TMPH_A_777811 TFJATS-Style4.cls March 5, 2013 20:57 PROOF COVER SHEET Author(s): Corneliu I. Oprea, Petre Panait, Boris F. Minaev, ˚ gren, Fanica Cimpoesu, Marilena Ferbinteanu and Hans A Mihai A. Gˆırt¸u Article title: Comparative computational IR, Raman and phosphorescence study of Ru- and Rh-based complexes Article no: 777811 Enclosures: 1) Query sheet 2) Article proofs Dear Author, 1. Please check these proofs carefully. It is the responsibility of the corresponding author to check these and approve or amend them. A second proof is not normally provided. Taylor & Francis cannot be held responsible for uncorrected errors, even if introduced during the production process. Once your corrections have been added to the article, it will be considered ready for publication. Please limit changes at this stage to the correction of errors. You should not make insignificant changes, improve prose style, add new material, or delete existing material at this stage. Making a large number of small, non-essential corrections can lead to errors being introduced. We therefore reserve the right not to make such corrections. For detailed guidance on how to check your proofs, please see http://journalauthors.tandf.co.uk/production/checkingproofs.asp. 2. Please review the table of contributors below and confirm that the first and last names are structured correctly and that the authors are listed in the correct order of contribution. This check is to ensure that your name will appear correctly online and when the article is indexed. Sequence Prefix Given name(s) Surname 1 Corneliu I. Oprea 2 Petre Panait 3 Boris F. Minaev 4 Hans ˚ gren A 5 Fanica Cimpoesu 6 Marilena Ferbinteanu 7 Gˆırt¸u Mihai A. Suffix TMPH_A_777811 TFJATS-Style4.cls March 5, 2013 20:57 Queries are marked in the margins of the proofs. AUTHOR QUERIES General query: You have warranted that you have secured the necessary written permission from the appropriate copyright owner for the reproduction of any text, illustration, or other material in your article. (Please see http://journalauthors.tandf.co.uk/preparation/permission.asp.) Please check that any required acknowledgements have been included to reflect this. Q1. AU: The word “four” has been been changed to “to” for grammar. Please check. Q2. AU: Please clarify what these characters should be. Q3. AU: As per journal style, “et al.” is not used in the references list. Please provide the relevant author names in order to complete the reference “57”. TMPH_A_777811 TFJATS-Style4.cls March 5, 2013 20:57 How to make corrections to your proofs using Adobe Acrobat Taylor & Francis now offer you a choice of options to help you make corrections to your proofs. Your PDF proof file has been enabled so that you can edit the proof directly using Adobe Acrobat. This is the simplest and best way for you to ensure that your corrections will be incorporated. If you wish to do this, please follow these instructions: 1. Save the file to your hard disk. 2. Check which version of Adobe Acrobat you have on your computer. You can do this by clicking on the “Help” tab, and then “About.” If Adobe Reader is not installed, you can get the latest version free from http://get.adobe.com/reader/. • If you have Adobe Reader 8 (or a later version), go to “Tools”/ “Comments & Markup”/ “Show Comments & Markup.” • If you have Acrobat Professional 7, go to “Tools”/ “Commenting”/ “Show Commenting Toolbar.” 3. Click “Text Edits.” You can then select any text and delete it, replace it, or insert new text as you need to. If you need to include new sections of text, it is also possible to add a comment to the proofs. To do this, use the Sticky Note tool in the task bar. Please also see our FAQs here: http://journalauthors.tandf.co.uk/production/index.asp. 4. Make sure that you save the file when you close the document before uploading it to CATS using the “Upload File” button on the online correction form. A full list of the comments and edits you have made can be viewed by clicking on the “Comments” tab in the bottom left-hand corner of the PDF. If you prefer, you can make your corrections using the CATS online correction form. TMPH_A_777811 TFJATS-Style4.cls March 5, 2013 20:57 Molecular Physics, 2013 Vol. 00, No. 00, 1–13, http://dx.doi.org/10.1080/00268976.2013.777811 Comparative computational IR, Raman and phosphorescence study of Ru- and Rh-based complexes ˚ grenb , Fanica Cimpoesud , Marilena Ferbinteanue and Corneliu I. Opreaa , Petre Panaita , Boris F. Minaevb,c,∗ , Hans A Mihai A. Gˆırt¸ua,∗ 5 a Department of Physics, Ovidius University of Constant¸a, Constant¸a, Romania; b Department of Theoretical Chemistry and Biology, Royal Institute of Technology, Stockholm, Sweden; c Department of Chemistry, B. Khmelnitsky National University, Cherkassy, Ukraine; d Department of Theoretical Chemistry, Institute of Physical Chemistry, Bucharest, Romania; e Department of Inorganic Chemistry, University of Bucharest, Bucharest, Romania (Received 16 December 2012; final version received 12 February 2013) 10 We report density functional theory (DFT) calculations providing the infrared and Raman spectra of [Ru(II)(bpy)3-n (dcbpy)n ]2 + and [Rh(III)(bpy)3-n (dcbpy)n ]3 + complexes, where bpy = 2,2’-bipyridyl, dcbpy = 4,4’dicarboxy-2,2’-bipyridyl, and n = 0, 1, 2, 3, studied in the context of dye-sensitised solar cells. We compare and contrast the role of the metallic ion and of the COOH groups on the vibration and phosphorescence properties of these complexes. The vibrational spectra are not very sensitive to the replacement of the metal ion, but the presence of carboxyl groups leads to a richer spectrum due to the additional bands caused by the COOH groups. Comparison with the limited experimental data available allowed the assignment of the Raman bands. The calculated phosphorescence lifetimes suffer only modest changes when the COOH groups are introduced but vary significantly when changing the metal ion, being two orders of magnitude larger for Rh(III) than for the Ru(II) complexes. 15 Keywords: ruthenium complexes; rhodium complexes; 2,2’-bipyridyl; 4,4’-dicarboxy-2,2’-bipyridyl; density functional theory; infrared spectra; Raman spectra; phosphorescence 20 1. Introduction 25 30 35 40 Based on its chemical stability, redox properties, excited state reactivity, luminescence emission, and excited state lifetime, [Ru(bpy)3 ]2 + (bpy = 2,2’-bipyridine) has provided the natural starting point for numerous studies in inorganic charge-transfer complexes, being one of the most extensively studied species [1]. Research on the larger class of transition metal-polypyridine complexes has stimulated the development of photocatalysis [2,3], photo-biochemistry [4,5], and photovoltaic devices [1,6]. The Gr¨atzel cells [7] are low-cost photovoltaic devices based on the light absorption in a dye sensitising a wide bandgap semiconductor, such as TiO2 , in the anatase form. The electron is transferred from the dye to the semiconductor and, through a transparent conducting oxide, passing the external load to the counter-electrode, where an electrolyte brings it back to the dye. The Ru(II)-polypyridine complexes have lead to photovoltaic conversion