Next Article in Journal
Cantharidin Impedes Activity of Glutathione S-Transferase in the Midgut of Helicoverpa armigera Hübner
Previous Article in Journal
Preparation and Evaluation of Dental Resin with Antibacterial and Radio-Opaque Functions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Role of the Conjugate Bridge in Electronic Structures and Related Properties of Tetrahydroquinoline for Dye Sensitized Solar Cells

1
State Key Laboratory of Gansu Advanced Non-ferrous Metal Materials, Lanzhou University of Technology, Lanzhou 730050, Gansu, China
2
Department of Applied Physics, Lanzhou University of Technology, Lanzhou 730050, Gansu, China
3
Gansu Computing Center, Lanzhou 730030, Gansu, China
4
Institute of Electronic Information Science and Technology, Lanzhou City University, Lanzhou 730070, Gansu, China
5
College of Physics and Electronic Engineering, Northwest Normal University, Lanzhou 730070, Gansu, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2013, 14(3), 5461-5481; https://doi.org/10.3390/ijms14035461
Submission received: 7 December 2012 / Revised: 11 January 2013 / Accepted: 4 February 2013 / Published: 8 March 2013
(This article belongs to the Section Physical Chemistry, Theoretical and Computational Chemistry)

Abstract

:
To understand the role of the conjugate bridge in modifying the properties of organic dye sensitizers in solar cells, the computations of the geometries and electronic structures for 10 kinds of tetrahydroquinoline dyes were performed using density functional theory (DFT), and the electronic absorption and fluorescence properties were investigated via time dependent DFT. The population analysis, molecular orbital energies, radiative lifetimes, exciton binding energies (EBE), and light harvesting efficiencies (LHE), as well as the free energy changes of electron injection (ΔGinject ) and dye regeneration ( Δ G dye regen ) were also addressed. The correlation of charge populations and experimental open-circuit voltage (Voc) indicates that more charges populated in acceptor groups correspond to larger Voc. The elongating of conjugate bridge by thiophene units generates the larger oscillator strength, higher LHE, larger absolute value of ΔGinject, and longer relative radiative lifetime, but it induces the decreasing of EBE and Δ G dye regen. So the extending of conjugate bridge with thiopene units in organic dye is an effective way to increase the harvest of solar light, and it is also favorable for electron injection due to their larger ΔGinject. While the inversely correlated relationship between EBE and LHE implies that the dyes with lower EBE produce more efficient light harvesting.

1. Introduction

As a novel photovoltaic device, dye sensitized solar cells (DSC) have attracted much attention because of many merits, such as easy fabrication, lower cost, etc.[15]. However, the photo-to-current conversion efficiency (PCE) of DSC is still not high enough to commercialize. The central interest of researching DSC is how to improve the PCE. It has been found that all of the main components of DSC, including dye sensitizers, anode, and cathode, as well as electrolyte, can affect the PCE. Particularly, the dye sensitizers, which take the function of light harvesting and photo-excited electron injection in DSC, have a significant influence on the PCE [611].
Up to now, the dye sensitizers of DSC can be classified into metal-organic complexes and metal-free organic dyes [710]. In metal-free organic dyes sensitizers, such as cyanines [1214], hemicyanines [15,16], triphenylmethanes [8,17,18], perylenes [1923], coumarins [2426], porphyrins [2732], squaraines [3335], indoline [36,37], and azulene-based dyes [38]etc., were developed since their high molar absorption coefficient, relatively simple synthetic procedure, various structures, and lower cost. Recently, the 12.3% PCE has been achieved by cosensitization of donor-π conjugate bridge-acceptor (D-π-A) zinc porphyrin dye YD2-o-C8 with another triphenylamine-based organic dye Y123 [39]. So the development of novel organic dye sensitizer is a promising method to improve the PCE of DSC.
For dye sensitizer with good performance in DSCs, the following properties are required [40]. First, the dye should have a broad absorption spectrum in the visible region to harvest more photons from solar radiation. Therefore, the wide absorption region and larger absorption coefficient are expected. Second, in terms of simplified schematic of the energetic processes occurring in conventional DSC and weak coupling limit, the dye should have the suitable energy level of ground state S0 and excited state S1. The level of S0 must locate below the redox couple of electrolyte, and therefore the electron acceptance of oxidized dye from electrolyte is energetically permitted. The level of S1 should be suitable above the conduction band of semiconductor in anode so that the electron in excited dyes can inject into the anode semiconductor. Also, the S1 should have significant charge transfer (CT) character. So, the further developments of novel dye sensitizers depend on the quantitative information of dye sensitizers [9,11,41], such as electronic structures, absorption properties, etc. The theoretical investigations of the physical properties of dye sensitizers pave the way to disclose the relationships among the performance, structures, and the properties. For instance, on the basis of density functional theory (DFT) calculations for the chemical structures and exciton binding energies (EBE) of several pure organic dyes, the novel organic triphenylamine-based dye sensitizer EB-01 was designed, and the over 9% of PCE was achieved by EB-01 sensitized DSC [42].
The metal-free organic dye sensitizers with good performance in DSC usually have a D-π-A structure [710]. In D-π-A dyes, the π-electron delocalization from the electron-donor to the electron-acceptor through the π-conjugated linker can affect the performance of DSC [4345]. Therefore, the π-conjugated linker should play an essential role in CT process and photophysical property in the D-π-A dyes. The tetrahydroquinoline dyes had been engineered and synthesized as sensitizers for applications in DSC, and the different length of thiophene-containing conjugate bridges were introduced to dye sensitizers and served as electron spacers [46,47]. To understand the role of conjugate bridge in modifying the properties of organic dye sensitizers in solar cells, the computations of the geometries and electronic structures for 10 kinds of tetrahydroquinoline dyes (the molecular structures presented in Table 1) were performed using DFT, and the electronic absorption and fluorescence properties were investigated via time dependent DFT (TDDFT) with polarizable continuum model (PCM) for solvent effects. The population analysis, molecular orbital energies, radiative lifetimes, EBE, light harvesting efficiencies (LHE), and the free energy changes of electron injection ( ΔGinject ) and dye regeneration ( Δ G dye regen ) are also studied.

2. Computational Methods

The accuracy of DFT calculations depend on the selected functional and basis sets. It has been demonstrated that the polarized split-valence 6-31G(d,p) basis sets are sufficient for calculating the excitation properties of organic dyes [48], and introducing additional diffuse functions in basis sets generate negligible effects on the electron density and hence on the accuracy of DFT and TDDFT results [49]. So the 6-31G(d,p) basis sets were adopted in the present work. Generally, for the dye sensitizers with good performance in DSC, the electronic excitations are CT processes. To select functional for the reliable description of excited properties of dyes, the dye sensitizer C1-1 was selected as the representative dye. The geometry in gas phase was optimized, and then it was further optimized in solvent. Based upon the optimized geometry in solvent, the absorption properties were investigated using TDDFT. The non-equilibrium version of the PCM [50,51] was adopted for calculating the solvent effects. The above mentioned calculations were performed with different functionals, including the hybrid functionals PBE0 [5254], M062X [55], and the long range corrected hybrid functionals CAM-B3LYP [56], LC-ωPBE [5760], and ωB97X [61]. Table 2 presents the comparison between calculated λmax using different functionals and experimental λmax for C1-1. The discrepancy between the experimental and calculated results with PBE0, CAM-B3LYP, LC-ωPBE, ωB97X, and M062X functionals are about 0.47, 0.02, 0.49, 0.34, and 0.03 eV, respectively. Apparently, the CAM-B3LYP and M062X functionals provide reliable results, predicting excitation energies agree well with the experimental data. Furthermore, the absolute deviation of CAM-B3LYP functional is slightly smaller than that of M062X. Therefore, CAM-B3LYP is the selected functional for calculating the geometries, electronic structures, electronic absorption, as well as fluorescence spectrum for other tetrahydroquinoline dyes. The solvent effects were also dealt with PCM method. All of the calculations were performed by using Gaussian09 package [62].

