCE) spectra shown in Figure 3 were collected to evaluate the DSSC

Figure 3. Incident photon-to-current conversion efficiency (IPCE) spectra of DSSCs based on TiO2 (black), ADDA−TiO2 (blue), PtTPO−TiO2 (red), and ADDA/PtTPO−TiO2 (green).

photovoltaic performance under monochromatic light. The DSSC made with only the PtTPO-sensitizing molecule was very weakly IPCE, as illustrated by the red line in Figure 3. This indicated that the alkyl chain prevented electron injection from PtTPO into the conduction band of TiO2. The short-circuit current density (JSC) measured in the DSSC based on ADDA/PtTPO−TiO2 (10:1), indicated by the green line in Figure 2, was higher than those of the ADDA− TiO2 and PtTPO−TiO2 DSSCs. The power-conversion efficiency of the DSSC with co-adsorbed ADDA and PtTPO increased to 0.72%. In this device, the IPCE decreased in the

2. RESULTS AND DISCUSSION ADDA−TiO2 and PtTPO−TiO2 electrodes were prepared by immersion for 1 day in solutions of the respective dyes. An ADDA/PtTPO−TiO2 (10:1) electrode (10:1 means of concentration ratio of dyes in the TiO2 electrode immersion solution) was fabricated by immersion in a solution containing 1 mM ADDA and 0.1 mM PtTPO for 1 day. The absorption spectra, obtained via UV−vis spectroscopy, of the PtTPO− TiO2 electrode showed two characteristic peaks, referred to as the Soret band and the Q band. The absorption spectrum of 11272

DOI: 10.1021/acsomega.9b01210 ACS Omega 2019, 4, 11271−11275

ACS Omega

Article

Table 1. Amount of ADDA and PtTPO Adsorbed on TiO2 Photoanode, PCE of the DSSCs Using Each Photoanode, and JUC concentration ratio ADDA/PtTPOa

ADDAb (nmol cm−2)

only ADDA 20:1 10:1 1:1 1:10 only PtTPO

70 65 58 29 12

PtTPOb (nmol cm−2)

ADDA/ PtTPO

PCEc,d (%)

JUCe (mA cm−2 sun−2)

6 8 20 32 44

10.8 7.25 1.45 0.375

0.60 0.57 0.72 0.29 0.06 0.04

0.018 0.036 0.009 0.011

a Concentration ratio of dyes in the TiO2 electrode immersion solution. bQuantities of the adsorbed dyes were determined by absorbance spectroscopy with alkaline solutions containing desorbed dyes from the stained TiO2 electrodes. cPCE was measured under AM 1.5G irradiation (100 mW cm−2). dDSSCs were fabricated with an electrolyte consisting of 0.6 M 1,2-dimethyl-3-propylimidazolium iodide (DMPII), 0.1 M Lil, 0.05 M I2, and 0.5 M 4-tert-butylpyridine (TBP) in MeCN. eJUC was determined by subtracting the photocurrent of PtTPO from that of ADDA/ PtTPO under AM 1.5G irradiation with a 490 nm long-pass filter.

It is well established that TTA-UC exhibits anti-Stokes fluorescence, for which intensities can vary between quadratic and linear with respect to excitation power when a noncoherent light source is used. Such a behavior is reflected in a change in slope from 2 at low intensity to 1 at high intensity, which is a clear indicator of a TTA-UC process.24−26 Photocurrents derived from TTA-UC also exhibit a change in slope.10−14 In Figure 4, the photocurrent of the DSSC made with ADDA/PtTPO−TiO2 (10:1) exhibited increases along two straight lines with slopes of 1.5 and 1.0 that bordered around 7 mW cm−2. On the other hand, the photocurrent of ADDA−TiO2 exhibited a linear dependence under an AM 1.5G solar simulator. This is to be expected for a singledyephotoexcitation-electron injection process, which is observed with standard DSSCs. The shift in Figure 4 from quasiquadratic to supralinear dependence indicated that the photocurrent observed at wavelengths longer than 490 nm was generated by the TTA-UC mechanism. The border between the two ranges, which is known as Ith,27 ideally occurs below a solar intensity of 1 sun in solar cells.14 In the DSSC demonstrated here, Ith occurred below this irradiation intensity. Under a light intensity of 1 sun, detection occurred at 76.9 mW cm−2 when the 490 nm long-pass filter was used. Moreover, stained TiO2 photoanodes at four different ADDA-to-PtTPO ratios were prepared by immersing TiO2 electrodes in solutions containing different ADDA concentrations and a constant concentration of PtTPO. Absorption measurements were made with the photoanodes (Figure S2). The dyes were then desorbed from the photoanodes into solutions, which were analyzed by UV−vis spectroscopy, and the concentration ratios of the dyes were determined from the absorbance data. The J−V curves and IPCE spectra (Figures S3 and S4) were obtained for the fabricated DSSCs. The PCE, JUC, and concentration ratios of the adsorbed dyes are summarized in Table 1. The highest PCE was observed in the DSSCs with ADDA/PtTPO−TiO2 (10:1). In the ADDA/ PtTPO−TiO2 (10:1) electrode, the quantities of the adsorbed ADDA and PtTPO were 58 and 8 nmol cm−2, respectively. Efficient upconversion via TTA in solution requires about 100 times more triplet acceptors than donors.27 In the devices that exhibited the maximum PCE and JUC values in this study, the quantity of ADDA triplet acceptors was about 7.25 times greater than that of the donor, PtTPO. In mesoporous ZrO2 films with co-adsorbed derivatives of platinum porphyrin and diphenylanthracene, the maximum UC emission intensity was observed when the quantity of sensitizers exceeded that of the emitters 10-fold.18 The TTA-UC system in this case did not require 100 times more acceptors than donors. In fact, the