efficiencies above 11% [8,9]. Various Ru(II) complexes containing carboxyl groups, as in 4,4’-dicarboxy-2,2’-bipyridine, dcbpy [10] or 5,5’dicarboxy-2,2’-bipyridine [11], have been studied because of the need for anchoring groups to ensure the adsorption of the dye to the oxide. The presence of the carboxyl group ∗ Corresponding author. Email: [email protected] C 2013 Taylor & Francis requires special synthesis and leads to modified chemical properties, dcbpy-based complexes being more difficult to obtain [10]. Ru(II) polypyridine complexes have been extensively studied both experimentally [1,12] and theoretically. Density functional theory (DFT) calculations have provided geometry optimisation and excited-state properties of the [Ru(bpy)3 ]2 + species [13–15], and allowed the study of ligand substituent effects in [Ru(L)3 ]2 + complexes (L = bpy, bpm, bpz) [16,17]. Other DFT studies reported the effect of the ligand substitution in the less symmetric [Ru(bpy)2 L]2 + family of complexes [18–23], including vibration spectra analysis and vibronic effects on the charge transfer [24,25]. The theoretical studies have also been extended to other Ru(II) complexes with various different heterocycles [26– 28], as well as to complexes based on other transition metal ions, such as Os [29,30], Ir [25,31], Re [32], Pt [33], etc. Despite this increasing number of theoretical studies many questions are still waiting for an answer. For instance, open questions that still need to be addressed relate to (1) the role of the transition-metal, the Ru(II) complexes outperforming by far other DSSC dyes, (2) the role of the carboxyl groups, and whether it is limited to just insuring the anchoring of the dye to the substrate, (3) the influence 45 50 55 60 65 TMPH_A_777811 TFJATS-Style4.cls 2 March 5, 2013 20:57 C.I. Oprea et al. of the vibronic perturbations on the electron transfer and (4) the effect of the radiative and non-radiative deexcitation transitions of the dyes on the device performance. The vibration spectra have been studied for [Ru(bpy)3 ]2 + [34–39] and other Ru(II) complexes with various ligands [40]. Theoretical studies of IR and Raman 75 spectra have addressed in particular the normal modes of the [Ru(bpy)3 ]2 + [41–44], taking advantage of the symmetry of the complex. No such studies are available for complexes with various dicarboxyl groups or with Rh(III) ions. We build here on a previous study [45] in 80 which we presented results of DFT calculations providing the structure, electronic properties, and UVVis absorption spectra of [Ru(II)(bpy)3-n (dcbpy)n ]2 + and [Rh(III)(bpy)3-n (dcbpy)n ]3 + , where dcbpy = 4,4’dicarboxy-2,2’-bipyridyl, bpy = 2,2’-bipyridyl, and n = 85 0, 1, 2, 3 complexes. The comparison with the Rh(III) complexes revealed the better matching with the solar spectrum of the absorption properties of the Ru(II) complexes. Of the complexes studied, the most suited as pigments for dyesensitised solar cells are the [Ru(II)(bpy)3-n (dcbpy)n ]2 + 90 complex with n = 1 and 2, based on their intense absorption band in the visible region, the presence of the anchoring groups allowing the bonding to the TiO2 substrate as well as the charge transfer, and the good energy level alignment with the conduction band edge of the semiconducting sub95 strate and the redox level of the electrolyte [45]. Here we report DFT calculations of the infrared (IR) and Raman spectra as well as phosphorescence studies of the Ru(II) and Rh(III) complexes mentioned above. We compare and contrast the role of the metallic ion and of 100 the COOH groups on the vibrational and phosphorescent properties of these complexes. Phosphorescence transitions and lifetimes have been studied from the perspective of new materials for organic light emitting devices [46], but such studies are also relevant 105 for photovoltaic cells, particularly DSSCs. First principle theoretical analysis of phosphorescence of organometallic compounds has recently become a realistic task with the 70 use of the quadratic response (QR) technique in the framework of the time-dependent (TD) DFT approach [47–49]. We report here correlations between the main features of 110 the electronic structures and the phosphorescence lifetimes and discuss the relevance for hybrid organic–inorganic photovoltaic cells. 2. Computational details The geometrical structures of the [Ru(II)(bpy) 2+ and [Rh(III)(bpy)3-n (dcbpy)n ]3 + com3-n (dcbpy)n ] plexes (n = 0, 1, 2, and 3) were reported previously [45], based on DFT calculations. Correspondingly, the IR and the Raman spectra were simulated using the same hybrid B3LYP exchange-correlation functional [50,51]. The Los Alamos effective core potential (ECP) [52] and double-ζ quality functions for valence electrons [53] were used by employing the LANL2DZ basis set. Optical absorption spectra of all complexes, including the lowest 50 singlet–singlet excitations and the lowest 6 singlet–triplet excitations, were simulated using the time dependent-DFT (TD-DFT) method [54], with the same B3LYP functional and ECP on the metallic ion, but including polarisation functions for the valence region of every atom via the DZVP basis sets [55]. Calculations were performed both in vacuum and in water solvent by employing the non-equilibrium conductor-like polarisable continuum model (C-PCM) [56]. The geometry optimisation, IR and Raman spectra, electronic structure and transitions in vacuum as well as in water solvent were calculated with the GAUSSIAN03 package [57], whereas the electronic transitions in vacuum and the phosphorescence lifetime calculations were performed with the Dalton program [58]. 3. Results and discussion 3.1. IR and Raman spectra The optimised geometries for the four Rh-based complexes, calculated previously [45] are shown in Figure 1. Details regarding the structures have already been presented [45]. B/w in print, colour online Figure 1. Optimised structures of the [Rh(III)(bpy)3-n (dcbpy)n ]3 + complexes, with n equal to a) 0, b) 1, c) 2, and d) 3. 115 120 125 130 135 140 TMPH_A_777811 TFJATS-Style4.cls March 5, 2013 20:57 Molecular Physics 145 150 155 160 165 170 175 Here we stress that the structures of the corresponding Ru-based complexes are similar, the role played by the transition metal ion being most obvious in the distance between the metal ion and the nitrogen atoms of the adjacent dcbpy or bpy, which is larger for M = Ru by about 0.02 ˚ . However, the metal ion does not affect substantially the A bpy or dcbpy ligands, as both the carbon–carbon and the nitrogen–carbon bond lengths are practically identical. For each metal ion, the difference between r(M–Ndcbpy ) and ˚ , which suggests r(M–Nbpy ) is very small, less than 0.005 A that the presence of the carboxyl groups has a negligible influence on those distances. The symmetry group of the n = 1 and n = 2 complexes is C2 , whereas for the n = 0 and n = 3 complexes it is D3 . The bipyridyl ligands have deviations from planarity of up to 2.15o and are positioned almost reciprocally perpendicular, the angles taking values in the range 87.73–96.78o . The simulated IR and Raman spectra of [Ru(dcbpy)3 ]2 + and [Rh(dcbpy)3 ]3 + , calculated at