Next Article in Journal
QSAR Modeling on Benzo[c]phenanthridine Analogues as Topoisomerase I Inhibitors and Anti-cancer Agents
Next Article in Special Issue
Metathesis Transformations of Natural Products: Cross-Metathesis of Natural Rubber and Mandarin Oil by Ru-Alkylidene Catalysts
Previous Article in Journal
Optimization of a Microwave-Coupled Enzymatic Digestion Process to Prepare Peanut Peptides
Previous Article in Special Issue
Synthesis of Electronically Modified Ru-Based Neutral 16 VE Allenylidene Olefin Metathesis Precatalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Development of a Method for the Preparation of Ruthenium Indenylidene-Ether Olefin Metathesis Catalysts

Department of Chemistry and Biochemistry, California State University, Northridge, 18111 Nordhoff Street, Northridge, CA 91330, USA
*
Author to whom correspondence should be addressed.
Molecules 2012, 17(5), 5675-5689; https://doi.org/10.3390/molecules17055675
Submission received: 28 March 2012 / Revised: 18 April 2012 / Accepted: 23 April 2012 / Published: 11 May 2012
(This article belongs to the Special Issue Olefin Metathesis and Its Application)

Abstract

:
The reactions between several derivatives of 1-(3,5-dimethoxyphenyl)-prop-2-yn-1-ol and different ruthenium starting materials [i.e., RuCl2(PPh3)3 and RuCl2(p-cymene)(L), where L is tricyclohexylphosphine di-t-butylmethylphosphine, dicyclohexylphenylphosphine, triisobutylphosphine, triisopropylphosphine, or tri-n-propylphosphine] are described. Several of these reactions allow for the easy, in-situ and atom-economic preparation of olefin metathesis catalysts. Organic precursor 1-(3,5-dimethoxyphenyl)-1-phenyl-prop-2-yn-1-ol led to the formation of active ruthenium indenylidene-ether complexes, while 1-(3,5-dimethoxyphenyl)-prop-2-yn-1-ol and 1-(3,5-dimethoxyphenyl)-1-methyl-prop-2-yn-1-ol did not. It was also found that a bulky and strong σ-donor phosphine ligand was required to impart good catalytic activity to the new ruthenium complexes.

1. Introduction

Metal alkylidene complexes have been the focus of intense research in synthetic chemistry [1]. Most notable are the Schrock molybdenum and tungsten alkylidene [2] and the Grubbs ruthenium alkylidene complexes [3], which are excellent catalysts for olefin metathesis and have enabled an astonishingly broad spectrum of applications in organic and polymer synthesis [4]. Molybdenum alkylidene complexes find many important uses including in enantioselective metathesis reactions [5], while certain tungsten alkylidene systems were recently shown to be suitable catalysts for Z-selective metathesis [6]. Ruthenium alkylidene systems have been the most widely used olefin metathesis catalysts in industrial and academic laboratories, because they combine robustness, functional group tolerance, excellent activity, and good selectivity including enantioselectivity [5,7,8,9] and Z-selectivity [10].
Many ruthenium-based olefin metathesis catalysts are commercially available including members of the ruthenium benzylidene (1) [11,12], benzylidene-ether (2) [13,14], and indenylidene (3) families (Figure 1) [15,16]. The preparations of these complexes are not straightforward, but involve several steps and require crystallization of the catalysts to remove phosphine byproducts. For example, the most practical method to produce the 1st generation Hoveyda catalyst 2a consists of first preparing and isolating the 1st generation Grubbs complex 1a in a two-step, one-pot process [11], and subsequently reacting 1a with an alkoxystyrene molecule to give 2a and one equivalent of tricyclohexylphosphine (PCy3) [13]. The end of the process requires isolating the catalyst 2a away from the liberated phosphine. In addition to being cumbersome, this catalyst synthesis is not atom economic, wasting several equivalents of triphenylphosphine (PPh3) and one equivalent of the expensive PCy3 ligand. Therefore, we became motivated to develop a direct, in-situ, and atom-economic method for the synthesis of olefin metathesis catalysts.
The preparation of the ruthenium indenylidene complex 3a is comparable with that of 1a in some respects, because it also involves a two-step, one-pot process followed by isolation of the catalyst away from the liberated PPh3. On the other hand, making 3a seems more attractive because it avoids the use of diazo compounds and does not require cooling of the reaction mixtures. However, an analysis of the history of 3a’s discovery and additional studies reveals that its formation is not as straightforward as it initially appears. Indeed, the reaction between RuCl2(PPh3)3 and 1,1-diphenylprop-2-yn-1-ol (HC≡CCPh2OH) was first studied by Hill and reported to yield allenylidene complex RuCl2(PPh3)2(=C=C=CPh2) [17]. Conversely, Nolan [18] and Fürstner [19] described that the same reaction under very similar conditions (reflux in THF for 2 h) yields the indenylidene complex RuCl2(PPh3)2(PhInd) (where PhInd is a 3-phenyl-1-indenylidene fragment). Schanz found the preparation of RuCl2(PPh3)2(PhInd) under the aforementioned conditions to be unreliable and showed that it affords the dinuclear species (PPh3)2ClRu(µ-Cl)3Ru(PPh3)2(=C=C=CPh2) [20]. It is also interesting to note that [RuCl2(p-cymene)]2 reacts with HC≡CCPh2OH in the presence of two equivalents of PCy3 under similar conditions as above to give RuCl2(PCy3)2(=C=C=CPh2), which does not rearrange to RuCl2(PCy3)2(PhInd) [18]. In contrast, the formation of RuCl2(PPh3)2(PhInd) from the reaction of RuCl2(PPh3)3 and HC≡CCPh2OH is thought to go through an allenylidene-to-indenylidene rearrangement involving an electrophilic aromatic substitution of one of the allenylidene’s phenyl groups [20]. A similar rearrangement was observed by Dixneuf in a cationic ruthenium allenylidene system [21].
Altogether, we were attracted by the use of propargyl alcohol derivatives to generate metal alkylidene complexes, but desired to design 1-phenylprop-2-yn-1-ol precursors that would favor the formation of metal indenylidene over metal allenylidene complexes. We reasoned that prop-2-yn-1-ol molecules containing an activated phenyl ring would be suitable precursors. We have previously communicated the convenient, atom-economic, and in-situ preparation of a new indenylidene-ether olefin metathesis catalyst using 1-(3,5-dimethoxyphenyl)-1-phenylprop-2-yn-1-ol [22]. Herein, we describe the path that led to the development of this synthesis as well as the reaction of other organic precursors with different ruthenium starting materials.

2. Results and Discussion

2.1. Synthesis of the 1-Phenylprop-2-yn-1-ol Derivatives

Two organic precursors, 1-(3,5-dimethoxyphenyl)-1-phenylprop-2-yn-1-ol (4a) and its isotopologue (4a-13C2), were initially prepared (Scheme 1). 3,5-Dimethoxybenzophenone, obtained by reacting 3,5-dimethoxybenzonitrile with phenylmagnesium chloride, was treated with lithium acetylide (LiC≡CH and Li13C≡13CH) at −78 °C in THF to give the non-labeled 4a and 13C-doubly-labeled 4a-13C2 organic precursors, respectively. The 1H-NMR spectrum of 4a in CDCl3 showed characteristic resonances for the alcohol group (singlet at 2.84 ppm) and the acetylenic proton (singlet at 2.93 ppm). The resonance for the protons of the methoxy groups appear as a singlet at 3.74 ppm. The 13C-NMR spectrum of 4a features a total of twelve peaks including two singlets at 86.45 and 75.56 ppm for the acetylenic carbon atoms. In 4a-13C2, the acetylenic proton is characterized by a doublet of doublets (1JC-H = 250.4 Hz and 2JC-H = 50.0 Hz) in the 1H-NMR spectrum and the acetylenic carbon atoms by two doublets (1JC-C = 171.5 Hz) in the 13C-NMR spectrum.
We were also interested in making organic precursors that were less bulky than 4a (vide infra) and whose synthesis would not require the use of n-butyllithium. Thus, 1-(3,5-dimethoxyphenyl)-prop-2-yn-1-ol (4b) and 1-(3,5-dimethoxyphenyl)-1-methylprop-2-yn-1-ol (4c) were straightforwardly produced by reacting ethynylmagnesium bromide with 3,5-dimethoxybenzaldehyde and 3,5-dimethoxyacetophenone, respectively (Scheme 2). The 1H-NMR spectrum of 4b in CDCl3 showed a singlet at 2.64 ppm for the acetylenic proton. The resonance for the protons of the methoxy groups appears as a singlet at 3.78 ppm, and that for the propargylic proton is found as a singlet at 5.37 ppm. The 1H-NMR spectrum of 4c in CDCl3 is characterized by a singlet at 2.66 ppm for the acetylenic proton, a singlet at 1.76 ppm for the propargylic methyl group, and a singlet at 3.78 ppm for the methoxy substituents.

2.2. Reaction of 1-Phenylprop-2-yn-1-ol Derivatives with Different Ruthenium Starting Materials

Compound 4a was reacted with RuCl2(PPh3)3 in refluxing THF-d8. 31P-NMR spectroscopy revealed that complete disappearance of RuCl2(PPh3)3 occurred after 3 h to give free PPh3 (singlet at −4.9 ppm) and two new species in a ~6:1 ratio (singlet at 53.4 ppm for the major species and singlet at 27.5 ppm for the minor species). The methoxy groups of both the major and minor species seemed to be inequivalent, according to 1H-NMR spectroscopy, suggesting that the products could be ruthenium indenylidene complexes. Nevertheless, a firmer elucidation of the nature of the organometallic products was hampered by the lack of information in the 1H-NMR spectrum and the low intensities of the low-field signals in the 13C-NMR spectrum. Thus, we began to explore the reaction of RuCl2(PPh3)3 with the 13C-doubly labeled organic precursor 4a-13C2 in refluxing THF-d8. After 3 h of reaction, the 13C{1H} spectrum of the mixture features a doublet of doublets at 288.9 ppm (1JCα-Cβ = 49.6 Hz; 2JCα-P = 12.2 Hz) for the 13Cα nucleus of the major species, consistent with the Cα nucleus coupling to one 31P nucleus and to the 13Cβ nucleus. Further down field, the spectrum exhibits a doublet of triplets at 301.1 ppm (1JCα-Cβ = 41.3 Hz; 2JCα-P = 14.2 Hz) for the 13Cα nucleus of the minor species, consistent with the Cα nucleus coupling to two 31P nuclei and to the 13Cβ nucleus. The spectrum also exhibits a doublet of doublets at 143.8 ppm for the 13Cβ nucleus of the major species and a doublet of triplets at 140.2 ppm for the 13Cβ nucleus of the minor species, but lacks any resonance in the 250–220 ppm region. These data suggest that no appreciable amount of ruthenium allenylidene complex was formed and are consistent with the major species being a mono-phosphine ruthenium indenylidene complex 5a-13C2 and the minor species a bis-phosphine ruthenium indenylidene 5b-13C2 (Scheme 3) [17,18,19,20]. The 31P-NMR spectrum shows a doublet of doublets at 53.4 ppm for the major species and a doublet of doublets at 27.5 ppm for the minor species. The configuration of the ligands around the ruthenium center, including whether the chlorides adopt a cis- or trans-arrangement, is not known based on the above data.
The mixture of 5a-13C2 and 5b-13C2 in THF-d8 obtained above was treated with three equivalents of PCy3 per ruthenium atom at room temperature to give a mixture of organometallic species containing a new mono-phosphine ruthenium indenylidene complex 6a-13C2 as a major component (Scheme 4). 6a-13C2 is characterized by a doublet at 48.6 ppm (2JCα-P = 11.1 Hz) in the 31P-NMR spectrum, and in the 13C-NMR spectrum, by a doublet of doublets at 287.0 ppm (1JCα-Cβ = 49.0 Hz; 2JCα-P = 11.2 Hz) for the 13Cα nucleus and a doublet at 129.2 ppm (1JCα-Cβ = 49.0 Hz) for the 13Cβ nucleus.
Complex 6a-13C2 can also be generated as the major product by reacting RuCl2(p-cymene)(PCy3) with organic precursor 4a-13C2 upon reflux in THF-d8 for 16 h (Scheme 5) [22]. This reaction also produces a minor product 6b-13C2, whose 31P-NMR spectrum features a doublet at 68.1 ppm (2JCα-P = 15.2 Hz) and whose 13C-NMR spectrum shows a doublet of doublets at 256.2.0 ppm (1JCα-Cβ = 47.8 Hz; 2JCα-P = 15.0 Hz) for the 13Cα nucleus and a doublet at 129.2 ppm (1JCα-Cβ = 48.9 Hz) for the 13Cβ nucleus. A mixture of non-labeled 6a/6b complexes was prepared on a gram scale by a similar procedure. However, efforts to isolate and purify the major and minor products 6a and 6b by silica gel chromatography and crystallization have been unsuccessful.
Bruneau and coworkers independently prepared a related complex by a similar approach utilizing 1,1-bis-(3,5-diisopropoxyphenyl)prop-2-yn-1-ol as an organic precursor [23]. The NMR spectroscopy data for their complex is very similar to those for our minor species (6b-13C2): The 31P-NMR spectrum shows a resonance at 68.2 ppm and the 13C-NMR spectrum shows resonances at 258.8 ppm (2JCα-P = 14.2 Hz) and 136.3 ppm for the 13Cα and 13Cβ nuclei, respectively. Additionally, Bruneau and coworkers obtained a crystal structure of their complex showing it to be a mono-PCy3 ruthenium indenylidene-ether complex where the phosphine and ether ligands are trans to each other. We propose that 6a and 6b may be isomers, where the minor species 6b possesses a structure similar to that of Bruneau’s complex (trans-phosphine-ether arrangement) and where the major complex 6a adopts a cis-phosphine-ether arrangement as the more stable isomer, due to the reduced steric of the methoxy group in 6 compared to the bulkier isopropoxy group in Bruneau’s complex. Similarly, a relatively unhindered ruthenium alkylidene-pyridine complex supported by a NHC ligand was shown to exist as a major isomer adopting a cis-NHC-pyridine arrangement and a minor isomer with a trans-NHC-pyridine configuration [24].
A solution of complexes 6a and 6b, prepared in situ using the non-labeled organic precursor 4a following a similar procedure as that shown in scheme 5, can be used without additional treatment to catalyze ring-closing metatheses (RCM) to produce 5-, 6-, and 7-membered disubstituted cycloalkenes with activities similar to those of the 1st generation Hoveyda catalyst 2a under standard conditions [22,25]. In order to explore the influence of the phosphine ligand on the activity of these in-situ generated ruthenium indenylidene-ether complexes, other RuCl2(p-cymene)(L) starting materials, where L is di-t-butylmethylphosphine [P(tBu)2Me], dicyclohexylphenylphosphine [P(Cy)2Ph], triisobutylphosphine [P(iBu)3], triisopropylphosphine [P(iPr)3], or tri-n-propylphosphine [P(nPr)3], were reacted with organic precursor 4a in refluxing THF-d8 for 16 h. 31P-NMR spectroscopy showed that each one of these reactions affords new organometallic species (Table 1). The activity of these complexes in ring-closing metathesis is described in Section 2.3.
We also explored the possibility of using 1-(3,5-dimethoxyphenyl)-prop-2-yn-1-ol (4b) and 1-(3,5-dimethoxyphenyl)-1-methylprop-2-yn-1-ol (4c) as organic precursors for the preparation of ruthenium indenylidene-ether complexes. This study was motived by two main factors. First, compounds 4b and 4c are easier to prepare than 4a (see above). Second, we hypothesized that less hindered indenylidene fragments may lead to faster-initiating catalysts based on a comparison of the RCM activity of our and Bruneau’s catalysts (Section 2.3). According to 1H- and 31P-NMR spectroscopy, RuCl2(p-cymene)(PCy3) does not react with either 4b or 4c in THF-d8 at room temperature after 18 h. Furthermore, heating the respective solutions at 40 °C for 18 h yields much starting materials and several unidentified new species. The reaction of RuCl2(p-cymene)(PCy3) with 4b or 4c in refluxing THF-d8 for 18 h affords complicated mixtures of products. Similarly, refluxing a solution of RuCl2(PPh3)3 and 4c for 3 h produces multiple species. On the other hand, RuCl2(PPh3)3 reacts quite cleanly with 4b to generate a new species characterized by four doublets of equal intensity at 50.4 ppm (2JP-P = 37.7 Hz), 47.1 ppm (2JP-P = 38.3 Hz), 40.9 ppm (2JP-P = 23.8 Hz), and 39.0 ppm (2JP-P = 24.0 Hz) in the 31P-NMR spectrum. This 31P-NMR signature is very similar to that of asymmetric dinuclear vinylidene complexes (P-P)ClRu(µ-Cl)3Ru(P-P)(=C=CHR) [26], and almost identical to that of the asymmetric bimetallic allenylidene compound (PPh3)2ClRu(µ-Cl)3Ru(PPh3)2(=C=C=CPh2) [17,20]. The 1H-NMR spectrum shows equivalent methoxy groups, but does not exhibit any signals corresponding to a Ru=C=CHC(OH)Ph fragment [27], leading us to believe that (PPh3)2ClRu(µ-Cl)3Ru(PPh3)2(=C=C=CHAr) (where Ar = 3,5-dimethoxyphenyl) may have been formed. In any case, the reaction of RuCl2(PPh3)3 and 4b does not form an indenylidene complex. Altogether, these results suggest that the derivatives of 1-phenylprop-2-yn-1-ol may need to bear two electron-withdrawing groups (e.g., aryl groups) in the propargylic position to be suitable precursors for the clean and efficient formation of ruthenium indenylidene complexes.

2.3. RCM Activity of the Ruthenium Indenylidene-Ether Complexes

As mentioned above, a solution of 6a/6b generated in situ efficiently promotes the formation of 5-, 6-, and 7-membered disubstituted cycloalkenes by RCM. Although the activity of our system is very similar to that of the 1st generation Hoveyda catalyst 2a in the RCM of diethyl diallylmalonate (DEDAM) under standard conditions [22,25], it is interesting to note that Bruneau’s catalyst exhibits a longer initiation period (Figure 2) [23]. Indeed, at the 30 min time point, Bruneau’s catalyst has converted about 33% of the substrate to product, while our catalyst has reached greater than 90% conversion. A possible explanation for this is that the bulkier indenylidene ligand of Bruneau’s catalyst hinders the rotation around the Ru=C bond of the ruthenium indenylidene fragment, a rotation that may be necessary for the formation of the rotamer that initiates the olefin metathesis reaction. A similar rotation takes place in the Hoveyda catalysts, [28,29] whose activation is thought to involve dissociative and associative interchange pathways [30].
The activity of the solutions generated by the reaction of RuCl2(p-cymene)(L) and organic precursor 4a (Table 1) was compared to that of the solution of 6a/6b (L = PCy3) in the RCM of DEDAM. The catalytic system 6a/6b bearing the PCy3 ligand is the most active, reaching greater than 97% conversion within 30 min at 1.0 mol % catalyst loading (Table 2; entry 1). The catalysts supported by P(tBu)2Me or P(iPr)3 ligands are less effective than 6a/6b (Table 2; compare entries 2 and 6 to entry 1), but are still able to achieve greater than 90% conversion within 60 min at 2.0 mol % catalyst loadings (Table 2; entries 3 and 7). Conversely, the complexes containing the P(Cy)2Ph, P(iBu)3 or P(nPr)3 ligands show very low activity in the RCM of DEDAM under the tested conditions (Table 2; entries 4, 5 and 8).
These results follow the ligand effects observed by Grubbs and coworkers, namely that larger and more electron-donating phosphine ligands lead to more active catalysts [31]. These trends are easily rationalized for the 1st generation Grubbs systems which enter the catalytic cycle via a dissociative substitution of a phosphine with an olefin [32]. Thus, the ligand dissociation is favored by the large steric hindrance and the strong trans influence of ligands such as PCy3, producing more active species. The explanation for the trends within the ruthenium indenylidene-ether set of catalysts is less intuitive. Nevertheless, it seems reasonable that phosphine ligands with stronger electron-donating abilities may turn over faster by accelerating the olefin binding step [33] or by stabilizing the metallacyclobutane intermediate and metathesis transition states [34], as it was originally proposed in the case of N-heterocyclic carbene (NHC) versus phosphine ligands [35]. It is also conceivable that bulkier phosphine ligands are required to influence the Ru=C bond rotation favoring the formation of the active rotamer and driving the reaction toward the less sterically hindered intermediates and transition states [33,34].

3. Experimental

3.1. General

NMR spectra were recorded on a Bruker 400 MHz NMR spectrometer running Xwin-NMR software. Chemical shifts are reported in parts per million (ppm) downfield from tetramethylsilane (TMS) with reference to internal solvent for 1H-NMR and 13C-NMR spectra. Chemical shifts are reported in parts per million (ppm) downfield from H3PO4 for 31P-NMR spectra. All glassware was oven dried. Unless noted otherwise, all reactions were conducted under an atmosphere of argon (in an argon-filled glove box or under argon using Schlenk techniques). All organic solvents were dried by passage through solvent purification columns containing activated molecular sieves. All other commercial chemicals were used as obtained. Organic precursors 4a and 4a-13C2 [22], and RuCl2(p-cymene)(L) [36] [where L = PCy3, P(tBu)2Me, PCy2Ph, P(iBu)3, P(iPr)3, and P(nPr)3] were prepared according to literature procedures.

3.2. Preparation of the 1-Phenylprop-2-yn-1-ol Derivatives 4b and 4c

3.2.1. Preparation of 1-(3,5-Dimethoxyphenyl)-prop-2-yn-1-ol (4b)

A dry 100 mL reaction flask equipped with a stir bar was charged with 3,5-dimethoxybenzaldehyde (1.0 g, 6.0 mmol) and anhydrous THF (15 mL) inside the glove box, capped with a septum, and taken out of the glove box. The mixture was placed in a 0 °C ice bath. A 0.5 M solution of ethynylmagnesium bromide in THF (20 mL, 10 mmol) was added dropwise under stirring. The reaction mixture was allowed to warm up to room temperature and stirred for 16 h. A 10% aqueous solution of NH4Cl (40 mL) was added and the mixture stirred for 30 min. The product was extracted with ether (3 × 30 mL) and the combined organic layers were washed with water (50 mL) and brine (50 mL) before being dried with anhydrous magnesium sulfate. The filtrate was dried in vacuo to afford 1-(3,5-dimethoxyphenyl)-prop-2-yn-1-ol as an orange oil in 91% yield. 1H-NMR (CDCl3): δ 6.70 (s, 2H), 6.41 (s, 1H), 5.37 (s, 1H), 3.78 (s, 6H), 2.64 (s, 1H), proton from OH group not observed. 13C{1H} NMR (CDCl3): δ 160.96, 142.44, 104.55, 100.57, 83.42, 74.71, 64.35, 55.41.

3.2.2. Preparation of 1-(3,5-Dimethoxyphenyl)-1-methylprop-2-yn-1-ol (4c)

Powdered potassium carbonate (27.3 g, 197.5 mmol) was added to a suspension of 3,5-dihydroxy-acetophenone (5.0 g, 32.9 mmol) in acetone (50 mL) and the mixture was stirred vigorously for 20 min. Methyl iodide (8.4 mL, 134.9 mmol) was added dropwise and the reaction mixture was refluxed for 16 h. The mixture was filtered and the solid washed with acetone. Water (100 mL) was added to the filtrate, and the product was extracted with ether (3 × 100 mL). The combined organic layers were washed with water (200 mL) before being dried with anhydrous magnesium sulfate. The filtrate was dried in vacuo to afford 3,5-dimethoxyacetophenone as a dark red oil in 93% yield. 1H-NMR (CDCl3): δ 7.09 (s, 2H), 6.65 (s, 1H), 3.82 (s, 6H), 2.57 (s, 3H). 13C-NMR (CDCl3): δ 197.76, 160.88, 139.10, 106.18, 105.34, 55.59, 26.72. A dry 100 mL reaction flask equipped with stir bar was charged with 3,5-dimethoxyacetophenone (0.5 g, 2.8 mmol) and anhydrous THF (10 mL) inside the glove box, capped with a septum, and taken out of the glove box. The mixture was placed in a 0 °C ice bath. A 0.5 M solution of ethynylmagnesium bromide in THF (9 mL, 4.5 mmol) was added dropwise under stirring. The reaction mixture was allowed to warm up to room temperature and stirred for 16 h. A 10% aqueous solution of NH4Cl (20 mL) was added and the mixture stirred for 30 min. The product was extracted with ether (3 × 30 mL) and the combined organic layers were washed with water (50 mL) and brine (100 mL) before being dried with anhydrous magnesium sulfate. The filtrate was dried in vacuo to afford 1-(3,5-dimethoxyphenyl)-1-methylprop-2-yn-1-ol as a brown oil in 84% yield. 1H-NMR (CDCl3): δ 6.82 (s, 2H), 6.39 (s, 1H), 3.80 (s, 6H), 2.66 (s, 1H), 1.76 (s, 3H), proton from OH group not observed. 13C{1H} NMR (CDCl3): δ 160.71, 147.65, 103.31, 99.66, 87.16, 72.96, 69.83, 55.29, 33.01.

3.3. Reaction between 1-Phenylprop-2-yn-1-ol Derivatives and Different Ruthenium Starting Materials

3.3.1. NMR Study of the Reaction between RuCl2(PPh3)3 and 1-(3,5-dimethoxyphenyl)-1-phenylprop-2-yn-1-ol (4a)

1-(3,5-Dimethoxyphenyl)-1-phenylprop-2-yn-1-ol, 4a, (12 mg, 0.0375 mmol, 1.2 equiv) was weighed in a 2 mL vial and RuCl2(PPh3)3 (30 mg, 0.03128 mmol, 1.0 equiv) was weighed in a separate 2 mL vial. The organic precursor 4a was dissolved with THF-d8 (0.5 mL) and this solution of 4a transferred to the vial containing the RuCl2(PPh3)3 starting material. The mixture was then transferred to a J-Young tube, which was capped, brought out of the glove box and placed for 3 h in an oil bath set at 70 °C. The reaction mixture was analyzed by 31P-NMR. 31P-NMR (161 MHz, THF-d8): δ 53.4 (s; major species), 27.5 (s; minor species), −4.9 (s, PPh3).

3.3.2. NMR study of the Reaction between RuCl2(PPh3)3 and 1-(3,5-Dimethoxyphenyl)-1-Phenylprop-2-yn-1-ol (4a-13C2)

The reaction between RuCl2(PPh3)3 and 4a-13C2 was set up following the same procedure used for the reaction between RuCl2(PPh3)3 and 4a (see sub-section 3.3.1.). The reaction mixture was analyzed by 13C and 31P-NMR. 13C{1H} NMR (100 MHz, THF-d8): δ 301.1 (dt, 1JCα-Cβ = 41.3 Hz; 2JCα-P = 14.2 Hz; Cα of minor species), 288.9 (dd, 1JCα-Cβ = 49.6 Hz; 2JCα-P = 12.2 Hz; Cα of major species), 143.8 (dd, 1JCα-Cβ = 49.4 Hz; 3JCβ-P = 3.7 Hz; Cβ of major species), 140.2 (dt, 1JCα-Cβ = 41.2 Hz; 3JCβ-P = 5.3 Hz; Cβ of minor species). 31P-NMR (161 MHz, THF-d8): δ 53.4 (dd, 2JCα-P = 12.0 Hz; 3JCβ-P = 3.1 Hz; major species), 27.5 (dd, 2JCα-P = 13.6 Hz; 3JCβ-P = 5.5 Hz; minor species), −4.9 (s, PPh3).

3.3.3. Gram-Scale Preparation of the 6a/6b Mixture

A 100 mL Schlenk tube equipped with a stir bar was charged with 1-(3,5-dimethoxyphenyl)-1-phenylprop-2-yn-1-ol, 4a, (730 mg, 2.72 mmol, 1.1 equiv), RuCl2(p-cymene)(PCy3) (1.45 g, 2.47 mmol), and THF (25 mL). The Schlenk tube was sealed, brought out of the glove box, and placed under stirring for 16 h in an oil bath set at 70 °C. The volatiles were removed under vacuum. The brown residue was then dissolved with toluene (5 mL) in air and the toluene solution was slowly added to pentane (200 mL) in a 500 mL Erlenmeyer flask under vigorous stirring. The brown precipitate was collected by filtration and dried under vacuum overnight to yield 1.11 g of brown crystalline material (65% yield). 31P-NMR (161 MHz, THF-d8): δ 48.6 (s; major species), 68.1 (s; minor species).

3.3.4. General Procedure for the Reaction between RuCl2(p-cymene)(L) and Organic Precursors 4a

A 2 mL vial was charged with RuCl2(p-cymene)(L) (50 mg), the organic precursor 4a (1.1 equiv), and THF-d8 (0.5 mL). The mixture was then transferred to a J-Young tube, which was capped, brought out of the glove box, and placed in an oil bath set at 70 °C for 16 h. 31P-NMR spectra were recorded (see Table 1 in Results and Discussion).

3.3.5. General Procedure for the Reaction between RuCl2(p-cymene)(PCy3) and Organic Precursors 4b and 4c

A 2 mL vial was charged with RuCl2(p-cymene)(PCy3) (30 mg, 0.051 mmol, 1.0 equiv), the organic precursor 4b or 4c (0.097 mmol, 1.9 equiv), and THF-d8 (0.5 mL). The mixture was then transferred to a J-Young tube, which was capped and brought out of the glove box. The reactions were monitored at room temperature, 40 °C and 70 °C by 1H-NMR and 31P-NMR spectroscopy (see Results and Discussion).

3.3.6. General Procedure for the Reaction between RuCl2(PPh3)3 and Organic Precursors 4b and 4c

RuCl2(PPh3)3 (30 mg, 0.031 mmol, 1.0 equiv) was weighed in a 2 mL vial and the organic precursor 4b or 4c (0.065 mmol, 2.1 equiv) was weighed in a separate 2 mL vial. The organic precursor was dissolved with THF-d8 (0.5 mL) and this solution was transferred to the vial containing the RuCl2(PPh3)3 starting material. The mixture was then transferred to a J-Young tube, which was capped, brought out of the glove box, and placed for 3 h in an oil bath set at 70 °C. The reactions were analyzed by 1H-NMR and 31P-NMR spectroscopy (see Results and Discussion).

3.4. General Procedure for the RCM of Diethyl Diallylmalonate (DEDAM)

A 0.1 M stock solution of DEDAM was prepared in the glove box by dissolving DEDAM (60 mg, 0.25 mmol) in 2.44 mL of CD2Cl2. A portion of this 0.1 M DEDAM solution (0.5 mL, 50 µmol) was transferred to a NMR tube equipped with a screw-cap septum top. Separately, a 4 mL conical vial was charged with RuCl2(p-cymene)(L) (0.085 mmol) and 4a (25 mg, 0.094 mmol, 1.1 equiv). The vial was then filled with THF to a 1.0 mL mark (calibrated) before dropping a spin vane in the solution. The vial was sealed, brought out of the glove box, and placed under stirring for 16 h in an oil bath set at 70 °C. A portion of this solution (6.0 µL for 1.0 mol %, and 12 µL for 2.0 mol % ruthenium loading) was added to the 0.1 M solution of DEDAM in CD2Cl2 (0.5 mL, 50 µmol) by injecting it with through the septum using a syringe outside the glove box. The NMR tube was then placed in an oil bath regulated at 40 °C and the reaction mixture was analyzed by 1H NMR spectroscopy after a period of time (30 and 60 min). The extent of conversion of the RCM reaction was determined by comparing the ratio of the integrals of the methylene protons in the substrate (dt, 2.61 ppm) with those in the product (s, 2.98 ppm).

4. Conclusions

Reactions between several derivatives of 1-(3,5-dimethoxyphenyl)-prop-2-yn-1-ol and different ruthenium starting materials (i.e., RuCl2(PPh3)3 and RuCl2(p-cymene)(L), where L is tricyclohexylphosphine, di-t-butylmethylphosphine, dicyclohexylphenylphosphine, triisobutylphosphine, triisopropylphosphine, or tri-n-propylphosphine) have been explored and have led to the development of a straightforward method for the preparation of new ruthenium indenylidene-ether ring-closing metathesis catalysts. The method involves the reaction between RuCl2(p-cymene)(PCy3) and 1-(3,5-dimethoxyphenyl)-1-phenylprop-2-yn-1-ol in refluxing THF and possesses many advantages. First, it is a one-step process from the starting materials. Second, it does not require the use of difficult-to-handle diazo compounds and does not need to be conducted at low temperature. Third, it consumes only one equivalent of the expensive PCy3 ligand and is altogether very atom-economic. Fourth, it does not produce any inhibiting byproducts, allowing the catalyst solution to be used without further treatment. The resulting catalyst promotes the formation of 5-, 6-, and 7-membered disubstituted cycloalkenes with activities comparable to that of the commercial 1st generation Hoveyda catalyst under standard conditions.
Additionally, it was shown that the use of 1-(3,5-dimethoxyphenyl)-prop-2-yn-1-ol and 1-(3,5-dimethoxyphenyl)-1-methyl-prop-2-yn-1-ol did not lead to the formation of ruthenium indenylidene complexes, indicating that the 1-phenylprop-2-yn-1-ol derivatives may need to bear two aryl groups in the propargylic position to be suitable precursors for these types of reactions.
Finally, a study of the effects of the phosphine ligand on the ring-closing metathesis activity of these new ruthenium complexes revealed that larger and more electron-donating phosphine ligands lead to more efficient catalysts.

Acknowledgments

Acknowledgment is made to the Donors of the American Chemical Society Petroleum Research Fund for partial support of this research (grant # 50352-UN13). We thank Materia, Inc. for their generous gift of [RuCl2(p-cymene)]2.

References and Notes

  1. Herndon, J.W. The chemistry of the carbon-transition metal double and triple bond. Annual survey covering the year 2009. Coord. Chem. Rev. 2011, 255, 3–100. [Google Scholar] [CrossRef]
  2. Schrock, R.R. Recent advances in high oxidation state Mo and W imido alkylidene chemistry. Chem. Rev. 2009, 109, 3211–3226. [Google Scholar] [CrossRef] [PubMed]
  3. Vougioukalakis, G.C.; Grubbs, R.H. Ruthenium-based heterocyclic carbene-coordinated olefin metathesis catalysts. Chem. Rev. 2010, 110, 1746–1787. [Google Scholar] [CrossRef] [PubMed]
  4. Grubbs, R.H. Handbook of Metathesis; Wiley-VCH: Weinheim, Germany, 2003. [Google Scholar]
  5. Hoveyda, A.H.; Malcolmson, S.J.; Meek, S.J.; Zhugralin, A.R. Catalytic enantioselective olefin metathesis in natural product synthesis. chiral metal-based complexes that deliver high enantioselectivity and more. Angew. Chem. Int. Ed. Engl. 2010, 49, 34–44. [Google Scholar] [CrossRef] [PubMed]
  6. Peryshkov, D.V.; Schrock, R.R.; Takase, M.K.; Muller, P.; Hoveyda, A.H. Z-Selective olefin metathesis reactions promoted by tungsten oxo alkylidene complexes. J. Am. Chem. Soc. 2011, 133, 20754–20757. [Google Scholar] [CrossRef] [PubMed]
  7. Seiders, T.J.; Ward, D.W.; Grubbs, R.H. Enantioselective ruthenium-catalyzed ring-closing metathesis. Org. Lett. 2001, 3, 3225–3228. [Google Scholar] [CrossRef] [PubMed]
  8. van Veldhuizen, J.J.; Garber, S.B.; Kingsbury, J.S.; Hoveyda, A.H. A recyclable chiral Ru catalyst for enantioselective olefin metathesis. Efficient catalytic asymmetric ring-opening/cross metathesis in air. J. Am. Chem. Soc. 2002, 124, 4954–4955. [Google Scholar] [CrossRef] [PubMed]
  9. Funk, T.W.; Berlin, J.M.; Grubbs, R.H. Highly active chiral ruthenium catalysts for asymmetric ring-closing olefin metathesis. J. Am. Chem. Soc. 2006, 128, 1840–1846. [Google Scholar] [CrossRef] [PubMed]
  10. Endo, K.; Grubbs, R.H. Chelated ruthenium catalysts for z-selective olefin metathesis. J. Am. Chem. Soc. 2011, 133, 8525–8527. [Google Scholar] [CrossRef] [PubMed]
  11. Schwab, P.; Grubbs, R.H.; Ziller, J.W. Synthesis and applications of RuCl2(=CHR')(PR3)2: The influence of the alkylidene moiety on metathesis activity. J. Am. Chem. Soc. 1996, 118, 100–110. [Google Scholar] [CrossRef]
  12. Scholl, M.; Ding, S.; Lee, C.W.; Grubbs, R.H. Synthesis and activity of a new generation of ruthenium-based olefin metathesis catalysts coordinated with 1,3-dimesityl-4,5-dihydroimidazol-2-ylidene ligands. Org. Lett. 1999, 1, 953–956. [Google Scholar] [CrossRef] [PubMed]
  13. Kingsbury, J.S.; Harrity, J.P.A.; Bonitatebus, P.J., Jr.; Hoveyda, A.H. A recyclable ru-based metathesis catalyst. J. Am. Chem. Soc. 1999, 121, 791–799. [Google Scholar] [CrossRef]
  14. 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]
  15. Fürstner, A.; Thiel, O.R.; Ackermann, L.; Schanz, H.-J.; Nolan, S.P. Ruthenium carbene complexes with N,N′-bis(mesityl)imidazol-2-ylidene ligands: RCM catalysts of extended scope. J. Org. Chem. 2000, 65, 2204–2207. [Google Scholar] [CrossRef] [PubMed]
  16. Monsaert, S.; Drozdzak, R.; Dragutan, V.; Dragutan, I.; Verpoort, F. Indenylidene-ruthenium complexes bearing saturated N-heterocyclic carbenes: Synthesis and catalytic investigation in olefin metathesis reactions. Eur. J. Inorg. Chem. 2008, 2008, 432–440. [Google Scholar] [CrossRef]
  17. Harlow, K.J.; Hill, A.F.; Wilton-Ely, J.D.E.T. The first co-ordinatively unsaturated Group 8 allenylidene complexes: Insights into Grubbs’ vs. Dixneuf-Furstner olefin metathesis catalysts. J. Chem. Soc. Dalton Trans. 1999, 285–292. [Google Scholar] [CrossRef]
  18. Schanz, H.-J.; Jafarpour, L.; Stevens, E.D.; Nolan, S.P. Coordinatively unsaturated 16-electron ruthenium allenylidene complexes: Synthetic, structural, and catalytic studies. Organometallics 1999, 18, 5187–5190. [Google Scholar] [CrossRef]
  19. Fuerstner, A.; Guth, O.; Duffels, A.; Seidel, G.; Liebl, M.; Gabor, B.; Mynott, R. Indenylidene complexes of ruthenium: Optimized synthesis, structure elucidation, and performance as catalysts for olefin metathesis-application to the synthesis of the ADE-ring system of nakadomarin A. Chem.-Eur. J. 2001, 7, 4811–4820. [Google Scholar] [CrossRef]
  20. Shaffer, E.A.; Chen, C.-L.; Beatty, A.M.; Valente, E.J.; Schanz, H.-J. Synthesis of ruthenium phenylindenylidene, carbyne, allenylidene and vinylmethylidene complexes from (PPh3)3-4RuCl2: A mechanistic and structural investigation. J. Organomet. Chem. 2007, 692, 5221–5233. [Google Scholar] [CrossRef]
  21. Castarlenas, R.; Vovard, C.; Fischmeister, C.; Dixneuf, P.H. Allenylidene-to-indenylidene rearrangement in arene-ruthenium complexes: A key step to highly active catalysts for olefin metathesis reactions. J. Am. Chem. Soc. 2006, 128, 4079–4089. [Google Scholar] [CrossRef] [PubMed]
  22. Jimenez, L.R.; Gallon, B.J.; Schrodi, Y. A most convenient and atom-economic preparation of a highly active ring-closing metathesis catalyst. Organometallics 2010, 29, 3471–3473. [Google Scholar] [CrossRef]
  23. Kabro, A.; Roisnel, T.; Fischmeister, C.; Bruneau, C. Ruthenium-Indenylidene olefin metathesis catalyst with enhanced thermal stability. Chem.-Eur. J. 2010, 16, 12255–12261. [Google Scholar] [CrossRef] [PubMed]
  24. Ung, T.; Hejl, A.; Grubbs, R.H.; Schrodi, Y. Latent ruthenium olefin metathesis catalysts that contain an n-heterocyclic carbene ligand. Organometallics 2004, 23, 5399–5401. [Google Scholar] [CrossRef]
  25. Ritter, T.; Hejl, A.; Wenzel, A.G.; Funk, T.W.; Grubbs, R.H. A standard system of characterization for olefin metathesis catalysts. Organometallics 2006, 25, 5740–5745. [Google Scholar] [CrossRef]
  26. Amoroso, D.; Yap, G.P.A.; Fogg, D.E. Deactivation of ruthenium metathesis catalysts via facile formation of face-bridged dimers. Organometallics 2002, 21, 3335–3343. [Google Scholar] [CrossRef]
  27. Bustelo, E.; Tenorio, M.J.; Puerta, M.C.; Valerga, P. Activation of alkynols by [Cp*RuCl(PEt3)2]: new intermediates and alternative dehydration products. X-ray crystal structures of [Cp*Ru{=C=CHC(=CH2)Ph}(PEt3)2][BPh4], [Cp*Ru(=C=C=CPh2)(PEt3)2][BPh4], and [Cp*Ru{C≡CC(PEt3)Me2}(PEt3)2][BPh4]. Organometallics 1999, 18, 4563–4573. [Google Scholar] [CrossRef]
  28. Vorfalt, T.; Wannowius, K.-J.; Plenio, H. Probing the mechanism of olefin metathesis in grubbs-hoveyda and grela type complexes. Angew. Chem. Int. Ed. Engl. 2010, 49, 5533–5536. [Google Scholar] [CrossRef] [PubMed]
  29. Ashworth, I.W.; Hillier, I.H.; Nelson, D.J.; Percy, J.M.; Vincent, M.A. What is the initiation step of the Grubbs-Hoveyda olefin metathesis catalyst? Chem. Commun. 2011, 47, 5428–5430. [Google Scholar] [CrossRef] [PubMed]
  30. Thiel, V.; Hendann, M.; Wannowius, K.-J.; Plenio, H. On the mechanism of the initiation reaction in Grubbs-Hoveyda complexes. J. Am. Chem. Soc. 2012, 134, 1104–1114. [Google Scholar] [CrossRef] [PubMed]
  31. Dias, E.L.; Nguyen, S.T.; Grubbs, R.H. Well-defined ruthenium olefin metathesis catalysts: Mechanism and activity. J. Am. Chem. Soc. 1997, 119, 3887–3897. [Google Scholar] [CrossRef]
  32. Sanford, M.S.; Ulman, M.; Grubbs, R.H. New insights into the mechanism of ruthenium-catalyzed olefin metathesis reactions. J. Am. Chem. Soc. 2001, 123, 749–750. [Google Scholar] [CrossRef] [PubMed]
  33. Mathew, J.; Suresh, C.H. Assessment of stereoelectronic effects in grubbs first-generation olefin metathesis catalysis using molecular electrostatic potential. Organometallics 2011, 30, 1438–1444. [Google Scholar] [CrossRef]
  34. Occhipinti, G.; Bjorsvik, H.-R.; Jensen, V.R. Quantitative structure-activity relationships of ruthenium catalysts for olefin metathesis. J. Am. Chem. Soc. 2006, 128, 6952–6964. [Google Scholar] [CrossRef] [PubMed]
  35. Sanford, M.S.; Love, J.A.; Grubbs, R.H. Mechanism and activity of ruthenium olefin metathesis catalysts. J. Am. Chem. Soc. 2001, 123, 6543–6554. [Google Scholar] [CrossRef] [PubMed]
  36. Demonceau, A.; Stumpf, A.W.; Saive, E.; Noels, A.F. Novel ruthenium-based catalyst systems for the ring-opening metathesis polymerization of low-strain cyclic olefins. Macromolecules 1997, 30, 3127–3136. [Google Scholar] [CrossRef]
Sample Availability: Not available.
Figure 1. Ruthenium alkylidene olefin metathesis catalysts.
Figure 1. Ruthenium alkylidene olefin metathesis catalysts.
Molecules 17 05675 g001
Scheme 1. Synthesis of 1-(3,5-dimethoxyphenyl)-1-phenylprop-2-yn-1-ol.
Scheme 1. Synthesis of 1-(3,5-dimethoxyphenyl)-1-phenylprop-2-yn-1-ol.
Molecules 17 05675 sch001
Scheme 2. Synthesis of smaller derivatives of 1-phenylprop-2-yn-1-ol.
Scheme 2. Synthesis of smaller derivatives of 1-phenylprop-2-yn-1-ol.
Molecules 17 05675 sch002
Scheme 3. Reaction between RuCl2(PPh3)3 and 4a-13C2.
Scheme 3. Reaction between RuCl2(PPh3)3 and 4a-13C2.
Molecules 17 05675 sch003
Scheme 4. Phosphine exchange on complexes 5a-13C2 and 5b-13C2.
Scheme 4. Phosphine exchange on complexes 5a-13C2 and 5b-13C2.
Molecules 17 05675 sch004
Scheme 5. Reaction between RuCl2(p-cymene)(PCy3) and 4a-13C2.
Scheme 5. Reaction between RuCl2(p-cymene)(PCy3) and 4a-13C2.
Molecules 17 05675 sch005
Figure 2. Activity comparison for catalysts 2a, 6a/6b, and Bruneau’s catalyst.
Figure 2. Activity comparison for catalysts 2a, 6a/6b, and Bruneau’s catalyst.
Molecules 17 05675 g002
Table 1. 31P-NMR shifts for RuCl2(p-cymene)(L) and for the major product of the reaction between RuCl2(p-cymene)(L) and 4a. a
Table 1. 31P-NMR shifts for RuCl2(p-cymene)(L) and for the major product of the reaction between RuCl2(p-cymene)(L) and 4a. a
Ligand (L)RuCl2(p-cymene)(L)
δ (ppm)
Major Product
δ (ppm)
P(tBu)2Me43.861.1
P(Cy)2Ph19.550.0
P(iBu)322.944.2
P(iPr)335.758.1
P(nPr)316.140.8
a THF-d8; reflux; 16 h.
Table 2. RCM of DEDAM with solutions prepared by reaction of RuCl2(p-cymene)(L) and 4a. a
Table 2. RCM of DEDAM with solutions prepared by reaction of RuCl2(p-cymene)(L) and 4a. a
Molecules 17 05675 i001
EntryLigand (L)Ru loading
(mol %)
Time
(min)
Conversion
(%) b
1PCy31.030> 97
2P(tBu)2Me1.06063
3P(tBu)2Me2.06092
4P(Cy)2Ph2.060> 3
5P(iBu)32.0604
6P(iPr)31.06068
7P(iPr)32.06095
8P(nPr)32.060> 3
a Solutions of RuCl2(p-cymene)(L) and 4a in THF were refluxed for 16 h; b Determined by 1H-NMR analysis of the crude reaction mixture; >97% is indicated when no substrate was detected and <3% when no product was detected.

Share and Cite

MDPI and ACS Style

Jimenez, L.R.; Tolentino, D.R.; Gallon, B.J.; Schrodi, Y. Development of a Method for the Preparation of Ruthenium Indenylidene-Ether Olefin Metathesis Catalysts. Molecules 2012, 17, 5675-5689. https://doi.org/10.3390/molecules17055675

AMA Style

Jimenez LR, Tolentino DR, Gallon BJ, Schrodi Y. Development of a Method for the Preparation of Ruthenium Indenylidene-Ether Olefin Metathesis Catalysts. Molecules. 2012; 17(5):5675-5689. https://doi.org/10.3390/molecules17055675

Chicago/Turabian Style

Jimenez, Leonel R., Daniel R. Tolentino, Benjamin J. Gallon, and Yann Schrodi. 2012. "Development of a Method for the Preparation of Ruthenium Indenylidene-Ether Olefin Metathesis Catalysts" Molecules 17, no. 5: 5675-5689. https://doi.org/10.3390/molecules17055675

Article Metrics

Back to TopTop