spectral range of 300−400 nm, which was the region of absorption by ADDA and PtTPO as indicated by their Soret bands. The decline in the IPCE between 300 and 400 nm was due to the decreasing amounts of adsorbed ADDA, as shown Table 1, as well as absorption competition between ADDA and PtTPO. In contrast, IPCE was enhanced in the spectral range from 500 to 550 nm, shown by the green line in Figure 3, which matched the absorption range of the PtTPO Q-band. However, the increase in the IPCE between 500 and 550 nm could not be solely attributed to electron injection from PtTPO to the conduction band of TiO2. The PtTPO−TiO2based DSSC was very weakly powered, and the ADDA/ PtTPO−TiO2 cell contained less adsorbed PtTPO than the PtTPO−TiO2 cell (Table 1). This indicated that ADDA/ PtTPO co-adsorbed on TiO2 generated a photocurrent by electron injection following photon conversion by TTA-UC. The prevailing figure of merit, JUC,4,23 was 0.036 mA cm−2 sun−2 for ADDA/PtTPO−TiO2 (10:1) owing to upconversion under illumination of 1 sun. This JUC was comparable to those of the TTA-UC-incorporated DSSCs fabricated with other approaches.23 Further support for TTA-UC photocurrent generation was obtained when we examined the dependence of the photocurrent on the excitation intensity in the DSSC employing the ADDA/PtTPO−TiO2 (10:1) electrode (Figure 4). The photocurrent was measured with a source meter under a solar simulator using a Xe lamp with a 490 nm long-pass filter to excite only the PtTPO as light-harvesting dyes on TiO2. Each intensity was measured with a Si photodiode.

Figure 4. Excitation intensity dependence spectra of DSSC photocurrent densities with respect to excitation intensities under an AM 1.5G solar simulator for ADDA/PtTPO−TiO2. A Xe lamp with a 490 nm long-pass filter was used for ADDA/PtTPO−TiO2. 11273

