Next Article in Journal
Selective Catalytic Oxidation of Benzyl Alcohol by MoO2 Nanoparticles
Previous Article in Journal
The Effects of the Crystalline Phase of Zirconia on C–O Activation and C–C Coupling in Converting Syngas into Aromatics
Previous Article in Special Issue
Effects of Substitution Pattern in Phosphite Ligands Used in Rhodium-Catalyzed Hydroformylation on Reactivity and Hydrolysis Stability
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Reactivity of Poly(propyleneimine) Dendrimers Functionalized with Cyclopentadienone N-Heterocyclic-Carbene Ruthenium(0) Complexes

1
Dipartimento di Chimica Industriale “Toso Montanari”, Università degli Studi di Bologna, viale Risorgimento 4, 40136 Bologna, Italy
2
Present address: Department of Chemistry and Applied Biosciences, Eidgenössische Technische Hochschule (ETH) Zürich, Vladimir-Prelog Weg, 1-5/10 8093 Zürich, Switzerland
3
Dipartimento di Scienze Chimiche e Geologiche, Università degli Studi di Modena e Reggio Emilia, via G. Campi 103, 41125 Modena, Italy
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(2), 264; https://doi.org/10.3390/catal10020264
Submission received: 23 January 2020 / Revised: 18 February 2020 / Accepted: 19 February 2020 / Published: 22 February 2020
(This article belongs to the Special Issue Ligand Design in Metal Chemistry: Reactivity and Catalysis)

Abstract

:
Ligand design in metal chemistry is a fundamental step when pursuing compounds with specific reactivity. In this paper, the functionalization of the OH group in the lateral chain of the N-heterocyclic-carbene (NHC) ligand bound to a bis-carbonyl cyclopentadienone NHC ruthenium(0) complex allowed the decoration of five generations of poly(propyleneimine) (PPIs) dendrimers with up to 64 organometallic moieties. The coupling was achieved by employing carbonyldiimidazole and the formation of carbamate linkages between dendritic peripheral NH2 and lateral OH groups on ruthenium complexes. The synthetic procedure, chemical purification, and spectroscopic characterization of the five generations of dendrimers (3g15) are here described. The ruthenium-modified dendrimers were activated as catalysts in the transfer hydrogenation of the model compound 4-fluoroacetophenone in the presence of cerium ammonium nitrate as their mononuclear congeners. The catalytic activity, being similar for the five generations, shows a decrease if compared to mononuclear complexes. This detrimental effect might be ascribed to the –CH2NH– functionalization, largely present in dendrimer skeleton and that can compete with the hydrogen transfer mechanism, but also partially to a dendritic effect caused by steric encumbrance.

Graphical Abstract

1. Introduction

Dendritic catalysts are functional macromolecules with precise and unambiguous structures, which, thanks to their monodisperse nature, maintain the advantages of homogeneous catalysts showing fast kinetic responses and easy tunability [1,2,3,4,5,6,7,8,9,10,11,12]. Furthermore, they can be removed from the reaction mixture by membrane or nanofiltration techniques and precipitation, exploiting their bigger sizes compared to products, and this confers the advantages of heterogeneous catalysts [13]. Catalytic sites grafted on the periphery of dendrimers can give rise to active, multivalent species that might result in high reaction rates. Several reviews report on the use of dendrimers in catalysis [1,2,3,4,5,6,7,8,9,10,11], but few of them are specifically dedicated to the dendrimer effect [14,15,16]. The consequences of dendritic effects are known to be like substrate activation and influence the reaction rate and selectivity. On one hand, a dendrimer brings together a large number of catalytically active species in a nanoobject, while, on the other hand, limitation to the access of substrate molecules can lead to a negative effect, i.e., lowering of reaction rates caused by bulk hindrance.
The majority of the reported dendritic catalysts for hydrogenation contain phosphines since they have proved to be optimal ligands for grafting metal complexes, thanks to their ability to firmly coordinate metal ions and lead to high catalytic performances [17,18,19,20,21,22,23]. Both negative [24] and positive [25,26,27] dendritic effects have been observed when hydrogenation is carried out with increasing sizes of dendrimers. On the contrary, less examples regarding transfer hydrogenation are reported [13,28,29]. Worthy of note is the tetrabranched phosphoranyl-terminated carbosilane ruthenium(II) derivative able to catalyze the transfer hydrogenation of cyclohexanone, where cyclohexadiene or formic acid acts as stoichiometric hydrogen donor species. This first-generation dendrimer, which contains the lowest number of peripheral ruthenium functionalities, namely four, was found to be less active than the mononuclear complex [28]. There are also examples in which the dendritic system is given by the surrounding part of the catalytic center [30,31], as in the case of the “green” application of the fluorinated dendritic chiral mono-N-tosylated-1,2-diphenylethylenediamine (FTsDPEN), which was employed in the asymmetric transfer hydrogenation of prochiral ketones in aqueous media catalyzed by ruthenium(II) centers [32].
Within the field of transfer hydrogenation, our group has recently developed novel ruthenium compounds for bifunctional catalysis [33,34] by combining cyclopentadienones, which cooperate with the metal centers in catalytic redox reactions [35,36,37,38], and N-heterocyclic carbenes (NHCs) [39,40], largely used and very versatile ligands due to their easy preparation and the tunability of their steric and electronic properties [41,42,43,44]. In particular, the reduction of the model compound 4-fluoroacetophenone catalyzed by complexes 1ac (Figure 1) can be activated by addition of cerium ammonium nitrate (CAN), which favors the release of CO [34].
NHCs are effective tools for the rational design of transition metal catalysts [45,46,47,48,49,50,51,52] in that their skeleton can be provided with suitable substituents for heterogenization (e.g., –OH in 1c, Figure 1). Despite the fact that NHCs can be considered ancillary ligands as outstanding as phosphines, examples of NHC-decorated dendrimers reported in the literature are still contained [53,54,55,56]. These include the NHC-rhodium dendrimer showing a positive dendrimer effect in hydrosilylation of ketones [54], and an NHC-ruthenium derivative employed in olefin metathesis, in which the catalytic center is bonded to the dendrimer through a Ru=C double bond involving the aromatic moiety without exploiting the ancillary NHC ligand [53]. In order to evaluate a possible dendrimer effect and inspired by previous works on dendritic systems [57], we report here on the development of an efficient method for the straightforward synthesis, chemical purification, and spectroscopic characterization of five generations of poly(propyleneimine) (PPIs) NHC-ruthenium(0) functionalized dendrimers with up to 64 organometallic moieties. Their catalytic activity in transfer hydrogenation has been tested for the reduction of 4-fluoroacetophenone in iPrOH, and the results obtained will be compared with the ones previously obtained for the mononuclear ruthenium complexes (Table 1) [34].

