Next Article in Journal
7-((5-Bromo-1H-indol-3-yl)(4-methoxyphenyl)methyl)-1,3,5-triaza-7-phosphaadamantan-7-ium Tetrafluoroborate
Previous Article in Journal
Diethyl 2-Cyano-3-oxosuccinate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Synthesis and X-ray Structures of Potential Light-Harvesting Ruthenium(II) Complexes

Department of Chemistry, University of South Dakota, Vermillion, SD 57069, USA
*
Author to whom correspondence should be addressed.
Molbank 2023, 2023(2), M1635; https://doi.org/10.3390/M1635
Submission received: 7 March 2023 / Revised: 13 April 2023 / Accepted: 20 April 2023 / Published: 27 April 2023

Abstract

:
We synthesized the luminescent ruthenium(II) polypyridyl complexes of type [Ru(bpy)2(L1)][ClO4]2 (1) (where L1 = 4,4-dicarboxy-2,2-bipyridine); [Ru(bpy)2(L2)][ClO4]2 (2); and [Ru(L2)3][ClO4]2 (3) (where L2 = 4,4-dimethanol-2,2-bipyridine). Photo-physical and electrochemical properties of the Ru(II) complexes were investigated along with the emission vs. pH. This reveals that the carboxylic acids in the 2,2-bipyridine ligand had a more important influence on the photophysical and electrochemical properties of the Ru(II) complexes than alcohol. The crystal structure of the Ru(II) complexes 13 is also discussed in this paper. The cyclic voltammetry of 13 yields a reversible RuIII/II wave that shifts 1.4–1.2 V. UV/Visible absorbance spectroscopy reveals that Metal-to-Ligand Charge Transfer (MLCT) transitions shift to lower energy upon deprotonation of the complex.

1. Introduction

Energy is a basic need and is essential for the earth’s life [1]. The increasing consumption of non-renewable energies such as fossil fuels has caused an adverse effect on ecosystems, causing pollution and global warming, leading to a focus on sustainable and renewable energy sources [2]. Two crucial renewable energy sources have been a significant focus in the 21st century for a clean and sustainable environment: solar and hydrogen fuel cells. Solar cells use the light energy from the sun, whereas hydrogen fuel cells combine hydrogen and oxygen to produce electricity. However, the cost and non-eco-friendly waste to construct the solar cells and generate hydrogen has limited the scope of these renewable sources of energy.
Ruthenium complexes are well-known as the photosensitizer for Dyes-Sensitized Solar Cells (DSSCs) and water splitting due to their high-oxidized state stability and photo-electro-chemical properties, making their practical application feasible [3]. At present, several ruthenium(II)-polypyridyl complexes have been employed as active materials for DSSCs with overall power conversation efficiencies of over 11% under standard illumination, which is attributed to a wide-absorption range (visible to near-infrared) of material [4]. The absorption range of ruthenium(II)-polypyridyl complexes can be tuned carefully to improve the optical properties by considering the HOMO and LUMO energy levels for their potential application in DSSCs.
In this regard, we have synthesized a couple of ruthenium(II)-polypyridyl complexes (cis-[Ru(bpy)2-(bpy-X)]) with different substituent groups (X = 4,4-dicarboxy-2,2-bipyridine; 4,4-methanol-2,2-bipyridine) on the 2,2-bipyridine ligand, as shown in Scheme 1, to study the effect of the substituent in photoelectric conversion. The spectroscopy and photochemistry of complexes [Ru(bpy)2L]2 have been of particular interest (for example, L = dicarboxy-4,4-bipyridine) because of their longer emission lifetimes and higher emission quantum yields. In such mixed-ligand complexes, the electron is largely localized on that ligand, which is more easily reduced. So, bis-(2,2-bipyridine)(4,4-dicarboxy-2,2-bipyridine)ruthenium(II) perchlorate (1), bis-(2,2-bipyridine)(4,4-methanol-2,2-bipyridine)ruthenium(II) perchlorate (2) as well as tris-(4,4-methanol-2,2-bipyridine)ruthenium(II) perchlorate (3) were synthesized. Detailed studies on the properties of all the complexes are reported in this paper. In addition, the photophysical and redox properties of such transition metal complexes can provide important information regarding the nature of the ground and excited states. These experiments aimed to investigate the photophysical and electrochemical properties of these complexes.