DFT/B3LYP/LANL2DZ level, are displayed in Figure 2. The value of the scaling factor used for the vibration frequencies was 0.9614 [59]. As shown in Figure 2, the IR and Raman spectra of the complexes with three dcbpy groups have in common both the high-frequency stretches of the O–H (3511 cm−1 ) and C–H (3130 cm−1 ) groups and the fingerprint region between 1000 and 1600 cm−1 . Below 1000 cm−1 the Raman bands are weak, for both metals, whereas the IR spectra still show intense features. From Figure 2, the effect of changing the transition metal ion appears stronger in the Raman spectra, with a significant optical activity drop in the fingerprint region. In the IR spectra, the extra positive charge in the Rh-based complex polarises more the system, enhancing the high-frequency C–H stretch (3511 cm−1 ) but weakening the metal–nitrogen stretch (1006 cm−1 ). On the 3 B/w in print, colour online Figure 2. Simulated IR (top) and Raman (bottom) spectra of both [Ru(dcbpy)3 ]2 + and [Rh(dcbpy)3 ]3 + complexes, calculated at the DFT/B3LYP/LANL2DZ level. The spectral lines were convoluted with Lorentzian distributions of 20 cm−1 linewidth. other hand, as expected, the larger mass of the Rh3 + ion leads to a shift to lower frequency of the vibration peaks, 180 especially in the high-frequency regime. The simulated IR and Raman spectra of [Ru(bpy)3-n (dcbpy)n ]2 + and [Rh(bpy)3-n (dcbpy)n ]3 + , n = 0, 1, 2, 3, calculated at DFT/B3LYP/LANL2DZ level, are displayed in Figures 3 and 4. The calculated frequencies 185 ν, the IR intensities, and the Raman optical activities of the main transitions in the spectrum are shown in Table 1 (see supplementary material) for the ground state of the [Ru(dcbpy)3 ]2 + complex, together with the assignment of the type of vibration. The atom labelling scheme is shown 190 in Figure 5. B/w in print, colour online Figure 3. Simulated IR spectra of a) [Ru(II)(bpy)3-n (dcbpy)n ]2 + , b) [Rh(III)(bpy)3-n (dcbpy)n ]3 + for n = 0, 1, 2 and 3, calculated at the DFT/B3LYP/LANL2DZ level. The spectral lines were convoluted with Lorentzian distributions of 20 cm−1 linewidth. TMPH_A_777811 TFJATS-Style4.cls March 5, 2013 20:57 4 C.I. Oprea et al. Table 1. Frequencies, IR intensities and Raman optical activities of the main transitions in the spectrum for the ground state of the [Ru(dcbpy)3 ]2 + complex, calculated at DFT/B3LYP/LANL2DZ level. Each mode is described by the corresponding types of vibration. The atom labelling scheme is shown in Figure 5. Mode ν (cm−1 ) 49 51 52 54 72 78 82 94 95 109 116 124 125 128 135 136 143 154 162 167 168 173 177 180 181 188 189 191 192 198 205 207 213 226 231 363 373 418 427 544 603 617 720 720 863 883 984 986 992 1006 1035 1062 1143 1254 1274 1275 1322 1326 1387 1411 1454 1455 1524 1525 1593 1639 1640 3130 3511 3512 IR–I (km/mol) R–AO (A4 /AMU) Type of vibration 20.6 51.8 36.3 0.0 74.5 81.2 277.7 0.8 11.2 0.7 38.8 18.8 4.1 19.9 0.0 848.9 212.7 186.2 1.4 73.2 0.2 235.9 3.0 253.4 0.1 23.0 1.8 33.2 0.1 1.5 582.7 5.1 5.5 385.9 2.6 3.8 0.0 38.8 21.0 0.1 10.7 5.1 47.0 0.1 45.6 0.1 0.2 104.2 432.2 679.8 0.6 16.2 14.9 428.4 43.6 115.7 25.3 281.4 0.1 314.2 184.9 425.9 75.4 236.9 1004.0 309.0 1060.1 99.4 14.6 785.3 ρ(chelate ring) + τ (OH) ρ(chelate ring) + τ (OH) ν(Ru–N) + δ(OCO) + ν(C4–C7); chelate ring breathing ν(Ru–N) + δ(OCO) + ν(C4–C7); chelate ring breathing δ(RuNC6) + δ(COH) + δ(CCO) δ(OCO) + δ(C2NC6) + δ(C3C4C5) ω(OH) δ(NRuN) + δ(C2C3C4) + δ(C5C6N) + ν(C2–C2’) + δ(COH) δ(NruN) + δ(C2C3C4) + δ(C5C6N) + ν(C2–C2’) + δ(COH) δ(C2NC6) + δ(C3C4C5) + ν(C2–C2’) + ν(C4–C7) + ν(C–O) ω(CH) ν(Ru–N) + δ(C2NC6) + δ(C2C3C4) + δ(C4C5C6) + τ (CH) ν(Ru–N) + δ(C2NC6) + δ(C2C3C4) + δ(C4C5C6) + τ (CH) ν(Ru–N) + δ(C2NC6) + δ(C2C3C4) + δ(C4C5C6) + τ (CH) ν(Ru–N) + δ(C2NC6) + δ(C2C3C4) + δ(C4C5C6) δ(C3C2N) + δ(COH) + ν(C–O) + δ(CCH) δ(CCH) + δ(COH) + ν(C–O) + δ(C2’C2C3) δ(COH) + δ(C2C3C4) + ν(C4–C7) + δ(CCH) δ(C3C2C2’) + δ(CCH) + ν(C2–N) δ(CCH) + δ(COH) + ν(C2–C2’) + δ(OCO) δ(CCH) + δ(COH) + ν(C2–C2’) + δ(OCO) δ(COH) + ν(C4–C7) + δ(CCH) + δ(OCO) δ(COH) + ν(C4–C7) + δ(CCH) + δ(OCO) + ν(C2–C2’) δ(CCH) + δ(C2’C2N) + ν(C2–C3) + ν(C5–C6) δ(C2’C2N) + δ(C4C3C) + ν(C4–C3) + ν(C2–C2’) + δ(CCH) ν(C2–C2’) + ν(C5–C6) + δ(CCH) + ν(C–N) + δ(C3C2N) ν(C2–C2’) + ν(C5–C6) + δ(CCH) + ν(C–N) + δ(C3C2N) ν(C–N) + ν(C4–C5) + δ(C3C4C) + δ(CCH) + δ(NRuN) ν(C–N) + ν(C4–C5) + δ(C3C4C) + δ(CCH) + δ(NRuN) ν(C2–C3) + ν(C5–C6) + δ(CCH) + δ(C3C4C5) + δ(C2’C2N) ν(C–O) + δ(CCO) + δ(COH) ν(C–O) + δ(CCO) + δ(COH) ν(C–H) ν(O–H) ν(O–H) B/w in print, colour online Figure 4. Simulated Raman spectra of a) [Ru(II)(bpy)3-n (dcbpy)n ]2 + , b) [Rh(III)(bpy)3-n (dcbpy)n ]3 + for n = 0, 1, 2 and 3, calculated at DFT/B3LYP/LANL2DZ level. The spectral lines were convoluted with Lorentzian distributions of 20 cm−1 linewidth. TMPH_A_777811 TFJATS-Style4.cls March 5, 2013 20:57 Molecular Physics B/w in print, colour online Figure 5. Scheme with the atom labelling used to describe the vibrations of the ligands. 195 200 205 210 215 220 225 230 The vibration spectra of both families of complexes under study are composed of 61, 67, 73, or 79 atoms, and have 177, 195, 213, and 231 normal modes of vibration for n = 0, 1, 2, and 3, respectively. Given the symmetry of the complexes, the decompositions in normal modes are: 177 = 30a1 + 29a2 + 59e, 195 = 98a + 97b, 213 = 107a + 106b, and 231 = 39a1 + 38a2 + 77e. For the less symmetric complexes, with n = 1 and 2, all modes are both IR- and Raman-active. In the case of the more symmetric n = 0 or 3 complexes, IR-active are the modes transforming as a2 and e, whereas those transforming as a1 and e are Raman-active [42]. A careful look at the values reported in Table 1, however, shows very small but still non-zero IR intensities for Raman-active modes and the reverse. This is due to the fact that the software maintained the largest Abelian group, C2 , instead of the full point group of the complex, C3 . To better understand the vibration modes and their symmetry we included in the supplementary materials animations corresponding to the most intense peaks in both the IR and Raman spectra. Figures 3 and 4 allow the analysis of the role played by the COOH groups. Compared to the prototypical [Ru(bpy)3 ]2 + the complexes with n = 1, 2, and 3 display a much richer spectrum due to the additional bands caused by the COOH groups. The obvious thing to note is the increase with the number of dcbpy ligands, n, of the vibration intensity of the modes involving COOH groups. The linear dependence with respect to n observed for the intensity of the IR peaks at 3511, 3130, 1639, 1525, 1387, 1322, 1274, 1143, 1062, 1035, 863, 720, 617, and 544 cm−1 , as well as the Raman peaks at 3512, 3130, 1640, 1593, 1525, 1455, 1411, 1326, 1275, 1254, 1143, 1062, 1006, 992, and 863 cm−1 , can provide a textbook example for the study of the vibration spectra. Also interesting is the inverse correlation, i.e. the decrease of the peak intensity with the number of COOH groups at 1437 and 775 cm−1 in the IR and at 1305 and 748 cm−1 in the Raman spectra. These bands are diminished when replacing the H atom bound to C4 with COOH. The bulky carboxyl groups hinder the out-of-plane wagging or twisting of the pyridine