Yuzhe Zhanga,
Tomohiro Higashino*a,
Keigo Namikawaa,
W. Ryan Osterloha and
Hiroshi Imahori
*abc
aDepartment of Molecular Engineering, Graduate School of Engineering, Kyoto University, Kyoto, 615-8510, Japan. E-mail: t-higa@scl.kyoto-u.ac.jp; imahori@scl.kyoto-u.ac.jp
bInstitute for Integrated Cell-Material Sciences (WPI-iCeMS), Kyoto University, Kyoto, 606-8501, Japan
cInstitute for Liberal Arts and Sciences (ILAS), Kyoto University, Kyoto, 606-8316, Japan
First published on 21st July 2025
Porphyrin dyes have garnered significant attention as promising photosensitizers for dye-sensitized solar cells (DSSCs) due to their exceptional light-harvesting capabilities and remarkable power conversion efficiencies (PCEs) when paired with cobalt(II/III) complex-based redox shuttles. Meanwhile, copper(I/II) complexes have emerged as new generation redox shuttles, achieving impressive open-circuit voltages (VOC) exceeding 1.0 V. However, porphyrin-based DSSCs using copper(I/II) redox shuttles have struggled with low-to-moderate PCEs, primarily due to insufficient driving forces for the dye regeneration process. In this study, we introduce FL1, a novel porphyrin dye featuring a fluorene moiety with reduced electron-donating properties, designed to ensure a sufficient driving force for dye regeneration using copper(I/II) complexes. Under optimized conditions, DSSC incorporating FL1 with a copper(I/II) complex utilizing 4,4′-dimethoxy-6,6′-dimethyl-2,2′-bipyridine [Cu(2MeOby)2][TFSI]/[Cu(2MeOby)2][TFSI]2 achieved a notable PCE of 8.30% with a VOC of 0.890 V. Furthermore, our investigation into counterion effects revealed that DSSCs employing [Cu(2MeOby)2][PF6]/[Cu(2MeOby)2][PF6]2 as a redox shuttle delivered the highest PCE of 9.06% with a VOC of 0.900 V, attributed to its superior diffusion coefficient. Finally, co-sensitized DSSCs featuring FL1 and XY1B achieved an outstanding PCE of 10.9%, while retaining a high VOC of 0.945 V, setting a new benchmark efficiency for porphyrin-based DSSCs utilizing copper(I/II) redox shuttles. This breakthrough highlights the immense potential of further refining porphyrin dyes and copper(I/II) redox shuttles through energy-level engineering to optimize the driving force for dye regeneration and propel advancements in DSSC technology.
Given the fundamental working principle of DSSCs, the sensitizer is a key player in facilitating light-harvesting and electron transfer (ET) processes at the dye/TiO2/electrolyte interface. To push the boundaries of DSSC efficiency, the development of sensitizers with strong light-harvesting capabilities across the visible and near-infrared spectrum is highly desirable.9–13 Moreover, the strategic incorporation of bulky substituents is crucial for suppressing charge recombination (CR) processes, thereby optimizing device performance. Porphyrin dyes have emerged as highly promising photosensitizers for DSSCs, thanks to their exceptional light-harvesting properties, adaptable electronic structures achieved through molecular engineering, and outstanding stability compared to conventional dyes.14–32 Notably, single push-pull-type porphyrin dyes designed with effective steric hindrance from bulky substituents have demonstrated impressive PCEs, reaching up to 13% when paired with tris(bipyridyl)CoII/III ([Co(bpy)3]2+/3+) as a redox shuttle. This progress underscores the immense potential of porphyrin-based sensitizers in advancing high-performance DSSCs.