3. Results and Discussion

3.1. Geometrical Structures

The optimized geometries of tetrahydroquinoline dyes C1-1, C1-2, C1-3, C1-4, C1-5, C2-1, C2-2, C2-3, C2-4, and C3-1 are shown in Figures 1 and 2. The selected geometrical parameters, including bond lengths, bond angles, and dihedrals are listed in Tables S1–S3 in supporting information (SI). The calculated geometrical data indicate that the geometrical parameters of same group (tetrahydroquinoline, vinylene, thienyl and dithieno[3,2-b;2′,3′-d]thienyl, etc.) in different dyes are very similar. This can be understood from the localized interaction of chemical bonds. The cyanoacrylic acid groups in the dyes are coplanar with the groups in conjugate bridge. The torsion angles among the thiophene-rings in bithiophene and dithieno-[3,2-b;2′,3′-d]thiophene groups are quite small. For instance, the torsion angle between thiophene-rings in bithiophene in C1-2 is about 1.7°. The conjugated effects, which can be enhanced by the quasi-planar structure of bridge and the coplanar character between acceptor moieties and conjugate bridges, extend the delocalization of electron, and it is favorable for efficient CT from the chromophore to the carboxyl group. Similar geometric characters are also appeared in the organic dye sensitizers JK-1 and JK-2 [63], JK-16 and JK-17 [64,65], and other organic dyes which contain cyanoacrylic acid group [6669].
To investigate the role of conjugate bridge length in the performance of dyes in DSCs, we define “L” as the distance between two special C atoms. One is the C atom in the carboxyl group. The other one is the C atom in tetrahydroquinoline moiety where the conjugate bridge connects. When the dyes adsorbed on the surface of semiconductor nanoparticles, the photo-induced CT occurs at the interface between dyes and semiconductor nanoparticles. Thus, the L can describe the CT distance and it can describe conjugate length to some extent. The Table S4 in SI lists the calculated L. The data indicate that the lengths of conjugate bridge with bithiophene and dithieno-[3,2-b;2′,3′-d]thiophene groups are very similar. On the basis of the calculated L and the reported photovoltaic experimental data [47], the dependence of open-circuit photovoltage (Voc), short-circuit photocurrent density (Jsc), and solar-to-electrical energy conversion efficiencies (η) on the length of conjugate bridges are plotted in Figure 3. It can be found that the longest L with terthiophene is unfavorable to improve dye performance in DSCs, while the shorter L with thiophene (C2-1) or thienylvinyl (C1-1) have better performance. It is understandable because the longer L may increase the possibility that the excited electrons decay to the electrolyte [70].

3.2. Electronic Structures

In order to analyze the charge distribution and the electron-transfer mechanism of D-π-A dyes, the Natural Bond Orbital (NBO) analysis has been performed, based upon the optimized structure of the ground state. The calculated NBO results are listed in Table 3. The positive charges of tetrahydroquinoline moieties represent the acts as an effective electron-donor unit. Contrarily, the negative NBO charges of cyanoacrylic acid reveal that the electrons trap in the electron-acceptor unit. The data also suggest the charges of similar acceptors in different dyes are very similar, but the charges of donors and conjugate bridges are more flexible. This means that the natural charges in dye sensitizers are dominated by acceptor moieties.
The open-circuit voltage Vocversus the charges of acceptor groups in tetrahydroquinoline dyes is presented in Figure 4 in order to investigate the relationship between these quantities. It can be found that more charges populated in acceptor groups correspond to larger Voc. This means more charges populated in acceptor are favorable to increase Voc. So, introducing strong electron-withdrawing group in acceptor of dye sensitizer is a possible way to increase Voc of DSC. This tendency is similar to our previous work for the dyes of coumarin NKX derivatives [70].
The highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) energies of better dye sensitizers are required to locate suitable values for matching the conduction band edge of semiconductor and redox potential of electrolyte in DSCs. The HOMO level corresponds to the oxidation potential of dye sensitizer [71], and the larger oxidation potential increase the driving force for the reduction of oxidized dye [72]. The HOMO and LUMO energies, as well as the HOMO-LUMO gaps (Eg) of tetrahydroquinoline dyes in solvent are presented in Figure 5, and the data are listed in Table S5 in SI. The ranges of HOMO, LUMO, and Eg for tetrahydroquinoline dyes in gas phase are about −5.85 to −6.26 eV, −1.32 to −1.77 eV and 4.09 to 4.94 eV, respectively, and the corresponding data in solvent are about −5.92 to −6.25 eV, −1.46 to −1.79 eV and 4.13 to 4.79 eV, respectively. For C1-n and C2-n (n = 1–3) dyes, the HOMO energies are increased and the LUMO energies are decreased in turn since the size of the thiophene-containing increases. Therefore, the Eg of the dyes are reduced in turn. Compared the C1-n and C2-n (n = 1–4) dyes with giving n, the corresponding π-conjugated linker of C1-n dyes have one vinylene group more than that of C2-n dyes, though they have same electron donor and acceptor groups, and then the HOMO energies of C1-n are more positive, the LUMO energies of C1-n are more negative, and the Eg of C1-n is smaller. The comparison of the dyes with bithiophene and terthiophene in conjugate bridges suggests that introducing rigid planar dithieno-[3,2-b;2′,3′-d] thiophene in conjugate bridges generate lower HOMO energies and broader Eg. From the data of C1-2 and C1-5, it can be found that inserting vinylene group in the middle of bithiophene elevate HOMO about 0.09 eV in gas phase (0.05 eV in solvent), decline LUMO about 0.09 eV in gas phase (0.06 eV in solvent), and thus reduce Eg about 0.17 eV in gas phase (0.11 eV in solvent). The comparison of C2-1 and C3-1 indicates that substitution of hydrogen with phenyl increase HOMO about 0.08 eV in solvent, and the reduced Eg by about 0.25 eV.

3.3. Absorption and Emission Spectra: UV-Vis and Fluorescence Spectra

The electronic absorption spectra of tetrahydroquinoline dyes C1-1, C1-2, C1-3, C1-5, C2-1, C2-2, and C3-1 were measured in ethanol solution, and the absorption spectra of the dyes C1-4, C2-3, and C2-4 were measured in DMF solution. The experimental and calculated electronic absorption λmax (nm/eV), as well as the λmax errors (nm/eV) between the experiment and the calculation are listed in Table 4. It can be found that the calculated results agree well with that of experiment. So, the results of CAM-B3LYP functional combined with PCM method are reliable for analyzing excited state properties of tetrahydroquinoline dyes. Apparently, with the elongating conjugate bridge by thiophene units, the corresponding λmax has a red-shift, which is favorable for matching the solar radiation spectra. The corresponding oscillator strength also increases in the order because the longer conjugate bridge extends the overlap of HOMO and LUMO.
To obtain the microscopic information about the electronic transitions, we check the corresponding MO properties. The absorptions in visible and near-UV region are the most important regions for photo to current conversion, so only the singlet→singlet transitions of the absorption bands with the wavelength longer than 300 nm and the oscillator strength larger than 0.1 are listed in Table 5. The isodensity plots of the frontier MOs that related to the absorptions in visible and near-UV region are presented in Table 6. The MOs indicate that the HOMO-1 of the dyes are delocalized in the molecules, and the HOMOs are mainly contributed from the framework of quinoline moiety to conjugate bridge groups, whereas the LUMOs and LUMO + 1 mainly locate on the acceptor cyanoacetic acid moieties and conjugate bridge groups. In addition, the molecular orbital overlaps of HOMO and LUMO are highly coupled to the π-conjugated linker. For these dyes, the maximum absorptions in UV/vis spectra are dominated by HOMO→LUMO π→π* transitions. Furthermore, the overlap between HOMO and LUMO of the dyes suggest the maximum absorptions have some local excited transitions in conjugate bridge. While the relocations of the HOMOs and LUMOs in the dyes support that the transitions at maximum absorptions have intramolecular charge transfer (IMCT) character. The further MO analysis indicates that the quinoline groups are effective chromophores in IMCT, and then they play an important role in the sensitization of DSC.
For the dye sensitizers with good performance in DSCs, the LHE are expected to be as high as possible to increase the photocurrent if the excited processes have CT character. LHE can be calculated as [73]:
LHE = 1 - 10 - A = 1 - 10 - f
where A is the absorption coefficient and f is oscillator strength of the excited state associated to the λ max ( 1 ). The calculated LHE of tetrahydroquinoline dyes are presented in Figure 6. It can be found that C1-3 and C2-3 have the largest LHE in C1-n (n = 1–5) and C2-n (n = 1–4) series, respectively. In addition, the elongating of conjugate bridge generates the larger oscillator strength and LHE because of the enhanced overlap between the ground states (mainly contributed by HOMO) and the first excited states (mainly contributed by LUMO). So, extending the conjugate bridge in D-π-A organic dye is an effective way to increase the harvest of solar light.
The EBE, as an important quantity for the efficiency of excitonic solar cells, determines the charge separation in solar cells [42]. The EBE can be calculated as the difference between the electronic and optical band gap energies [74]. The electronic band gap is calculated as the energy difference between the HOMO and LUMO levels, while the first excitation energy is adopted as the optical gap [42,75]. The calculated EBE (in eV) of tetrahydroquinoline dyes are shown in Figure 6. It indicates that the elongating of conjugate bridge generates the decreasing of EBE, which is favorable for photo-to-current energy conversion via the dissociation of a bound exciton (hole/electron pair). Furthermore, the smallest EBE of C1-3 and C2-3 dyes suggest the exciton can dissociate more effectively than that in other dyes. More interestingly, the figure reveals that the EBE and the LHE are inversely correlated, implying that the dyes with lower EBE produce more efficient light harvesting. This agrees with the relationship between the EBE and quantum yield [42].
Concerning fluorescence, the calculated and experimental emission maxima λ and Stokes shift (SS) for the tetrahydroquinoline dyes are listed in Table 7. For most of the dyes, except C1-1 and C2-1, the computed emission maxima are in good agreement with the experimental data because the discrepancy is lower than 0.3 eV. Particularly, for C2-3 and C1-4, a quantitatively agreement with the experimental results is obtained since the errors are about 0.03 and 0.08 eV, respectively. While in the case of C1-1 and C2-1, the errors (0.34 and 0.45 eV, respectively) are larger than that of others. This is probably to be ascribed to the neglect of direct solute-solvent interactions [76]. After state specific correction, the calculated emission maxima of C1-1 and C2-1 dyes are about 2.14 and 2.41 eV, respectively, and therefore the errors of emission maxima are reduced to 0.22 and 0.32 eV, respectively. Compared with emission maxima, the SS has similar errors. It can be found that, except C1-3, elongating conjugate bridge generates red shift of emission maxima, larger oscillator strength, and increasing of SS due to the enhanced flexibility by the longer conjugate linker, allowing for a larger structural relaxation from the excited state S1 to the ground state S0.
Radiative lifetimes for spontaneous emission from S1 were computed from fluorescene energies (Efluo) and oscillator strength, which represents the transition probability (f in a.u.) as [76]:
τ = 2 π ɛ 0 m e 2 c 3 e 4 ( E fluo ) 2 f
where, me is the electron mass, ɛ0 is the vacuum permittivity, is the reduced Plank constant, c is the speed of light, and e is the elementary charge. To investigate the role of conjugate bridge in radiative lifetime, rather than accurate reproduce the radiative lifetime, we define relative radiative lifetime (τrel) of the dye molecule by assuming C1-1 radiative lifetime (ιC1–1) as reference value:
τ rel = τ τ C 1 - 1 = ( E fluoC 1 - 1 ) 2 f C 1 - 1 ( E fluoi ) 2 f i
where EfluoC1–1, Efluoi, fC1–1, and fi are fluorescene energies and oscillator strength of C1-1 and the selected dye molecule, respectively. The calculated relative radiative lifetime of the tetrahydroquinoline dyes are listed in Table 6. Similar to oscillator strength, the relative radiative lifetime is increasing with the elongating conjugate bridge by thiophene unit. The shortest lifetime of C1-4 and C2-4 may be ascribed to their planar conjugation (by dithienothiophene group), favoring charge recombination.

3.4. The Driving Force of Electron Injection and Dye Regeneration

The electron injection from the excited dyes to semiconductor conduction band and the dye regeneration processes can be described as CT reaction. The driving force, which is defined as the potential difference between the oxidized dyes and the electrolyte, determines the yield of electron collection and dye regeneration. Furthermore, the oxidation potential of the dyes must be more positive than that of electrolyte, thus ensuring enough driving force for a fast and efficient regeneration of the dye cation radical and avoiding the geminate charge recombination between oxidized dye sensitizers and the nanocrystalline TiO2 film where the photo-excited electrons are injected. In terms of the Marcus theory for electron transfer [77], the CT rate constants can be affected by free energy change related to the reaction. The free energy change for electron injection ( ΔGinject ) affects the electron injection rate and therefore the Jsc in DSCs. ΔGinject can be viewed as the electron injection driving force [44]. According to Preat’s method [73], assuming the electron injection occurs from the unrelaxed excited state of the dye, the ΔGinject can be calculated by the following Equation:
Δ G i n j e c t = E O X d y e * - E C B S C
where E O X d y e * is the oxidation potential of the dye in the excited state and E C B S C is the reduction potential of the conduction band of the semiconductor. The reported E C B S C = 4.0 eV for TiO2[78] was adopted in this work. The E O X d y e * can be calculated as following [79]:
E O X d y e * = E O X d y e - λ max
in which E O X d y e is the redox potential of the ground state and λmax is the absorption maximum with IMCT character. Whereas the free energy change of dye regeneration ( Δ G dye regen ) can affect on the rate constant of redox process between the oxidized dyes and electrolyte. The Δ G dye regen can be calculated as
Δ G d y e r e g e n = E O X d y e - E r e d o x e l e c t r o l y t e
where E redox electrolyte is the redox potential of electrolyte. The E redox electrolyte of commonly used redox couple iodide/triiodide is about 4.85 eV (0.35 V vs. NHE) [80]. The calculated E O X d y e , E O X d y e *, ΔGinject, and Δ G dye regen for tetrahydroquinoline dyes are listed in Table 8.
In terms of the calculated results, the negative ΔGinject of the tetrahydroquinoline dyes indicates that the excited state with IMCT character lies above the TiO2 conduction band edge. For C1-n (n = 1–5) dyes, the absolute value of ΔGinject for C1-3 is larger than that of other dyes, and the absolute value of ΔGinject increases with the elongating of conjugate bridges with thiophene units. The ΔGinject of C1-4 which has dithienothiophene in conjugate linker is almost as same as that of C1-3. In addition, the data of C1-5 means that inserting a vinyl unit between bithiophene in conjugate bridges induces a slight decreasing of ΔGinject (about 0.02 eV). Furthermore, with elongating conjugate bridges, the Δ G dye regen of the dyes decrease in the corresponding order. So, introducing longer conjugate bridges in dyes can suppress electron recombination. For C2-n (n = 1–4) dyes, the variation of ΔGinject is only about 0.05 eV (from -0.59 to -0.64 eV). The C2-3 and C2-4 have largest absolute value of ΔGinject, and the dependence of Δ G dye regen on the number of thiophene unit in conjugate bridge is similar to that of C1-n (n = 1–4) dyes. Comparing the data of C1-n and C2-n (n = 1–4), it is found that inserting vinyl units between quinoline and thiophene generate a reduction of absolute value ΔGinject of and Δ G dye regen.

4. Conclusions

Conjugate bridge of organic dye sensitizer for solar cells can modify their electronic structure based properties, and therefore the performance in DSCs. To understand the role of conjugate bridge in the modification of electronic structure based properties, the computations of the geometries and electronic structures for 10 kinds of tetrahydroquinoline dyes were performed using DFT, and the electronic absorption and fluorescence properties were investigated via TDFT with PCM for solvent effects. The population analysis, molecular orbital energies, radiative lifetimes, EBE, LHE, ΔGinject, and Δ G dye regen were also addressed.
The population analysis of NBO results indicates that more charges populated in acceptor groups correspond to larger Voc. This means introducing strong electron-withdrawing group in the acceptor of dye sensitizer is a possible way to increase Voc of DSC. The absorption properties and MO analysis indicate that the quinoline groups are effective chromophores in IMCT, and they take the function of electron injection in DSCs.
The elongating of a conjugate bridge generates the larger oscillator strength, higher LHE, larger the absolute value of ΔGinject, and longer the relative radiative lifetime, as well as the decreasing of EBE and the Δ G dye regen. So extending the conjugate bridge with thiopene units in D-π-A organic dye is an effective way to increase the harvest of solar light, also it is favorable for fast electron injection due to larger ΔGinject. While the inversely correlated relationship between EBE and LHE imply that the dyes with lower EBE have more efficient light harvesting. The decreasing of Δ G dye regen with the elongating conjugate bridge implies that introducing longer conjugate bridges can suppress electron recombination.

Acknowledgments

This work supported by the Basic Scientific Research Foundation for Gansu Universities of China (Grant No. 1210ZTC055), National Natural Science Foundation of China (Grant Nos. 11164016, 11164015), and the Scientific Developmental Foundation of Lanzhou University of Technology.

