Previous Article in Journal
Smart Inorganic Nanomaterials for Tumor Microenvironment Modulation
Previous Article in Special Issue
Recent Advances in Zinc Complexes for Stereoselective Ring-Opening Polymerization and Copolymerization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Catalytic Activity of Rhenium(I) Tricarbonyl Complexes Containing Polypyridine and Phosphorus–Nitrogen Ligands in the Hydrogen Transfer of Acetophenone

by
César Zúñiga
1,2,*,
Mauricio Fuentealba
3,
Elizabeth Olave
3,
Diego P. Oyarzún
4,
Andrés Aracena
2,5,
Mauricio Yañez-S
6,
Plinio Cantero-López
7,8,9 and
Pedro A. Aguirre
10,*
1
Instituto de Ciencias Naturales, Facultad de Medicina Veterinaria y Agronomía, Universidad de Las Américas, Sede Santiago, Campus Providencia, Av. Manuel Montt 948, Providencia, Santiago 7500000, Chile
2
Centro de Investigación en Ciencias Biológicas y Químicas, Universidad de Las Américas, Av. Manuel Montt 948, Providencia, Santiago 7500000, Chile
3
Instituto de Química, Pontificia Universidad Católica de Valparaíso, Av. Universidad 330, Curauma, Valparaíso 2340000, Chile
4
Departamento de Química y Biología, Facultad de Ciencias Naturales, Universidad de Atacama, Copayapu 485, Copiapó 1530000, Chile
5
Instituto de Ciencias Naturales, Facultad de Medicina Veterinaria y Agronomía, Universidad de Las Américas, Sede Santiago, Campus La Florida, Av. Walker Martínez 1360, La Florida, Santiago 8240000, Chile
6
Departamento de Cs. Biológicas y Químicas, Facultad de Recursos Naturales, Universidad Católica de Temuco, Av. Rudecindo Ortega 2950, Campus San Juan Pablo II, Temuco 4780000, Chile
7
Departamento de Ciencias Químicas, Facultad de Ciencias Exactas, Universidad Andres Bello, Viña del Mar 2531015, Chile
8
Center of Applied Nanoscience (CANS), Facultad de Ciencias Exactas, Universidad Andres Bello, Santiago 8370186, Chile
9
Relativistic Molecular Physics Group (ReMoPh), PhD Program in Molecular Physical Chemistry, Facultad de Ciencias Exactas, Universidad Andres Bello, Santiago 8370146, Chile
10
Facultad de Ciencias Químicas y Farmacéuticas, Universidad de Chile, Olivos 1007, Independencia, Santiago 8380544, Chile
*
Authors to whom correspondence should be addressed.
Inorganics 2025, 13(10), 338; https://doi.org/10.3390/inorganics13100338 (registering DOI)
Submission received: 8 September 2025 / Revised: 11 October 2025 / Accepted: 13 October 2025 / Published: 14 October 2025

Abstract

This work reports the synthesis and characterization of a novel rhenium(I) complex incorporating a phosphorus–nitrogen (P,N) ligand. The compound crystallizes in a distorted octahedral geometry, as confirmed by single-crystal X-ray diffraction analysis. The complexes were evaluated as catalysts in the transfer hydrogenation of acetophenone using 2-propanol as the hydrogen source. Comparative studies with other rhenium(I) complexes bearing polypyridine ligands revealed high catalytic performance, achieving conversions between 68% and 99%. These results highlight the promising potential of P,N-ligand rhenium complexes in homogeneous transfer hydrogenation catalysis. The optimal reaction time was found to be 4 h for the complexes studied, with only the fac-[Re(CO)3(PN)Cl] complex showing improved conversion upon extending the reaction time to 7 h, likely due to the donor effects provided by the P,N-ligand.

1. Introduction

Over the past several decades, the synthesis and design of transition metal complexes featuring pyridine-derived and hetero-bidentate phosphorus-nitrogen (P,N) ligands [1,2,3] have garnered significant interest due to their diverse applications in various fields of chemistry [4,5,6]. Among these, rhenium carbonyl complexes are particularly promising, owing to their potential utility in areas such as microscopy, light-emitting diodes (OLEDs) [7,8,9], luminescence-based materials [10,11,12,13], photosensitizers in solar cells [14], sensors, and biomedical technologies. These applications are largely attributed to their unique photophysical and photochemical properties, which can be finely tuned through appropriate ligand design [15,16,17,18,19]. Moreover, rhenium complexes have been extensively explored in the context of electrocatalytic and photocatalytic CO2 reduction [20,21,22,23,24].
Transition metal complexes are also widely recognized for their catalytic activity in transfer hydrogenation (TH) reactions involving ketones and imines [25,26,27,28,29]. Significant advances have been made using catalysts based on ruthenium [30,31,32,33,34,35,36,37,38], rhodium [39,40,41], iron [42], and iridium [43,44,45,46]. In transfer hydrogenation, a hydrogen atom is delivered from a donor molecule—typically an alcohol—to a substrate, rather than using molecular hydrogen. The mechanism of hydrogen transfer can vary, often involving the concerted or stepwise transfer of a hydride and proton, and is highly dependent on the nature of the metal center, ligand environment, and reaction conditions [25]. Despite the relatively high cost of some catalysts, TH reactions remain attractive due to their operational simplicity, avoidance of pressurized hydrogen gas, and high efficiency under mild conditions.
Rhenium(I) complexes bearing polypyridine or phosphorus-based ligands have been comparatively less explored in the context of TH catalysis. Notably, Landwehr et al. [47] reported the synthesis and catalytic evaluation of [Re(H)(NO)(PR3)(C5H4OH)] complexes in the transfer hydrogenation of ketones and imines using 2-propanol as a hydrogen donor.
This article reports the catalytic activity of a series of Re(I) complexes containing either a polypyridine ligand or a phosphorus–nitrogen (PN) ligand in the hydrogen transfer reaction. The hydrogen transfer was performed using acetophenone as the substrate, and the conversion was studied at different reaction times. In addition, the article describes the synthesis and characterization of a new fac-[Re(CO)3(PN)Cl] complex. The rhenium complexes exhibited catalytic activity in the hydrogen transfer of acetophenone, achieving conversions between 68% and 99% after 7 h of reaction. The results suggest that when coordinated to a Re(I) carbonyl precursor, the ligands promote the hydrogen transfer reaction due to their donor effect.