2. Results and Discussion

2.1. NMR Spectroscopy

The 1H and 13C NMR spectra for compounds 13 were recorded in CD3CN. Compounds 13 showed appropriate numbers of multiplets in the aromatic region, and the –COOH proton did not show in the 1H NMR spectrum of compound 1. Functionalized BPY’s aromatic protons are de-shielded. Compounds 2 and 3 exhibited peaks at 4.74 ppm and 5.73 pm due to the –CH2OH group. Further, as expected, all resonances were sharp and had well-defined splitting patterns. The aromatic and aliphatic proton ratios matched well with the proposed structure. Compounds 13 produced the expected number of signals in the 13C NMR. The NMR spectra of compounds 13 are shown in Figures S1–S6.

2.2. Crystallization and Structure Determination

The structures of [Ru(bpy)2{bpy(COOH)2}]2+ (1), [Ru(bpy)2{bpy(CH2OH)2}]2+ (2) and [Ru{bpy(CH2OH)2}3]2+ (3) were studied by X-ray diffraction, as shown in Figure 1. The crystallographic data are given in Table 1, and selected bond distances and angles are listed in Table 2.
The crystal structure of 1 has previously been reported without counter-ions or heavily hydrated with different cell parameters, and ruthenium(II) has been reported to be balanced by the deprotonation of carboxylic acid [5,6,7]. In this case, the ruthenium(II) of complex 1 was balanced with the presence of perchlorate anions as counter-ions and solvated with acetonitrile (Figure S10). The bond length of C=O (1.199 Å), C-O (1.328 Å), average Ru-N distance (2.065 Å), average bite angle of N-Ru-N (78.70°), and other average angles of N-Ru-N (92.08° and 172.81°) are similar to the crystal structure reported previously [5].
The crystal structure of 2 has previously been reported with different cell parameters, with hexafluorophosphate as counter-ions and as solvated with water and acetone [8]. The average bond length of C-O (1.341 Å), average Ru-N distance (2.056 Å), average bite angle of N-Ru-N (78.82°), and other average angles of N-Ru-N (92.08° and 174.30°) are similar to those reported for other bipyridyl-coordinated [Ru(bpy)3]2+ and [Ru(bpy)2(bpy(OH)2)]2+ complexes [5,9,10]. Here, the ruthenium(II) is balanced with the two perchlorate anions.
The crystal structure of 3 is reported here for the first time. The average bond length of C-O, average Ru-N length, average bite angle, and other angles of N-Ru-N of 3 are presented in Table 1 and are found to be similar to the crystal structure of 2. However, the three hydroxymethyl groups are disordered; ruthenium(II) is balanced with two perchlorate anions and solvated with acetonitrile and water molecules.

2.3. Optical Properties

2.3.1. Absorbance Spectroscopy

UV/Visible absorption data were collected for 1, 2, and 3 using a 5 × 10−5 molar solution in water (Figure 2). The observed absorbance bands were like those seen for [Ru(bpy)3]2+ in water. Several intense transitions in the wavelength range from 240 to 300 nm are assigned to π-π* transitions. The electronic transitions that appear at wavelengths higher than 300 nm in [Ru(bpy)3]2+ result from many overlapping MLCT bands from the metal-centered d-orbitals to the ligand π* orbitals and are, therefore, assigned similarly for 13. The lowest energy MLCT transition observed for compounds 13 occurs at λmax = ~460 nm, which are like the corresponding MLCT transitions in [Ru(bpy)3]2+max = 451 nm). Compound 1’s MLCT band is slightly more red-shifted than compounds 2 and 3 due to the electron-withdrawing –COOH group.
These wavelength shifts scale with the ligands’ electron-donating ability, which destabilizes the filled d-orbitals, resulting in lower transition energies. In addition, these results follow the same trend observed for the RuIII/II redox potential as a function of ligands. Upon deprotonation of 23 with aqueous t-butylammonium hydroxide or aqueous NaOH in an aqueous solution to make -CH2O, the spectral region between 300 and 600 nm did not change significantly, as shown in Figures S11 and S12.

