Influence of TiO2 Particle Size on Dye ... - ACS Publications

Mar 12, 2018 - Center for Nanoparticle Research, Institute for Basic Science (IBS), Seoul 08826, Republic of Korea. ‡. School of Chemical and Biolog...
1 downloads 6 Views 1MB Size
Subscriber access provided by Kaohsiung Medical University

C: Energy Conversion and Storage; Energy and Charge Transport 2

Influence of TiO Particle Size on Dye-Sensitized Solar Cells Employing Organic Sensitizer and Cobalt (III/II) Redox Electrolyte Yoon Jun Son, Jin Soo Kang, Jungjin Yoon, Jin Kim, Juwon Jeong, Jiho Kang, Myeong Jae Lee, Hyun S Park, and Yung-Eun Sung J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.7b12206 • Publication Date (Web): 12 Mar 2018 Downloaded from http://pubs.acs.org on March 12, 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 31 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

Influence of TiO2 Particle Size on Dye-Sensitized Solar Cells Employing Organic Sensitizer and Cobalt (III/II) Redox Electrolyte Yoon Jun Son,†,‡,# Jin Soo Kang,†,‡,# Jungjin Yoon,§ Jin Kim,†,‡ Juwon Jeong,†,‡ Jiho Kang,†,‡ Myeong Jae Lee,†,‡,┴ Hyun S. Park,*,|| and Yung-Eun Sung*,†,‡ †

Center for Nanoparticle Research, Institute for Basic Science (IBS), Seoul 08826, Republic of

Korea. ‡

School of Chemical and Biological Engineering, Seoul National University, Seoul 08826,

Republic of Korea. §

Global Frontier Center for Multiscale Energy Systems and Department of Mechanical and

Aerospace Engineering, Seoul National University, Seoul 08826, Republic of Korea. ||

Fuel Cell Research Center, Korea Institute of Science and Technology (KIST), Seoul 02792,

Republic of Korea. ┴

School of Advanced Materials Engineering, Kookmin University, Seoul 02707, Republic of

Korea.

#

These authors contributed equally to this work.

*E-mail: [email protected] (Y.-E. Sung), [email protected] (H. S. Park) 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

ABSTRACT Dye-sensitized solar cells (DSSCs) are highly efficient and reliable photovoltaic devices that are based on nanostructured semiconductor photoelectrodes. From their inception in 1991, colloidal TiO2 nanoparticles (NPs) with the large surface area have manifested the highest performances, and the particle size of around 20 nm is generally regarded as the optimized condition. However, though there have been reports on the influences of particles sizes in conventional DSSCs employing iodide redox electrolyte, the size effects in DSSCs with state-of-the-art cobalt electrolyte have not been investigated. In this research, systematic analyses on DSSCs with cobalt electrolyte are carried out by using various sizes of NPs (20-30 nm), and the highest performance is obtained in the case of 30 nm-sized TiO2 NPs, indicating that there is a reversed power conversion efficiency (PCE) trend when compared with those with iodide counterpart. Detailed investigations on various factors – light harvesting, charge injection, dye regeneration, and charge collection – reveal that TiO2 particles with size range of 20 nm to 30 nm do not have a notable difference in charge injection, dye regeneration, and even in light harvesting efficiency. It is experimentally verified that superior charge collection property is the sole origin of the higher performance, suggesting that charge collection should be prioritized for designing nanostructured TiO2 photoelectrodes for DSSCs employing cobalt redox electrolytes.

2 ACS Paragon Plus Environment

Page 2 of 31

Page 3 of 31 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

 INTRODUCTION On account of global energy and environmental crisis, development of alternatives for fossil fuels is one of the most important tasks that mankind is facing these days. Among various sustainable energy sources, solar energy is regarded as a promising candidate due to its large abundance and availability. There have been efforts to utilize incident sunlight by solar-toelectrical energy conversion using solar cells, and beside Si and thin film photovoltaics, dyesensitized solar cells (DSSCs) have drawn great attention due to their high performance and reliability in addition to economical advantages.1-8 In DSSCs, mesoscopic oxide semiconductor films are often used as photoelectrodes, and dye molecules adsorbed on the surface of the oxide serve as the light harvesters. This photoelectrode design enables effective photon absorption by large surface area for dye loading. TiO2 is widely used in DSSCs because of its favorable band positions, high stability, and low cost.9-12 TiO2 nanomaterials with various morphologies, such as nanoparticles (NPs),6,13 nanotubes,14-17 and nanowires18-20 have been applied to the photoanode of DSSCs; however, NPs manifested superior performance when compared with others because they provide larger surface areas for light harvesting. In DSSCs employing TiO2 NPs, the size of the particles has significant influences on photovoltaic performances of DSSCs, and there have been a number of studies to unveil the size effect in DSSCs. Yanagida et al. reported that the size of TiO2 NPs affects electron diffusion and recombination processes in TiO2 film, leading to a substantial difference in charge collection properties.21 Cao et al. revealed that larger TiO2 NPs show low electron transport resistance and high recombination resistance, which result in a superior charge collection efficiency.22 However, he insisted that increasing the size of TiO2 NPs over a certain level is detrimental to the overall performances, mainly due to the drop in available surface area for dye-loading and light harvesting. These observations suggested that the balance between charge collection and light harvesting is important for highly efficient power conversion, and in this respect, TiO2 NPs with the size of around 20 nm is often regarded as the optimum size to achieve the highest performance. Additionally, there have been investigations on the effect of particle size on electron injection from dye to TiO2 NPs. Yartsev et al. reported that charge injection is faster for larger NPs,23 but Tachiya et al. revealed that electron injection decreases from 90% to 30-70% when the size of NPs becomes larger than 50 nm when conventional Ru sensitizer is utilized.24 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

