Open Access Article
Rinki Deka,
Bikash Kalita and
Dhruba Jyoti Kalita
*
Department of Chemistry, Gauhati University, Guwahati-781014, India. E-mail: dhrubajyoti.kalita@gauhati.ac.in
First published on 17th December 2025
Harvesting solar energy from sunlight to generate electricity is considered one of the most important technologies to address future sustainability. Dye-sensitized solar cells (DSSCs) have attracted tremendous interest and attention over the past two decades due to their potential advantages for being implemented on large areas and using light-weight flexible substrates. The electronic, optical, and photovoltaic properties of dyes are pivotal for efficient solar energy conversion, and these properties can be finely tuned by structural modifications. In this study, we have designed a series of D–D–π–A architectured dyes, employing coumarin–thiophene–cyanoacrylic acid as the D–π–A core with the integration of eight efficient auxiliary donor units. Density functional theory (DFT) and time-dependent DFT (TD-DFT) computations have been employed to elucidate the electronic structure and optical characteristics of the dyes. Through DFT and TD-DFT simulations, we investigate the impact of various auxiliary donors on geometrical configurations, electronic structures, and optical properties. Our findings reveal that the incorporation of double donors not only enhances electron-donating capabilities but also impedes aggregation between dye molecules, thereby preventing recombination of injected electrons with the I−/I3− in the electrolyte at TiO2 semiconducting surfaces. This study underscores the effectiveness of incorporating auxiliary donor groups into organic dyes as a promising strategy for the development of high-performance, metal-free organic dyes tailored for photovoltaic applications.
DSSCs are semiconductor photovoltaic devices that directly transform solar radiation into electric current.21 A basic DSSC consists of a molecular dye, a mesoporous wide-band-gap semiconductor oxide, a redox couple electrolyte and a collector/counter electrode. The dye molecule (photosensitizer) absorbs sunlight and injects its excited electrons into the conduction band of the semiconductor and becomes oxidized. The redox electrolyte acts as an intermediate to transfer holes from the dye to the counter electrode for regeneration of the dye.22 Dyes with good light-harvesting ability and appropriate energy levels may lead to adequate photoelectric conversion efficiency.23–25 The most stable and efficient dyes for DSSCs are metal–organic dyes, such as those containing ruthenium and osmium, but they are costly, difficult to synthesize, and hazardous by nature. Owing to the drawbacks of metal organic dyes, novel donor–π–acceptor (D–π–A)-type metal-free organic dyes have been described to enhance the efficiency of DSSCs.26 Thus metal-free organic dyes have been synthesized and used as sensitizers in DSSCs because of their chemical adaptability, simple synthetic approaches to various molecular structures, ease of purification, low material cost, high molar extinction coefficients, and high levels of solar spectral absorption in the visible range.27,28 To increase the efficiency of DSSC devices, significant research and development is going into manufacturing DSSCs based on organic compounds with a variety of molecular structures, including D–π–A, D–A–A, D–π–D, A–π–A, and D–π–A–π–D. The D–π–A architecture can facilitate effective intramolecular charge transfer (ICT) from the donor group of the dye to its acceptor unit through the conjugated π-spacer, which is necessary for effective electron injection into the semiconductor.29 In the typical D–π–A architecture, the D units determine the optical characteristics of the dye. The molecular engineering of different donor units has led to the development of various sensitizers that enhance photovoltaic device performance and modulate the absorption characteristics of the dyes.30 Double donors may provide the sensitizers with an expanded absorption region, improved power conversion efficiency, higher molar extinction coefficients and increased light absorption capacities.13 In addition, double donors have influenced both molecular energy levels and light-harvesting ability. In 2008, Ning et al. introduced an innovative design for D–D–π–A structured dyes, employing a starburst triarylamine group as an electron donor. This configuration significantly boosted photovoltaic efficiency when compared to utilizing a single triphenylamine unit. Such an approach holds promise in enhancing the stability of solar cells.31 In addition to increasing light absorption capacity, the extra electron donor in the D–D–π–A structure also helps to prevent dye aggregation.32,33 Moreover, Dai et al. have designed an M45 sensitizer based on the D–D–π–A architecture that has an overall power conversion efficiency of 9.02%.34 Recently, Lin et al. designed double-donor-based dyes (ME101–ME106) with promising outcomes.35 These findings indicate that D–D–π–A-based sensitizers have significant benefits in DSSC fields.