2.3.2. Emission Spectroscopy

The emission spectra of 13 (Figure 3) in acetonitrile showed a nice Gaussian curve with an emission maximum at 660 nm, 610 nm, and 618 nm, respectively. Compound 2’s emission is ~50 nm more blue-shifted than compound 1 due to the –CH2OH group. The emission maximum of compound 3 is close to compound 1.

2.4. Cyclic Voltammetry

It is important to know the details of the redox process of Ru3+/2+ to understand the redox chemistry of the ruthenium complexes 13, which is presented in Figures S7–S9. The redox chemistry of complexes 13 in CH3CN is presented in Table 3 as compared with Ru(bpy)32+. The cyclic voltammogram of Ru(bpy)32+ corresponds to the reversible one metal-based Ru3+/2+ oxidation process and three one-electron BPY-based reduction processes [11]. We observe similar redox chemistry for all complexes; however, in compound 3, the three one-electron reduction processes are merged together into a single broad peak at −1.34 V, as shown in Figure 4, which is due to the effect of the substituent in the bipyridine unit.
In mixed-ligand complexes such as compounds 12, these electron transitions upon optical absorption would occur between the metal center and the ligand, which is most easily reducible. The electron-withdrawing character of the carboxylic acid group would shift the reduction potential of the ligand positively relative to that of the unsubstituted BPY ligand, as reported earlier [12]. Furthermore, compound 1 showed a promising applicability10 towards Dye-Sensitized Solar Cells. From 1 to 3, the oxidation potentials decreased with an increasing pKa. The data presented in Table 3 on the complexes with carboxylic acid groups and alcoholic functional groups exposes several interesting effects on the Ru→L luminescence.

3. Materials and Methods

3.1. Materials

4,4′-bis(hydroxymethyl)-2,2′-bipyridine,4,4′-dicarboxy-2,2′-bipyridine, [Ru(BPY)2Cl2].xH2O, lithium perchlorate, tetrabutylammonium perchlorate (TBAP), tetrabutylammonium hydroxide and 70% perchloric acid were purchased from Aldrich and used without purification. The perchlorate salts used in the selectivity studies were dried at 100 °C under a vacuum over Drierite to minimize the effects of hydration. CH3CN, THF, DMF, and CH2Cl2 were purchased from Aldrich and purified using a PURE SOLVTM solvent purification system. HPLC-grade anhydrous acetonitrile (Fisher/Acros) was used in all spectroscopic studies.
Caution: Although we have experienced no difficulties with these perchlorate salts, they should be treated as potentially explosive and handled with care.

3.2. Physical Measurements

The 1H and 13C NMR spectra were obtained using Bruker 400 MHz instruments at room temperature and using deuterated solvents. Absorbance data were collected using a Varian Cary 50 BIO UV-visible spectrophotometer. Luminescence titrations were conducted using a Fluoromax–4 spectrofluorometer. Mass spectrometry was conducted using a Varian 500-MS IT ESI mass spectrometer. The cyclic voltammograms were recorded using a CH instruments 660 electrochemical workstation. Elemental analyses were conducted using an Exeter CE-440 Elemental analyzer. Melting points were determined using open capillaries and were uncorrected.

3.3. Single-Crystal X-ray Structure Determination

X-ray quality crystals of compounds 13 was obtained by the diffusion of diethyl ether into an acetonitrile solution. Crystallographic data for 13 was collected at 100 K using a Bruker SMART APEX II diffractometer by MoKα radiation. The data reduction and refinement were completed using the WinGX suite of crystallographic software [13,14]. Structures were solved using SIR97 [15]. All hydrogen atoms were placed in ideal positions and refined as riding atoms with relative isotropic displacement parameters. Table 1 lists additional crystallographic and refinement information.