2. Results

2.1. Synthesis and Characterization of the Ruthenium-Decorated Dendrimers

The new ruthenium(0) complex 2 with CO2-imidazole (CO2Im) functionalized NHC ligand was prepared in 84% yield by reacting 1c with a slight excess of carbonyldiimidazole (CDI:1c = 1.1:1) in CH2Cl2 at room temperature for 2 h, as shown in Scheme 1. The compound 2 was then easily purified by washing the imidazolium co-product with water.
The off-white solid 2 is air stable and was characterized by analytical measurements (Figures S1–S4 in the Supplementary Materials, SM). The 1H and 13C NMR spectra were registered in CDCl3 and they are comparable to the ones of the initial complex 1c for the unaltered structural parts [34]. In particular, the downfield-shifted resonance of the coordinated carbon atom of the carbene (δ = 174.80 ppm) appeared in the 13C NMR spectrum. In addition, the imidazole moiety displays signals at chemical shifts (1H NMR: δ = 8.03, 7.33, 7.08 ppm; 13C NMR: δ = 136.95, 130.88, 116.97 ppm) close to those of the starting material CDI, while the C=O group resonates at 147.89 ppm, typical for a carbamate moiety (–OC(O)N–). In the infrared spectrum in CH2Cl2, CO stretching bands (νCO = 2007 and 1948 cm–1) are consistent with those of 1c, and the stretching at 1768 cm–1 is ascribable to the carbamate group formed during the reaction. Further evidences of the occurred functionalization reaction were given by Electron-Spray Ionization mass spectrometry (MS-ESI) measurements, where the protonated molecular ion [M + H]+ at m/z 823 could be detected.
The reaction in CH2Cl2 at room temperature for 2 h of a 4.8 fold excess of 2 with the dendrimer DAB-dendr-(NH2)4 (g1) led to the formation of the dendrimer 3g1 decorated with four ruthenium(0) moieties (Scheme 2). The CO2Im group of 2 easily reacted with DAB-dendr-(NH2)n (n = 8, 16, 32, 64, gn, see Scheme S1 in the SM), up to the fifth generation, under the same experimental conditions yielding the fully functionalized new dendritic organometallic dendrimers 3gn (Figure 2) in acceptable yields. The excess of 2 and the side products could be easily removed from the crude product by extraction with water and washing with Et2O. All five generation macromolecules 3gn were isolated as light brown solids and were revealed to be stable to air and moisture.
The dendritic compounds are soluble in THF and CH2Cl2, sparingly soluble in Et2O, and totally insoluble in hexane, toluene as well as aqueous solvents. They were characterized by 1H and 13C NMR, IR spectroscopy and mass spectrometry, when feasible (Figures S5–S21 in the SM). Considering the NMR spectra registered in CDCl3, they show the signals due to the diaminobutane-based PPIs skeleton (1H: four broad multiplets in the range 3.8–1.3 ppm; 13C: four resonances in the range 63–25 ppm) and those given by the anchored peripheral NHC ruthenium moieties (see Materials and Methods section). Even if all the proton signals appear as broad peaks typical of a polymer-like structure, it is clearly detectable that the resonances of the two methylene groups belonging to the –NHCCH2CH2OC(O)NH– unit of the organometallic moieties are at 3.77 and 3.59 ppm, which are more shielded than those of 2 (4.05 and 3.99 ppm in –NHCCH2CH2OC(O)Im). The signal of the NH proton of the carbamate linkers close to the surface of the PPIs dendrimer moves along the whole spectrum and it is often not clearly detected due to either that it can resonate in the same region of aromatic signals or that it could be engaged in hydrogen bonds like –NH⋯(O)CO– or –NH⋯cyclopentadienone, fact that becomes more relevant upon increasing the dendritic generation where the interacting groups get in closer contact. The IR spectra of all five dendritic species in CH2Cl2 solution show a strong absorption at 1718 cm−1 attributable to both C=O and –O2CNH– stretching bands. The structure of the first generation dendrimer 3g1 (Scheme 2) could be also confirmed by MS-ESI, which shows the molecular ion with sodium [M + Na]+ at m/z 3355. The high molecular weights of the upper generations 3g25 prevented the use of MS-ESI technique for mass detection.

2.2. Catalytic Transfer Hydrogenation

The five generations of functionalized dendrimers 3g15 were tested as precursors under catalytic transfer hydrogenation conditions employing 4-fluoroacetophenone as model substrate in refluxing iPrOH (hydrogen source). Catalytic runs were performed in order to investigate the stability of the peripheral organometallic moieties on the PPIs dendrimers and to detect any possible dendritic effect. Results obtained are reported in Table 2, and in all cases, selectivity is complete and conversion corresponds to yield.
3g15 did not show any catalytic activity in the absence of additives, as already observed in the case of the mononuclear congeners [34]. Addition of 1 equivalent of CAN, with respect to Ru loading, promotes CO release [34,35] and leads to the pre-catalyst activation [58] resulting in some catalytic activity. Nevertheless, only 15–17% conversion could be reached after 24 h for all the five generations, which allow us to discard any positive dendritic effect. These conversions are lower than those obtained with the mononuclear alkyl-substituted NHC complexes 1a and 1b but similar to 1c (see Table 1). This peculiar behavior could be then ascribed to –CH2NH– dendrimer functionalization, which might compete with the hydrogen transfer mechanism as indeed depicted previously for the –CH2OH group in 1c [34].
A similar outcome was found by adding 10 eq. of pyridine for each ruthenium(0) center to 3g1 (entry 6 of Table 2), confirming the activation effect of the latter additive stated for pyridyl functionalized NHC ligands [58]. Although the pre-catalyst activation could be achieved also by pyridine, this was not used to test the following generations of dendrimer due to the high amount of additive required compared to CAN and the limited enhancement in the conversion (20% vs 17%).

3. Materials and Methods

3.1. General Information

Solvents dichloromethane (CH2Cl2), diethyl ether (Et2O), acetonitrile (CH3CN) and toluene were dried and distilled prior to use. Acetone was degassed and stored under inert atmosphere on molecular sieves. The other solvents iPrOH, heptane, hexane, CDCl3, D2O and toluene-d8 (Sigma Aldrich) were employed without further purification. DAB-dendr-(NH2)n g15 (n = 4, 8, 16, 32 and 64), 4-fluoroacetophenone, CDI and CAN (Merck KGaA, Darmstadt, Germany: Sigma-Aldrich Products sold through Sigma-Aldrich, Italy) were employed as purchased. 1c [39] was prepared as previously reported. The NMR spectra were recorded using Varian Inova 300 (Palo Alto, California - US) (1H: 300.1 MHz, 13C: 75.5 MHz), Varian Mercury Plus VX 400 (Oxford, UK) (1H: 399.9 MHz; 13C: 100.6 MHz, 19F: 376.5 MHz), Varian Inova 600 (Oxford, UK) (1H: 599.7 MHz, 13C: 150.8 MHz) spectrometers at 298 K; chemical shifts were referenced internally to residual solvent peaks. Full 1H and 13C NMR assignments were done, when necessary, by gHSQC and gHMBC NMR experiments using standard Varian pulse sequences. Infrared spectra were recorded at 298 K on a Perkin-Elmer (Llantrisant, UK) Spectrum Two spectrophotometer. MS-ESI+ spectra were recorded with Waters (Milford, MA, USA) Micromass ZQ 4000 spectrometer with samples dissolved in MeOH or CH3CN. Elemental analyses were performed with the Thermo Scientific Flash 2000 with CHNS analyser (Bremen, Germany) instrument.

3.2. Synthesis of Dicarbonyl(η4-3,4-bis(4-methoxyphenyl)-2,5-diphenylcyclopenta-2,4-dienone) (1-methyl-3-(2-CO2Im-ethyl)imidazol-ylidene)ruthenium (2)

1c (1.44 g, 1.98 mmol) and CDI (0.97 g, 5.99 mmol) were dissolved in anhydrous CH2Cl2 under nitrogen atmosphere and the reaction mixture was stirred at room temperature for 2 h. The solution was then extracted with water (3 × 20 mL), and traces of water were removed from organic phase with Na2SO4. The solvent was then removed under reduced pressure and the white/brown product, identified as 2, was obtained with an 84% yield. 1H NMR (399.9 MHz, 298 K, CDCl3): δ 8.03 (s, 1H, Im); 7.78 (m, 4H, CHaryl), 7.33 (m, 1H, Im), 7.16–7.12 (m, 8H, CHaryl), 7.08 (m, 1H, Im), 7.06 (m, 2H, CHaryl), 6.96 (d, J = 2.0 Hz, 1H, CHNHC), 6.85 (d, J = 2.0 Hz, 1H, CHNHC), 6.65 (m, 4H, CHaryl), 4.05 (m, 2H, –CH2–), 3.99 (m, 2H, –CH2–), 3.72 (s, 6H, –OCH3), 3.14 (s, 3H, –NCH3) ppm. 13C NMR (100.6 MHz, 298 K, CDCl3): δ 202.15 (CO), 174.80 (Ccarbene), 169.41 (C=O, Cp), 158.64 (–COCH3), 147.89 (–OC(O)N–), 136.95 (CH, Im), 134.91 (Cqaryl), 133.55 (CHaryl), 130.88 (CH, Im), 129.36 (CHaryl), 127.63 (CHaryl), 125.63 (Cqaryl), 124.70 (CHNHC), 124.37 (CHaryl), 121.76 (CHNHC), 116.97 (CH, Im), 112.96 (CHaryl), 104.09 (C2,5, Cp), 78.88 (C3,4, Cp), 67.02 (–CH2–), 55.00 (–OCH3), 49.20 (–CH2–), 38.41 (CH3, NHC). IR (CH2Cl2): ν(carbonyl CO): 2007 and 1948, ν(C=O, –O2CIm): 1768 cm–1. MS-ESI+ (MeOH): m/z 823 [M + H]+, 845 [M + Na]+. Anal. Calcd (%) for C43H36N4O7Ru: C, 62, 84; H, 4, 42; N, 6, 82. Found: C, 62, 81; H, 4, 43; N, 6, 83. 1H and 13C NMR, IR and MS-ESI spectra of 2 are reported in Figures S1–S4 in the SM.

3.3. Synthesis of Functionalized Dendrimers 3g1–5

g1 (0.021 g, 0.06 mmol) and 2 (0.260 g, 0.31 mmol) were dissolved in anhydrous CH2Cl2 (10 mL) under inert atmosphere and the reaction mixture was stirred at room temperature for two days. The solution was then extracted with water (3 × 20 mL) and traces of water were removed from organic phase with MgSO4. Organic solvent was removed under reduced pressure and the solid crude was washed with Et2O (3 × 15 mL). The final yellow solid was identified as 3g1 (n = 4) and obtained with 42% yield. 1H NMR (399.9 MHz, 298 K, CDCl3): δ 7.72–6.60 (CHaryl), 6.81 (m, 4H, CHNHC), 6.62 (m, 4H, CHNHC), 3.77 (m, 8H, –CH2–), 3.68 (s, 24H, –OCH3), 3.59 (m, 8H, –CH2–), 3.09 (s, 12H, –NCH3), 3.07, 2.97, 2.36, 1.57, 1.36 (br, 32H, CH2, DAB-dendr) ppm. 13C NMR (100.6 MHz, 298 K, CDCl3): δ 202.06 (C=O), 172.67 (Ccarbene), 169.56 (C=O, Cp), 158.53 (–COCH3), 156.00 (–OC(O)N–), 135.04 (Cqaryl), 133.55 (CHaryl), 129.40 (CHaryl), 127.45 (CHaryl), 125.50 (Cqaryl), 124.44 (CHaryl), 124.36 (CHNHC), 121.80 (CHNHC), 112.85 (CHaryl), 103.73 (C2,5, Cp), 78.97 (C3,4, Cp), 63.29 (–CH2–), 54.96 (–OCH3), 49.75 (–CH2–), 38.25 (CH3, NHC), 53.73, 51.65, 39.51, 26.99, 24.88 (CH2, DAB-dendr) ppm. IR (CH2Cl2): ν(carbonyl CO): 2006 and 1947, ν(C=O, –NHC(O)O–): 1718, ν(C=C): 1601 and 1518, ν(C=O, Cp): 1582 cm–1. MS-ESI+ (CH3CN): m/z 3355 [M + Na]+. Anal. Calcd (%) for C176H168N14O28Ru4: C, 63, 45; H, 5, 08; N, 5, 89. Found: C, 63, 46; H, 5, 08; N, 5, 88. 1H and 13C NMR, IR and MS-ESI spectra of 3g1 are reported in Figures S5–S9 in the SM.
The upper generations of dendrimers, peripherally decorated with n ruthenium(0) moieties, 3g25 with n = 8, 16, 32 and 64, respectively, were prepared by applying the same procedure, but they required seven days of reaction time. 1H and 13C NMR spectra are reported in Figures S10–S21 in the SM.
3g2, prepared from g2 (0.020 g, 0.03 mmol) and 2 (0.26 g, 0.31 mmol), yield of 25%. 1H NMR (399.9 MHz, 298 K, CDCl3): δ 7.75–6.61 (CHaryl), 6.96 (m, 8H, CHNHC), 6.79 (m, 8H, CHNHC), 3.77 (m, 16H, –CH2–), 3.70 (s, 48H, –OCH3), 3.59 (m, 16H, –CH2–), 3.10 (s, 24H, –NCH3), 3.20–2.80, 2.60–2.20, 1.96–1.42 (br, 80H, CH2, DAB-dendr) ppm. 13C NMR (100.6 MHz, 298 K, CDCl3): δ 202.18 (C=O), 173.96 (Ccarbene), 169.52 (C=O, Cp), 158.61 (–COCH3), 156.14 (–OC(O)N–), 135.00 (Cqaryl), 133.58 (CHaryl), 129.49 (CHaryl), 127.59 (CHaryl), 125.78 (Cqaryl), 124.55 (CHaryl), 124.30 (CHNHC), 122.43 (CHNHC), 112.95 (CHaryl), 103.65 (C2,5, Cp), 78.68 (C3,4, Cp), 63.29 (–CH2–), 55.02 (–OCH3), 50.00 (–CH2–), 38.31 (CH3, NHC), 53.47, 51.54, 39.51, 26.84, 23.63 (CH2, DAB-dendr) ppm. IR (CH2Cl2): ν(carbonyl CO): 2008 and 1948; ν(C=O, –NHC(O)O–): 1719; ν(C=C): 1603 and 1518; ν(C=O, Cp): 1577 cm–1. Anal. Calcd (%) for C360H352N30O56Ru8: C, 63, 55; H, 5, 21; N, 6, 18. Found: C, 63, 54; H, 5, 19; N, 6, 19.
3g3, prepared from g3 (0.021 g, 0.01 mmol) and 3 (0.17 g, 0.21 mmol), yield of 32%. 1H NMR (399.9 MHz, 298 K, CDCl3): δ 7.76–6.61 (CHaryl), 6.96 (m, 16H, CHNHC), 6.78 (m, 16H, CHNHC), 3.77 (m, 32H, –CH2–), 3.71 (s, 96H, –OCH3), 3.67 (m, 32H, –CH2–), 3.10 (s, 48H, –NCH3), 3.20–1.50 (br, 176H, CH2, DAB-dendr) ppm. 13C NMR (100.6 MHz, 298 K, CDCl3): δ 202.17 (C=O), 173.94 (Ccarbene), 169.46 (C=O, Cp), 158.60 (–COCH3), 153.57 (–OC(O)N–), 135.00 (Cqaryl), 133.57 (CHaryl), 129.29 (CHaryl), 127.59 (CHaryl), 125.52 (Cqaryl), 124.54 (CHaryl), 124.29 (CHNHC), 122.38 (CHNHC), 112.95 (CHaryl), 103.98 (C2,5, Cp), 78.67 (C3,4, Cp), 63.69 (–CH2–), 55.01 (–OCH3), 50.01 (–CH2–), 38.30 (CH3, NHC), 53.39, 51.39, 38.41, 26.78, 24.02 (CH2, DAB-dendr) ppm. IR (CH2Cl2): ν(carbonyl CO): 2008 and 1949, ν(C=O, –NHC(O)O–): 1718, ν(C=C): 1602 and 1517, ν(C=O, Cp): 1577 cm–1. Anal. Calcd (%) for C728H720N62O112Ru16: C, 63, 61; H, 5, 28; N, 6, 32. Found: C, 63, 60; H, 5, 30; N, 6, 33.
3g4, prepared from g4 (0.020 g, 6.00 µmol) and 3 (0.20 g, 0.25 mmol), yield of 41%. 1H NMR (399.9 MHz, 298 K, CDCl3): δ 7.75–6.60 (CHaryl), 6.96 (m, 32H, CHNHC), 6.79 (m, 32H, CHNHC), 3.77 (m, 64H, –CH2–), 3.71 (s, 192H, –OCH3), 3.67 (m, 64H, –CH2–), 3.10 (s, 96H, –NCH3), 3.20–1.50 (br, 368H, CH2, DAB-dendr) ppm. 13C NMR (100.6 MHz, 298 K, CDCl3): δ 202.16 (C=O), 173.88 (Ccarbene), 169.38 (C=O, Cp), 158.60 (–COCH3), 153.58 (–OC(O)N–), 134.97 (Cqaryl), 133.57 (CHaryl), 129.31 (CHaryl), 127.59 (CHaryl), 125.54 (Cqaryl), 124.53 (CHaryl), 124.30 (CHNHC), 122.38 (CHNHC), 112.95 (CHaryl), 103.96 (C2,5, Cp), 78.75 (C3,4, Cp), 63.68 (–CH2–), 55.02 (–OCH3), 49.99 (–CH2–), 38.32 (CH3, NHC), 52.47, 51.22, 38.43, 24.10, 19.73 (CH2, DAB-dendr) ppm. IR (CH2Cl2): ν(carbonyl CO): 2009 and 1947, ν(C=O, –NHC(O)O–): 1718, ν(C=C): 1603 and 1518, ν(C=O, Cp): 1576 cm–1. Anal. Calcd (%) for C1464H1456N126O224Ru32: C, 63, 63; H, 5, 31; N, 6, 39. Found: C, 63, 61; H, 5, 31 N, 6, 39.
3g5, prepared from g5 (0.021 g, 3.00 µmol) and 3 (0.20 g, 0.25 mmol), yield of 43%. 1H NMR (399.9 MHz, 298 K, CDCl3): δ 7.75–6.58 (CHaryl), 6.96 (m, 64H, CHNHC), 6.78 (m, 64H, CHNHC), 3.77 (m, 128H, –CH2–), 3.71 (s, 384H, –OCH3), 3.66 (m, 128H, –CH2–), 3.10 (s, 192H, –NCH3), 3.20–1.50 (br, 752H, CH2, DAB-dendr). 13C NMR (100.6 MHz, 298 K, CDCl3): δ 202.16 (C=O), 173.91 (Ccarbene), 169.43 (C=O, Cp), 158.60 (–COCH3), 153.57 (–OC(O)N–), 134.98 (Cqaryl), 133.57 (CHaryl), 129.30 (CHaryl), 127.59 (CHaryl), 125.53 (Cqaryl), 124.54 (CHaryl), 124.30 (CHNHC), 122.38 (CHNHC), 112.95 (CHaryl), 103.97 (C2,5, Cp), 78.70 (C3,4, Cp), 63.80 (–CH2–), 55.02 (–OCH3), 49.92 (–CH2–), 38.31 (CH3, NHC), 52.46, 51.35, 38.39, 26.70, 24.09 (CH2, DAB-dendr). IR (CH2Cl2): ν(carbonyl CO): 2006 and 1950, ν(C=O, –NHC(O)O–): 1717, ν(C=C): 1602 and 1518, ν(C=O, Cp): 1577 cm–1. Anal. Calcd (%) for C2936H2928N254O448Ru64: C, 63, 64; H, 5, 33; N, 6, 42. Found: C, 63, 64; H, 5, 35; N, 6, 41.

3.4. General Procedure for Transfer Hydrogenation

Functionalized dendrimer (5 mol % [Ru]), CAN (1 eq.) or pyridine (10 eq.) and iPrOH (3 mL) were stirred at reflux for 15 min. Then 4-fluoroacetophenone (36 μL, 300 μmol) was added and sampling was done at regular intervals by withdrawing aliquots (about 0.05 mL) of the reaction mixture and diluting them with addition of CDCl3 (0.5 mL). Conversions were determined by 19F NMR spectroscopy by monitoring signals at δ = −105.35 (4-fluoroacetophenone) and −115.70 ppm (1-(4-fluorophenyl)ethanol). Further details are given in Figures S22–S23 and Tables S1–S2 in the SM.

4. Conclusions

Five generations of novel cyclopentadienone-NHC-ruthenium(0)-functionalized PPIs dendrimers 3g15 with up to 64 organometallic peripheral moieties have been here successfully prepared and provide a precious route for the immobilization of hydroxy-functionalized NHC complexes onto organic and inorganic amino-decorated supports. The dendrimers 3g15 also represent one of the few examples in which NHC ligands are exploited as ancillary ligands for the functionalization of high-generation dendrimers. The stability shown by these dendrimers highlights a significant opportunity to obtain nanosized multinuclear systems, which have the potential to work in between homogeneous and heterogeneous catalytic regimes.
Dendrimers 3g15, thanks to the ruthenium(0) centers present at the periphery, are active in the transfer hydrogenation of 4-fluoroacetophenone in the presence of additives like CAN or pyridine. In line with what previously observed for the corresponding mononuclear species 1c, the limited catalytic activity is more likely due to dendrimer –CH2NH– functionalization rather than ascribable to a sterically induced detrimental dendrimer effect. Nevertheless, taking advantage of the synthetic strategy here described, further studies will be devoted in the future to improve the catalytic activity, for example by modifying the dendritic linker.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/2/264/s1, Scheme S1: Poly(propyleneimine) (PPIs) dendrimers used as support for ruthenium complexes in this work; Figures S1–S4: 1H and 13C NMR, IR and MS-ESI spectra of 2; Figures S5–S21: 1H and 13C NMR, IR and MS-ESI spectra of 3g1–5; Figures S22 and S23: 19F NMR spectra for catalytic transfer hydrogenation of 4-fluoroacetophenone with 3g15; Tables S1 and S2: 19F NMR conversion data for catalytic transfer hydrogenation of 4-fluoroacetophenone with 3g15.

Author Contributions

Conceptualization, R.M., C.C., and M.C.C.; methodology, C.C. and R.C.; validation, L.R., and A.C.; founding acquisition, V.Z.; data curation, R.M., C.C., and R.C.; writing—original draft preparation, R.M.; writing—review and editing, L.R., M.C.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Italian Ministero dell’Istruzione, dell’Università e della Ricerca (MIUR) PRIN 2015 project (20154 × 9ATP_003) entitled “Towards a Sustainable Chemistry: Design of Innovative Metal-Ligand Systems for Catalysis and Energy Applications,” and by the Università degli Studi di Bologna.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Wang, D.; Astruc, D. Dendritic catalysis—Basic concepts and recent trends. Coord. Chem. Rev. 2013, 257, 2317–2334. [Google Scholar] [CrossRef]
  2. Astruc, D. Electron-transfer processes in dendrimers and their implication in biology, catalysis, sensing and nanotechnology. Nat. Chem. 2012, 4, 255–267. [Google Scholar] [CrossRef] [PubMed]
  3. Méry, D.; Astruc, D. Dendritic catalysis: Major concepts and recent progress. Coord. Chem. Rev. 2006, 250, 1965–1979. [Google Scholar] [CrossRef]
  4. Wang, D.; Deraedt, C.; Ruiz, J.; Astruc, D. Magnetic and Dendritic Catalysts. Acc. Chem. Res. 2015, 48, 1871–1880. [Google Scholar] [CrossRef] [PubMed]
  5. Astruc, D.; Lu, F.; Aranzaes, J.R. Nanoparticles as Recyclable Catalysts: The Frontier between Homogeneous and Heterogeneous Catalysis. Angew. Chem. Int. Ed. 2005, 44, 7852–7872. [Google Scholar] [CrossRef]
  6. Astruc, D.; Boisselier, E.; Ornelas, C. Dendrimers Designed for Functions: From Physical, Photophysical, and Supramolecular Properties to Applications in Sensing, Catalysis, Molecular Electronics, Photonics, and Nanomedicine. Chem. Rev. 2010, 110, 1857–1959. [Google Scholar] [CrossRef]
  7. Oosterom, G.E.; Reek, J.N.H.; Kamer, P.C.J.; van Leeuwen, P.W.N.M. Transition Metal Catalysis Using Functionalized Dendrimers. Angew. Chem. Int. Ed. 2001, 40, 1828–1849. [Google Scholar] [CrossRef]
  8. Van Heerbeek, R.; Kamer, P.C.J.; van Leeuwen, P.W.N.M.; Reek, J.N.H. Dendrimers as Support for Recoverable Catalysts and Reagents. Chem. Rev. 2002, 102, 3717–3756. [Google Scholar] [CrossRef]
  9. Reek, J.N.H.; Arévalo, S.; van Heerbeek, R.; Kamer, P.C.J.; van Leeuwen, P.W.N.M. Dendrimers in Catalysis. In Advances in Catalysis; Elsevier: San Diego, CA, USA, 2006; Volume 49, pp. 71–151. ISBN 978-0-12-007849-3. [Google Scholar]
  10. Dykes, G.M. Dendrimers: A review of their appeal and applications. J. Chem. Technol. Biotechnol. 2001, 76, 903–918. [Google Scholar] [CrossRef]
  11. Hwang, S.-H.; Shreiner, C.D.; Moorefield, C.N.; Newkome, G.R. Recent progress and applications for metallodendrimers. New J. Chem. 2007, 31, 1192. [Google Scholar] [CrossRef]
  12. Twyman, L.J.; King, A.S.H.; Martin, I.K. Catalysis inside dendrimers. Chem. Soc. Rev. 2002, 31, 69–82. [Google Scholar] [CrossRef] [PubMed]
  13. Baráth, E. Selective Reduction of Carbonyl Compounds via (Asymmetric) Transfer Hydrogenation on Heterogeneous Catalysts. Synthesis 2020, 52, 504–520. [Google Scholar] [CrossRef]
  14. Caminade, A.-M.; Ouali, A.; Laurent, R.; Turrin, C.-O.; Majoral, J.-P. The dendritic effect illustrated with phosphorus dendrimers. Chem. Soc. Rev. 2015, 44, 3890–3899. [Google Scholar] [CrossRef] [PubMed]
  15. Tomalia, D.A. Dendritic effects: Dependency of dendritic nano-periodic property patterns on critical nanoscale design parameters (CNDPs). New J. Chem. 2012, 36, 264–281. [Google Scholar] [CrossRef]
  16. Helms, B.; Fréchet, J.M.J. The Dendrimer Effect in Homogeneous Catalysis. Adv. Synth. Catal. 2006, 348, 1125–1148. [Google Scholar] [CrossRef]
  17. Kakkar, A.K. Dendritic polymers: From efficient catalysis to drug delivery. Macromol. Symp. 2003, 196, 145–154. [Google Scholar] [CrossRef]
  18. Yi, B.; He, H.-P.; Fan, Q.-H. Synthesis and application of peripherally alkyl-functionalized dendritic pyrphos ligands: Homogeneous-supported catalysts for enantioselective hydrogenation. J. Mol. Catal. A Chem. 2010, 315, 82–85. [Google Scholar] [CrossRef]
  19. Yu, J.; RajanBabu, T.V.; Parquette, J.R. Conformationally Driven Asymmetric Induction of a Catalytic Dendrimer. J. Am. Chem. Soc. 2008, 130, 7845–7847. [Google Scholar] [CrossRef]
  20. Findeis, R.A.; Gade, L.H. Tripodal Phosphane Ligands with Novel Linker Units and Their Rhodium Complexes as Building Blocks for Dendrimer Catalysts. Eur. J. Inorg. Chem. 2003, 2003, 99–110. [Google Scholar] [CrossRef]
  21. Natarajan, B.; Jayaraman, N. Synthesis and studies of Rh (I) catalysts within and across poly(alkyl aryl ether) dendrimers. J. Organomet. Chem. 2011, 696, 722–730. [Google Scholar] [CrossRef]
  22. Rodríguez, L.I.; Rossell, O.; Seco, M.; Muller, G. Carbosilane dendrimers peripherally functionalized with P-stereogenic diphosphine ligands and related rhodium complexes. J. Organomet. Chem. 2009, 694, 1938–1942. [Google Scholar] [CrossRef]
  23. Rodríguez, L.-I.; Rossell, O.; Seco, M.; Muller, G. Rhodium or ruthenium units peripherally coordinated to carbosilane dendrimers functionalized with P-stereogenic monophosphines. J. Organomet. Chem. 2007, 692, 851–858. [Google Scholar] [CrossRef]
  24. Kassube, J.K.; Gade, L.H. Immobilisation of the Pyrphos Ligand on Soluble Hyperbranched Supports and Use in Rhodium-Catalysed Hydrogenation in Ionic Liquids. Adv. Synth. Catal. 2009, 351, 739–749. [Google Scholar] [CrossRef]
  25. Tang, W.-J.; Huang, Y.-Y.; He, Y.-M.; Fan, Q.-H. Dendritic MonoPhos: Synthesis and application in Rh-catalyzed asymmetric hydrogenation. Tetrahedron Asymm. 2006, 17, 536–543. [Google Scholar] [CrossRef]
  26. Zhang, F.; Li, Y.; Li, Z.-W.; He, Y.-M.; Zhu, S.-F.; Fan, Q.-H.; Zhou, Q.-L. Modular chiral dendritic monodentate phosphoramidite ligands for Rh (II)-catalyzed asymmetric hydrogenation: Unprecedented enhancement of enantioselectivity. Chem. Commun. 2008, 45, 6048. [Google Scholar] [CrossRef]
  27. Wang, Z.-J.; Deng, G.-J.; Li, Y.; He, Y.-M.; Tang, W.-J.; Fan, Q.-H. Enantioselective Hydrogenation of Quinolines Catalyzed by Ir (BINAP)-Cored Dendrimers: Dramatic Enhancement of Catalytic Activity. Org. Lett. 2007, 9, 1243–1246. [Google Scholar] [CrossRef]
  28. Chen, Y.-C.; Wu, T.-F.; Deng, J.-G.; Liu, H.; Cui, X.; Zhu, J.; Jiang, Y.-Z.; Choi, M.C.K.; Chan, A.S.C. Multiple Dendritic Catalysts for Asymmetric Transfer Hydrogenation. J. Org. Chem. 2002, 67, 5301–5306. [Google Scholar] [CrossRef]
  29. Chen, Y.-C.; Wu, T.-F.; Jiang, L.; Deng, J.-G.; Liu, H.; Zhu, J.; Jiang, Y.-Z. Synthesis of Dendritic Catalysts and Application in Asymmetric Transfer Hydrogenation. J. Org. Chem. 2005, 70, 1006–1010. [Google Scholar] [CrossRef]
  30. Chen, Y.-C.; Wu, T.-F.; Deng, J.-G.; Liu, H.; Jiang, Y.-Z.; Choi, M.C.K.; Chan, A.S.C. Dendritic catalysts for asymmetric transfer hydrogenation. Chem. Commun. 2001, 14, 1488–1489. [Google Scholar] [CrossRef]
  31. Liu, P.N.; Chen, Y.C.; Li, X.Q.; Tu, Y.Q.; Gen Deng, J. Dendritic catalysts for asymmetric transfer hydrogenation based (1S,2R)-norephedrine derived ligands. Tetrahedron Asymmetry 2003, 14, 2481–2485. [Google Scholar] [CrossRef]
  32. Wang, W.; Wang, Q. A fluorinated dendritic TsDPEN-Ru (ii) catalyst for asymmetric transfer hydrogenation of prochiral ketones in aqueous media. Chem. Commun. 2010, 46, 4616. [Google Scholar] [CrossRef] [PubMed]
  33. Cesari, C.; Mazzoni, R.; Müller-Bunz, H.; Albrecht, M. Ruthenium (0) complexes with triazolylidene spectator ligands: Oxidative activation for (de)hydrogenation catalysis. J. Organomet. Chem. 2015, 793, 256–262. [Google Scholar] [CrossRef]
  34. Cesari, C.; Cingolani, A.; Parise, C.; Zacchini, S.; Zanotti, V.; Cassani, M.C.; Mazzoni, R. Ruthenium hydroxycyclopentadienyl N-heterocyclic carbene complexes as transfer hydrogenation catalysts. RSC Adv. 2015, 5, 94707–94718. [Google Scholar] [CrossRef]
  35. Warner, M.C.; Casey, C.P.; Bäckvall, J.-E. Shvo’s Catalyst in Hydrogen Transfer Reactions. In Bifunctional Molecular Catalysis, Topics in Organometallic Chemistry; Ikariya, T., Shibasaki, M., Eds.; Springer: Berlin/Heidelberg, Germany, 2011; Volume 37, pp. 85–125. ISBN 978-3-642-20730-3. [Google Scholar]
  36. Conley, B.L.; Pennington-Boggio, M.K.; Boz, E.; Williams, T.J. Discovery, Applications, and Catalytic Mechanisms of Shvo’s Catalyst. Chem. Rev. 2010, 110, 2294–2312. [Google Scholar] [CrossRef]
  37. Pasini, T.; Solinas, G.; Zanotti, V.; Albonetti, S.; Cavani, F.; Vaccari, A.; Mazzanti, A.; Ranieri, S.; Mazzoni, R. Substrate and product role in the Shvo’s catalyzed selective hydrogenation of the platform bio-based chemical 5-hydroxymethylfurfural. Dalton Trans. 2014, 43, 10224–10234. [Google Scholar] [CrossRef]
  38. Cesari, C.; Sambri, L.; Zacchini, S.; Zanotti, V.; Mazzoni, R. Microwave-Assisted Synthesis of Functionalized Shvo-Type Complexes. Organometallics 2014, 33, 2814–2819. [Google Scholar] [CrossRef]
  39. Cesari, C.; Conti, S.; Zacchini, S.; Zanotti, V.; Cassani, M.C.; Mazzoni, R. Sterically driven synthesis of ruthenium and ruthenium–silver N-heterocyclic carbene complexes. Dalton Trans. 2014, 43, 17240–17243. [Google Scholar] [CrossRef]
  40. Cingolani, A.; Cesari, C.; Zacchini, S.; Zanotti, V.; Cassani, M.C.; Mazzoni, R. Straightforward synthesis of iron cyclopentadienone N-heterocyclic carbene complexes. Dalton Trans. 2015, 44, 19063–19067. [Google Scholar] [CrossRef]
  41. Hahn, F.E.; Jahnke, M.C. Heterocyclic Carbenes: Synthesis and Coordination Chemistry. Angew. Chem. Int. Ed. 2008, 47, 3122–3172. [Google Scholar] [CrossRef]
  42. Melaimi, M.; Soleilhavoup, M.; Bertrand, G. Stable Cyclic Carbenes and Related Species beyond Diaminocarbenes. Angew. Chem. Int. Ed. 2010, 49, 8810–8849. [Google Scholar] [CrossRef] [Green Version]
  43. Benhamou, L.; Chardon, E.; Lavigne, G.; Bellemin-Laponnaz, S.; César, V. Synthetic Routes to N-Heterocyclic Carbene Precursors. Chem. Rev. 2011, 111, 2705–2733. [Google Scholar] [CrossRef] [PubMed]
  44. Schaper, L.-A.; Hock, S.J.; Herrmann, W.A.; Kühn, F.E. Synthesis and Application of Water-Soluble NHC Transition-Metal Complexes. Angew. Chem. Int. Ed. 2013, 52, 270–289. [Google Scholar] [CrossRef] [PubMed]
  45. Bourissou, D.; Guerret, O.; Gabbaï, F.P.; Bertrand, G. Stable Carbenes. Chem. Rev. 2000, 100, 39–92. [Google Scholar] [CrossRef] [PubMed]
  46. Herrmann, W.A. N-Heterocyclic Carbenes: A New Concept in Organometallic Catalysis. Angew. Chem. Int. Ed. 2002, 41, 1290–1309. [Google Scholar] [CrossRef]
  47. Crudden, C.M.; Allen, D.P. Stability and reactivity of N-heterocyclic carbene complexes. Coord. Chem. Rev. 2004, 248, 2247–2273. [Google Scholar] [CrossRef]
  48. Poyatos, M.; Mata, J.A.; Peris, E. Complexes with Poly (N-heterocyclic carbene) Ligands: Structural Features and Catalytic Applications. Chem. Rev. 2009, 109, 3677–3707. [Google Scholar] [CrossRef]
  49. Schuster, O.; Yang, L.; Raubenheimer, H.G.; Albrecht, M. Beyond Conventional N-Heterocyclic Carbenes: Abnormal, Remote, and Other Classes of NHC Ligands with Reduced Heteroatom Stabilization. Chem. Rev. 2009, 109, 3445–3478. [Google Scholar] [CrossRef] [Green Version]
  50. Mercs, L.; Albrecht, M. Beyond catalysis: N-heterocyclic carbene complexes as components for medicinal, luminescent, and functional materials applications. Chem. Soc. Rev. 2010, 39, 1903. [Google Scholar] [CrossRef]
  51. Hopkinson, M.N.; Richter, C.; Schedler, M.; Glorius, F. An overview of N-heterocyclic carbenes. Nature 2014, 510, 485–496. [Google Scholar] [CrossRef]
  52. Nolan, S.P. (Ed.) N-Heterocyclic Carbenes: Effective Tools for Organometallic Synthesis, 1st ed.; Wiley-VCH: Weinheim, Germany, 2014; ISBN 978-3-527-33490-2. [Google Scholar]
  53. Garber, S.B.; Kingsbury, J.S.; Gray, B.L.; Hoveyda, A.H. Efficient and Recyclable Monomeric and Dendritic Ru-Based Metathesis Catalysts. J. Am. Chem. Soc. 2000, 122, 8168–8179. [Google Scholar] [CrossRef]
  54. Fujihara, T.; Obora, Y.; Tokunaga, M.; Sato, H.; Tsuji, Y. Dendrimer N-heterocyclic carbene complexes with rhodium (I) at the core. Chem. Commun. 2005, 4526. [Google Scholar] [CrossRef] [PubMed]
  55. Virboul, M.A.N.; Lutz, M.; Siegler, M.A.; Spek, A.L.; van Koten, G.; Klein Gebbink, R.J.M. One-Pot Synthesis and Immobilisation of Sulfonate-Tethered N-Heterocyclic Carbene Complexes on Polycationic Dendrimers. Chem. Eur. J. 2009, 15, 9981–9986. [Google Scholar] [CrossRef] [PubMed]
  56. Anjali, J.K.; Sreekumar, K. Heterogeneous High-Loading Hyperbranched Polyglycidol with Peripheral NHC–Pd Complex: Synthesis and Application as Catalyst in Suzuki Coupling Reaction. Catal. Lett. 2019, 149, 1952–1964. [Google Scholar] [CrossRef]
  57. Busetto, L.; Cassani, M.C.; van Leeuwen, P.W.N.M.; Mazzoni, R. Synthesis of new poly(propylenimine) dendrimers DAB-dendr-[NH(O)COCH2CH2OC(O)C5H4Rh(NBD)]n {n = 4, 8, 16, 32, 64} functionalized with alkoxycarbonylcyclopentadienyl complexes of rhodium (I). Dalton Trans. 2004, 17, 2767–2770. [Google Scholar] [CrossRef]
  58. Cesari, C.; Mazzoni, R.; Matteucci, E.; Baschieri, A.; Sambri, L.; Mella, M.; Tagliabue, A.; Basile, F.L.; Lucarelli, C. Hydrogen Transfer Activation via Stabilization of Coordinatively Vacant Sites: Tuning Long-Range π-System Electronic Interaction between Ru(0) and NHC Pendants. Organometallics 2019, 38, 1041–1051. [Google Scholar] [CrossRef]
Figure 1. Cyclopentadienone NHC ruthenium(0) complexes 1ac.
Figure 1. Cyclopentadienone NHC ruthenium(0) complexes 1ac.
Catalysts 10 00264 g001
Scheme 1. Synthesis of the ruthenium(0) complex 2 with CO2Im-functionalized NHC ligand.
Scheme 1. Synthesis of the ruthenium(0) complex 2 with CO2Im-functionalized NHC ligand.
Catalysts 10 00264 sch001
Scheme 2. Synthesis of the ruthenium-functionalized first-generation poly(propyleneimine) (PPIs) dendrimer 3g1.
Scheme 2. Synthesis of the ruthenium-functionalized first-generation poly(propyleneimine) (PPIs) dendrimer 3g1.
Catalysts 10 00264 sch002
Figure 2. Structure of the second-to-fifth generation dendrimers functionalized with n ruthenium(0) complexes: 3g2 (n = 8), 3g3 (n = 16), 3g4 (n = 32) and 3g5 (n = 64).
Figure 2. Structure of the second-to-fifth generation dendrimers functionalized with n ruthenium(0) complexes: 3g2 (n = 8), 3g3 (n = 16), 3g4 (n = 32) and 3g5 (n = 64).
Catalysts 10 00264 g002
Table 1. Catalytic transfer hydrogenation of 4-fluoroacetophenone with ruthenium catalysts [Ru].1
Table 1. Catalytic transfer hydrogenation of 4-fluoroacetophenone with ruthenium catalysts [Ru].1
Catalysts 10 00264 i001
Entry[Ru]AdditiveConversion (%) 8 hConversion (%) 24 h
11aCAN2561
21bCAN1087
31cCAN1025
1 General condition: ruthenium complex [Ru] (5 mol %), iPrOH (3 mL), reflux (82 °C); conversions were determined by 19F NMR spectroscopy; 1 mol eq. of CAN per ruthenium center was added.
Table 2. Catalytic transfer hydrogenation of 4-fluoroacetophenone with ruthenium dendrimers 3g15 as catalyst [Ru].1.
Table 2. Catalytic transfer hydrogenation of 4-fluoroacetophenone with ruthenium dendrimers 3g15 as catalyst [Ru].1.
Catalysts 10 00264 i002
Entry[Ru]AdditiveConversion (%) 8 hConversion (%) 24 h
13g1CAN 2917
23g2CAN 2917
33g3CAN 2715
43g4CAN 2415
53g5CAN 2616
63g1pyridine 3020
1 General conditions: ruthenium complex [Ru] (5 mol %), iPrOH (3 mL), reflux (82 °C); conversions were determined by 19F NMR spectroscopy; 2 1 mol eq. of CAN per ruthenium center; 3 10 mol eq. of pyridine per ruthenium center.

Share and Cite

MDPI and ACS Style

Cesari, C.; Conti, R.; Cingolani, A.; Zanotti, V.; Cassani, M.C.; Rigamonti, L.; Mazzoni, R. Synthesis and Reactivity of Poly(propyleneimine) Dendrimers Functionalized with Cyclopentadienone N-Heterocyclic-Carbene Ruthenium(0) Complexes. Catalysts 2020, 10, 264. https://doi.org/10.3390/catal10020264

AMA Style

Cesari C, Conti R, Cingolani A, Zanotti V, Cassani MC, Rigamonti L, Mazzoni R. Synthesis and Reactivity of Poly(propyleneimine) Dendrimers Functionalized with Cyclopentadienone N-Heterocyclic-Carbene Ruthenium(0) Complexes. Catalysts. 2020; 10(2):264. https://doi.org/10.3390/catal10020264

Chicago/Turabian Style

Cesari, Cristiana, Riccardo Conti, Andrea Cingolani, Valerio Zanotti, Maria Cristina Cassani, Luca Rigamonti, and Rita Mazzoni. 2020. "Synthesis and Reactivity of Poly(propyleneimine) Dendrimers Functionalized with Cyclopentadienone N-Heterocyclic-Carbene Ruthenium(0) Complexes" Catalysts 10, no. 2: 264. https://doi.org/10.3390/catal10020264

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