DOI: 10.1021/acsomega.9b01210 ACS Omega 2019, 4, 11271−11275

ACS Omega

Article

CH2Cl2 and H2O. The organic phase was washed with saturated aqueous NaCl (thrice) and dried over Na2SO4. The solvent was removed under vacuum, and the crude product was purified by column chromatography on a silica gel by eluting with CH2Cl2 to give a deep purple solid. Yield: 46.5 mg (83%). Purified product and 8-aminooctanoic acid (14.7 mg, 92.3 μmol) were dissolved in a mixture of CH2Cl2 and MeOH (v/v = 1:1) at 0 °C and stirred at room temperature for 8 h. The obtained product was partitioned between CH2Cl2 and H2O. The organic phase was washed with saturated aqueous NaCl (thrice) and dried over Na2SO4. The solvent was removed under vacuum to give the corresponding free-base porphyrin with an alkyl carboxylic acid chain. 4.4. Pt(II)−8-[4-(10,15,20-Triphenylporphyrin-5-yl)benzamido]octanoic Acid (Metalloporphyrin with an Alkyl Carboxylic Acid Chain). Free-base porphyrin with an alkyl carboxylic acid chain (49.2 mg, 61.5 μmol) and PtCl2 (33 mg, 124 μmol) were dissolved in benzonitrile and refluxed at 200 °C for 1 day. The reaction progress was monitored by changes in the Soret and Q-bands characteristic of the freebase porphyrin reactant and the metalloporphyrin product. After workup, the reaction mixture was evaporated, and then the product was collected as an orange solid. 1 H NMR (CDCl3, ppm): δ 8.76 (s, 6H), 8.68 (d, 2H), 8.50 (d, 2H), 8.24 (d, 2H), 8.15 (d, 6H), 7.75 (m, 9H). MALDITOF (DCTB, m/z): calcd for C53H43N5O3Pt, 992.29; found, 992.30 [M + H+]. 4.5. Device Fabrication. Double-layered nanocrystalline TiO2 photoanodes were prepared by coating two different nanocrystalline TiO2 pastes (Solaronix, T/SP, particle size: 20 nm and R/SP, particle size 400 nm). The transparent nanocrystalline TiO2 paste was coated onto fluorine-doped SnO2 (FTO) conducting glass plates by screen printing at the size of 4 mm × 4 mm, and the electrodes were sintered at 500 °C for 30 min in air. The electrodes were treated in 40 mM TiCl4 aqueous solution at 70 °C for 30 min and washed with water. And then, the light-scattering layer was screen-printed onto the transparent layer and sintered again in the same way as before. The thickness of the TiO2 film is 12 μm (9 μm transparent layer and 3 μm scattering layer), according to the measurement with SURFCOM 1400D (Accretech, Tokyo Seimitsu Co., Ltd.). The nanocrystalline TiO2 electrodes were immersed in each dye solution (1 mM ADDA and PtTPO 0.1 mM in DMSO) for 1 day at room temperature. The photoanode and Pt counter electrode were assembled as a sandwich using 30 μm thick Surlyn spacer. The electrolyte solution was introduced into the cell from predrilled hole on the Pt counter electrode. The electrolyte consisted of 0.6 M 1,2-dimethyl-3-propylimidazolium iodide (DMPImI), 0.05 M I2, 0.1 M LiI, and 0.5 M 4-tert-butylpyridine (TBP) in acetonitrile. 4.6. Photovoltaic Measurements. The photovoltaic performances of the fabricated solar cells were evaluated by the incident photon-to-current conversion efficiency (IPCE) spectra and the photocurrent density−voltage (J−V) measurements on the cells with the area of 0.16 cm2. The IPCE spectra were recorded with S10AC system (Peccell Technologies, Inc., Japan), and the J−V curves were recorded under AM 1.5 one sun condition (100 mW cm−2) with YSS-150A (Yamashita Denso, Japan). The light intensity was calibrated with reference to a Si solar cell (BS-520, Bunkokeiki Co., Ltd., Japan).

IPCE in the DSSC we fabricated with 20:1 ADDA/PtPTO− TiO2 was lower in the spectral range from 500 to 550 nm than the IPCE in the DSSC made with 10:1 ADDA/PtTPO−TiO2. In contrast, the IPCE at ∼515 nm was negligible in the DSSC made with 1:10 ADDA/PtTPO−TiO2. Its IPCE at 515 nm was about the same as that of the DSSC made with PtTPO−TiO2. This may have been due to the nonradiative decay from the ADDA triplet state prior to TTA. The transferred triplet electron would quench without TTA because the likelihood of one excited ADDA molecule being close enough to another excited ADDA molecule for TTA-UC to occur would be small. This was because PtTPO was present in a much greater quantity than ADDA. Consequently, the photocurrent at 515 nm was negligible, even though two dyes were present on the TiO2 electrode.

