Sie sind auf Seite 1von 6

Job/Unit: I20474

/KAP1

Date: 17-07-12 17:53:33

Pages: 6

SHORT COMMUNICATION
DOI: 10.1002/ejic.201200474

Cationic Iridium Complexes Coordinated with Coumarin Dyes Sensitizers for Visible-Light-Driven Hydrogen Generation
Shin-ya Takizawa,*[a] Csar Prez-Bolvar,[b] Pavel Anzenbacher, Jr.,[b] and Shigeru Murata*[a]
Keywords: Iridium / Photochemistry / Sensitizers / Hydrogen / Cobalt
Cationic iridium(III) complexes comprising two cyclometalating 3-(2-benzothiazolyl)-7-(diethylamino)coumarin (Coumarin 6) ligands and one diimine ancillary ligand (N N) were prepared [N N = 2,2 -bipyridyl (2), 1,10-phenanthroline (3), 2,2 -biquinoline (4)] and examined for their photophysical properties. The synergic effect of the intraligand chargetransfer (ILCT) transition in the chromophoric Coumarin 6 moiety and the metal-ligand-to-ligand CT (MLLCT) and/or ligand-to-ligand CT (LLCT) transitions associated with the diimine ligand rendered these complexes strongly absorptive in the visible region. Their enhanced absorptivity is significant when compared to the corresponding electroneutral complex 1 with acetylacetonate. Specifically, the molar extinction coefficient of complex 2 reached 129000 M1 cm1 at 483 nm, which is an unprecedentedly high value relative to those of other recently reported Ir-based dyes. Ir sensitizers 24 were tested for visible-light-driven hydrogen generation by using an IrCo photocatalytic system, where triethylamine and tris(2,2 -bipyridyl)dichloridocobalt were employed as a sacrificial electron donor and a water-reduction catalyst, respectively. As a result, complexes 2 and 3 were found to be highly effective sensitizers displaying turnover numbers up to 1500.

Introduction
Cyclometalated iridium(III) complexes have garnered wide attention as emitting materials owing to their advantageous photophysical properties such as high intersystemcrossing efficiency, high phosphorescence quantum yield, and long lifetime of the triplet excited states.[1] It has been shown that judicious ligand selection or ligand modification enables tuning of their properties including redox potentials, triplet energy, emission color, solubility, and stability. Such tunability can optimize the performance of Ir complexes for each target application.[2] While the most frequently studied were the electroneutral Ir materials, bis-cyclometalated cationic IrIII complexes have also been the subject of numerous studies in applications including light-emitting electrochemical cells (LECs),[3] singlet-oxygen generation,[4] bioimaging,[5] and as photoredox catalysts for organic synthesis[6] to name a few. Interestingly, these complexes are known to act as sensitizers for photoinduced hydrogen (H2) generation from water, which has been recognized as one of the promising solar energy conversion systems. Recently, Bernhard, Beller, Sun, and Wang have independently reported a variety of photoinduced H2 generation systems using iridium(III) bis(2phenylpyridinato-N,C2 )-2,2 -bipyridine hexafluorophosphate [Ir(ppy)2(bpy)]PF6 and its derivatives.[7] To increase the water solubility and the environmentally benign nature of the catalytic systems, the cationic Ir complexes can also be modified by changing the counter anions.[8] The application of Ir complexes, however, could be mired by their poor visible-light-harvesting ability because of the small molar extinction coefficients at long wavelengths. To maximize the utilization of solar energy through efficient light absorption, new cationic complexes 24 comprising two strongly absorbing 3-(2-benzothiazolyl)-7-(diethylamino)coumarin (Coumarin 6) ligands and a diimine ancillary ligand were prepared. Here we report on the unique photophysical properties of 24 and their successful application in visible-light-driven H2 generation. Figure 1 shows the structures of electroneutral complex 1[2a,9] and cationic complexes 24, as well as their absorption spectra featuring an intense absorption band in the visible region.
1

[a] Department of Basic Science, Graduate School of Arts and Sciences, The University of Tokyo Meguro-ku, Tokyo 153-8902, Japan Fax: +81-3-5454-6998 E-mail: ctaki@mail.ecc.u-tokyo.ac.jp cmura@mail.ecc.u-tokyo.ac.jp Homepage: http://maildbs.c.u-tokyo.ac.jp/~smurata/index-en.html [b] Department of Chemistry and Center for Photochemical Sciences, Bowling Green State University Bowling Green, OH 43403, USA Supporting information for this article is available on the WWW under http://dx.doi.org/10.1002/ejic.201200474.
Eur. J. Inorg. Chem. 0000, 00