On the other hand, Saffri reported that in the case of DSSCs employing porphyrin sensitizers, larger TiO2 NPs are not only more efficient for charge injection but also beneficial for overall photovoltaic performance.25 In DSSCs employing TiO2 photoelectrodes, triiodide/iodide (I3–/I–) redox electrolytes are often used due to their favorable dye regeneration properties for conventional Ru-based dyes and large charge recombination resistance at TiO2/electrolyte interfaces.26 However, I3–/I– redox reactions occur at 0.35 V vs. NHE, which is an unnecessarily high energy level to reduce the oxidized sensitizers of which HOMO level is located at around 1.1 V vs. NHE, limiting photovoltage of DSSCs as a consequence.27 Moreover, I3–/I– have a number of drawbacks for use in DSSCs, such as significant light absorption in the visible light region and corrosive characteristic, which allows small choice for materials selection in electrodes.28 As an alternative to I3–/I– redox system, cobalt-based redox couples were introduced into DSSCs, and substantial improvements were achieved by using organic donor-π-acceptor (D-π-A) dyes or porphyrinderived sensitizers, which are compatible with cobalt redox electrolytes.5,6,29,30 Though these redox couples have the disadvantage of fast interfacial recombination kinetics with conduction band electrons at TiO2 surface, the introduction of long alkyl chain to molecular sensitizers successfully addressed the recombination problems.31 While there have been numerous studies on the size effect of TiO2 NPs on DSSCs employing I3–/I– redox electrolyte, same kind of approaches have not been reported for the case of cobalt redox couples, though they have served as state-of-the-art redox electrolyte for DSSCs. In this study, systematic investigations were performed on the size effect of TiO2 NPs on the photovoltaic performances of DSSC employing cobalt-based redox electrolyte. We used typical sensitizer (Y123 dye) and electrolyte (cobalt bipyridine ([Co(bpy)3]3+/2+) redox electrolyte) in this work, and the particles sizes were varied from around 20 to 30 nm. Unlike the case of I3–/I–based DSSCs, the highest photovoltaic performance was achieved at 30 nm- sized TiO2 NPs when [Co(bpy)3]3+/2+ redox couple was utilized in DSSCs. The effect of particle size was then studied in terms of light harvesting, charge injection, charge collection, and dye regeneration using various optical and photoelectrochemical measurements and techniques. From these systematic analyses, we could verify that increasing the particles size up to 30 nm brought the significant advance in charge collection without deteriorating other terms including light harvesting. 4 ACS Paragon Plus Environment

Page 4 of 31

Page 5 of 31 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

 EXPERIMENTAL SECTION Preparation of TiO2 Photoanodes. All of the reagents were research grade products obtained from Sigma-Aldrich (otherwise stated) and were used without further purification. For the preparation of electrodes, fluorine-doped tin oxide (FTO) glass (TEC-8, Pilkington) was ultrasonically cleaned using acetone, ethanol, and deionized (DI) water for 10 min each. Colloidal TiO2 pastes with different particle sizes (DN-GPS-18TS, DN-GPS-22TS, DN-GPS30TS, Dyenamo) were cast onto FTO via doctor-blading technique. After drying the deposited TiO2 films on FTO glass at 60 °C for 5 h, scattering layer (Ti-Nanooxide R/SP, SOLARONIX) was cast onto TiO2 films followed by heat treatment at 500 °C for 30 min in air. TiCl4-post treatment was conducted by immersing the TiO2 electrodes in a 16 mM TiCl4 aqueous solution at 70 °C for 30 min, and the TiO2 electrodes were annealed once again at 450 °C for 30 min in air. Then the TiO2 electrodes were immersed in a mixture solution of tert-butanol and acetonitrile (1:1 in vol. ratio) containing 0.1 mM of Y123 dye (DN-F05Y, Dyenamo) and 5 mM of chenodeoxycholic acid at 30 °C for 24 h. For the preparation of Ru-dye sensitized solar cells, the TiO2 films were sensitized by immersing them in an ethanol solution containing 0.5 mM N719 dye (Ru 535-bisTBA, Solaronix) at 30 °C for 24 h. Fabrication of DSSCs. Counter electrodes (CEs) for DSSCs were prepared by spin casting 300 µL of isopropanol containing 50 mM H2PtCl6 followed by heat treatment at 400 °C for 30 min in air. The [Co(bpy)3]3+/2+ redox electrolyte was composed of 0.22 M Co(bpy)3(PF6)2 (DN-C01, Dyenamo), 0.033 M Co(bpy)3(PF6)3 (DN-C02, Dyenamo), 0.1 M LiClO4, and 0.2 M 4-tertbutylpyridine in acetonitrile solvent. The I3–/I– redox electrolyte was composed of 0.6 M 1-butyl3-methylimidazolium iodide, 30 mM I2, 0.1 M guanidinium thiocyanate, and 0.5 M 4-tertbutylpyridine in a mixture of acetonitrile and valeronitrile (85:15 in vol. ratio). The prepared photoanodes and CEs were assembled using thermoplastic sealants (Surlyn, Dupont), and the redox electrolytes were injected into the assembled cell through pre-drilled holes. Physicochemical and Photoelectrochemical Characterizations. The morphologies and size of TiO2 nanoparticles were characterized by using a transmission electron microscope (TEM, Tecnai F20, FEI). The thickness of TiO2 films was measured using a scanning electron 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

Page 6 of 31

microscope (SEM, AURIGA, ZEISS). X-ray diffraction (XRD) patterns were obtained by using Rigaku D-MAX2500-PC equipped with a Cu Kα radiation source. N2 sorption analyses were performed using Micrometrics ASAP 2020, and optical measurements were carried out using spectrophotometer (Lambda 45, Perkin-Elmer). For the measurement of dye adsorption, Y123 dye-sensitized photoanodes (active area: 0.45 × 0.45 cm2) were dipped into 5 mL of a mixture solution of ethanol and toluene (1:1 vol. ratio) containing 0.1 M NaOH, and the mixture solution was stirred until the dye completely desorbed into the liquid. Photoelectrochemical Measurements. The photocurrent density–voltage (J–V) measurements were performed by using a solar simulator (XIL model 05A50KS source measure units, SERIC) and a potentiostat (Multistat 1480, Solartron). One sun illumination condition was verified by a reference Si cell which was calibrated by National Institute of Advanced Industrial Science and Technology (AIST) in Japan. Incident photon-to-current conversion efficiency (IPCE) data were collected at the wavelength range of 350 nm to 800 nm with a spectral resolution of 10 nm by using

a

quantum

efficiency

measurement

equipment

(QEX7,

PV

Measurements).

Photoluminescence (PL) spectra were obtained using a spectrofluorometer (Scientific FluoroMax-4, Horiba). Electrochemical impedance spectroscopy (EIS) measurements were performed by using Zahner Zennium Electrochemical Workstation with AC amplitude of 10 mV. For

intensity-modulated

photocurrent

spectroscopy

(IMPS)

and

intensity-modulated

photovoltage spectroscopy (IMVS) analyses, a diode laser (λ=680 nm) was used as the light source, and sinusoidally modulated light with the modulation depth of around 10% of the overall light intensity was illuminated on the photoanode side of DSSC. IMPS and IMVS measurements were carried out at four different DC light intensities (25, 50, 75, 100 W/m2) by using Zahner Zennium Electrochemical Workstation and Zahner PP211 potentiostat.