3. CONCLUSIONS We have demonstrated the co-adsorption of a TTA-UC donor containing an alkyl carboxylic acid chain (PtTPO) and an acceptor (ADDA) on mesoporous TiO2 to integrate TTA-UC into a DSSC. The DSSC based on our co-adsorbed TiO2 photoanode was able to exhibit photocurrent by electron injection derived from TTA-UC under 1 sun (AM 1.5G) irradiation. Overall, the power-conversion efficiency of the DSSC with co-adsorbed ADDA and PtTPO was increased to 0.72%, and the photocurrent contributed by TTA-UC, JUC, was 0.036 mA cm−2. This photocurrent was comparable to that of other TTA-UC DSSCs. The IPCE spectra and excitation intensity dependence analysis indicated that the photocurrent under irradiation at wavelengths of >490 nm was generated through the TTA-UC mechanism. However, our device is still insufficient in terms of photon-to-electron efficiency when compared to other DSSCs. A molecular design that overcomes this insufficiency may be a breakthrough for improving the efficiency of the integrated TTA-UC DSSCs. 4. EXPERIMENTAL SECTION 4.1. General. NMR spectra were recorded on a Bruker Biospin AV400M spectrometer at 400 MHz for 1H in CDCl3 solution. UV−vis spectra and fluorescence spectra were measured on a Hitachi High-Tech Science U-3900 and a JASCO FP6600. Matrix assisted laser desorption ionizationtime of flight (MALDI-TOF) mass spectra were obtained on a JEOL JMS-S3000 SpiralTOF spectrometer with DCTB as the matrix. All of the chemicals were purchased from commercial suppliers and used without purification. Column chromatography was performed with silica gel (Wakogel C-300). 4.2. Synthesis. The synthesis of 4,4′-(anthracene-9,10diyl)dibenzoic acid (ADDA) was carried out based on ref 28. Pt(II)−8-[4-(10,15,20-triphenylporphyrin-5-yl)benzamido]octanoic acid (PtTPO) were prepared under modified condition of literature procedure. 4.3. 8-[4-(10,15,20-Triphenylporphyrin-5-yl)benzamido]octanoic Acid (Free-Base Porphyrin with an Alkyl Carboxylic Acid Chain). 5-(4-Carboxyphenyl)10,15,20-triphenylporphyrin (free-base porphyrin, 49.8 mg, 75.6 μmol) and N-hydroxysuccinimide (13.1 mg, 113.4 μmol) in a mixture of CH2Cl2 and THF (v/v = 2:1) at 0 °C was treated with 1-(3-dimethylaminopropyl)-3-ethylcarbodiimide hydrochloride (21.7 mg, 113.4 μmol), and the mixture was stirred at the same temperature for 12 h. The reaction was quenched with H2O, and the mixture was partitioned between 11274

DOI: 10.1021/acsomega.9b01210 ACS Omega 2019, 4, 11271−11275

ACS Omega



Article

(15) Dexter, D. L. A Theory of Sensitized Luminescence in Solids. J. Chem. Phys. 1953, 21, 836−850. (16) Monguzzi, A.; Tubino, R.; Meinardi, F. Upconversion-Induced Delayed Fluorescence in Multicomponent Organic System: Role of Dexter Energy Transfer. Phys. Rev. B: Condens. Matter Mater. Phys. 2008, 77, No. 155122. (17) Lissau, J. S.; Gardner, J. M.; Morandeira, A. Photon Upconversion on Dye-Sensitized Nanostructured ZrO2 Films. J. Phys. Chem. C 2011, 115, 23226−23232. (18) Lissau, J. S.; Nauroozi, D.; Santoni, M.-P.; Ott, S.; Gardner, J. M.; Morandeira, A. Photon Upconversion from Chemically Bound Triplet Sensitizers and Emitters on Mesoporous ZrO2: Implications for Solar Energy Conversion. J. Phys. Chem. C 2015, 119, 25792− 25806. (19) Zhou, Y.; Ayad, S.; Ruchlin, C.; Posey, V.; Hill, S. P.; Wu, Q.; Hanson, K. Examining the role of acceptor molecule structure in selfassembled bilayers: surface loading, stability, energy transfer, and upconverted emission. Phys. Chem. Chem. Phys. 2018, 20, 20513. (20) Takekuma, Y.; Ochiai, T.; Nagata, M. Immobilization of Rhodamine B Isothiocyanate on TiO2 for Light Harvesting in Zinc Phthalocyanine Dye-sensitized Solar Cells. Chem. Lett. 2018, 47, 225−227. (21) Lin, Y.-J.; Chen, J.-W.; Hsiao, P.-T.; Tung, Y.-L.; Chang, C.-C.; Chen, C.-M. Efficiency improvement of dye-sensitized solar cells by in situ fluorescence resonance energy transfer. J. Mater. Chem. A 2017, 5, 9081−9089. (22) Kondo, M.; Yukari, N.; Kaoru, F.; Morio, N.; Yoshiharu, S.; Takehisa, D.; Kouji, I.; Alastair, T. G.; Richard, J. C.; Mamoru, N. Self-Assembled Monolayer of Light-Harvesting Core Complexes from Photosynthetic Bacteria on a Gold Electrode Modified with Alkanethiols. Biomacromolecules 2007, 8, 2457−2463. (23) Dilbeck, T.; Hanson, K. Molecular Photon Upconversion Solar Cells Using Multilayer Assemblies: Progress and Prospects. J. Phys. Chem. Lett. 2018, 9, 5810−5821. (24) Monguzzi, A.; Mezyk, J.; Scotognella, F.; Tubino, R.; Meinardi, F. Upconversion-induced fluorescence in multicomponent systems: Steady-state excitation power threshold. Phys. Rev. B 2008, 78, No. 195112. (25) 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. (26) Haefele, A.; Blumhoff, J.; Khnayzer, R. S.; Castellano, F. N. Getting to the (Square) Root of the Problem: How to Make Noncoherent Pumped Upconversion Linear. J. Phys. Chem. Lett. 2012, 3, 299−303. (27) Monguzzi, A.; Tubino, R.; Hoseinkhani, S.; Campione, M.; Meinardi, F. Low power, non-coherent sensitized photon upconversion: modelling and perspectives. Phys. Chem. Chem. Phys. 2012, 14, 4322−4332. (28) Lissau, J. S.; Nauroozi, D.; Santoni, M. P.; Edvinsson, T.; Ott, S.; Gardner, J. M.; Morandeira, A. What Limits Photon Upconversion on Mesoporous Thin Films Sensitized by Solution-Phase Absorbers? J. Phys. Chem. C 2015, 119, 4550−4564.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.9b01210.