The work of Liang Han et al., where they successfully synthesized a parent dye based on a triphenylamine-coumarin moiety, has inspired us to investigate the tuning of dye properties by modifying the auxiliary donor moiety.13 In this work, we have modified the end donor unit in a D–D–π–A-based architecture. In this context, we have utilized coumarin–thiophene–cyanoacrylic acid as the D–π–A component and integrated an additional end auxiliary donor unit for enhanced efficiency, viz., phenoxazine (D1), indacenodithiophene (D2), dimethoxy-substituted indoline (D3), methoxy-substituted triphenyl amine (D4), methoxy-substituted indoline (D5), N-annulated indenoperylene (D6), dimethoxy-substituted triphenylamine (D7), and methoxy-substituted diphenyl amine (D8). We have used density functional theory (DFT) and time-dependent DFT (TD-DFT) to determine the electronic structure and optical absorption characteristics of the dyes. Based upon the obtained results, we have studied the role of different auxiliary donors in tuning the geometries, electronic structures, and optical properties. The key parameters, including dihedral angle, dipole moment, HOMO and LUMO energies with their differences (ΔH–L values), ionization potential, electron affinity, ground-state oxidation potential, excited-state oxidation potential, light harvesting capacity, absorption properties, reorganization energies, charge transfer rate, etc., have been extensively evaluated. Additionally, we have adopted systematic investigations to gauge the effect of the dye/TiO2 on the electronic and optical characteristics. To assess the performance of the adsorption systems in DSSCs, we have evaluated the geometry, electron injection, absorption spectra, and excited state of the dye/TiO2 cluster system. To the best of our knowledge, design and theoretical study of these molecules based on the D–D–π–A architecture have been reported very rarely. The obtained results offer significant guidance for developing novel dyes based on double donors that will result in highly efficient DSSCs. The structures of all the designed dyes are presented in Fig. 1.
To monitor the performance of the solar cells, the designed dyes have been further investigated after their binding to the TiO2 semiconducting surface. The most common phases of TiO2 are rutile, anatase, and brookite, with anatase being the most widely studied due to its significance in photocatalysis and surface chemistry. In this study, we have adopted the anatase Ti5O10 model to represent the TiO2 semiconducting surface. We have utilized a variety of acetic acid derivatives and cyanoacrylic acid derivatives as anchoring groups, which covalently bond to the semiconductor surface. These derivatives are ideal for enabling effective adsorption of dye-sensitizers onto the semiconducting surface. The DFT formalism has been employed to model Ti5O10 at the atomic level, which allowed for accurate simulations of the atomic structure and electronic properties of the Ti5O10 surface. In this modeling, the positions of the Ti and O atoms have been optimized to replicate realistic surface structures, ensuring that the stoichiometry and morphology align with experimental observations. Each Ti atom is sixfold coordinated, while the O atoms are threefold coordinated. DFT calculations indicate that the anatase (101) surface is the most active, as it contains unpassivated Ti and O atoms, making it widely used for simulating the interface electron transfer process in DSSCs.37,38 For the optimization of the dye–TiO2 complexes, the split valence approach has been applied. The B3LYP-D3/6-31G(d) level of theory has been employed for atoms other than the Ti atom and for the Ti atom the B3LYP-D3 functional along with the LANL2DZ basis set has been employed.39
| IP = E+(M°) − E°(M°), | (1) |
| EA = E°(M°) − E−(M°), | (2) |
The reorganization energy (λ) refers to the change in energy due to the structural reorganization of the dye molecule to reduce the impact of its excess charge.39 There are typically two contributions of λ, the outer-sphere and the inner-sphere contribution. The outer-sphere contribution is caused by electron/nucleus relaxation or polarization of the surrounding medium, while the inner-sphere contribution is caused by the geometry relaxation process, which is connected with the charge received or released by the dye molecule. In this work, we have taken into account only the λ values contributed by the inner-sphere portion. The λ values for cationic species (λ+) and anionic species (λ−) can be calculated by using eqn (3) and (4), respectively:43
| λ+ = [E+(M°) − E°(M°)] − [E+(M+) − E°(M+)], | (3) |
