DOI:
10.1039/C5RA00587F
(Paper)
RSC Adv., 2015,
5, 25079-25088
Molecular engineering of quinoxaline dyes toward more efficient sensitizers for dye-sensitized solar cells†
Received
11th January 2015
, Accepted 23rd February 2015
First published on 23rd February 2015
Abstract
D–A–π–A-featured organic dyes incorporating diphenylquinoxaline unit (such as IQ4) have shown great potential in anti-aggregation and broadening spectral response in the field of dye-sensitized solar cells (DSSCs). The crucial restriction for quinoxaline-based cell to attain higher efficiency is the relatively low photocurrent density (JSC). In the present work, three novel push–pull dyes only differing in electron donors, have been designed based on the IQ4 backbone, in order to further improve the light-harvesting capability of quinoxaline dyes and to examine the donor influence on dye performance. Theoretical analysis of the factors correlated with the JSC and open-circuit photovoltage (VOC) demonstrate that, relative to the parent IQ4 dye, the NIQ4 dye bearing the elegant N-annulated perylene donor shows a good performance in light harvesting, electron injection, and dye regeneration, indicating an increased JSC potential for the related cell. Furthermore, despite possessing a smaller vertical dipole moment, the improved blocking effect of NIQ4 not only prevents unfavorable self-aggregation, but also effectively inhibits the parasitic back-recombination. Therefore, the NIQ4 is proposed to be a potential dye in DSSC applications.
1. Introduction
Dye-sensitized solar cells (DSSC) are attracting widespread attention1–5 as an alternative inexpensive and environmentally friendly method for sunlight-to electricity conversion since the seminal work of O'Regan and Grätzel in 1991.6 The performance of DSSCs depends on the nature of sensitizer(s), photoanode, counter electrode, electrolyte, and the complicated interactions between the above components. In the past decades, great efforts have been made to explore the structure–property relationship7–12 and screen optimal materials for efficient DSSCs.13,14
The vital role in light harvesting and electron injection at dye/semiconductor interface has led to an explosion in dye sensitizer research. There are two types of dyes commonly used, namely, metal–organic15 and organic dyes.2,16 Particularly, organic dyes with D–π–A configuration are popular for their non-toxicity, high molar extinction coefficients, flexible molecular engineering, and material abundance. Experimentally, the organic dyes C259 and C239 co-sensitized DSSC has been reported to achieve the record efficiency of 12.8%.17 In addition, the D–A–π–A-featured organic dyes incorporating an additional electron-withdrawing unit have been designed to solve the self-aggregation problem between dyes and further broaden the spectral response.18–27 In this field, dyes with quinoxaline unit seem to be impressive.28–31 For instance, the DSSC based on IQ4 exhibits desirable short-circuit current density (Jsc) of 17.55 mA cm−2 as well as high open-circuit voltage (Voc) of 0.74 V with a fill factor (FF) of 0.71 under AM 1.5 illumination. This should be mainly ascribed to the strong anti-aggregation ability, the improved photostability and thermal stability of the 2,3-diphenylquinoxaline unit in IQ4.32 However, the relatively low photocurrent density restricts such cell to attain higher efficiency.33 As well-known, the nature of the electron donor within an organic dye not only affects the regeneration rate of an oxidized dye, the interaction between the dye-TiO2 layer and the electrolyte, but also plays a significant role in modulating molecular energy levels and light-harvesting capability of dyes.34,35 Therefore, the scrutiny of efficient organic dyes based on judicious selection of electron donors is much beneficial.36–41 Besides the commonly used bulky triarylamine-based donor2,42,43 and indoline moiety,44 favorable electron-donating ability of the donor has also been observed for JD21 containing the ullazine group,45 and WW-6 functionalized by the N-annulated perylene (NP) building block.46
In the present work, three donor moieties (triarylamine, ullazine, and NP groups, respectively) were introduced based on the IQ4 backbone, in order to further improve the light-harvesting ability of quinoxaline dyes, and to examine the donor influence on dye performance (Scheme 1). Theoretical analysis demonstrates that the solar cell based on NIQ4 with NP group is expected to exhibit a higher JSC and a comparable VOC, compared to the reference IQ4 dye. Therefore, the NIQ4 dye is proposed to be a promising candidate in DSSC applications.
 |
| Scheme 1 Chemical structures of the investigated dyes. | |
2. Computational details
All calculations have been performed with Gaussian 09 program package.47 The ground (S0) and the first excited-state (S1) geometries of the free and adsorbed dyes were optimized using the density functional theory (DFT) B3LYP functional48–50 coupled with the 6-31G(d, p) basis set for C, H, O, N, S atoms, and a LANL2DZ basis set for Ti atom. Frequency calculations were then carried out to confirm the nature of obtained minima at the same theory level as geometry optimizations. To get the reorganization energy (λ) which could affect the efficiency of electron injection, the cationic state of free dyes were optimized at the UB3LYP/6-31G(d, p) level. It has been reported that the B3LYP functional works very well in the description of the electronic structure51–56 but could not give an accurate representation of charge–transfer excited states,57 and this can be significantly improved by means of the long-range corrected functionals (e.g. CAM-B3LYP58).54–56 Therefore, the absorption spectra associated with the excited state of free and adsorbed dyes were simulated using the CAM-B3LYP functional with 6-31G(d, p) and LANL2DZ basis sets based on the optimized ground-state geometries, and the spectral curves were obtained within the SWizard program (http://www.sg-chem.net/swizard/)59 using the pseudo-Voigt model (a convolution of the Gaussian and Lorentzian functions) with full-width at half-height of 3000 cm−1. The charge transfer process intra-dyes and inter-dye/titanium surface at the maximum absorption (λmax) were qualitatively displayed by electron density differences (EDD) maps and quantitatively described by natural bond orbital (NBO) analysis. The corresponding charge transfer parameters were obtained by Multiwfn 3.3.6 program.60 All calculations above were carried out with bulk solvent effect of dichloromethane (CH2Cl2) included by using the conducting polarizable continuum model (C-PCM).61–63
3. Results and discussion
The power conversion efficiency (η) of a solar cell is determined by JSC, VOC, and fill factor (FF), as compared to the incident solar power (Pinc): |
 | (1) |
JSC in DSSCs can be described by:64–66
|
 | (2) |
In which, SI denotes the solar radiation intensity,
e is the unit charge,
h is the Planck constant,
c is the light speed in vacuum, and IPCE is monochromatic incident phonon to current conversion efficiency, it can be produced by using the following equation:
65,67 |
IPCE(λ) = LHE(λ)Φinjηregηcoll
| (3) |
Where, LHE(
λ) is the light-harvesting efficiency at a given wavelength,
Φinj represents the electron injection efficiency,
ηreg is the efficiency for dye regeneration, and
ηcoll denotes the charge collection efficiency. Note that IPCE is sometimes expressed as the product of LHE, injection efficiency, and charge collection efficiency.