2. Experimental Section

2.1. General Procedure

All reactions were carried out under an argon atmosphere. Organic solvents were dried using standard procedures prior to use. [Re(CO)5Cl] and diphenyl-2-pyridylphosphine were purchased from Aldrich and used without further purification. Infrared spectra were recorded using a PerkinElmer FTIR “Spectrum Two” spectrometer (Shelton, CT, USA) equipped with a UATR accessory. Samples were directly placed on the diamond crystal, pressed to 30% of the total supported pressure, and scanned in the range of 4000–500 cm−1 with a resolution of 1 cm−1. 1H and 31P{1H} NMR spectra were acquired on a Bruker AVANCE 400 MHz spectrometer (Billerica, MA, USA). All chemical shifts are reported in parts per million (ppm) relative to tetramethylsilane (TMS) as the internal standard. Elemental analyses were conducted using a CE Instruments EA 1108 elemental analyzer (Cornaredo, Italy). The following rhenium complexes were synthesized according to previously reported procedures: [Re(CO)3(PyCOOH)2Cl] [48], [Re(CO)3(bpy)Cl] [49], [Re(CO)3(Phen)Cl] [50], [Re(CO)3(Phen-NO2)Cl] [51], [Re(CO)3(dppz-NO2)Cl] [51], and [Re(CO)3(dppz-CN)Cl] [52,53].
Catalytic tests were conducted under a nitrogen atmosphere in a 30 mL glass reactor equipped with a reflux condenser. Typically, the rhenium complex (0.01 mmol) was dissolved in 2-propanol (8 mL) and heated to reflux under vigorous stirring. Sodium hydroxide (1 mL, 0.1 M) was then added dropwise, followed by acetophenone (10 mmol). The catalytic reaction’s progress was monitored by gas chromatography (GC), with periodic sampling being performed every 10 min. GC analysis was carried out with a Pekin Elmer Clarus 580 chromatograph instrument (Shelton, CT, USA) equipped with an FID detector, using an Elite-5 capillary column and nitrogen as carrier gas. The GC analyses were carried out using an internal standard to establish the linearity between the conversion percentage and the concentration of the products formed in the catalytic reaction. GC-mass spectra were run to confirm the identity of the products on a MAT 95 XP Thermo Electron (Bremen, Germany).

2.2. Synthesis of Fac-[Re(CO)3(PN)Cl] Complex

Rhenium pentacarbonyl chloride (180 mg, 0.50 mmol) and diphenyl-2-pyridylphosphine (132 mg, 0.50 mmol) were dissolved in chloroform (30 mL) and refluxed for 24 h (Scheme 1). After cooling to room temperature, the solvent was removed under reduced pressure. The resulting solid was redissolved in chloroform (5 mL) and precipitated by the addition of diethyl ether. The precipitate was cooled at 4 °C, filtered, washed with small amounts of cold diethyl ether, and dried under vacuum. Yield: 134 mg (54%). Anal. Calc. for ReC20H14NPO3Cl: C, 42.22; H, 2.48; N, 2.46. Found: C, 42.77; H, 2.63; N, 2.49%. FT-IR (ATR, cm−1): νC=O = 2022, 1917, 1893.1H NMR ((CD3)2CO) δ (ppm), J (Hz): 8.90 (d, J = 5.2, 1H-Py), 8.28 (t, J = 7.9, 1H-Py), 8.10 (dd, J = 7.8, 3.5, 1H-Py), 7.92–7.68 (m, 5H aromatic protons), 7.65–745 (m, 6H aromatic protons-Py). 31P{1H} NMR ((CD3)2CO) δ (ppm): −31.01. Single crystals suitable for X-ray diffraction were grown via vapor diffusion of diethyl ether into a dichloromethane solution of the complex.

2.3. X-Ray Crystal Structure Determination for Fac-[Re(CO)3(PN)Cl] Complex

X-ray suitable crystals were obtained as described above. A suitable single crystal was mounted using MiTeGen MicroMounts (Ithaca, NY, USA). Table 1 shows experimental and crystallographic data for the obtained complex. On the other hand, selected bond distances and angles are reported in Table 2. Intensity data were collected at room temperature on a Bruker D8 QUEST diffractometer (Karlsruhe, Germany) equipped with a bidimensional CMOS Photon100 detector (Madison, WI, USA), using graphite monochromated Mo-Kα radiation. The diffraction frames were integrated by means of the APEX3 package and were corrected for absorptions with SADABS. The solution and refinement for the Re(I) complex was carried out with Olex2 [54]. The structure was solved with Patterson Method using the ShelXS software v2018/3 [55]. The complete structures were refined by the full matrix least-squares procedures of the reflection intensities (F2) with the ShelXL [55]. All non-hydrogens were refined with anisotropic displacement coefficients, and all hydrogen atoms were placed in idealized locations.

3. Results and Discussion

3.1. Synthesis and Characterization of Fac-[Re(CO)3(PN)Cl] Complex

The new Rhenium(I) complex was prepared by mixing one equivalent of [Re(CO)5Cl] compound and one equivalent of the diphenyl-2-pyridylphosphine in refluxing chloroform for 24 h, as shown in Scheme 1. The obtained complex is stable in air, white microcrystalline solid, and soluble in common organic solvents. In the structure of the synthesized new rhenium(I) complex, bidentate coordination of the ligand to the metal is proposed. This occurs between the phosphorus atom and the nitrogen atom from the pyridine fragment of the ligand, and a stable four-membered chelate ring is formed.
The complex features a nearly ideal octahedral geometry around the rhenium center, with a facial (fac) arrangement of three carbonyl ligands. The characteristic fac-tricarbonyl configuration is supported by the IR spectrum, displaying one strong band at 2022 cm−1 and two lower-frequency bands at 1917 and 1893 cm−1, consistent with symmetric and asymmetric CO stretching vibrations.
The new rhenium(I) complex was fully characterized by standard spectroscopic methods (1H and 31P{1H} NMR) and gave satisfactory analysis. In the 31P{1H} NMR spectrum, the coordinated phosphine of fac-[Re(CO)3(PN)Cl] gives rise to a singlet at δ = −31.01 ppm, whereas the free diphenyl-2-pyridylphosphine ligand exhibits a resonance at δ = −4.1 ppm. This significant upfield shift of approximately 27 ppm clearly indicates strong coordination to the rhenium center. The large nuclear shielding observed in the complex is fully consistent with chelation of the P,N ligand, as reported in related Re(I) tricarbonyl complexes containing donor phosphine ligands [56]. This structural assignment is further confirmed by X-ray crystallographic analysis.

3.2. X-Ray Crystallographic Study

The molecular structure of the fac-[Re(CO)3(PN)Cl] complex, complete with an atom-numbering scheme, is illustrated in Figure 1. The compound crystallized in the monoclinic space group P21/c. Selected bond lengths and angles pertinent to the structure are presented in Table 2.
The structural analysis reveals a distorted octahedral coordination sphere with three carbonyl ligands arranged in a fac conformation, accompanied by a chloride ligand and the PN bidentate ligand. The bond angles between the carbonyl ligands C1-Re1-C2, C3-Re1-C1, and C3-Re1-C2 are 90.16 (15), 89.91 (12), and 87.59 (13) degrees, confirming the facial conformation. The bite angle of the PN ligand is 65.13 (5), corresponding to the N1-Re1-P1 bond angle; this is the most significant distortion from the ideal octahedral angle (90°) due to the rigidity of the ligand, which results in a deviation of the octahedral distortion parameters. The octahedral deviation parameters for the coordination sphere of fac-[Re(CO)3(PN)Cl] complex is Σ = 77.6 (4)° ( ( Σ = i = 1 12 φ i   90 ) [57] and ϴ = 168.7 (4)° [58] parameters ϴ = i = 1 24 θ i   60 ) .
On the other hand, Table 3 shows the intermolecular interactions of the fac-[Re(CO)3(PN)Cl] complex. The crystal structure is stabilized by some weak hydrogen bonds (according to Desiraju’s classification). This corresponds to the intermolecular interaction between O1 of the carbonyl ligand and H7 (+x, 3/2 − y, −1/2 + z). Furthermore, it is noted that the Cl1 ligand engages in a dual intermolecular interaction with the hydrogens from the adjacent asymmetric units H6 (+x, 3/2 − y, −1/2 + z) and H5 (2 − x, 1 − y, 1 − z). A Hirshfeld surface analysis was carried out to adequately describe the intermolecular interactions.

