Next Article in Journal
Preparation of Fe3O4/TiO2/C Nanocomposites and Their Application in Fenton-Like Catalysis for Dye Decoloration
Next Article in Special Issue
Practical Pd(TFA)2-Catalyzed Aerobic [4+1] Annulation for the Synthesis of Pyrroles via “One-Pot” Cascade Reactions
Previous Article in Journal
Facile Synthesis of Bimetallic Pt-Ag/Graphene Composite and Its Electro-Photo-Synergistic Catalytic Properties for Methanol Oxidation
Previous Article in Special Issue
Highly Efficient Tetranuclear ZnII2LnIII2 Catalysts for the Friedel–Crafts Alkylation of Indoles and Nitrostyrenes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Characterization, and Catalytic Hydrogenation Activity of New N-Acyl-Benzotriazole Rh(I) and Ru(III) Complexes in [bmim][BF4]

Chemistry Department, Faculty of Science, Anadolu University, Eskişehir 26470, Turkey
*
Author to whom correspondence should be addressed.
Catalysts 2016, 6(9), 147; https://doi.org/10.3390/catal6090147
Submission received: 10 August 2016 / Revised: 6 September 2016 / Accepted: 14 September 2016 / Published: 20 September 2016
(This article belongs to the Special Issue Organometallic Catalysis for Organic Synthesis)

Abstract

:
The hydrogenation activity of new N-acyl-benzotriazole Rh(I) and Ru(III) complexes in ionic liquid media is reported in this study. Both complexes were completely soluble in 1-butyl-3-methylimidazolium tetrafluoroborate, [bmim][BF4], and they were able to catalyze the hydrogenation of styrene and 1-octene. While ethylbenzene conversion in styrene hydrogenation reached 84% when the Ru complex was used, 100% conversion was obtained with the Rh complex at 393 K in 6 h. Additionally, total conversion in 1-octene hydrogenation reached 100% with the Rh complex in [bmim][BF4] media. The hydrogenation of styrene and 1-octene in dimethyl sulfoxide (DMSO) and toluene was also studied to compare the solvent effect on catalytic system. The effect of some catalytic parameters such as temperature, H2 (g) pressure, and catalyst amount on the conversion was examined, and it was found that the conversion increased parallel to the increasing temperature and H2 pressure. The recyclability of catalysts was also investigated, and it was revealed that the Rh complex in particular maintained the activity for at least 10 cycles.

Graphical Abstract

1. Introduction

As a reaction media, organic solvents are main consumption inputs in synthetic chemistry. In a chemical reaction, decreasing, or if possible completely removing, the usage of organic solvents is an important goal of green chemistry. For this reason, scientists have discovered several alternatives to conventional reaction media. Among these, ionic liquids are coming into prominence for several of their properties such as thermal stability, low vapor pressure, non-flammability, and so on [1,2,3]. Ionic liquids (ILs) are non-molecular solvents with low melting points (<100 °C). They are able to dissolve apolar and polar organic, inorganic, and polymeric materials. Various organic reactions are successfully carried out in ionic liquids. Additionally, they are used as reaction media in reactions with transition metals such as oxidation [4,5], hydrogenation [6,7], hydroformylation [8,9], oligomerization [10,11], and so on [12,13]. Chauvin et al. performed the first hydrogenation reaction in ILs ([bmim][BF4] or [bmim][SbF6]) with [Rh(nbd)(Ph3P)2]PF6 (nbd = norbornadiene) complex and reported a higher activity than that in acetone media [14].
Researchers have developed a wide range of metal complexes for use as a catalyst in ILs media. Heterocyclic compounds with at least one O, S, or N atom in the aromatic ring are attractive ligands for transition metal catalysis. Nitrogen containing molecules such as benzimidazole, benzotriazole, indole, and triazole can be coordinated to a metal center as mono-, bi-, ter-, and tetra-dentate ligands. One of these ligands is the 1H-benzotriazole (C6H5N3) molecule, which is an aromatic heterocyclic compound made up of benzene and a triazole ring. The synthesis and catalytic application of different benzotriazole metal complexes are reported in related literature. Verma et al. synthesized and characterized cheap and thermally stable benzotriazole derivative ligands and their palladium complexes and reported these complexes as efficient catalysts in C–C (Suzuki, Heck, Fujiwara–Moritani, Sonogashira), C–N, and C–S coupling reactions [15]. Das et al. synthesized and characterized 1-(phenylselenomethyl)-1H-benzotriazole (L1) and 1-(4-metoxyphenyltelluromethyl)-1H-benzotriazole (L2) ligands as well as their Pd(II) and Ru(II) complexes. They reported that both complexes are suitable for C–C coupling reactions (Heck and Suzuki–Miyaura) and catalytic oxidation of alcohols [16].
We herein describe the synthesis and characterization of new Rh(I) and Ru(III) complexes ([Rh(COD)L]Cl and [Ru(L)(H2O)Cl3]) including a N-acyl-benzotriazole ligand. We examined the catalytic activity of complexes in hydrogenation reactions in [bmim][BF4] used as reaction media. The effect of temperature, H2 pressure, catalyst amount, and organic solvent on the catalytic activity of complexes was determined. In addition, the reusability of catalysts in styrene hydrogenation was investigated as well.

2. Results and Discussion

2.1. Characterization of Ligand and Complexes

2.1.1. Ligand Characterization

Ligand (L) was obtained with a nucleophilic reaction between 1H-benzotriazole and thiophene-2,5-dicarboxylic acid (Figure 1). In the FT-IR spectrum of ligand, the O–H stretching vibration of thiophene-2,5-dicarboxylic acid at 3000–3100 cm−1 disappeared after the reaction with benzotriazole. The carbonyl stretching vibration of free dicarboxylic acid shifted from 1662 to 1695 cm−1. Additionally, C=N stretching vibration in triazole ring was observed at 1368 and 1223 cm−1 values (see Supplementary Materials, Figure S1).
In the 1H NMR spectra, thiophene protons were observed at δ = 8.65 ppm as a singlet peak. The doublet peaks at δ = 8.46 and 8.25 ppm and triplet peaks at δ = 7.78 and 7.62 ppm were associated with the benzotriazole moiety of the ligand. Due to its symmetrical nature, the ligand with 18 carbon atoms was observed to have nine carbon signals in its C NMR spectrum. The carbonyl carbons were observed at δ = 159.02 ppm, and the carbon atoms belonging to the thiophene ring were observed at δ = 146.03 and 137.2 ppm. The other carbons were observed to be between δ = 141.7 and 114.8 ppm.

2.1.2. Rhodium Complex Characterization

In the FT-IR spectra, the Rh–N stretching vibration was observed at 423 cm−1 as a new peak, which indicated that the ligand coordinated to the metal center via the second nitrogen atom, N(2), of triazole moiety [17]. C=O and C–N vibrations of the ligand at 1689 and 1222 cm−1 were also observed in complex spectrum (see Supplementary Materials, Figure S2).
In the 1H NMR of the Rh complex, the singlet proton peak of thiophene shifted to δ = 8.56 ppm after the coordination to metal center. Two doublet peaks of benzotriazole shifted to δ = 8.39 ppm as a doublet of doublets. Additionally, two triplet peaks shifted to δ = 7.89 and 7.71 ppm in the complex spectrum. The two new peaks observed at δ = 4.26 and 2.14 ppm are related to the 1,5-cyclooctadiene (COD) protons [18]. Additionally, the two singlet peaks at δ = 3.62 and 1.76 ppm are related to the solvent protons. The carbonyl (C=O) carbon signal was observed at 159.13 ppm, and the carbon atoms of the COD (cyclooctadiene) group were observed at 127.59 and 27.93 ppm. The other carbon signals appeared between 145.8 and 115.01 ppm.
In the UV–Vis spectrum of the ligand, the three absorption bands were observed at 262, 281, and 335 nm, which could be attributed to π π * and n π * intra-ligand charge transfer transitions. The spectrum of the Rh(I) complex displayed three absorption bands at 252, 281, and 329 nm, which could be attributed to ligand metal charge transfer (LMCT) transitions. No d d transition was observed at 200–800 nm.
The mass spectra of Rh(I) complex showed the molecular ion peaks at m/z 661.4 and agree very well with the formula weights of the complex including the solvent (THF, tetrahydrofuran) molecule. This confirms the proposed structures of this complex as a 1:1; M/L complex (see Supplementary Materials, Figure S4). The magnetic susceptilibity of the Rh complex was measured according to the Evans method and was found to be zero. This result indicated that the complex is diamagnetic (μs = 0) consistent with square planar geometry. Eventually, the possible structure of the Rh complex is suggested as in Figure 2.

2.1.3. Ruthenium Complex Characterization

In the FT-IR spectrum of the Ru complex, the Ru–N streching vibration was observed at 434 cm−1, which demonstrated that the ligand coordinated to the metal center via the nitrogen atom of the triazole moiety. The Ru–O strecthing vibration was observed at 450 cm−1 in the IR spectra, and the signal at 4.95 ppm in the 1H NMR spectra indicated that the aqua ligand coordinated to the metal center via the oxygen atom. Additionally, the carbonyl (C=O) and C–N stretching vibrations of the ligand were observed at 1705, 1368, and 1251 cm−1, respectively (see Supplementary Materials, Figure S3).
In the 1H NMR spectra of the ruthenium complex, the aromatic proton signals of the thiophene moiety shifted from δ = 8.65 ppm to δ = 7.73 ppm, and the proton signals of the benzotriazole moiety shifted to higher frequencies (δ = 7.93 ppm and 7.46 ppm) after the M–L interaction. According to the literature, the singlet signal at δ = 3.42 ppm is dedicated to the aqua ligand in the complex structure [19]. In the 13C NMR spectra, the C=O group signals were observed at 162.85 ppm, and the other aromatic carbon signals appeared between 140.2 and 115.4 ppm.
In the UV–Vis spectrum of the Ru complex, three absorbtion bands were observed at 272, 362, and 532 nm in dimethyl sulfoxide (DMSO). The intensive band at 272 nm results from intramolecular π π * transitions. Free ligand n π * transitions shifted from 335 to 362 nm due to a nitrogen-metal charge transfer. Additionally, the d d transition of octahedral complex was observed at 532 nm.
Ru(III) complex showed the molecular ion peaks at m/z 645.1. The calculated mass of Ru(III) complex was 644.9. The observed data were in good agreement with the proposed molecular formula that is [Ru(L)(H2O)Cl3]. In addition to the molecular ion peaks, the spectra exhibit other peaks assignable to various fragments arising from the thermal cleavage of the complexes (see Supplementary Materials, Figure S5). The magnetic susceptibility of the complex was measured as 1.26 BM, which means that d orbitals had one unpaired electron around the metal center (t2g5eg0) and that the complex geometry was octahedral.

2.2. Hydrogenation Activity of Complexes

In order to investigate the solubility of complexes in IL media, 10 mg of catalyst and 1 mL of [bmim][BF4] were put in a high-pressure reactor. The solubility of the catalyst was viewed through a sapphire window of the reactor. Both complexes were found to be homogenously soluble in [bmim][BF4] at certain pressures (10–60 bar) and at certain temperatures (323–393 K). In a typical hydrogenation reaction, catalyst, substrate, and IL were added to the reactor, and the reaction vessel was then purged with argon gas. The solution was heated to the reaction temperature, and H2 (g) was pressurized into the reactor. After the reaction, the products were extracted with hexane and analyzed with gas chromatography (GC).

2.2.1. Catalytic Activity of Rhodium Catalyst

The hydrogenation activity of the Rh complex was tested on two different substrates (styrene and 1-octene). The experimental results are presented in Table 1 and Table 2. For the purpose of investigating the temperature effect on styrene hydrogenation, the experiments were performed at six different temperature values (Entries 1–6). The best ethyl benzene conversion was found to be 39.7% at 393 K in 1 h (Entry 6). When the reaction time was increased to 6 h, the conversion value reached 100% (Entry 7). In order to determine the solvent effect on the catalytic system, the experiments were also performed in DMSO and toluene under optimum reaction conditions (393 K, 10 bar H2, 6 h). The total conversion reached 100% in toluene, but it was only 22.9% in DMSO (Entries 7–9). The results indicated that rhodium complex activity was found higher in toluene and ionic liquid than in DMSO. Because of these results, it can be stated that the potential coordination of DMSO molecules to the metal center changes the catalytic route and that the effectiveness of the catalyst thus decreases in DMSO media. In order to determine the H2 pressure effect on styrene hydrogenation, the catalytic reactions were conducted under 10-, 30-, and 60-bar pressures at 343 K in 1 h. The product conversion and turnover frequency (TOF) reached its highest value under 60-bar H2 (Entry 11).
We studied three different ns/nc ratios to determine the effect of the catalyst amount at 393 K and 10-bar H2 on styrene hydrogenation. It was observed that the total conversion increased in line with the increasing catalyst amount and that the highest TOF value was obtained with a ns/nc ratio of 531 in 6 h (Entry 12). The results can be seen in Table 1.
The reusability of the catalyst was also tested in styrene hydrogenation during the 10 cycles at 393 K and 10-bar H2 pressure for 6 h. After the catalytic reaction, the substrate and product(s) were removed from the IL with n-hexane. Afterward, fresh substrate was added to the reaction media to operate the next cycle. Following 10 cycles, an approximate decrease of 10% was observed in the activity of the Rh catalyst. Figure 3 presents the results obtained from the recycle experiments.
In order to evaluate the activity of the rhodium complex in 1-octene hydrogenation, catalytic experiments were performed at temperatures ranging from 323 to 383 K (Entries 1–7) under 10-bar H2 pressure for 1 h (Table 2). It was found that the total conversion gradually increased in line with the increasing temperature and reached 100% at 383 K (Entry 7). Additionally, the isomerization products of 1-octene were observed during these experiments. The distribution of the products can be seen in Table 2. The product distribution at 383 K consisted of 62.5% n-octane, 26.6% 2-octene, and 10.9% 3-octene, so selectivity to n-octane of catalyst was found to be higher compared with the other isomerization products. 1-octene hydrogenation reactions were also performed in DMSO and toluene. The total conversion was found to be 55.2% and 81.3% at 383 K in DMSO and toluene, respectively (Entries 8 and 9).

2.2.2. Catalytic Activity of Ruthenium Catalyst

The catalytic activity of the ruthenium catalyst was investigated between 363 and 393 K temperatures in styrene hydrogenation. The experimental results are presented in Table 3. The highest ethyl benzene conversion value (34.8%) was obtained at 393 K in 1 h (Entry 4). When the reaction time was increased to 6 h, the conversion reached 84.3% under the same conditions (Entry 5). The catalytic activity of the complex was also tested in organic solvents. It was found that the ethyl benzene conversion was only 19.1% in DMSO, while it was 43.1% in toluene at 393 K (Entries 6 and 7). For this reason, the order of the Ru complex activity in solvents could be suggested to be as [bmim][BF4] > toluene > DMSO.
In order to investigate the H2 pressure effect on styrene hydrogenation, catalytic experiments were conducted at 10-, 30-, and 60-bar H2 pressures at 373 K. The conversion value was found to increase sharply from 43.1% under 10-bar H2 pressure to 95.7% under 30 bar and to reach 100% at 60-bar in 1 h (Entry 9). For the purpose of determining the ns/nc ratio effect, catalytic experiments were performed with three different values (515, 172, and 103) at 393 K. The best ethyl benzene conversion (84.3%) was obtained with ns/nc = 103 for 6 h (Entry 12).
Reusability tests of the ruthenium catalyst in styrene hydrogenation were performed at 393 K and 10-bar H2 for 6 h during the 10 cycles (Figure 4). The ethyl benzene conversion found to be 84.3% in the first cycle decreased approximately 20% in the last cycle.
The ruthenium catalyst was found ineffective for 1-octene hydrogenation between 373 and 393 K under 10-bar H2 pressure (Table 4). The total conversion only reached about 5% in 6 h at 393 K (Entry 3). In addition, the complex activity was tested in DMSO and toluene, and the best activity was obtained in toluene (total conversion = 53.7%, Entry 5).
We also investigated the performance of metal precursors ([Rh2Cl2(COD)2] and RuCl3·xH2O) in [bmim][BF4]. These studies are intended to provide a fundamental understanding of the effect of the ligand on activity. For styrene hydrogenation reaction, lower TOF values were obtained with both metal precursors. However, RuCl3·xH2O showed activity in 1-octene hydrogenation, while our complex was found to be ineffective. Therefore, we observed that the ligand has a positive effect on the catalytic reaction in IL.

3. Materials and Methods

3.1. General

All the reagents were commercially obtained (Sigma-Aldrich Co., St. Louis, MO, USA and Merck Co., Darmstadt, Germany) at the highest quality and used without further purification. The synthesis of ligand and complexes was carried out under a nitrogen atmosphere using standard reaction techniques. The catalytic reactions were performed in a 100-mL stainless steel high-pressure reactor (Parr inst., Moline, IL, USA) containing a magnetic bar, temperature controller, and a sapphire window for viewing. 1H NMR and 13C NMR spectra were recorded on a Bruker 500 MHz DPX FT (Bruker Co., Ettlingen, Germany) spectrometer. Chemical shifts were expressed in δ (ppm) using Tetramethyl Silane (TMS) as the internal standard. Elemental analysis was conducted using a Elementar Vario EL III microanalyzer (Elementar Analysensysteme GmbH, Langenselbold, Germany) device. FT-IR spectra were recorded with a Perkin Elmer Spectrum 100 Spectrometer (Perkin Elmer Inc., Waltham, MA, USA) using KBr discs. Mass spectra were recorded on an Agilent LC-MSD Trap SL spectrophotometer (Agilent Tech., Santa Clara, CA, USA). Melting points were measured with a Stuart SMP-30 melting point apparatus (Biby Scientific Ltd., Staffordshire, UK). Magnetic susceptibility measurements were performed with the Sherwood Scientific MK-1 device (Sherwood Scientific, Cambridge, UK). The product conversions in the catalytic reactions were determined with Thermo Finnigan Trace GC (Thermo Fischer Scientific Inc., Waltham, MA, USA) using a Permabond SE-54-DF-0.25 25 m × 0.32 mm ID column and using helium as carrier gas.

3.2. Synthesis of Ligand and Complexes

3.2.1. Ligand Synthesis

1H-benzotriazole (5.6 g, 0.047 mol) was dissolved in 20 mL of THF and then cooled to 273 K. Thionyl chloride (1.0 mL) was added dropwise to this solution under a nitrogen atmosphere for 30 min. A thiophene-2,5-dicarboxylic acid (1.0 g, 0.0058 mol) solution in 30 mL of THF was added to the first solution and stirred at room temperature for 24 h. At the end of the reaction, the solution was filtered, and the residue was dried under vacuum. For purification, the residue was dissolved in CH2Cl2 and extracted with 6N HCl (3 × 10 mL). The organic phase was dried over anhydrous MgSO4 and evaporated with a rotary evaporator. The ligand was obtained as a white solid (Figure 1). Yield: 82% (1.8 g), mp: 513 K. Anal. Calcd (%) for C18H10N6O2S: C = 57.8; H = 2.7; S = 8.6; N = 22.5. Found: C = 58.1; H = 2.5; S = 8.3; N = 22.8. FT-IR (cm−1): 1695, 1368, 1223. 1H NMR: (500 MHz, CDCl3, δ ppm): 8.65 (2H, s), 8.46 (2H, d, 3J = 8.251), 8.25 (2H, d, 3J = 8.251), 7.78 (2H, t, 3J = 7.46), 7.62 (2H, t, 3J = 7.46). 13C NMR: (500 MHz, CDCl3, δ ppm): 159.02, 146.03, 141.72, 137.2, 131.98, 130.94, 126.84, 120.60, 114.83.

3.2.2. Synthesis of [Rh(COD)L]Cl Catalyst

For the synthesis of the rhodium catalyst, the ligand (L) (304 mg, 0.8 mmol) was dissolved in THF (20 mL), and a Rh2 (COD)2Cl2 (200 mg, 0.4 mmol) solution in 5 mL of THF was added slowly to the ligand solution. The mixture was stirred at 323 K for 72 h. At the end of the reaction, the reaction mixture was filtered, and the yellow product was washed with cold THF and dried under vacuum (Figure 2). Yield: 73% (470 mg) mp: 573–578 K. Anal. Calcd (%) for C26H22ClN6O2SRh: C = 50.3; H = 3.6; S = 5.2; N = 13.5. Found: C = 50.1; H = 3.7; S = 5.3; N = 13.3. LC–MS, m/z (calcd/found): 655.1/661.4 FT-IR (cm−1): 1689, 1222, 423. 1H NMR: (500 MHz, d6-DMSO, δ ppm): 8.56 (2H, s), 8.39 (6H, dd, 3J = 24.09, 8.28), 7.89 (4H, t, 3J = 7.68), 7.71 (4H, t, 3J = 7.62), 4.26 (2H, t, 3J = 7.03), 2.14 (4H, m). 13C NMR: (500 MHz, d6-DMSO, δ ppm): 159.13, 145.87, 141.24, 137.77, 132.02, 131.77, 127.59, 120.88, 115.01, 30.67, 27.98.

3.2.3. Synthesis of [Ru(L)(H2O)Cl3]

The ligand (183 mg, 0.48 mmol) was dissolved in a mixture of ethanol and THF (1:1, 20 mL), and a RuCl3·xH2O (100 mg, 0.48 mol) solution in THF (5 mL) was added dropwise to the solution. The mixture was stirred at 353 K for 48 h. The precipitated black solid was filtered, washed with cold ethanol, and dried under vacuum (Figure 2). Yield: 63% (146 mg), mp > 598 K (decomp.). Anal. Calcd (%) for C18H12Cl3N6O3SRu: C = 36.4; H = 2.2; S = 5.3; N = 14.0. Found: C = 36.3; H = 2.1; S = 5.3; N = 14.1. LC–MS, m/z (calcd/found): 644.9/645.1 FT-IR (cm−1): 1705, 450, 434. 1H NMR: (500 MHz, d6-DMSO, δ ppm): 7.93 (3H, q, 3J = 4.51), 7.80 (1H, d, 3J = 3.80), 7.75 (1H, d, 3J = 3.80), 7.73 (1H, s), 7.46 (4H, q, 3J = 4.56), 4.95 (2H) 13C NMR: (500 MHz, d6-DMSO, δ ppm): 162.97, 140.27, 137.19, 134.13, 133.71, 129.01, 128.32, 125.86, 115.46.

4. Conclusions

In the present study, the Rh(I) and Ru(III) complexes consisting of N-acyl-benzotriazole ligand were synthesized and characterized. These complexes were found to be effective catalysts in styrene and 1-octene hydrogenation in IL ([BMIM][BF4]). The results revealed that the rhodium complex activity was higher than the ruthenium complex activity both in styrene hydrogenation and in 1-octene hydrogenation. In styrene hydrogenation, the best ethyl benzene conversions were obtained in ionic liquid as well as in the apolar organic solvent (toluene) with the Rh complex. Moreover, n-octane selectivity of the Rh complex was found to be higher than the other isomerization products (2-octene and 3-octene). It was also found that the catalytic activity of the two complexes increased parallel to the increasing reaction temperature. At the same time, the increasing hydrogen pressure had a positive effect on the catalytic reactions.
The present study primarily aimed at immobilizing synthesized homogenous catalysts in ionic liquid media so as to recover and reuse the complexes. Based on these results, it can be stated that rhodium complexes can be reused up to ten times with little activity loss in styrene hydrogenation reactions.

Supplementary Materials

The following are available online at www.mdpi.com/2073-4344/6/9/147/s1, Figure S1: Free ligand FT-IR spectrum, Figure S2: Rh complex FT-IR spectrum, Figure S3: Ru complex FT-IR spectrum, Figure S4: Rh complex LC–MS spectrum, Figure S5: Ru complex LC–MS spectrum.

Acknowledgments

We thank the Anadolu University Plant Drug and Scientific Research Center (AUBIBAM) for NMR measurements and The Scientific and Technological Research Council of Turkey (TUBITAK), 2211-Doctorate Grant Programme, for its financial support.

Author Contributions

Hakan Ünver performed all experiments and wrote the manuscript. The characterizations of ligand and complexes were performed by Filiz Yılmaz and contributed to the writing of the paper. Both authors have given approval to the final version of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wilkes, J.S. Properties of ionic liquid solvents for catalysis. J. Mol. Catal. A 2004, 214, 11–17. [Google Scholar] [CrossRef]
  2. Muskawar, P.; Thenmozhi, K.; Bhagat, P. Designing of thermally stable amide functionalized benzimidazolium perchlorate ionic liquid for transamidation of primary carboxamides. Appl. Catal. A 2015, 493, 158–167. [Google Scholar] [CrossRef]
  3. Lee, Y.-S.; Kim, D. Cycling performance of lithium polymer cells assembled by in situ polymerization of a non-flammable ionic liquid monomer. Electrochim. Acta 2013, 106, 460–464. [Google Scholar] [CrossRef]
  4. Li, X.; Cao, R. Selective oxidation of alcohols with H2O2 catalyzed by long chain multi-SO3H functionalized heteropolyanion-based ionic liquids under solvent-free conditions. Catal. Commun. 2015, 69, 5–10. [Google Scholar] [CrossRef]
  5. Lentini, S.; Galloni, P.; Bosch, I. Ionic liquids as reaction media in catalytic oxidations with manganese and iron pyridyl triazacyclononane complexes. Inorg. Chim. Acta 2014, 410, 60–64. [Google Scholar] [CrossRef] [Green Version]
  6. Arras, J.; Ruppert, D. Supported ruthenium catalysed selective hydrogenation of citral in presence of [NTf2] based ionic liquids. Appl. Catal. A 2009, 371, 73–77. [Google Scholar] [CrossRef]
  7. Floris, T.; Kluson, P. Quaternary ammonium salts ionic liquids for immobilization of chiral Ru-BINAP complexes in asymmetric hydrogenation of β-ketoesters. Appl. Catal. A 2009, 366, 160–165. [Google Scholar] [CrossRef]
  8. Norinder, J.; Rodrigues, C. Tandem hydroformylation–acetalization with a ruthenium catalyst immobilized in ionic liquids. J. Mol. Catal. A 2014, 391, 139–143. [Google Scholar] [CrossRef]
  9. Wessercheid, P.; Waffenschmidt, H. Ionic liquids in regioselective platinum-catalysed hydroformylation. J. Mol. Catal. A 2000, 164, 61–67. [Google Scholar] [CrossRef]
  10. Azizov, A.G.; Kalbalieva, E.S. Oligomerization of hexene-1 in the presence of catalytic dystems on the basis of ionic liquids. Petroleum Chem. 2010, 50, 57–65. [Google Scholar] [CrossRef]
  11. Stenzel, O.; Brüll, R. Oligomerization of olefins in a chloroaluminate ionic liquid. J. Mol. Catal. A 2003, 192, 217–222. [Google Scholar] [CrossRef]
  12. Dupont, J.; de Souza, R.F.; Suarez, P.A. Ionic liquid (molten salt) phase organometallic catalysis. Chem. Rev. 2002, 102, 3667–3692. [Google Scholar] [CrossRef] [PubMed]
  13. Welton, T. Room-temperature ionic liquids. Solvents for synthesis and catalysis. Chem. Rev. 1999, 99, 2071–2084. [Google Scholar] [PubMed]
  14. Chauvin, Y.; Mussmann, L.; Olivier, H. A novel class of versatile solvents for two-phase catalysis: Hydrogenation, isomerization, and hydroformylation of alkenes catalyzed by rhodium complexes in liquid 1,3-dialkylimidazoliurn salts. Angew. Chem. Int. Ed. 1995, 34, 23–24. [Google Scholar] [CrossRef]
  15. Verma, A.; Jha, R.; Chaudhary, R. 2-(1-Benzotriazolyl)pyridine: A robust bidentate ligand for the palladium catalyzes C–C (Suzuki, Heck, Fujiwara-Moritani, Sonogashira), C–N and C–S coupling reactions. Adv. Synth. Catal. 2013, 355, 421–438. [Google Scholar]
  16. Das, D.; Singh, P.; Singh, A.K. Palladium and half sandwich ruthenium(II) complexes of selenated and tellurated benzotriazoles: Synthesis, structural aspects and catalytic applications. J. Organomet. Chem. 2010, 695, 955–962. [Google Scholar] [CrossRef]
  17. Richardson, C.; Stell, P.J. Benzotriazole as a structural component in chelating and bridging heterocyclic ligands; ruthenium, palladium, copper and silver complexes. Dalton Trans. 2003. [Google Scholar] [CrossRef]
  18. Christiansen, A.; Selent, D.; Spannenberg, A. Reaction of secondary phosphine oxides with rhodium(I). Organometallics 2010, 29, 3139–3145. [Google Scholar] [CrossRef]
  19. Takahashi, Y. Synthesis of various cationic ruthenium-aqua complexes with hydrotris(3,5-diisopropylpyrazolyl)borate ligand, TpiPrRu(OH2)n(L)3-n·(CF3SO3)(THF)x, and their structures supported by hydrogen-bonding networks. Inorg. Chem. 1998, 37, 3186–3194. [Google Scholar] [CrossRef]
Figure 1. Synthesis of N-acyl-benzotriazole ligand. (THF = Tetrahydrofuran, L = Ligand).
Figure 1. Synthesis of N-acyl-benzotriazole ligand. (THF = Tetrahydrofuran, L = Ligand).
Catalysts 06 00147 g001
Figure 2. Synthesis and possible structures of the complexes.
Figure 2. Synthesis and possible structures of the complexes.
Catalysts 06 00147 g002
Figure 3. Reusability of [Rh(COD)L]Cl catalyst in styrene hydrogenation. Reaction conditions: T = 393 K, PH2= 10 bar, nsub. = 8.69 × 10−4, ncat. = 8.09 × 10−6 mol, ns/nc = 106, t = 6 h.
Figure 3. Reusability of [Rh(COD)L]Cl catalyst in styrene hydrogenation. Reaction conditions: T = 393 K, PH2= 10 bar, nsub. = 8.69 × 10−4, ncat. = 8.09 × 10−6 mol, ns/nc = 106, t = 6 h.
Catalysts 06 00147 g003
Figure 4. Reusability of [Ru(L)(H2O)Cl3] catalyst in styrene hydrogenation. Reaction conditions: T = 393 K, PH2 = 10 bar, nsub. = 8.69 × 10−4, ncat. = 8.35 × 10−6 mol, ns/nc = 103, t = 6 h.
Figure 4. Reusability of [Ru(L)(H2O)Cl3] catalyst in styrene hydrogenation. Reaction conditions: T = 393 K, PH2 = 10 bar, nsub. = 8.69 × 10−4, ncat. = 8.35 × 10−6 mol, ns/nc = 103, t = 6 h.
Catalysts 06 00147 g004
Table 1. Styrene hydrogenation with [Rh(COD)L]Cl catalyst a. TOF (Turnover Frequency); TON (Turnover Number).
Table 1. Styrene hydrogenation with [Rh(COD)L]Cl catalyst a. TOF (Turnover Frequency); TON (Turnover Number).
Catalysts 06 00147 i001
EntryT (K)PH2 (bar)t (h)ns/ncTotal Conversion (%)TONTOF
134310110618.32020
235310110619.22121
336310110624.42626
437310110624.62626
538310110635.83838
639310110639.74343
739310610610010718
8 b39310610622.9254
9 c39310610610010718
1034330110641.64545
1134360110663.46868
1239310653143.523339
1339310617782.414825
14 d39310610678.68514
a Reaction conditions: ncat. = 8.09 × 10−6 mol, nstyr. = 8.69 × 10−4 mol, Vsolv. = 0.5 mL, PH2 = 10 bar; b DMSO; c toluene; d cat. = [Rh2Cl2(COD)2].
Table 2. 1-Octene hydrogenation with [Rh(COD)L]Cl catalyst a.
Table 2. 1-Octene hydrogenation with [Rh(COD)L]Cl catalyst a.
Catalysts 06 00147 i002
EntryT (K)Total Conversion (%)Products (%)TONTOF
(n-Octane)(2-Octene)(3-Octene)
132310.53.74.22.58383
233320.85.99.15.81616
334328.915.27.76.02323
435336.818.211.17.52929
536365.130.720.414.15151
637386.746.024.316.46868
738310062.526.610.97979
8 b38355.227.217.710.34343
9 c38381.352.318.410.66464
10 d38362.931.318.213.44949
a Reaction conditions: ncat. = 8.09 × 10−6 mol, n1-oct. = 6.37 × 10−4 mol, ns/nc = 79, Vsolv. = 0.5 mL, PH2 = 10 bar, t = 1 h; b DMSO; c toluene; d cat. = [Rh2Cl2(COD)2].
Table 3. Styrene hydrogenation with [Ru(L)(H2O)Cl3] catalyst a.
Table 3. Styrene hydrogenation with [Ru(L)(H2O)Cl3] catalyst a.
EntryT (K)PH2 (bar)t (h)ns/ncTotal Conversion (%)TONTOF
136310110312.61313
237310110322.22323
338310110322.92424
439310110334.83636
539310610384.38815
6 b39310610319.1203
7 c39310610343.1457
837330110395.7100100
9373601103100104104
1039310651547.224641
1139310617254.99516
12 d39310610371.67512
a Reaction conditions: ncat. = 8.69 × 10−6 mol, nstyr. = 8.69 × 10−4 mol, Vsolv. = 0.5 mL, PH2 = 10 bar; b DMSO; c toluene; d cat. = RuCl3·xH2O.
Table 4. 1-Octene hydrogenation with [Ru(L)(H2O)Cl3] catalyst a.
Table 4. 1-Octene hydrogenation with [Ru(L)(H2O)Cl3] catalyst a.
EntryT (K)Total Conversion (%)Products (%)TONTOF
(n-Octane)(2-Octene)(3-Octene)
13732.52.50020.3
23833.23.20020.4
33934.44.40030.6
4 b39361.80050.8
5 c39353.711.528.413.8417
6 d39375.565.87.22.55810
a Reaction conditions: ncat. = 8.69 × 10−6 mol, n1-oct. = 6.37 × 10−4 mol, ns/nc = 76, Vsolv. = 0.5 mL, PH2 =10 bar, t =6 h; b DMSO; c toluene; d cat. = RuCl3·xH2O.

Share and Cite

MDPI and ACS Style

Ünver, H.; Yılmaz, F. Synthesis, Characterization, and Catalytic Hydrogenation Activity of New N-Acyl-Benzotriazole Rh(I) and Ru(III) Complexes in [bmim][BF4]. Catalysts 2016, 6, 147. https://doi.org/10.3390/catal6090147

AMA Style

Ünver H, Yılmaz F. Synthesis, Characterization, and Catalytic Hydrogenation Activity of New N-Acyl-Benzotriazole Rh(I) and Ru(III) Complexes in [bmim][BF4]. Catalysts. 2016; 6(9):147. https://doi.org/10.3390/catal6090147

Chicago/Turabian Style

Ünver, Hakan, and Filiz Yılmaz. 2016. "Synthesis, Characterization, and Catalytic Hydrogenation Activity of New N-Acyl-Benzotriazole Rh(I) and Ru(III) Complexes in [bmim][BF4]" Catalysts 6, no. 9: 147. https://doi.org/10.3390/catal6090147

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