Meanwhile, the redox shuttle serves as a crucial component in DSSCs, as the open-circuit voltage (VOC) is primarily determined by the potential difference between the quasi-Fermi level of TiO2 and the redox potential (Eredox) of the redox shuttle.33–35 Over the past few decades, iodide/triiodide (I−/I3−) and [Co(bpy)3]2+/3+ have been the most widely utilized redox shuttles. However, DSSCs incorporating these traditional shuttles have faced a VOC limitation, with values restricted to below 1.0 V. To address this constraint, copper(I/II) complexes have emerged as new-generation redox shuttles, offering a more positive Eredox and significantly enhancing performance. DSSCs employing suitable complementary dyes and [Cu(tmby)2]+/2+ (tmby = 4,4′-6,6′-dimethyl-2,2′-bipyridine) have achieved PCEs of up to 15%, coupled with a higher VOC exceeding 1.0 V.8,36–44 This progress emphasizes the transformative potential of copper-based redox shuttles in propelling DSSC efficiency to new heights.
In this context, DSSCs incorporating highly efficient porphyrin dyes alongside copper redox shuttles present a promising pathway for further enhancing PCE. However, despite their potential, the PCEs of porphyrin-based DSSCs using copper(I/II) complex-based redox shuttles remain below 7%, even in the case of LG4 (Fig. 1),45–47 falling significantly behind other organic dye-based DSSCs in performance. These low-to-moderate PCEs are primarily attributed to the challenge of achieving a sufficient driving force for effective dye regeneration (−ΔGreg ≥ 0.2 eV). Conventional push–pull-type porphyrin dyes typically exhibit relatively low oxidation potentials (Eox) of 0.7–0.8 V vs. NHE, limiting their compatibility with copper(I/II) complexes. In our previous studies, we demonstrated that increasing the −ΔGreg value through energy-level engineering is an effective strategy for boosting PCEs in porphyrin-based DSSCs.46 Thus, porphyrin dyes with elevated Eox values (≥0.9 V vs. NHE) would be well-suited for DSSCs utilizing copper(I/II) complexes. Additionally, ligand modification of copper(I/II) complexes emerges as another promising approach for optimizing −ΔGreg values (Fig. 2).47,48 Incorporating electron-donating substituents into the bipyridine framework shifts the Eredox value negatively, a favorable adjustment for improving porphyrin-based DSSC performance. These advancements collectively pave the way for further refining porphyrin-copper systems, driving DSSCs toward higher efficiencies.
![]() | ||
Fig. 2 Energy-level engineering of both dyes and redox shuttles for achieving an efficient driving force for dye regeneration (−ΔGreg) toward enhanced cell performance. |
Herein, we present FL1, a novel porphyrin dye incorporating a 9,9-dioctylfluorene moiety, along with its photovoltaic performance in DSSCs employing copper(I/II) complex-based redox shuttles (Fig. 1). The reduced electron-donating nature of the 9,9-dioctylfluorene moiety, compared to the phenothiazine moiety in previously reported LG4,45 was expected to enhance the −ΔGreg value (Fig. 2), while the two bulky octyl chains at the 9-position of the fluorene skeleton were anticipated to provide a strong blocking effect against the redox shuttle. Furthermore, we aimed to further increase the −ΔGreg value by pairing FL1 with a copper(I/II) complex incorporating 4,4′-dimethoxy-6,6′-dimethyl-2,2′-bipyridine (2MeOby), [Cu(2MeOby)2]+/2+, featuring electron-donating methoxy groups (Fig. 1). Additionally, we also investigated the effect of counterions because the counterion effect of the ruthenium dyes (i.e. N719) has been reported.49 Actually, counterions play a critical role in the diffusion behavior of copper(I/II) complexes in organic solvents,50 significantly influencing the PCE of DSSCs, as high diffusion coefficients are essential for efficient charge transport within the electrolyte.50,51 To optimize charge transport, we examined three counterions, trifluoromethane-sulfonimide ([TFSI]−), tetrafluoroborate ([BF4]−), and hexafluorophosphate ([PF6]−). Through rational molecular engineering and a systematic exploration of the counterion effect, we successfully enhanced the PCE, with FL1-based DSSCs achieving an impressive PCE of 9.06%. Moreover, co-sensitization with XY1B further pushed the PCE beyond 10% while retaining a high VOC of 0.945 V, marking the first instance of porphyrin-based DSSCs with copper(I/II) complex-based redox shuttles surpassing this milestone.