rings, shifting the mode to 5 lower frequencies. An increasing number of COOH groups will shift more and more of the energy available for those modes, leading to a decrease of the peak activity from n = 0 to 3. We now proceed for a more systematic analysis of the most important features in the spectra of [Ru(bpy)3-n (dcbpy)n ]2 + , noting the similarity between the spectra of the Ru(II) and Rh(III) complexes, as shown in Figure 2. We focus on the Ru(II) complexes both for their practical importance and for the availability of some experimental and calculated spectra which allow a more straightforward comparison with our results. We start with the O–H stretch observed at 3511 cm−1 in both the IR and Raman spectra. The bands can be attributed to the O–H in-phase and out-of-phase stretches, as it is well established [60]. The bands are missing from the spectrum of the parent, n = 0, compound and have an intensity that increases proportional to n for the other complexes. The O–H groups are also present in the scissor modes observed in the IR fingerprint region, at 1143, 1062, and 1035 cm−1 , as well as the wagging modes at 617 cm−1 . Furthermore, the presence of the carboxyl groups has a clear mark in the strong C = O stretching mode at 1639 cm−1 in the IR and 1640 cm−1 in the Raman spectra, which occurs together with the O–C–O and C–O–H bending modes. The scissor modes δ(COH) and δ(OCO) can also be found in the fingerprint region, at 1322 cm−1 in IR and 1326 cm−1 in Raman spectra. At lower frequencies, at 601 cm−1 strong O–C–O scissoring modes are present only in the IR spectra of the n ≥ 1 complexes. The low-intensity C–H stretches are present in all spectra, regardless of n, at 3130 cm−1 in the IR and Raman spectra. Bending C–H modes are observed particularly in the IR spectra at 1143 and 1062 cm−1 as well as in the Raman spectra, at 1593, 1455, and 1254 cm−1 . However, the bending modes present in the fingerprint region are not pure, but mixed with other contributions. For instance, the vibrations of the entire pyridine ring, as well as the deformations of the chelate ring also contribute to the C–H bending modes. The aromatic pyridine rings can withstand vibrations of the C = C and C = N double bonds. The C = C bonds leave a clear mark, regardless of n, through the stretching modes located in the Raman spectra at 1593 (the most intense band) and 1455 cm−1 and in the IR spectra at 1387 cm−1 . The C = N bonds display a strong Raman stretching mode at 1525 cm−1 . At 1524 cm−1 the activity is weaker and the modes are active also in the IR. The pyridine rings suffer vibrations as an entity at many frequencies. Various types of C–C–C scissor modes are active at frequencies such as 1254, 1143, and 1062 cm−1 . The former mode is stronger in the Raman whereas the other two are more intense in the IR spectra. The C–N–C scissor modes of the pyridine rings occur at lower frequencies, such as 1006, 992, 986, and 984 cm−1 together with stretches of 235 240 245 250 255 260 265 270 275 280 285 TMPH_A_777811 TFJATS-Style4.cls March 5, 2013 20:57 6 290 295 300 305 310 315 C.I. Oprea et al. the metal–nitrogen bonds. The first three are particularly strong in the Raman spectra. The C–C σ -bonds connecting the pyridine rings, experience stretching vibrations at various frequencies, the most important being 1595, 1455, 1411, and 1325 cm−1 , the first one being the most important, weaker in the IR but stronger in the Raman spectra. The C–C interring vibrations are mixed in the spectrum with δ(CCC) and δ(CCH) modes of the pyridine rings, which occur together with various vibration modes of the 5-membered chelate ring. The metal–nitrogen vibrations are important for the induction of vibronic perturbations in the electron shells of the excited complexes. The Ru–N stretching modes occur first in the Raman spectra, with a strong signal at 1006 cm−1 and a weaker optical activity at 992 and 986 cm−1 . As discussed above, the metal–nitrogen stretching comes together with a C–N–C scissor bending as well as chelate ring deformations. At 984 cm−1 the corresponding IR signal is stronger. At lower frequencies, the deformation vibrations of the metallic environment occur together with a bending of the ligands. At 427 and 418 cm−1 we find chelate ring breathing bands, the former present only in the Raman, whereas the later active in both IR and Raman. At lower frequencies, 373 and 363 cm−1 , the chelate rocking modes are active mostly in the IR. In the remaining of this section we will compare our results to the experimental data available. For a proper comparison, we represent in Figure 6 the Raman intensity, calculated keeping in mind not just the part intrinsic to the scattering molecule (Gaussian’s Raman optical activity), but also the factor that depends on temperature (which affects the population of the scattering vibrational state), B/w in print, colour online Figure 6. Simulated Raman intensity of [Ru(II)(bpy)3-n (dcbpy)n ]2 + , for n = 1 and 3, calculated [61] at DFT/B3LYP/LANL2DZ level, at room temperature and a laser excitation wavelength of 532 nm [39]. The spectral lines were convoluted with Lorentzian distributions of 20 cm−1 linewidth. and the factor that is dependent on the exciting laser frequency [61]. The relative height of the peaks is different in Figure 6 with respect to Figure 4a mainly because of the factor depending on the excitation frequency, which tends to increase the low-frequency bands. As mentioned in the introduction, except for the case of [Ru(bpy)3 ]2 + [34–44] and a paper on [Ru(dcbpy)3 ]2 + [39] there are no published experimental spectra or theoretical results for the IR and Raman spectra of the complexes calculated in this study. Therefore, for comparison, we will also use some other data available for similar bipyridyl complexes [40,62–64]. The key bands in the experimental Raman spectra available of [Ru(dcbpy)3 ]2 + [39] are located at 1618, 1546, 1480, 1433, 1373, 1298, and 1276 cm−1 . Similarly, for [Ru(bpy)2 (dcbpy)]2 + [39], the main peaks observed experimentally are situated at 1542, 1375, and 1272 cm−1 . Our calculations make possible the assignment of these bands as shown below. The 1618 cm−1 experimental peak likely corresponds to the very intense 1593 cm−1 mode attributed to the ν(C2– C3) and ν(C5–C6) stretches with some contributions from δ(CCH) and δ(C3C4C5) scissoring vibrations. Correspondingly, in the Raman spectrum of the [Ru(bpy)3]2 + complex [41,62], the bands of the ν(C2–C3) stretching vibrations of pyridine are observed at lower frequencies, of 1608 cm−1 , whereas using the same scaling factor of 0.9614 [59] for the entire frequency range, we obtain 1578 cm−1 . The presence of the COOH groups in the [Ru(dcbpy)3]2 + complex shifts that band to higher frequencies both in the experimental and the calculated spectra. The bands at 1546 cm−1 for [Ru(dcbpy)3]2 + , at 1542 cm−1 for [Ru(bpy)2 (dcbpy)]2 + [39], and at 1563 cm−1 for [Ru(bpy)3]2 + [41], may correspond to the vibrations of the aromatic C = C and C = N bonds of the pyridine rings, ν(C–N) and ν(C4–C5). Our results follow a similar trend: 1525 cm−1 , 1522 cm−1 , and 1533 cm−1 , respectively. Further, the