 RESULTS AND DISCUSSION Physical Characterization of TiO2 Electrodes. For the preparation of mesoscopic TiO2 photoelectrodes, colloidal TiO2 pastes containing various sizes of nanoparticles (DN-GPS-18TS, DN-GPS-22TS, DN-GPS-30TS, Dyenamo) and scattering layer (TiO2 particles of 200-400 nm; Ti-Nanooxide R/SP, Solaronix) were sequentially cast onto FTO glass substrate and were 6 ACS Paragon Plus Environment

Page 7 of 31 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

thermally sintered at 500 °C. The thickness of the TiO2 films was controlled for accurate investigations on the size effects of TiO2 NPs. Figure S1 shows the cross-sectional SEM images of the TiO2 electrodes. The thicknesses of the mesoscopic layers and scattering layers above them were around 5.8 µm 3 µm, respectively. This thickness matched well with the optimized range for the usage of cobalt bipyridine redox electrolyte.31-33 Figures 1a-c show the TEM images of the TiO2 NPs with different sizes. The particle sizes were measured based on the observations from TEM analyses, and the average diameters of three different TiO2 NPs calculated from 100 particles were 20.7 ± 4.4 nm (Figure 1d), 24.1 ± 3.1 nm (Figure 1e), and 29.6 ± 4.1 nm (Figure 1f) as summarized in Table 1. The TiO2 electrodes employing three different sizes of NPs were hereafter denoted as T1 (smallest particles), T2 (intermediate particles), and T3 (largest particles). The phase of TiO2 in the photoelectrodes was characterized by obtaining X-ray diffraction (XRD) patterns (Figure S2), and all of the signals besides those from FTO substrate (JCPDS 41-1445) were able to be assigned as the peaks originated from the anatase TiO2 (JCDPS 21-1272). In order to measure the specific surface area and pore characteristics, N2 sorption experiments were performed, and Figure S3 shows the isotherms for the TiO2 NPs in T1, T2, and T3. From the type-V behavior of the isotherm, it could be verified that the TiO2 electrodes are mesoporous. The Brunauer-Emmett-Teller (BET) surface areas were calculated as 73.5, 63.6, and 39.3 m2/g for the cases of T1, T2, and T3, respectively, and these values clearly show that there is a steep decrease in specific surface area as the particle size increases. Pore volumes were also obtained by Barret-Joyner-Halenda (BJH) methods (see Table 1), and porosities of T1, T2, and T3 were calculated using the pore volumes and the following equation: P = Vp / (ρ-1 + Vp)

(1)

where Vp is cumulative pore volume and ρ-1 is the reciprocal of the density of anatase TiO2 (0.257 cm3 g-1).22 The calculated porosities of T1, T2, and T3 were 65.19%, 65.24%, and 62.43%, respectively. It was previously reported that the porosity of TiO2 NPs, which is related to their coordination number, have significant influences on the electron transport properties.34 Mesoscopic TiO2 film with a porosity of 50-65% is known to have a coordination number of around 4, and this value is reported to be sufficiently high for effective electron transport.22 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

Given that the porosities of our TiO2 electrodes are within this range regardless of the particle size, it was expected that there are sufficient pathways for electrons to travel through the sintered TiO2 NPs. Therefore, we could exclude the effect of limitations in charge transport caused by the porosity of TiO2, and this indicates that T1, T2, and T3 are favorable for systematic studies on the size effects of TiO2 NPs in DSSCs. The detailed properties of TiO2 NPs obtained from the TEM and N2 sorption experiments are summarized in Table 1. The effective surface areas of the TiO2 electrodes comprising T1, T2, and T3 were measured by dye-adsorption experiments. Since this study was focused on the DSSCs employing [Co(bpy)3]3+/2+ electrolytes, Y123 dye that is well-suited to the cobalt bipyridine redox couple was used. Dye molecules were first chemisorbed on the surface of TiO2 followed by detachment of the sensitizers using a mixture solution of ethanol and toluene containing NaOH. The amount of the adsorbed dye molecules were characterized by measuring the absorbance at 477 nm, and the results are displayed in Table S1 and Figure S4. The loading amounts of Y123 dyes on T1, T2, and T3 were 185, 122, and 69 nmol per unit geometric area of the TiO2 film (1 cm2), respectively. As observed in the BET analyses, the effective surface area became smaller in the case of larger TiO2 NPs. Size effect of TiO2 nanoparticle on DSSC performance. The DSSCs employing T1, T2, and T3 were prepared by sandwich-like assembly of Y123 dye-sensitized photoelectrodes and Ptbased counter electrodes, by using 25 µm-thick thermoplastic sealants (Surlyn, DuPont), and acetonitrile-based [Co(bpy)3]3+/2+ redox electrolyte was then injected into the cell through predrilled holes. Figure 2a shows the J–V curves of the DSSCs measured under standard AM 1.5G illumination with a light intensity of 100 mW/cm2 (one sun condition), and the photovoltaic performance parameters are summarized in Table 2. The DSSC employing T3 manifested the highest power conversion efficiency (PCE) followed by those with T2 and T1. It was notable that the trend in PCE was opposite to the number of adsorbed dye molecules per unit active area, because larger loading of sensitizers often leads to an enhanced light harvesting. Our observation was also opposite to the trend in DSSCs employing N719 dye and I3–/I– redox electrolyte (Table S2 and Figure S5), which is the most widely studied system. According to the previous studies, around 20 nm-sized TiO2 NPs show the highest performance in N719/iodide-based DSSCs,2,4,35 and further increase or decrease in particles sizes results in poorer light harvesting (by less 8 ACS Paragon Plus Environment

Page 8 of 31

Page 9 of 31 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

amount of dye-loading) or inferior charge collection properties (due to the increase in the number of grain boundaries), respectively.15,19,36-38 From the photovoltaic parameters of the DSSCs (Table 2), it was observed that open-circuit voltage (Voc) increased with increasing particle size. In order to understand this Voc trend, we conducted the J–V measurements under dark condition (Figure S6) and EIS analyses (Figure S7). As the particle size was increased, higher overpotential for dark currents in J–V and larger charge recombination resistances in EIS were observable, implying that the trend in Voc resulted from the superior charge collection characteristic of largesized TiO2 NPs regarding the interfacial charge recombination. Nevertheless, we could deduce that the increase in PCE in the case of larger TiO2 NPs was mainly originated from the enhancement in short-circuit current density (Jsc) rather than the other parameters; Voc and fill factor (FF). The difference in Jsc was further confirmed by IPCE spectra (Figure 2b). Jsc can be calculated from the IPCE spectra and the AM 1.5G spectrum using the following equation: Jsc = ∫ IPCE(λ) e φph, AM1.5G(λ) dλ

(2)

where e is the elementary charge and φph, AM1.5G is the photon flux in AM 1.5 G illumination. Since the elementary charge and the AM 1.5G spectrum are constant, the difference in Jsc could be regarded as a consequence of different IPCE values. In DSSCs, IPCE is generally expressed in terms of four efficiency factors as the following equation: IPCE(λ) = LHE(λ) × φinj(λ) × φreg × ηcc(λ)

(3)

where LHE is light-harvesting efficiency, φinj is quantum yield for electron injection from excited sensitizer to TiO2 conduction band, φreg is regeneration efficiency of oxidized dyes and ηcc is collection efficiency for photogenerated charge carriers. In order to understand the origin of different Jsc values in DSSCs employing TiO2 NPs with different sizes, we investigated the effect of particle sizes on these four efficiency factors separately. First, the effect of particles sizes on light harvesting efficiency (LHE) was investigated. LHEs of photoanodes in DSSCs have often been studied based on transmittance; however, significant reflection occurs when scattering layer is incorporated. Therefore, LHE was 9 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 10 of 31

calculated with considerations on reflectance, based on the following equations:39,40 I0 = I(absorbed) + I(transmitted) + I(refected)