References

  1. Oregan, B.; Gratzel, M. A low-cost, high-efficiency solar-cell based on dye-sensitized colloidal TiO2 films. Nature 1991, 353, 737–740. [Google Scholar]
  2. Gratzel, M. Photoelectrochemical cells. Nature 2001, 414, 338–344. [Google Scholar]
  3. Gratzel, M. Dye-sensitized solar cells. J. Photochem. Photobiol. C 2003, 4, 145–153. [Google Scholar]
  4. Nazeeruddin, M.K.; Klein, C.; Liska, P.; Gratzel, M. Synthesis of novel ruthenium sensitizers and their application in dye-sensitized solar cells. Coord. Chem. Rev 2005, 249, 1460–1467. [Google Scholar]
  5. Li, B.; Wang, L.; Kang, B.; Wang, P.; Qiu, Y. Review of recent progress in solid-state dye-sensitized solar cells. Sol. Energy Mater. Sol. Cells 2006, 90, 549–573. [Google Scholar]
  6. Gratzel, M. Recent advances in sensitized mesoscopic solar cells. Acc. Chem. Res 2009, 42, 1788–1798. [Google Scholar]
  7. Mishra, A.; Fischer, M.K.R.; Bauerle, P. Metal-free organic dyes for dye-sensitized solar cells: From structure: Property relationships to design rules. Angew. Chem. Int. Ed 2009, 48, 2474–2499. [Google Scholar]
  8. Ning, Z.J.; Tian, H. Triarylamine: A promising core unit for efficient photovoltaic materials. Chem. Commun. 2009, 5483–5495. [Google Scholar]
  9. Robertson, N. Optimizing dyes for dye-sensitized solar cells. Angew. Chem. Int. Ed 2006, 45, 2338–2345. [Google Scholar]
  10. Zakeeruddin, S.M.; Gratzel, M. Solvent-free ionic liquid electrolytes for mesoscopic dye-sensitized solar cells. Adv. Funct. Mater 2009, 19, 2187–2202. [Google Scholar]
  11. Clifford, J.N.; Martinez-Ferrero, E.; Viterisi, A.; Palomares, E. Sensitizer molecular structure-device efficiency relationship in dye sensitized solar cells. Chem. Soc. Rev 2011, 40, 1635–1646. [Google Scholar]
  12. Chen, X.Y.; Guo, J.H.; Peng, X.J.; Guo, M.; Xu, Y.Q.; Shi, L.; Liang, C.L.; Wang, L.; Gao, Y.L.; Sun, S.G.; Cai, S.M. Novel cyanine dyes with different methine chains as sensitizers for nanocrystalline solar cell. J. Photochem. Photobiol. A 2005, 171, 231–236. [Google Scholar]
  13. Sayama, K.; Tsukagoshi, S.; Mori, T.; Hara, K.; Ohga, Y.; Shinpou, A.; Abe, Y.; Suga, S.; Arakawa, H. Efficient sensitization of nanocrystalline TiO2 films with cyanine and merocyanine organic dyes. Sol. Energy Mater. Sol. Cells 2003, 80, 47–71. [Google Scholar]
  14. Wu, W.J.; Hua, J.L.; Jin, Y.H.; Zhan, W.H.; Tian, H. Photovoltaic properties of three new cyanine dyes for dye-sensitized solar cells. Photochem. Photobiol. Sci 2008, 7, 63–68. [Google Scholar]
  15. Chen, Y.S.; Li, C.; Zeng, Z.H.; Wang, W.B.; Wang, X.S.; Zhang, B.W. Efficient electron injection due to a special adsorbing group’s combination of carboxyl and hydroxyl: Dye-sensitized solar cells based on new hemicyanine dyes. J. Mater. Chem 2005, 15, 1654–1661. [Google Scholar]
  16. Wang, Z.S.; Li, F.Y.; Huang, C.H. Photocurrent enhancement of hemicyanine dyes containing RSO3 group through treating TiO2 films with hydrochloric acid. J. Phys. Chem. B 2001, 105, 9210–9217. [Google Scholar]
  17. Liang, M.; Xu, W.; Cai, F.S.; Chen, P.Q.; Peng, B.; Chen, J.; Li, Z.M. New triphenylamine-based organic dyes for efficient dye-sensitized solar cells. J. Phys. Chem. C 2007, 111, 4465–4472. [Google Scholar]
  18. Shen, P.; Tang, Y.H.; Jiang, S.H.; Chen, H.J.; Zheng, X.Y.; Wang, X.Y.; Zhao, B.; Tan, S.T. Efficient triphenylamine-based dyes featuring dual-role carbazole, fluorene and spirobifluorene moieties. Org. Electron 2011, 12, 125–135. [Google Scholar]
  19. Cappel, U.B.; Karlsson, M.H.; Pschirer, N.G.; Eickemeyer, F.; Schoneboom, J.; Erk, P.; Boschloo, G.; Hagfeldt, A. A broadly absorbing perylene dye for solid-state dye-sensitized solar cells. J. Phys. Chem. C 2009, 113, 14595–14597. [Google Scholar]
  20. Jiao, C.J.; Zu, N.N.; Huang, K.W.; Wang, P.; Wu, J.S. Perylene anhydride fused porphyrins as near-infrared sensitizers for dye-sensitized solar cells. Org. Lett 2011, 13, 3652–3655. [Google Scholar]
  21. Li, C.; Schoneboom, J.; Liu, Z.H.; Pschirer, N.G.; Erk, P.; Herrmann, A.; Mullen, K. Rainbow perylene monoimides: Easy control of optical properties. Chem. Eur. J 2009, 15, 878–884. [Google Scholar]
  22. Li, C.; Yum, J.H.; Moon, S.J.; Herrmann, A.; Eickemeyer, F.; Pschirer, N.G.; Erk, P.; Schoeboom, J.; Mullen, K.; Gratzel, M.; et al. An improved perylene sensitizer for solar cell applications. ChemSusChem 2008, 1, 615–618. [Google Scholar]
  23. Tian, H.; Liu, P.H.; Zhu, W.H.; Gao, E.Q.; Da-Jun, W.A.; Cai, S.M. Synthesis of novel multi-chromophoric soluble perylene derivatives and their photosensitizing properties with wide spectral response for SnO2 nanoporous electrode. J. Mater. Chem 2000, 10, 2708–2715. [Google Scholar]
  24. Hara, K.; Sato, T.; Katoh, R.; Furube, A.; Ohga, Y.; Shinpo, A.; Suga, S.; Sayama, K.; Sugihara, H.; Arakawa, H. Molecular design of coumarin dyes for efficient dye-sensitized solar cells. J. Phys. Chem. B 2003, 107, 597–606. [Google Scholar]
  25. Koops, S.E.; Barnes, P.R.F.; O’Regan, B.C.; Durrant, J.R. Kinetic competition in a coumarin dye-sensitized solar cell: Injection and recombination limitations upon device performance. J. Phys. Chem. C 2010, 114, 8054–8061. [Google Scholar]
  26. Li, X.G.; Lu, H.J.; Wang, S.R.; Guo, J.J.; Li, J. Sensitizers of dye-sensitized solar cells. Prog. Chem 2011, 23, 569–588. [Google Scholar]
  27. Campbell, W.M.; Burrell, A.K.; Officer, D.L.; Jolley, K.W. Porphyrins as light harvesters in the dye-sensitised TiO2 solar cell. Coord. Chem. Rev 2004, 248, 1363–1379. [Google Scholar]
  28. Hsieh, C.P.; Lu, H.P.; Chiu, C.L.; Lee, C.W.; Chuang, S.H.; Mai, C.L.; Yen, W.N.; Hsu, S.J.; Diau, E.W.G.; Yeh, C.Y. Synthesis and characterization of porphyrin sensitizers with various electron-donating substituents for highly efficient dye-sensitized solar cells. J. Mater. Chem 2010, 20, 1127–1134. [Google Scholar]
  29. Imahori, H.; Umeyama, T.; Ito, S. Large pi-aromatic molecules as potential sensitizers for highly efficient dye-sensitized solar cells. Acc. Chem. Res 2009, 42, 1809–1818. [Google Scholar]
  30. Lee, C.W.; Lu, H.P.; Lan, C.M.; Huang, Y.L.; Liang, Y.R.; Yen, W.N.; Liu, Y.C.; Lin, Y.S.; Diau, E.W.G.; Yeh, C.Y. Novel zinc porphyrin sensitizers for dye-sensitized solar cells: Synthesis and spectral, electrochemical, and photovoltaic properties. Chem. Eur. J 2009, 15, 1403–1412. [Google Scholar]
  31. Pasunooti, K.K.; Song, J.L.; Chai, H.; Amaladass, P.; Deng, W.Q.; Liu, X.W. Synthesis, characterization and application of trans-D-B-A-porphyrin based dyes in dye-sensitized solar cells. J. Photochem. Photobiol. A 2011, 218, 219–225. [Google Scholar]
  32. Wu, S.L.; Lu, H.P.; Yu, H.T.; Chuang, S.H.; Chiu, C.L.; Lee, C.W.; Diau, E.W.G.; Yeh, C.Y. Design and characterization of porphyrin sensitizers with a push-pull framework for highly efficient dye-sensitized solar cells. Energy Environ. Sci 2010, 3, 949–955. [Google Scholar]
  33. Geiger, T.; Kuster, S.; Yum, J.H.; Moon, S.J.; Nazeeruddin, M.K.; Gratzel, M.; Nuesch, F. Molecular design of unsymmetrical squaraine dyes for high efficiency conversion of low energy photons into electrons using TiO2 nanocrystalline films. Adv. Funct. Mater 2009, 19, 2720–2727. [Google Scholar]
  34. Paek, S.; Choi, H.; Kim, C.; Cho, N.; So, S.; Song, K.; Nazeeruddin, M.K.; Ko, J. Efficient and stable panchromatic squaraine dyes for dye-sensitized solar cells. Chem. Commun 2011, 47, 2874–2876. [Google Scholar]
  35. Silvestri, F.; Lopez-Duarte, I.; Seitz, W.; Beverina, L.; Martinez-Diaz, M.V.; Marks, T.J.; Guldi, D.M.; Pagani, G.A.; Torres, T. A squaraine-phthalocyanine ensemble: Towards molecular panchromatic sensitizers in solar cells. Chem. Commun. 2009, 4500–4502. [Google Scholar]
  36. Cheng, H.M.; Hsieh, W.F. Electron transfer properties of organic dye-sensitized solar cells based on indoline sensitizers with ZnO nanoparticles. Nanotechnology 2010, 21. [Google Scholar] [CrossRef]
  37. Kuang, D.; Uchida, S.; Humphry-Baker, R.; Zakeeruddin, S.M.; Gratzel, M. Organic dye-sensitized ionic liquid based solar cells: Remarkable enhancement in performance through molecular design of indoline sensitizers. Angew. Chem. Int. Ed 2008, 47, 1923–1927. [Google Scholar]
  38. Zhang, X.H.; Li, C.; Wang, W.B.; Cheng, X.X.; Wang, X.S.; Zhang, B.W. Photophysical, electrochemical, and photoelectrochemical properties of new azulene-based dye molecules. J. Mater. Chem 2007, 17, 642–649. [Google Scholar]
  39. 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. [Google Scholar]
  40. Wong, B.M.; Cordaro, J.G. Coumarin dyes for dye-sensitized solar cells: A long-range-corrected density functional study. J. Chem. Phys. 2008, 129, 214703:1–214703:8. [Google Scholar]
  41. De Angelis, F.; Fantacci, S.; Selloni, A.; Nazeeruddin, M.K.; Grätzel, M. Time-dependent density functional theory investigations on the excited states of Ru(II)-Dye-Sensitized TiO2 nanoparticles: The role of sensitizer protonation. J. Am. Chem. Soc 2007, 129, 14156–14157. [Google Scholar]
  42. Kim, B.-G.; Zhen, C.-G.; Jeong, E.J.; Kieffer, J.; Kim, J. Organic dye design tools for efficient photocurrent generation in dye-sensitized solar cells: Exciton binding energy and electron acceptors. Adv. Funct. Mater 2012, 22, 1606–1612. [Google Scholar]
  43. Fan, W.; Tan, D.; Deng, W. Theoretical investigation of triphenylamine dye/titanium dioxide interface for dye-sensitized solar cells. Phys. Chem. Chem. Phys 2011, 13, 16159–16167. [Google Scholar]
  44. Fan, W.; Tan, D.; Deng, W.-Q. Acene-modified triphenylamine dyes for dye-sensitized solar cells: A computational study. ChemPhysChem 2012, 13, 2051–2060. [Google Scholar]
  45. Tai, C.K.; Chen, Y.J.; Chang, H.W.; Yeh, P.L.; Wang, B.C. DFT and TD-DFT investigations of metal-free dye sensitizers for solar cells: Effects of electron donors and π-conjugated linker. Comput. Theor. Chem 2011, 971, 42–50. [Google Scholar]
  46. Chen, R.; Yang, X.; Tian, H.; Sun, L. Tetrahydroquinoline dyes with different spacers for organic dye-sensitized solar cells. J. Photochem. Photobiol. A 2007, 189, 295–300. [Google Scholar]
  47. Chen, R.; Yang, X.; Tian, H.; Wang, X.; Hagfeldt, A.; Sun, L. Effect of tetrahydroquinoline dyes structure on the performance of organic dye-sensitized solar cells. Chem. Mater 2007, 19, 4007–4015. [Google Scholar]
  48. Jacquemin, D.; Perpète, E.A.; Ciofini, I.; Adamo, C. Accurate simulation of optical properties in dyes. Acc. Chem. Res 2008, 42, 326–334. [Google Scholar]
  49. Pastore, M.; Mosconi, E.; de Angelis, F.; Grätzel, M. A computational investigation of organic dyes for dye-sensitized solar cells: Benchmark, strategies, and open issues. J. Phys. Chem. C 2010, 114, 7205–7212. [Google Scholar]
  50. Cossi, M.; Rega, N.; Scalmani, G.; Barone, V. Energies, structures, and electronic properties of molecules in solution with the C-PCM solvation model. J. Comput. Chem 2003, 24, 669–681. [Google Scholar]
  51. Barone, V.; Cossi, M. Quantum calculation of molecular energies and energy gradients in solution by a conductor solvent model. J. Phys. Chem. A 1998, 102, 1995–2001. [Google Scholar]
  52. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. [Phys. Rev. Lett. 77, 3865 (1996)]. Phys. Rev. Lett 1997, 78, 1396. [Google Scholar]
  53. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett 1996, 77, 3865–3868. [Google Scholar]
  54. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys 1999, 110, 6158–6170. [Google Scholar]
  55. Zhao, Y.; Truhlar, D.G. Comparative DFT study of van der waals complexes: Rare-gas dimers, alkaline-earth dimers, zinc dimer, and zinc-rare-gas dimers. J. Phys. Chem. A 2006, 110, 5121–5129. [Google Scholar]
  56. Yanai, T.; Tew, D.P.; Handy, N.C. A new hybrid exchange–correlation functional using the Coulomb-attenuating method (CAM-B3LYP). Chem. Phys. Lett 2004, 393, 51–57. [Google Scholar]
  57. Vydrov, O.A.; Scuseria, G.E.; Perdew, J.P. Tests of functionals for systems with fractional electron number. J. Chem. Phys. 2007, 126, 154109:1–154109:9. [Google Scholar]
  58. Vydrov, O.A.; Scuseria, G.E. Assessment of a long-range corrected hybrid functional. J. Chem. Phys. 2006, 125, 234109:1–234109:9. [Google Scholar]
  59. Vydrov, O.A.; Heyd, J.; Krukau, A.V.; Scuseria, G.E. Importance of short-range versus long-range Hartree-Fock exchange for the performance of hybrid density functionals. J. Chem. Phys. 2006, 125, 074106:1–074106:9. [Google Scholar]
  60. Tawada, Y.; Tsuneda, T.; Yanagisawa, S.; Yanai, T.; Hirao, K. A long-range-corrected time-dependent density functional theory. J. Chem. Phys 2004, 120, 8425–8433. [Google Scholar]
  61. Chai, J.D.; Head-Gordon, M. Systematic optimization of long-range corrected hybrid density functionals. J. Chem. Phys. 2008, 128, 084106:1–084106:15. [Google Scholar]
  62. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian; Gaussian Inc: Wallingford, CT, USA, 2010. [Google Scholar]
  63. Kim, S.; Lee, J.K.; Kang, S.O.; Ko, J.; Yum, J.H.; Fantacci, S.; de Angelis, F.; di Censo, D.; Nazeeruddin, M.K.; Gratzel, M. Molecular engineering of organic sensitizers for solar cell applications. J. Am. Chem. Soc 2006, 128, 16701–16707. [Google Scholar]
  64. Zhang, C.R.; Wu, Y.Z.; Chen, Y.H.; Chen, H.S. Geometries, electronic structures and related properties of organic dye sensitizers JK16 and JK17. Acta Phys. Chim. Sin 2009, 25, 53–60. [Google Scholar]
  65. Choi, H.; Lee, J.K.; Song, K.; Kang, S.O.; Ko, J. Novel organic dyes containing bis-dimethylfluorenyl amino benzo[b]thiophene for highly efficient dye-sensitized solar cell. Tetrahedron 2007, 63, 3115–3121. [Google Scholar]
  66. Zhang, C.R.; Liu, Z.J.; Chen, Y.H.; Chen, H.S.; Wu, Y.Z.; Yuan, L.H. DFT and TDDFT study on organic dye sensitizers D5, DST and DSS for solar cells. J. Mol. Struc. Theochem 2009, 899, 86–93. [Google Scholar]
  67. Zhang, C.R.; Liu, Z.J.; Chen, Y.H.; Chen, H.S.; Wu, Y.Z.; Feng, W.J.; Wang, D.B. DFT and TD-DFT study on structure and properties of organic dye sensitizer TA-St-CA. Curr. Appl. Phys 2010, 10, 77–83. [Google Scholar]
  68. Jung, I.; Lee, J.K.; Song, K.H.; Song, K.; Kang, S.O.; Ko, J. Synthesis and photovoltaic properties of efficient organic dyes containing the benzo b furan moiety for solar cells. J. Org. Chem 2007, 72, 3652–3658. [Google Scholar]
  69. Casanova, D. The role of the p Linker in Donor-p-Acceptor organic dyes for high-performance sensitized solar cells. ChemPhysChem 2011, 12, 2979–2988. [Google Scholar]
  70. Zhang, C.-R.; Liu, L.; Liu, Z.-J.; Shen, Y.-L.; Sun, Y.-T.; Wu, Y.-Z.; Chen, Y.-H.; Yuan, L.-H.; Wang, W.; Chen, H.-S. Electronic structures and optical properties of organic dye sensitizer NKX derivatives for solar cells: A theoretical approach. J. Mol. Graphics Modell 2012, 38, 419–429. [Google Scholar]
  71. Nazeeruddin, M.K.; Kay, A.; Rodicio, I.; Humphrybaker, R.; Muller, E.; Liska, P.; Vlachopoulos, N.; Gratzel, M. Conversion of light to electricity by cis-X2bis(2,2′-bipyridyl-4,4′-dicarboxylate)Ruthenium(II) charge-transfer sensitizers (X = Cl, Br, I, CN, and SCN) on nanocrystalline TiO2 electrodes. J. Am. Chem. Soc 1993, 115, 6382–6390. [Google Scholar]
  72. Kurashige, Y.; Nakajima, T.; Kurashige, S.; Hirao, K.; Nishikitani, Y. Theoretical investigation of the excited states of coumarin dyes for dye-sensitized solar cells. J. Phys. Chem. A 2007, 111, 5544–5548. [Google Scholar]
  73. Preat, J.; Michaux, C.; Jacquemin, D.; Perpete, E.A. Enhanced efficiency of organic dye-sensitized solar cells: Triphenylamine derivatives. J. Phys. Chem. C 2009, 113, 16821–16833. [Google Scholar]
  74. Gregg, B.A. Excitonic solar cells. J. Phys. Chem. B 2003, 107, 4688–4698. [Google Scholar]
  75. Scholes, G.D.; Rumbles, G. Excitons in nanoscale systems. Nat. Mater 2006, 5, 683–696. [Google Scholar]
  76. Le Bahers, T.; Pauporte, T.; Scalmani, G.; Adamo, C.; Ciofini, I. A TD-DFT investigation of ground and excited state properties in indoline dyes used for dye-sensitized solar cells. Phys. Chem. Chem. Phys 2009, 11, 11276–11284. [Google Scholar]
  77. Marcus, R.A. Electron transfer reactions in chemistry. Theory and experiment. Rev. Mod. Phys 1993, 65, 599–610. [Google Scholar]
  78. Asbury, J.B.; Wang, Y.-Q.; Hao, E.; Ghosh, H.N.; Lian, T. Evidences of hot excited state electron injection from sensitizer molecules to TiO2 nanocrystalline thin films. Res. Chem. Intermed 2001, 27, 393–406. [Google Scholar]
  79. Katoh, R.; Furube, A.; Yoshihara, T.; Hara, K.; Fujihashi, G.; Takano, S.; Murata, S.; Arakawa, H.; Tachiya, M. Efficiencies of electron injection from excited N3 dye into nanocrystalline semiconductor (ZrO2, TiO2, ZnO, Nb2O5, SnO2, In2O3) films. J. Phys. Chem. B 2004, 108, 4818–4822. [Google Scholar]
  80. Boschloo, G.; Hagfeldt, A. Characteristics of the iodide/triiodide redox mediator in dye-sensitized solar cells. Acc. Chem. Res 2009, 42, 1819–1826. [Google Scholar]
Figure 1. The optimized geometrical structures of dyes (a) C1-1, (b) C1-2, (c) C1-3, (d) C1-4, and (e) C1-5. (CAM-B3LYP/6-31G(d,p) in gas phase, the blue colored spheres: N; the red colored spheres: O; the yellow colored spheres: S; the other dark colored spheres: C; the hydrogen atoms have been omitted for clarity.)
Figure 1. The optimized geometrical structures of dyes (a) C1-1, (b) C1-2, (c) C1-3, (d) C1-4, and (e) C1-5. (CAM-B3LYP/6-31G(d,p) in gas phase, the blue colored spheres: N; the red colored spheres: O; the yellow colored spheres: S; the other dark colored spheres: C; the hydrogen atoms have been omitted for clarity.)
Ijms 14 05461f1
Figure 2. The optimized geometrical structures of dyes (a) C2-1, (b) C2-2, (c) C2-3, (d) C2-4, and (e) C3-1. (CAM-B3LYP/6-31G(d,p) in gas phase, the blue colored spheres: N; the red colored spheres: O; the yellow colored spheres: S; the other dark colored spheres: C; the hydrogen atoms have been omitted for clarity.)
Figure 2. The optimized geometrical structures of dyes (a) C2-1, (b) C2-2, (c) C2-3, (d) C2-4, and (e) C3-1. (CAM-B3LYP/6-31G(d,p) in gas phase, the blue colored spheres: N; the red colored spheres: O; the yellow colored spheres: S; the other dark colored spheres: C; the hydrogen atoms have been omitted for clarity.)
Ijms 14 05461f2
Figure 3. The open-circuit photovoltage (Voc), short-circuit photocurrent density (Jsc), and solar-to-electrical energy conversion efficiencies (η) versus the length of conjugate bridges. (a) C1-n dyes, n = 1–5; (b) C2-m dyes, m = 1–4.
Figure 3. The open-circuit photovoltage (Voc), short-circuit photocurrent density (Jsc), and solar-to-electrical energy conversion efficiencies (η) versus the length of conjugate bridges. (a) C1-n dyes, n = 1–5; (b) C2-m dyes, m = 1–4.
Ijms 14 05461f3
Figure 4. The open-circuit voltage Vocversus the charges of acceptor groups in tetrahydroquinoline dyes.
Figure 4. The open-circuit voltage Vocversus the charges of acceptor groups in tetrahydroquinoline dyes.
Ijms 14 05461f4
Figure 5. The calculated frontier molecular orbital energies of tetrahydroquinoline dyes (CAM-B3LYP/6-31G(d,p)).
Figure 5. The calculated frontier molecular orbital energies of tetrahydroquinoline dyes (CAM-B3LYP/6-31G(d,p)).
Ijms 14 05461f5
Figure 6. The calculated light harvesting efficiencies (LHE) and the exciton binding energies (EBE, in eV) of tetrahydroquinoline dyes.
Figure 6. The calculated light harvesting efficiencies (LHE) and the exciton binding energies (EBE, in eV) of tetrahydroquinoline dyes.
Ijms 14 05461f6
Table 1. The chemical structures of the tetrahydroquinoline dyes.
Table 1. The chemical structures of the tetrahydroquinoline dyes.
Ijms 14 05461f7
DyesX=Y=DyesX=Y=
C1-1CH3Ijms 14 05461f8C2-1CH3Ijms 14 05461f9
C1-2CH3Ijms 14 05461f10C2-2CH3Ijms 14 05461f11
C1-3CH3Ijms 14 05461f12C2-3CH3Ijms 14 05461f13
C1-4CH3Ijms 14 05461f14C2-4CH3Ijms 14 05461f15
C1-5CH3Ijms 14 05461f16C3-1PhIjms 14 05461f17
Table 2. Experimental absorption maxima, the computed excitation energies λmax (nm/eV) and oscillator strength f of the lowest excited state for the dye C1-1 in ethanol solution, provided by PCM-TDDFT/ 6-31G (d,p).
Table 2. Experimental absorption maxima, the computed excitation energies λmax (nm/eV) and oscillator strength f of the lowest excited state for the dye C1-1 in ethanol solution, provided by PCM-TDDFT/ 6-31G (d,p).
PBE0CAM-B3LYPLC-ωPBEωB97XM062XExperiment
λmax569/2.18472/2.63395/3.14414/2.99474/2.62468/2.65
f1.36041.42301.41891.41481.3526
Table 3. The charge populations of the ground state (S0) for tetrahydroquinoline dyes.
Table 3. The charge populations of the ground state (S0) for tetrahydroquinoline dyes.
DyesDonating groupπ-Conjugated linkerAcceptor group
C1-10.1550.025−0.180
C1-20.1310.039−0.170
C1-30.1150.047−0.162
C1-40.1380.027−0.164
C1-50.1260.043−0.169
C2-10.202−0.015−0.187
C2-20.1630.009−0.172
C2-30.1490.015−0.164
C2-40.171−0.005−0.165
C3-10.1510.029−0.180
Table 4. The experimental and calculated electronic absorption λmax (nm/eV), as well as the λmax errors (nm/eV) between the experiment and the calculation of tetrahydroquinoline dyes.
Table 4. The experimental and calculated electronic absorption λmax (nm/eV), as well as the λmax errors (nm/eV) between the experiment and the calculation of tetrahydroquinoline dyes.
DyesExperimental λmaxCalculated λmaxerror
C1-1468/2.65472/2.636/0.02
C1-2472/2.63483/2.5511/0.08
C1-3475/2.61476/2.611/0.00
C1-4467/2.65470/2.643/0.01
C1-5492/2.52499/2.487/0.04
C2-1441/2.81434/2.867/0.05
C2-2462/2.68457/2.715/0.03
C2-3455/2.72460/2.695/0.03
C2-4444/2.79441/2.813/0.02
C2-5470/2.64465/2.675/0.03
Table 5. The calculated excitation energies (eV/nm), electronic transition configurations and oscillator strengths (f) for tetrahydroquinoline dyes in solution (CAM-B3LYP/6-31G(d,p)).
Table 5. The calculated excitation energies (eV/nm), electronic transition configurations and oscillator strengths (f) for tetrahydroquinoline dyes in solution (CAM-B3LYP/6-31G(d,p)).
DyesStateConfigurations composition with |coeff.| > 0.2 (transition oribitals)Excitation energyf
C1-110.6617(H → L)2.63/4721.4230
20.6370(H − 1 → L)3.97/3130.1305

C1-210.6080(H → L); -0.2892(H − 1 → L)2.55/4831.7838
20.5426(H − 1 → L); −0.3226(H → L + 2); −0.2109(H − 3 → L)3.60/3440.0022
30.5499(H → L + 1); 0.2860(H → L); 0.2298(H − 1 → L)3.90/3180.3394

C1-310.5272(H → L); −0.3657(H − 1 → L); −0.2453(H → L + 1)2.61/4762.0586
20.4624(H → L + 1); −0.4117(H − 1 → L); 0.2571(H − 2 → L)3.37/3670.2063
30.4336(H → L); 0.3773(H → L + 1); 0.2856(H − 1 → L)3.70/3350.3175

C1-410.6138(H → L); −0.2753(H − 1 → L)2.64/4701.8212
20.5204(H − 1 → L); −0.3108(H → L + 1); 0.2258(H − 2 → L)3.65/3400.1647
30.6496(H − 2 → L)3.83/3240.1734
40.5542(H → L + 1);0.2836(H → L); 0.2423(H − 1 → L)3.98/3120.3454

C1-510.5885(H → L); −0.3119(H − 1 → L)2.48/4991.9599
20.5146(H − 1 → L); 0.3505(H → L + 1); 0.2427(H − 2 → L)3.43/3610.0147
30.5168(H → L + 1); −0.3192(H → L); −0.2396(H − 1 → L)3.74/3310.3021

C2-110.6732(H → L); −0.1707(H − 1 → L)2.86/4341.1582

C2-210.6210(H → L); −0.2866(H − 1 → L)2.71/4571.4695
20.5942(H − 1 → L); 0.2351(H → L)3.80/3260.0264

C2-310.5470(H → L); −0.3771(H − 1 → L)2.69/4601.7513
20.4771(H − 1 → L); 0.3489(H → L + 1); 0.26105(H − 2 → L); 0.2354(H → L)3.57/3470.0684
30.4921(H → L + 1); −0.3547(H → L + 1); −0.20695(H − 1 → L)3.89/3190.3308

C2-410.6227(H → L); −0.2849(H − 1 → L)2.81/4411.5097
20.5318(H − 2 → L); −0.3719(H − 1 → L)3.80/3270.0072
30.4731(H − 1 → L);0.4340(H − 2 → L)3.93/3160.2227

C3-110.6618(H → L)2.67/4651.4440
20.6327(H − 1 → L)4.00/3100.1103
Table 6. Isodensity plots (isodensity contour = 0.02 a.u.) of the frontier molecular orbitals of the tetrahydroquinoline dyes.
Table 6. Isodensity plots (isodensity contour = 0.02 a.u.) of the frontier molecular orbitals of the tetrahydroquinoline dyes.
DyesHOMO − 1HOMOLUMOLUMO + 1
C1-1Ijms 14 05461f18Ijms 14 05461f19Ijms 14 05461f20Ijms 14 05461f21
C1-2Ijms 14 05461f22Ijms 14 05461f23Ijms 14 05461f24Ijms 14 05461f25
C1-3Ijms 14 05461f26Ijms 14 05461f27Ijms 14 05461f28Ijms 14 05461f29
C1-4Ijms 14 05461f30Ijms 14 05461f31Ijms 14 05461f32Ijms 14 05461f33
C1-5Ijms 14 05461f34Ijms 14 05461f35Ijms 14 05461f36Ijms 14 05461f37
C2-1Ijms 14 05461f38Ijms 14 05461f39Ijms 14 05461f40Ijms 14 05461f41
C2-2Ijms 14 05461f42Ijms 14 05461f43Ijms 14 05461f44Ijms 14 05461f45
C2-3Ijms 14 05461f46Ijms 14 05461f47Ijms 14 05461f48Ijms 14 05461f49
C2-4Ijms 14 05461f50Ijms 14 05461f51Ijms 14 05461f52Ijms 14 05461f53
C3-1Ijms 14 05461f54Ijms 14 05461f55Ijms 14 05461f56Ijms 14 05461f57
Table 7. Fluorescence maxima (λ in nm/eV), oscillator strength (f in a.u.) and relative radiative lifetimes ( τrel ) of tetrahydroquinoline dyes in solution with respect to the experimental results. The corresponding Stokes shifts (SS, in eV) are also reported.
Table 7. Fluorescence maxima (λ in nm/eV), oscillator strength (f in a.u.) and relative radiative lifetimes ( τrel ) of tetrahydroquinoline dyes in solution with respect to the experimental results. The corresponding Stokes shifts (SS, in eV) are also reported.
DyesComputedExperimental a

λFιrelSSλSS
C1-1549/2.261.61661.000.37647/1.920.73
C1-2640/1.942.12511.030.61670/1.850.78
C1-3701/1.772.35721.120.84615/2.020.59
C1-4618/2.012.17580.940.63643/1.930.72
C1-5687/1.802.37871.070.68656/1.890.63
C2-1488/2.541.25621.020.32594/2.090.72
C2-2583/2.131.74571.040.58652/1.900.78
C2-3662/1.872.11311.180.82652/1.900.82
C2-4558/2.221.77990.940.59612/2.030.76
C3-1551/2.251.65100.990.42635/1.950.69
athe experimental results were reported in [47].
Table 8. The calculated redox potential (in eV), ΔGinject (in eV), and Δ G dye regen (in eV) for the tetrahydroquinoline dyes in the solution.
Table 8. The calculated redox potential (in eV), ΔGinject (in eV), and Δ G dye regen (in eV) for the tetrahydroquinoline dyes in the solution.
Dye E O X d y e E O X d y e *ΔGinject Δ G dye regen
C1-16.093.46−0.541.24
C1-25.973.42−0.581.12
C1-35.963.35−0.651.11
C1-46.003.36−0.641.15
C1-55.923.44−0.561.07
C2-16.253.39−0.611.40
C2-26.123.41−0.591.27
C2-36.053.36−0.641.20
C2-46.173.36−0.641.32
C3-16.173.50−0.501.32

Share and Cite

MDPI and ACS Style

Zhang, C.-R.; Liu, L.; Zhe, J.-W.; Jin, N.-Z.; Ma, Y.; Yuan, L.-H.; Zhang, M.-L.; Wu, Y.-Z.; Liu, Z.-J.; Chen, H.-S. The Role of the Conjugate Bridge in Electronic Structures and Related Properties of Tetrahydroquinoline for Dye Sensitized Solar Cells. Int. J. Mol. Sci. 2013, 14, 5461-5481. https://doi.org/10.3390/ijms14035461

AMA Style

Zhang C-R, Liu L, Zhe J-W, Jin N-Z, Ma Y, Yuan L-H, Zhang M-L, Wu Y-Z, Liu Z-J, Chen H-S. The Role of the Conjugate Bridge in Electronic Structures and Related Properties of Tetrahydroquinoline for Dye Sensitized Solar Cells. International Journal of Molecular Sciences. 2013; 14(3):5461-5481. https://doi.org/10.3390/ijms14035461

Chicago/Turabian Style

Zhang, Cai-Rong, Li Liu, Jian-Wu Zhe, Neng-Zhi Jin, Yao Ma, Li-Hua Yuan, Mei-Lin Zhang, You-Zhi Wu, Zi-Jiang Liu, and Hong-Shan Chen. 2013. "The Role of the Conjugate Bridge in Electronic Structures and Related Properties of Tetrahydroquinoline for Dye Sensitized Solar Cells" International Journal of Molecular Sciences 14, no. 3: 5461-5481. https://doi.org/10.3390/ijms14035461

Article Metrics

Back to TopTop