peaks at 1480 cm−1 for the complex with n = 3 [39] and at 1491 cm−1 for the one with n = 0 [41] can be correlated with the vibrations of the interring C–C σ -bond. Our calculations reflect the same trend, providing frequencies of 1455 and 1467 cm−1 , respectively. The next main experimental band for [Ru(dcbpy)3]2 + , located at 1433 cm−1 [39], may be attributed to the pyridine ring deformations, particularly the δ(C2’C2N), δ(C4C3C), and δ(CCH) scissor modes together with the ν(C4–C3), ν(C4–C7), and ν(C2–C2’) stretches, with a calculated maximum at 1411 cm−1 . We note that the corresponding band of [Ru(bpy)3]2 + , is very weak in the Raman spectrum. Further down in the frequencies, the bands of the complex with the n = 3, situated at 1373 cm−1 , or for the one with n = 1 at 1375 cm−1 [39], again have no clear correspondent in the experimental spectrum of n = 0 [41]. This band is more difficult to assign, as the modes that are 320 325 330 335 340 345 350 355 360 365 370 TMPH_A_777811 TFJATS-Style4.cls March 5, 2013 20:57 Molecular Physics 375 380 385 390 395 400 405 410 415 420 relatively close in frequency are strongly active in the IR and only very weak in the Raman spectra. A mode with a high Raman intensity is found at 1326 and 1325 cm−1 for the two complexes, respectively, and it consists of vibrations involving the COOH groups, such as the δ(COH) and δ(OCO) scissoring modes and the ν(C4–C7) stretches. We also note a reversed situation, where an intense Raman band of [Ru(bpy)3]2 + [41], located at 1320 cm−1 , has no clear assignment in the [Ru(dcbpy)3]2 + spectra [39]. The intense band due to both the interring C–C σ -bond stretch and the in-plane bending of the CH group which is peaked at 1305 cm−1 in our calculated spectra. Similar interring vibrations with a pronounced contribution from δ(CCH) vibrations, have been observed at 1317 cm−1 in the spectrum of the [Ru(bpy)2(BIK)]2 + complex [62]. The band at 1298 cm−1 in the experimental spectrum of the n = 3 complex [39] can be assigned to the 1275 cm−1 theoretical mode, consisting of various scissor modes, δ(CCH), δ(COH), and δ(OCO), together with an interring stretch, ν(C2–C2’). The last main peaks in the experimental Raman spectra of [Ru(dcbpy)3]2 + and [Ru(bpy)2 (dcbpy)]2 + [39] are found at 1276 and 1272 cm−1 , respectively. The calculated spectra show intense maxima at 1254 and 1248 cm−1 , respectively, attributed mainly to the δ(C3C2C2’) and δ(CCH) bending modes, together with a C–N stretch. Unfortunately, the range of the experimental spectra available does not extend below 1200 cm−1 , missing some important features, particularly at 1006 cm−1 where we calculated that the Ru–N stretching vibrations occur. Other such calculated stretching modes can be found at 992, 986, and 418 cm−1 . These vibrations are the most important for the induction of vibronic perturbations in the electron shells of the excited complexes. The Ru–N stretching vibrations in the experimental spectra of [Ru(bpy)3]2 + are located at 371 cm−1 and correspond to the 353 cm−1 in our results, occurring together with a chelate ring breathing [65]. We note that although our results describe accurately the experimental data available, the exact position of the peaks is systematically shifted towards lower frequencies by a factor of roughly 0.984 compared to the experimental data available [39]. The use of a standard scaling factor of 0.9614 [59] over the entire range of frequencies leads to an underestimation of the peak values. Based on our results, a scaling factor of 0.9771 would have described more accurately the actual positions of the vibration bands. For comparisons of various scaling factors for different IR regions see references [66–69]. 3.2. Phosphorescence study Spectral properties of the dye complexes are studied in 425 the framework of time-dependent density functional theory (TD DFT) [47]. In this scheme the excitation energies ∇ω can be determined from the poles of the ground state 7 linear polarisation propagator, det E [2] − ωS [2] = 0, (1) where E[2] and S[2] are the electronic Hessian and overlap matrices, respectively. The transition moments can 430 be calculated from the residues of Equation (1) for the singlet–triplet transitions. In the case of singlet–triplet transitions the zeroth-order contribution vanishes due to spinorthogonality. The first contribution to the transition moment then comes from the first-order perturbation theory 435 when the spin-orbit (SO) coupling operator is treated as a perturbation. The sublevels of the triplet state are considered to be energy degenerate in the first order, so the singlet–triplet transition moment reduces to: ∞ S0 |μˆ α |Sn Sn |Hˆ SO |T1k S0 |μˆ α |T1k = Mαk = E(Sn ) − E(T1 ) n=0 + ∞ S0 |Hˆ SO |Tn Tn |μˆ α |T k 1 n=1 E(Tn ) − E(S0 ) , (2) where the summation over intermediate triplet states Tn 440 includes all three sublevels of each triplet state. The radiative rate constant Ak and the phosphorescence lifetime from one of the three sublevels (indexed by k = x, y, z) of the lowest triplet state T1k is given by Ak = 3 k 2 1 4 3 M , = αo E k α τk 3t0 α∈{x,y,z} (3) where t0 = (4π ε0 )2 3 /me e4 , α 0 is the fine-structure con- 445 stant, Ek is the transition energy, and Mα k is the α-axis projection of the electric dipole transition moment between the ground state and the k-spin level of the triplet state. The radiative lifetime of the triplet state in the high-temperature 450 limit is estimated by [70] 1 1 1 = . τ 3 α∈{x,y,z} τk (4) It corresponds to averaging over all three sublevels when they are in thermal equilibrium. We note that Equation (4) is not strictly valid for the triplet states with high zero-field splitting (ZFS) [71] but is applicable [72] tothe complexes Q1 discussed here, for which the six d electrons are paired. In 455 the present work the SOC operator in Equation (2) is used in a semi-empirical effective single-electron approximation [73], in which the number of electrons is reduced and the two-electron part of SO coupling is removed, which greatly 460 simplifies the calculations. The electronic transitions in vacuum and the phosphorescence lifetime calculations performed in the gas phase TMPH_A_777811 TFJATS-Style4.cls March 5, 2013 20:57 8 C.I. Oprea et al. Table 2. Radiative phosphorescence lifetimes, τ (μs), of [Ru(II)(bpy)3-n (dcbpy)n ]2 + and [Rh(III)(bpy)3-n (dcbpy)n ]3 + complexes (n = 0, 1, 2, and 3), calculated with quadratic response DFT. ES-T (eV) is the S0 –T1 transition energy. Calculations are performed at the S0 optimised geometry in gas phase. The experimental values are from reference [1] and reference [76] for the Ru(II) and Rh(III) complexes, respectively. RT = room temperature. M N ES-T calc. ES-T exp. 77 K Ru(II) Rh(III) 465 470 475 480 485 490 τ exp. RT 77 K Principal configuration RT 0 2.38 2.04 13.9 0.60 1 2.19 2.02 10.5 0.46 2 2.23 14.0 3 2.28 11.5 0 2.91 1 2.84 4.99 103 2 2.83 4.53 103 3 2.83 5.10 103 2.77 2.70 5.93 103 with Dalton [58], with linear and quadratic response functions, respectively [74,75], are reported in Table 2 for all six complexes. The values of the triplet–singlet transition energies are compared, where data are available for experimental values obtained at 77 K and/or at room temperature. The calculated S0 –T1 transition energies are systematically larger than the experimental ones especially those measured at room temperature, by at most 14%. The values determined at liquid nitrogen temperatures are much closer to the calculated ones, within 5%. Such differences occur when computing vertical transitions. Calculations of the transition energies were also performed using GAUSSIAN03 [57] with the same DFT functional and basis sets and we found that the energy differences are smaller than 0.15%. We note that in the case of the Rh(III) ion the transition energies are systematically larger than those for the Ru(II) complexes. The composition of the S–T transitions reveals a ligand-to-metal charge-transfer nature in the case of Ru(II) and a ligand-to-ligand charge-transfer character for the Rh(III) species, which is consistent with our previous work regarding UV-Vis absorption based on singlet–singlet transitions [45]. The calculated phosphorescence lifetimes are very different from Rh(III) to Ru(II) complexes, variations of two orders of magnitude being observed. A probable explanation is related to the nature of the states involved in the transitions from the perspective of the spin–orbit interaction. In the case of the Rh(III) complexes the metal ion has only a small contribution to only one of the two states (less 2.0 103 0.015 Expansion of the special part of the triplet state wave function LUMO + 1 (dt2g π ∗ bpy ) → HOMO (dz2 ) LUMO (dt2g π ∗ dcbpy ) → HOMO (dz2 ) LUMO + 1 (dt2g π ∗ dcbpy ) → HOMO (dz2 ) LUMO + 1 (dt2g π ∗ dcbpy ) → HOMO (dz2 ) LUMO (dt2g π ∗ bpy ) → HOMO-2 (π bpy ) LUMO (dt2g π ∗ dcbpy ) → HOMO-2 (π bpy ) LUMO + 1 (ddz2 π ∗ dcbpy ) → HOMO-4 (π bpy ) LUMO (dt2g π ∗ dcbpy ) → HOMO-7 (π dcbpy ) Coeff. 0.668 0.649 0.679 0.672 0.422 0.592 0.415 0.308 than 6% of the electron density), which leads to a weak SO coupling, a small transition rate and a long lifetime (see Table 3). In contrast, in the case of the Ru(II) complexes the states participating in the transition have a more significant electron density on the metal ion (∼60%), leading to stronger SO interactions and shorter phosphorescence lifetimes. The presence of the carboxyl groups within each family of complexes leads to differences in the lifetimes within 25%, with no apparent systematic trend when considering the COOH content. The agreement between the calculated and experimental phosphorescence lifetimes is worsening with increasing temperature. The data available for the [Rh(III)(bpy)3 ]3 + complex show that our calculations are much closer to the 77 K value than that at the room temperature one [76]. This is because the phosphorescence lifetime measured at the room temperature includes a large contribution from the non-radiative quenching [67]. We also note that the calculated transition energy is in better agreement with the value measured at the lower temperature. As no 77 K data is known to us for the case of Ru(II) complexes, in order to check the reliability of our results we note that both at 77 and ∼300 K experimental phosphorescence lifetimes values are available for similar systems. For instance, for [Ru(II)(dmbpy)3 ]2 + the lifetimes are 4.6 and 0.95 μs, whereas for [Ru(II)(bpy)2 (dmbpy)]2 + , 5.2 and 0.44 μs, at 77 K and RT, respectively [77,78]. Moreover, more recently, another complex, [Ru(II)(bpy)2 (L1)]2 + , was reported with 5.23 and 0.22 μs, correspondingly [79]. The 495 500 505 510 515 520 TMPH_A_777811 TFJATS-Style4.cls March 5, 2013 20:57 Molecular Physics 9 Table 3. Electron density (in %) on the various building blocks of [Ru(II)(bpy)3-n (dcbpy)n ]2 + and [Rh(III)(bpy)3-n (dcbpy)n ]3 + complexes (n = 0, 1, 2, and 3), calculated by DFT/B3LYP/LANL2-DZVP for the key molecular orbitals. [Ru(II)(bpy)3-n (dcbpy)n ]2 + n=0 M bpy –COOH n=1 n=2 n=3 HOMO LUMO + 1 HOMO LUMO HOMO LUMO + 1 HOMO LUMO + 1 62.41 37.59 – 8.19 91.81 – 61.44 38.26 0.30 5.75 86.98 7.28 60.44 38.96 0.60 9.98 83.35 6.67 59.57 39.53 0.90 9.64 83.77 6.59 [Rh(III)(bpy)3-n (dcbpy)n ]3 + n=0 M bpy –COOH n=1 n=2 n=3 HOMO-2 LUMO HOMO-2 LUMO HOMO-2 LUMO HOMO-2 LUMO 3.18 96.82 – 5.25 94.75 – 0.43 99.51 0.06 4.49 95.45 0.06 0.37 89.43 10.20 3.32 90.73 5.95 0.35 89.29 10.36 5.49 88.92 5.59 room temperature values of these systems are comparable to the ones reported for [Ru(II)(bpy)3 ]2 + [1,76] which leads 525 us to speculate that the 77 K values might also be similar. If that was true the measured values are still lower than the ones we calculated for [Ru(II)(bpy)3-n (dcbpy)n ]2 + but in reasonable agreement, especially considering the fact that our DFT approach does not take into account the influence 530 of the temperature. It is of interest that a comparative analysis of all complexes with respect to the requirements for the dyes in order to be used in dye-sensitised solar cells. From this perspective, the energy level alignment between the substrate, the dye, and the electrolyte is crucial for a good photovoltaic 535 conversion [80,81]. Figure 7 displays an energy diagram with the conduction and valence band edges of the TiO2 [80], the energy levels of the singlet ground states, the lowest triplet excited state, and the sixth triplet excited state of all the eight complexes under study, calculated for con- 540 sistency in water solvent, as well as the redox level of the I3− /I− electrolyte [81]. B/w in print, colour online Figure 7. Energy diagram showing the HOMO level (blue) which is taken as an origin for the singlet ground states energy levels, the lowest triplet excited state (red) and the sixth triplet excited state (orange) levels of the various [Ru(II)(bpy)3-n (dcbpy)n ]2 + and [Rh(III)(bpy)3-n (dcbpy)n ]3 + complexes for n = 0, 1, 2 and 3, calculated in solution by TD-DFT/B3LYP/LANL2-DZVP methods. The conduction (red) and valence band (blue) edges of the TiO2 [80], as well as the redox level of the I3− /I− electrolyte [81] are also shown. Next to the transition line is stated, in eV, the energy of the first (sixth) singlet–triplet transition for each complex. TMPH_A_777811 TFJATS-Style4.cls March 5, 2013 20:57 10 C.I. Oprea et al. B/w in print, colour online Figure 8. Isodensity surfaces (0.03 e/bohr3 ) of selected molecular orbitals of the [Ru(II)(bpy)2 (dcbpy)]2 + complex: a) singlet ground state, S0 , b) triplet excited state, T3 , c) triplet excited state, T5 , and of the [Rh(III)(bpy)2 (dcbpy)]3 + complex: d) singlet ground state, S0 , e) triplet excited state, T3 , f) triplet excited state, T5 . We find that, except for the n = 0 and 1 Ru(II) complexes, all dyes having the first six triplet excited states 545 lie below the CB edge of the semiconductor. Therefore, for [Ru(II)(bpy)3 ]2 + and [Ru(II)(bpy)2 (dcbpy)]2 + some of the triplet excited states may contribute to the electron injection into the TiO2 conduction band. In the other cases, the low-lying triplet states cannot participate in charge injec550 tion, the photoelectron being lost by radiative deexcitation to the ground state. Such phosphorescence processes are detrimental to the efficiency of photovoltaic devices [82]. We display in Figure 8 the singlet ground state and two triplet excited states of [Ru(II)(bpy)2 (dcbpy)]2 + , to illus555 trate the likelihood of charge transfer to the TiO2 nanoparticle. The singlet ground state is fairly localised on the metal ion, whereas the triplet excited states are delocalised over the bipyridyl ligand, with a modest contribution from the t2g -like orbitals. The electron transfer to the semiconducting oxide is more likely from the orbital with high electron 560 density on the dcbpy ligand, due to the likely binding to titanium through the anchoring COO- groups. For contrast, we also display in Figure 8 the electron density of the corresponding MOs of [Rh(III)(bpy)2 (dcbpy)]3 + . The electron transfer is highly 565 unlikely because of the poor alignment with the conduction band edge of the oxide (see Figure 7). The very weak electron density on the Rh(III) ion illustrates the data presented in Table 3, for explaining the long phosphorescence 570 lifetimes. To strengthen this argument, we provide in Table 3 the Mulliken charge for the closed shell ground state of all Table 4. Mulliken charge on the various building blocks of [Ru(II)(bpy)3-n (dcbpy)n ]2 + and [Rh(III)(bpy)3-n (dcbpy)n ]3 + complexes (n = 0, 1, 2, and 3), calculated by DFT/B3LYP/LANL2-DZVP for the singlet ground state. Ru Q2 M bpy dcbpy 2 COOH Rh n=0 n=1 n=2 n=3 n=0 n=1 n=2 n=3 0.892 0.370 – – 0.889 0.387 0.522 −0.184 0.890 0.400 0.533 −0.178 0.886 – 0.543 −0.172 0.780 0.740 – – 0.774 0.745 0.779 −0.044 0.771 0.750 0.783 −0.043 0.764 – 0.787 −0.042 TMPH_A_777811 TFJATS-Style4.cls March 5, 2013 20:57 Molecular Physics complexes. In the case of Ru(II) complexes, the charge is almost equally distributed between the metal ion (∼0.89) and 575 the three ligands (∼3∗ 0.37), whereas for the Rh(III) complexes the charge is located mostly on the ligands (∼3∗ 0.74 compared to ∼0.78 on the metal). As expected, the extra electron leaving the Rh(III) ion is distributed on the ligands. It is worthwhile to note in Table 4 the charge density on 580 the COOH groups. The highest negative charge density on carboxyl groups occurs in the case of the Ru(II) systems, particularly for the n = 1 complex. This result is relevant to DSSCs, showing that the charge transfer is optimised in the case of [Ru(II)(bpy)2 (dcbpy)]2 + . 585 590 595 600 605 610 615 620 625 4. Conclusions Based on DFT calculations we simulated the IR and Raman spectra of [Ru(bpy)3-n (dcbpy)n ]2 + and [Rh(bpy)3-n (dcbpy)n ]3 + complexes. We found that the spectra are not very sensitive to the replacement of the metal ion, the overall qualitative behaviour being similar. The effect of changing the transition metal ion appears stronger in the Raman spectra, with a significant optical activity drop in the fingerprint region. The replacement of Ru(II) with Rh(III) leads in the IR spectra to the enhancement of the high-frequency C–H stretch and a weakening of the metal–nitrogen stretch. Also, as expected, the larger mass of the Rh(III) ion leads to a shift to lower frequency of the vibration peaks, especially in the high-frequency regime. The presence of carboxyl groups leads to a richer spectrum due to the additional bands caused by the COOH groups. The vibration intensity of the modes involving COOH groups increases proportionally with the number of dcbpy ligands, n, providing a textbook example for the study of vibration spectra. Also, some bands are inhibited, as the replacement of the H atom bound to C4 with bulky carboxyl groups hinder the out-of-plane wagging or twisting of the pyridine rings. The comparison with the limited experimental data available allowed the assignment of the Raman bands for [Ru(dcbpy)3 ]2 + and [Ru(bpy)2 (dcbpy)]2 + . Moreover, our results are compatible with previous experimental and theoretical studies of [Ru(bpy)3 ]2 + as well as various other complexes with substituted ligands. Although our results describe accurately the experimental data available, the exact position of the peaks is systematically shifted towards lower frequencies compared to the experimental data available. The use of a standard scaling factor of 0.9614 over the entire range of frequencies leads to an underestimation of the peak values, whereas a scaling factor of 0.9771 would have described more accurately the actual positions of the vibration bands. The composition of the transitions between the triplet excited states and the singlet ground state reveals a ligandto-metal charge-transfer in the case of Ru(II) and a ligandto-ligand charge-transfer for Rh(III), consistent with our 11 previous work regarding absorption singlet-to-singlet transitions. The calculated phosphorescence lifetimes are two orders of magnitude larger for Rh(III) than for Ru(II) complexes, likely due to the nature of the states involved in the transitions. In the case of the Ru(II) complexes the 630 states participating in the transition have a more significant electron density on the metal ion, leading to a stronger SO interaction and shorter phosphorescence lifetimes. For [Ru(II)(bpy)3 ]2 + and [Ru(II)(bpy)2 (dcbpy)]2 + some of the triplet excited states may have some contribution to the elec- 635 tron injection into the TiO2 conduction band, whereas in the other cases the photoelectron is, likely, lost by radiative deexcitation to the ground state. Our present study strengthens our previous conclusion that, of the eight dyes studied, the Ru(II) complexes with n = 1 or 2, [Ru(II)(bpy)2 (dcbpy)]2 + 640 in particular, are most suited for being used in DSSCs. Acknowledgements The authors acknowledge partial financial support as follows: M.F., F.C., and M.A.G from ANCS, grant PN2-Capacitati-M3 contract 517/2011, C.I.O. from CNCS, grant PN2-RU-PD contract 172/2010 and B.F.M from the Ministry of Education and Science of Ukraine. The authors thank Professor Ionel Humelnicu, for useful discussions. 645 References [1] A. Juris, V. Balzani, F. Barigelletti, S. Campagna, P. Belser, and A. von Zelewski, Coord. Chem. Rev. 84, 85 (1988). [2] B. Gholamkhass, H. Mametsuka, K. Koike, T. Tanabe, M. Furue, and O. Ishitani, Inorg. Chem. 44, 2326 (2005). [3] F. Lachaud, A. Quaranta, Y. Pellegrin, P. Dorlet, M.-F. Charlot, S. Un, W. Leibl, and A. Aukauloo, Angew. Chem. Int. Ed. 44, 1536 (2005). [4] J.K. Barton, Science 233, 727 (1986); A.E. Friedman, J.C. Chambron, J.P. Sauvage, N.J. Turro, and J.K. Barton, J. Amer. Chem. Soc. 112, 4960 (1990). [5] Y. Xiong and L.N. Ji, Coord. Chem. Rev. 185–186, 711 (1999). [6] V. Balzani, G. Bergamini, and P. Ceroni, Coord. Chem. Rev. 252, 2456 (2008). [7] M. Gratzel, Nature 414, 338 (2001). [8] M.K. Nazeeruddin, F. De Angelis, S. Fantacci, A. Selloni, G. Viscardi, P. Liska, S. Ito, B. Takeru, and M. Gratzel, J. Am. Chem. Soc. 127, 16835 (2005). [9] F. Gao, Y. Wang, D. Shi, J. Zhang, M. Wang, X. Jing, R. Humphry-Baker, P. Wang, S.M. Zakeeruddin, and M. Gratzel, J. Am. Chem. Soc. 130, 10720 (2008). [10] O. Kohle, S. Ruile, and M. Gratzel, Inorg. Chem. 35, 4779 (1996). [11] P.A. Anderson, G.F. Strouse, J.A. Treadway, R.K. Keene, and T.J. Meyer, Inorg. Chem. 33, 3863 (1994). [12] V. Balzani, G. Bergamini, F. Marchioni, and P. Ceroni, Coor. Chem. Rev. 250, 1254 (2006). [13] M. Buchs and C. Daul, Chimia 52, 163 (1998). [14] F. Alary, J.