(4)

LHE = I(absorbed) / I0 = [I0 – I(transmitted) – I(refected)] / I0 = 1 – T – R

(5)

where I0 is the intensity of incident light, and T and R are the transmittance and reflectance of photoanodes, respectively. We measured transmittance and reflectance in the wavelength region of 350-800 nm and calculated LHE using equation (5), and the results are displayed in Figures 3a and b. Three photoanodes employing different sizes of TiO2 NPs had similar LHE in the measured wavelength region (Figure 3b), indicating that the light harvesting shows saturated behaviors in the cases of T1, T2, and T3. We have to state here that there was a loss of photons in reflectance measurements possibly due to the effect of scattering layer and glass/air interface. This was confirmed by the transmittance and reflectance spectra of FTO glass and unsensitized TiO2 film with scattering layer (Figure S8). Figure S8 shows that the sum of transmittance and reflectance values is lower than 100% even at wavelength ranges that are optically inert to FTO and TiO2, implying a substantial loss of photons by unavoidable reflection toward incident beam source. Nevertheless, it was clear that the LHE was similar in the cases of T1, T2, and T3, and we performed further discussion based on the obtained LHE. The spectra showing the maximum photocurrent density with regard to the wavelength was obtained using the measured LHE and the AM 1.5G spectra in Figure 3c, and the results are depicted in Figure 3d. Additionally, the maximum Jsc value was calculated by integrating the ideal spectral photocurrent densities, and similar values of around 17.7 mA/cm2 were obtained for T1, T2 and T3 (see Table S3). This result is hard to understand when only the particle size and the corresponding dye-loading amount are considered, because dye-loading was significantly lower when larger TiO2 NPs were used. However, the measured similar LHE of three photoanodes can be explained by superior absorption properties (extinction coefficient) of Y123 sensitizer compared to typical Ru-complex sensitizer (e.g. N719 dye).33,41 In addition, the presence of light scattering layer, which extends the optical path length within TiO2 film must have contributed to additional enhancement in light absorption at photoelectrodes, mitigating the particle size effect on LHE as a consequence.5,42-46 In short, it could be concluded that the effect of particles sizes on LHE was insignificant within the range of 20-30 nm. 10 ACS Paragon Plus Environment

Page 11 of 31 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

Second, in order to compare the kinetics of photo-induced electron injection for the cases of T1, T2 and T3, PL measurements were performed on dye-sensitized TiO2 photoanodes. During the PL measurements, dye molecules on TiO2 electrode absorb the incident photons from laser, produce photo-induced charge carriers, which can either be injected into TiO2 conduction band or release energy by radiative recombination.47 Excitation of dyes was carried out using 463 nm laser, as this wavelength belongs to the absorption range of Y123 dye.48 Figure 4a shows the PL spectra of Y123 dye-sensitized T1, T2, and T3, from where no notable difference was observable. Then we measured time-resolved photoluminescence (TRPL) decay measurement at the wavelength of maximum PL intensity (660 nm), to compare the electron injection kinetics when different sized TiO2 NPs are used. TRPL decay curves of T1, T2, and T3, which are presented in Figure 4b, were fitted according to the third order exponential decay model on the following equation: y(t) = y0 + A1 e-t/τ1 + A2 e-t/τ2 + A3 e-t/τ3

(6)

Among the parameters in the decay model, τ1 is the fast decay component term, which is related to the kinetics of charge injection.49 A lower τ1 value means less charge recombination in dye due to superior electron injection from dye into TiO2 electrode.50-51 The τ1 values obtained from T1, T2, and T3 were 0.57, 0.56, and 0.61 ns, respectively, and no significant difference could be revealed (detailed fitting results for other constants are summarized in Table S4). This means that the quantum yield of charge injection at the dye/TiO2 interface was almost independent of the size of TiO2 NPs. Therefore, we could conclude that TiO2 particle size does not affect charge injection properties in the range of 20-30 nm. Third, the effect of particle size on dye regeneration efficiency was investigated. In DSSCs, redox couples are responsible for the regeneration of oxidized dye molecules, and mass transfer properties of the redox couples dictate the dye regeneration kinetics when same dye and redox electrolytes are used. Cobalt-based redox couples often suffer from poor mass transfer within the mesoporous TiO2 film, mainly due to the steric hindrance originated from their large molecular size. This leads to sluggish dye regeneration and low photocurrent generation.52-53 Therefore, it was expected that any difference in φreg for the cases of T1, T2 and T3 would be solely influenced by the mass transfer of [Co(bpy)3]3+/2+ redox couples. To compare the mass transport 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

properties, we performed photocurrent transient measurements by measuring time-transient Jsc value with on and off of one sun illumination. It has been reported in the previous publications that initial value of photocurrent (measured instantly after the light absorption) decreases to a saturated value as the redox species near the surface of photoanodes are continuously consumed.54 The mass transfer properties can be compared by observing the magnitude of photocurrent decay after the initiation of light illumination. As depicted in Figure 5, DSSCs employing T1, T2, and T3 showed an instant drop of photocurrent densities for around 150 ms until reaching saturated values. However, the magnitudes of photocurrent decay for three DSSCs were very small compared to their saturated Jsc values (Jsat) as well as comparable to each other, revealing that the effect of mass transport limitation would be too trivial to bring about a significant difference in φreg. Furthermore, we quantitatively calculated regeneration efficiency, which is defined as the ratio of saturated photocurrent density to initial peak current density,55 and also obtained the Warburg impedance values from the EIS spectra measured under the forward bias of 0.7 V (Table S5). From the insignificant gaps in regeneration efficiencies and Warburg impedances, we could additionally confirm that the difference in mass transport limitations regarding the particle sizes is negligible within the 20–30 nm range. This might be ascribable to the similar porosities (60-65%) of T1, T2, and T3 as presented in Table 1, which is a sufficiently high value for efficient mass transfer of [Co(bpy)3]3+/2+ redox couples.55,56 Another notable observation in time transient photocurrents is that the initial maximum value of photocurrent density (Jmax) was quite different as the size of TiO2 NPs were varied, in the order of T3 > T2 > T1. Jmax is not associated with mass transfer properties of the redox couples, but rather dependent on other factors such as light harvesting, charge injection, and charge collection.31,55 Therefore, the effect of particles size on the dye regeneration term could be ruled out. Lastly, the effect of particle size in charge collection was taken into account. In order to compare ηcc in DSSCs employing T1, T2, and T3, electron dynamics in each TiO2 electrode were investigated by IMPS and IMVS measurements. IMPS and IMVS spectra were obtained at shortcircuit and open-circuit conditions, respectively, using 670 nm laser source with variations in light intensity; 25, 50, 75, and 100 W/m2. First, electron transport time (τd), which shows the time for an injected electron to reach current collector through semiconductor film, and electron diffusion coefficient (Dn) within TiO2 were calculated from the IMPS results based on the 12 ACS Paragon Plus Environment