Hirshfeld Surface Analysis

In this study, the assessment of Hirshfeld surfaces and their corresponding two-dimensional fingerprint plots was conducted employing the CrystalExplorer software v17.5 [59,60,61]. Input for the analysis was sourced from the crystallographic information file (CIF) pertinent to each molecular complex under investigation.
The intermolecular interactions of the crystal structures of fac-[Re(CO)3(PN)Cl] were quantified using Hirshfeld surface analysis (Figure 2). The more essential contributions of the different interactions are presented in Figure 3. For fac-[Re(CO)3(PN)Cl] complex, the fingerprint plots show that the dominant interaction is H⋯H (32.1%). In the same way, the complex exhibited a significant O⋯H contribution (26.3%) corresponding to intermolecular interactions by hydrogen bonds between the oxygen of carbonyl ligands to hydrogen atoms from the C-H aromatic ring (see above). Furthermore, the Cl⋯H interaction exhibits a percentage of the contribution of 10.7%, corresponding to the weak molecular interaction between the chloro ligand and C-H hydrogen atoms. Finally, the H⋯C and O⋯C interactions, accounting for percentages of 17.7% and 10.4%, respectively, represent the C-H⋯π and CO⋯π interactions observed within the crystalline arrangement. A detailed description of the methodology and the definition of the dnorm parameter, and the full set of fingerprint plots and interaction contributions are provided in the Supplementary Information.

3.3. Catalytic Activity

Reaction aliquots were taken every 1 h and analyzed by gas chromatography (GC) to determine conversion and selectivity. The hydrogen transfer reaction is illustrated in Scheme 2, and the catalytic results are summarized in Table 4.
In homogeneous catalysis, the ligand plays a crucial role not only in stabilizing the metal center but also in modulating the reaction selectivity. In this study, both phosphorus–nitrogen and polypyridine ligands were evaluated, each providing strong donor effects to the rhenium center. For transfer hydrogenation, formation of a metal hydride is generally required for catalytic activity, and such formation is favored when electron-donating ligands are present.
The results in Table 4 indicate that all tested Re(I) complexes were active, achieving conversions between 68% and 99% (after 7 h of reaction time), with 100% selectivity toward the alcohol product. Comparison between the precursor [Re(CO)5Cl] and its derivatives demonstrate that donor ligands significantly enhance catalytic performance, likely by stabilizing the active metal species under reaction conditions. Without such stabilization, the rhenium center may undergo reduction to Re(0), leading to catalyst deactivation.
Turnover numbers (TONs) suggest gradual deactivation over time, as values do not scale proportionally with reaction duration. Among the tested complexes, [Re(CO)3(bpy)Cl] and fac-[Re(CO)3(PN)Cl} exhibited the highest activity after 4 h, with turnover frequencies (TOFs) of 220 h−1 and 200 h−1, respectively. The complex [Re(CO)3(bpy)Cl] does not increase its activity when the reaction time is extended to 7 h, however complex fac-[Re(CO)3(PN)Cl] shows a conversion of 99% when the reaction time is extended. This suggests that the complex [Re(CO)3(bpy)Cl] undergoes deactivation after 4 h of reaction, this is confirmed by the low increase in TON, which rises from 650 to only 700. All catalysts show maximum activity after 4 h of reaction, and only complex fac-[Re(CO)3(PN)Cl] shows a significant increase when the reaction time is 7 h. The activity for the complex fac-[Re(CO)3(PN)Cl] improves from 80% to 99%. This could indicate that the donor effects of this ligand stabilize the complex fac-[Re(CO)3(PN)Cl] under the reaction conditions studied. An experiment was designed to detect the Re–H species by 1H NMR spectroscopy. However, it was not possible to characterize this intermediate, most likely because the Re–H species decomposes rapidly, as suggested by the bubbling observed when transferring the sample from the reactor to the NMR tube. In contrast, our group has successfully detected other transition metal hydrides, such as ruthenium complexes, which are more stable and can be characterized by NMR [62].
For comparison, Landwehr et al. [47] reported 98% conversion in only 1 h using [Re(H)(NO)(PR3)(C5H4OH)] (R = iPr3, Cy) as catalyst in acetophenone transfer hydrogenation. However, their experiments employed a substrate-to-catalyst ratio of 200:1, whereas our study used a ratio of 1000:1 (five times higher) yielding comparable TOFs. This observation reinforces that rhenium carbonyl complexes require initial hydride formation to become catalytically active under mild conditions.
In addition to the catalytic performance, the solubility and stability of the rhenium(I) complexes were evaluated. The complexes are soluble under the catalytic conditions employed, with an estimated solubility of ca. 10 mg/15 mL of isopropanol. This level of solubility was sufficient to carry out the catalytic reactions efficiently and ensured the reproducibility of the experimental results. Furthermore, the complexes exhibited good stability under transfer hydrogenation conditions. As shown in Table 4, the turnover numbers (TONs) increase consistently with reaction time, indicating that the catalysts remain active throughout the reaction with only minor loss of activity. These observations confirm that both solubility and stability contribute to the reliable catalytic performance of the complexes studied.
Overall, the rhenium complexes studied here, particularly those with strong donor ligands, demonstrate promising activity and selectivity in transfer hydrogenation, while offering advantages over noble metal catalysts such as ruthenium and rhodium in terms of cost and ease of synthesis.