Absorption spectra of stained TiO2 films, photovoltaic measurements, and schematic representation of the power-generation processes (PDF)

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Morio Nagata: 0000-0001-6949-713X Notes

The authors declare no competing financial interest.



REFERENCES

(1) Yanai, N.; Kimizuka, N. Recent emergence of photon upconversion based on triplet energy migration in molecular assemblies. Chem. Commun. 2016, 52, 5354−5370. (2) Simon, Y. C.; Weder, C. Low-power photon upconversion through triplet−triplet annihilation in polymers. J. Mater. Chem. 2012, 22, 20817. (3) Shockley, W.; Queisser, H. J. Detailed Balance Limit of Efficiency of p-n Junction Solar Cells. J. Appl. Phys. 1961, 32, 510− 519. (4) 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. (5) Tayebjee, M. J.; McCamey, D. R.; Schmidt, T. W. Beyond Shockley-Queisser: Molecular Approaches to High-Efficiency Photovoltaics. J. Phys. Chem. Lett. 2015, 6, 2367−2378. (6) O’Regan, B.; Gratzel, M. A low-cost, high-efficiency solar cell based on dye-sensitized colloidal TiO2 films. Nature 1991, 353, 737− 740. (7) Nattestad, A.; Cheng, Y. Y.; MacQueen, R. W.; Schulze, T. F.; Thompson, F. W.; Mozer, A. J.; Fuckel, B.; Khoury, T.; Crossley, M. J.; Lips, K.; Wallace, G. G.; Schmidt, T. W. Dye-Sensitized Solar Cell with Integrated Triplet-Triplet Annihilation Upconversion System. J. Phys. Chem. Lett. 2013, 4, 2073−2078. (8) Cheng, Y. Y.; Nattestad, A.; Schulze, T. F.; MacQueen, R. W.; Fückel, B.; Lips, K.; Wallace, G. G.; Khoury, T.; Crossley, M. J.; Schmidt, T. W. Increased upconversion performance for thin film solar cells: a trimolecular composition. Chem. Sci. 2016, 7, 559−568. (9) Simpson, C.; Clarke, T. M.; MacQueen, R. W.; Cheng, Y. Y.; Trevitt, A. J.; Mozer, A. J.; Wagner, P.; Schmidt, T. W.; Nattestad, A. An intermediate band dye-sensitised solar cell using triplet-triplet annihilation. Phys. Chem. Chem. Phys. 2015, 17, 24826−24830. (10) Ahmad, S.; Liu, J.; Gong, C.; Zhao, J.; Sun, L. Photon UpConversion via Epitaxial Surface-Supported Metal−Organic Framework Thin Films with Enhanced Photocurrent. ACS Appl. Energy Mater. 2018, 1, 249−253. (11) Hill, S. P.; Banerjee, T.; Dilbeck, T.; Hanson, K. Photon upconversion and photocurrent generation via self-assembly at organic-inorganic interfaces. J. Phys. Chem. Lett. 2015, 6, 4510−4517. (12) Hill, S. P.; Dilbeck, T.; Baduell, E.; Hanson, K. Integrated Photon Upconversion Solar Cell via Molecular Self-Assembled Bilayers. ACS Energy Lett. 2016, 1, 3−8. (13) Dilbeck, T.; Hill, S. P.; Hanson, K. Harnessing molecular photon upconversion at sub-solar irradiance using dual sensitized selfassembled trilayers. J. Mater. Chem. A 2017, 5, 11652−11660. (14) Hill, S. P.; Hanson, K. Harnessing Molecular Photon Upconversion in a Solar Cell at Sub-solar Irradiance: Role of the Redox Mediator. J. Am. Chem. Soc. 2017, 139, 10988−10991. 11275

DOI: 10.1021/acsomega.9b01210 ACS Omega 2019, 4, 11271−11275