Page 12 of 31

Page 13 of 31 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

following equations:57 τd = 1 / 2πfmin(IMPS)

(7)

Dn = d2 / 2.5τd

(8)

where fmin (IMPS) is the characteristic frequency at the minimum of the imaginary component, d is the thickness of TiO2 film, which is around 5.8 µm for all TiO2 films in this work. The calculated τd and Dn values of DSSCs employing T1, T2, and T3 are respectively displayed in Figures 6a and b. It was clear from the results that as the light intensity increases, τd decreases and Dn increases, indicating that electron diffusion is facilitated in TiO2 under larger photon flux. This observation matches well with those in the previous reports that at high light intensity, the traps in TiO2 are filled with electrons and trapping/detrapping process is thereby restrained, making electron transport faster as a consequence.58 Regardless of the light intensity, it was notable that DSSCs employing larger NPs had shorter τd and larger Dn than those with smaller NPs, suggesting that electron transport is faster in TiO2 electrode when the size of NPs is increased. This can be attributed to the different number of defect sites; mesoscopic film with smaller NPs, which are nanocrystals, has a larger number of grain boundaries for electrons to pass through until they arrive at the current collector.21,59 Trapping of charge carriers at the defect sites in grain boundaries retard the transport of electrons in mesoscopic TiO2 films.60 Then the electron lifetime (τn), which is the time it takes for an electron in semiconductor film to recombine with a hole/electron acceptor at semiconductor/electrolyte interface, was calculated from the IMVS results using the following equation:58 τn = 1 / 2πfmin(IMVS)

(9)

where fmin (IMVS) is the characteristic frequency at the minimum of the imaginary component. A larger τn means a slower recombination rate of photoelectrons at semiconductor/electrolyte interface, which is favorable for efficient charge collection. As depicted in Figure 6c, τn was the largest in the case of T3 followed by those of T2 and T1, which matches well with the trend observed in charge recombination resistances in EIS analyses (Figure S7). This means that smaller TiO2 NPs are more vulnerable to interfacial charge recombination of conduction band 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

Page 14 of 31

electrons. This also seems to be related to the different number of grain boundaries in TiO2 films employing TiO2 NPs of various sizes, because charge recombination occurs more often in the defect sites, which are densely located in grain boundaries. In the previous report by Cao and coworkers, DSSCs employing conventional N719 dye and iodide redox electrolyte manifested smaller charge recombination resistance at TiO2/electrolyte interface when smaller TiO2 NPs were used (unless too small for mass transfer of iodide species in the electrolyte); however, τn, which can be expressed by the product of charge recombination resistance and chemical capacitance, was larger in relatively small-sized NPs, and larger chemical capacitance in smaller TiO2 particles was suggested as the origin of the increased τn.22 Contrary to the case of iodide redox electrolyte, usage of cobalt bipyridine species leads to significant interfacial charge recombination with electrons in TiO2 due to their fast electron transfer kinetics as an outersphere redox couple.31,61 Therefore, the effect of charge recombination resistance seems to be larger than that of chemical capacitance in our system, and this difference between iodide and cobalt redox electrolytes is expected to be the origin of opposite trends in τn when different redox species are utilized. Based on the Dn and τn obtained from IMPS and IMVS analyses, respectively, electron diffusion length (Ln), which indicates the average travel distance of electrons within TiO2, was calculated based on the following equation:58 Ln = (Dn × τn)1/2

(10)

Ln determined by the competition between electron transport and recombination is directly correlated to ηcc, and a larger Ln represents superior charge collection efficiency in mesoscopic TiO2 electrodes.62,63 Figure 6d shows that Ln increases as the size of NPs becomes larger, and it could thereby be verified that the different Jsc values in DSSCs employing T1, T2, and T3 were originated from the difference in charge collection properties, which were significantly affected by the size of TiO2 NPs. Furthermore, to quantitatively compare the relationship between Jsc values and charge collection properties, charge collection efficiency was calculated using the following equation:4 ηcc = 1- (τtr/τn) 14 ACS Paragon Plus Environment

(11)

Page 15 of 31 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

From the calculated ηcc values in Table S6, it was confirmed that the ratio of change in Jsc and ηcc from T1 to T3 almost coincide with each other under four different illumination conditions, adding credibility to our result that difference in Jsc values are mainly determined by different charge collection properties of three electrodes.

 CONCLUSIONS In this work, we investigated the size effect of TiO2 NPs on DSSC employing cobalt-based redox electrolyte. Mesoscopic TiO2 photoelectrodes with 20-30 nm sized NPs were introduced to DSSCs employing [Co(bpy)3]3+/2+ redox electrolyte and Y123 dye, and reversed trend in photovoltaic performance was observed when compared with conventional systems based on I3– /I– redox couples and Ru-complex sensitizers; the highest PCE was obtained in the case of 30 nm NPs. The size effects were further examined by verifying the influences of NP sizes on four different efficiency factors – light harvesting, charge injection, charge collection, and dye regeneration – using various optical and electrochemical techniques. From these investigations, it was verified that larger TiO2 NPs with smaller surface area do not show any notable drop in LHE, but significant improvement in charge collection was observed. Given that charge injection and dye regeneration were not affected by the size of TiO2 NPs, we could conclude that the higher performance at 30 nm was ascribable to the enhanced charge collection within TiO2 by a reduced number of grain boundaries. Since our study is the first investigation on the effect of TiO2 NP sizes on DSSCs employing cobalt-based redox electrolyte, and our observation strongly demonstrate that improving charge collection is the key to enhanced photovoltaic performance, it is strongly anticipated that this work will provide significant insights into the design of mesoscopic TiO2 photoelectrodes.

 ASSOCIATED CONTENT Supporting Information. The Supporting Information is available free of charge on the ACS Publications website at DOI: Additional experimental results of SEM images, XRD characterization, BET characterization, 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

Dark J-V, and EIS analysis.

 AUTHOR INFORMATION Corresponding Authors *[email protected] *[email protected] Author Contributions #

Y. J. S. and J. S. K. contributed equally to this work.

Notes The authors declare no competing financial interests.

 ACKNOWLEDGEMENTS Y.-E. S. acknowledges that this work was supported by the Institute for Basic Science (IBS) in Republic of Korea (Project Code: IBS-R006-D1). H. S. P. acknowledges supports from the National Research Foundation of Korea (2016M3D1A1021142) funded by the Ministry of Science, ICT & Future Planning of Korea. Photoluminescence (PL) spectra and UV-Vis optical measurements were performed at the Global Frontier R&D Program on Center for Multiscale Energy System in Republic of Korea.