4. Conclusions

We have developed a synthetic procedure for obtaining a novel rhenium(I) tricarbonyl complex incorporating a phosphorus–nitrogen ligand (PN = diphenyl-2-pyridylphosphine). The resulting compound, fac-[Re(CO)3(PN)Cl], crystallizes in a distorted octahedral geometry, as confirmed by single-crystal X-ray diffraction analysis.
Catalytic studies demonstrated that fac-[Re(CO)3(PN)Cl] is an effective homogeneous catalyst for the transfer hydrogenation of acetophenone using 2-propanol as a hydrogen donor. High conversion (99%) and complete selectivity (100%) toward the alcohol product were achieved under mild conditions, with a substrate-to-catalyst ratio of 1000:1. Comparable activity was observed for the polypyridine complex [Re(CO)3(bpy)Cl], which exhibited turnover frequency of TOF (220 h−1) and the fac-[Re(CO)3(PN)Cl] (200 h−1) reported at 4 h of reaction time.
The catalytic performance of these rhenium complexes, particularly given the high substrate-to-catalyst ratio, is in line with previously reported systems employing more noble metals, while offering advantages in terms of cost and accessibility. The combination of ease of synthesis, structural stability, and high activity highlights P,N- and polypyridine-ligated rhenium(I) complexes as promising candidates for sustainable homogeneous transfer hydrogenation catalysis.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/inorganics13100338/s1, Figure S1: 1H NMR spectrum of fac-[Re(CO)3(PN)Cl] in acetone-d6; Figure S2: 31P{1H} NMR spectrum of fac-[Re(CO)3(PN)Cl] in (CD3)2CO; Figure S3: FT-IR spectrum of fac-[Re(CO)3(PN)Cl]; Figure S4: FT-IR spectrum of [Re(CO)3(bpy)Cl]; Figure S5: 1H NMR spectrum of fac-[Re(CO)3(Phen)Cl] in CDCl3; Figure S6: FT-IR spectrum of fac-[Re(CO)3(Phen)Cl]; Figure S7: 1H NMR spectrum of fac-[Re(CO)3(Phen-NO2)Cl] in CDCl3; Figure S8: FT-IR spectrum of fac-[Re(CO)3(Phen-NO2)Cl]; Figure S9: 1H NMR spectrum of fac-[Re(CO)3(dppz-NO2)Cl] in DMSO-d6; Figure S10: FT-IR spectrum of fac-[Re(CO)3(dppz-NO2)Cl]; Figure S11: 1H NMR spectrum of fac-[Re(CO)3(dppz-CN)Cl] in CDCl3; Figure S12: FT-IR spectrum of fac-[Re(CO)3(dppz-CN)Cl]; Figure S13: 1H NMR spectrum of fac-[Re(CO)3(PyCOOH)2Cl] in (CD3)2CO; Figure S14: FT-IR spectrum of fac-[Re(CO)3(PyCOOH)2Cl]. Figure S15: Full set of 2D fingerprint plots of fac-[Re(CO)3(PN)Cl] complex for intermolecular contacts. Supplementary data CCDC reference number 1584232 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge at https://www.ccdc.cam.ac.uk/structures (accessed 12 October 2025) or from the Cambridge Crystallographic Data Centre, 12, Union Road, Cambridge CB2 1EZ, UK; Fax: (internat.) +44 1223/336 033; Email: deposit@ccdc.cam.ac.uk.