3.4. Experimental Procedure

Synthesis of Ruthenium(II) Complexes (13)

Synthesis of [bis(2,2′-bipyridine)(4,4′-dicarboxy-2,2′-bipyridine)ruthenium(II)] perchlorate (1):
A slight modification was done in the available procedure9 to synthesize compound 1. A total of 1-g (1.85 mmol) of Ru(BPY)2Cl2.xH2O was mixed with little excess (0.6 g, 2.55 mmol) of 4,4′-dicarboxyl-2,2′-bipyridine along with 0.6 g (7.14 mmol) of sodium bicarbonate in a round bottom flask. The solution was refluxed in 30 mL of water and 20 mL of methanol for 2–3 h under an inert atmosphere. It cooled down on its own after the heating period. A saturated aqueous lithium perchlorate was added to the reaction mixture, and the solution was brought to acidic levels (pH = 5–6) using 70% perchloric acid. Red powder with very high purity obtained over time was filtered and dried under a vacuum. The yield is 70%, and the melting point is over 300 °C. A small portion of the red powder was dissolved in acetonitrile, and diethyl ether was diffused into the solution. Dark red crystals of complex 1 were obtained over time. The elemental analyses calculated for C32H24N6O12RuCl2.3CH3CN included: C, 46.59; H, 3.37; and N, 12.86 %. The following further values were found: C, 46.17; H, 3.21; and N, 12.47 %. The following were found for 1H NMR (CD3CN at 25 °C): 7.39–7.45 (m, 4H, Ar-H); 7.61–7.72 (m, 4H, Ar-H); 7.81–7.83 (m, 2H, Ar-H); 7.93–7.95 (m, 2H, Ar-H); 8.05–8.11 (m, 4H, Ar-H); 8.51–8.53 (m, 4H, Ar-H); and 9.03 (s, 2H, Ar-H). The following were found for 13C NMR (CD3CN at 25 °C): 124.9, 125.4, 127.8, 128.7, 128.8, 139.3, 139.8, 152.6, 152.8, 153.8, 157.6, 157.8, 158.7, and 165.1.
Synthesis of [bis(2,2′-bipyridine)(4,4′-bis(hydroxymethyl)-2,2′-bipyridine)ruthenium(II)] perchlorate (2):
Ru(BPY)2Cl2.xH2O (0.5 g, 0.93 mmol) was mixed with 0.22 g (1.01 mmol) of 4,4′-bis(hydroxymethyl)-2,2′-bipyridine in a round bottom flask that contained 50 mL of absolute ethanol. The solution was refluxed for 6 h under an inert atmosphere and cooled down to room temperature. A saturated aqueous lithium perchlorate was added to the reaction mixture, and the solution was kept in the refrigerator. Red powder was obtained over time, filtered, and dried under a vacuum. The yield was 65%, and the melting point was over 250 °C. The red powder was dissolved in a minimum amount of acetonitrile, and diethyl ether was diffused into the solution. Orange-red crystals of complex 2 were obtained over time. The elemental analyses were calculated for C32H28N6O10RuCl2.CH3CH2OCH2CH3 included: C, 46.39; H, 3.38; and N, 10.14 %. The following further values were found: C, 46.26; H, 3.30; and N, 9.98 %. The following were found for 1H NMR (CD3CN at 25 °C): 4.73–4.75 (d, 2H, CH2-O); 5.72–5.74 (t, 2H, OH); 7.45–7.54 (m, 6H, Ar-H); 7.64–7.66 (m, 2H, Ar-H); 7.73–7.76 (m, 4H, Ar-H); 8.14–8.18 (m, 4H, Ar-H); 8.69 (s, 2H, Ar-H); and 8.82–8.84 (m, 4H, Ar-H). The following were found for 13C NMR (CD3CN at 25 °C): 61.2; 121.3; 124.4; 124.9; 137.8; 150.7; 151.1; 151.2; 154.3; 156.0; and 156.6.
Synthesis of [(4,4′-bis(hydroxymethyl)-2,2′-bipyridine)ruthenium(II)] perchlorate (3):
Ru(DMSO)4Cl2 (0.3 g, 0.62 mmol) was mixed with a 3.3 equivalent of 4,4′-bis(hydroxymethyl)-2,2′-bipyridine (0.42 g) in a round bottom flask that contained 30 mL of ethylene glycol. The solution was refluxed overnight under an inert atmosphere and cooled down to room temperature. A saturated aqueous lithium perchlorate was added to the reaction mixture, and the solution was kept in the refrigerator. Red powder was obtained over time, filtered, and dried under a vacuum. The yield was 50%, and the melting point was over 250 °C. The red powder was dissolved in a minimum amount of acetonitrile, and diethyl ether was diffused into the solution. Red crystals of complex 3 were obtained over time. The elemental analyses were calculated for C36H42N6O14RuCl2 included: C, 45.30; H, 4.40; and N, 8.80 %. The following further values were found: C, 45.17; H, 4.31; and N, 8.65 %. The following were found for 1H NMR (CD3CN at 25 °C): 3.96–3.99 (t, 2H, OH); 4.77–4.79 (d, 2H, CH2-O); 7.45–7.54 (d, 6H, Ar-H); 7.33–7.35 (d, 6H, Ar-H); and 8.47 (s, 6H, Ar-H). The following were found for 13C NMR (CD3CN at 25 °C): 62.6; 122.2; 125.6; 152.1; 154.8; and 157.8.