37,68 Here, we separate the regeneration efficiency from the whole charge collection efficiency, because the complex charge collection efficiency is determined by the time constant for transport and recombination of the conduction band electrons injected into the nanocrystalline TiO
2 film,
65 which are not the scope of this work. We aim at figuring out the structural aspects related to regeneration efficiency.
As for VOC in DSSCs, it can be calculated by:69
|
 | (4) |
Here
q is the unit charge,
ECB is the conduction band edge (CBE) of the semiconductor,
k is the Boltzmann constant,
T is the absolute temperature,
nc is the number of electrons in the conduction band,
NCB is the density of accessible states in the conduction band,
68 and
Eredox is the electrolyte redox potential. The shift of
ECB (ΔCB) upon dye adsorption can be roughly expressed as:
70 |
 | (5) |
In this expression,
γ is the surface concentration of dyes,
μnormal denotes the dipole moment of individual dye molecules perpendicular to the surface of the semiconductor,
ε0 and
ε are the vacuum permittivity and the dielectric permittivity, respectively.
Based on these theoretical criterions, the parameters regulating JSC and VOC were separately evaluated in the following section to examine the electron donor influence on dye performance.
3.1 Evaluating Jsc
3.1.1 Light-harvesting efficiency of dye molecules. It is known that the electron-donating capacity of donor group plays an important role in determining the energetic positions of frontier molecular orbitals of dyes, which is correlated with the light-harvesting efficiency of a dye and the charge transfer kinetics at the titanium/dye/electrolyte interface, such as dye regeneration and charge recombination. Thus, at the first beginning of our study, the electron-donating capability of the four electron-donor groups (i.e. indoline, triarylamine, ullazine, and N-annulated perylene (NP) units) was evaluated by computing the energy levels of their highest occupied molecular orbitals (HOMOs). As depicted in Fig. S1,† the general order of the HOMO level for the four donors is NP (−4.93 eV) < indoline (−4.87 eV) < ullazine (−4.81 eV) < triarylamine (−4.77 eV). The HOMO in NP is the lowest in energy, indicating the weakest electron-donating ability of the NP unit. This characteristic of NP may be helpful in dye regeneration. Meanwhile, the triarylamine group exhibits the strongest electron-donating ability, which may narrow the energy gap of the frontier molecular orbitals and further red-shift the absorption spectrum of the corresponding dye, in comparison with dyes based on the other three groups.It is conceivable that the molecules investigated here should have several possible conformations, particularly in the π-linker of quinoxaline dyes. Taking IQ4 as the model dye, three possible trans-/cis-conformations (A, B, C) have been optimized at the B3LYP/6-31G(d, p) level. With the same theory level, frequency analysis was performed to confirm the nature of obtained minima. It is shown in Fig. 1 that IQ4-B with a linear π-linker is the most stable conformer with the lowest relative energy (ΔE = 0.00 kcal mol−1). This is consistent with an original intention in the design of dye sensitizers, that is, a developed sensitizer should stand perpendicularly on (rather than lie on) the surface of semiconductor. Based on this, the geometry optimizations of the newly developed push–pull dyes were carried out with the replacement of indoline unit in IQ4 by triarylamine (TIQ4), ullazine (UIQ4), and NP (NIQ4) donor moieties, respectively. The bond lengths and dihedral angles between the electron-donor group and the 2,3-diphenylquinoxaline segment are listed in Table S1.† There exist small differences (within 0.02 Å) in the distance between the two link atoms, and the largest torsional angle of −51.0° between the NP donor group and quinoxaline segment. This twist in NIQ4 may result in a larger localization of π-electron at donor moiety from HOMO and a decreased electron density overlap between HOMO and the lowest unoccupied molecular orbital (LUMO), endowing an adverse contribution on the absorption coefficient of the related dye.34
 |
| Fig. 1 The optimized geometries and relative energies of IQ4 isomers with respect to the IQ4-B species (ΔE is in kcal mol−1). | |
Based on the optimized geometries, time-dependent DFT (TDDFT) calculations were done at the CAM-B3LYP/6-31G(d, p) (LANL2DZ) level to obtain UV-Vis spectra and light-harvesting efficiency (LHE) of dyes. LHE at the certain wavelength (LHE(λ)) can be calculated according to:71,72
Where,
ε(
λ) is the molar absorption coefficient at given wavelength,
b is the thickness of TiO
2 film, and
c is the dye concentration on TiO
2 surface. Here,
b and
c are assumed to be 10 μm and 300 mmol L
−1, respectively.