0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Job/Unit: I20474

/KAP1

Date: 17-07-12 17:53:33

Pages: 6

SHORT COMMUNICATION

S.-y. Takizawa, C. Prez-Bolvar, P. Anzenbacher, Jr., S. Murata

Figure 1. Top: Chemical structure of Ir complexes 14, in which the second Coumarin 6 ligand is replaced with a symbol for clarity of the structure. Bottom: UV/Vis absorption spectra of 14 in CH2Cl2/MeOH (9:1 v/v).

Results and Discussion


Complexes 24 were successfully prepared by a two-step reaction, via the di--chlorido-bridged IrIII dimer of Coumarin 6, and characterized by 1H NMR spectroscopy, MS, and elemental analysis (see the Supporting Information for details). Complex 1[2a,9] with the acetylacetonate ancillary ligand was prepared for comparison. The photophysical and electrochemical data for the complexes are listed in Table 1. The UV/Vis absorption spectra of 24 show intense absorption bands in the visible region (430600 nm). Importantly, the molar absorptivity of complexes 2, 3 with 2,2 -bipyridyl and 1,10-phenanthroline ancillary ligands was much larger than that of electroneutral complex 1. Specifically, the molar extinction coefficient of 2 reached 129000 m1 cm1 at 483 nm, which is an unprecedentedly high value relative to those of other recently reported Ir-

based dyes.[4b,10] The enhanced absorptivity of 2 and 3 is presumably due to the combined effects of spin-allowed intraligand charge-transfer (ILCT) in the Coumarin ligand, metal-ligand-to-ligand CT (MLLCT) and/or ligand-to-ligand CT (LLCT) related to the diimine ligands. In complex 1, the absorption bands observed at 445 and 473 nm correspond to the ILCT of the Coumarin ligand as acetylacetonate does not contribute to the lowest energy electronic transition.[11] In addition to the enhanced extinction coefficients of the new cationic complexes, their endabsorption wavelengths were found to be redshifted relative to that of 1 (see Figure 1, bottom). In particular, the shift of complex 4 with 2,2 -biquinoline was the largest because of the extended -conjugation in the ancillary ligand. This is consistent with the lowest HOMOLUMO energy gap (E), caused by a positive shift in the reduction potential as observed by cyclic voltammetry (Table 1). In contradistinction, both the oxidation and reduction potentials of 1 3 were almost independent of the type of the ancillary ligands. It was found, however, that the absorption peak of complex 4 at 480 nm was exceptionally lower than those of the other cationic complexes 2, 3 (Figure 1 and Table 1) and close to that of electroneutral complex 1. This is interpreted by the assumption that the ILCT nature of the Coumarin 6 ligand is dominant in this spectral region (ca. 480 nm) whereas the MLLCT or LLCT transition, which is associated with the biquinoline ligand, likely participates in the large redshift of the end absorption. At room temperature, complexes 1, 2, and 3 exhibited phosphorescence maxima at 565, 589, and 589 nm, respectively (Figure S1), and the corresponding lifetimes were 6.57, 6.21, and 7.06 s. Notably, these lifetimes are considerably longer than that of [Ir(ppy)2(bpy)]PF6 (0.39 s),[7a] suggesting that the new complexes would be favorable for efficient electron transfer processes. The relatively long-lived triplet excited state, as well as the vibronic fine structure observed in the phosphorescence spectra of 13, indicates that the excited states of these complexes possess 3LC (triplet ligand-centered) nature rather than MLCT character.[2a] In contrast, complex 4 was not emissive at room temperature, possibly because of the existence of a fast nonradiative decay pathway, the origin of which is not clear at present. Conceivably, the energy gap law might account for the large nonradiative rate constant at room temperature, as the phosphorescence of 4 detected at 77 K was significantly redshifted relative to those of 13 (see Table 1 and Figure S1).[12]

Table 1. Photophysical and electrochemical properties of Ir complexes 14. Complex abs /nm ()[a] em /nm 1 2 3 4 473 483 485 480 (83200) (129000) (126000) (90000) 565 589 589 ND[f] 298 K[a] 0.60 0.26 0.18
[d]

77 K[a] /s 6.57 6.21 7.06 em /nm 560 576 579 644 /s 21.9 91.7 87.1 49.5 (40.9 %); 12.1 (59.1 %)

E1/2ox /V[b] 0.70[e] 0.68 0.66 0.73

E1/2red /V[b] ND[f] 1.62 1.57 1.18

E /V[c] 2.30 2.23 1.91

[a] In CH2Cl2/MeOH (9:1 v/v). [b] Determined by CV in CH3CN containing 0.1 m nBu4NClO4 and referenced against ferrocene/ferrocenium (Fc/Fc+). [c] E = E1/2ox. E1/2red. [d] Ref.[2a] [e] Determined in CH2Cl2 because of the low solubility in CH3CN. [f] ND: not detected. 2
www.eurjic.org 0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Eur. J. Inorg. Chem. 0000, 00

Job/Unit: I20474

/KAP1

Date: 17-07-12 17:53:33

Pages: 6

Cationic Iridium Complexes Coordinated with Coumarin Dyes

Ir sensitizers 24 were tested for visible-light-driven H2 generation by using a reported IrCo system,[7a,13] where triethylamine (TEA) and [Co(bpy)3]Cl2[7a,13,14] were employed as a sacrificial electron donor and a water-reduction catalyst, respectively. Each reaction contained 20.5 m of the sensitizer, 0.5 m TEA, and 2.56 mm [Co(bpy)3]Cl2 in CH3CN/H2O (1:1 v/v). Each solution (3 mL) was bubbled with argon for 20 min and then irradiated with visible light ( 440 nm) with a 500 W xenon arc lamp. Also, [Ir(ppy)2(bpy)]PF6 and a well-known sensitizer, [Ru(bpy)3]Cl2, were examined. The insolubility of complex 1 in CH3CN/H2O (1:1 v/v) did not allow for the photoinduced H2 generation reaction to take place under the conditions used for 24, thereby precluding a meaningful comparison. The time profile of light-driven H2 generation is shown in Figure 2, whilst Table 2 summarizes the quantity of H2 evolved and the turnover numbers (TON) with respect to the sensitizer and the Co catalyst. The reaction with 2 produced 0.511 mL of H2 corresponding to 1480 TON of 2, which is indicative of the high photosensitizing activity of this cationic complex. The reaction with 3 exhibited a rate of H2 evolution comparable to that of the system with 2, and the TON with respect to 3 was 1530. This result agrees well with the fact that sensitizers 2 and 3 have almost identical photophysical and electrochemical properties. Furthermore, experiments with the known sensitizers [Ir(ppy)2(bpy)]PF6 and [Ru(bpy)3]Cl2 resulted in the generation of lower amounts of H2 and lower turnover numbers (see Figure 2 and Table 2). Together, these experiments demonstrate the utility of new sensitizers 2 and 3 for visible-light-driven H2 generation.

Table 2. Results of the photoinduced H2 generation reactions.[a]


Sensitizer 2 3 4 [Ir(ppy)2(bpy)]PF6[b] [Ru(bpy)3]Cl2[b] H2 evolved /mol TONSens[c] 45.6 46.7 7.63 9.75 0.0309 1480 1530 248 315 1.01 TONCo[d] 11.9 12.2 1.99 2.54 0.00805

[a] Reaction conditions: TEA (0.5 m)/sensitizer (20.5 m)/[Co(bpy)3]Cl2 (2.56 mm)/HCl (0.24 m) in CH3CN/H2O (1:1 v/v, 3 mL), 440 nm, 6 h. [b] Irradiated with light of 390 nm. [c] Turnover numbers with respect to sensitizer. [d] Turnover numbers with respect to [Co(bpy)3]Cl2.

ceeds via a reductive quenching pathway[7a,7b] accompanied by a catalytic cycle of the Co complex.[14e,14f] First, a photoexcited IrIII complex undergoes reductive quenching by TEA, and the resultant IrII species reduces the CoII catalyst to give the CoI species. Ultimately, reaction of this Cobased active intermediate with protons may lead to H2 generation. It has been confirmed that the reduction potentials of the sensitizers excited states (Ered* = +0.99 V for 2; +1.03 V for 3,[15] vs. SCE) are high enough to promote oxidation of TEA (Eox = +0.71 V[16] vs. SCE), while the oxidation potentials of the IrII reductants (1.16 V for 2; 1.11 V for 3, vs. SCE) are low enough to promote the reduction of CoII (0.98 V[14d] vs. SCE).

Figure 2. Time profile of H2 generation for the photoreactions with complexes 24, [Ir(ppy)2(bpy)]PF6, and [Ru(bpy)3]Cl2. Reaction conditions: TEA (0.5 m)/sensitizer (20.5 m)/[Co(bpy)3]Cl2 (2.56 mm)/HCl (0.24 m) in CH3CN/H2O (1:1 v/v, 3 mL), 440 nm (sensitizers 24), 390 nm {[Ir(ppy)2(bpy)]PF6, [Ru(bpy)3]Cl2}.

When UV/Vis absorption spectra were recorded in the course of the reaction with 2, a broad absorption band appeared at around 610 nm as seen in Figure 3, indicating that a CoI species is formed in consequence of the reduction of the CoII catalyst.[14a,14b,14c] Considering that the concentration of TEA is much higher than that of [Co(bpy)3]Cl2, we propose that the photoinduced H2 generation reaction proEur. J. Inorg. Chem. 0000, 00

Figure 3. Top: Changes in the absorption spectra during the photoinduced H2 generation reaction with sensitizer 2. Bottom: Color change of the solution. (a) Before irradiation. (b) After irradiation for 1 h. (c) Exposed to air after 6 h of irradiation.

Further insights were obtained from looking at the changes in the absorption spectra (Figure 3). The absorbance at 610 nm gradually decreased after 3 h of irradiation, implying that the Co catalyst is partially degraded in the course of the reaction. However, its TON was determined to be 11.9, and this number demonstrates that
www.eurjic.org

0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Job/Unit: I20474

/KAP1

Date: 17-07-12 17:53:33

Pages: 6

SHORT COMMUNICATION
Entry 1 2 3[b] 4[c] 5[b] 6[b] 7[d] Electron donor 0.5 m (TEA) 0.5 m (TEA) 0.5 m (TEA) 0.05 m (TEA) 0.5 m (TEA) 0.5 m (TEOA) Sensitizer 2 20.5 m 20.5 m 20.5 m 20.5 m 20.5 m 20.5 m [Co(bpy)3]Cl2 2.56 mm 0.281 mm 2.56 mm 2.56 mm 2.56 mm 2.56 mm

S.-y. Takizawa, C. Prez-Bolvar, P. Anzenbacher, Jr., S. Murata

Table 3. Photoinduced H2 generation reactions with sensitizer 2 under various conditions.[a] H2 evolved /mol 45.6 39.0 0.284 22.8 0.00787 0 8.58 TONSens[e] 1480 1270 9.19 741 280 TONCo[f] 11.9 92.6 5.94 2.23

[a] Reaction conditions: electron donor/sensitizer 2 (20.5 m)/[Co(bpy)3]Cl2/HCl (0.24 m) in CH3CN/H2O (1:1 v/v, 3 mL), 440 nm, 6 h. [b] Irradiation time: 3 h. [c] HCl (0.024 m). [d] Irradiation time: 5.5 h. [e] Turnover numbers with respect to sensitizer 2. [f] Turnover numbers with respect to [Co(bpy)3]Cl2.

[Co(bpy)3]Cl2 can be recycled in this system despite its much higher concentration compared with that of the sensitizer. On the other hand, future investigations including the use of more durable catalysts[7b,7h,14f] could improve the catalytic systems with sensitizers 2 and 3. It is also important to note that the broad absorption band in the range 500800 nm disappeared completely when the reaction mixture was exposed to air after 6 h of irradiation and the dark-colored solution turned transparent yellow, as shown in Figure 3. This observation rules out the possibility that this system may produce heterogeneous catalytic colloidal metal or metal oxide sites during irradiation. Thus, the observed broad absorption is reasonably attributed to a reactive Co-based intermediate, which is most probably the CoI species.[14a,14b,14c] Nevertheless, more extensive research is required to reveal the catalytic mechanism of [Co(bpy)3]Cl2 that generates H2. In contradiction to the high activity of 2 and 3, complex 4 showed low sensitizing activity. Most likely, this originates from the fast nonradiative decay of the triplet excited state competing with the reductive quenching by TEA and preventing efficient H2 generation. Another plausible explanation is that the electron transfer from the IrII species (0.72 V vs. SCE) to the CoII catalyst (0.98 V[14d] vs. SCE) may not be favorable in view of their oxidation and reduction potentials, even though the photoexcited state of 4 (Ered* = +1.21 V vs. SCE) is energetic enough to be quenched by TEA. This suggests that the selection of an appropriate ancillary ligand is essential for achieving high photosensitizing ability in the new cationic Ir-based catalytic system. In order to obtain further information about this photocatalytic system, the influence of the concentration of the components on H2 evolution was investigated (Table 3). Notably, decreasing the concentration of [Co(bpy)3]Cl2 from 2.56 mm to 0.281 mm did not suppress the hydrogen evolution significantly; the catalytic cycle turned out to be durable even at the low concentration of the Co complex (compare entry 2 with entry 1). Irradiation of the solution in the absence of the Co complex proved to produce little hydrogen (entry 3). These results again highlight the crucial catalytic role of [Co(bpy)3]Cl2. Other control experiments confirmed that the three components are requisites for efficient photoinduced hydrogen generation, as shown in entries 5 and 6. On the other hand, when the TEA concentra4
www.eurjic.org

tion was reduced to 0.05 m from 0.5 m, half the amount of H2 evolved (compare entry 4 with entry 1). This is because the reductive quenching of photoexcited complex 2 is partially suppressed by the lowered electron-donor concentration. Finally, triethanolamine (TEOA),[7a,7b,13] which has often been used in place of TEA, was also evaluated as an electron donor. However, the system with TEOA produced less H2 relative to the reaction with TEA (entry 7 in Table 3, Figure S10). This result is rationalized by the fact that the phosphorescence of 2 was quenched by TEA whereas TEOA did not quench the phosphorescence. In short, TEOA is an unfavorable electron donor for reductively quenching the photoexcited state of 2 in this photoinduced H2 generation system.

Conclusions
We have synthesized new cationic IrIII complexes comprising two cyclometalating 3-(2-benzothiazolyl)-7-(diethylamino)coumarin ligands and one diimine ancillary ligand as sensitizers for visible-light-driven H2 generation. Due to the synergic effect of ILCT transitions in the chromophoric Coumarin 6 moiety and MLLCT/LLCT transitions related to the diimine ligand, these complexes show strong absorption in the visible region. Their enhanced absorptivity is significant when compared to the corresponding electroneutral acetylacetonate complex. Such a dramatic enhancement has not been reported in other IrIII complexes. Moreover, these cationic Ir complexes are highly effective sensitizers for visible-light-driven H2 generation and display turnover numbers up to 1500. It is expected that this work would pave a new avenue for the development of useful Ir sensitizers for practical solar energy conversion systems. Further optimization of the new complexes and reaction conditions is currently underway.

Experimental Section
Experimental details are given in the Supporting Information. Supporting Information (see footnote on the first page of this article): Experimental details; phosphorescence spectra and cyclic voltammograms of complexes 14; changes in the absorption spectra during the photoinduced H2 generation reactions with 3, 4, [Ir(ppy)2(bpy)]PF6, and [Ru(bpy)3]Cl2. UV/Vis absorption spectra
Eur. J. Inorg. Chem. 0000, 00

0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Job/Unit: I20474

/KAP1

Date: 17-07-12 17:53:33

Pages: 6

Cationic Iridium Complexes Coordinated with Coumarin Dyes of [Ir(ppy)2(bpy)]PF6, [Ru(bpy)3]Cl2, and Coumarin 6; time profile of H2 generation for the photoreactions with 2 under various conditions; quenching experiment of 2 with TEA. [6] J. M. R. Narayanam, C. R. J. Stephenson, Chem. Soc. Rev. 2011, 40, 102113. [7] a) J. I. Goldsmith, W. R. Hudson, M. S. Lowry, T. H. Anderson, S. Bernhard, J. Am. Chem. Soc. 2005, 127, 75027510; b) P. N. Curtin, L. L. Tinker, C. M. Burgess, E. D. Cline, S. Bernhard, Inorg. Chem. 2009, 48, 1049810506; c) S. Metz, S. Bernhard, Chem. Commun. 2010, 46, 75517553; d) B. F. DiSalle, S. Bernhard, J. Am. Chem. Soc. 2011, 133, 1181911821; e) F. Gartner, B. Sundararaju, A.-E. Surkus, A. Boddien, B. Loges, H. Junge, P. H. Dixneuf, M. Beller, Angew. Chem. 2009, 121, 10147; Angew. Chem. Int. Ed. 2009, 48, 99629965; f) F. Gartner, D. Cozzula, S. Losse, A. Boddien, G. Anilkumar, H. Junge, T. Schulz, N. Marquet, A. Spannenberg, S. Gladiali, M. Beller, Chem. Eur. J. 2011, 17, 69987006; g) P. Zhang, M. Wang, Y. Na, X. Li, Y. Jiang, L. Sun, Dalton Trans. 2010, 39, 12041206; h) P. Zhang, P.-A. Jacques, M. Chavarot-Kerlidou, M. Wang, L. Sun, M. Fontecave, V. Artero, Inorg. Chem. 2012, 51, 21152120. [8] a) A. Guerrero-Martinez, Y. Vida, D. Dominguez-Gutierrez, R. Q. Albuquerque, L. De Cola, Inorg. Chem. 2008, 47, 9131 9133; b) R. V. Kiran, C. F. Hogan, B. D. James, D. J. D. Wilson, Eur. J. Inorg. Chem. 2011, 48164825. [9] S. M. Borisov, I. Klimant, Anal. Chem. 2007, 79, 75017509. [10] a) K. Hanson, A. Tamayo, V. V. Diev, M. T. Whited, P. I. Djurovich, M. E. Thompson, Inorg. Chem. 2010, 49, 60776804; b) J. Sun, W. Wu, H. Guo, J. Zhao, Eur. J. Inorg. Chem. 2011, 31653173. [11] W. Wu, W. Wu, S. Ji, H. Guo, J. Zhao, Dalton Trans. 2011, 40, 59535963. [12] a) J. V. Caspar, T. J. Meyer, Inorg. Chem. 1983, 22, 24442453; b) J. S. Wilson, N. Chawdhury, M. R. A. Al-Mandhary, M. Younus, M. S. Khan, P. R. Raithby, A. Kohler, R. H. Friend, J. Am. Chem. Soc. 2001, 123, 94129417; c) R. Pohl, P. Anzenbacher Jr., Org. Lett. 2003, 5, 27692772. [13] S. Jasimuddin, T. Yamada, K. Fukuju, J. Otsuki, K. Sakai, Chem. Commun. 2010, 46, 84668468. [14] a) M. G. Simic, M. Z. Hoffman, R. P. Cheney, Q. G. Mulazzani, J. Phys. Chem. 1979, 83, 439443; b) C. V. Krishnan, N. Sutin, J. Am. Chem. Soc. 1981, 103, 21412142; c) C. Creutz, H. A. Schwarz, N. Sutin, J. Am. Chem. Soc. 1984, 106, 3036 3037; d) C. V. Krishnan, B. S. Brunschwig, C. Creutz, N. Sutin, J. Am. Chem. Soc. 1985, 107, 20052015; e) J. Dong, M. Wang, P. Zhang, S. Yang, J. Liu, X. Li, L. Sun, J. Phys. Chem. C 2011, 115, 1508915096; f) V. Artero, M. Chavarot-Kerlidou, M. Fontecave, Angew. Chem. Int. Ed. 2011, 50, 72387266. [15] The excited-state reduction potentials (Ered*) of complexes 2 and 3 were estimated by using Ered* = E1/2red + E00 (E1/2red: ground-state reduction potential vs. SCE, E00: triplet energy estimated from the phosphorescence data at 77 K). [16] L. R. Heeb, K. S. Peters, J. Phys. Chem. B 2008, 112, 219226. Received: May 8, 2012 Published Online:

Acknowledgment
This work was partially supported by the Japan Society for the Promotion of Science through a Grant-in-Aid for Scientific Research (No. 20550119). [1] H. Yersin (Ed.), Highly Efficient OLEDs with Phosphorescent Materials, Wiley-VCH, Weinheim, Germany, 2008. [2] a) S. Lamansky, P. Djurovich, D. Murphy, F. Abdel-Razzaq, H.-E. Lee, C. Adachi, P. E. Burrows, S. R. Forrest, M. E. Thompson, J. Am. Chem. Soc. 2001, 123, 43044312; b) K. D. Glusac, S. Jiang, K. S. Schanze, Chem. Commun. 2002, 2504 2505; c) R. Ragni, E. A. Plummer, K. Brunner, J. W. Hofstraat, F. Babudri, G. M. Farinola, F. Naso, L. De Cola, J. Mater. Chem. 2006, 16, 11611170; d) S. Takizawa, J. Nishida, T. Tsuzuki, S. Tokito, Y. Yamashita, Inorg. Chem. 2007, 46, 4308 4319; e) M. Lepeltier, T. K.-M. Lee, K. K.-W. Lo, L. Toupet, H. Le Bozec, V. Guerchais, Eur. J. Inorg. Chem. 2007, 2734 2747; f) C. Ulbricht, B. Beyer, C. Friebe, A. Winter, U. S. Schubert, Adv. Mater. 2009, 21, 44184441; g) D. Hanss, J. C. Freys, G. Bernardinelli, O. S. Wenger, Eur. J. Inorg. Chem. 2009, 4850 4859; h) Y. Chi, P.-T. Chou, Chem. Soc. Rev. 2010, 39, 638 655; i) K. Tsuchiya, S. Yagai, A. Kitamura, T. Karatsu, K. Endo, J. Mizukami, S. Akiyama, M. Yabe, Eur. J. Inorg. Chem. 2010, 926933; j) N. Tian, D. Lenkeit, S. Pelz, L. H. Fischer, D. Escudero, R. Schiewek, D. Klink, O. J. Schmitz, L. Gonzlez, M. Schferling, E. Holder, Eur. J. Inorg. Chem. 2010, 48754885. [3] a) M. S. Lowry, J. I. Goldsmith, J. D. Slinker, R. Rohl, R. A. Pascal Jr., G. G. Malliaras, S. Bernhard, Chem. Mater. 2005, 17, 57125719; b) A. B. Tamayo, S. Garon, T. Sajoto, P. I. Djurovich, I. M. Tsyba, R. Bau, M. E. Thompson, Inorg. Chem. 2005, 44, 87238732; c) Md. K. Nazeeruddin, R. T. Wegh, Z. Zhou, C. Klein, Q. Wang, F. De Angelis, S. Fantacci, M. Gratzel, Inorg. Chem. 2006, 45, 92459250; d) C. Rothe, C.-J. Chiang, V. Jankus, K. Abdullah, X. Zeng, R. Jitchati, A. S. Batsanov, M. R. Bryce, A. P. Monkman, Adv. Funct. Mater. 2009, 19, 20382044; e) R. D. Costa, E. Orti, H. J. Bolink, S. Graber, S. Schaffner, M. Neuburger, C. E. Housecroft, E. C. Constable, Adv. Funct. Mater. 2009, 19, 34563463. [4] a) S. Takizawa, R. Aboshi, S. Murata, Photochem. Photobiol. Sci. 2011, 10, 895903; b) J. Sun, J. Zhao, H. Guo, W. Wu, Chem. Commun. 2012, 48, 41694171. [5] Q. Zhao, C. Huang, F. Li, Chem. Soc. Rev. 2011, 40, 2508 2524.

Eur. J. Inorg. Chem. 0000, 00

0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

www.eurjic.org

Job/Unit: I20474

/KAP1

Date: 17-07-12 17:53:33

Pages: 6

SHORT COMMUNICATION
Cationic iridium(III) complexes with Coumarin 6 ligands were prepared and used as sensitizers for visible-light-driven hydrogen generation. Intense absorption in the visible region was observed, and high turnover numbers in hydrogen generation were achieved.

S.-y. Takizawa, C. Prez-Bolvar, P. Anzenbacher, Jr., S. Murata

Iridium Photosensitizers
S.-y. Takizawa,* C. Prez-Bolvar, P. Anzenbacher, Jr., S. Murata* ....... 16 Cationic Iridium Complexes Coordinated with Coumarin Dyes Sensitizers for Visible-Light-Driven Hydrogen Generation Keywords: Iridium / Photochemistry / Sensitizers / Hydrogen / Cobalt

www.eurjic.org

0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Eur. J. Inorg. Chem. 0000, 00

Das könnte Ihnen auch gefallen