 REFERENCES (1) O’Regan, B.; Grätzel, M. A low-cost, high-efficiency solar cell based on dye-sensitized colloidal TiO2 films. Nature 1991, 353, 737-740. (2) Grätzel, M. Photoelectrochemical cells. Nature 2001, 414, 338–344. (3) Jin, H.-Y.; Kim, J.-Y.; Lee, J. A.; Lee, K.; Yoo, K.; Lee, D.-K.; Kim, B.; Kim, J. Y.; Kim, H.; Son, H. J. et al. Rapid sintering of TiO2 photoelectrodes using intense pulsed white light for flexible dye-sensitized solar cells. Appl. Phys. Lett 2014, 104, 143902. (4) Hagfeldt, A.; Boschloo, G.; Sun, L.; Kloo, L.; Pettersson, H. Dye-sensitized solar cells. 16 ACS Paragon Plus Environment

Page 16 of 31

Page 17 of 31 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

Chem. Rev. 2010, 110, 6595–6663. (5) Yella, A.; Lee, H.-W.; Tsao, H. N.; Yi, C.; Chandiran, A. K.; Nazeeruddin, M. K.; Diau, E. W.-G.; Yeh, C.-Y.; Zakeeruddin, S. M.; Grätzel, M. Porphyrin-sensitized solar cells with cobalt (II/III)-based redox electrolyte exceed 12 percent efficiency. Science 2011, 334, 629–634. (6) Kakiage, K.; Aoyama, Y.; Yano, T.; Oya, K.; Fujisawa, J.-i.; Hanaya, M. Highly-efficient dye-sensitized solar cells with collaborative sensitization by silyl-anchor and carboxy-anchor dyes. Chem. Commun. 2015, 51, 15894–15897. (7) Song, Q.; Zhu, Y.; Zheng, H.; Zhang, F.; Wu, M. Fabricating TiO2 film with a facile spray-coating technique for dye-sensitized solar cells. Mater. Design 2016, 98, 108–112. (8) Chung, D. Y.; Son, Y. J.; Yoo, J. M.; Kang, J. S.; Ahn, C. Y.; Park, S.; Sung, Y.-E. Coffee waste-derived hierarchical porous carbon as a highly active and durable electrocatalyst for electrochemical energy applications. ACS Appl. Mater. Interfaces 2017, 9, 41303-41313. (9) Kang, S. H.; Lim, J.-W.; Kim, H. S.; Kim, J.-Y.; Chung, Y.-H; Sung, Y.-E. Photo and electrochemical characteristics dependent on the phase ratio of nanocolumnar structured TiO2 films. Chem. Mater. 2009, 21, 2777–2788. (10) Liu, Y.; Che, R.; Chen, G.; Fan, J.; Sun, Z.; Wu, Z.; Wang, M.; Li, B.; Wei, J.; Wei, Y. et al. Radially oriented mesoporous TiO2 microspheres with single-crystal-like anatase walls for high-efficiency optoelectronic devices. Sci. Adv. 2015, 1, e1500166. (11) Kusumawati, Y.; Hosni, M.; Martoprawiro, M. A.; Cassaignon, S.; Pauporté, Th. Charge transport and recombination in TiO2 brookite-based photoelectrodes. J. Phys. Chem. C 2014, 118, 23459–23467. (12) Ito, S.; Chen, P.; Comte, P.; Nazeeruddin, M. K.; Liska, P.; Pe´chy, P.; Grätzel, M. Fabrication of screen-printing pastes from TiO2 powders for dye-sensitised solar cells Prog. Photovolt: Res. Appl. 2007, 15, 603–612. (13) Hardin, B. E.; Snaith, H. J.; McGehee, M. D. The renaissance of dye-sensitized solar cells. Nat. Photonics 2012, 6, 162-169. (14) Macák, J. M.; Tsuchiya, H.; Ghicov, A.; Schmuki, P. Dye-sensitized anodic TiO2 nanotubes. Electrochem. Commum. 2005, 7, 1133–1137. (15) Mor, G. K.; Shankar, K.; Paulose, M.; Varghese, O. K.; Grimes, C. A. Use of highlyordered TiO2 nanotube arrays in dye-sensitized solar cells. Nano Lett. 2006, 6, 215–218. (16) Kim, J.-Y.; Lee, K.-H.; Shin, J.; Park, S. H.; Kang, J. S.; Han, K. S.; Sung, M. M.; Pinna, 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

N.; Sung, Y.-E. Highly ordered and vertically oriented TiO2/Al2O3 nanotube electrodes for application in dye-sensitized solar cells. Nanotechnology 2014, 25, 504003. (17) Kang, J. S.; Choi, H.; Kim, J.; Park, H.; Kim, J.-Y.; Choi, J.-W.; Yu, S.-H.; Lee, K. J.; Kang, Y. S.; Park, S. H. et al. Multidimensional anodized titanium foam photoelectrode for efficient utilization of photons in mesoscopic solar cells. Small 2017, 13, 1701458. (18) Feng, X.; Shankar, K.; Varghese, O. K.; Paulose, M.; Latempa, T. J.; Grimes, C. A. Vertically aligned single crystal TiO2 nanowire arrays grown directly on transparent conducting oxide coated glass: synthesis details and applications. Nano Lett. 2008, 8, 3781–3786. (19) Kang, S. H.; Choi, S.-H.; Kang, M.-S.; Kim, J.-Y.; Kim, H.-S.; Hyeon, T.; Sung, Y.-E. Nanorod-based dye-sensitized solar cells with improved charge collection efficiency. Adv. Mater. 2008, 20, 54–58. (20) Adachi, M.; Murata, Y.; Takao, J.; Jiu, J.; Sakamoto, M.; Wang, F. Highly efficient dyesensitized solar cells with a titania thin-film electrode composed of a network structure of singlecrystal-like TiO2 nanowires made by the “oriented attachment” mechanism. J. Am. Chem. Soc. 2004, 126, 14943–14949. (21) Nakade, S.; Saito, Y.; Kubo, W.; Kitamura, T.; Wada, Y.; Yanagida, S. Influence of TiO2 nanoparticle size on electron diffusion and recombination in dye-sensitized TiO2 solar cells. J. Phys. Chem. B 2003, 107, 8607–8611. (22) Park, K.; Zhang, Q.; Myers, D.; Cao, G. Charge transport properties in TiO2 network with different particle sizes for dye sensitized solar cells. ACS Appl. Mater. Interfaces 2013, 5, 1044−1052. (23) Benko1, G.; Skårman, B.; Wallenberg, R.; Hagfeldt, A.; Sundstro1m, V.; Yartsev, A.P. Particle size and crystallinity dependent electron injection in fluorescein 27-sensitized TiO2 films. J. Phys. Chem. B 2003, 107, 1370-1375. (24) Du, L.; Furube, A.; Hara, K.; Katoh, R.; Tachiya, M. Mechanism of particle size effect on electron injection efficiency in ruthenium dye-sensitized TiO2 nanoparticle films, J. Phys. Chem. C 2010, 114, 8135–8143. (25) Tahay, P.; Afshani, M. B. G.; Alavi, A.; Parsa, Z.; Safari, N. Interrelationship between TiO2 nanoparticle size and kind/size of dyes in the mechanism and conversion efficiency of dye sensitized solar cells. Phys. Chem. Chem. Phys. 2017, 19, 11187-11196. 18 ACS Paragon Plus Environment