64,73,74 The obtained results were shown in
Table 1 and
Fig. 2. Though the calculation (
λcalmax = 468 nm) significantly underestimates the absorption of the reference IQ4 dye (
λexpmax = 529 nm), we believe that the results obtained under the same level should give a useful indication to the optical property of several similar dyes. Furthermore, it is shown in
Fig. 2 that all designed dyes (TIQ4 and NIQ4) except for UIQ4 exhibit broad absorption in the visible region. This may eventually lead to a broad spectral response with high LHE values for TIQ4 and NIQ4. In this case, the NIQ4 dye functionalized by
N-annulated perylene group seems to be the winner molecule with the most red-shifted (
λcalmax = 477 nm) and broadest spectrum (the LHE reaches close to 100% in the range of 300–570 nm). The light-harvesting capability of dyes follows an order of NIQ4 > TIQ4 > IQ4 > UIQ4, which is not in line with the electron-donating ability of donor groups.
Table 1 The calculated maximum absorption wavelengths (λcalmax), oscillator strengths (f), transition natures of dyes (H = HOMO, L = LUMO), and electron density difference (EDD) maps between ground and the first excited state (the blue and purple zones correspond to density decrement and increment, respectively)
Dye |
λcalmax/nm |
f |
Assignment |
S0 → S1 |
The experimental value was taken from ref. 32. |
UIQ4 |
426 |
0.31 |
H → L (66%) |
 |
TIQ4 |
459 |
1.33 |
H → L (59%) |
 |
IQ4 |
468 |
1.28 |
H → L (69%) |
 |
IQ4a |
529 |
— |
— |
— |
NIQ4 |
477 |
1.43 |
H→ L (66%) |
 |
IQ4-TiO2 |
519 |
1.09 |
H → L (77%) |
 |
NIQ4-TiO2 |
544 |
0.55 |
H → L (81%) |
 |
 |
| Fig. 2 The absorption spectra (solid line) and light-harvesting efficiency curves (dash line) for IQ4 and its derivatives. | |
Further investigation shows that the λcalmax for all dyes is of S0 → S1 character (Table 1), which mainly corresponds to the electron transition from HOMO to LUMO. The composition75 and the spatial distribution of the related frontier orbitals have been shown in Table 2 and Fig. 3. It can be seen in Table 2 that the electron-donors contribute over 80% to HOMOs of dyes, and the π-linkers and acceptor parts are mainly responsible for LUMO. This may be ascribed to the relatively larger twisted angles between donor group and quinoxaline segment in dyes (Table S1†). Moreover, the donors of the newly designed dyes contribute more (>90%) to the HOMO than that of IQ4. The spatial distribution of the frontier orbitals in Fig. 3 is found to be obviously consistent with the orbital composition analysis, where the HOMOs of all dyes show localization mainly at electron-donor moiety, while LUMOs delocalize over the π-conjugated bridge and acceptor part. This will result in efficient charge separations, and can be visually described by the electronic density difference (EDD) plots in Table 1. From the EDD maps, we can see that the electron densities decrease particularly in electron-donor components and increase in the π-linkers and 2-cyanoacrylic acid anchor. In order to quantitatively describe the intramolecular charge transfer (CT) of dyes at λcalmax, the distance of charge transfer (D), the amount of transferred charge (Δq), and t parameter (defining the through space CT character of a transition) were calculated within Multiwfn 3.3.6 program, and have been summarized in Table 3. The optimal CT performance is obtained for NIQ4 that allows the transfer of 0.95 e− over 3.88 Å with good through space character (t = 3.33 Å).76 Furthermore, the computed energetic positions of LUMOs for all dyes are more positive than the CB of TiO2 (−4.0 eV vs. vacuum), ensuring an effective electron injection from the excited dye to titanium surface, and the HOMO of NIQ4 shows the largest amount of lowering compared to I−/I3− redox potential (−4.8 eV vs. vacuum), indicating a favorable dye regeneration process (Fig. 4).77
Table 2 Molecular orbital composition (in %) of the HOMO and LUMO orbitals for the investigated dyes
Dye |
Orbital |
Donor |
Quinoxaline |
Thiophene |
Anchor |
UIQ4 |
HOMO |
99.5 |
0.50 |
0.00 |
0.00 |
LUMO |
2.68 |
32.6 |
29.1 |
35.6 |
TIQ4 |
HOMO |
90.5 |
6.50 |
1.84 |
1.15 |
LUMO |
4.34 |
31.5 |
28.9 |
35.2 |
IQ4 |
HOMO |
85.0 |
10.1 |
3.00 |
1.88 |
LUMO |
4.36 |
31.1 |
29.2 |
35.4 |
NIQ4 |
HOMO |
91.0 |
6.19 |
1.72 |
1.08 |
LUMO |
4.61 |
31.9 |
28.8 |
34.6 |
 |
| Fig. 3 The frontier molecular orbitals of dyes corresponding to the maximum absorption. | |
Table 3 Ground and excited-state oxidation potentials, vertical transition energies, driving forces for dye regeneration and electron injection, reorganization energies, and charge transfer parameters of dyes
Dye |
EdyeOX/eV |
Eλmax/eV |
Edye*OX/eV |
ΔGreg/eV |
λreg/eV |
ΔGinj/eV |
qCT/e− |
dCT/Å |
t/Å |
The experimental values were taken from ref. 32. |
UIQ4 |
−4.86 |
−2.91 |
−1.95 |
0.06 |
0.22 |
2.05 |
0.80 |
2.10 |
6.83 |
TIQ4 |
−4.85 |
−2.70 |
−2.15 |
0.05 |
0.25 |
1.85 |
0.83 |
3.72 |
4.80 |
IQ4 |
−4.91 |
−2.65 |
−2.26 |
0.11 |
0.11 |
1.74 |
0.76 |
4.12 |
2.70 |
IQ4a |
−5.07 |
−2.34 |
−2.98 |
— |
— |
— |
— |
— |
— |
NIQ4 |
−4.94 |
−2.60 |
−2.34 |
0.14 |
0.14 |
1.66 |
0.95 |
3.88 |
3.33 |
 |
| Fig. 4 The energy-level arrangements of dyes together with the experimental conduction band edge of TiO2 and the redox potential of I−/I3−. | |
Based on the discussions for free dyes, the reference dye IQ4 and the designed dye NIQ4 were chosen to explore the interfacial interactions between dye molecules and semiconductor surface. To save the cost, a simple dye-Ti(OH)3·H2O model was adopted, which has been successfully employed to simulate the photoelectric properties of dye-TiO2 complex.38,68,78,79 Using the same methods as that for free dyes, the electronic and optical properties of dye-TiO2 complex were simulated. The related results have been given in Table 1 and Fig. 5. It can be found that both IQ4 and NIQ4 in bound state red-shift the absorption spectra and broaden the spectral response compared to those in isolated state. Though weaker in strength at λcalmax, the adsorbed NIQ4 exhibits more red-shift at λcalmax (544 nm) and absorbs more efficiently in solar spectrum (LHE close to 100% from 300 to 700 nm), suggesting a stronger electron coupling between the excited NIQ4 and the chelated Ti atom. The EDD plots in Table 1 give a clear picture for the charge carrier flux of adsorbed dyes at λcalmax, where electron densities accelerate mostly at the dye/titanium interface. The titanium surface was also expanded to a (TiO2)9 cluster to confirm the reliability of the simple Ti(OH)3·H2O model. It is revealed that the trend of absorption spectra for dye-(TiO2)9 complexes is similar to that obtained with the simple dye-Ti(OH)3·H2O model (Fig. S2 and Table S2†).
 |
| Fig. 5 The absorption spectra (solid line) and light-harvesting efficiency curves (dash line) for adsorbed IQ4 (black) and NIQ4 (red). | |
3.1.2 Electron injection efficiency. Electron injection efficiency (Φinj) associated with JSC can be indirectly characterized by the electron injection driving force (ΔGinj) from excited dyes to the CB band of semiconductor. ΔGinj can be determined by the following equation, assuming that the electron injection occurs from the unrelaxed excited state:80 |
 | (7) |
In which, ESCCB is the conduction band edge of TiO2 (ESCCB = −4.0 eV vs. vacuum), and Edye*OX is the excited state oxidation potential and estimated by: |
 | (8) |
In which, EdyeOX is the ground state oxidation potential and Eλmax is the vertical transition energy from the ground state to excited state. We list these values in Table 3. It can be seen that the EdyeOX of IQ4 (−5.07 eV) is well reproduced by our calculation (−4.91 eV), whereas, the overestimation of the vertical transition energy for IQ4 directly leads to a larger deviation of the calculated Edye*OX (−2.26 eV) from experimental measurements (−2.98 eV), and thus a larger ΔGinj (1.74 eV). Moreover, it reveals in Table 3 that the replacement of the indoline unit in IQ4 by triarylamine (TIQ4), ullazine (UIQ4), and NP (NIQ4) moieties, respectively, displays a more significant effect on Edye*OX than that on EdyeOX. Though the ΔGinj value of IQ4 is the smallest in all investigated dyes, it is sufficient to inject electrons from excited dyes to CB edge of TiO2.77
3.1.3 Dye regeneration efficiency. For an effective DSSC, the dye regeneration completed by the reducing agent is vital in achieving a higher JSC. A simplified, but informative, way to investigate dye regeneration relies on the analysis of the spin density distributions for the oxidized systems.35 Obtained by the (U)DFT/B3LYP calculations, the contour plots of the spin densities for such systems together with the contributions from the donor moiety have been summarized in Table 4. As depicted in Table 4, the spin density for all oxidized dyes mainly localizes at the donor moiety of the dye, and the contribution from the electron donor shows a higher value for the newly designed dyes (UIQ4˙+ (99.7%), TIQ4˙+ (97.2%), and NIQ4˙+ (95.0%)), relative to IQ4˙+ (89.2%). This effect could enhance the interaction between the designed dyes and the I−/I3− redox couple, and therefore the rate of dye regeneration, compared to IQ4˙+, where holes reside on the π-bridge in a larger proportion.
Table 4 Contour plots of the spin density for the oxidized species of the investigated dyes resulting from (U)DFT/B3LYP/C–PCM(CH2Cl2) calculations (values given in parentheses represent the spin density contribution of the donor moiety)
The quantitative way to measure the efficiency of dye regeneration is computing the internal reorganization energy (λreg)34 and the regeneration driving force (ΔGreg) for dyestuff. The λreg can be calculated by:
|
λreg = [E+0 − E++] + [E0+ − E0]
| (9) |
In which,
E++ and
E0 are optimized energies of the cationic and neutral forms of a dye,
E+0 is the energy of the cation at the neutral geometry, and
E0+ is the energy of the neutral molecule at the cationic geometry. The Δ
Greg is determined by the energy difference between the redox potential of I
−/I
3− (
Eelectrolyteredox = − 4.8 eV) and
EdyeOX:
|
ΔGreg = Eelectrolyteredox − Edyeox
| (10) |
One can find in Table 3 that the NIQ4 dye presents the largest driving force of 0.14 eV for dye regeneration as well as the comparable reorganization energy (0.14 eV) to IQ4 (0.11 eV), which together with the favorable spin density distribution of the cationic radical, sufficient ΔGinj, and better light-harvesting ability may lead to a higher Jsc for NIQ4-sensitized solar cell. As for UIQ4 and TIQ4, the relatively larger λreg (UIQ4/0.22 eV, TIQ4/0.25 eV) and smaller ΔGreg (UIQ4/0.06 eV, TIQ4/0.05 eV), despite the favorable spin density contribution from donor moieties may decrease the efficiency of dye regeneration, and thus result in a lower Jsc for the related cell. In this context, it is safe to say that the N-annulated perylene unit instead of indoline in quinoxaline dyes is beneficial in enhancing the short-circuit current density.
3.2 Evaluating Voc
Besides the Jsc, a potential cell should also possess a high open-circuit photovoltage (Voc) value to reach high power conversion efficiency (η). The Voc is directly relevant with the CB energy (ECB). It is known that the adsorption of a dye molecule on TiO2 surface can induce a shift of the conduction band profile of TiO2 (ΔCB), and this effect can be divided into two aspects, namely, charge transfer (CT) and electrostatic (EL) potential.81 In the present work, the investigation of electrostatic effect was simplified by calculating the dipole moment of the free dyes at the geometries of dyes adsorbed onto TiO2 surface (μnormal),82 while the exploration of CT at dye/TiO2 interface can be qualitatively or semi-quantitatively realized through the analysis of the difference of charge distribution in the semiconductor between the ground and the first excited state of dye-TiO2 complex.68,79 A larger μnormal pointing outward the semi-conductor surface and a more charge distribution in TiO2 will lead to a larger ΔCB. Considering the bidentate chelating mode between IQ4/NIQ4 and TiO2, the C2 axis of the carboxylate in the dye is set parallel to the x-axis, and the yz plane is parallel to the semiconductor surface. Thus, the μx of dye molecules is referred to as μnormal. The calculated results are shown in Table 5. The μnormal of IQ4 (10.3 D) is almost the double value of NIQ4 (5.78 D). As for CT effect, the natural population analysis (NPA) demonstrates that the number of photo-injected electrons in TiO2 for NIQ4 (0.041 e) is slightly more than that for IQ4 (0.037 e), which says in favor of NIQ4 as a good electron-injection precursor.
Table 5 The natural bond orbital analysis (atomic charge) of the ground (S0) and the first excited state (S1) of dye-Ti(OH)3·H2O adducts (in e)
Dye |
S0 |
S1 |
Δqb |
donor |
π-spacer |
Anchor |
Ti(OH)3·H2O |
μnormala |
donor |
π-spacer |
anchor |
Ti(OH)3·H2O |
μnormal is the vertical dipole moment (Debye) of the free dyes at the geometries of dyes adsorbed onto TiO2 surface. Δq is the charge density differences on Ti(OH)3·H2O between S0 and S1. |
IQ4 |
0.102 |
0.125 |
−0.657 |
0.430 |
10.3 |
0.0495 |
0.231 |
−0.673 |
0.393 |
0.037 |
NIQ4 |
0.0434 |
0.163 |
−0.642 |
0.436 |
5.78 |
0.0454 |
0.229 |
−0.669 |
0.395 |
0.041 |
On the other hand, charge recombination between TiO2-injected electrons and the oxidized dye or the electrolyte (I−/I3−) can substantially decrease the charge density in the semiconductor, and thus leads to a reduced Voc value. Differences in the recombination kinetics between TiO2-injected electrons and the dye cationic radicals can be reflected from the twisted dihedral angle between the dye donor and π-conjugated moiety. Looking at the geometry parameters of oxidized dyes reported in Table S1,† the selected dihedral angle is 38.9° for IQ4˙+ and 49.3° for NIQ4˙+. The more twisted angle between NP unit and quinoxaline in NIQ4 is believed to hamper the back-recombination from the injected-electrons to dye donor, exerting a desirable effect on charge density in TiO2. In addition, the improved blocking effect of NIQ4 (the bulky N-annulated perylene moiety and bis(2,4-dihexyloxy)benzene substituted quinoxaline36) compared to IQ4 not only prevents unfavorable aggregation between dyes, but also can form a compact dye layer to effectively inhibit the approach of the redox couple to the TiO2 surface,4 consequently an enhanced device performance. The retarded electron–hole recombination in NIQ4 is supposed to compensate well the undesired smaller μnormal, and therefore makes NIQ4 endow a comparable contribution on the photovoltage compared to the reported IQ4 dye.
To this point, we can conclude that NIQ4 is a potential candidate for DSSC applications. A solar cell sensitized by NIQ4 is expected to exhibit a higher JSC and a comparable VOC to IQ4, and thus an improved DSSC photovoltaic performance.
4. Conclusions
In this study, with IQ4 as the prototype dye, three novel quinoxaline sensitizers (UIQ4, TIQ4, and NIQ4) were designed with electron donor modification, aiming to identify a promising dye for DSSC applications. Factors correlated with the short-circuit current density (JSC) and open-circuit photovoltage (VOC), including light-harvesting efficiency (LHE), electron injection efficiency (ΔGinj), dye regeneration efficiency (ΔGreg), the vertical dipole moment (μnormal), and the number of photo-injected electrons in TiO2, respectively, have been studied in detail by means of DFT and TDDFT approaches. It is found that compared to the reported IQ4 dye bearing indoline donor, the NIQ4 dye functionalized by N-annulated perylene (NP) donor displays the more red-shifted (λmax = 477 nm) absorption and broader spectral domain (the LHE reaches close to 100% in the range of 300–570 nm). The good light-harvesting and electron injection ability together with the comparable regeneration efficiency of NIQ4 compared to those of IQ4 may eventually enhance the JSC potential of NIQ4-based cell. Furthermore, despite possessing a smaller μnormal, the improved blocking effect of NIQ4 not only prevents unfavorable aggregation between dyes, but also effectively impedes the parasitic back-recombination. Therefore, we can conclude that a solar cell fabricated with NIQ4 is expected to exhibit a higher JSC and a comparable VOC to that with IQ4, and thus an improved DSSC photovoltaic performance.
The results of this paper emphasize the significance of the electron donor in controlling the kinetic parameters, which are responsible for the device performance. We hope our work will provide a rational guideline for the future design of push–pull organic dyes towards more efficient dye-sensitized solar cells.
Acknowledgements
The authors gratefully acknowledge financial support from the Major State Basic Research Development Programs of China (2011CBA00701), the National Natural Science Foundation of China (21473010, 21303007), the Excellent Young Scholars Research Fund of Beijing Institute of Technology (2013YR1917), and Beijing Key Laboratory for Chemical Power Source and Green Catalysis (2013CX02031). This work is also supported by the opening project of State Key Laboratory of Explosion Science and Technology (Beijing Institute of Technology) (ZDKT12-03).
References
- B. E. Hardin, H. J. Snaith and M. D. McGehee, Nat. Photonics, 2012, 6, 162–169 CrossRef CAS.
- M. F. Xu, M. Zhang, M. Pastore, R. Z. Li, F. De Angelis and P. Wang, Chem. Sci., 2012, 3, 976–983 RSC.
- A. K. Chandiran, M. Abdi-Jalebi, M. K. Nazeeruddin and M. Grätzel, ACS Nano, 2014, 8, 2261–2268 CrossRef CAS PubMed.
- Z. Ning, Y. Fu and H. Tian, Energy Environ. Sci., 2010, 3, 1170–1181 CAS.
- H. N. Tsao, C. Yi, T. Moehl, J.-H. Yum, S. M. Zakeeruddin, M. K. Nazeeruddin and M. Grätzel, ChemSusChem, 2011, 4, 591–594 CrossRef CAS PubMed.
- B. O'Regan and M. Grätzel, Nature, 1991, 353, 737–740 CrossRef.
- L.-N. Yang, Z.-Z. Sun, S.-L. Chen and Z.-S. Li, ChemPhysChem, 2014, 15, 458–466 CrossRef CAS PubMed.
- S. Feng, Q.-S. Li, L.-N. Yang, Z.-Z. Sun, T. A. Niehaus and Z.-S. Li, J. Power Sources, 2015, 273, 282–289 CrossRef CAS PubMed.
- P.-P. Sun, Q.-S. Li, L.-N. Yang, Z.-Z. Sun and Z.-S. Li, Phys. Chem. Chem. Phys., 2014, 16, 21827–21837 RSC.
- L.-N. Yang, H.-Y. Zhou, P.-P. Sun, S.-L. Chen and Z.-S. Li, ChemPhysChem, 2015, 16, 601–606 CrossRef CAS PubMed.
- Q. Chai, W. Li, Y. Wu, K. Pei, J. Liu, Z. Geng, H. Tian and W. Zhu, ACS Appl. Mater. Interfaces, 2014, 6, 14621–14630 CAS.
- X. Xu, D. Huang, K. Cao, M. Wang, S. M. Zakeeruddin and M. Grätzel, Sci. Rep., 2013, 3, 1489 Search PubMed.
- J.-Y. Chang, S. C. Chang, S.-H. Tzing and C.-H. Li, ACS Appl. Mater. Interfaces, 2014, 6, 22272–22281 CAS.
- J. Mao, J. Yang, J. Teuscher, T. Moehl, C. Yi, R. Humphry-Baker, P. Comte, C. Grätzel, J. Hua, S. M. Zakeeruddin, H. Tian and M. Grätzel, J. Phys. Chem. C, 2014, 118, 17090–17099 CAS.
- A. Yella, H. W. Lee, H. N. Tsao, C. Yi, A. K. Chandiran, M. K. Nazeeruddin, E. W. Diau, C. Y. Yeh, S. M. Zakeeruddin and M. Grätzel, Science, 2011, 334, 629–634 CrossRef CAS PubMed.
- M. Xu, D. Zhou, N. Cai, J. Liu, R. Li and P. Wang, Energy Environ. Sci., 2011, 4, 4735 CAS.
- M. Zhang, Y. Wang, M. Xu, W. Ma, R. Li and P. Wang, Energy Environ. Sci., 2013, 6, 2944–2949 CAS.
- W. Zhu, Y. Wu, S. Wang, W. Li, X. Li, J. Chen, Z.-s. Wang and H. Tian, Adv. Funct. Mater., 2011, 21, 756–763 CrossRef CAS.
- S. Qu, C. Qin, A. Islam, Y. Wu, W. Zhu, J. Hua, H. Tian and L. Han, Chem. Commun., 2012, 48, 6972–6974 RSC.
- J.-J. Kim, H. Choi, J.-W. Lee, M.-S. Kang, K. Song, S. O. Kang and J. Ko, J. Mater. Chem., 2008, 18, 5223–5229 RSC.
- Q. Feng, X. Lu, G. Zhou and Z.-S. Wang, Phys. Chem. Chem. Phys., 2012, 14, 7993–7999 RSC.
- Y. Wu, X. Zhang, W. Li, Z.-S. Wang, H. Tian and W. Zhu, Adv. Energy Mater., 2012, 2, 149–156 CrossRef CAS.
- S. Haid, M. Marszalek, A. Mishra, M. Wielopolski, J. Teuscher, J.-E. Moser, R. Humphry-Baker, S. M. Zakeeruddin, M. Grätzel and P. Bäuerle, Adv. Funct. Mater., 2012, 22, 1291–1302 CrossRef CAS.
- W. Li, Y. Wu, Q. Zhang, H. Tian and W. Zhu, ACS Appl. Mater. Interfaces, 2012, 4, 1822–1830 CAS.
- M. Velusamy, K. R. Justin Thomas, J. T. Lin, Y.-C. Hsu and K.-C. Ho, Org. Lett., 2005, 7, 1899–1902 CrossRef CAS PubMed.
- Y. Cui, Y. Wu, X. Lu, X. Zhang, G. Zhou, F. B. Miapeh, W. Zhu and Z.-S. Wang, Chem. Mater., 2011, 23, 4394–4401 CrossRef CAS.
- Z.-M. Tang, T. Lei, K.-J. Jiang, Y.-L. Song and J. Pei, Chem.–Asian J., 2010, 5, 1911–1917 CrossRef CAS PubMed.
- J. Shi, Z. Chai, J. Su, J. Chen, R. Tang, K. Fan, L. Zhang, H. Han, J. Qin, T. Peng, Q. Li and Z. Li, Dyes Pigm., 2013, 98, 405–413 CrossRef CAS PubMed.
- K. Pei, Y. Wu, W. Wu, Q. Zhang, B. Chen, H. Tian and W. Zhu, Chem.–Eur. J., 2012, 18, 8190–8200 CrossRef CAS PubMed.
- D. W. Chang, H. J. Lee, J. H. Kim, S. Y. Park, S.-M. Park, L. Dai and J.-B. Baek, Org. Lett., 2011, 13, 3880–3883 CrossRef CAS PubMed.
- X. Lu, Q. Feng, T. Lan, G. Zhou and Z.-S. Wang, Chem. Mater., 2012, 24, 3179–3187 CrossRef CAS.
- K. Pei, Y. Wu, A. Islam, Q. Zhang, L. Han, H. Tian and W. Zhu, ACS Appl. Mater. Interfaces, 2013, 5, 4986–4995 CAS.
- K. Pei, Y. Wu, A. Islam, S. Zhu, L. Han, Z. Geng and W. Zhu, J. Phys. Chem. C, 2014, 118, 16552–16561 CAS.
- J. Zhang, L. Yang, M. Zhang and P. Wang, RSC Adv., 2013, 3, 6030 RSC.
- A. Yella, R. Humphry-Baker, B. F. E. Curchod, N. Ashari Astani, J. Teuscher, L. E. Polander, S. Mathew, J.-E. Moser, I. Tavernelli, U. Rothlisberger, M. Grätzel, M. K. Nazeeruddin and J. Frey, Chem. Mater., 2013, 25, 2733–2739 CrossRef CAS.
- J. Yang, P. Ganesan, J. Teuscher, T. Moehl, Y. J. Kim, C. Yi, P. Comte, K. Pei, T. W. Holcombe, M. K. Nazeeruddin, J. Hua, S. M. Zakeeruddin, H. Tian and M. Grätzel, J. Am. Chem. Soc., 2014, 136, 5722–5730 CrossRef CAS PubMed.
- J. Zhang, H.-B. Li, Y. Geng, S.-Z. Wen, R.-L. Zhong, Y. Wu, Q. Fu and Z.-M. Su, Dyes Pigm., 2013, 99, 127–135 CrossRef CAS PubMed.
- A. K. Biswas, S. Barik, A. Sen, A. Das and B. Ganguly, J. Phys. Chem. C, 2014, 118, 20763–20771 CAS.
- A. Dualeh, F. De Angelis, S. Fantacci, T. Moehl, C. Yi, F. Kessler, E. Baranoff, M. K. Nazeeruddin and M. Grätzel, J. Phys. Chem. C, 2011, 116, 1572–1578 Search PubMed.
- A. Baheti, K. R. Justin Thomas, C. Li, C. Lee and K. Ho, ACS Appl. Mater. Interfaces, 2015, 7, 2249–2262 CAS.
- Q. Chai, W. Li, S. Zhu, Q. Zhang and W. Zhu, ACS Sustainable Chem. Eng., 2013, 2, 239–247 CrossRef.
- S. M. Feldt, E. A. Gibson, E. Gabrielsson, L. Sun, G. Boschloo and A. Hagfeldt, J. Am. Chem. Soc., 2010, 132, 16714–16724 CrossRef CAS PubMed.
- L.-N. Yang, Z.-Z. Sun, S.-L. Chen and Z.-S. Li, Dyes Pigm., 2013, 99, 29–35 CrossRef CAS PubMed.
- D. Kuang, S. Uchida, R. Humphry-Baker, S. M. Zakeeruddin and M. Grätzel, Angew. Chem., Int. Ed., 2008, 47, 1923–1927 CrossRef CAS PubMed.
- J. H. Delcamp, A. Yella, T. W. Holcombe, M. K. Nazeeruddin and M. Grätzel, Angew. Chem., Int. Ed., 2013, 52, 376–380 CrossRef CAS PubMed.
- J. Luo, M. Xu, R. Li, K. W. Huang, C. Jiang, Q. Qi, W. Zeng, J. Zhang, C. Chi, P. Wang and J. Wu, J. Am. Chem. Soc., 2014, 136, 265–272 CrossRef CAS PubMed.
- M. J. Frisch. G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery Jr, J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian 09, Revision D. 01, Gaussian, Inc., Wallingford CT, 2009 Search PubMed.
- A. D. Becke, J. Chem. Phys., 1993, 98, 1372–1377 CrossRef CAS PubMed.
- A. D. Becke, J. Chem. Phys., 1993, 98, 5648–5652 CrossRef CAS PubMed.
- C. T. Lee, W. T. Yang and R. G. Parr, Phys. Rev. B: Condens. Matter Mater. Phys., 1988, 37, 785–789 CrossRef CAS.
- H. Li, Y. Li and M. Chen, RSC Adv., 2014, 4, 57916–57922 RSC.
- N. Nath Ghosh, A. Chakraborty, S. Pal, A. Pramanik and P. Sarkar, Phys. Chem. Chem. Phys., 2014, 16, 25280–25287 RSC.
- K. Chaitanya, X.-H. Ju and B. M. Heron, RSC Adv., 2015, 5, 3978–3998 RSC.
- W. Li, L. G. C. Rego, F.-Q. Bai, J. Wang, R. Jia, L.-M. Xie and H.-X. Zhang, J. Phys. Chem. Lett., 2014, 5, 3992–3999 CrossRef CAS.
- H.-Q. Xia, C.-P. Kong, J. Wang, F.-Q. Bai and H.-X. Zhang, RSC Adv., 2014, 4, 50338–50350 RSC.
- H.-H. G. Tsai, C.-J. Tan and W.-H. Tseng, J. Phys. Chem. C, 2015, 119, 4431–4443 CAS.
- N. Martsinovich and A. Troisi, Energy Environ. Sci., 2011, 4, 4473–4495 CAS.
- T. Yanai, D. P. Tew and N. C. Handy, Chem. Phys. Lett., 2004, 393, 51–57 CrossRef CAS PubMed.
- S. I. Gorelsky, SWizard program, University of Ottawa, Ottawa, Canada, 2012, http://www.sg-chem.net/ Search PubMed.
- J. Liu, X. Yang and L. Sun, Chem. Commun., 2013, 49, 11785–11787 RSC.
- V. Barone and M. Cossi, J. Phys. Chem. A, 1998, 102, 1995–2001 CrossRef CAS.
- R. Cammi, B. Mennucci and J. Tomasi, J. Phys. Chem. A, 1999, 103, 9100–9108 CrossRef CAS.
- J. Tomasi, B. Mennucci and R. Cammi, Chem. Rev., 2005, 105, 2999–3093 CrossRef CAS PubMed.
- W. Ma, Y. Jiao and S. Meng, J. Phys. Chem. C, 2014, 118, 16447–16457 CAS.
- A. Yella, H.-W. Lee, H. N. Tsao, C. Yi, A. K. Chandiran, M. K. Nazeeruddin, E. W.-G. Diau, C.-Y. Yeh, S. M. Zakeeruddin and M. Grätzel, Science, 2011, 334, 629–634 CrossRef CAS PubMed.
- H.-B. Li, J. Zhang, Y. Wu, J.-L. Jin, Y.-A. Duan, Z.-M. Su and Y. Geng, Dyes Pigm., 2014, 108, 106–114 CrossRef CAS PubMed.
- J. Zhang, H.-B. Li, J.-Z. Zhang, Y. Wu, Y. Geng, Q. Fu and Z.-M. Su, J. Mater. Chem. A, 2013, 1, 14000–14007 CAS.
- J. Zhang, H.-B. Li, S.-L. Sun, Y. Geng, Y. Wu and Z.-M. Su, J. Mater. Chem., 2012, 22, 568–576 RSC.
- T. Marinado, K. Nonomura, J. Nissfolk, M. K. Karlsson, D. P. Hagberg, L. Sun, S. Mori and A. Hagfeldt, Langmuir, 2009, 26, 2592–2598 CrossRef PubMed.
- S. Rühle, M. Greenshtein, S. G. Chen, A. Merson, H. Pizem, C. S. Sukenik, D. Cahen and A. Zaban, J. Phys. Chem. B, 2005, 109, 18907–18913 CrossRef PubMed.
- S. Ardo and G. J. Meyer, Chem. Soc. Rev., 2009, 38, 115–164 RSC.
- G. M. Hasselman, D. F. Watson, J. R. Stromberg, D. F. Bocian, D. Holten, J. S. Lindsey and G. J. Meyer, J. Phys. Chem. B, 2006, 110, 25430–25440 CrossRef CAS PubMed.
- Y. Jiao, W. Ma and S. Meng, Chem. Phys. Lett., 2013, 586, 97–99 CrossRef CAS PubMed.
- J.-Z. Zhang, J. Zhang, H.-B. Li, Y. Wu, H.-L. Xu, M. Zhang, Y. Geng and Z.-M. Su, J. Power Sources, 2014, 267, 300–308 CrossRef CAS PubMed.
- T. Lu and F. W. Chen, Acta Chim. Sin., 2011, 69, 2393–2406 CAS.
- T. Le Bahers, T. Pauporté, P. P. Lainé, F. Labat, C. Adamo and I. Ciofini, J. Phys. Chem. Lett., 2013, 4, 1044–1050 CrossRef CAS.
- B. Lim, G. Y. Margulis, J.-H. Yum, E. L. Unger, B. E. Hardin, M. Grätzel, M. D. McGehee and A. Sellinger, Org. Lett., 2013, 15, 784–787 CrossRef CAS PubMed.
- B. Peng, S. Yang, L. Li, F. Cheng and J. Chen, J. Chem. Phys., 2010, 132, 034305 CrossRef PubMed.
- W.-L. Ding, D.-M. Wang, Z.-Y. Geng, X.-L. Zhao and W.-B. Xu, Dyes Pigm., 2013, 98, 125–135 CrossRef CAS PubMed.
- J. Preat, C. Michaux, D. Jacquemin and E. A. Perpete, J. Phys. Chem. C, 2009, 113, 16821–16833 CAS.
- E. Ronca, M. Pastore, L. Belpassi, F. Tarantelli and F. De Angelis, Energy Environ. Sci., 2013, 6, 183–193 CAS.
- L.-N. Yang, Z.-Z. Sun, Q.-S. Li, S.-L. Chen, Z.-S. Li and T. A. Niehaus, J. Power Sources, 2014, 268, 137–145 CrossRef CAS PubMed.
Footnote |
† Electronic supplementary information (ESI) available: Energy-level graphic of the donor moiety, optical property of dye-Ti(OH)3·H2O and dye-(TiO2)9 complexes, selected geometry parameters for the neutral and oxidized forms of dyes, cartesian coordinates of the optimized structures. See DOI: 10.1039/c5ra00587f |
|
This journal is © The Royal Society of Chemistry 2015 |
Click here to see how this site uses Cookies. View our privacy policy here.