-L. Heully, L. Bijeire, and P. Vicendo, Inorg. Chem. 46, 3154 (2007). [15] T. Guillon, M. Boggio-Pasqua, F. Alary, J.-L. Heully, E. Lebon, P. Sutra, and A. Igau, Inorg. Chem. 49, 8862 (2010). 650 655 660 665 670 675 680 TMPH_A_777811 12 685 690 695 700 705 710 715 720 725 730 735 740 745 TFJATS-Style4.cls March 5, 2013 20:57 C.I. Oprea et al. [16] K.C. Zheng, J.P. Wang, W.L. Peng, X.W. Liu, and F.C. Yun, J. Molec. Struct (Theochem) 582, 1 (2002). [17] S.R. Stoyanov, J.M. Villegas, and D.P. Rillema, Inorg. Chem. 41, 2941 (2002). [18] L.-C. Xu, J. Li, S. Shi, K.-C. Zheng, and L.-N. Ji, J. Molec. Struct (Theochem) 855, 77 (2008). [19] J.-F. Yin, D. Bhattacharya, P. Thanasekaran, C.-P. Hsu, T.-W. Tseng, and K.-L. Lu, Inorg. Chim. Acta 362, 5064 (2009). [20] Y.-J. Chen, J.F. Endicott, and V. Swayambunathan, Chem. Phys. 326, 79 (2006). [21] W. Henry, W.R. Browne, K.L. Ronayne, N.M. O’Boyle, J.G. Vos, and J.J. McGarvey, J. Molec. Struct. 735–736, 123 (2005). [22] M.-F. Charlot, Y. Pellegrin, A. Quaranta, W. Leibl, and A. Aukauloo, Chem. Eur. J. 12, 796 (2006). [23] G.V. Baryshnikov, B.F. Minaev, and V.A. Minaeva. Optics and Spectrosc. 110, 393 (2011). [24] B.F. Minaev, V.A. Minaeva, G.V. Baryshnikov, M.A. Gˆırt¸u, ˚ gren, Rus. J. Appl. Chem. 82, 1211 (2009). and H. A [25] B.F. Minaev, V. Minaeva, G. Baryshnikov, M. Gˆırt¸u, and H. ˚ gren, Mol. Simulat. 37, 670 (2011). A [26] J. Bossert and C. Daniel, Coord. Chem. Rev. 252, 2493 (2008). [27] F. Alary, M. Boggio-Pasqua, J.-L. Heully, C.J. Marsden, and P. Vicendo, Inorg. Chem. 47, 5259 (2008). [28] A.L. Kaledin, Z. Huang, Q. Yin, E.L. Dunphy, E.C. Constable, C.E. Housecroft, Y.V. Geletii, T. Lian, C.L. Hill, and D.G. Musaev, J. Phys. Chem. A 114, 6284 (2010). [29] X. Zhou, A.-M. Ren, and J.-K. Feng, J. Organometallic Chem. 690, 338 (2005). [30] Z.-Z. Xie and W.-H. Fang, J Molecular Structure: THEOCHEM 717, 179 (2005). ˚ gren, and F. De Angelis, Chem. Phys. 358, [31] B. Minaev, H. A 245 (2009). [32] Y. Gao, S. Sun, and K. Han, Spectrochimica Acta Part A 71, 2016 (2009). [33] C.I. Oprea, F. Moscalu, A. Dumbrava, S. Ioannou, A. Nicolaides, and M.A. Gˆırt¸u, J. Optoelectr. Adv. Mater. 11, 1773 (2009). [34] P.G. Bradley, N. Kress, B.A. Hornberger, R.F. Dallinger, and W.H. Woodruf, J. Am. Chem. Soc. 103, 7441 (1981). [35] A. Basu, H.D. Gafney, and T.C. Strekas, Inorg. Chem. 21, 2231 (1982). [36] L.K. Orman and J.B. Hopkins, Chem. Phys. Lett. 149, 375 (1988). [37] T. Yabe, D.R. Anderson, L.K. Orman, Y.J. Chang, and J.B. Hopkins, J. Phys. Chem. 93, 2302 (1989). [38] Y.J. Chang, X. Xu, T. Yabe, S.-C. Yu, D.R. Anderson, L.K. Orman, and J.B. Hopkins, J. Phys. Chem. 94, 729 (1990). [39] T. Yabe, L.K. Orman, D.R. Anderson, S.-C. Yu, X. Xu, and J.B. Hopkins, J. Phys. Chem. 94, 7128 (1990). [40] J.R. Schoonover and G.F. Strouse, Chem. Rev. 98, 1335 (1998). [41] P.K. Mallick, G.D. Danzer, D.P. Strommen, and J.R. Kincaid, J. Phys. Chem. 92, 5628 (1988). [42] G.D. Danzer and J.R. Kincaid, J. Phys. Chem. 94, 3976 (1990). [43] D.P. Strommen, P.K. Mallick, G.D. Danzer, R.S. Lumpkin, and J.R. Kincaid, J. Phys. Chem. 94, 1357 (1990). [44] B.D. Alexander, T.J. Dines, and R.W. Longhurst, Chem. Phys. 352, 19 (2008). [45] C.I. Oprea, B. Frecus, B.F. Minaev, and M.A. Gˆırt¸u, Molec. Phys. 109, 2511 (2011). [46] R.C. Evans, P. Douglas, and C.J. Winscom, Coord. Chem. Rev. 250, 2093 (2006). [47] I. Tunnell, Z. Rinkevicius, P. Salek, O. Vahtras, T. Helgaker, ˚ gren, J. Chem. Phys. 119, 11024 (2003). and H. A ˚ gren, Chem. Phys. 315, 215 (2005). [48] B. Minaev and H. A 750 ˚ gren, Chem. [49] E. Jansson, B. Minaev, S. Schrader, and H. A Phys. 333, 157 (2007). [50] A.D. Becke, J. Chem. Phys. 98, 5648 (1993). [51] C. Lee, W. Yang, and R.G. Parr, Phys. Rev. B 37, 785 (1998). [52] P.J. Hay and W.R. Wadt, J. Chem. Phys. 82, 299 (1985). 755 [53] T.H. Dunning, Jr. and P.J. Hay, in Modern Theoretical Chemistry, edited by H.F. Schaefer, , (Plenum, New York, 1976), Vol. 3, pp. 1–28. [54] M.E. Casida, C. Jamorski, K.C. Casida, and D.R. Salahub, J. Chem. Phys. 108, 4439 (1998). 760 [55] N. Godbout, D.R. Salahub, J. Andzelm, and E. Wimmer, Can. J. Chem. 70, 560 (1992). [56] V. Barone and M. Cossi, J. Phys. Chem. A 102, 1995 (1998); J. Tomasi, B. Mennucci, and R. Cammi, Chem. Rev. 105, 2999 (2005). 765 [57] M.J. Frisch et al., Gaussian 03, Revision C.02 (Gaussian, Inc., Wallingford CT, 2004). Q3 [58] DALTON, a molecular electronic structure program, Release Dalton2011 (2011). <http://daltonprogram.org> [59] A.P. Scott and L. Radom, J. Phys. Chem. 100, 16502 (1996). 770 [60] R.M. Silverstein, G.C. Bassler, and T.C. Morrill, Spectrometric Identification of Organic Compounds, 4th ed. (John Wiley and Sons, New York, 1981). [61] S.D. Williams, T.J. Johnson, T.P. Gibbons, and C.L. Kitchens, Theor. Chem. Acc. 117, 283 (2007). 775 [62] T.J. Dines and R.D. Peacock, J. Chem. Soc. Faraday Trans. 84, 3445 (1988). [63] T.A. Heimer, C.A. Bignozzi, and G.J. Meyer, J. Phys. Chem., 97, 11987 (1993). [64] M. Muniz-Miranda, J. Raman Spectrosc. 31, 637 (2003). 780 [65] B.F. Minaev, V.A. Minaeva, and G.V. Baryshnikov, Russ. J. General Chem. 81, 2332 (2011). [66] M.L. Laury, M.J. Carlson, and A.K. Wilson, J. Comput. Chem. 33, 2380 (2012). [67] B.F. Minaev, Opt. Spectrosc. 98, 336 (2005). 785 [68] B.F. Minaev, V.A. Minaeva, and G.V. Baryshnikov, Russ. J. General Chem. 81, 2332 (2011). [69] V.A. Minaeva, B.F. Minaev, G.V. Baryshnikov, H. Agren, and M. Pittelkow, Vib. Spectrosc., 61, 156 (2012). [70] W. Finkenzeller and H. Yersin, Chem. Phys. Lett. 377, 299 790 (2003). [71] H. Yersin, Highly Efficient OLEDs with Phosphorescent Materials (Wiley-VCH, 2008). ˚ gren, and F. De Angelis, Chem. Phys. 358, [72] B.F. Minaev, H. A 245 (2009). 795 [73] S. Koseki, M. Schmidt, and M. Gordon, J. Phys. Chem. A 102, 10430 (1998). ˚ gren, O. Vahtras, H. Koch, P. Jørgensen, and T. Hel[74] H. A gaker, J. Chem. Phys. 98, 6417 (1993). ˚ gren, O. Vahtras, and B. Minaev, Adv. Quantum Chem. 800 [75] H. A 27, 71 (1996). [76] M. Maestri, D. Sandrini, V. Balzani, U. Maeder, and A. von Zelewsky, Inorg. Chem. 26, 1323 (1987). [77] M.J. Cook, A.P. Lewis, G.S.G. McAuliffe, V. Skarda, A.J. Thomson, J.L. Glasper, and D.J. Robbins, J. Chem. Soc. 805 Perkin Trans. 2, 1293 (1984). [78] K. Nakamaru, Bull. Chem. Soc. Jpn. 55, 2697 (1982). [79] S. Stagni, E. Orselli, A. Palazzi, L. De Cola, S. Zacchini, C. Femoni, M. Marcaccio, F. Paolucci, and S. Zanarini, Inorg. Chem. 46, 9126 (2007). 810 [80] F. de Angelis, S. Fantacci, and A. Selloni, Nanotechnol. 19, 424002 (2008). TMPH_A_777811 TFJATS-Style4.cls March 5, 2013 20:57 Molecular Physics 815 [81] L.M. Peter, J. Phys. Chem. C 111, 6601 (2007). [82] T.W. Hamann, R.A. Jensen, A.B.F. Martinson, H. Van Ryswykac, and J.T. Hupp, Energy Environ. Sci. 1, 66 (2008) Supplementary information See the supplementary materials for tables of the frequencies, IR and Raman intensities and assignment of the vibration modes, 13 tables of the triplet–singlet transition energies and phosphorescence lifetimes in vacuum, and tables of the triplet–singlet transition energies in water solvent for all eight complexes: [Ru(bpy)3-n(dcbpy)n]2 + and [Rh(bpy)3-n(dcbpy)n]3 + , with n = 0, 1, 2, and 3. 820
© Copyright 2024