Page 18 of 31

Page 19 of 31 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

(26) Chiba, Y.; Islam, A.; Watanabe, Y.; Komiya, R.; Koide, N.; Han, L. Dye-sensitized solar cells with conversion efficiency of 11.1%. Jpn. J. Appl. Phys. 2006, 45, 24-28. (27) Ardo, S.; Meyer, G. J. Photodriven heterogeneous charge transfer with transition-metal compounds anchored to TiO2 semiconductor surfaces. Chem. Soc. Rev. 2009, 38, 115-164. (28) Boschloo, G.; Hagfeldt, A. Characteristics of the iodide/triiodide redox mediator in dyesensitized solar cells. Acc. Chem. Res. 2009, 42, 1819–1826. (29) Sapp, S. A.; Elliott, C. M.; Contado, C.; Caramori, S.; Bignozzi, C. A. Substituted polypyridine complexes of cobalt(II/III) as efficient electron-transfer mediators in dye-sensitized solar cells. J. Am. Chem. Soc. 2002, 124, 11215–11222. (30) Nusbaumer, H.; Zakeeruddin, S. M.; Moser, J. E.; Grätzel, M. An alternative efficient redox couple for the dye-sensitized solar cell system. Chem. Eur. J. 2003, 9, 3756–3763. (31) Feldt, S. M.; Gibson, E. A.; Gabrielsson, E.; Sun, L.; Boschloo, G.; Hagfeldt, A. Design of organic dyes and cobalt polypyridine redox mediators for high-efficiency dye-sensitized solar cells. J. Am. Chem. Soc. 2010, 132, 16714–16724. (32) Mathew, S.; Yella, A.; Gao, P.; Humphry-Baker, R.; Curchod, B. F. E.; Ashari-Astani, N.; Tavernelli, I.; Rothlisberger, U.; Nazeeruddin, M. K.; Grätzel, M. Dye-sensitized solar cells with 13% efficiency achieved through the molecular engineering of porphyrin sensitizers. Nat. Chem. 2014, 6, 242–247. (33) Yum, J. H.; Baranoff, E.; Kessler, F.; Moehl1, T.; Ahmad, S.; Bessho1, T.; Marchioro, A.; Ghadiri1, E.; Moser, J.-E.; Yi, C. et al. A cobalt complex redox shuttle for dye-sensitized solar cells with high open-circuit potentials. Nat. Commun. 2012, 3, 631. (34) Lagemaat, J. V. D.; Benkstein, K. D.; Frank, A. J. Relation between particle coordination number and porosity in nanoparticle films: Implications to dye-sensitized solar cells. J. Phys. Chem. B 2001, 105, 12433-12436. (35) Hagfeldt, A.; Grätzel, M. Molecular Photovoltaics. Acc. Chem. Res. 2000, 33, 269–277. (36) Jang, S.; Kang, J. S.; Lee, J.-K.; Kim, S. M.; Son, Y. J.; Lim, A.; Cho, H.; Kim, J.; Jeong, J.; Lee, G. et al. Enhanced light harvesting in mesoscopic solar cells by multilevel multiscale patterned photoelectrodes with superpositioned optical properties. Adv. Funct. Mater. 2016, 26, 6584–6592. (37) Zhu, K.; Neale, N. R.; Miedaner, A.; Frank, A. J. Enhanced charge-collection efficiencies and light scattering in dye-sensitized solar cells using oriented TiO2 nanotubes arrays. Nano Lett. 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

2007, 7, 69-74. (38) So, S.; Hwang, I.; Schmuki, P. Hierarchical DSSC structures based on “single walled” TiO2 nanotube arrays reach a back-side illumination solar light conversion efficiency of 8%. Energy Environ. Sci. 2015, 8, 849-854. (39) Ghadiri, E.; Zakeeruddin, S. M.; Hagfeldt, A.; Grätzel, M.; Moser, J.-E. Ultrafast charge separation dynamics in opaque, operational dye-sensitized solar cells revealed by femtosecond diffuse reflectance spectroscopy. Sci. Rep. 2016, 6, 24465. (40) Tachibana, Y.; Hara, K.; Sayama, K.; Arakawa, H. Quantitative analysis of lightharvesting efficiency and electron-transfer yield in ruthenium-dye-sensitized nanocrystalline TiO2 solar cells. Chem. Mater. 2002, 14, 2527-2535. (41) Ahn, S. H.; Kim, D. J.; Chi, W. S.; Kim, J. H. Hierarchical double-shell nanostructures of TiO2 nanosheets on SnO2 hollow spheres for high-efficiency, solid-state, dye-sensitized solar cells. Adv. Funct. Mater. 2014, 24, 5037–5044. (42) Kim, J.; Lee, H.; Kim, D. Y.; Seo, Y. Resonant multiple light scattering for enhanced photon harvesting in dye-sensitized solar cells. Adv. Mater. 2014, 26, 5192-5197. (43) Koo, H.-J.; Kim, Y. J.; Lee, Y. H.; Lee, W. I.; Kim, K.; Park, N.-G. Nano-embossed hollow spherical TiO2 as bifunctional material for high-efficiency dye-sensitized solar cells. Adv. Mater. 2008, 20, 195-199. (44) Zhang, Q.; Myers, D.; Lan, J.; Jenekhe, S. A.; Cao, G. Applications of light scattering in dye-sensitized solar cells. Phys. Chem. Chem. Phys. 2012, 14, 14982–14998. (45) Kang, J. S.; Lim, J.; Rho, W.-Y.; Kim, J.; Moon, D.-S.; Jeong, J.; Jung, D.; Choi, J.-W.; Lee, J.-K.; Sung, Y.-E. Wrinkled silica/titania nanoparticles with tunable interwrinkle distances for efficient utilization of photons in dye-sensitized solar cells. Sci. Rep. 2016, 6, 30829. (46) Chang, S.; Li, Q.; Xiao, X.; Wong, K. Y.; Chen, T. Enhancement of low energy sunlight harvesting in dye-sensitized solar cells using plasmonic gold nanorods. Energy Environ. Sci. 2012, 5, 9444-9448. (47) Lim, S. P.; Pandikumar, A.; Lim, H. N.; Ramaraj, R.; Huang, N. M. Boosting photovoltaic performance of dye-sensitized solar cells using silver nanoparticle-decorated N,Sco-doped-TiO2 photoanode. Sci. Rep. 2015, 5, 11922. (48) Tsao, H. N.; Yi, C.; Moehl, T.; Yum, J.-H.; Zakeeruddin, S. M.; Nazeeruddin, M. K.; Grätzel, M. Cyclopentadithiophene bridged donor–acceptor dyes achieve high power conversion 20 ACS Paragon Plus Environment