Author Contributions

Conceptualization, C.Z., M.F. and P.A.A.; methodology, C.Z. and P.A.A.; software, M.F. and E.O.; validation, P.C.-L., D.P.O. and C.Z.; formal analysis, M.F., E.O. and A.A.; investigation, C.Z., M.F. and M.Y.-S.; resources, C.Z., M.F., D.P.O. and P.A.A.; data curation, E.O., M.Y.-S., A.A.; writing—original draft preparation, C.Z., M.F. and P.A.A.; writing—review and editing, C.Z., M.F., A.A., M.Y.-S., D.P.O. and P.A.A.; visualization, C.Z., M.F. and D.P.O.; supervision, C.Z. and P.A.A.; project administration, C.Z., M.F. and P.A.A.; funding acquisition, C.Z., M.F., D.P.O., P.C.-L. and P.A.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Universidad de Las Américas, Chile, Proyecto PIR202414, Vicerrectoría de Investigación y Postgrado, Universidad de Las Américas, Chile. The authors acknowledge financial support from the FONDEQUIP, grant no. EQM120095 (Single crystal X-ray diffractometer), the Vicerrectoría de Investigación y Estudios Avanzados, Pontificia Universidad Católica de Valparaíso, Chile (M.F). E.O. thanks the CONICYT (Chile) for support of a graduate fellowship and finally M.Y.-S. thanks FONDECYT 1231041.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Graham, T.W.; Cariou, R.P.-Y.; Sánchez-Nieves, J.; Allen, A.E.; Udachin, K.A.; Regragui, R.; Carty, A.J. Reactivity of Terminal Electrophilic Phosphinidene Complexes: Synthesis of the First Rhenium Phosphinidene, [Re(CO)51-PNiPr2)][AlCl4], and Novel Reactions with Azobenzene. Organometallics 2005, 24, 2023–2026. [Google Scholar] [CrossRef]
  2. Cowley, A.R.; Dilworth, J.R.; Nairn, A.K.; Robbie, A.J. Preparation and Characterization of Diarylphosphazene and Diarylphosphinohydrazide Complexes of Titanium, Tungsten and Ruthenium and Phosphorylketimido Complexes of Rhenium. Dalton Trans. 2005, 4, 680. [Google Scholar] [CrossRef] [PubMed]
  3. Saucedo Anaya, S.A.; Hagenbach, A.; Abram, U. Tricarbonylrhenium(I) and -Technetium(I) Complexes with Bis(2-Pyridyl)Phenylphosphine and Tris(2-Pyridyl)Phosphine. Polyhedron 2008, 27, 3587–3592. [Google Scholar] [CrossRef]
  4. Villafañe, F. Re I (CO)3 Complexes with Diimine Ligands Synthesized in Situ. Coord. Chem. Rev. 2017, 339, 128–137. [Google Scholar] [CrossRef]
  5. Edwards, P.G.; Newman, P.D.; Stasch, A. Coordination Chemistry of an Asymmetric P,N,O Tridentate Ligand Containing Primary Phosphine, Amine and Alcohol Donors. J. Organomet. Chem. 2011, 696, 1652–1658. [Google Scholar] [CrossRef]
  6. Correia, J.D.G.; Domingos, Â.; Santos, I.; Alberto, R.; Ortner, K. Re Tricarbonyl Complexes with Ligands Containing P,N,N and P,N,O Donor Atom Sets: Synthesis and Structural Characterization. Inorg. Chem. 2001, 40, 5147–5151. [Google Scholar] [CrossRef]
  7. Ranjan, S.; Lin, S.-Y.; Hwang, K.-C.; Chi, Y.; Ching, W.-L.; Liu, C.-S.; Tao, Y.-T.; Chien, C.-H.; Peng, S.-M.; Lee, G.-H. Realizing Green Phosphorescent Light-Emitting Materials from Rhenium(I) Pyrazolato Diimine Complexes. Inorg. Chem. 2003, 42, 1248–1255. [Google Scholar] [CrossRef]
  8. Yıldız, C.A.; Güney, E.; Nasif, V.; Karakaş, D.; Erkan, S. Investigation of Substituent Effect on Rhenium Complexes by DFT Methods: Structural Analysis, IR Spectrum, Quantum Chemical Parameter, NLO and OLED Properties, Molecular Docking. J. Mol. Struct. 2023, 1278, 134835. [Google Scholar] [CrossRef]
  9. Hu, Y.-X.; Zhao, G.-W.; Dong, Y.; Lü, Y.-L.; Li, X.; Zhang, D.-Y. New Rhenium(I) Complex with Thiadiazole-Annelated 1,10-Phenanthroline for Highly Efficient Phosphorescent OLEDs. Dyes Pigments 2017, 137, 569–575. [Google Scholar] [CrossRef]
  10. Gupta, D.; Sathiyendiran, M. Rhenium-Carbonyl-Based Supramolecular Coordination Complexes: Synthesis, Structure and Properties. ChemistrySelect 2018, 3, 7439–7458. [Google Scholar] [CrossRef]
  11. Luengo, A.; Fernández-Moreira, V.; Marzo, I.; Gimeno, M.C. Trackable Metallodrugs Combining Luminescent Re(I) and Bioactive Au(I) Fragments. Inorg. Chem. 2017, 56, 15159–15170. [Google Scholar] [CrossRef] [PubMed]
  12. Kamecka, A.; Prachnio, K.; Kapturkiewicz, A. The Luminescence Properties of [Re(CO)2(P^P)(N^N)]+ Complexes: Comparison with Their [Re(CO)3(N^N)(Cl)] Analogues. J. Lumin. 2018, 203, 409–419. [Google Scholar] [CrossRef]
  13. Ray, S.; Rajak, K.K. Rhenium(I) Complexes Incorporating Pyrene Bearing N, N Ligand: Luminescent Based Sensors for DNA. J. Organomet. Chem. 2024, 1009, 123077. [Google Scholar] [CrossRef]
  14. Veronese, L.; Quartapelle Procopio, E.; Moehl, T.; Panigati, M.; Nonomura, K.; Hagfeldt, A. Triarylamine-Based Hydrido-Carboxylate Rhenium(I) Complexes as Photosensitizers for Dye-Sensitized Solar Cells. Phys. Chem. Chem. Phys. 2019, 21, 7534–7543. [Google Scholar] [CrossRef]
  15. Takeda, H.; Koike, K.; Morimoto, T.; Inumaru, H.; Ishitani, O. Photochemistry and Photocatalysis of Rhenium(I) Diimine Complexes. In Advances in Inorganic Chemistry; Elsevier: Amsterdam, The Netherlands, 2011; Volume 63, pp. 137–186. ISBN 978-0-12-385904-4. [Google Scholar]
  16. Vlček, A.; Busby, M. Ultrafast Ligand-to-Ligand Electron and Energy Transfer in the Complexes Fac-[ReI(L)(CO)3(Bpy)]n+. Coord. Chem. Rev. 2006, 250, 1755–1762. [Google Scholar] [CrossRef]
  17. Kurtz, D.A.; Brereton, K.R.; Ruoff, K.P.; Tang, H.M.; Felton, G.A.N.; Miller, A.J.M.; Dempsey, J.L. Bathochromic Shifts in Rhenium Carbonyl Dyes Induced through Destabilization of Occupied Orbitals. Inorg. Chem. 2018, 57, 5389–5399. [Google Scholar] [CrossRef]
  18. Acosta, A.; Antipán, J.; Fernández, M.; Prado, G.; Sandoval-Altamirano, C.; Günther, G.; Gutiérrez-Urrutia, I.; Poblete-Castro, I.; Vega, A.; Pizarro, N. Photochemistry of P,N-Bidentate Rhenium(I) Tricarbonyl Complexes: Reactive Species Generation and Potential Application for Antibacterial Photodynamic Therapy. RSC Adv. 2021, 11, 31959–31966. [Google Scholar] [CrossRef]
  19. Pizarro, N.; Duque, M.; Chamorro, E.; Nonell, S.; Manzur, J.; De La Fuente, J.R.; Günther, G.; Cepeda-Plaza, M.; Vega, A. Dual Emission of a Novel (P,N) ReI Complex: A Computational and Experimental Study on [P,N-{(C6H5)2(C5H4N)P}Re(CO)3Br]. J. Phys. Chem. A 2015, 119, 3929–3935. [Google Scholar] [CrossRef]
  20. Nguyen, P.N.; Dao, T.-B.-N.; Tran, T.T.; Tran, N.-A.T.; Nguyen, T.A.; Phan, T.-D.L.; Nguyen, L.P.; Dang, V.Q.; Nguyen, T.M.; Dang, N.N. Electrocatalytic CO2 Reduction by [Re(CO)3Cl(3-(Pyridin-2-Yl)-5-Phenyl-1,2,4-Triazole)] and [Re(CO)3Cl(3-(2-Pyridyl)-1,2,4-Triazole)]. ACS Omega 2022, 7, 34089–34097. [Google Scholar] [CrossRef]
  21. Clark, M.L.; Cheung, P.L.; Lessio, M.; Carter, E.A.; Kubiak, C.P. Kinetic and Mechanistic Effects of Bipyridine (Bpy) Substituent, Labile Ligand, and Brønsted Acid on Electrocatalytic CO2 Reduction by Re(Bpy) Complexes. ACS Catal. 2018, 8, 2021–2029. [Google Scholar] [CrossRef]
  22. Willkomm, J.; Bouzidi, S.; Bertin, E.; Birss, V.I.; Piers, W.E. Aqueous CO2 Reduction by a Re(Bipyridine)-Polypyrrole Film Deposited on Colloid-Imprinted Carbon. ACS Catal. 2021, 11, 1096–1105. [Google Scholar] [CrossRef]
  23. Murata, K.; Tanaka, H.; Ishii, K. Electrochemical Reduction of CO2 by a Gas-Diffusion Electrode Composed of Fac-Re(Diimine)(CO)3Cl and Carbon Nanotubes. J. Phys. Chem. C 2019, 123, 12073–12080. [Google Scholar] [CrossRef]