| λ− = [E°(M−) − E−(M−)] − [E°(M°) − E−(M°)]. | (4) |
The power conversion efficiency (PCE or η) of a photovoltaic device is generally expressed using the following equation:42
![]() | (5) |
The mathematical expression for Jsc can be defined as42:
![]() | (6) |
Light harvesting capacity (LHC) is an important parameter that helps in calculating the efficiency of a dye, obtained using the following equation:42
| LHC = 1 − 10−fosc, | (7) |
| ΔGinj = ESOP − ECB. | (8) |
The calculation of ESOP relative to the CB of TiO2 is a key predictive tool for designing of new dyes. For effective electron injection, the ESOP of the dye should be greater than the CB of TiO2 (i.e., −4.0 eV). The ESOP values can be estimated by using eqn (9):42
| ESOP = GSOP + Eg. | (9) |
| GSOP = E°(M°) − E+(M°). | (10) |
It is necessary that the GSOP value of the dye lies below the redox potential of the electrolyte I−/I3− (−4.8 eV).44 Several investigations have indicated that the photovoltaic performance of a DSSC is highly dependent on the driving force of regeneration of the dye. Dye regeneration, quantified by ΔGreg, is a process in which the oxidized dye after electron injection is regenerated by the redox electrolyte. ΔGreg can be evaluated by using eqn (11):39
| ΔGreg = Eredox(I−/I3−) − GSOP. | (11) |
The charge transfer rates (kCT) can be calculated using Marcus theory and are represented by eqn (12):43
![]() | (12) |
![]() | (13) |
![]() | (14) |
![]() | (15) |
| Dyes | ϕ1 (°) | ϕ2 (°) | ϕ3 (°) |
|---|---|---|---|
| D1 | 7.44 | 7.29 | 0.68 |
| D2 | 8.09 | 9.88 | 1.12 |
| D3 | 7.81 | 9.26 | 0.94 |
| D4 | 7.74 | 8.38 | 0.84 |
| D5 | 7.84 | 9.40 | 0.95 |
| D6 | 7.93 | 9.51 | 1.09 |
| D7 | 7.50 | 8.32 | 0.79 |
| D8 | 7.96 | 9.68 | 1.06 |
From Table 1, it has been observed that all the designed dyes are almost planar in structure since the dihedral angles between their D–D, D–π and π–A moieties are effectively small (below 10°). The dihedral angles between the auxiliary donor and donor unit (ϕ1) are in the range of 8° to 9°, for the donor and π unit (ϕ2) are in the range of 7° to 10° and for the π and acceptor unit (ϕ3) are in the range of 0° to 2° for all the dyes. Considering the dihedral angles between all the fragments are less than 10°, this indicates that each auxiliary donor along with the primary coumarin donor has its π-conjugated thiophene bridge located coplanar with the cyanoacrylic acid group, acting as both the acceptor and the anchoring group on the TiO2 surface. These coplanar structures between the donor, linker and acceptor led to a strong conjugation effect as well as allowing effective injection into the conduction band of TiO2 through the cyanoacrylic acid group. Among the studied dyes, dye D1, having a phenoxazine auxiliary donor, possesses the lowest values of dihedral angle among all the fragments. Consequently, dye D1 possesses the most planar structure among all the designed dyes. Analyzing their structural properties, we can conclude that incorporation of an efficient auxiliary donor can enhance the extent of electron delocalization of the sensitizer, which in turn demonstrates the ease of charge transfer.
| Dyes | HOMO (eV) | LUMO (eV) | ΔH–L (eV) |
|---|---|---|---|
| D1 | −4.92 | −2.88 | 2.04 |
| D2 | −5.22 | −2.95 | 2.27 |
| D3 | −5.04 | −2.92 | 2.12 |
| D4 | −4.93 | −2.84 | 2.09 |
| D5 | −5.05 | −2.88 | 2.17 |
| D6 | −5.06 | −2.81 | 2.25 |
| D7 | −4.93 | −2.86 | 2.07 |
| D8 | −5.10 | −2.90 | 2.20 |
Table 2 shows that the studied dyes have ΔH–L values ranging from 2.04 eV to 2.27 eV. This range falls within the reported 1 eV to 4 eV range needed for semiconducting devices.48 This suggests that our designed dyes are favourable for DSSCs. The calculated ΔH–L values are in the following order: D2 > D6 > D8 > D5 > D3 > D4 > D7 > D1. It is evident that dye D1 exhibits the lowest ΔH–L value (2.04 eV), while dye D2 shows the highest (2.27 eV). Compared to the triphenylamine-based donors (D7 and D4), the phenoxazine donor in dye D1 demonstrates a stronger electron-donating capability. This is attributed to the presence of electron-rich oxygen and nitrogen atoms within its heterocyclic ring, which significantly enhance its electron-donating ability.49 However, increasing the number of methoxy groups on the triphenylamine unit enhances its electron-donating ability, thereby enabling fine-tuning of the target properties. The stronger π-delocalization facilitated by the cyclopentadithiophene and benzothiadiazole moieties further boosts the donor strength of the indoline units (as seen in D3 and D5). Consequently, these compounds demonstrate potential as promising candidates for photovoltaic applications.
As frontier molecular orbital (FMO) analysis is an effective approach for exploring the optoelectronic characteristics, we have represented the FMO diagrams for all the dye systems in Fig. 3. From Fig. 3, it is observed that the HOMOs are delocalized over the coumarin and auxiliary donor parts. On the other hand, the LUMOs are delocalized over the acceptor part, with some contribution from the π-bridge unit. This distribution pattern of HOMO and LUMO over the donor, acceptor and linker clearly demonstrates the occurrence of charge separation in our designed dyes and demands valuable consideration for future applications.
| Dyes | Donor | π-bridge | Acceptor |
|---|---|---|---|
| D1 | 0.333 | −0.112 | −0.221 |
| D2 | 0.286 | −0.107 | −0.179 |
| D3 | 0.311 | −0.104 | −0.207 |
| D4 | 0.313 | −0.101 | −0.212 |
| D5 | 0.303 | −0.103 | −0.200 |
| D6 | 0.291 | −0.105 | −0.186 |
| D7 | 0.330 | −0.112 | −0.218 |
| D8 | 0.299 | −0.102 | −0.197 |
| Dyes | GSOP | ESOP | ΔGreg | ΔGinj |
|---|---|---|---|---|
| D1 | −6.06 | −2.94 | 1.26 | 1.05 |
| D2 | −6.44 | −3.29 | 1.64 | 0.70 |
| D3 | −6.24 | −3.09 | 1.44 | 0.90 |
| D4 | −6.08 | −3.01 | 1.28 | 0.98 |
| D5 | −6.24 | −3.11 | 1.44 | 0.88 |
| D6 | −6.37 | −3.26 | 1.57 | 0.73 |
| D7 | −6.07 | −2.95 | 1.27 | 1.04 |
| D8 | −6.18 | −3.25 | 1.38 | 0.74 |
![]() | ||
| Fig. 5 Plot of band alignment of the designed dyes with respect to the CB of TiO2 and the redox potential of I−/I3−. | ||
From Table 4 and Fig. 5, it has been observed that the GSOP values of the dyes lie below the redox potential of the I−/I3− electrolyte couple and the ESOP values lie above the conduction band of the TiO2. From this observation, it is obvious that our designed dyes exhibit spontaneous electron injection from the excited state of the dyes to the conduction band of the TiO2 semiconducting surface and thus spontaneous regeneration of dyes from the electrolyte couple. This demonstrates that all our designed dyes have the ability to serve as potential candidates for DSSC fabrication.
The estimation of the rate of electron injection (ΔGinj) and dye regeneration (ΔGreg) is very important for studying photovoltaic characteristics. It is notable that higher ΔGinj values of the dyes indicate higher efficiency towards electron injection.42,56 As shown in Table 4, the values of ΔGinj for all the dyes are in the range of 0.70 to 1.05 eV, implying that the electrons of the dyes have been effectively injected from the anchoring group into the conduction band of the TiO2 semiconductor surface under photoexcitation. Also, according to Table 4, the ΔGreg values of all the designed dyes are positive. This indicates that the redox level of the electrolyte is higher than the ground state of the dyes and hence responsible for the suppression of electron recombination. For the dye to regenerate more quickly and efficiently, the ΔGreg value has to be low.58 Here, all computed ΔGreg values are between 1.26–1.64 eV. In this regard, we can conclude that the studied dyes are promising towards dye regeneration and electron injection.
The ionization potential (IP) and electron affinity (EA) are two exemplifying criteria for demonstrating the charge transfer characteristics. The amount of energy needed for ionization of a molecule is termed as IP, which is the energy difference between the cationic and neutral states. Conversely, EA is the ability to make anions by accepting electrons, which is the difference in energy at the ground states between neutral and anionic molecules.56 Both of these parameters can be calculated using eqn (1) and (2), respectively, and the corresponding values are given in Table 5. It is already well known that a low IP speeds up the creation of holes by facilitating the removal of electrons. Furthermore, a high value of EA indicates that it will be challenging to remove the conduction-band electrons. The EA value of the dye molecule explains the recombination between the injected electron and the oxidized dye species. A lower EA value ensures easy removal of electrons from the conduction band. Therefore, in order to transport electrons to the semiconducting surface effectively, a dye sensitizer has to have a low EA value.29 Table 5 illustrates that IP is found to be between 6.06 and 6.44 eV for all the designed dyes. This suggests that the dyes have superior stability against oxidation. Conversely, the projected range for the EA values is 1.59–1.85 eV. This indicates that our designed dyes meet the requirements for becoming a suitable dye sensitizer.
| Dyes | IP | EA |
|---|---|---|
| D1 | 6.06 | 1.59 |
| D2 | 6.44 | 1.85 |
| D3 | 6.24 | 1.63 |
| D4 | 6.08 | 1.60 |
| D5 | 6.24 | 1.64 |
| D6 | 6.37 | 1.72 |
| D7 | 6.07 | 1.59 |
| D8 | 6.18 | 1.70 |
| Dyes | Eg (eV) | λmax (nm) | fosc | Transitions | LHC | µ (Debye) |
|---|---|---|---|---|---|---|
| D1 | 2.02 | 522 | 2.04 | H → L (83%) | 0.990 | 13.23 |
| D2 | 3.15 | 400 | 1.26 | H → L (61%) | 0.945 | 10.92 |
| D3 | 2.37 | 481 | 1.79 | H → L (70%) | 0.983 | 12.46 |
| D4 | 2.21 | 498 | 1.93 | H → L (73%) | 0.988 | 12.76 |
| D5 | 2.52 | 470 | 1.67 | H → L (68%) | 0.978 | 12.20 |
| D6 | 2.97 | 420 | 1.49 | H → L (65%) | 0.967 | 11.05 |
| D7 | 2.09 | 506 | 2.01 | H → L (80%) | 0.990 | 13.01 |
| D8 | 2.71 | 434 | 1.55 | H → L (67%) | 0.971 | 11.46 |
Specifically, fosc represents the probability of electromagnetic radiation absorption during transition between the energy levels of a particular molecule.59 From Table 6, it has been observed that all the designed dyes possess comparatively high fosc values. Among all the designed dyes, the D1 dye has the highest absorption maximum (522 nm) with maximum oscillator strength (2.04). This may be due to the (+I) effect of the phenoxazine donor group attached to the D1 dye. The presence of strong electron-donating and strong electron-accepting groups can reduce the HOMO–LUMO gap, thereby facilitating efficient electron transfer from the donor to the acceptor group. This can result in a redshift of the absorption maximum and an increase in oscillator strength. Thus, the high oscillator frequency and high absorption maxima exhibited by all designed dyes may be due to the +I effect of the donor moieties at the peripheral positions. Furthermore, it is evident that the electronic transitions with the highest wavelength of absorption occur primarily due to the H → L transitions (61–83%) for all the dyes. The corresponding UV-visible spectra for the designed dyes have been provided in Fig. 6.
The dipole moment (µ) is a measure of the separation of positive and negative charges within a molecule that gives information about its polarity. Higher values of µ indicate that the designed dyes are polar in nature.56 In donor–acceptor systems, where a dye is designed to donate electrons from donor to acceptor moieties, a higher dipole moment can enhance the separation of charges upon absorption of light. From Tables 6, it has been observed that all the designed dyes possess comparatively high values of µ. Among them, D1 and D7 have higher values of µ due to the phenoxazine and dimethoxy-substituted triphenylamine donor groups.
We have calculated LHC to predict the Jsc values using eqn (6). From eqn (6), it is evident that a high LHC value will lead to a high Jsc value. Table 6 shows that all the dyes have relatively high LHC values, ranging from 0.945 to 0.990. Furthermore, D1 and D7 possess the highest LHC values (i.e., 0.990) among all studied dyes. Based on the LHC values, D1 and D7 are expected to have the greatest Jsc values among all the designed dyes, which can lead to maximum η.
The reorganization energy (λ) is another crucial factor that influences the efficiency of DSSCs. The energy cost due to the conformational change during photoexcitation is defined as the reorganization energy.39 Marcus electron transfer theory relates the total reorganization energy to the rate of the electron transfer, which is expressed by eqn (12). We have calculated the reorganization energy for both holes (λ+) and electrons (λ−) to understand the charge transportation in the designed dyes and their respective values have been reported in Table 7. For effective charge transportation, the λ value (λ+ or λ−) needs to be low. Thus, it is necessary to have a low value of λ for a high charge transfer rate. A lower λ+ value reflects the hole transporting nature of the dyes and conversely, a lower λ− value reflects the electron transporting nature of the dyes.
| Dyes | λ+ | λ− | λtot |
|---|---|---|---|
| D1 | 0.059 | 0.164 | 0.223 |
| D2 | 0.250 | 0.356 | 0.606 |
| D3 | 0.196 | 0.249 | 0.445 |
| D4 | 0.126 | 0.217 | 0.343 |
| D5 | 0.226 | 0.288 | 0.514 |
| D6 | 0.243 | 0.318 | 0.561 |
| D7 | 0.068 | 0.176 | 0.244 |
| D8 | 0.239 | 0.316 | 0.555 |
From Table 7, it has been observed that for all the designed dyes, the λ− values are higher than the λ+ values. As a result, all of our designed dyes behave as hole transporting materials. Additionally, it is noted that, among all the designed dyes, dyes D1 and D7 possess lower λ+ values compared to the rest. As a result, among all designed dyes, D1 and D7 are expected to have facile hole transportation.
We have also calculated the total reorganization energy (λtot) values (i.e., the sum of the values of λ+ and λ−) and these are reported in Table 7. For achieving greater electron injection, i.e., to obtain higher current density, the λtot value must be smaller. This will reduce recombination.42 From Table 7, it is observed that the designed D1 and D7 dyes possess lower values of λtot compared to the other dyes. This reveals better electron–hole separation efficiency of the D1 and D7 dyes, which may have slower recombination processes compared to the other designed dyes.
To gauge the electronic coupling matrix element (V), we have considered the π-stacking in a cofacial arrangement of the two adjacent dyes. The π-stacking orientation of the dyes has been evaluated by arranging two adjacent dyes in a face-to-face manner at a distance of 3.5 Å. The representative arrangement of two stacked dyes (for all the dye systems) has been presented in Fig. 7 and the calculated V values (using eqn (13)) are reported in Table 8. Using these V values, we have calculated the charge transfer rates for holes (kCT+) and electrons (kCT−) (using eqn (12)) and reported these in Table 8.
| Dyes | V+ (eV) | V− (eV) | kCT+ × 1013 (s−1) | kCT− × 1013 (s−1) | µhop+ (cm2 V−1 s−1) | µhop− (cm2 V−1 s−1) |
|---|---|---|---|---|---|---|
| D1 | 0.032 | 0.029 | 47.26 | 3.178 | 6.656 | 4.721 |
| D2 | 0.012 | 0.011 | 12.06 | 1.103 | 0.251 | 0.069 |
| D3 | 0.257 | 0.151 | 18.13 | 8.116 | 3.964 | 0.411 |
| D4 | 0.262 | 0.155 | 20.48 | 2.152 | 4.318 | 0.432 |
| D5 | 0.077 | 0.029 | 17.53 | 5.994 | 0.373 | 1.265 |
| D6 | 0.071 | 0.053 | 10.48 | 0.169 | 0.296 | 0.043 |
| D7 | 0.048 | 0.016 | 46.56 | 1.712 | 5.974 | 0.355 |
| D8 | 0.082 | 0.077 | 16.66 | 3.495 | 0.323 | 0.474 |
From Table 8, it is evident that the kCT+ values of the compounds are higher than their corresponding kCT− values. This result is in accordance with the reorganization energy values. Moreover, it is also apparent from Table 8 that dye D1 and D7 exhibit comparatively higher kCT+ values due to their lower value of λ+ among all the designed dyes. Another significant measure is the hopping mobility (µhop), which helps in determining the conducting capacity of the organic dyes. A high µhop value signifies higher electronic coupling between the adjacent dyes, which in turn indicates a better conducting capacity of the organic dyes. The calculated µhop values for holes (µhop+) and electrons (µhop−) are reported in Table 8. From this table, it is observed that among all the studied dyes, D1 and D7 possess higher values of µhop+. These values are consistent with the observed kCT+ values for these dyes. Therefore, we are hopeful that our designed dyes may serve as potential candidates for the fabrication of optoelectronic devices in the near future.
As demonstrated in Fig. 10, the HOMOs in all the dye clusters are mainly delocalized over the donor portion of the dyes. Conversely, the LUMOs are primarily delocalized over the acceptor portion with some contributions from the π-bridging unit as well as the Ti5O10 cluster. Thus, this analysis demonstrates the existence of an efficient charge transfer mechanism in the dye–Ti5O10 clusters.
As a part of our calculations, we have also calculated the Ti–O bond lengths of the designed dyes and the results are presented in Table 9. The representation of the Ti–O bond length in the representative D1–Ti5O10 cluster is presented in Fig. 11. Table 9 shows that all the designed dyes have Ti–O bond lengths ranging from 2.034–2.049 Å. These values are almost in agreement with the theoretically reported Ti–O bond lengths (2.03–2.24 Å) for various dye–TiO2 clusters.42 These observations suggest that all the designed dyes undergo chemisorption on the TiO2 surface.
| Dye–Ti5O10 | Ti–Oa (Å) | Ti–Ob (Å) |
|---|---|---|
| D1–Ti5O10 | 2.034 | 2.046 |
| D2–Ti5O10 | 2.037 | 2.040 |
| D3–Ti5O10 | 2.035 | 2.049 |
| D4–Ti5O10 | 2.036 | 2.040 |
| D5–Ti5O10 | 2.035 | 2.049 |
| D6–Ti5O10 | 2.036 | 2.040 |
| D7–Ti5O10 | 2.037 | 2.040 |
| D8–Ti5O10 | 2.036 | 2.049 |
Additionally, we have computed the ΔH–L values and ground-state dipole moments (µg) for all the designed dye–Ti5O10 clusters. The results have been reported in Table 10.
| Dye–Ti5O10 | ΔH–L (eV) | µg (Debye) |
|---|---|---|
| D1–Ti5O10 | 1.57 | 20.41 |
| D2–Ti5O10 | 1.86 | 15.06 |
| D3–Ti5O10 | 1.66 | 18.06 |
| D4–Ti5O10 | 1.60 | 18.56 |
| D5–Ti5O10 | 1.74 | 17.55 |
| D6–Ti5O10 | 1.85 | 16.05 |
| D7–Ti5O10 | 1.58 | 19.18 |
| D8–Ti5O10 | 1.80 | 16.75 |
From the comparison of Tables 2 and 10, it has been observed that the ΔH–L values of the dye–Ti5O10 clusters are lower than those of the isolated dyes. Moreover, from the comparison of Tables 6 and 10, it has been observed that the dipole-moment values of the dye–Ti5O10 clusters (µg) are higher than those of the isolated dyes (µ). These observations indicate that binding of the dyes to the Ti5O10 semiconducting surface leads to the enhancement of their charge transport characteristics.
The molecular electrostatic potential surface (MEPS) is a popular way of visualizing the electrostatic nature of dye molecules. The MEPS serves as a useful resource, offering qualitative insights into charge transportation from the donor moiety to the acceptor moiety through the π-linker.65 The positive potential increases in the following color order: red < orange < yellow < green < blue. Here, the blue and red colors depict the electron-deficient and electron-rich regions, respectively. We have presented the MEPS plots of the designed dye–Ti5O10 clusters in Fig. 12.
From Fig. 12, it is observed that for all designed dye–Ti5O10 clusters, the positive charge is spread over the dyes and the negative charge is spread over the Ti5O10 surface. This observation signifies the charge transfer characteristics of all the designed dyes.
In order to study the absorption properties of the dye–Ti5O10 clusters, we have computed the excitation energies (Eg), maximum absorption wavelengths (λmax), oscillator strengths (fosc), transitions, and contributions of the frontier orbitals. The results are shown in Table 11.
| Dyes | Eg (eV) | λmax (nm) | fosc | Transitions | Eb (eV) |
|---|---|---|---|---|---|
| D1–Ti5O10 | 2.15 | 536 | 2.10 | H → L (90%) | 0.20 |
| D2–Ti5O10 | 3.21 | 465 | 1.63 | H → L (68%) | 0.30 |
| D3–Ti5O10 | 2.42 | 515 | 1.90 | H → L (73%) | 0.26 |
| D4–Ti5O10 | 2.37 | 520 | 2.01 | H → L (78%) | 0.28 |
| D5–Ti5O10 | 2.60 | 507 | 1.82 | H → L (70%) | 0.29 |
| D6–Ti5O10 | 2.99 | 470 | 1.71 | H → L (69%) | 0.23 |
| D7–Ti5O10 | 2.18 | 530 | 2.06 | H → L (87%) | 0.22 |
| D8–Ti5O10 | 2.84 | 460 | 1.76 | H → L (71%) | 0.28 |
Table 11 illustrates that the dye–Ti5O10 clusters experienced higher maximum absorption wavelength (λmax) values and lower excitation energies (Eg). This finding suggests that the adsorbed dyes have a red shift as compared to the isolated dyes and follow the same trend. The absorption maxima for all the dye–Ti5O10 clusters are primarily due to the H → L electronic transition with contributions of 68–90%. The corresponding spectra of the dye–Ti5O10 clusters have been presented in Fig. 13.
For the TDM analysis of the dye clusters, all dye-cluster systems have been divided into two parts: the dye and the Ti5O10 cluster. The Multiwfn 3.8 program has been used to generate TDM plots for the dye clusters in the gas phase. Fig. 14 presents the TDM plots for the dye-clusters at an isosurface value of 0.002. The plots clearly show that the charge density is extensively distributed across both the dye and the Ti5O10 cluster, appearing predominantly along the diagonal. Analysis of all designed dyes confirms efficient electron transfer from the dye to the Ti5O10 cluster. The visual representation indicates a green region over the dye and a blue region over the Ti5O10 cluster, highlighting significant electron transfer from the dye to the TiO2 semiconductor surface.
The exciton binding energy (EBE) plays a crucial role in assessing the effectiveness of solar cells. It is directly related to the separation of charges within solar cells.66 The working principle of DSSCs involves the absorption of sunlight by dye molecules, leading to the generation of excitons (i.e., electron–hole pairs). These excitons subsequently split upon reaching the surface of the semiconductor. In this context, EBE is defined as the amount of energy needed to separate the electron and hole from an exciton.67,68 This ease of charge separation is crucial in applications of dye-sensitized solar cells, where efficient conversion of light energy into electrical energy relies on effective charge transfer processes. It has been reported that the optimal value of Eb for organic semiconducting materials falls within the range of 0.2 to 1 eV.69 The expression for Eb can be written as Eb = ΔH–L − E1, where E1 denotes the optical band gap.69,70 The EBE values have been reported in Table 11.
Table 11 indicates that all dye–Ti5O10 clusters have low Eb values, which in turn indicates easy charge separation. Among the eight dye–Ti5O10 clusters, the dye D1–Ti5O10 cluster has a lower Eb value. This suggests the electron transport from the dye to the Ti5O10 semiconducting surface will be easier in the D1 dye.
In order to assess the viability of our designed dyes, we have used the Ti5O10 surface. For dye–Ti5O10 clusters, it can be inferred that the µg values are higher and ΔH–L values are lower than those of their isolated counterparts. This confirms the enhancement of the charge transport properties of the designed dyes upon binding the dye to the Ti5O10 semiconducting surface. Furthermore, the absorption spectra of the dye–Ti5O10 clusters demonstrate a rise in the λmax values. This suggests the presence of a red shift in comparison to the isolated dyes. Consequently, the dyes exhibit enhanced performance upon binding to the TiO2 surface. In short, we can conclude that our designed dyes serve as viable options for the fabrication of DSSCs.
Supplementary information: optimized coordinates of the studied dyes, HOMO, LUMO, ΔH–L, and absorption energies of the test compound, the optimized structure of the test compound, and the justification for selecting Ti5O10 clusters have been provided. See DOI: https://doi.org/10.1039/d5ra07959d.
| This journal is © The Royal Society of Chemistry 2025 |