Page 20 of 31

Page 21 of 31 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

efficiencies in dye-sensitized solar cells based on the tris-cobalt bipyridine redox couple. ChemSusChem 2011, 4, 591–594. (49) Kim, M.-C.; Kim, B. J.; Son, D.-Y.; Park, N.-G.; Jung, H. S.; Choi, M. Observation of enhanced hole extraction in Br concentration gradient perovskite materials. Nano Lett. 2016, 16, 5756-5763. (50) Yang, Y.; Ri, K.; Mei, A.; Liu, L.; Hu, M.; Liu, T.; Li, X.; Han, H. The size effect of TiO2 nanoparticles on a printable mesoscopic perovskite solar cell. J. Mater. Chem. A 2015, 3, 9103-9107. (51) Heo, J. H.; Han, H. J.; Kim, D.; Ahn, T. K.; Im, S. H. Hysteresis-less inverted CH3NH3PbI3 planar perovskite hybrid solar cells with 18.1% power conversion efficiency. Energy Environ. Sci. 2015, 8, 1602-1608. (52) Nusbaumer, H.; Moser, J.-E.; Zakeeruddin, S. M.; Nazeeruddin, M. K.; Grätzel, M. CoII(dbbip)22+ complex rivals tri-iodide/iodide redox mediator in dye-sensitized photovoltaic cells. J. Phys. Chem. B 2001, 105, 10461-10464. (53) Safdari, M.; Lohse, P. W.; Häggman, L.; Frykstrand, S.; H¨ogberg, D.; Rutland, M.; Asencio, R. A.; Gardner, J.; Kloo, L.; Hagfeldt, A. et al. Investigation of cobalt redox mediators and effects of TiO2 film topology in dye-sensitized solar cells. RSC Adv. 2016, 6, 56580-56588. (54) Kim, J.-Y.; Lee, K. J.; Kang, S. H.; Shin, J.; Sung, Y.-E. Enhanced photovoltaic properties of a cobalt bipyridyl redox electrolyte in dye-sensitized solar cells employing vertically aligned TiO2 nanotube electrodes. J. Phys. Chem. C 2011, 115, 19979–19985. (55) Kim, H.-S.; Ko, S.-B.; Jang, I-H.; Park, N.-G. Improvement of mass transport of the [Co(bpy)3]II/III redox couple by controlling nanostructure of TiO2 films in dye-sensitized solar cells. Chem. Commun. 2011, 47, 12637–12639. (56) Yella, A.; Mathew, S.; Aghazada, S.; Comte, P.; Grätzel, M.; Nazeeruddin, M. K. Dyesensitized solar cells using cobalt electrolytes: the influence of porosity and pore size to achieve high-efficiency. J. Mater. Chem. C, 2017, 5, 2833-2843. (57) Nakade, S.; Kanzaki, T.; Wada, Y.; Yanagida, S. Stepped light-induced transient measurements of photocurrent and voltage in dye-sensitized solar cells: application for highly viscous electrolyte systems. Langmuir 2005, 21, 10803-10807. (58) Kru1ger, J.; Plass, R.; Grätzel, M.; Cameron, P. J.; Peter, L. M. Charge transport and back reaction in solid-state dye-sensitized solar cells: a study using intensity-modulated photovoltage 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

and photocurrent spectroscopy. J. Phys. Chem. B 2003, 107, 7536-7539. (59) Jankulovska, M.; Berger, T.; Wong, S. S.; Gómez, R.; Lana-Villarreal, T. Trap states in TiO2 films made of nanowires, nanotubes or nanoparticles: an electrochemical study. ChemPhysChem, 2012, 13, 3008–3017. (60) O'Regan, B. C.; Durrant, J. R. Kinetic and energetic paradigms for dye-sensitized solar cells: moving from the ideal to the real. Acc. Chem. Res. 2009, 42, 1799-1808. (61) Klahr, B. M.; Hamann, T. W. Performance enhancement and limitations of cobalt bipyridyl redox shuttles in dye- sensitized solar cells. J. Phys. Chem. C 2009, 113, 14040-14045. (62) Peter, L. M.; Wijayantha, K. G. U. Electron transport and back reaction in dye sensitised nanocrystalline photovoltaic cells. Electrochim. Acta 2000, 45, 4543-4551. (63) Lee, B. H.; Song, M. Y.; Jang, S.-Y.; Jo, S. M.; Kwak, S.-Y.; Kim, D.-Y. Charge transport characteristics of high efficiency dye-sensitized solar cells based on electrospun TiO2 nanorod photoelectrodes. J. Phys. Chem. C 2009, 113, 21453–21457.

22 ACS Paragon Plus Environment

Page 22 of 31

Page 23 of 31 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

Table 1. Particle sizes, surface areas, and pore characteristics of T1, T2, and T3.

Average

Surface

Cumulative

Mean

Particle Size

Area

Pore Volume

Pore Size

[nm]

[m2/g]

[cm3/g]

[nm]

T1

20.7

73.5

0.481

19.8

65.19

T2

24.1

63.6

0.482

21.1

65.24

T3

29.6

39.3

0.427

27.4

62.43

23 ACS Paragon Plus Environment

Porosity [%]

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 24 of 31

Table 2. J–V parameters obtained from the results from Figure 2a.

Voc

Jsc

FF

PCE

[V]

[mA/cm2]

[%]

[%]

T1

0.788

12.2

69.2

6.64

T2

0.789

12.5

69.6

6.89

T3

0.794

13.4

68.5

7.29

24 ACS Paragon Plus Environment

Page 25 of 31 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

Figure 1. TEM images of TiO2 nanoparticles from (a) T1, (b) T2, and (c) T3. The size distribution of the nanoparticles in the cases of (d) T1, (e) T2, and (f) T3, which were calculated based on 100 particles in the TEM images.

25 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

Figure 2. (a) J–V curves and (b) IPCE spectra of DSSCs employing T1, T2, and T3.

26 ACS Paragon Plus Environment

Page 26 of 31

Page 27 of 31 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

Figure 3. (a) Transmittance (dotted line), reflectance (dash line), and sum of transmittance and reflectance spectra (solid line) and (b) LHE spectra of Y123 dye-sensitized photoelectrodes. (c) AM 1.5 G solar spectrum. (d) Maximum photocurrent density spectra obtained using LHE and AM 1.5 G spectra.

27 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

Figure 4. (a) Steady-state PL spectra and (b) TRPL decay curves of T1, T2, and T3. PL measurements were measured using a 463 nm laser source.

28 ACS Paragon Plus Environment

Page 28 of 31

Page 29 of 31 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

Figure 5. Transient photocurrent density of DSSCs employing T1, T2, and T3 at short-circuit condition.

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

Figure 6. (a) Electron transport time, (b) electron diffusion coefficient, (c) electron lifetime, and (d) electron diffusion length in TiO2 photoelectrodes of DSSCs employing T1, T2, and T3 obtained by IMPS and IMVS measurements.

30 ACS Paragon Plus Environment

Page 30 of 31

Page 31 of 31 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

TOC Graphic

31 ACS Paragon Plus Environment