  24. Nguyen, P.N.; Pham, H.P.; Dang, Q.V.; Pham, K.D.; Doan, G.N.; Ho, T.H.; Nguyen, T.M.; Dang, N.N. Low-Intensity-Visible-Light-Driven Photocatalytic CO2 Reduction by Rhenium Tricarbonyl Complexes Based on Pyridyl-Triazole Ligands. J. Organomet. Chem. 2025, 1023, 123438. [Google Scholar] [CrossRef]
  25. Baráth, E. Hydrogen Transfer Reactions of Carbonyls, Alkynes, and Alkenes with Noble Metals in the Presence of Alcohols/Ethers and Amines as Hydrogen Donors. Catalysts 2018, 8, 671. [Google Scholar] [CrossRef]
  26. Santana, C.G.; Krische, M.J. From Hydrogenation to Transfer Hydrogenation to Hydrogen Auto-Transfer in Enantioselective Metal-Catalyzed Carbonyl Reductive Coupling: Past, Present, and Future. ACS Catal. 2021, 11, 5572–5585. [Google Scholar] [CrossRef]
  27. Popovici, I.; Tannoux, T.; Bourcier, S.; Casaretto, N.; Auffrant, A. Coordination of NNN Iminophosphorane Ligands to MnII, CoII, and FeII and Application in the Transfer Hydrogenation of Ketones. Inorg. Chem. Commun. 2025, 180, 114917. [Google Scholar] [CrossRef]
  28. Namlı, M.; Işık, U.; Kantar, C.; Aydemir, M. Application of Phthalocyanine Complexes of Cu(II), Co(II), Ni(II), and Zn(II) as Catalysts in the Transfer Hydrogenation of Acetophenone and Its Derivatives. J. Mol. Struct. 2024, 1312, 138447. [Google Scholar] [CrossRef]
  29. Vermaak, V.; Vosloo, H.C.M.; Swarts, A.J. The Development and Application of Homogeneous Nickel Catalysts for Transfer Hydrogenation and Related Reactions. Coord. Chem. Rev. 2024, 507, 215716. [Google Scholar] [CrossRef]
  30. Venkatachalam, G.; Ramesh, R. Ruthenium(III) Bis-Bidentate Schiff Base Complexes Mediated Transfer Hydrogenation of Imines. Inorg. Chem. Commun. 2006, 9, 703–707. [Google Scholar] [CrossRef]
  31. Dayan, S.; Ozpozan, N.K.; Özdemir, N.; Dayan, O. Synthesis of Some Ruthenium(II)–Schiff Base Complexes Bearing Sulfonamide Fragment: New Catalysts for Transfer Hydrogenation of Ketones. J. Organomet. Chem. 2014, 770, 21–28. [Google Scholar] [CrossRef]
  32. Prakash, G.; Viswanathamurthi, P. New Ruthenium(II) Carbonyl Complexes Bearing Disulfide Schiff Base Ligands and Their Applications as Catalyst for Some Organic Transformations. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2014, 129, 352–358. [Google Scholar] [CrossRef]
  33. Singh, P.; Singh, A.K. Transfer Hydrogenation of Ketones and Catalytic Oxidation of Alcohols with Half-Sandwich Complexes of Ruthenium(II) Designed Using Benzene and Tridentate (S, N, E) Type Ligands (E = S, Se, Te). Organometallics 2010, 29, 6433–6442. [Google Scholar] [CrossRef]
  34. Moya, S.A.; Vidal, M.; Brown, K.; Negrete-Vergara, C.; Abarca, G.; Aguirre, P. Ruthenium Carbonyl Compounds Containing Polypyridine Ligands as Catalysts in the Reaction of N-Benzylideneaniline Hydrogenation. Inorg. Chem. Commun. 2012, 22, 146–148. [Google Scholar] [CrossRef]
  35. Moya, S.A.; Araya, J.C.; Gajardo, J.; Guerchais, V.; Le Bozec, H.; Toupet, L.; Aguirre, P. Synthesis and Characterization of Ruthenium (II) Carbonyl Complexes Containing Naphthyridine and Acetylacetonate Ligands and Their Catalytic Activity in the Hydrogen Transfer Reaction. Inorg. Chem. Commun. 2013, 27, 108–110. [Google Scholar] [CrossRef]
  36. Revathi, S.; Ghatak, T. Ethanol as Hydrogen Donor: An Efficient Transfer Hydrogenation of Aldehydes, Ketones, and Nitroarenes with H-Bonded Ru(II)-N-Heterocyclic Iminium Complex. J. Catal. 2024, 429, 115207. [Google Scholar] [CrossRef]
  37. Subaramanian, M.; Sivakumar, G.; Landge, V.G.; Kumar, R.; Natte, K.; Jagadeesh, R.V.; Balaraman, E. General and Selective Homogeneous Ru-Catalyzed Transfer Hydrogenation, Deuteration, and Methylation of Functional Compounds Using Methanol. J. Catal. 2023, 425, 386–405. [Google Scholar] [CrossRef]
  38. Liu, T.; Wang, L.; Wu, K.; Wang, Q.; Yu, Z. Mono- and Multinuclear Pincer-Type Ru(II) Complex Catalysts and Their Catalytic Applications. Inorg. Chim. Acta 2023, 551, 121458. [Google Scholar] [CrossRef]
  39. Prakash, O.; Singh, P.; Mukherjee, G.; Singh, A.K. Efficient Catalysis of Transfer Hydrogenation of Ketones and Oxidation of Alcohols with Newly Designed Half-Sandwich Rhodium(III) and Iridium(III) Complexes of Half-Pincer Chalcogenated Pyridines. Organometallics 2012, 31, 3379–3388. [Google Scholar] [CrossRef]
  40. Prakash, O.; Joshi, H.; Sharma, K.N.; Gupta, P.L.; Singh, A.K. Transfer Hydrogenation (pH Independent) of Ketones and Aldehydes in Water with Glycerol: Ru, Rh, and Ir Catalysts with a COOH Group near the Metal on a (Phenylthio)Methyl-2-Pyridine Scaffold. Organometallics 2014, 33, 3804–3812. [Google Scholar] [CrossRef]
  41. Ak, B.; Aydemir, M.; Durap, F.; Meriç, N.; Baysal, A. The First Application of C2-Symmetric Ferrocenyl Phosphinite Ligands for Rhodium-Catalyzed Asymmetric Transfer Hydrogenation of Various Ketones. Inorganica Chim. Acta 2015, 438, 42–51. [Google Scholar] [CrossRef]
  42. Archana, B.; Sinha, A. Unprecedented Homogeneous Transfer Hydrogenation Reaction of Aldehydes and Ketones by Μ-Mono Oxo Bridged Iron (III) Complex. Inorg. Chem. Commun. 2025, 180, 114891. [Google Scholar] [CrossRef]
  43. Işık, U.; Meriç, N.; Kayan, C.; Kılıç, A.; Belyankova, Y.; Zazybin, A.; Aydemir, M. Synthesis of Half-Sandwich Ruthenium(II) and Iridium(III) Complexes Containing Imidazole-Based Phosphinite Ligands and Their Use in Catalytic Transfer Hydrogenation of Acetophenone with Isopropanol. J. Organomet. Chem. 2023, 998, 122800. [Google Scholar] [CrossRef]
  44. Zhu, X.-H.; Cai, L.-H.; Wang, C.-X.; Wang, Y.-N.; Guo, X.-Q.; Hou, X.-F. Efficient and Versatile Transfer Hydrogenation Catalysts: Iridium (III) and Ruthenium (II) Complexes with 4-Acetylbenzyl-N-Heterocyclic Carbenes. J. Mol. Catal. Chem. 2014, 393, 134–141. [Google Scholar] [CrossRef]
  45. Smith, I.G.; Zgrabik, J.C.; Gutauskas, A.C.; Gray, D.L.; Domski, G.J. Synthesis, Characterization, and Catalytic Behavior of Mono- and Bimetallic Ruthenium(II) and Iridium(III) Complexes Supported by Pyridine-Functionalized N -Heterocyclic Carbene Ligands. Inorg. Chem. Commun. 2017, 81, 27–32. [Google Scholar] [CrossRef]
  46. Ak, B.; Aydemir, M.; Durap, F.; Meriç, N.; Elma, D.; Baysal, A. Highly Efficient Iridium Catalysts Based on C2-Symmetric Ferrocenyl Phosphinite Ligands for Asymmetric Transfer Hydrogenations of Aromatic Ketones. Tetrahedron Asymmetry 2015, 26, 1307–1313. [Google Scholar] [CrossRef]
  47. Landwehr, A.; Dudle, B.; Fox, T.; Blacque, O.; Berke, H. Bifunctional Rhenium Complexes for the Catalytic Transfer-Hydrogenation Reactions of Ketones and Imines. Chem.—Eur. J. 2012, 18, 5701–5714. [Google Scholar] [CrossRef]
  48. Zúñiga, C.; Oyarzún, D.P.; Martin-Transaco, R.; Yáñez-S, M.; Tello, A.; Fuentealba, M.; Cantero-López, P.; Arratia-Pérez, R. Synthesis, Characterization and Relativistic DFT Studies of fac-Re(CO)3(Isonicotinic Acid)2Cl Complex. Chem. Phys. Lett. 2017, 688, 66–73. [Google Scholar] [CrossRef]
  49. Portenkirchner, E.; Oppelt, K.; Ulbricht, C.; Egbe, D.A.M.; Neugebauer, H.; Knör, G.; Sariciftci, N.S. Electrocatalytic and Photocatalytic Reduction of Carbon Dioxide to Carbon Monoxide Using the Alkynyl-Substituted Rhenium(I) Complex (5,5′-Bisphenylethynyl-2,2′-Bipyridyl)Re(CO)3Cl. J. Organomet. Chem. 2012, 716, 19–25. [Google Scholar] [CrossRef]
  50. Frin, K.P.M.; Murakami Iha, N.Y. Modulation of Trans-to-Cis Photoisomerization and Photoluminescence of 1,2-Bis(4-Pyridyl)Ethylene or 4-Styrylpyridine Coordinated to fac-Tricarbonyl(5-Chloro-1,10-Phenathroline)Rhenium(I). Inorg. Chim. Acta 2011, 376, 531–537. [Google Scholar] [CrossRef]
  51. Sanhueza, L.; Barrera, M.; Crivelli, I. Experimental and Theoretical Studies of a New Donor–Acceptor Re(I) Complexes Using Nitropolypyridil Ligand. Analysis of the NLO Potential Response. Polyhedron 2013, 57, 94–104. [Google Scholar] [CrossRef]
  52. Lundin, N.J.; Walsh, P.J.; Howell, S.L.; Blackman, A.G.; Gordon, K.C. A Synthetic, Structural, Spectroscopic and DFT Study of ReI, CuI, RuII and IrIII Complexes Containing Functionalised Dipyrido [3,2-a:2′,3′-c]Phenazine (Dppz). Chem. Eur. J. 2008, 14, 11573–11583. [Google Scholar] [CrossRef]
  53. Zúñiga, C.; Crivelli, I.; Loeb, B. Synthesis, Characterization, Spectroscopic and Electrochemical Studies of Donor–Acceptor Ruthenium(II) Polypyridine Ligand Derivatives with Potential NLO Applications. Polyhedron 2015, 85, 511–518. [Google Scholar] [CrossRef]
  54. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A Complete Structure Solution, Refinement and Analysis Program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  55. Sheldrick, G.M. A Short History of SHELX. Acta Crystallogr. Sect. A 2008, 64, 112–122. [Google Scholar] [CrossRef] [PubMed]
  56. Mayberry, D.D.; Nesterov, V.N.; Richmond, M.G. Ambidentate Ligand Reactivity with the Rhenium(I) Compounds [BrRe(CO)4]2 and Cis-BrRe(CO)4 L: A Kinetic and Mechanistic Study. Eur. J. Inorg. Chem. 2017, 2017, 3990–3998. [Google Scholar] [CrossRef]
  57. McCusker, J.K.; Rheingold, A.L.; Hendrickson, D.N. Variable-Temperature Studies of Laser-Initiated 5T21A1 Intersystem Crossing in Spin-Crossover Complexes: Empirical Correlations between Activation Parameters and Ligand Structure in a Series of Polypyridyl Ferrous Complexes. Inorg. Chem. 1996, 35, 2100–2112. [Google Scholar]
  58. Pritchard, R.; Barrett, S.A.; Kilner, C.A.; Halcrow, M.A. The Influence of Ligand Conformation on the Thermal Spin Transitions in Iron(III) Saltrien Complexes. Dalton Trans. 2008, 24, 3159. [Google Scholar] [CrossRef]
  59. McKinnon, J.J.; Mitchell, A.S.; Spackman, M.A. Hirshfeld Surfaces: A New Tool for Visualising and Exploring Molecular Crystals. Chem.—Eur. J. 1998, 4, 2136–2141. [Google Scholar] [CrossRef]
  60. McKinnon, J.J.; Spackman, M.A.; Mitchell, A.S. Novel Tools for Visualizing and Exploring Intermolecular Interactions in Molecular Crystals. Acta Crystallogr. B 2004, 60, 627–668. [Google Scholar] [CrossRef]
  61. Spackman, M.A.; McKinnon, J.J. Fingerprinting Intermolecular Interactions in Molecular Crystals. CrystEngComm 2002, 4, 378–392. [Google Scholar] [CrossRef]
  62. Parra-Melipán, S.; López, V.; Moya, S.A.; Valdebenito, G.; Aranda, B.; Aguirre, P. Valorization of Furfural Using Ruthenium (II) Complexes Containing Phosphorus-Nitrogen Ligands under Homogeneous Transfer Hydrogen Condition. Mol. Catal. 2021, 513, 111729. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of complex fac-[Re(CO)3(PN)Cl].
Scheme 1. Synthesis of complex fac-[Re(CO)3(PN)Cl].
Inorganics 13 00338 sch001
Figure 1. ORTEP of the fac-[Re(CO)3(PN)Cl] complex. The thermal ellipsoids were drawn with a 30% probability. All the hydrogen atoms are omitted for the sake of clarity.
Figure 1. ORTEP of the fac-[Re(CO)3(PN)Cl] complex. The thermal ellipsoids were drawn with a 30% probability. All the hydrogen atoms are omitted for the sake of clarity.
Inorganics 13 00338 g001
Figure 2. Hirshfeld surface mapped with dnorm over molecular structure of fac-[Re(CO)3(PN)Cl] complex.
Figure 2. Hirshfeld surface mapped with dnorm over molecular structure of fac-[Re(CO)3(PN)Cl] complex.
Inorganics 13 00338 g002
Figure 3. The intermolecular hydrogen bonds of fac-[Re(CO)3(PN)Cl] crystal structure. Symmetry codes: i (+x, 3/2 − y, −1/2 + z); ii (2 − x, 1 − y, 1 − z).
Figure 3. The intermolecular hydrogen bonds of fac-[Re(CO)3(PN)Cl] crystal structure. Symmetry codes: i (+x, 3/2 − y, −1/2 + z); ii (2 − x, 1 − y, 1 − z).
Inorganics 13 00338 g003
Scheme 2. Hydrogenation of acetophenone by a rhenium(I) complex.
Scheme 2. Hydrogenation of acetophenone by a rhenium(I) complex.
Inorganics 13 00338 sch002
Table 1. Crystal data and structure refinement for fac-[Re(CO)3(PN)Cl] complex.
Table 1. Crystal data and structure refinement for fac-[Re(CO)3(PN)Cl] complex.
CompoundFac-[Re(CO)3(PN)Cl]
Empirical formulaC20H14ClNO3PRe
Formula weight568.94
Temperature/K298
Crystal systemmonoclinic
Space groupP21/c
a/Å10.3912 (14)
b/Å12.4141 (16)
c/Å16.222 (2)
α/°90
β/°94.930 (5)
γ/°90
Volume/Å32084.9 (5)
Z4
ρcalcg/cm31.813
μ/mm−16.052
F (000)1088.0
Crystal size/mm30.3 × 0.256 × 0.152
RadiationMoKα (λ = 0.71073)
2Θ range for data collection/°5.04 to 52.864
Index ranges−13 ≤ h ≤ 12, −15 ≤ k ≤ 15, −20 ≤ l ≤ 20
Reflections collected37,744
Independent reflections4274 [Rint = 0.0272, Rsigma = 0.0134]
Data/restraints/parameters4274/15/239
Goodness-of-fit on F21.085
Final R indexes [I >= 2σ (I)]R1 = 0.0165, wR2 = 0.0338
Final R indexes [all data]R1 = 0.0213, wR2 = 0.0354
Largest diff. peak/hole/e Å−30.58/−0.67
Table 2. Bond distances and angles of the fac-[Re(CO)3(PN)Cl] complex.
Table 2. Bond distances and angles of the fac-[Re(CO)3(PN)Cl] complex.
Bond Distances (Å)
Re1-Cl12.4817 (7)
Re1-P12.4800 (7)
Re1-N12.201 (2)
Re1-C11.911 (3)
Re1-C21.930 (3)
Re1-C31.907 (3)
Bond Angles (°)
P1-Re1-Cl184.43 (2)
N1-Re1-Cl183.72 (5)
N1-Re1-P165.13 (5)
C1-Re1-Cl192.82 (9)
C1-Re1-P1104.54 (9)
C1-Re1-N1169.32 (10)
C1-Re1-C290.16 (15)
C2-Re1-Cl193.82 (10)
C2-Re1-P1165.25 (12)
C2-Re1-N1100.13 (13)
C3-Re1-Cl1176.93 (9)
C3-Re1-P193.52 (9)
C3-Re1-N193.35 (10)
C3-Re1-C189.91 (12)
C3-Re1-C287.59 (13)
Table 3. Intermolecular hydrogen-bonding interaction parameters for compound fac-[Re(CO)3(PN)Cl].
Table 3. Intermolecular hydrogen-bonding interaction parameters for compound fac-[Re(CO)3(PN)Cl].
D-H⋯Ad(D-H)/Åd(H-A)/Åd(D-A)/ÅD-H-A/°
C7 i-H7 i⋯O10.932.553.347 (4)143.7
C6 ii-H6 ii⋯Cl10.932.873.794 (3)170.4
C5 ii-H5 ii⋯Cl10.932.873.528 (4)128.6
Symmetry codes: i (+x, 3/2 − y,−1/2 + z); ii (2 − x,1 − y,1 − z).
Table 4. Transfer hydrogenation of acetophenone catalyzed by Re(I) complexes.
Table 4. Transfer hydrogenation of acetophenone catalyzed by Re(I) complexes.
% Conversion (TON, TOF h−1)
EntryComplex1 h2 h3 h4 h7 h
1[Re(CO)5Cl]3 (30, 30)7 (70, 35)8 (80, 27)8 (80, 27)8 (80, 27)
2[Re(CO)3(PN)Cl]15 (150,150)37 (370, 185)65 (650, 217)80 (800, 200)99 (990, 141)
3[Re(CO)3(PyCOOH)2Cl]24 (240,240)44 (440, 220)56 (560, 187)65 (650, 163)70 (700, 100)
4[Re(CO)3(bpy)Cl]33 (330, 330)53 (530, 265)70 (700, 233)88 (880, 220)90 (900, 129)
5[Re(CO)3(Phen)Cl]30 (300, 300)47 (470, 235)55 (550, 183)68 (680, 170)71 (710, 101)
6[Re(CO)3(Phen-NO2)Cl]36 (360, 360)52 (520, 260)63(630, 210)66 (660, 165)68 (680, 97)
7[Re(CO)3(dppz-NO2)Cl]34 (340, 340)65 (650, 325)71(710, 240)68 (680, 170)69 (690, 98)
8[Re(CO)3(dppz-CN)Cl]29 (290, 290)42 (420, 210)56 (560, 187)62 (620, 155)68 (680, 97)
Condition: Substrate/catalyst = 1000/1; Catalyst = 0.01 mmol; NaOH/metal = 10/1; solvent = 2-propanol; temperature = Reflux.
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

Zúñiga, C.; Fuentealba, M.; Olave, E.; Oyarzún, D.P.; Aracena, A.; Yañez-S, M.; Cantero-López, P.; Aguirre, P.A. Catalytic Activity of Rhenium(I) Tricarbonyl Complexes Containing Polypyridine and Phosphorus–Nitrogen Ligands in the Hydrogen Transfer of Acetophenone. Inorganics 2025, 13, 338. https://doi.org/10.3390/inorganics13100338

AMA Style

Zúñiga C, Fuentealba M, Olave E, Oyarzún DP, Aracena A, Yañez-S M, Cantero-López P, Aguirre PA. Catalytic Activity of Rhenium(I) Tricarbonyl Complexes Containing Polypyridine and Phosphorus–Nitrogen Ligands in the Hydrogen Transfer of Acetophenone. Inorganics. 2025; 13(10):338. https://doi.org/10.3390/inorganics13100338

Chicago/Turabian Style

Zúñiga, César, Mauricio Fuentealba, Elizabeth Olave, Diego P. Oyarzún, Andrés Aracena, Mauricio Yañez-S, Plinio Cantero-López, and Pedro A. Aguirre. 2025. "Catalytic Activity of Rhenium(I) Tricarbonyl Complexes Containing Polypyridine and Phosphorus–Nitrogen Ligands in the Hydrogen Transfer of Acetophenone" Inorganics 13, no. 10: 338. https://doi.org/10.3390/inorganics13100338

APA Style

Zúñiga, C., Fuentealba, M., Olave, E., Oyarzún, D. P., Aracena, A., Yañez-S, M., Cantero-López, P., & Aguirre, P. A. (2025). Catalytic Activity of Rhenium(I) Tricarbonyl Complexes Containing Polypyridine and Phosphorus–Nitrogen Ligands in the Hydrogen Transfer of Acetophenone. Inorganics, 13(10), 338. https://doi.org/10.3390/inorganics13100338

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

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop