You are on page 1of 7

8541

2009, 113, 85418547 Published on Web 04/27/2009

Simple Triarylamine-Based Dyes Containing Fluorene and Biphenyl Linkers for Efcient Dye-Sensitized Solar Cells
Abhishek Baheti, Payal Tyagi, K. R. Justin Thomas,*, Ying-Chan Hsu, and Jiann Tsuen Lin
Department of Chemistry, Indian Institute of Technology Roorkee, Roorkee-247 667, India, and Institute of Chemistry, Academia Sinica, 115 Nankang, Taipei, Taiwan ReceiVed: March 12, 2009; ReVised Manuscript ReceiVed: April 20, 2009

New dipolar dyes featuring arylamine donors, uorene or biphenyl conjugation, and cyanoacetic acid acceptor units have been synthesized and characterized by spectral, optical, and electrochemical measurements. Absorption measurements indicate that the uorene conjugation is benecial for red shifting the charge transfer transition and increasing the optical density. On the contrary, the biphenyl linker leads to a blue-shifted hypochromic charge-transfer transition due to the inefciency in the -conjugation. For a particular linker, the diphenylamine unit acts as an efcient donor when compared to the 1-naphthylphenylamine moiety. However, the red shift observed in the absorption prole for the diphenylamine-containing dyes is compensated for by a reduction in the molar extinction coefcient, that is, the naphthylamine derivatives are benecial for increasing the molar extinction coefcient. Electrochemical investigations show that the uorene-based dyes are more redox stable than the corresponding biphenyl derivatives. Effects of deprotonation on the absorption maxima and redox potential have been investigated by absorption and electrochemical measurements. These studies reveal a blue shift in the absorption prole and a cathodic shift for the oxidation potential originating due to deprotonation of the dyes. The dye-sensitized solar cells fabricated using these dyes exhibit moderate to good efciencies, and the uorene-based dyes are better than the biphenyl analogues. It is attributed to the diminished charge separation in the biphenyl derivatives due to the nonplanar structural feature which hinders donor-acceptor interactions and electron delocalization across the molecule.
Introduction Research to develop new dyes suitable for application in photovoltaics1 and dye-sensitized solar cells (DSSCs)2,3 has gained much impetus in the recent years, owing to the demand for renewable energy resources. Contrary to the fossil fuels, solar energy is available aplenty in most of the countries and may be generously used for electricity production.4 Solar cells which convert the solar energy into electrical energy are conventionally fabricated using silicon-based materials. Recently, it has been demonstrated that solar cells can also be constructed from organometallic and organic dyes. Use of organometallic or organic dyes in solar cells offers several advantages over the silicon-based inorganic materials, namely, ease for fabrication, reduction in production cost, and functional tuning by chemical modication. Gratzel and co-workers5,6 have explored a large variety of ruthenium complexes for application in dye-sensitized solar cells and observed that thiocyanate-containing ruthenium(II) complexes perform exceptionally well in nanocrystalline anatase TiO2-based dye-sensitized solar cells. Among the organometallic complexes, cis-bis(isothiocyanato)bis(2,2-bipyridyl-4,4-dicarboxylato)-ruthenium(II) (N3)6f and the related derivatives have been applied in DSSC more successfully with
* To whom correspondence should be addressed. Phone: +91-1332285376. E-mail: krjt8fcy@iitr.ernet.in. Indian Institute of Technology Roorkee. Academia Sinica.

highest power efciency () values (9-12%).7 Recent research has established the properties required for the dyes to effectively function in dye-sensitized solar cells. It is now accepted that a dye with wide solar spectral absorption, a high molar extinction coefcient, efcient charge separation, redox stability, and suitable functional groups to interact with the electron sink such as TiO2 is necessary for efcient energy harvesting. Organic dyes featuring donor-acceptor architecture have been synthesized utilizing various conjugating motifs for application in dye-sensitized solar cells.8-13 Amine donors and cyanoacrylic acid acceptor based dyes have been developed and intensively studied as sensitizers for dye-sensitized solar cells. It has been demonstrated that the conjugating segment must be appropriate to separate the HOMO located on the amine and the LUMO located on the acceptor unit. This must ensure an efcient intramolecular charge separation with high photocurrent generation and slow recombination. Several aromatic segments have been incorporated as the linker in dyes and found benecial for the photocurrent generation and intramolecular charge separation. Thiophene,9 coumarin,10 benzothiadiazole,11 and phenylene12 derivatives have been synthesized and found to exhibit interesting photocurrent generation. Thienyluorene conjugated dyes have also been reported to show excellent dyesensitized solar cell efciency.13 However, increasing the conjugation has been found to be detrimental for charge separation and energy harnessing.13 This is probably due to the reduced interaction between the donor and acceptor upon 2009 American Chemical Society