4. Conclusions

The ruthenium(II) polypyridyl complexes were synthesized and characterized using spectroscopic and electrochemical techniques. The photo-physical and electrochemical properties of the Ru(II) complexes reveal the influence of substituents on 2,2-bipyridine. The DSSC property of compound 1 was tested and reported, and we are currently exploring the application of compounds 2 and 3 as well. The cyclic voltammetry of complexes 13 reveals the Ru3+/2+ oxidation wave that shifts 1.4–1.2 V to lower energy levels, whereas the Metal-to-Ligand Charge Transfer (MLCT) transitions shift to lower energy levels upon deprotonation of the complex. This observation contrasts with mixed-ligand systems containing deprotonate groups, such as -CH2OH, that demonstrate different types of electronic transitions assigned as mixed Metal-Ligand-to-Ligand Charge Transfer (MLCT).

Supplementary Materials

The following supporting information can be downloaded online. Figure S1. 1H NMR of compound 1. Figure S2. 1H NMR of compound 2. Figure S3. 1H NMR of compound 3. Figure S4. 13C NMR of compound 1. Figure S5. 13C NMR of compound 2. Figure S6. 13C NMR of compound 3. Figure S7. CV of compound 1. Referenced vs. Ag/AgCl, glassy carbon, 1 mM in 0.1 M tetrabutylammonium perchlorate. Figure S8. CV of compound 2. Referenced vs. Ag/AgCl, glassy carbon, 1 mM in 0.1 M tetrabutylammonium perchlorate. Figure S9. CV of compound 3. Referenced vs. Ag/AgCl, glassy carbon, 1 mM in 0.1 M tetrabutylammonium perchlorate. Figure S10. ORTEP of compound 1, along with solvents and anions. Figure S11. UV and emission spectra of 13 (5 × 10−5 M) in water with an excess of tetrabutylammonium hydroxide and the excitation wavelength = 450 nm. Figure S12. UV and emission spectra of 13 (5 × 10−5 M) in water with an excess of sodium hydroxide, and the excitation wavelength = 450 nm.

Author Contributions

Conceptualization, K.M. and A.G.S.; methodology, K.M.; software, A.H. and M.A.; validation, K.M., A.H. and M.A.; formal analysis, N.N., T.J.H. and K.A.G.; investigation, N.N., T.J.H. and K.A.G.; resources, K.M.; data curation, N.N., T.J.H. and K.A.G.; writing—original draft preparation, K.M.; writing—review and editing, A.H.; visualization, K.M.; supervision, K.M.; project administration, K.M.; funding acquisition, A.G.S. All authors have read and agreed to the published version of the manuscript.