Additionally, copper complexes incorporating [TFSI]− as a counterion were synthesized (Scheme S2†). The reaction of CuI with the ligand 2MeOby,53 followed by counterion exchange using LiTFSI, produced the copper(I) complex [Cu(2MeOby)2][TFSI]. In contrast, the direct reaction of Cu(TFSI)2 with 2MeOby resulted in the copper(II) complex [Cu(2MeOby)2][TFSI]2 without counterion exchange. Copper(I) complexes [Cu(2MeOby)2][X] (X = BF4 or PF6) were obtained by reacting [Cu(CH3CN)4][BF4] or [Cu(CH3CN)4][PF6] with 2MeOby (Scheme S3†). Meanwhile, the copper(II) complexes [Cu(2MeOby)2][X]2 (X = BF4 or PF6) were prepared via a two-step process: first, CuSO4·5H2O was reacted with 2MeOby, followed by counterion exchange with the respective ammonium salts (NH4BF4 or NH4PF6). Characterization of these compounds, including their copper complexes, was performed using 1H and 13C NMR, FT-IR spectroscopy, and high-resolution mass spectrometry (Fig. S1–S9†). Notably, the as-prepared copper(II) complexes contained varying proportions of their copper(I) counterparts due to unavoidable auto-reduction under ambient conditions, as Cl− ions were absent.54,55 Since Kitagawa et al. reported that the auto-reduction of copper(II) complexes occurred in ethanol,54 methanol in the reaction solvent may be a reductant for auto-reduction. On the other hand, we obtained pure copper(II) complexes by the bulk electrolysis of the corresponding copper(I) complexes, following established literature protocols.56,57 The progress of electrochemical oxidation was monitored using the current, and the electrolysis was considered complete when the relative current approached zero, indicating that the copper(I) complexes had been entirely converted to their copper(II) counterparts (Fig. S16 and S17†). Since the as-prepared copper(II) complexes contain no other impurities except for the corresponding copper(I) complexes, the actual compositions of the as-prepared copper(II) complexes were determined as [Cu(2MeOby)2][X]n (n = 1.6–1.9; X = TFSI, BF4, or PF6) by absorbance at 748 nm in acetonitrile (Fig. S19†), using the absorption coefficient of pure copper(II) complexes (vide infra). It is noteworthy that the actual compositions of the as-prepared copper(II) complexes in our laboratory were also determined as [Cu(tmby)2][TFSI]1.8 in a similar manner.
Single crystals suitable for X-ray diffraction analysis were successfully grown via vapor diffusion of Et2O into acetonitrile solutions of [Cu(2MeOby)2][TFSI] or [Cu(2MeOby)2][BF4] and by vapor diffusion of n-butanol into acetonitrile solutions of [Cu(2MeOby)2][PF6]. The resulting crystal structures (Fig. S10†) revealed Cu–N bond distances of 2.01–2.07 Å, confirming that the counterions had negligible influence on the geometry of the copper(I) complexes. The crystal data of the copper(I) complexes are also summarized in Table S1.†
![]() | ||
Fig. 3 UV/vis/NIR absorption spectra of FL1 (red), LG4 (blue), and XY1B18 (black) in THF. |
Dye | λabsa/nm (ε/103 M−1 cm−1) | λemb/nm | Eoxc/V | E0–0/eV | d/V | −ΔGinje/eV | −ΔGregf/eV | −ΔGregg/eV |
---|---|---|---|---|---|---|---|---|
a Wavelengths for Soret and Q-band maxima in THF.b Wavelengths for emission maxima in THF by excitation at Soret band maxima.c Determined by cyclic voltammetry (CV) measurements of the porphyrin-adsorbed TiO2 films (vs. the normal hydrogen electrode (NHE)).d Determined by adding the E0–0 value to the Eox (vs. NHE).e Driving forces for electron injection from the porphyrin singlet excited state (1ZnP*) to the conduction band (CB) of TiO2 (−0.5 V vs. NHE).f Driving forces for the regeneration of oxidized porphyrin dyes (ZnP˙+) by the [Cu(tmby)2]+/2+ redox couple (+0.88 V vs. NHE).g Driving forces for the regeneration of oxidized porphyrin dyes (ZnP˙+) by the [Cu(2MeOby)2]+/2+ redox couple (+0.82 V vs. NHE). | ||||||||
FL1 | 460 (226), 668 (72) | 681 | 1.07 | 1.90 | −0.83 | 0.33 | 0.19 | 0.25 |
LG4 | 460 (206), 673 (69) | 688 | 0.98 | 1.87 | −0.89 | 0.39 | 0.10 | 0.16 |
The electrochemical properties of the porphyrin dyes were evaluated using cyclic voltammetry (CV) and differential pulse voltammetry (DPV) in THF with 0.1 M tetrabutylammonium hexafluorophosphate (n-Bu4NPF6) as a supporting electrolyte. The first oxidation potential (Eox) of FL1 was measured at +0.93 V vs. NHE (Fig. S13†), which is more positive than that of LG4 (+0.86 V vs. NHE), reinforcing the fluorene moiety's weaker electron-donating ability. To further assess the ET processes at the dye/TiO2/electrolyte interface, CV measurements were performed for dye-adsorbed TiO2 films (FL1/TiO2 and LG4/TiO2) as working electrodes. The Eox values of FL1/TiO2 and LG4/TiO2 were determined to be +1.07 V and +0.98 V vs. NHE, respectively (Fig. S14† and Table 1). Considering the redox potential (Eredox) of the copper(I/II) complexes [Cu(tmby)2][TFSI]/[Cu(tmby)2][TFSI]2 (+0.88 V vs. NHE, Fig. S20a†), the driving forces for dye regeneration (−ΔGreg) of the porphyrin radical cation (ZnP˙+) were calculated to be 0.19 eV for FL1 and 0.10 eV for LG4. Since previous studies suggest that −ΔGreg should be comparable to or larger than 0.20 eV for efficient dye regeneration via copper(I/II) redox shuttles,36 the larger −ΔGreg value for FL1 than for LG4 is expected to enhance dye regeneration efficiency, contributing to improved PCE. Additionally, the driving forces for electron injection (−ΔGinj) from the porphyrin excited singlet state (1ZnP*) to the conduction band (CB) of TiO2 (−0.5 V vs. NHE) were estimated using their excited-state oxidation potentials calculated using their oxidation potentials and optical HOMO–LUMO gaps. Both FL1 and LG4 possess sufficient −ΔGinj values (≥0.30 eV) to facilitate efficient electron injection into the CB of TiO2.
To obtain insight into the ground state geometries and electronic structures of the porphyrin dyes, we performed density functional theory (DFT) calculations on model porphyrins at the B3LYP/6-31G(d) level (Fig. 4). Both FL1 and LG4 adopt a highly planar structure, facilitating an extended π-system for effective charge delocalization. The two alkyl chains at the 9-position of the fluorene moiety in FL1 serve a dual purpose: providing a blocking effect and suppressing dye aggregation due to the enhanced steric hindrance (Fig. S15†). The highest occupied molecular orbitals (HOMOs) display significant orbital distribution on the porphyrin core and donor moieties, while the orbital distributions of the lowest unoccupied molecular orbitals (LUMOs) are mainly localized on the electron-withdrawing cyanoacrylic acid anchoring groups and porphyrin core. The HOMO level of FL1 (−4.81 eV) is slightly more negative than that of LG4 (−4.74 eV), agreeing with the weaker electron-donating ability of the fluorene moiety than the phenothiazine moiety. The calculated HOMO–LUMO energy gaps were determined to be 2.01 eV for LG4 and 2.06 eV for FL1, aligning well with the optical HOMO–LUMO gaps and electrochemical properties of these porphyrin dyes.
λabsa/nm (ε/M−1 cm−1) | Db/10−5 cm2 s−1 | |
---|---|---|
a Wavelength of the absorption peaks in acetonitrile.b Diffusion coefficients determined via chronoamperometry. | ||
[Cu(2MeOby)2][TFSI] | 334 (4700), 445 (4500) | 1.23 |
[Cu(2MeOby)2][TFSI]2 | 385 (1800), 748 (110) | — |
[Cu(2MeOby)2][BF4] | 334 (4500), 445 (4500) | 1.30 |
[Cu(2MeOby)2][BF4]2 | 385 (1700), 748 (130) | — |
[Cu(2MeOby)2][PF6] | 334 (4700), 445 (4600) | 2.41 |
[Cu(2MeOby)2][PF6]2 | 385 (1700), 748 (110) | — |
To further explore the electrochemical behavior, we conducted CV measurements for both copper(I) and copper(II) complexes in acetonitrile (Fig. S20†). The copper(I) complex [Cu(2MeOby)2][TFSI] displayed a reversible oxidation peak at +0.82 V vs. NHE. Notably, the copper(II) complex [Cu(2MeOby)2][TFSI]2 exhibited a reversible reduction peak at precisely the same potential, +0.82 V, identical to the oxidation potential of its copper(I) counterpart. Consequently, the redox potential of [Cu(2MeOby)2][TFSI]/[Cu(2MeOby)2][TFSI]2 was determined to be +0.82 V vs. NHE, which is lower than that of [Cu(tmby)2][TFSI]/[Cu(tmby)2][TFSI]2 attributable to the electron-rich nature of the 2MeOby ligand with methoxy substitutions. Furthermore, the redox potentials of [Cu(2MeOby)2][BF4]/[Cu(2MeOby)2][BF4]2 and [Cu(2MeOby)2][PF6]/[Cu(2MeOby)2][PF6]2 were similarly determined to be +0.82 V vs. NHE, reinforcing the notion that counterions have negligible impact on the redox behavior of the copper(I/II) complexes. Considering the Eox values of FL1/TiO2 (+1.07 V vs. NHE) and LG4/TiO2 (+0.98 V vs. NHE), the −ΔGreg values for FL1 and LG4 with [Cu(2MeOby)2]+/2+ were estimated to be 0.25 and 0.16 eV, respectively. The increase in −ΔGreg values is expected to accelerate the dye regeneration process, which, in turn, should enhance the PCE employing [Cu(2MeOby)2]+/2+ as a redox shuttle.
Dye | Electrolyte | JIPCEb/mA cm−2 | JSC/mA cm−2 | VOC/V | ff | PCE/% |
---|---|---|---|---|---|---|
a Photovoltaic parameters derived from the highest PCEs. The values in parentheses denote average values from three or four independent experiments. Error bars represent the standard error of the mean.b The JIPCE values were obtained by integrating the IPCE plots, and the relative discrepancies from the corresponding JSC values obtained from the J–V characteristics are provided in the parentheses. | ||||||
LG4 | 0.2 M [Cu(tmby)2][TFSI] | 9.50 (0.7%) | 9.57 (9.35 ± 0.4) | 0.842 (0.844 ± 0.011) | 0.716 (0.716 ± 0.003) | 5.77 (5.65 ± 0.2) |
0.05 M [Cu(tmby)2][TFSI]2 | ||||||
LG4 | 0.2 M [Cu(2MeOby)2][TFSI] | 10.0 (0.7%) | 10.07 (9.96 ± 0.2) | 0.821 (0.819 ± 0.002) | 0.762 (0.763 ± 0.004) | 6.30 (6.22 ± 0.1) |
0.05 M [Cu(2MeOby)2][TFSI]2 | ||||||
FL1 | 0.2 M [Cu(tmby)2][TFSI] | 10.8 (3.5%) | 11.2 (10.7 ± 0.6) | 0.902 (0.895 ± 0.013) | 0.767 (0.757 ± 0.019) | 7.75 (7.28 ± 0.5) |
0.05 M [Cu(tmby)2][TFSI]2 | ||||||
FL1 | 0.2 M [Cu(2MeOby)2][TFSI] | 12.3 (2.4%) | 12.6 (12.1 ± 0.4) | 0.887 (0.893 ± 0.01) | 0.730 (0.719 ± 0.009) | 8.15 (7.78 ± 0.3) |
0.05 M [Cu(2MeOby)2][TFSI]2 | ||||||
FL1 | 0.1 M [Cu(tmby)2][TFSI] | 10.2 (1.9%) | 10.4 (10.1 ± 0.4) | 0.935 (0.919 ± 0.024) | 0.720 (0.735 ± 0.03) | 7.00 (6.84 ± 0.2) |
0.025 M [Cu(tmby)2][TFSI]2 | ||||||
FL1 | 0.1 M [Cu(2MeOby)2][TFSI] | 11.6 (6.5%) | 12.4 (12.3 ± 0.5) | 0.890 (0.895 ± 0.006) | 0.752 (0.748 ± 0.008) | 8.30 (8.20 ± 0.2) |
0.025 M [Cu(2MeOby)2][TFSI]2 | ||||||
FL1 | 0.1 M [Cu(2MeOby)2][BF4] | 12.0 (4.7%) | 12.6 (12.4 ± 0.6) | 0.913 (0.916 ± 0.003) | 0.777 (0.762 ± 0.02) | 8.94 (8.63 ± 0.3) |
0.025 M [Cu(2MeOby)2][BF4]2 | ||||||
FL1 | 0.1 M [Cu(2MeOby)2][PF6] | 13.0 (1.5%) | 13.2 (12.7 ± 0.4) | 0.900 (0.910 ± 0.01) | 0.763 (0.762 ± 0.006) | 9.06 (8.80 ± 0.2) |
0.025 M [Cu(2MeOby)2][PF6]2 | ||||||
XY1B | 0.1 M [Cu(2MeOby)2][PF6] | 14.1 (2.1%) | 14.4 (14.6 ± 0.2) | 0.961 (0.963 ± 0.002) | 0.752 (0.73 ± 0.02) | 10.4 (10.3 ± 0.1) |
0.025 M [Cu(2MeOby)2][PF6]2 | ||||||
FL1 + XY1B | 0.1 M [Cu(2MeOby)2][PF6] | 14.4 (5.3%) | 15.2 (15.3 ± 0.1) | 0.945 (0.940 ± 0.01) | 0.756 (0.752 ± 0.005) | 10.9 (10.8 ± 0.1) |
0.025 M [Cu(2MeOby)2][PF6]2 |
The photocurrent–voltage characteristics of DSSCs under optimized conditions are presented in Fig. 5, with the corresponding photovoltaic data listed in Table 3. The superior PCEs observed for FL1 compared to LG4 primarily stem from its higher JSC and VOC values. The enhanced JSC values are attributed to improved incident photon-to-current efficiencies (IPCEs), as illustrated in Fig. 6b. The IPCE values follow the trend: LG4 with [Cu(tmby)2][TFSI]/[Cu(tmby)2][TFSI]2 < LG4 with [Cu(2MeOby)2][TFSI]/[Cu(2MeOby)2][TFSI]2 < FL1 with [Cu(tmby)2][TFSI]2/[Cu(tmby)2][TFSI]2 < FL1 with [Cu(2MeOby)2][TFSI]/[Cu(2MeOby)2][TFSI]2. Indeed, the JIPCE values derived from photocurrent action spectra are compatible with the measured JSC values. To further elucidate the differences in IPCE, we analyzed light-harvesting efficiency (LHE), electron injection efficiency (ϕinj), and charge collection efficiency (ηcol), based on the equation: IPCE = LHE × ϕinj × ηcol. The absorption spectra of dye-adsorbed TiO2 films, recorded without a light-scattering layer, revealed nearly identical absorption profiles for FL1 and LG4 (Fig. 6a). Taking into account the excellent absorption coefficients of the porphyrin dyes and the presence of a light-scattering layer in the actual device, the LHE values approach 100% for both dyes. Although self-quenching of the excited state due to dye aggregation on TiO2 could potentially reduce ϕinj, co-adsorption of CDCA effectively mitigates dye aggregation effects. Consequently, the higher IPCE values observed for FL1 compared to LG4 are primarily attributed to its superior ηcol values.
Given that the ηcol value correlates with the dye regeneration efficiency (φreg), we conducted microsecond time-resolved transient absorption (TA) measurements on the porphyrin-sensitized TiO2 films. The decay profiles of the porphyrin radical cation (ZnP˙+) at 800 nm were analyzed, as the electrochemical oxidation of the porphyrin dyes distinctly revealed the characteristic absorption at 700–850 nm (Fig. S24†). The φreg values were derived from the lifetimes of ZnP˙+ in the absence and presence of copper(I/II) redox shuttles (Fig. S25 and Table S2†). The estimated φreg values follow the trend: LG4 with [Cu(tmby)2][TFSI]/[Cu(tmby)2][TFSI]2 (58%) < LG4 with [Cu(2MeOby)2][TFSI]/[Cu(2MeOby)2][TFSI]2 (66%) < FL1 with [Cu(tmby)2][TFSI]/[Cu(tmby)2][TFSI]2 (71%) < FL1 with [Cu(2MeOby)2][TFSI]/[Cu(2MeOby)2][TFSI]2 (84%). This order aligns well with the trend observed for the −ΔGreg values. The highest φreg value of 84% for FL1 with [Cu(2MeOby)2][TFSI]/[Cu(2MeOby)2][TFSI]2 underscores the critical role of achieving a −ΔGreg value of 0.2 eV for an efficient dye regeneration process, even in DSSCs employing copper(I/II) redox shuttles.
To further assess CR between electrons in the CB of TiO2 and the redox shuttle, we examined current–voltage characteristics under dark conditions (Fig. S26†). The more positive voltage onsets observed for FL1 compared to LG4 indicate an effective suppression of the CR process, likely due to an enhanced blocking effect. Dye loading amounts (Γ) for FL1 and LG4, under the optimized sensitization conditions, were determined to be 9.5 × 10−11 and 1.1 × 10−10 mol cm−2, respectively. These values were obtained by measuring the absorbance of porphyrin dyes dissolved from the dye-sensitized TiO2 films into a THF/H2O (v/v = 1/1) solution with 0.1 M NaOH. The smaller Γ value for FL1 compared to LG4 suggests that the increased steric hindrance of FL1 plays a critical role in enhancing the blocking effect and suppressing dye aggregation on TiO2, primarily due to the presence of alkyl chains on the fluorene moiety. Additionally, the ET process at the dye/TiO2/electrolyte interface was investigated using electrical impedance spectroscopy (EIS) under AM 1.5 illumination under open-circuit conditions (Fig. S27†). The charge transfer resistance (Rp) at the TiO2/dye/electrolyte interface followed this trend: (32.2 Ω for LG4 with [Cu(2MeOby)2][TFSI]/[Cu(2MeOby)2][TFSI]2 < 36.8 Ω for LG4 with [Cu(tmby)2][TFSI]/[Cu(tmby)2][TFSI]2 < 42.5 Ω for FL1 with [Cu(2MeOby)2][TFSI]/[Cu(2MeOby)2][TFSI]2 < 45.0 Ω for FL1 with [Cu(tmby)2][TFSI]/[Cu(tmby)2][TFSI]2). This order is consistent with that of the observed VOC values, reinforcing the correlation between increased Rp values for FL1 and higher VOC values in the DSSC utilizing FL1 compared to LG4.
To demonstrate the potential of our new porphyrin dye FL1 for DSSCs using copper(I/II) complex-based redox shuttles, we also measured photovoltaic performances of the DSSCs with FL1 and LG4 using the standard I−/I3− redox shuttle (Fig. S28, S29 and Table S3†). Although the DSSCs using the I−/I3− redox shuttle exhibited higher JSC values than those using [Cu(L)2][TFSI]/[Cu(L)2][TFSI]2 (L = tmby or 2MeOby) due to their larger driving forces for dye regeneration, the less positive redox potential of I−/I3− (+0.40 V vs. NHE) resulted in significantly lower VOC values. Overall, the PCEs of the DSSCs using the I−/I3− redox shuttle (5.84% for LG4 and 6.74% for FL1) are lower than those using the [Cu(2MeOby)2][TFSI]/[Cu(2MeOby)2][TFSI]2 redox shuttle. These results clearly corroborate that the combination of the porphyrin dye and the copper(I/II) complex-based redox shuttle with a sufficient −ΔGreg value is an effective means to achieve further enhancement in PCE.
Since counterions significantly affect diffusion properties (vide supra), they are also expected to impact the photovoltaic performance of DSSCs utilizing copper(I/II) complexes. To investigate this effect, we examined different counterions at optimal concentrations of 0.1 M [Cu(2MeOby)2][X] and 0.025 M [Cu(2MeOby)2][X]2 (X = TFSI, BF4, or PF6) (Fig. 7a and Table 3). DSSCs incorporating [Cu(2MeOby)2][X]/[Cu(2MeOby)2][X]2 (X = BF4 or PF6) attained higher PCEs (8.94% for [Cu(2MeOby)2][BF4]/[Cu(2MeOby)2][BF4]2 and 9.06% for [Cu(2MeOby)2][PF6]/[Cu(2MeOby)2][PF6]2) compared to [Cu(2MeOby)2][TFSI]/[Cu(2MeOby)2][TFSI]2. Remarkably, the PCE of the DSSC employing FL1 with [Cu(2MeOby)2][PF6]/[Cu(2MeOby)2][PF6]2 is the highest reported among porphyrin-based DSSCs with copper(I/II) redox shuttles. The superior PCE observed for [Cu(2MeOby)2][PF6]/[Cu(2MeOby)2][PF6]2 can be attributed to its significantly improved JSC value, in line with its higher IPCE values relative to those of [Cu(2MeOby)2][BF4]/[Cu(2MeOby)2][BF4]2 and [Cu(2MeOby)2][TFSI]/[Cu(2MeOby)2][TFSI]2 (Fig. 7b). To further elucidate this enhancement, we estimated φreg values through time-resolved TA measurements conducted in the presence of copper(I/II) complexes at low concentrations (0.1 M CuI and 0.025 M CuII) (Fig. S31 and Table S4†). The φreg values followed the trend: [Cu(2MeOby)2][TFSI]/[Cu(2MeOby)2][TFSI]2 (83%) < [Cu(2MeOby)2][BF4]/[Cu(2MeOby)2][BF4]2 (84%) < [Cu(2MeOby)2][PF6]/[Cu(2MeOby)2][PF6]2 (87%). Since counterions do not directly impact the Eredox value, the highest diffusion coefficient of [Cu(2MeOby)2][PF6] is likely responsible for the superior φreg, IPCE, and JSC values, ultimately leading to the highest PCE. Even though the maximum IPCE value of the DSSC using [Cu(2MeOby)2][PF6]/[Cu(2MeOby)2][PF6]2 exceeded 60% due to the highest φreg value, there remains room for further improvement in the JSC value. Considering that the enhanced blocking effect can suppress the unfavorable CR process,17,18 the introduction of bulky substituents is expected to increase the ηcol value, leading to a higher JSC value.
We also measured current–voltage characteristics under dark conditions (Fig. S32†) to evaluate the impact of counterions on VOC values. The onsets of dark current for [Cu(2MeOby)2][BF4]/[Cu(2MeOby)2][BF4]2 and [Cu(2MeOby)2][PF6]/[Cu(2MeOby)2][PF6]2 were more positive than those for [Cu(2MeOby)2][TFSI]/[Cu(2MeOby)2][TFSI]2. Thus, the slightly improved VOC values observed for [Cu(2MeOby)2][BF4]/[Cu(2MeOby)2][BF4]2 and [Cu(2MeOby)2][PF6]/[Cu(2MeOby)2][PF6]2 can be ascribed to more effective suppression of CR between electrons in the CB of TiO2 and the redox shuttle. Although the precise mechanism remains unclear at this stage, these findings suggest that counterions significantly influence interfacial behavior at the TiO2/dye/electrolyte interfaces, with the [TFSI]− anion potentially accelerating undesirable CR processes.
Footnote |
† Electronic supplementary information (ESI) available: Experimental section, synthetic details, optical and electrochemical properties of porphyrin dyes and copper complexes, photovoltaic properties, HR-MS spectra, and NMR spectra. CCDC 2451025–2451027. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d5sc03537f |
This journal is © The Royal Society of Chemistry 2025 |