10.1021/jp902206g CCC: $40.75

8542

J. Phys. Chem. C, Vol. 113, No. 20, 2009

Letters 123.40, 122.61, 121.88, 119.54, 117.75, 100.19. HR MS (ESI) m/z calcd. for C33H28O2N2: 484.2151; Found: 484.2157. (E)-2-Cyano-3-(9,9-diethyl-7-(naphthalen-1-yl(phenyl)amino)9H-uoren-2-yl)acrylic Acid (4d). Red solid. Yield: 75%. IR (KBr, cm-1): 2225 (CtN). 1H NMR (CDCl3, 500 MHz): 0.32 (t, J ) 7.3 Hz, 6 H), 1.77-1.84 (m, 2H), 1.93-2.00 (m, 2H), 6.93 (dd, J ) 10.0 Hz, 2.3 Hz, 1H), 6.98 (d, J ) 2.5 Hz, 2 H), 7.02 (t, J ) 7.5 Hz, 1H), 7.13 (d, J ) 8.0 Hz, 2H), 7.24-7.36 (m, 4H), 7.44-7.51 (m, 2H), 7.55 (d, J ) 8.5 Hz, 1H), 7.66 (d, J ) 8.0 Hz, 1H), 7.80 (d, J ) 8.5 Hz, 1H), 7.89 (t, J ) 8.3 Hz, 2H), 7.95 (dd, J ) 8.0 Hz, 1.0 Hz, 1H), 7.99 (s, 1H), 8.33 (s, 1H). 13C NMR (CDCl3, 125.770 MHz): 155.48, 153.00, 150.54, 149.81, 148.30, 147.26, 143.35, 135.28, 133.35, 131.74, 130.34, 129.26, 129.31, 128.47, 126.98, 126.65, 126.34, 126.32, 126.17, 124.93, 124.17, 122.81, 122.53, 121.68, 120.53, 119.23, 116.65, 115.58, 100.00. HR MS (ESI) m/z calcd. for C37H30O2N2: 534.2307; Found: 534.2315. 3. Results and Discussion 3.1. Synthesis and Characterization. The synthetic routes leading to the formation of the four new dyes based on uorene and biphenyl cores are illustrated in Scheme 1. The synthesis began with 4,4-dibromobiphenyl (1a) or 2,7-dibromo-9,9diethyluorene (1b), which was reacted with the appropriate diarylamine employing Hartwigs palladium-catalyzed C-N coupling protocol.14 This generated the monoaminated biphenyl or uorene derivatives (2a-2d) in moderate yields (55-65%). Lithiation of the compounds 2a-2d with a stoichiometric amount of n-butyl lithium at -78 C followed by quenching with N-formyl piperidine and subsequent acidic hydrolysis produced the required aldehydes (3a-3d) in reasonable quantities (48-72%). In the last step, the aldehyde derivatives 3a-3d were converted to the desired dyes 4a-4d by reaction with cyanoacetic acid in reuxing acetic acid in the presence of ammonium acetate. The target compounds were characterized by NMR spectroscopy, elemental analysis, and high-resolution mass spectrometry and were found to be consistent with the proposed structures. The dyes were orange or red in color and freely soluble in common solvents including dichloromethane, tetrahydrofuran, methanol, dimethylformamide, and toluene. The dilute solutions of the dyes appeared dark yellow. 3.2. Photophysical Properties. The absorption spectra recorded for the dyes in dichloromethane solutions are displayed in Figure 1. All of the dyes exhibit a prominent peak around 395-424 nm (Table 1 and Figure 1) with molar extinction coefcient between 20.9 103 to 32.9 103 M-1 cm-1. This is attributed to the charge-transfer transition from the amine donor to the cyanoacrylic acid acceptor moiety. This assignment is supported by the fact that the dyes show negative solvatochromism for this transition and red-shifted absorption (Figure S1, Supporting Information) when compared to the parent amine (2) and precursor aldehydes (3). The blue shift observed for the dyes in the polar solvents (for instance, see Figure S2, Supporting Information) probably indicates that the dyes are well solvated in the ground state. The absorption maxima of the biphenyl-based dyes (4a and 4b) display a 14-19 nm hypsochromic shift when compared to that of the uorene-based analogues (4c and 4d). This blue shift probably stems from the stabilization of HOMO orbitals, which increases the band gap and leads to the reduction in the absorption wavelength. It is also possible that the twisting in the excited state hampers the donor-acceptor interaction in the biphenyl-based dyes and subsequently may lead to an inefcient electron injection.8a,15 Alternatively, as the uorene segment is

increasing the linker length. Additionally an elongated spacer may also lead to nonplanar structures which will forbid effective charge migration. In this paper, we report simple and efcient dyes based on uorene. Thiophene that separated the uorene and acceptor segment in an earlier design has been stripped off to decrease the linker length. Thiophene being an electron-rich segment might not help to place the HOMO and LUMO well separated. We have also prepared the biphenyl analogues to study the effect of planar and rigid structural components on the charge separation efciency. Fluorene appears to be a better -conjugator when compared to the biphenyl group due to the presence of a methylene bridge which renders rigidity and ensures coplanarity. On this basis, it was expected that the uorenebased dyes may lead to efcient charge separation and power efciency. Due to the presence of aromatic trigonal amine units, both the planar uorene and nonplanar biphenyl segments cannot engage in - stacking; thus, dye aggregation may also be eliminated. Aggregation may affect the absorption of light and also promote intermolecular quenching processes detrimental for the generation of electrons. 2. Experimental Section 2.1. Synthesis of the Dyes (4a-4d). The synthesis of all of the dyes were accomplished by following the procedure illustrated below for 4a. (E)-2-Cyano-3-(4-(diphenylamino)biphenyl-4-yl)acrylic Acid (4a). The aldehyde, 4-(diphenylamino)biphenyl-4-carbaldehyde (3a, 0.70 g, 2 mmol), 2-cyanoacetic acid (0.19 g, 2.2 mmol), ammonium acetate (0.05 g), and glacial acetic acid (5 mL) were mixed together and heated at 120 C with efcient stirring for 8 h. The red solution was cooled to induce precipitation. The precipitate formed was ltered and thoroughly washed with hexane and a cold hexane/diethyl ether (5:1) mixture. A red solid obtained was puried further by crystallization from an ether/hexane mixture. Yield: 76%. IR (KBr, cm-1): 2218 (CtN). 1 H NMR (CDCl3, 400 MHz): 7.05-7.08 (m, 2H), 7.11-7.15 (m, 6H), 7.24-7.30 (m, 4H), 7.49-7.53 (m, 2H), 7.70-7.72 (d, J ) 8.48 Hz, 2H), 8.06-8.08 (d, J ) 8.47 Hz, 2H), 8.30 (s, 1H). 13C NMR (CDCl3, 125.770 MHz): 156.00, 148.70, 147.22, 132.25, 132.11, 129.46, 127.94, 127.01, 125.04, 123.65, 122.87, 100.00. HR MS (FAB) m/z calcd. for C28H20O2N2: 416.1525; Found: 416.1530. (E)-2-Cyano-3-(4-(naphthalen-1-yl(phenyl)amino)biphenyl4-yl)acrylic Acid (4b). Red solid. Yield: 82%. IR (KBr, cm-1): 2227 (CtN). 1H NMR (CDCl3, 400 MHz): 6.98-7.04 (m, 3H), 7.13-7.15 (d, J ) 7.69 Hz, 2H), 7.22-7.25 (m, 2H), 7.35-7.39 (m, 2H), 7.45-7.50 (m, 4H), 7.65-7.69 (m, 2H), 7.79-7.81 (d, J ) 8.18 Hz, 1H), 7.88-7.94 (m, 2H), 8.03-8.05 (d, J ) 8.39 Hz, 2H), 8.28 (s, 1H). 13C NMR (CDCl3, 125.770 MHz): 167.29, 156.04, 149.27, 147.84, 146.34, 142.86, 135.33, 132.27, 131.18, 131.10, 130.42, 129.27, 128.53, 128.00, 127.91, 127.47, 127.02, 126.88, 126.79, 126.43, 126.33, 124.04, 123.07, 122.91, 122.72, 120.89, 120.69, 115.62, 100.55. HR MS (FAB) m/z calcd. for C32H22O2N2: 466.1681; Found: 466.1687. (E)-2-Cyano-3-(7-(diphenylamino)-9,9-diethyl-9H-uoren2-yl)acrylic Acid (4c). Red solid. Yield: 71%. IR (KBr, cm-1): 2221 (CtN). 1H NMR (CDCl3, 400 MHz): 0.11 (t, J ) 7.3 Hz, 6H), 1.60-1.79 (m, 4H), 6.74-6.89 (m, 8H), 6.98-7.05 (m, 4H), 7.36 (d, J ) 8.3 Hz, 1H), 7.45 (d, J ) 8.0 Hz, 1H), 7.73 (d, J ) 7.7 Hz, 1H), 7.78 (s, 1H), 8.10 (s, 1H). 13C NMR (CDCl3, 125.770 MHz): 157.02, 153.22, 150.88, 149.39, 147.93, 147.54, 134.01, 132.24, 129.40, 128.96, 125.42, 124.69,

Letters SCHEME 1: Synthetic Pathway Leading to the Dyes 4a-4d

J. Phys. Chem. C, Vol. 113, No. 20, 2009 8543

TABLE 1: Electro-optical Properties of the Dyes


compound
4a 4b 4c 4d

max, nm (max, M-1 cm-1 103)a


405 (20.9) 395 (23.5) 424 (25.6), 305 (26.0) 409 (32.9), 292 (18.9)

TiO abs 2,

nm
394 395 416 398

em, nmb
493 479 487 488

Stokes shift, cm-1


4407 4440 3051 3958

Eox (Ep), mVa


600c 608c 541 (106) 560 (144)

HOMO, eVd
5.400 5.408 5.341 5.360

LUMO, eVd
2.844 2.825 2.910 2.855

E0-0, eV
2.556 2.583 2.431 2.505

E0-0*, Ve
-1.19 -1.21 -1.12 -1.18

a Measured for approximately 2.0 10-5 M dichloromethane solutions. b Measured in toluene. c Measured for dichloromethane solutions using tetrabutylammonium hexauorophosphate (TBAHP) as the supporting electrolyte at a scan rate of 100 mV/s. d Deduced from the equations HOMO ) Eox + 4.8 and LUMO ) HOMO - E0-0. e Excited-state oxidation potential versus NHE.

more electron-rich than the biphenyl unit, it may reinforce the donor-acceptor interaction by substantiating the donor strength of the amino group.16 Within the class, the dye containing diphenylamine donor possesses a red-shifted charge-transfer absorption with respect to the dye with the 1-naphthylphenylamine donor. This is probably caused by the competitive delocalization of amine lone pair electrons by the naphthalene unit, which diminishes the donor-acceptor interaction appreciably. These interpretations are further supported by the following observations: (a) the biphenyl-based dyes (4a and 4b) possess comparatively lower molar extinction coefcients than the uorene-based analogues (4c and 4d). The extinction coefcient of the charge-transfer transition band in the series increases in the order 4a < 4b < 4c < 4d. (b) Additionally, the dyes containing 1-naphthylphenylamine donors (4b and 4d) display hypochromic and hypsochromic effects when compared to the analogous diphenylamine-based dyes. The blue shift in absorption maximum and the increase in extinction coefcient for the charge-transfer transition band may not arise due to H-aggregation in the present dyes as they are less at and may not be prone for facile H aggregation.17 Also, at the low concentrations (2.0 10-5 M) used in the present study, no dye aggregation is detected, and the dyes obeyed Beers law (see Supporting Information). The decrease of extinction coefcient for the biphenyl dyes may be attributed to the decrease in the coplanarity between the electron donor and the electron acceptor, which would eventually lower the magnitude of the transition probability. An alternative explanation for the blue shift observed for the dyes in polar aprotic solvents may be explained on the basis of protonation and deprotonation of the dyes. It is possible that the carboxylic acid is partially deprotonated in the ground state.

Figure 1. Absorption spectra of the dyes (4a-4d) recorded in dichloromethane.

Such a deprotonation would diminish the electron-accepting ability of the acceptor moiety, which in turn may substantially decrease the donor-acceptor interaction in the dye. This will manifest as a blue shift in the absorption spectrum. This hypothesis is further supported by the fact that addition of a base such as triethylamine to the toluene solution of the dye (4d) blue shifted the absorption peak and that addition of an acid (triuoroacetic acid) to the toluene solution of the dye led to a red shift (Figure 2 and Table 2). These observations suggest the presence of a protonation-deprotonation equilibrium for the dyes in solution. It is speculated that in toluene solution, the dyes are predominantly in the protonated form (DH). However,

8544

J. Phys. Chem. C, Vol. 113, No. 20, 2009

Letters

Figure 2. Absorption spectra of 4d in toluene and dimethylformamide before and after the addition of triethylamine and triuoroacetic acid, respectively.

Figure 3. Emission spectra of the dyes (4a-4d) recorded for toluene solutions.

TABLE 2: Absorption Data for the Dyes Recorded in Toluene and Dimethylformamide in the Presence of Triethylamine/Triuoroacetic Acid
max, nm (max, M-1 cm-1 103) dye toluene toluene + TEA toluene + TFA DMF DMF + TFA
4a 4b 4c 4d 426 (18.4) 420 (17.9) 450 (23.6) 438 (27.6) 380 (23.0) 379 (24.7) 399 (26.7) 401 (32.7) 451 (21.4) 448 (20.0) 479 (26.4) 477 (32.0) 373 (29.0) 372 (31.3) 389 (30.6) 393 (35.4) 415 (24.4) 411 (24.6) 437 (25.1) 438 (31.2)

addition of triethylamine deprotonated the dye and shifted the equilibrium to the deprotonated form (D-) side, while the addition of triuoroacetic acid reversed the equilibrium and increased the amount of protonated form in the solution. The pKa of the conjugate acid of triphenylamine is -5.0, while that of triuoroacetic acid is 0.5. It is clear from this data that the triuoroacetic acid is a weaker acid than the protonated triphenylamine. Therefore, it is unlikely that the amine unit will be protonated in the presence of triuoroacetic acid. Absorption spectra of the dyes recorded in dimethylformamide displayed a more blue-shifted absorption prole than that of the toluene solution. On this basis, it was expected that in basic dimethylformamide solution, the deprotonated form might be present in larger quantities and that addition of triuoroacetic acid would push the equilibrium to the protonated form side. As speculated, addition of triuoroacetic acid to the dimethylformamide solution of the dye resulted in a red shift for the absorption maximum (Figure 2). The same behavior was noticed for the other three dyes also (Table 2). All of the dyes except 4b when adsorbed on the TiO2 surface displayed a blue-shifted absorption prole, which probably arises due to the deprotonation of the carboxylic acid group.6d,e,18 The reason for the anomalous behavior observed for 4b is unclear.

Figure 4. Cyclic voltammograms of the uorene-based dyes 4c and 4d.

DH (Nonpolar Solvents)

y z \
Acid

Base

D- + H+ (Polar Solvents)

The emission spectra of the dyes were recorded in toluene solutions and are depicted in Figure 3. The dyes emit a greenishyellow color in toluene; however, their emission is completely quenched in polar solvents such as dimethylformamide. This probably indicates the presence of a noticeable dipole in the donor-acceptor compounds due to extensive polarization, which

may interact with the solvent dipole and trigger a nonradiative relaxation of the excited state by electron transfer. 3.3. Electrochemical Properties. The redox behavior of the dyes was examined in dichloromethane using cyclic and differential pulse voltammetric methods. All of the dyes show an oxidation wave attributable to the removal of an electron from the amine segment. The cyclic voltammograms observed for the uorene-based dyes (4c and 4d) are displayed in Figure 4. The uorene-based dyes (4c and 4d) displayed a cathodically shifted quasi-reversible oxidation process when compared to the biphenyl-based dyes (4a and 4b). The biphenyl-based dyes exhibited irreversible oxidation waves. The exact oxidation potential for the biphenyl dyes was extracted from the differential pulse voltammograms. The negatively shifted oxidation observed for the uorene-based dyes is attributed to the electron richness of the uorene unit, which may increase the oxidation propensity of the amine unit. Incorporation of an additional uorene segment has been demonstrated to signicantly decrease the oxidation potential of the amines.16,19 Substitution of uorene unit for the other aromatic units (phenyl, naphthyl) in the triarylamines has been reported to decrease the oxidation potential dramatically. The oxidation potential of the present

Letters

J. Phys. Chem. C, Vol. 113, No. 20, 2009 8545


TABLE 3: DSSC Performance Parameters of the Dyesa
dye 4a 4b 4c 4d FlDye N719 VOC, V 0.68 0.69 0.70 0.73 0.69 0.73 JSC, mA cm-2 10.73 11.18 11.28 12.28 17.00 15.81 , % 4.98 5.24 5.23 6.00 7.19 7.53 ff 0.68 0.68 0.66 0.67 0.62 0.65

a DSSC performance measurements were conducted with TiO2 photoelectrodes with an approximately 18 m thickness and 0.25 cm2 working area on the FTO (15 /sq) substrates.

Figure 5. Differential pulse voltammograms recorded for 4d in dichloromethane solutions before (violet) and after (blue) the addition of triuoroacetic acid.

dyes is anodically shifted when compared to the simple triarylamines without an electron-accepting cyanoacrylic acid unit.16,20 This clearly indicates the existence of electronic communication between the amine donor and cyanoacrylic acid acceptor. It is interesting to note that the oxidation potentials observed for 4c and 4d are more positive than that observed for FlDye, which contains a thiophene unit between the uorene and acceptor units.13 Stronger interaction between the donor and acceptor will make the oxidation of the amine difcult. Reduced donor strength observed for the naphthyl-containing amines in the absorption spectra (vide supra) is also corroborated by the electrochemical studies. The naphthylamine-containing dyes displayed more positive oxidation potentials than those of the diphenylamine-based dyes. The presence of acid-base equilibrium for the dyes in the solutions can also be identied by electrochemical studies. It was believed that the addition of triuoroacetic acid to the dye solution would shift the equilibrium to the protonated form side and ensure larger amounts of protonated form available at the electrode surface. The oxidation potential of the protonated form will be anodically shifted more than that of the deprotonated form due to the enhanced donor-acceptor interaction in the former. As expected, the oxidation potential of the dyes increased upon the addition of triuoroacetic acid to the dichloromethane solutions (Figure 5). Protonation at the amine segment may also bring such a change in oxidation potential, however with a larger magnitude. In view of the above inferences, which suggests that deprotonation may shift the oxidation potential of the dye cathodically compared to the protonated state and that deprotonation may occur upon anchoring the dyes on the TiO2 surface, interpretations about the feasibility of charge separation for the dyes attached to the TiO2 surface and electron injection dynamics must be done with due allowances. To identify the feasibility of electron injection from the dye to TiO2, we have estimated the excited-state redox potential (E* ) ox of the sensitizers from the rst oxidation potential (Eox) at the ground state and the zero-zero electronic transition energy (E0-0) as given below7e

Figure 6. I-V curves of the DSSCs based on the dyes 4a-4d and the reference compound N719.

E * ) Eox - E0-0 ox
The optical band gap was derived from the absorption edge. For an efcient electron injection to the TiO2 electrode, the dyes

must possess an excited-state potential lower than the conduction band edge energy. This will ensure an energetically favorable downhill electron injection. The E* values observed for the dyes ox (-1.12 to -1.21 V versus NHE) are more negative than the conduction band edge energy level of the TiO2 electrode (-0.5 V versus NHE).10d The E* values predict a facile electron ox injection from the dye to the TiO2 electrode. An alteration of redox potential by 0.2 V due to deprotonation is not going to signicantly change the electron ow from the present dyes. The dye regeneration is essential to achieve a high stability and operational lifetime for the DSSCs. Efcient dye regeneration is possible if the ground-state redox potential of the dye is more positive than that of the electrolyte redox couple. For the dyes reported here, the oxidation potentials lie in the range of 1.31-1.38 V versus NHE. These values are anodically shifted to that of the electrolyte (I-/I3-) redox couple (0.4 V versus NHE) and suggest that the dye regeneration is feasible with the selected electrolyte. 3.4. Performance of Dye-Sensitized Solar Cells. Due to the promising photophysical and redox properties of the dyes, we have tested their utility in dye-sensitized solar cells. The dyesensitized solar cells were constructed by using 4a-4d as the sensitizer for nanocrystalline anatase TiO2. Typical solar cells, with an effective area of 0.25 cm2, were fabricated with an electrolyte composed of 0.05 M I2/0.5 M LiI/0.5 M tertbutylpyridine in acetonitrile solution. The device performance statistics under AM 1.5 illumination are collected in Table 3. Figure 6 shows the photocurrent-voltage (I-V) curves of the cells. The incident photon-to-current conversion efciencies (IPCE) of the dyes on TiO2 are plotted in Figure 7. The devices of the dyes exhibited good conversion efciencies, ranging from 66 to 80% of the standard device fabricated using the ruthenium

8546

J. Phys. Chem. C, Vol. 113, No. 20, 2009

Letters behavior between the biphenyl- and uorene-based dyes manifest in the device performance of the dyes in the DSSCs. Fluorene-based dyes show better solar cell parameters than those of the biphenyl analogues, owing to the efcient donor-acceptor interaction and charge migration in the former arising due to the coplanar structure and effective conjugation, and the latter helps to enhance the absorption ability of the dyes. The higher light-to-electricity efciency realized for the simple dyes reported here reveals that metal-free organic dyes are promising in the development of DSSCs. Computational and experimental investigations to unravel the electron-transfer and recombination process in the dyes are in progress. Acknowledgment. The Department of Science and Technology (Grant No. SR/S1/OC-11/2007), New Delhi, and the Indian Institute of Technology Roorkee (Initiation Grant) have supported this work. A.B. thanks the Indian Institute of Technology Roorkee for the MHRD assistantship.

Figure 7. IPCE plots of the DSSCs fabricated using the dyes 4a-4d and N719.

dye N7196f and measured under similar conditions. The DSSC efciency observed for 4a is signicantly higher than that reported for the phenyl conjugated analogue, 2-cyano-3-(4(diphenylamino)phenyl)acrylic acid.21 This is attributed to the extension of conjugation in 4a by a phenyl segment, which presumably benets the molecule by contributing to the increase in molar extinction coefcient (20900 versus 15800 M-1 cm-1) and probably to the retardation of the charge recombination. In general, the uorene-based dyes are exhibiting comparatively higher device efciencies than the biphenyl-based dyes. This may probably stem from the higher molar extinction coefcient of the uorene-based dyes, which can generate more excited electrons under the same wavelength light irradiation for uorene-based dyes than the biphenyl-based dyes. The DSSC based on 4d shows better device characteristics, with an open circuit voltage of 0.73 V, a short circuit photocurrent density of 12.28 mA cm-2, a ll factor of 0.67, and an overall lightto-electricity conversion efciency of 6.0% under a 1.5 AM irradiation. For a given conjugation, the naphthylamine-containing dyes (4b and 4d) are showing more promising device efciencies than the diphenylamine-featuring dyes (4a and 4c). This may also be rationalized as a result of originating from the enhanced absorption capability of the naphthylaminecontaining dyes (4b and 4d). The IPCE spectrum of 4d shows a maximum value close to 75%, which is consistent with the larger optical density observed in the absorption spectrum. The Voc values of the uorene-based dyes (4c and 4d) are larger than those realized for biphenyl-based dyes (4a and 4b). This may also help to improve the efciency of the DSSCs. 4. Conclusions In this paper, we have reported the synthesis of hitherto unknown simple and efcient organic dyes featuring biphenyl and uorene conjugation suitable for dye-sensitized solar cells based on nanocrystalline anatase titanium dioxide. The nature of the conjugation signicantly affects the absorption and electrochemical properties of the dyes. The differences observed for the biphenyl and uorene dyes in the absorption properties are attributed to the planarity and rigidity of the uorene unit, while the facile oxidation of uorene-based dyes originate from the electron richness of the uorene segment. The signicant differences noticed in the absorption and electrochemical

Supporting Information Available: 1H NMR of the newly synthesized compounds and absorption proles of the dyes recorded in toluene and dimethylformamide and in the presence of triuoroacetic acid or triethylamine. This material is available free of charge via the Internet at http://pubs.acs.org. References and Notes
(1) (a) Balzani, V.; Credi, A.; Venturi, M. ChemSusChem 2008, 1, 26 58. (b) Butler, D. Nature 2008, 454, 558. (c) Lewis, N. S. Science 2007, 315, 798. (d) Hagfeldt, A.; Gratzel, M. Acc. Chem. Res. 2000, 33, 269. (2) Robertson, N. Angew. Chem., Int. Ed. 2006, 45, 2338. (3) (a) Snaith, H. J.; Schmidt-Mende, L. AdV. Mater. 2007, 19, 3187. (b) de la Torre, G.; Claessens, C. G.; Torres, T. Chem. Commun. 2007, 2000. (4) Armaroli, N.; Balzani, V. Angew. Chem., Int. Ed. 2007, 46, 52. (5) (a) Jang, S. R.; Yum, J. H.; Klein, C.; Kim, K. J.; Wagner, P.; Ofcer, D.; Gratzel, M.; Nazeeruddin, M. K. J. Phys. Chem. C 2009, 113, 1998. (b) Choi, H.; Baik, C.; Kim, S.; Kang, M. S.; Xu, X.; Kang, H. S.; Kang, S. O.; Ko, J; Nazeeruddin, M. K.; Gratzel, M. New. J. Chem. 2008, 32, 2233. (c) Abbotto, A.; Barolo, C.; Bellotto, L.; De Angelis, F; Gratzel, M.; Manfredi, N.; Marinzi, C.; Fantacci, S.; Yum, J. H.; Nazeeruddin, M. K. Chem. Commun. 2008, 5318. (d) Gao, F. F.; Wang, Y.; Zhang, J.; Shi, D.; Wang, M. K.; Humphry-Baker, R.; Wang, P.; Zakeeruddin, S. M.; Gratzel, M. Chem. Commun. 2008, 2635. (e) Nazeeruddin, M. K.; Gratzel, M. Struct. Bonding (Berlin) 2007, 123, 113. (f) Gratzel, M. Chem. Lett. 2005, 34, 8. (g) Gratzel, M. Nature 2001, 414, 338. (h) ORegan, B.; Gratzel, M. Nature 1991, 353, 737. (6) (a) Wang, M.; Chen, P.; Humphry-Baker, R.; Zakeeruddin, S. M.; Gratzel, M. ChemPhysChem 2009, 10, 290. (b) Lee, C.; Yum, J. H.; Choi, H.; Kang, S. O.; Ko, J.; Humphry-Baker, R.; Gratzel, M.; Nazeeruddin, M. K. Inorg. Chem. 2008, 47, 2267. (c) Nazeeruddin, M. K.; De Angelis, F.; Fantacci, S.; Selloni, A.; Viscardi, G.; Liska, P.; Ito, S.; Takeru, B.; Gratzel, M. J. Am. Chem. Soc. 2005, 127, 16835. (d) Nazeeruddin, M. K.; Pechy, P.; Renouard, T.; Zakeeruddin, S. M.; Humphry-Baker, R.; Comte, P.; Liska, P.; Cevey, L.; Costa, E.; Shklover, V.; Spiccia, L.; Deacon, G. B.; Bignozzi, C. A.; Gratzel, M. J. Am. Chem. Soc. 2001, 123, 1613. (e) Nazeeruddin, M. K.; Zakeeruddin, S. M.; Humphry-Baker, R.; Jirousek, M.; Liska, P.; Vlachopoulos, N.; Shklover, V.; Fisher, C. H.; Gratzel, M. Inorg. Chem. 1999, 38, 6298. (f) Nazeeruddin, M. K.; Kay, A.; Rodicio, I.; Humphry-Baker, R.; Muller, E.; Liska, P.; Vlachopoulos, N.; Gratzel, M. J. Am. Chem. Soc. 1993, 115, 6382. (7) (a) Ito, S.; Miura, H.; Uchida, S.; Takata, M.; Sumioka, K.; Liska, P.; Comte, P.; Pechy, P.; Gratzel, M. Chem. Commun. 2008, 5194. (b) Hwang, S.; Lee, J. H.; Park, C.; Lee, H.; Kim, C.; Park, C.; Lee, M.-H.; Lee, W.; Park, J.; Kim, K.; Park, N.-G.; Kim, C. Chem. Commun. 2007, 4887. (c) Karthikeyan, C. S.; Wietasch, H.; Thelakkat, M. AdV. Mater. 2007, 19, 1091. (d) Kuang, D. B.; Klein, C.; Ito, S.; Moser, J. E.; HumphryBaker, R.; Zakeeruddin, S. M.; Gratzel, M. AdV. Funct. Mater. 2007, 17, 154. (e) Klein, C.; Nazeeruddin, M. K.; Liska, P.; Di Censo, D.; Hirata, N.; Palomares, E.; Durrant, J. R.; Gratzel, M. Inorg. Chem. 2005, 44, 178. (f) Klein, C.; Nazeeruddin, M. K.; Di Censo, D.; Liska, P.; Gratzel, M. Inorg. Chem. 2004, 43, 4216. (g) Wang, P.; Humphry-Baker, R.; Moser, J. E.; Zakeeruddin, S. M; Gratzel, M. Chem. Mater. 2004, 16, 3246. (8) (a) Hagberg, D. P.; Marinado, T.; Karlsson, K. M.; Nonomura, K.; Qin, P.; Boschloo, G.; Brinck, T.; Hagfeldt, A.; Sun, L. J. Org. Chem. 2007, 72, 9550. (b) Ito, S.; Zakeeruddin, S. M.; Humphry-Baker, R.; Liska,

Letters
P.; Charvet, R.; Comte, P.; Nazeeruddin, M. K.; Pechy, P.; Takada, M.; Miura, H.; Uchida, S.; Gratzel, M. AdV. Mater. 2006, 18, 1202. (c) Hara, K.; Sato, T.; Katoh, R.; Furube, A.; Yoshihara, T.; Murai, M.; Kurashige, M.; Ito, S.; Shinpo, A.; Suga, S.; Arakawa, H. AdV. Funct. Mater. 2005, 15, 246. (d) Schmidt-Mende, L.; Bach, U.; Humphry-Baker, R.; Horiuchi, T.; Miura, H.; Ito, S.; Uchida, S.; Gratzel, M. AdV. Mater. 2005, 17, 813. (e) Horiuchi, T.; Miura, H.; Sumioka, K.; Uchida, S. J. Am. Chem. Soc. 2004, 126, 12218. (f) Kitamura, T.; Ikeda, M.; Shigaki, K.; Inoue, T.; Anderson, N. A.; Ai, X.; Lian, T.; Yanagida, S. Chem. Mater. 2004, 16, 1806. (g) Horiuchi, T.; Miura, H.; Uchida, S. Chem. Commun. 2003, 3036. (h) Hara, K.; Kurashige, M.; Ito, S.; Shinpo, A.; Suga, S.; Sayama, K.; Arakawa, H. Chem. Commun. 2003, 252. (i) Yao, Q.-H.; Shan, L.; Li, F.Y.; Yin, D.-D.; Huang, C.-H. New J. Chem. 2003, 27, 1277. (j) Sayama, K.; Tsukagoshi, S.; Hara, K.; Ohga, Y.; Shinpou, A.; Abe, Y.; Suga, S.; Arakawa, H. J. Phys. Chem. B 2002, 106, 1363. (k) Sayama, K.; Hara, K.; Mori, N.; Satsuki, M.; Suga, S.; Tsukagoshi, S.; Abe, Y.; Sugihara, H.; Arakawa, H. Chem. Commun. 2000, 1173. (l) Wang, Z.-S.; Li, F.-Y.; Huang, C.-H. Chem. Commun. 2000, 2063. (9) (a) Thomas, K. R. J.; Hsu, Y. C.; Lin, J. T.; Lee, K. M.; Ho, K. C.; Lai, C. H.; Cheng, Y. M.; Chou, P. T. Chem. Mater. 2008, 20, 1830. (b) Liu, W.-H.; Wu, I.-C.; Lai, C.-H.; Lai, C.-H.; Chou, P.-T.; Li, Y.-T.; Chen, C.-L.; Hsu, Y.-Y.; Chi, Y. Chem. Commun. 2008, 5152. (c) Wang, M.; Xu, M.; Shi, D.; Li, R.; Gao, F.; Zhang, G.; Yi, Z.; Humphry-Baker, R.; Wang, P.; Zakeeruddin, S. M.; Gratzel, M. AdV. Mater. 2008, 20, 4460. (d) Qin, H.; Wenger, S.; Xu, M.; Gao, F.; Jing, X.; Wang, P.; Zakeeruddin, S. M.; Gratzel, M. J. Am. Chem. Soc. 2008, 130, 9202. (e) Kim, S.; Lee, J. W.; Kang, S. O.; Ko, J.; Yum, J.-H.; Fantacci, S.; De Angellis, F.; Di Censo, D.; Nazeeruddin, M. K.; Gratzel, M. J. Am. Chem. Soc. 2006, 128, 16701. (10) (a) Wang, W. S.; Cui, Y.; Dan-Oh, Y.; Kasada, C.; Shinpo, A.; Hara, K. J. Phys. Chem. C 2008, 112, 17011. (b) Wang, W. S.; Cui, Y.; Dan-Oh, Y.; Kasada, C.; Shinpo, A.; Hara, K. J. Phys. C 2007, 111, 7224. (c) Wang, W. S.; Cui, Y.; Hara, K.; Dan-Oh, Y.; Kasada, C.; Shinpo, A. AdV. Mater. 2007, 19, 1138. (d) Hara, K.; Sato, T.; Katoh, R.; Furube, A.; Ohga, Y.; Shinpo, A.; Suga, S.; Sayama, K.; Sugihara, H.; Arakawa, H. J. Phys. Chem. B 2003, 107, 597. (e) Hara, K.; Kurashige, M.; Dan-Oh, Y.; Kasada, C.; Shinpo, A.; Suga, S.; Sayama, K.; Arakawa, H. New J. Chem. 2003, 27, 783. (f) Hara, K.; Sayama, K.; Ohga, Y.; Shinpo, A.; Suga, S.; Arakawa, H. Chem. Commun. 2001, 569. (11) (a) Liu, X. Z.; Zhu, R.; Zhang, Y.; Liu, B.; Ramakrishna, S. Chem. Commun. 2008, 3789. (b) Velusamy, M.; Thomas, K. R. J.; Lin, J. T.; Hsu, Y.-C.; Ho, K.-C. Org. Lett. 2005, 7, 1899. (12) (a) Zhao, G.; Wu, G.; He, C.; Bai, F. Q.; Xi, H.; Zhang, H. X.; Li, Y. J. Phys. Chem. C 2009, 113, 2636. (b) Tao, Y.; Wang, Q.; Shang, Y.; Yang, C.; Ao, L.; Qin, J.; Ma, D.; Shuai, Z. Chem. Commun. 2009, 77. (c) Yen, Y. S.; Hsu, Y. C.; Lin, J. T.; Chang, C. W.; Hsu, C. P.; Yin, D. J. J. Phys. Chem. C 2008, 112, 12557. (d) Qin, P.; Zhu, H.; Edvinsson, T.; Boschloo, G.; Hagfeldt, A.; Sun, L. J. Am. Chem. Soc. 2008, 130, 8570. (e) Li, G.; Jiang, K J.; Li, Y. F.; Li, S. L.; Yang, L. M. J. Phys. Chem. C 2008, 112, 11591. (f) Shi, D.; Cao, Y.; Pootrakulchote, N.; Yi, Z.; Xu, M.; Zakeeruddin, S. M.; Gratzel, M.; Wang, P. J. Phys. Chem. C 2008, 112, 17478. (g) Xu, M.; Li, R.; Pootrakulchote, N.; Shi, D.; Guo, J.; Yi, Z.;

J. Phys. Chem. C, Vol. 113, No. 20, 2009 8547


Zakeeruddin, S. M.; Gratzel, M.; Wang, P. J. Phys. Chem. C 2008, 112, 19770. (h) Tian, H.; Yang, X.; Chen, R.; Zhang, R.; Hagfeldt, A.; Sun, L. J. Phys. Chem. C 2008, 112, 11023. (i) Budy, S. M.; Suresh, S.; Spraul, B. K.; Smith, J. D. W. J. Phys. Chem. C 2008, 112, 8099. (j) Holzapfel, M.; Lambert, C. J. Phys. Chem. C 2008, 112, 1227. (k) Ning, Z.; Zhang, Q.; Wu, W.; Pei, H.; Liu, B.; Tian, H. J. Org. Chem. 2008, 73, 3791. (l) Zhou, G.; Pschirer, N.; Schoneboom, J. C.; Eickemeyer, F.; Baumgarten, M.; Mullen, K. Chem. Mater. 2008, 20, 1808. (m) Zhou, G.; Baumgarten, M.; Mullen, K. J. Am. Chem. Soc. 2008, 130, 12477. (n) Fron, E.; Bell, T. D. M.; Vooren, A. V.; Schweitzer, G.; Cornil, J.; Beljonne, D.; Toele, P.; Jacob, Jo.; Mullen, K.; Hofkens, J.; Auweraer, M. V.; Schryver, F. C. D J. Am. Chem. Soc. 2007, 129, 610. (o) Zhou, G.; Baumgarten, M.; Mullen, K. J. Am. Chem. Soc. 2007, 129, 12211. (p) Johansson, E. M. J.; Edvinsson, T.; Odelius, M.; Hagberg, D. P.; Sun, L.; Hagfeldt, A.; Siegbahn, H.; Rensmo, H. J. Phys. Chem. C 2007, 111, 8580. (q) Edvinsson, T.; Li, C.; Pschirer, N.; Schneboom, J.; Eickemeyer, F.; Sens, R.; Boschloo, G.; Herrmann, A.; Mllen, K.; Hagfeldt, A. J. Phys. Chem. C 2007, 111, 15137. (r) Hwang, S.; Lee, J. H.; Park, C.; Lee, H.; Kim, C.; Park, C.; Lee, M. H.; Lee, W.; Park, J.; Kim, K.; Park, N. G.; Kim, C. Chem. Commun. 2007, 4887. (s) Liang, M.; Xu, W.; Cai, F.; Chen, P.; Peng, B.; Chen, J.; Li, Z. J. Phys. Chem. C 2007, 111, 4465. (t) Liang, M.; Xu, W.; Cai, F.; Chen, P.; Peng, B.; Chen, J.; Li, Z. J. Phys. Chem. C 2007, 111, 4465. (u) Amthor, S.; Lambert, C.; Dmmler, S.; Fischer, I.; Schelter, J. J. Phys. Chem. A 2006, 110, 5204. (13) Thomas, K. R. J.; Lin, J. T.; Hsu, Y.-C.; Ho, K.-C. Chem. Commun. 2005, 4098. (14) (a) Hartwig, J. F.; Kawatsura, M.; Hauck, S. I.; Shaughnessy, K. H.; Alcazar-Roman, L. M. J. Org. Chem. 1999, 64, 5575. (b) Driver, M. S.; Hartwig, J. F. J. Am. Chem. Soc. 1996, 118, 7217. (15) Li, G.; Jiang, K. J.; Li, Y. F.; Li, S. L.; Yang, L. M. J. Phys. Chem. C 2008, 112, 11591. (16) Hreha, R. D.; George, C. P.; Haldi, A.; Domercq, B.; Malagoli, M.; Barlow, S.; Bredas, J. L.; Kippelen, B.; Marder, S. R. AdV. Funct. Mater. 2004, 13, 967. (17) (a) Mann, J. R.; Gannon, M. K.; Fitzgibbons, T. C.; Detty, M. R.; Watson, D. F. J. Phys. Chem. C 2008, 112, 13057. (b) Bujdak, J.; Iyi, N. J. Phys. Chem. B 2005, 109, 4608. (18) Yen, Y. S.; Hsu, Y. C.; Lin, J. T.; Chang, C. W.; Hsu, C. P.; Yin, D. J. J. Phys. Chem. C 2008, 112, 12557. (b) Hagberg, D. P.; Edvinsson, T.; Marinado, T.; Boschloo, G.; Hagfeldt, A.; Sun, L. Chem. Commun. 2006, 2245. (19) (a) Li, Z. H.; Wong, M. S.; Fukutani, H.; Tao, Y. Org. Lett. 2006, 8, 4271. (b) Wong, W. Y.; Zhou, G. J.; Yu, X. M.; Kwok, H. S.; Tang, B. Z. AdV. Funct. Mater. 2006, 16, 838. (c) Thomas, K. R. J.; Velusamy, M.; Lin, J. T.; Chuen, C. H.; Tao, Y. T. Chem. Mater. 2005, 17, 1860. (20) Li, Z. H.; Wong, M. S.; Tao, Y.; DLorio, M. J. Org. Chem. 2004, 69, 921. (21) Xu, W.; Peng, B.; Chen, J.; Liang, M.; Cai, F. J. Phys. Chem. C 2008, 112, 874.

JP902206G

You might also like