Funding

N.N., T.J.H. and K.A.G. gratefully acknowledge support for summer REU (NSF CHE 1460872) research at USD. NSF-EPSCoR (EPS-0554609) and the South Dakota Governor’s 2010 Initiative are also appreciated for the purchase of a Bruker SMART APEX II CCD diffractometer. NSF-URC (CHE-0532242) also provided funding for the purchase of the elemental analyzer. The 400 MHz Bruker NMR was also provided by funding from NSF-MRI-CHE-1229035. The authors would like to thank the Center for Fluorinated Functional Materials (CFFM) for financial support and NSF-MRI-CHE-1919637 for the purchase of a Bruker Dual-Source Single-Crystal X-ray Diffractometer.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The crystallographic data (CCDC entry: 1443902, 1857593, 1857586) can be obtained free of charge via www.ccdc.cam.ac.uk/conts/retrieving.html and can be accessed on or after 1 August 2023 (or from the Cambridge Crystallographic Data Centre, 12, Union Road, Cambridge CB21EZ, UK; fax: +44-1223-336-033; or [email protected]). Supplementary information is available on the journal website as a softcopy.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lior, N. Energy resources and use: The present situation and possible paths to the future. Energy 2008, 33, 842–857. [Google Scholar] [CrossRef]
  2. Qin, Y.; Peng, Q. Ruthenium sensitizers and their applications in dye-sensitized solar cells. Int. J. Photoenergy 2012, 2012, 842–857. [Google Scholar] [CrossRef]
  3. Kohle, O.; Grätzel, M.; Meyer, A.F.; Meyer, T.B. The photovoltaic stability of, bis (isothiocyanato) rutheniurn (II)-bis-2, 2′ bipyridine-4, 4′-dicarboxylic acid and related sensitizers. Adv. Mater. 1997, 9, 904–906. [Google Scholar] [CrossRef]
  4. Nazeeruddin, M.K.; Pechy, P.; Renouard, T.; Zakeeruddin, S.M.; Humphry-Baker, R.; Comte, P.; Liska, P.; Cevey, L.; Costa, E.; Shklover, V. Engineering of efficient panchromatic sensitizers for nanocrystalline TiO2-based solar cells. J. Am. Chem. Soc. 2001, 123, 1613–1624. [Google Scholar] [CrossRef] [PubMed]
  5. Beauvilliers, E.E.; Meyer, G.J. Evidence for Cation-Controlled Excited-State Localization in a Ruthenium Polypyridyl Compound. Inorg. Chem. 2016, 55, 7517–7526. [Google Scholar] [CrossRef] [PubMed]
  6. Caspar, R.; Amouri, H.; Gruselle, M.; Cordier, C.; Malézieux, B.; Duval, R.; Leveque, H. Efficient asymmetric synthesis of Δ-and Λ-enantiomers of (bipyridyl) ruthenium complexes and crystallographic analysis of Δ-bis (2, 2′-bipyridine)(2, 2′-bipyridine-4, 4′-dicarboxylato) ruthenium: Diastereoselective homo-and heterochiral ion pairing revisited. Eur. J. Inorg. Chem. 2003, 2003, 499–505. [Google Scholar]
  7. Pang, J.; Di, Z.; Qin, J.-S.; Yuan, S.; Lollar, C.T.; Li, J.; Zhang, P.; Wu, M.; Yuan, D.; Hong, M. Precisely embedding active sites into a mesoporous Zr-framework through linker installation for high-efficiency photocatalysis. J. Am. Chem. Soc. 2020, 142, 15020–15026. [Google Scholar] [CrossRef] [PubMed]
  8. Guillo, P.; Hamelin, O.; Pécaut, J.; Ménage, S. Complexation to [Ru (bpy)2]2+: The trick to functionalize 3, 3′-disubstituted-2, 2′-bipyridine. Tetrahedron Lett. 2013, 54, 840–842. [Google Scholar] [CrossRef]
  9. Rillema, D.P.; Jones, D.S. Structure of tris (2, 2′-bipyridyl) ruthenium (II) hexafluorophosphate, [Ru(bipy)3][PF6]2; X-ray crystallographic determination. J. Chem. Soc. Chem. Commun. 1979, 19, 849–851. [Google Scholar] [CrossRef]
  10. Hansen, L.E.; Glowacki, E.R.; Arnold, D.L.; Bernt, G.J.; Chi, B.; Fites, R.J.; Freeburg, R.A.; Rothschild, R.F.; Krieg, M.C.; Howard, W.A. Syntheses and characterization of some chloro, methoxy, and mercapto derivatives of [Ru (η2-2, 2′-bipyridine)3]2+2PF6: Crystal and molecular structures of [Ru (η2-2, 2′-bipyridine) 2 (η2-4, 4′-(X) 2-2, 2′-bipyridine)]2+ 2PF6(X= Cl, OCH3). Inorg. Chim. Acta 2003, 348, 91–96. [Google Scholar] [CrossRef]
  11. Rillema, D.P.; Allen, G.; Meyer, T.; Conrad, D. Redox properties of ruthenium (II) tris chelate complexes containing the ligands 2, 2’-bipyrazine, 2, 2’-bipyridine, and 2, 2’-bipyrimidine. Inorg. Chem. 1983, 22, 1617–1622. [Google Scholar] [CrossRef]
  12. Fuentes, M.J.; Bognanno, R.J.; Dougherty, W.G.; Boyko, W.J.; Kassel, W.S.; Dudley, T.J.; Paul, J.J. Structural, electronic and acid/base properties of [Ru (bpy (OH)2)3]2+(bpy(OH)2= 4, 4′-dihydroxy-2, 2′-bipyridine). Dalton Trans. 2012, 41, 12514–12523. [Google Scholar] [CrossRef] [PubMed]
  13. Farrugia, L.J. WinGX suite for small-molecule single-crystal crystallography. J. Appl. Crystallogr. 1999, 32, 837–838. [Google Scholar] [CrossRef]
  14. Sheldrick, G.M. SHELXL-97. Program for Crystal Structure Refinement; ScienceOpen, Inc.: Burlington, MA, USA, 1997. [Google Scholar]
  15. Altomare, A.; Burla, M.C.; Camalli, M.; Cascarano, G.L.; Giacovazzo, C.; Guagliardi, A.; Moliterni, A.G.; Polidori, G.; Spagna, R. SIR97: A new tool for crystal structure determination and refinement. J. Appl. Crystallogr. 1999, 32, 115–119. [Google Scholar] [CrossRef]
Scheme 1. Reaction scheme for the synthesis of ruthenium(II) complexes (13).
Scheme 1. Reaction scheme for the synthesis of ruthenium(II) complexes (13).
Molbank 2023 m1635 sch001
Figure 1. ORTEP (50% ellipsoid) of compounds 13. Selected atoms are labeled. Anions and solvents are not shown here for clarity.
Figure 1. ORTEP (50% ellipsoid) of compounds 13. Selected atoms are labeled. Anions and solvents are not shown here for clarity.
Molbank 2023 m1635 g001
Figure 2. UV spectrum of compounds 13. A 5 × 10−5 molar of 13 in water was used for the studies.
Figure 2. UV spectrum of compounds 13. A 5 × 10−5 molar of 13 in water was used for the studies.
Molbank 2023 m1635 g002
Figure 3. Emission spectra of compounds 13. A 5 × 10−5 molar of 13 in water was used for the studies, and the excitation wavelength = 450 nm.
Figure 3. Emission spectra of compounds 13. A 5 × 10−5 molar of 13 in water was used for the studies, and the excitation wavelength = 450 nm.
Molbank 2023 m1635 g003
Figure 4. Cyclic voltammogram of compounds 2 and 3. Referenced vs. Ag/AgCl, glassy carbon, 1 mM in 0.1 M tetrabutylammonium perchlorate.
Figure 4. Cyclic voltammogram of compounds 2 and 3. Referenced vs. Ag/AgCl, glassy carbon, 1 mM in 0.1 M tetrabutylammonium perchlorate.
Molbank 2023 m1635 g004
Table 1. Crystallographic data for compounds 13.
Table 1. Crystallographic data for compounds 13.
Compounds123
Empirical formulaC38H33Cl2N9O12RuC32H28Cl2N6O10RuC76H78N14O31Ru2Cl4
Formula weight979.70828.292027.46
WavelengthMoKα 0.71073MoKα 0.71073MoKα 0.71073
SystemSMART APEXIISMART APEXIISMART APEXII
Temperature, K100(2)100(2)100(2)
Crystal systemtriclinicmonoclinicTriclinic
Space groupP-1P 1 21/c 1P-1
a, Å8.993(3)8.8451(6)10.7926(4)
b, Å14.987(5)30.857(2)11.1969(4)
c, Å15.291(5)14.0432(9)19.3405(8)
α, °93.301(4)9084.03
β, °93.474(4)99.2170(10)80.87
γ, °97.352(4)9062.94
Volume, Å32035.8(12)3783.4(4)2053.54(14)
Z281
Density (calc)
g·cm−3
1.5981.4411.639
Absorb. Coef. Mm−10.5910.6151036
F(000)99616631036
θ range2.51–24.082.42–24.862.26–27.27
Index ranges±10, ±17, ±17±10, ±37, ±16±13, ±14, ±24
Reflections collected183313899124520
Independent reflections653869869261
Observed reflections531156107809
Max/Min trans.0.737–0.943 0.866–0.943
Data/restr./param.6538/0/5626986/2/4629261/0/582
Goodness-of-fit1.0671.1071.071
Final R indices [I > 2σ(I)]0.03900.05220.0623
R indices (all data)0.05390.06400.0751
CCDC Number144390218575931857586
Table 2. Important bond lengths (Å) and angles (°) of compounds 13.
Table 2. Important bond lengths (Å) and angles (°) of compounds 13.
123
C=O (double bond)1.198 Å; 1.199 Å--
C-O (single bond)1.311 Å; 1.346 Å1.279 Å; 1.404 Å1.385 Å (avg)
Ru-N (avg)2.065 Å2.056 Å2.058 Å
N-Ru-N (avg)
(bite angle)
78.20°78.82°78.60°
N-Ru-N (avg)
(other angles)
92.08° & 172.81°92.08° & 174.30°93.07° & 174.00°
Table 3. Electrochemical data; referenced vs. Ag/AgCl, glassy carbon, 1 mM in 0.1 M tetrabutylammonium perchlorate at room temperature.
Table 3. Electrochemical data; referenced vs. Ag/AgCl, glassy carbon, 1 mM in 0.1 M tetrabutylammonium perchlorate at room temperature.
CompoundSolventEA1/2 (V)
BPY reductionOxidation
BPY0/−1BPY−1/−2BPY−2/−3Ru3+/2+
1CH3CN−1.39−1.58−1.87+1.38
2CH3CN−1.31−1.48−1.736+1.29
3CH3CN−1.34+1.19
([Ru(bpy)3]2+) [11]CH3CN−1.31−1.50−1.77+1.27
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mariappan, K.; Hussain, A.; Nisly, N.; Henning, T.J.; Goerl, K.A.; Alaparthi, M.; Sykes, A.G. Synthesis and X-ray Structures of Potential Light-Harvesting Ruthenium(II) Complexes. Molbank 2023, 2023, M1635. https://doi.org/10.3390/M1635

AMA Style

Mariappan K, Hussain A, Nisly N, Henning TJ, Goerl KA, Alaparthi M, Sykes AG. Synthesis and X-ray Structures of Potential Light-Harvesting Ruthenium(II) Complexes. Molbank. 2023; 2023(2):M1635. https://doi.org/10.3390/M1635

Chicago/Turabian Style

Mariappan, Kadarkaraisamy, Anwar Hussain, Nathaniel Nisly, Tanner J. Henning, Kathryn A. Goerl, Madhubabu Alaparthi, and Andrew G. Sykes. 2023. "Synthesis and X-ray Structures of Potential Light-Harvesting Ruthenium(II) Complexes" Molbank 2023, no. 2: M1635. https://doi.org/10.3390/M1635

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop