Next Article in Journal
A Planar Disk Electrode Chip Based on MWCNT/CS/Pb2+ Ionophore IV Nanomaterial Membrane for Trace Level Pb2+ Detection
Next Article in Special Issue
Biomass Derived N-Doped Porous Carbon Made from Reed Straw for an Enhanced Supercapacitor
Previous Article in Journal
Dispersive Micro-Solid Phase Extraction Using a Graphene Oxide Nanosheet with Neocuproine and Batocuproine for the Preconcentration of Traces of Metal Ions in Food Samples
Previous Article in Special Issue
Microstructure, Physical and Biological Properties, and BSA Binding Investigation of Electrospun Nanofibers Made of Poly(AA-co-ACMO) Copolymer and Polyurethane
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Novel Route to Cationic Palladium(II)–Cyclopentadienyl Complexes Containing Phosphine Ligands and Their Catalytic Activities

by
Dmitry S. Suslov
1,*,
Mikhail V. Bykov
1,
Marina V. Pakhomova
1,
Timur S. Orlov
1,2,
Zorikto D. Abramov
1,
Anastasia V. Suchkova
1,
Igor A. Ushakov
3,
Pavel A. Abramov
4,5 and
Alexander S. Novikov
6,7,*
1
Research Institute of Oil and Coal Chemical Synthesis, Irkutsk State University, ul. K. Marksa, 1, Irkutsk 664003, Russia
2
School of High Technologies, National Research Irkutsk State Technical University, Lermontov St., 83, Irkutsk 664074, Russia
3
A.E. Favorsky Irkutsk Institute of Chemistry SB RAS, Favorsky St., 1, Irkutsk 664033, Russia
4
Nikolaev Institute of Inorganic Chemistry SB RAS, pr-kt Akad. Lavrentieva, 3, Novosibirsk 630090, Russia
5
Research School of Chemistry and Applied Biomedical Sciences, Tomsk Polytechnic University, Tomsk 634034, Russia
6
Institute of Chemistry, Saint Petersburg State University, Universitetskaya Nab., 7/9, Saint Petersburg 199034, Russia
7
Research Institute of Chemistry, Peoples’ Friendship University of Russia (RUDN University), Miklukho-Maklaya St., 6, Moscow 117198, Russia
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(10), 4141; https://doi.org/10.3390/molecules28104141
Submission received: 5 May 2023 / Revised: 12 May 2023 / Accepted: 15 May 2023 / Published: 17 May 2023
(This article belongs to the Special Issue Feature Papers in Applied Chemistry 2.0)

Abstract

:
The Pd(II) complexes [Pd(Cp)(L)n]m[BF4]m were synthesized via the reaction of cationic acetylacetonate complexes with cyclopentadiene in the presence of BF3∙OEt2 (n = 2, m = 1: L = PPh3 (1), P(p-Tol)3, tris(ortho-methoxyphenyl)phosphine (TOMPP), tri-2-furylphosphine, tri-2-thienylphosphine; n = 1, m = 1: L = dppf, dppp (2), dppb (3), 1,5-bis(diphenylphosphino)pentane; n = 1, m = 2 or 3: 1,6-bis(diphenylphosphino)hexane). Complexes 13 were characterized using X-ray diffractometry. The inspection of the crystal structures of the complexes enabled the recognition of (Cp)⋯(Ph-group) and (Cp)⋯(CH2-group) interactions, which are of C–H…π nature. The presence of these interactions was confirmed theoretically via DFT calculations using QTAIM analysis. The intermolecular interactions in the X-ray structures are non-covalent in origin with an estimated energy of 0.3–1.6 kcal/mol. The cationic palladium catalyst precursors with monophosphines were found to be active catalysts for the telomerization of 1,3-butadiene with methanol (TON up to 2.4∙104 mol 1,3-butadiene per mol Pd with chemoselectivity of 82%). Complex [Pd(Cp)(TOMPP)2]BF4 was found to be an efficient catalyst for the polymerization of phenylacetylene (PA) (catalyst activities up to 8.9 × 103 gPA·(molPd·h)−1 were observed)

Graphical Abstract

1. Introduction

The elucidation of ferrocene’s structure in 1952 by Wilkinson, Woodward, and Fischer [1,2] inspired chemists to extensively study transition metal complexes with a cyclopentadienyl (Cp) ligand. This ligand exhibits unique electronic properties and often behaves as a stable three-coordinate spectator ligand [3]. Furthermore, Cp proved to be a good supporting ligand for catalyst design [4]. Although in most cases, industrial application requires the use of heterogeneous, supported catalysts, homogeneous and metal nanoparticle catalysts [5,6,7], which may present an intermediate option between homogeneous and heterogeneous catalysts, continue to evolve [8]. A famous example of this is metallocene catalysts (with titanium-family metals) that have been applied in the industrial production of α-olefin polymers owing to their single reaction active center and high catalytic activity [9,10,11,12].
Although a number of η5-cyclopentadienyl palladium complexes have been reported, they are relatively few compared to those of other platinum-group metals [4,13]. Most of these are neutral palladium(II) species (e.g., see [13,14,15,16,17,18,19,20,21,22,23,24]), including the most well-known [25] (η3-allyl)(η5-cyclopentadienyl)palladium(II)complex [21,22]. Cationic palladium complexes containing the η5-cyclopentadienyl ligand are rare [17,26,27,28,29,30,31,32,33,34,35,36] (Scheme 1), and only six of them have been characterized using X-ray diffraction methods [26,27,28,32,34,35]. Consequently, the chemistry of η5-Cp Pd compounds is not well developed. There is an opinion [19] that this is due to a lack of suitable methods to introduce the Cp ligand to palladium. The main method of synthesis involves the interaction of CpTl with palladium halides. Cyclopentadienylthallium is a highly toxic chemical [37] which is sometimes used in large excess [19]. Another synthetic route, reported by Roberts et al. [35], includes the reaction of cyclopentadiene with the dicationic solvent complexes of palladium to produce monocationic η5-cyclopentadienyl complexes.
In this paper, we present a novel route for the syntheses of cationic η5-cyclopentadienyl palladium(II) complexes, featuring mono- and bidentate tertiary phosphines ligands. This chemistry is summarized in Scheme 2. Ten new compounds have been fully characterized, and the crystal structures of [Pd(Cp)(PPh3)2]BF4, [Pd(Cp)(dppp)]BF4, and [Pd(Cp)(dppb)]BF4 have been determined via X-ray diffraction. NMR spectroscopy features of the prepared cationic complexes are also discussed. Additionally, we disclose our findings on the catalytic activity of the novel complexes in the polymerization of phenylacetylene (PA) and telomerization of butadiene (BD) with methanol.

2. Results and Discussion

2.1. Synthesis of Palladium(II) Complexes

We have found that complexes 110 (Scheme 2) can be prepared through reacting cationic acetylacetonate complexes with cyclopentadiene in the presence of BF3∙OEt2. The starting palladium complexes are easily prepared in high yields from readily available [Pd(acac)(MeCN)2]BF4 [38,39]. Use of methanol as a solvent was found to be more convenient than non-coordinating dichloromethane or 1,2-dichloroethane. For example, the reaction of [Pd(acac)(PPh3)2]BF4 with 1 eq. of BF3∙OEt2 and 1 eq. of CpH in CH2Cl2 led to the formation of desired complex 1 within 1 h. However, in this case, the reaction was complicated by a side reaction. In the IR and 1H NMR spectra of the precipitate containing complex 1, signals that can be attributed to cyclopentadiene/dicyclopentadiene (DCPD) oligomers were observed. As a consequence, the preparation of complex 1 required additional recrystallization of the reaction product from a MeCN/Et2O mixture, which reduced the yield of the complex. The review [40] indicated that DCPD can be oligomerized using initiators of carbocationic polymerization. We carried out the reaction between cyclopentadiene (CpH) and BF3∙OEt2 in the absence of a palladium compound and obtained a white powder. According to IR and 1H NMR data, this can be attributed to a mixture of exo-2,3- and exo-2,7-polyDCPD [41,42]. The side process of oligomerization does not occur in methanol, but a longer reaction time and the use of an excess of CpH and BF3∙OEt2 are required to increase the yield of complex 1. On the other hand, the formation of complex 1 was not observed in more coordinating solvents such as MeCN or DMSO.
The new compounds (110) were characterized using multinuclear and two-dimensional (COSY, NOESY, HMBC, HSQC) NMR, IR, and UV spectroscopy; ESI-MS; and elemental analysis (see Supporting Information File (SI) for details). The obtained complexes appear to be stable solids even in air at ambient temperature, although they decompose slowly in solution. For example, solutions of complexes 110 in acetonitrile or MeCN/toluene mixtures were stable for days at room temperature, while those in MeOH, CH2Cl2, or CHCl3 decomposed within 10 h, forming an unidentified black precipitate.
In all cases, coordination of the phosphines caused a downfield shift Δ(δcomplex − δligand) of the 31P{1H} NMR signals in the range of 34.4 to 55.4 ppm (Table 1). It should be noted that the 31P{1H} and 1H NMR spectra of complex 3 are severely broadened at room temperature (e.g., ΔνP(FWHM) = 365 Hz at δP = 8.2 ppm), indicating that the complex shows dynamic exchange, most likely between “TOMPP-Pd”-rotamers in solution. The NMR signal broadening observed for complex 3 is comparable to that observed for cis-[Pd((1−3η)-but-2-en-1-yl)(TOMPP)2]BF4 [43].
The δ values obtained from 1H and 13C NMR spectra of compounds 110 are consistent with NMR data for known palladium complexes with the same phosphine ligands [38,50,51,52,53,54]. In each case, the 1H NMR signal for the cyclopentadienyl group appears as a triplet (JHP = 2.1–2.3 Hz), due to coupling to the two phosphorus atoms (Table 2). This was additionally confirmed in the phosphorus-decoupled 1H{31P} NMR spectra, where these signals appeared as singlets. The 1H NMR chemical shifts of the C5H5 group were dependent on the nature of the phosphine ligand. The C5H5 resonance in complex 3, for example, appears at 5.12 ppm (shielded by an aromatic ring current) when using two bulky tris(ortho-methoxyphenyl)phosphine ligands (cone angle θ =176° [55]), while the same group signal for complex 4 with non-bulky tris(2-furyl)phosphine was observed at 6.04 ppm. In addition, the 13C{1H} NMR spectra of compounds 19 revealed diagnostic carbon peaks of Cp ligands as triplets at 100–105 ppm (for complex 10, a broadened asymmetric singlet from CCp was observed at 101.5 ppm). In the aromatic region of the 13C{1H} NMR spectra, the signals corresponding to the ipso-C, ortho-C, and meta-C of the PAr-moieties were observed as the expected virtual triplets [56], whereas signals for para-C were still observed as singlets.
The NMR spectra of compounds 9 and 10 suggest that they exist in different isomeric forms in solution. In particular, the 1H and 13C NMR spectra of compound 10 display three distinct sets of resonances in C5H5 group region. The major isomer gave resonances at δ 5.78 (t, J = 2.0 Hz, C5H5) and 101.51 (s, br., C5H5), while the minor isomers were assigned to resonances at δ 5.76–5.69 (m), 5.61 (t, J = 2.0 Hz, C5H5), and 102.18 (s, br., C5H5). The 31P{1H} NMR spectrum of complex 10 displayed a singlet at δ 28.61 for the major isomer, while singlets at δ 28.39–28.22, 27.11 were observed for the minor isomers in a 6:4 intensity ratio. We assume that the isomers observed in solution are due to presence of μ-P-type coordination dimers and oligomers. The ESI-MS isotope distribution data (Figure S76, SI) for compound 10 do not contradict this assumption. The distribution can be interpreted in favor of the presence of bi- and trinuclear palladium complexes in solution. A quantitative calculation of the corresponding intensoids in the enviPat program [57] gave the ratio of dimers and trimers as 4 to 6. From ESI-MS and NMR data for complex 9, mononuclear compounds were mainly found in solution, and the suitable isotope distribution ratio was modeled as [M1]+:[M2]2+:[M3]3+ = 9.5:0.1:0.4. In the case of compounds 18, ESI-MS data confirmed the formation of the expected mononuclear complex cations.
Although the mechanism of synthesis of cationic cyclopentadienyl palladium complexes 110 from acetylacetonate precursors is not known, it is assumed that the process occurs stepwise, with an early step being the reaction of boron trifluoride molecules with the acetylacetonate ligand to produce a κ1-C-acac-Pd species, which then reacts with cyclopentadiene. Evidence supporting this hypothesis has been obtained via examining the stoichiometric reaction of [Pd(κ2-O,O′-acac)(TOMPP)2]BF4 with BF3∙OEt2 (Scheme 3). When [Pd(κ2-O,O′-acac)(TOMPP)2]BF4 is reacted with 2 equivalents of BF3∙OEt2 in the presence of 5 equivalents of MeCN as a stabilizing agent in CH2Cl2, a yellow-orange precipitate is formed. Although this intermediate could not be obtained pure, 1H, 13C, 31P, 19F, and 11B NMR and ESI-MS data (see SI) suggest it is mostly [Pd(κ1-C-acac∙BF3)(MeCN)(TOMPP)2]BF4. The presence of carbon-ligated acac is indicated by the doublet peaks (J = 3.9 Hz) of the methine carbon of the acac ligand at 101.96 ppm in the 13C{1H} NMR spectrum. The 1H NMR spectrum of the residue in CDCl3 shows signals at 2.30 and 6.00 ppm from the methyl and methine protons of the κ1-C-acac ligand and characteristic resonance at 3.81 ppm from the 2-methoxy-group of the TOMPP ligand. The 31P{1H} NMR spectrum shows a singlet from the major product at 32.7 ppm shifted downfield when compared to [Pd(acac)(TOMPP)2]BF4 (17.5 ppm [50]). The obtained precipitate also displays characteristic resonances in the 13C{1H} NMR spectrum for CO and CH3 groups of the κ1-C-acac ligand (at 192.49 and 24.22 ppm). The 19F NMR spectrum shows two major sets of signals with an intensity ratio of approximately 3:4, corresponding to the BF3 and [BF4] structural fragments in the intermediate. Furthermore, upon addition of cyclopentadiene to the intermediate in CH2Cl2, green complex 3 is formed along with BF2(acac), which was detected in the GC-MS spectra of the solution.

2.2. X-ray Crystal Structures of 1, 7, and 8

Single crystals of compounds 1, 7, and 8 suitable for X-ray crystallography were obtained via slow diffusion of toluene into acetonitrile solutions of the complexes. Figure 1 shows the molecular structures of 1, 7, and 8. Selected interatomic distances and angles are given in the figure captions. In all cases, the asymmetric unit of the crystals contains the complex cation, the [BF4] anion, and a solvent molecule. The molecular structures of these complexes reveal that the planes containing the metal and the two-ligand phosphorus atoms are almost perpendicular (88.05°, 89.33°, 88.05° for 1, 7, and 8, respectively) to the Cp ring planes. The alignments of the ML2 planes relative to the Cp rings are typically described as “eclipsed” or “staggered” conformations [27,34], which differ from each other due to a rotation of the cyclopentadienyl ring around the metal-ring centroid vector (Scheme 4). Our study of the [PdCp(PPh3)2]+ cation 1 (Figure 1) reveals that it adopts eclipsed conformation. In the molecular structure of complex 1, the cyclopentadienyl C–C bond lengths radiating from C2 (1.447(3) and 1.417(3) Å) are long, and the bond lengths of the C3–C4–C5–C1 sequence show characteristic [27,34] butadiene-like bond length alternation (1.399(3), 1.454(3), and 1.382(4) Å). The variations in the Pd–C distances are also typical of the pattern expected for an eclipse conformation; thus, the Pd–C2 distance of 2.268(2) Å is much shorter than the Pd1–C1 and Pd1–C3 distances (2.372(2) and 2.329(2) Å, respectively). Structural results for compounds 7 and 8 reveal that cations [PdCp(dppp)]+ and [PdCp(dppb)]+ adopt a geometry that is closer to staggered than eclipsed. For instance, in complex 8, the structural features include three shorter-than-average bond distances (C26–C25 = 1.407(8), C24–C23 = 1.370(8), and C22–C26 = 1.407(7) Å) and two longer bond lengths (C25–C24 = 1.436(7) and C23–C22 = 1.428(7) Å). In addition, Pd1–C22 and Pd1–C25 are shorter than the other metal–carbon bonds (Figure 1). The Pd–P distances and P–Pd–P bond angles in complex 7 (Pd1–P1(a) = 2.244 Å, ∠P1–Pd–P1a = 94.99°) are in good agreement with the corresponding values in complex 8 (Pd1–P1 = 2.263(1) Å, Pd1–P2 = 2.258(1) Å, ∠P1–Pd–P2 = 96.12(4)°), as well as with those previously reported for cationic cyclopentadienyl palladium-diphosphine complexes [26,27,34].
In the crystal packing of all presented complexes C–H…π and π-π contacts play one of the main structure-directed roles. Cp and Ph-groups give a lot of possibilities to realize a whole potential of this kind interactions. While all presented structures are different due to the difference in the contact types provided by the geometry of coordinated phosphines. This aspect plays the most important role in the crystal packing formation for the presented complexes. In the case of PPh3 (1) intramolecular π-π interactions bound Ph-groups of the adjacent phosphine ligands. Intermolecular contacts are of C–H…π nature and involved Ph-groups and Cp ligands of adjacent molecules (Figure 2). In the crystal structure of complex 7, the situation changes dramatically due to the presence of the (CH2)3 linker in the diphosphine ligand. We found strong CH2…π intermolecular contact-directed formation of 1D supramolecular structures (Figure 3a) with suppression of π–π interactions. On the other hand, complex 8 has practically the same (CH2)4 linker (one CH2-group longer) which is practically inactive in the formation of intermolecular interactions. Otherwise π–π and C–H…π contacts strongly direct orientation of Ph-groups of the diphosphine ligands (Figure 3b). This fact can be of electronic nature caused by the linker elongation.
Moreover, H…F contacts between cationic Pd complexes and BF4 anions also present in all presented structures. Some crystal packing projections are summarized in the Supporting Information (Figures S1–S3, SI).
In order to confirm or disprove the hypothesis on the existence of intermolecular interactions C–H⋯π in the X-ray structures 1, 7, and 8 and approximately quantify the strength of these supramolecular contacts from a theoretical viewpoint, DFT calculations followed by a topological analysis of the electron density distribution using the QTAIM approach [58] were carried out (see Computational Details and Table S2 in Supporting Information). Results of QTAIM analysis are summarized in Table 3. The contour line diagram of the Laplacian of electron density distribution ∇2ρ(r), bond paths, and selected zero-flux surfaces as well as a visualization of electron localization function (ELF) and reduced density gradient (RDG) analyses for intermolecular interactions C–H⋯π in the X-ray structure 7 are shown in Figure 4.
QTAIM analysis of model structures demonstrates the presence of bond critical points (3, −1) for intermolecular interactions C–H⋯π in the X-ray structures 1, 7, and 8 (Table 3). The low magnitude of the electron density (0.002–0.009 a.u.), positive values of the Laplacian of electron density (0.008–0.031 a.u.), and zero or very close to zero positive energy density (0.000–0.002 a.u.) in these bond critical points (3, −1) and estimated strength of appropriate short contacts (0.3–1.6 kcal/mol) are typical for very weak noncovalent interactions involving π-systems [59,60,61,62,63,64,65,66,67]. The balance between the Lagrangian kinetic energy G(r) and potential energy density V(r) at the bond critical points (3, −1) (the ratio –G(r)/V(r) ≥1) reveals that a covalent contribution in all intermolecular interactions C–H⋯π in the X-ray structures 1, 7, and 8 is absent [68] (Table 3). The Laplacian of electron density is typically decomposed into the sum of contributions along the three principal axes of maximal variation, giving the three eigenvalues of the Hessian matrix (λ1, λ2 and λ3), and the sign of λ2 can be utilized to distinguish bonding (attractive, λ2 < 0) weak interactions from non-bonding ones (repulsive, λ2 > 0) [69,70]. Thus, intermolecular interactions C–H⋯π in the X-ray structures 1, 7, and 8 are attractive (Table 3).
Table 3. Values of the density of all electrons—ρ(r), Laplacian of electron density—∇2ρ(r) and appropriate λ2 eigenvalues, energy density—Hb, potential energy density—V(r), and Lagrangian kinetic energy—G(r) (a.u.) at the bond critical points (3, −1), corresponding to intermolecular interactions C–H⋯π in the X-ray structures 1, 7, and 8, and estimated strength for these contacts Eint (kcal/mol).
Table 3. Values of the density of all electrons—ρ(r), Laplacian of electron density—∇2ρ(r) and appropriate λ2 eigenvalues, energy density—Hb, potential energy density—V(r), and Lagrangian kinetic energy—G(r) (a.u.) at the bond critical points (3, −1), corresponding to intermolecular interactions C–H⋯π in the X-ray structures 1, 7, and 8, and estimated strength for these contacts Eint (kcal/mol).
Length of Intermolecular Contact in Å *ρ(r)2ρ(r)λ2HbV(r)G(r)Eint **
1
2.6860.0080.030−0.0080.002−0.0040.0061.3
2.8810.0060.020−0.0060.001−0.0030.0040.9
2.7990.0060.018−0.0060.001−0.0020.0030.6
2.9610.0040.015−0.0040.001−0.0020.0030.6
3.3220.0020.008−0.0020.000−0.0010.0010.3
2.9860.0040.016−0.0040.001−0.0020.0030.6
3.0530.0040.015−0.0040.001−0.0020.0030.6
7
3.0340.0040.013−0.0040.001−0.0010.0020.3
2.9660.0040.015−0.0040.001−0.0020.0030.6
3.0770.0040.013−0.0040.001−0.0010.0020.3
3.0340.0040.013−0.0040.001−0.0010.0020.3
3.0770.0040.013−0.0040.001−0.0010.0020.3
3.0280.0050.017−0.0050.001−0.0020.0030.6
8
2.7780.0060.023−0.0060.001−0.0030.0040.9
2.8770.0060.019−0.0060.001−0.0020.0030.6
2.6620.0090.031−0.0090.001−0.0050.0061.6
2.9520.0050.019−0.0050.002−0.0020.0040.6
3.0430.0030.011−0.0030.001−0.0010.0020.3
* The “classic” Bondi’s [71] van der Waals radii for H and C atoms are 1.20 and 1.70 Å, respectively, and “modern” values of van der Waals radii suggested by Alvarez [72] H and C atoms are 1.20 and 1.77 Å, respectively. ** Eint ≈ −V(r)/2 [73].

2.3. Catalytic Studies

The cationic palladium complexes 110 (Scheme 2) were tested in the telomerization of 1,3-butadiene with MeOH (Scheme 5). Palladium-catalyzed telomerization, which involves the linear dimerization of 1,3-dienes with the simultaneous addition of a nucleophile, is an efficient organic transformation that complies with most green chemistry principles [74,75]. The telomerization of 1,3-butadiene (BD) with methanol to produce 1-methoxy-2,7-octadiene as the main product is a commercial route (developed by Dow Chemical) to synthesize 1-octene [76]. The selectivity of the reaction depends on the nature of the nucleophiles and the ligands coordinated to the reactive palladium center (Scheme 5) [74,75,77,78,79]. Previous studies have shown that several palladium complexes with monodentate phosphines[51,80,81,82,83,84,85], diphosphines [86], and NHC (NHC—N-heterocyclic carbene) ligands [87,88,89] are effective catalysts for the telomerization of BD with methanol.
As shown in Table 4, the selectivity and BD conversion largely depended on the nature of the catalyst. The reaction was performed in the presence of an excess of nucleophiles relative to the diene ([MeOH]0:[BD]0 = 1), so turnover number (TON) is based on butadiene conversion. As one can see from Table 4, for phosphine-ligated complexes, TON decreased in the following order: 1, 2 > 3 > 5 > 4610. For triarylphosphines as ligands, this correlates with the decreasing basicity of the phosphine ligand (cf. [90]), except TOMPP, which is more basic than PPh3 [91]. However, the ease of oxidation of the phosphine ligand also increases with its basicity, and therefore, the loss of productivity of the catalyst could be explained by a larger loss of the phosphine ligand through oxidation during the catalysis as proposed by van Leeuwen et al. [76]. Complexes 610 with diphosphines were not suitable for this telomerization reaction. It is known that diphosphines often perform worse than monophosphine catalysts as ligands for the telomerization of dienes with alcohols [76,86]. Next, the most interesting pre-catalyst 1 was studied with lower palladium concentration. When only 0.0021 mol% of Pd loading is used, higher turnovers are obtained (23,700) with practically the same selectivity as in entry 1. It should be noted that at 90 °C, the conversion of butadiene and the selectivity decrease (62% vs. 82% 1-MOD selectivity at 90 °C and 70 °C, respectively), showing that the catalyst formed from compound 1 is unstable at high temperatures. Finally, with a [BD]0/[Pd]0 ratio of 120,000:1 the conversion of BD as well as selectivity substantially dropped, and some amounts of vinylcyclohexene as a result of the Diels–Alder reaction were detected. For compound 1, the catalyst selectivities obtained are similar to those reported by van Leeuwen et al. [84], who used a somewhat similar catalyst system Pd0(dvd)2/2PPh3/5NaOMe (dvd—tetramethyldivinyldisiloxane), or by Beller et al. [80], who utilized conventional Pd(OAc)2/3PPh3/100NEt3, albeit with TON being 30% lower in our case (23,500 vs. 30,000–34,000 reported in [80,84]). Notably, these conversions and selectivities were achieved under solventless conditions and without the use of any added base.
For compound 1, the obtained catalyst TONs and selectivity are as might be expected from results by Beller et al. using catalyst system Pd(OAc)2/3PPh3.
At this point, we decided to test the catalytic activity of compounds 110 towards another type of substrate, and the addition polymerization of phenylacetylene (PA) was examined (Scheme 6, Table 5). Polyacetylenes are used in applications such as photonics, light-emitting diodes, conductors or semi-conductors, sensors, gas separation membranes, and chiral materials [92,93]. Poly(phenylacetylene) (PPA) is mainly produced from catalyzed reactions of phenylacetylene with early- and late-transition metal complexes, driven either via metathesis or insertion polymerization mechanisms [92]. Rhodium-based late-transition metal catalysts are commonly used, but they are associated with a high-cost disadvantage [94,95,96]. Therefore, there is a need to develop other catalytic systems, with Pd-based catalysts attracting most attention [92,93,97,98,99,100]. It has been established [92,93,97] that efficient palladium catalysts for the polymerization of phenylacetylene also fall under the category of cationic organometallic compounds bearing bulky phosphine ligands. The most notable examples are [(dppf)Pd(MeCN)(CH3)]OTf [101], developed by Darkwa, Pollack et al., and [(tBuXPhos)Pd(Me)](BArf) bearing a Buchwald-type dialkylbiarylphosphine ligand, proposed by Chen, Daugulis, and Brookhart [93]. Therefore, we expected that compounds 110 would exhibit interesting catalytic properties during screening.
As shown in Table 5, the polymer yields and molecular weights largely depended on the nature of the catalyst. Typically, low-molecular-weight powdery products were obtained. However, cationic pre-catalyst 3 with a bulky TOMPP ligand produced high-molecular-weight polyphenylacetylene with significant product yield. This is consistent with the known hypothesis that sterically bulky phosphine [93,97,99,103,104] (including TOMPP) or NHC-carbene [105,106] ligands are crucial for catalyst activity in the polymerization of substituted acetylenes. Compared with known Pd(II) catalysts [93,97,99,103,104], catalyst 3, described here, showed moderate to high activity.
In order to achieve more stable reaction conditions using catalyst 3, we investigated the polymerization of PA while varying temperature, ratio of [PA]0:[Pd]0, reaction time, and solvent. Polymerization temperature had a significant effect on catalytic activity and the polydispersity index (PDI) of the resulting polymers. As temperature (entries 11–14, Table 5) increased, yield and activity were enhanced; however, molecular weight and selectivity decreased significantly. Increasing the temperature above 80 °C led to significantly decreased productivity resulting from catalyst deactivation. The thermal stability of compound 3 in the polymerization of phenylacetylene, therefore, was modest. Gel permeation chromatography curves (Figures S95–S100, SI) of the obtained PPAs at 25–50 °C displayed bimodal distribution, which is characteristic of Pd initiators [93,99,101,107,108]. The effect of reaction time on conversion, activity, and molecular weight is summarized in the entries in Table 5. The polymerization was quenched at the respective time using acidified methanol. As shown, conversion increased with time, and approximately the same molecular weight was observed independent of conversion. The PDI and Mn were larger than expected for living polymerization based on monomer weight, which may indicate that catalyst 3 exhibits slow initiation. Polymerization proceeded in all solvents tested (see entries 22–27, Table 5). However, polymerization was substantially slower in acetonitrile, presumably due to its strong coordination to the palladium center, which inhibits coordination with the monomer. The reaction in THF afforded optimal results in terms of molecular weight (Mw = 23.2 kDa), isolated yield (87%), and high TON (435). The PPA samples obtained with catalyst 3 showed a sharp absorption peak in the infrared spectra at 740 cm−1 and a broad peak at 890 cm−1 (Figure S101, SI) characteristic of cis-PPA [102]. The 1H NMR spectra of the polymers also show a sharp singlet at δ 5.86 (s, 1H) due to the vinylic protons in the polymer, and a set of broad peaks at δ 6.72 (m, 2H) and δ 6.98 (m, 3H) ppm, which are associated with a head-to-tail structure of a cis-transoidal PPA (Figure S102, SI) [94,101]. Regarding the formation of the active species with such catalysts, we suggest that the reaction involves the transformation of the η5-cyclopentadienyl ligand to the η1-Cp-isomer. In the next step, insertion of the coordinated PA into the Pd–C bond occurs. Therefore, the 1H NMR spectra of PPA show the presence of a weak peak at 3.6 ppm, indicative of the Cp fragment from the cyclopentadienyl end group in the obtained polyphenylacetylene samples (Figure S102, SI).

3. Materials and Methods

3.1. General Procedures and Materials

All air- and/or moisture-sensitive compounds were manipulated using standard high-vacuum line, Schlenk, or cannula techniques under an argon atmosphere. Argon (Arnika-Prom-Service, Angarsk, Russia) was purified before feeding to the reactor via passing through columns packed with oxygen scavenger and molecular sieve 4A (Aldrich, Steinheim, Germany), respectively. Diethyl ether (99%), petroleum ether, THF (99.9%), and toluene (99.5%) (ZAO Vekton, Saint Petersburg, Russia) were distilled from sodium−benzophenone. CH2Cl2 (99.5%) (ZAO Vekton), CH3CN (99.9%) (ZAO Vekton, Saint Petersburg, Russia), phenylacetylene (98%) (ZAO Vekton, Saint Petersburg, Russia), and methanol (99%) (ZAO Vekton, Saint Petersburg, Russia) were distilled from CaH2. Solvents were stored over molecular sieves. Cyclopentadiene was freshly prepared through a thermal retro-Diels–Alder reaction from commercially obtained dicyclopentadiene (95%) (Acros Organics, Geel, Belgium). Other chemicals were purchased from Acros Organics, Sigma-Aldrich, and ABCR and employed without drying or any further purification. All glassware was dried for at least 2 h in a 150 °C oven and cooled under an argon atmosphere. Pd(acac)2 was synthesized according to a procedure from the literature [109] and recrystallized from acetone. The cationic acetylacetonate complexes were prepared according to known procedures from the literature [38,50,51,52,53,54]. All other reagents were obtained commercially and used as received. NMR spectra were recorded on a Bruker DPX-400 (400 MHz) spectrometer (at room temperature) (Karlsruhe, Germany) and Oxford X-Pulse benchtop (60 MHz) spectrometer (at 40 °C) (Abingdon, Oxon, UK). IR spectra were recorded on a Simex Infralum FT 801 spectrometer (Novosibirsk, Russia). Ultraviolet−visible light (UV-vis) spectra were recorded on an OKB Spectr SF-2000 spectrophotometer (Saint Petersburg, Russia). GC-FID and GC–MS analyses were performed on a Chromatec Crystall 5000.2 (Yoshkar-Ola, Russia) (SGE BPX-5 capillary column, Ringwood, Australia, internal standard benzene) and Shimadzu QP2010 Ultra (Kyoto, Japan) (GSBP-5MS capillary column, Delaware City, DE, USA), respectively. The molecular weights and polydispersity of the polymers were estimated in THF using a Thermo-Dionex HPLC system (Sunnyvale, CA, USA) (Phenomonex, Phenogel Linear column, Torrance, CA, USA). A set of monodisperse polystyrene standards (Agilent, Shropshire, UK) covering the molecular weight range of 103 to 106 was used.

3.2. General Procedure for 1,3-Butadiene Telomerization

All the reactions were carried out in a custom-made stainless-steel autoclave with a volume of 20 mL. The 1,3-butadiene (39 mmol) was first condensed into the reactor and its volume was controlled. The catalyst was then added directly into the reactor pot as a solution in CH2Cl2. After that, methanol (39 mmol) was added to the reactor via a syringe. The reactor was then closed and placed in an oil bath at a temperature of 70 °C. The reaction was stirred magnetically at 700 rpm for 2 h. The reaction was stopped using an ice bath. The catalytic reactions were analyzed using GC-FID, and the response factor for the telomers was determined using the pure fraction of telomers obtained (identified via GC–MS) after vacuum distillation of the reaction mixture.

3.3. General Procedure for Phenylacetylene Polymerization

All polymerizations were conducted in a 5 mL glass reactor equipped with a magnetic stirrer under an Ar atmosphere in an oil bath. The Pd catalyst was stirred in 1 mL of solvent, after which 1 mL of PA (9.1 mmol) was added. The reaction mixture was then stirred for the designated time, and subsequently poured into a large amount of acidified methanol to precipitate the polymer as a yellow powder. The product was isolated via filtration and dried under vacuum to reach a constant weight. If necessary, the polymer was purified through dissolving it in THF and precipitating it with methanol to obtain a fine yellow powder. The polymer samples were characterized using 1H NMR, FT-IR, and GPC analysis. PPA with high cis-content: 1H NMR (CDCl3): δ 6.97 (m, 3H), 6.73 (m, 2H), 5.86 (s, 1H).

3.4. Synthesis of Palladium Complexes

3.4.1. Preparation of (η5-Cyclopentadienyl)bis(triphenylphosphine-κP)palladium(II) Tetrafluoroborate, [Pd(η5-C5H5)(PPh3)2]BF4 (1)

[Pd(acac)(PPh3)2]BF4 (0.400 g, 0.490 mmol) was dissolved in 10 mL of MeOH, and BF3∙OEt2 (0.30 mL, 2.45 mmol) was added to this solution. The obtained yellow solution was stirred for 1.0 h at room temperature, and then cooled to 0 °C. To the resulting reaction mixture, cyclopentadiene (0.22 mL, 2.45 mmol) was added and stirred for 8.0 h at room temperature. The resulting reaction mixture was concentrated to 5 mL under vacuum. Addition of diethyl ether (20 mL) formed a purple precipitate, which was collected, washed with diethyl ether (2 × 10 mL), and dried (8 h) under vacuum to afford complex 1 as a purple powder (327 mg, 85.3%). Anal. Calcd for C41H35BF4P2Pd: C, 62.90; H, 4.51. Found: C, 63.03; H, 4.44. ESI-MS (positive ion mode, MeCN): m/z 695.12 [M+]. 1H NMR (400 MHz, CDCl3, 25 °C): δ 7.45–7.38 (m, 6H, HPh), 7.38–7.28 (m, 24H, HPh), 5.51 (t, J = 2.1 Hz, 5H, HCp).13C{1H} NMR (101 MHz, CDCl3, 25 °C): δ 133.73 (observed as virtual triplets (vt) due to virtual coupling [56,110], J = 6.1 Hz, CPh,ortho), 131.57 (s, CPh,para), 130.65 (vt, 1J(P,Cipso) = 24.5 Hz), 128.97 (vt, J = 5.6 Hz, CPh,meta), 104.27 (t, J = 2.3 Hz, CCp). 13C{1H, gated decoupling mode} NMR (101 MHz, CDCl3, 25 °C): δ 133.73 (d, JCH = 163 Hz, CPh,ortho), 131.57 (d, JCH = 163 Hz, CPh,para), the signals for ipso-C of PPh3 were not clearly observed, 128.97 (d, JCH = 165 Hz, CPh,meta), 104.27 (d, JCH = 178 Hz, CCp). 31P{1H} NMR (162 MHz, CDCl3, 25 °C): δ 33.77 (s). 19F NMR (376 MHz, CDCl3, 25 °C): δ −153.58, −153.64 (intensity ratio of approximately 20:80 corresponding to the natural abundances of 10B and 11B, respectively). IR (KBr, ν ~ , cm−1): 3115, 3075 (sh), 3055 [ν(C–H from CH, Ph and Cp)]; 1585, 1573, 1481, 1436 [νas(C=C, Ph) and phenyl ring deformations]; 1398, 1352 [νas(C=C, Cp) and cyclopentadienyl ring deformations]; 1312, 1284, 1185, 1162, 1122 δ(C–H, Ph and Cp, plane); 1084 δ(C–H, Ph and Cp, plane); 1098(sh), 1062, 1035(sh) νas(B–F); 1026, 999, 928, 831, 752 δ(C–H, Ph and Cp, off-plane); 795 [partially resolved in the IR bands of the symmetric νs(B–F) vibrations, which appear in the spectrum due to violation of the symmetry of the environment]; 703, 695 ν(P–CAr, Ph); 617 (rings deformations). UV-vis (C2H4Cl2, λmax/nm, (εmax/L mol−1 cm−1)): 350 (15,660).

3.4.2. Preparation of (η5-Cyclopentadienyl)bis[tris(4-methylphenyl)phosphine-κP]palladium(II) Tetrafluoroborate, [Pd(η5-C5H5)(P(p-MeC6H4)3)2]BF4 (2)

Preparation of 2 in CH2Cl2 as Solvent

[Pd(acac)(P(p-MeC6H4)3)2]BF4 (0.150 g, 0.167 mmol) was dissolved in 5 mL of CH2Cl2, and to this solution was added 10% v/v solution of BF3∙OEt2 in CH2Cl2 (0.23 mL, 0.183 mmol). The obtained yellow solution was stirred for 0.5 h at room temperature, and then cooled to 0 °C. To the resulting reaction mixture was added cyclopentadiene (0.08 mL, 0.832 mmol) and stirred for 1.0 h at room temperature. The resulting reaction mixture was concentrated to ~3 mL under vacuum. Addition of diethyl ether (10 mL) formed a purple precipitate, which was collected, washed with diethyl ether (2 × 10 mL), dissolved in MeCN (5 mL), filtered and dried (8 h) under vacuum to afford complex 2 as a purple powder (104 mg, 71.5%). Anal. Calcd for C47H47BF4P2Pd: C, 65.11; H, 5.46. Found: C, 65,22; H, 5,58. ESI-MS (positive ion mode, MeCN): m/z 779.22 [M+]. 1H NMR (400 MHz, CD3CN, 25 °C): δ 7.25–7.16 (m, 12H, HPh), 7.16–7.08 (m, 12H, HPh), 5.49 (t, J = 2.1 Hz, 5H, HCp), 2.35 (s, 18H, CH3). 13C{1H} NMR (101 MHz, CD3CN, 25 °C): δ 142.06 (s, CPh,para),133.67 (vt, J = 6.3 Hz, CPh,ortho), 129.24 (vt, J = 5.8 Hz, CPh,meta), 127.79 (vt, 1J(P,Cipso) = 26.5 Hz), 103.71 (t, J = 1.9 Hz, CCp), 20.35 (s, CH3).31P{1H} NMR (162 MHz, CD3CN, 25 °C): δ 32.89 (s). IR (KBr, ν ~ , cm−1): 3110, 3042, 3023 [ν(C–H from CH, Ph and Cp)]; 1597, 1541, 1497, 1448 [νas(C=C, Ph) and phenyl ring deformations]; 1398, 1354 [νas(C=C, Cp) and cyclopentadienyl ring deformations]; 1310, 1281, 1211, 1268, 1191, 1016 δ(C–H, Ph and Cp, plane); 1094, 1055, 1035 νas(B–F); 831, 804, 761 δ(C–H, Ph and Cp, off-plane); 795 [partially resolved in the IR bands of the symmetric νs(B–F) vibrations, which appear in the spectrum due to violation of the symmetry of the environment]; 710 ν(P–CAr, Ph); 615 (rings deformations). UV-vis (C2H4Cl2, λmax/nm, (εmax/L mol−1 cm−1)): 353 (16,200).

Preparation of 2 in MeOH as Solvent

[Pd(acac)(P(p-MeC6H4)3)2]BF4 (0.150 g, 0.167 mmol) was dissolved in 10 mL of MeOH, and BF3∙OEt2 (0.10 mL, 0.84 mmol) was added to this solution. The obtained yellow solution was stirred for 0.25 h at room temperature, and then cooled to 0 °C. To the resulting reaction mixture, cyclopentadiene (0.14 mL, 1.67 mmol) was added and stirred for 8.0 h at room temperature. The resulting reaction mixture was concentrated to ~3 mL under vacuum. Addition of diethyl ether (10 mL) formed a purple precipitate, which was collected, washed with diethyl ether (2 × 10 mL), and dried (8 h) under vacuum to afford complex 2 as a purple powder (120 mg, 82.5%). Anal. Calcd for C47H47BF4P2Pd: C, 65.11; H, 5.46. Found: C, 65.25; H, 5.53. 1H NMR (60 MHz, CD3CN, 40 °C): δ 7.27–7.09 (m, 24H, HPh), 5.50 (t, J = 2.1 Hz, 5H, HCp), 2.36 (s, 18H, CH3).

3.4.3. Preparation of (η5-Cyclopentadienyl)bis[tris(2-methoxyphenyl)phosphine-κP]palladium(II) Tetrafluoroborate, [Pd(η5-C5H5)(TOMPP)2]BF4 (TOMPP = Tris(2-methoxyphenyl)phosphine) (3)

[Pd(acac)(TOMPP)2]BF4 (0.338 g, 0.339 mmol) was dissolved in 10 mL of MeOH, and BF3∙OEt2 (0.20 mL, 1.70 mmol) was added to this solution. The obtained yellow solution was stirred for 0.25 h at room temperature, and then cooled to 0 °C. Cyclopentadiene (0.14 mL, 1.70 mmol) was added to the resulting reaction mixture and stirred for 2.0 h at room temperature. Addition of diethyl ether formed a green precipitate, which was collected, washed with diethyl ether (2 × 10 mL), and dried (8 h) under vacuum to afford complex 3 as a green powder (231 mg, 71.0%). Anal. Calcd for C47H47BF4O6P2Pd: C, 58.62; H, 4.92. Found: C, 59.09; H, 5.15. ESI-MS (positive ion mode, MeCN): m/z 875.19 [M+]. 1H NMR (400 MHz, CD3CN, 25 °C): δ 7.61–7.05 (m, 12H, HPh), 7.02–6.73 (m, 12H, HPh), 5.12 (t, J = 2.2 Hz, 5H, HCp), 3.75–2.84 (m, br., 18H, CH3O). 13C{1H} NMR (101 MHz, CD3CN, 25 °C): δ 159.99 (s, CPh,2-OMe), 132.34 (s, CPh), 119.99 (d, 1J(P,Cipso) =11.5 Hz), 111.64 (s, CPh), 105.38 (t, br., J = 2.2 Hz, CCp), 54.73 (s, CH3O). 31P{1H} NMR (162 MHz, CDCl3, 25 °C): δ 8.22 (br., Δν(FWHM) = 365 Hz). IR (KBr, ν ~ , cm−1): 3064, 3011 (sh) [ν(C–H from CH, Ar and Cp)]; 1586, 1574, 1518, 1457, 1431 [νas(C=C, Ph) and aryl ring deformations]; 1396, 1372 [νas(C=C, Cp) and cyclopentadienyl ring deformations]; 1277, 1134 δ(C–H, Ar and Cp, plane); 1249, 1165 (νAr-O-C); 1180 (δC–H from O-CH3); 1071, 1056, 1018 νas(B–F); 960, 943, 904, 850, 828, 798, 754 δ(C–H, Ar and Cp, off-plane); 795 (partially resolved in the IR bands of the symmetric νs(B–F) vibrations, which appear in the spectrum due to violation of the symmetry of the environment); 686, 668 ν(P–CAr, Ph); 617 (rings deformations). UV-vis (C2H4Cl2, λmax/nm, (εmax/L mol−1 cm−1)): 283(20,700), 370 (11,300).

3.4.4. Preparation of (η5-Cyclopentadienyl)bis[tris(2-furyl)phosphine-κP]palladium(II) Tetrafluoroborate, [Pd(η5-C5H5)(TFP)2]BF4 (TFP = Tris(2-furyl)phosphine) (4)

[Pd(acac)(TFP)2]BF4 (0.173 g, 0.229 mmol) was dissolved in a mixture of 5 mL of MeOH and 2 mL of CH2Cl2, and BF3∙OEt2 (0.14 mL, 1.15 mmol) was added to this solution. The obtained yellow solution was stirred for 0.5 h at room temperature, and then cooled to 0 °C. Cyclopentadiene (0.1 mL, 1.15 mmol) was added to the resulting reaction mixture and stirred for 2.0 h at room temperature. The resulting reaction mixture was concentrated to ~3 mL under vacuum. Addition of n-pentane formed a brown precipitate, which was collected, washed with diethyl ether (2 × 10 mL), and dried (8 h) under vacuum to afford complex 4 as a brown powder (123 mg, 74.1%). Anal. Calcd for C29H23BF4O6P2Pd: C, 48.20; H, 3.21. Found: C, 47.37; H, 3.15. ESI-MS (positive ion mode, MeCN): m/z 635.00 [M+]. 1H NMR (400 MHz, CD3CN, 25 °C): δ 7.79–7.74 (m, 6H, HFur5) (Fur–furyl ring), 6.84–6.78 (m, 6H, HFur3), 6.58–6.48 (m, 6H, HFur4), 6.04 (t, J = 2.3 Hz, 5H, HCp). 13C NMR (101 MHz, CD3CN, 25 °C): δ 151.10 (vt, J = 3.3 Hz, CFur5), 141.95 (vt, 1J(P,C) = 44.6 Hz CFur2), 124.35 (vt, J = 10.2 Hz, CFur3), 111.90 (vt, J = 4.1 Hz, CFur4), 102.68 (t, J = 2.3 Hz, CCp). 31P{1H} NMR (162 MHz, CD3CN, 25 °C): δ −23.99 (s). IR (KBr, ν ~ , cm−1): 3151, 3132, 3089 ν(C–H from Fur and Cp); 1549, 1456, 1368 ν(C=C from Fur) and ring deformations; 1396 (ν(C=C, Cp) and cyclopentadienyl ring deformations); 1284, 1216, 1163, δ(C–H from Fur and Cp, plane) and ring deformations; 1127, 1059, 1034(sh) νas(B–F); 1010 ring deformations and δ(C–H from Fur, plane); 908, 882, 830, 757, 669, 631 δ(C–H, Fur and Cp, off-plane). UV-vis (C2H4Cl2, λmax/nm, (εmax/L mol−1 cm−1)): 349 (14,400).

3.4.5. Preparation of (η5-Cyclopentadienyl)bis[tris(2-thienyl)phosphine-κP]palladium(II) Tetrafluoroborate, [Pd(η5-C5H5)(TTP)2]BF4 (TTP = Tris(2-thienyl)phosphine) (5)

[Pd(acac)(TTP)2]BF4 (0.132 g, 0.154 mmol) was dissolved in 7 mL of MeOH, and cyclopentadiene (0.08 mL, 0.8 mmol) was added to this solution. The obtained solution was stirred for 0.5 h at room temperature, and then cooled to 0 °C. ∙OEt2 (0.1 mL, 0.772 mmol) was added to the resulting reaction mixture and stirred for 72 h at room temperature. The resulting brown precipitate was filtered, washed with diethyl ether (2 × 10 mL), and dried (8 h) under vacuum to afford complex 5 as a purple powder (21 mg, 16.5%). ESI-MS (positive ion mode, MeCN): m/z 732.86 [M+]. 1H NMR (400 MHz, CD3CN, 25 °C): δ 7.88–7.82 (m, 6H, HThi5) (Thi–thienyl ring), 7.38–7.32 (m, 6H, HThi3), 7.14 (t, J = 4.3 Hz, 6H, HThi4), 5.80 (t, J = 2.2 Hz, 5H, HCp). 13C NMR (101 MHz, CD3CN, 25 °C): δ 139.51 (vt, J = 7.2 Hz, CThi5), 136.51 (vt, J = 2.5 Hz, CThi4), 134.13 (vt, J = 29.9 Hz, CThi2), 129.85 (vt, J = 6.9 Hz, CThi3), 105.27 (t, J = 2.4 Hz, CCp). 31P{1H} NMR (162 MHz, CD3CN, 25 °C): δ −1.84 (s). Due to lack of sample EA, FTIR and UV-vis spectra were not obtained.

3.4.6. Preparation of (η5-Cyclopentadienyl)(1,1′-bis(diphenylphosphino)ferrocene-κ2P,P′)palladium(II) Tetrafluoroborate, [Pd(η5-C5H5)(dppf)]BF4 (Dppf = 1,1′-Bis(diphenylphosphino)ferrocene) (6)

[Pd(acac)(dppf)]BF4 (0.300 g, 0.354 mmol) was dissolved in 10 mL of MeOH and BF3∙OEt2 (0.22 mL, 1.77 mmol) was added to this solution. The obtained red solution was stirred for 0.25 h at room temperature, and then cooled to 0 °C. Cyclopentadiene (0.15 mL, 1.77 mmol) was added to the resulting reaction mixture and stirred for 8.0 h at room temperature. The resulting reaction mixture was concentrated to ~3 mL under vacuum. Addition of Et2O formed precipitate, which was collected, washed with diethyl ether (2 × 10 mL), and dried (8 h) under vacuum to afford complex 6 as a dark red powder (180 mg, 62.0%). Anal. Calcd for C39H33BF4FeP2Pd: C, 57.64; H, 4.09. Found: C, 57.05; H, 4.27. ESI-MS (positive ion mode, MeCN): m/z 525.04 [M+]. 1H NMR (400 MHz, CD3CN, 25 °C): δ 7.70–7.60 (m, 12H, HPh), 7.59–7.49 (m, 8H, HPh,meta), 5.45 (t, J = 2.1 Hz, 5H, HCp), 4.59–4.52 (m, 4H, HCp′(dppf),β), 4.39–4.34 (m, 4H, HCp′(dppf),α). 13C NMR (101 MHz, CD3CN, 25 °C): δ 134.04 (vt, J = 24.7 Hz, CPh,ipso), 132.94 (vt, J = 6.7 Hz, CPh,ortho), 131.49 (s, CPh,para), 128.66 (vt, J = 5.7 Hz, CPh,meta), 102.86 (s, CCp), 76.46 (vt, J = 5.8 Hz, CCp′(dppf),α), 74.12 (vt, J = 3.7 Hz, CCp′(dppf),β), 72.43 (vt, J = 33.0 Hz, CCp′(dppf),ipso). 31P{1H} NMR (162 MHz, CD3CN, 25 °C): δ 38.51 (s). IR (KBr, ν ~ , cm−1): 3117, 3100, 3050 (ν(C–H from CH, Ph and Cp)); 1480, 1436 (νas(C=C, Ph) and phenyl ring deformations); 1389, 1355 (νas(C=C, Cp) and cyclopentadienyl ring deformations); 1307, 1282, 1182, 1174, 1167 δ(C–H, Ph and Cp, plane); 1082, 1058, 1037 νas(B–F); 829, 800, 752, δ(C–H, Ph and Cp, off-plane); 696 ν(P–CAr, Ph); 631 (rings deformations). UV-vis (C2H4Cl2, λmax/nm, (εmax/L mol−1 cm−1)): 341 (9000).

3.4.7. Preparation of (η5-Cyclopentadienyl)(1,3-bis(diphenylphosphino)propane-κ2P,P′)palladium(II) Tetrafluoroborate, [Pd(η5-C5H5)(Dppp)]BF4 (Dppp = 1,3-Bis(diphenylphosphino)propane) (7)

[Pd(acac)(dppp)]BF4 (0.100 g, 0.142 mmol) was dissolved in mixture of 5 mL of MeOH and 2 mL of CH2Cl2, and BF3∙OEt2 (0.09 mL, 0.709 mmol) was added to this solution. The obtained yellow solution was stirred for 0.5 h at room temperature, and then cooled to 0 °C. Cyclopentadiene (0.06 mL, 0.709 mmol) was added to the resulting reaction mixture and stirred for 2.0 h at room temperature. The resulting reaction mixture was concentrated to ~3 mL under vacuum. Addition of n-pentane formed a pink precipitate, which was collected, washed with diethyl ether (2 × 10 mL), and dried (8 h) under vacuum to afford complex 7 as a pink powder (71 mg, 74.2%). Anal. Calcd for C32H31BF4P2Pd: C, 57.30; H, 4.66. Found: C, 56.95; H, 4.78. ESI-MS (positive ion mode, MeCN): m/z 583.09 [M+]. 1H NMR (400 MHz, CD3CN, 25 °C): δ 7.61–7.42 (m, 20H, HPh), 5.51 (t, J = 2.1 Hz, 5H, HCp), 2.78–2.57 (m, CH2, 4H), 2.17 (s, CH2, overlapped with impurities in CD3CN). 13C NMR (101 MHz, CD3CN, 25 °C): δ 132.11 (vt, J = 5.9 Hz, CPh,ortho), 131.49 (vt, J = 25.0 Hz, CPh,ipso), 131.19 (s, CPh,para), 128.67 (vt, J = 5.6 Hz, CPh,meta), 100.86 (t, J = 1.9 Hz, CCp), 24.08 (vt, J = 19.8 Hz, CH2), 17.92 (s, CH2). 31P{1H} NMR (162 MHz, CD3CN, 25 °C): δ 16.65 (s). IR (KBr, ν ~ , cm−1): 3107, 3086, 3055 (ν(C–H from CH, Ph and Cp)); 2933, 2910, 2862 (ν(C–H from CH2)); 1586, 1573, 1452, 1483, 1436 (νas(C=C, Ph) and phenyl ring deformations); 1400, 1346 (νas(C=C, Cp) and cyclopentadienyl ring deformations); 1309, 1280, 1187, 1155, 1122(sh), 1100 δ(C–H, Ph and Cp, plane); 1082, 1059, 1036 νas(B–F); 999, 971, 932 δ(C–H, Ph and Cp, off-plane); 911, 874, 857 δ(C–H from CH2, rocking); 831 δ(C–H, Ph and Cp, off-plane); 791 partially resolved in the IR bands of the symmetric νs(B–F) vibrations, which appear in the spectrum due to violation of the symmetry of the environment; 751 δ(C–H, Ph and Cp, off-plane); 698, 673 ν(P–C); 617 (rings deformations). UV-vis (C2H4Cl2, λmax/nm, (εmax/L mol−1 cm−1)): 329 (15,500).

3.4.8. Preparation of (η5-Cyclopentadienyl)(1,4-bis(diphenylphosphino)butane-κ2P,P′)palladium(II) Tetrafluoroborate, [Pd(η5-C5H5)(Dppb)]BF4 (Dppb = 1,4-Bis(diphenylphosphino)butane) (8)

[Pd(acac)(dppb)]BF4 (0.100 g, 0.139 mmol) was dissolved in 5 mL of MeOH, and BF3∙OEt2 (0.09 mL, 0.70 mmol) was added to this solution. The obtained yellow solution was stirred for 0.5 h at room temperature, and then cooled to 0 °C. Cyclopentadiene (0.06 mL, 0.70 mmol) was added to the resulting reaction mixture and stirred for 2.0 h at room temperature. The resulting reaction mixture was concentrated to ~3 mL under vacuum. Addition of diethyl ether formed a purple precipitate, which was collected, washed with diethyl ether (2 × 10 mL), and dried (8 h) under vacuum to afford complex 8 as a purple powder (60 mg, 60.0%). Anal. Calcd for C33H33BF4P2Pd: C, 57.88; H, 4.86. Found: C, 57.36; H, 4.96. ESI-MS (positive ion mode, MeCN): m/z 597.11 [M+]. 1H NMR (400 MHz, CD3CN, 25 °C): δ 7.68–7.49 (m, 20H, HPh), 5.39 (t, J = 2.1 Hz, 5H, HCp), 2.60 (s, br, 4H, CH2), 1.83–1.63 (m, 4H, CH2). 13C NMR (101 MHz, CD3CN, 25 °C): δ 133.98 (vt, J = 24.5 Hz, CPh,ipso), 132.51 (vt, J = 5.8 Hz, CPh,ortho), 131.59 (s, CPh,para), 129.15 (vt, J = 5.4 Hz, CPh,meta), 102.33 (t, J = 1.9 Hz, CCp), 27.40 (vt, J = 16.4 Hz, CH2), 22.99 (s, CH2). 31P{1H} NMR (162 MHz, CD3CN, 25 °C): δ 33.06 (s). IR (KBr, ν ~ , cm−1): 3111, 3086, 3056 ν(C–H from CH, Ph and Cp); 2940, 2923, 2864 ν(C–H from CH2); 1586, 1573, 1483, 1452 (sh.), 1436 (νas(C=C, Ph) and phenyl ring deformations); 1399, 1352 (νas(C=C, Cp) and cyclopentadienyl ring deformations); 1308, 1281, 1232, 1188, 1163 δ(C–H, Ph and Cp, plane); 1099, 1055, 1036 νas(B–F); 1000 δ(C–H, Ph and Cp, off-plane); 895, 877 δ(C–H from CH2, rocking); 832 δ(C–H, Ph and Cp, off-plane); 793 (partially resolved in the IR bands of the symmetric νs(B–F) vibrations, which appear in the spectrum due to violation of the symmetry of the environment); 746 δ(C–H, Ph and Cp, off-plane); 698, 664 ν(P–C); 617 (rings deformations). UV-vis (C2H4Cl2, λmax/nm, (εmax/L mol−1 cm−1)): 334 (13,900).

3.4.9. Preparation of (η5-Cyclopentadienyl)(1,5-bis(diphenylphosphino)pentane-κ2P,P′)palladium(II) Tetrafluoroborate, [Pd(η5-C5H5)(Dpppt)]BF4 (Dpppt = 1,5-Bis(diphenylphosphino)pentane) (9)

[Pd(acac)(dpppt)]BF4 (0.100 g, 0.136 mmol) was dissolved in 6 mL of MeOH, and BF3∙OEt2 (0.08 mL, 0.68 mmol) was added to this solution. The obtained yellow solution was stirred for 0.5 h at room temperature, and then cooled to 0 °C. Cyclopentadiene (0.06 mL, 0.70 mmol) was added to the resulting reaction mixture and stirred for 2.0 h at room temperature. Addition of diethyl ether formed a purple precipitate, which was collected, washed with diethyl ether (2 × 10 mL), and dried (8 h) under vacuum to afford complex 9 as a purple powder (45 mg, 48.0%). Anal. Calcd for C34H35BF4P2Pd: C, 58.44; H, 5.05. Found: C, 59.96; H, 4.95. ESI-MS (positive ion mode, MeCN): m/z 611.12 [M+] the obtained isotopic pattern can be interpreted as [M1]+:[M2]2+:[M3]3+ = 9.5:0.1:0.4. 1H NMR (400 MHz, CD3CN, 25 °C): δ 7.61–7.26 (m, 20H, HPh), 5.79 (t, J = 2.0 Hz, 0.3H, isomer (dimer), HCp), 5.72 (t, J = 2.0 Hz, 0.7H, isomer (trimer), HCp), 5.44 (t, J = 2.0 Hz, 4H, major isomer (monomer), HCp), 2.54 (tt, J = 9.6, 4.8 Hz, 4H, CH2), 2.30–2.10 (m, 2H, CH2, overlapped with impurities in CD3CN) 1.84–1.69 (m, 4H, CH2). 13C NMR (101 MHz, CD3CN, 25 °C): δ 133.57 (vt, J = 23.7 Hz, CPh,ipso), 132.35 (vt, J = 5.7 Hz, CPh,ortho), 131.34 (s, CPh,para), 129.03 (vt, J = 5.4 Hz, CPh,meta), 103.32 (t, J = 2.0 Hz, major isomer, CCp), 103.02 (s, br, minor isomer, CCp), 26.74 (vt, J = 15.8 Hz, P–CH2), 22.56 (t, J = 5.9 Hz, CH2), 22.16 (s, CH2). 13C{1H, gated decoupling mode} NMR (101 MHz, CD3CN, 25 °C): δ 132.35 (d, JCH = 161 Hz, 8C, CPh,ortho), 131.34 (d, JCH = 163 Hz, 4C, CPh,para), 129.03 (d, JCH = 164 Hz, 8C, CPh,meta), the signals for ipso-C of PPh2-moiety were not observed, 103.32 (dt, JCH = 177 Hz, J = 6 Hz, 5C, CCp), 26.74, (td, JCH = 125 Hz, J = 16 Hz, 2C, CH2), 22.56, (t, JCH = 125 Hz, 1C, CH2), 22.16 (t, JCH = 127 Hz, 2C, CH2). 31P{1H} NMR (162 MHz, CD3CN, 25 °C): δ 28.76 (s, 0.04P), 28.25 (s, 0.16P), 19.08 (s, 0.8P). IR (KBr, ν ~ , cm−1): 3107, 3057 ν(C–H from CH, Ph and Cp); 2932, 2863 ν(C–H from CH2); 1586, 1572, 1483, 1453, 1436 (νas(C=C, Ph) and phenyl ring deformations); 1403, 1352 (νas(C=C, Cp) and cyclopentadienyl ring deformations); 1311, 1282, 1187, 1163 δ(C–H, Ph and Cp, plane); 1097, 1056, 1036 νas(B–F); 999 δ(C–H, Ph and Cp, off-plane); 928, 894 δ(C–H from CH2, rocking); 830, 804 δ(C–H, Ph and Cp, off-plane); 792 (partially resolved in the IR bands of the symmetric νs(B–F) vibrations, which appear in the spectrum due to violation of the symmetry of the environment); 745 δ(C–H, Ph and Cp, off-plane); 697, 652ν(P–C); 616(rings deformations). UV-vis (C2H4Cl2, λmax/nm, (εmax/L mol−1 cm−1)): 339 (11,800).

3.4.10. Preparation of {[Pd(η5-C5H5)(Dpphx)]BF4}n (Dpphx = 1,6-Bis(diphenylphosphino)hexane, n = 2, 3) (10) (Mixture of di-μ-(1,6-Bis(diphenylphosphino)hexane-κ2P,P′)-bis[(η5-cyclopentadienyl)palladium(II)] Bis(tetrafluoroborate) and Tri-μ-(1,6-Bis(diphenylphosphino)hexane-κ2P,P′)-tris[(η5-cyclopentadienyl)palladium(II)] Tris(tetrafluoroborate))

[Pd(acac)(dpphx)]n[BF4]n (0.227 g, 0.304 mmol) (n = 2, as previously reported [52]) was dissolved 10 mL of MeOH, and BF3∙OEt2 (0.19 mL, 1.52 mmol) was added to this solution. The obtained yellow solution was stirred for 0.5 h at room temperature, and then cooled to 0 °C. Cyclopentadiene (0.14 mL, 1.52 mmol) was added to the resulting reaction mixture and stirred for 2.0 h at room temperature. Addition of diethyl ether/n-pentane mixture formed a purple precipitate, which was collected, washed with diethyl ether (2 × 10 mL), and dried (8 h) under vacuum to afford complex 10 as a purple powder (178 mg, 82.0%). The product contained Et2O as adduct ([Et2O]:[Pd] ≈ 1:3) via 1H NMR analysis. Anal. Calcd for C35H37BF4P2Pd: C, 58.97; H, 5.23. Found: C, 59.25; H, 5.48. ESI-MS (positive ion mode, MeCN): m/z 625.14 [M+], the obtained isotopic pattern can be interpreted as [M1]+:[M2]2+:[M3]3+ = 0:4:6. 1H NMR (400 MHz, CD3CN, 25 °C): δ 7.72–7.21 (m, 20H, HPh), 5.78 (t, J = 2.0 Hz, 3H, major isomer (trimer), HCp), 5.76–5.69 (m, 1.5H, minor isomer (dimer), HCp), 5.61 (t, J = 2.0 Hz, 0.5H, minor isomer, HCp), 2.34–2.06 (m, 3H, CH2), 1.99–1.46 (m, 3H, CH2, overlapped with impurities in CD3CN), 1.41–1.21 (m, 6H, CH2). 13C NMR (101 MHz, CD3CN, 25 °C): δ (the signals for ipso-C of PPh2-moiety were not observed) 131.93 (vt, J = 5.6 Hz, CPh,ortho), 130.57 (s, CPh,para), 128.11 (s, J CPh,meta), 102.18 (s, minor isomer (dimer), CCp), 101.51 (s, major isomer (trimer), CCp), 29.16 (s, CH2), 25.46 (s, CH2). 31P{1H} NMR (162 MHz, CD3CN, 25 °C): δ 28.61 (s, 0.6P), 28.39–28.22 (m, 0.3P), 27.11(s, 0.1P). IR (KBr, ν ~ , cm−1): 3107, 3055 [ν(C–H from CH, Ph and Cp)]; 2933, 2864 (ν(C–H from CH2)); 1483, 1458, 1436 (νas(C=C, Ph) and phenyl ring deformations); 1401, 1352 (νas(C=C, Cp) and cyclopentadienyl ring deformations); 1312, 1281, 1186, 1161 δ(C–H, Ph and Cp, plane); 1099, 1056, 1036 νas(B–F); 998 δ(C–H, Ph and Cp, off-plane); 929 δ(C–H from CH2, rocking); 831 δ(C–H, Ph and Cp, off-plane); 790 (partially resolved in the IR bands of the symmetric νs(B–F) vibrations, which appear in the spectrum due to violation of the symmetry of the environment); 745 δ(C–H, Ph and Cp, off-plane); 697 ν(P–C); 616(rings deformations). UV-vis (C2H4Cl2, λmax/nm, (εmax/L mol−1 cm−1)): 333 (15,000).

3.5. Reaction of [Pd(κ2-O,O′-Acac)(TOMPP)2]BF4 with BF3∙OEt2

A solution of [Pd(κ2-O,O′-acac)(TOMPP)2]BF4 (100 mg), 5 eq. of MeCN, and 2 eq. of BF3∙OEt2 in 5 mL of CH2Cl2 was stirred at room temperature for 2 h. The resulting yellow solution was concentrated under vacuum and a yellow-orange precipitate was obtained, which was collected and dried under vacuum. Yield: 85 mg. No chemical analyses were completed. ESI-MS (positive ion mode, MeCN): m/z 1023.13. 1H NMR (400 MHz, CDCl3, 25 °C): 7.70–7.40 (m, 6H, HAri), 7.30–7.10 (m, 6H, HAri), 7.10–6.65 (m, 12H, HAri), 6.00 (s, 1H, CH(κ1-C-acac)), 3.81 (s, 15H, HOMe) 3.70–2.10 (m, 3H, HOMe), 2.30 (s, 6H, CH3(κ1-C-acac)), 2.02(s, 2.5H, CH3(MeCN)).13C{1H} NMR (101 MHz, CDCl3, 25 °C) δ 192.50 (s, C=O), 160.21 (s, CAr2), 135.97 (s, CAr6), 134.52 (s, CAr4), 122.56 (s (br.), CAr5,), 113.36 (br., CAr3), 112.20 (d (br.), 1J(P,CAr1) =67.1 Hz, CAr1), 101.96 (d, 2J(P,C) = 3.9 Hz), CH(κ1-C-acac)), 58.76 (s, COMe), 24.21 (s, CH3(κ1-C-acac)), 2.07 (s, MeCN). 31P{1H} NMR (162 MHz, CDCl3, 25 °C): δ 32.72 (s). 19F NMR (376 MHz, CDCl3, 25 °C): δ {−138.58 (0.7F), −138.64 (2.3F); 3F, BF3∙L}, {−152.85 (1.1F), −152.90 (2.9F); 4F, BF4} (intensity ratio ≈ 20:80 corresponds to the natural abundances of 10B and 11B, respectively). 11B NMR (128 MHz, CDCl3, 25 °C): δ 0.56 (1B, BF3∙L), −1.07 (1B, BF4).

3.6. X-ray Crystallographic Studies

Crystallographic data and refinement details are given in Table S1. The diffraction data for complexes 1, 7, and 8 were collected on a Bruker D8 Venture diffractometer with a CMOS PHOTON III detector and IµS 3.0 source (MoKα radiation, λ = 0.71073 Å) at 150 K. The φ- and ω-scan techniques were employed. Absorption correction was applied using SADABS (Bruker Apex3 software suite: Apex3, SADABS-2016/2 and SAINT, version 2018.7-2; Bruker AXS Inc.: Madison, WI, 2017). Structures were solved using SHELXT [111] and refined via full-matrix least-squares treatment against |F|2 in anisotropic approximation with SHELX 2014/7 [112] in the ShelXle program [113]. H-atoms were refined in geometrically calculated positions.
CCDC 2258771, 2258772, and 2258773 contain the supplementary crystallographic data for compounds 1, 7, and 8. These data can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html accessed on 12 May 2023, or from the Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: (+44) 1223-336-033; or e-mail: [email protected].

3.7. Computational Details

The single-point calculations based on the experimental X-ray geometries of compounds 1, 7, and 8 have been carried out at the DFT level of theory (ωB97XD/CEP-121G) with the help of the Gaussian-09 [114] program package. The topological analysis of the electron density distribution has been performed using the Multiwfn program (version 3.7) [115]. The Cartesian atomic coordinates for model structures are presented in Table S2.

4. Conclusions

We have demonstrated that cationic η5-cyclopentadienyl palladium complexes with tertiary phosphorus donor ligands can be prepared through reacting cationic acetylacetonate complexes with cyclopentadiene in the presence of BF3∙OEt2. Ten new arylphosphine-ligated cationic palladium cyclopentadienyl complexes [Pd(Cp)(L)n]m[BF4]m (n = 2, m = 1: L = PPh3, P(p-Tol)3, tris(ortho-methoxyphenyl)phosphine, tri-2-furylphosphine, tri-2-thienylphosphine; n = 1, m = 1: L = dppf, dppp, dppb, 1,5-bis(diphenylphosphino)pentane; n = 1, m = 2 or 3: 1,6-bis(diphenylphosphino)hexane) were synthesized and characterized. Three palladium(II) complexes (1, 7, and 8) have been structurally characterized using X-ray crystallography. In the molecular structures of complexes 1, 7, and 8, the Cp ring is nearly perpendicular to the PdP2 plane. In the cation of complex 1, the cyclopentadienyl ring adopts a nearly eclipsed orientation, while in complexes 7 and 8, it adopts a staggered orientation with respect to the perpendicular PdP2 fragment. Inspection of the crystal structures, as well as DFT calculations using QTAIM analysis, made it possible to recognize (Cp)⋯(Ph-group) and (Cp)⋯(CH2-group) non-covalent interactions, which are of C–H…π nature. The catalytic potential of [Pd(Cp)(L)2][BF4] (L—monodentate phosphine ligands) was demonstrated in the industrially important telomerization of 1,3–butadiene with methanol. Complexes containing PPh3 or P(p-Tol)3 catalyzed the formation of methoxyocta-2,7-dienes with good yields. The other studied palladium complexes showed substantially lower activities. In the presence of 0.0021 mol% palladium loading, the desired telomers were obtained with a chemoselectivity of 82% and TON of up to 2.4∙104 mol BD per mol Pd. We have also demonstrated that complex [Pd(Cp)(TOMPP)2]BF4 is an efficient catalyst for the polymerization of phenylacetylene. Specifically, catalyst activities of (1.1–8.9) × 103 gPA·(molPd·h)−1 were observed in neat monomer or THF as solvent, and obtained PPA featured >90% of cis double bond content.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28104141/s1, spectral data (Figures S4–S94), XRD results (Figures S1–S3, Table S1), Cartesian coordinates for the calculated structures (Table S2), and polymers characterization results (Figures S95–S103).

Author Contributions

Conceptualization, D.S.S. and M.V.B.; investigation, D.S.S., M.V.B., M.V.P., Z.D.A., T.S.O., A.V.S., I.A.U. (NMR), P.A.A. (XRD) and A.S.N. (DFT); writing—original draft preparation, D.S.S.; writing—review and editing, M.V.B., P.A.A., A.S.N. and T.S.O. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by a grant from the Russian Science Foundation No. 22-23-00862, https://rscf.ru/en/project/22-23-00862/ (accessed on 12 May 2023).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials. Additionally, CIFs are openly available in www.ccdc.cam.ac.uk/data_request/cif (accessed on 12 May 2023).

Acknowledgments

We thank A.V. Kuzmin for ESI-MS measurements. We thank D.A. Kraev for experimental support. These studies were performed utilizing equipment from the Baikal Analytical Center of Collective Use SB RAS (http://ckp-rf.ru/ckp/3050/, accessed on 12 May 2023), the Center for Collective Use of Analytical Equipment of Irkutsk State University (http://ckp-rf.ru/ckp/3264/, accessed on 12 May 2023), and the Shared Research Facilities for Physical and Chemical Ultramicroanalysis LIN SB RAS. The quantum chemical calculations were supported by the RUDN University Strategic Academic Leadership Program. P.A.A. thanks Ministry of Science and Higher Education of RF for the access to SCXRD facilities.

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.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Pfab, W.; Fischer, E.O. Zur Kristallstruktur Der Di-Cyclopentadienyl-Verbindungen Des Zweiwertigen Eisens, Kobalts Und Nickels. Z. Anorg. Allg. Chem. 1953, 274, 316–322. [Google Scholar] [CrossRef]
  2. Wilkinson, G.; Rosenblum, M.; Whiting, M.C.; Woodward, R.B. The Structure of Iron Bis-Cyclopentadienyl. J. Am. Chem. Soc. 1952, 74, 2125–2126. [Google Scholar] [CrossRef]
  3. Trost, B.M.; Ryan, M.C. Indenylmetal Catalysis in Organic Synthesis. Angew. Chem. Int. Ed. 2017, 56, 2862–2879. [Google Scholar] [CrossRef] [PubMed]
  4. Kharitonov, V.B.; Muratov, D.V.; Loginov, D.A. Cyclopentadienyl Complexes of Group 9 Metals in the Total Synthesis of Natural Products. Coord. Chem. Rev. 2022, 471, 214744. [Google Scholar] [CrossRef]
  5. Yu, H.; Xu, Y.; Havener, K.; Zhang, M.; Zhang, L.; Wu, W.; Huang, K. Temperature-Controlled Selectivity of Hydrogenation and Hydrodeoxygenation of Biomass by Superhydrophilic Nitrogen/Oxygen Co-Doped Porous Carbon Nanosphere Supported Pd Nanoparticles. Small 2022, 18, 2106893. [Google Scholar] [CrossRef]
  6. Yu, H.; Xu, Y.; Zhang, M.; Zhang, L.; Wu, W.; Huang, K. Magnetically-Separable Cobalt Catalyst Embedded in Metal Nitrate-Promoted Hierarchically Porous N-Doped Carbon Nanospheres for Hydrodeoxygenation of Lignin-Derived Species. Fuel 2023, 331, 125917. [Google Scholar] [CrossRef]
  7. Yu, H.; Zhang, L.; Gao, S.; Wang, H.; He, Z.; Xu, Y.; Huang, K. In Situ Encapsulated Ultrafine Pd Nanoparticles in Nitrogen-Doped Porous Carbon Derived from Hyper-Crosslinked Polymers Effectively Catalyse Hydrogenation. J. Catal. 2021, 396, 342–350. [Google Scholar] [CrossRef]
  8. Van Leeuwen, P.W.N.M.; Chadwick, J.C. Homogeneous Catalysts: Activity, Stability, Deactivation; Wiley-VCH: Weinheim, Germany, 2011; ISBN 978-3-527-32329-6. [Google Scholar]
  9. Kaminsky, W. Discovery of Methylaluminoxane as Cocatalyst for Olefin Polymerization. Macromolecules 2012, 45, 3289–3297. [Google Scholar] [CrossRef]
  10. Kaminsky, W. Polyolefins: 50 Years after Ziegler and Natta II; Kaminsky, W., Ed.; Advances in Polymer Science; Springer: Berlin/Heidelberg, Germany, 2013; Volume 258, ISBN 978-3-642-40804-5. [Google Scholar]
  11. Alt, H.G.; Köppl, A. Effect of the Nature of Metallocene Complexes of Group IV Metals on Their Performance in Catalytic Ethylene and Propylene Polymerization. Chem. Rev. 2000, 100, 1205–1222. [Google Scholar] [CrossRef]
  12. Zhao, Y.; Xu, X.; Wang, Y.; Liu, T.; Li, H.; Zhang, Y.; Wang, L.; Wang, X.; Zhao, S.; Luo, Y. Ancillary Ligand Effects on α-Olefin Polymerization Catalyzed by Zirconium Metallocene: A Computational Study. RSC Adv. 2022, 12, 21111–21121. [Google Scholar] [CrossRef]
  13. Poli, R. Monocyclopentadienyl Halide Complexes of the D- and f-Block Elements. Chem. Rev. 1991, 91, 509–551. [Google Scholar] [CrossRef]
  14. Bielinski, E.A.; Dai, W.; Guard, L.M.; Hazari, N.; Takase, M.K. Synthesis, Properties, and Reactivity of Palladium and Nickel NHC Complexes Supported by Combinations of Allyl, Cyclopentadienyl, and Indenyl Ligands. Organometallics 2013, 32, 4025–4037. [Google Scholar] [CrossRef]
  15. Butler, I.R. Transition Metal Complexes of Cyclopentadienyl Ligands. In Organometallic Chemistry; Royal Society of Chemistry: Cambridge, UK, 2001; pp. 442–478. [Google Scholar]
  16. Chalkley, M.J.; Guard, L.M.; Hazari, N.; Hofmann, P.; Hruszkewycz, D.P.; Schmeier, T.J.; Takase, M.K. Synthesis, Electronic Structure, and Reactivity of Palladium(I) Dimers with Bridging Allyl, Cyclopentadienyl, and Indenyl Ligands. Organometallics 2013, 32, 4223–4238. [Google Scholar] [CrossRef]
  17. Cross, R.J.; Wardle, R. Cyclopentadienyls of Palladium and Platinum. J. Chem. Soc. A Inorg. Phys. Theor. 1971, 2000–2007. [Google Scholar] [CrossRef]
  18. Felkin, H.; Kevin Turner, G. Dimeric Palladium (I) Complexes Containing the Bridging Cyclopentadienyl Group. J. Organomet. Chem. 1977, 129, 429–436. [Google Scholar] [CrossRef]
  19. Grushin, V.V.; Bensimon, C.; Alper, H. Condensation of Cyclopentadiene with Bridging Hydroxo Organopalladium and -Platinum Dimers: A Novel Simple Entry to.Eta.5-Cyclopentadienyl Complexes of Palladium and Platinum. Organometallics 1993, 12, 2737–2740. [Google Scholar] [CrossRef]
  20. Gubin, S.P.; Rubezhov, A.Z.; Winch, B.L.; Nesmeyanov, A.N. Cleavage of the C5H5-Palladium Bond in Cyclopentadienylallylpalladium. Tetrahedron Lett. 1964, 5, 2881–2887. [Google Scholar] [CrossRef]
  21. McClellan, W.R.; Hoehn, H.H.; Cripps, H.N.; Muetterties, E.L.; Howk, B.W. π-Allyl Derivatives of Transition Metals. J. Am. Chem. Soc. 1961, 83, 1601–1607. [Google Scholar] [CrossRef]
  22. Shaw, B.L. Allyl(Cyclopentadienyl)Palladium(II). Proc. Chem. Soc. 1960, 7, 247. [Google Scholar] [CrossRef]
  23. Smidt, J.; Jira, R. Verbindung des Cyclopentadiens mit Palladium. Angew. Chem. 1959, 71, 651. [Google Scholar] [CrossRef]
  24. Werner, H.; Kraus, H.-J.; Schubert, U.; Ackermann, K.; Hofmann, P. Strukturdynamische organometall-komplexe. J. Organomet. Chem. 1983, 250, 517–536. [Google Scholar] [CrossRef]
  25. Torregrosa, R.R.P. (η3-Allyl)(η5-Cyclopentadienyl)Palladium. In Encyclopedia of Reagents for Organic Synthesis; Charette, A., Bode, J., Rovis, T., Shenvi, R., Eds.; John Wiley & Sons, Ltd.: Chichester, UK, 2013; p. rn01566. ISBN 978-0-471-93623-7. [Google Scholar]
  26. Bachechi, F.; Lehmann, R.; Venanzi, L.M. Crystal and Molecular Structure of [Pd(η5-C5H5)(Bis-1,2-Diphenylphosphinoethane)][CF3SO3]. J. Crystallogr. Spectrosc. Res. 1988, 18, 721–728. [Google Scholar] [CrossRef]
  27. Cross, R.J.; Hoyle, R.W.; Kennedy, A.R.; Manojlović-Muir, L.; Muir, K.W. Distortion of Cyclopentadienyl Rings in η5-Cyclopentadienyl-Palladium Complexes: Crystal Structures of [Pd(C5H5)Cl(PMe2Ph)] and [Pd(C5H5)(Ph2PCH2CH2PPh2)][PF6]. J. Organomet. Chem. 1994, 468, 265–271. [Google Scholar] [CrossRef]
  28. Fallis, S.; Rodriguez, L.; Anderson, G.K.; Rath, N.P. Synthesis and Reactions of Cationic Palladium and Platinum Cyclopentadienyl Complexes. Molecular Structure of (H5-Cyclopentadienyl)[1, 2-Bis-(Diphenylphosphino)Ethane]Platinum(II) Triflate. Organometallics 1993, 12, 3851–3855. [Google Scholar] [CrossRef]
  29. Gusev, O.V.; Morozova, L.N.; Peganova, T.A.; Petrovskii, P.V.; Ustynyuk, N.A.; Maitlis, P.M. Synthesis of η5-1,2,3,4,5-Pentamethylcyclopentadienyl-Platinum Complexes. J. Organomet. Chem. 1994, 472, 359–363. [Google Scholar] [CrossRef]
  30. Gusev, O.V.; Morozova, L.N.; Peterleitner, M.G.; Peregudova, S.M.; Petrovskii, P.V.; Ustynyuk, N.A.; Maitlis, P.M. Synthesis of Palladium Cyclopentadienyl Complexes. Decamethylpalladocene Dication [Pd(η5-C5Me5)]2+. J. Organomet. Chem. 1996, 509, 95–99. [Google Scholar] [CrossRef]
  31. Johnson, B.F.G.; Lewis, J.; White, D.A. Reactions of Co-Ordinated Ligands. Part V. Reactions of Triphenylmethyl Tetrafluoroborate and Fluoroboric Acid with a Variety of Enyl Metal Complexes. J. Chem. Soc. A Inorg. Phys. Theor. 1970, 1738–1745. [Google Scholar] [CrossRef]
  32. Kurosawa, H.; Majima, T.; Asada, N. Synthesis, Structures, Stabilities, and Reactions of Cationic Olefin Complexes of Palladium(II) Containing the.Eta.5-Cyclopentadienyl Ligand. J. Am. Chem. Soc. 1980, 102, 6996–7003. [Google Scholar] [CrossRef]
  33. Maitlis, P.M.; Efraty, A.; Games, M.L. Cyclobutadiene-Metal Complexes. J. Organomet. Chem. 1964, 2, 284–286. [Google Scholar] [CrossRef]
  34. Manojlović-Muir, L.; Cross, R.J.; Hoyle, R.W. Structure of (η5-Cyclopentadienyl)Bis(Dimethylphenylphosphine)Palladium(II) Perchlorate. Acta Crystallogr. Sect. C Cryst. Struct. Commun. 1993, 49, 1603–1606. [Google Scholar] [CrossRef]
  35. Roberts, N.K.; Skelton, B.W.; White, A.H.; Wild, S.B. Cationic η5-Cyclopentadienylpalladium(II Complexes. Compounds of Type [Pd(η5-C5H5)L2]PF6 Containing Tertiary Phosphines, Arsines, and Stibines. Crystal and Molecular Structure of [Pd(η5-C5H5)(SbPh3)2]PF6 ·CH2Cl2. J. Chem. Soc. Dalton Trans. 1982, 12, 2093–2097. [Google Scholar] [CrossRef]
  36. White, D.A. Cationic Complexes of Cycloocta-1,5-Diene with Palladium(II) and Platjnum(II). Synth. React. Inorg. Met.-Org. Chem. 1971, 1, 133–139. [Google Scholar] [CrossRef]
  37. Pike, R.D. Thallium: Organometallic Chemistry—Based in Part on the Article Thallium: Organometallic Chemistry by William S. Rees, Jr. & Gertrud Kräuter Which Appeared in the Encyclopedia of Inorganic Chemistry, First Edition. In Encyclopedia of Inorganic and Bioinorganic Chemistry; John Wiley & Sons, Ltd.: Chichester, UK, 2011. [Google Scholar]
  38. Suslov, D.S.; Bykov, M.V.; Abramov, P.A.; Pahomova, M.V.; Ushakov, I.A.; Voronov, V.K.; Tkach, V.S. [Pd(Acac)(MeCN)2]BF4: Air-Tolerant, Activator-Free Catalyst for Alkene Dimerization and Polymerization. RSC Adv. 2015, 5, 104467–104471. [Google Scholar] [CrossRef]
  39. Guibert, I.; Neibecker, D.; Tkatchenko, I. Stoicheiometric Dimerisation of Methyl Acrylate Mediated by Pd(Acac)2·HBF4 Systems and Its Relevance to the Mechanism of Catalytic Dimerisation of Acrylates (Hacac = MeCOCH2COMe). J. Chem. Soc. Chem. Commun. 1989, 1850–1852. [Google Scholar] [CrossRef]
  40. Bermeshev, M.V.; Chapala, P.P. Addition Polymerization of Functionalized Norbornenes as a Powerful Tool for Assembling Molecular Moieties of New Polymers with Versatile Properties. Prog. Polym. Sci. 2018, 84, 1–46. [Google Scholar] [CrossRef]
  41. Mehler, C.; Risse, W. Addition Polymerization of Norbornene Catalyzed by Palladium(2+) Compounds. A Polymerization Reaction with Rare Chain Transfer and Chain Termination. Macromolecules 1992, 25, 4226–4228. [Google Scholar] [CrossRef]
  42. Mehler, C.; Risse, W. Pd(II)-Catalyzed Polymerization of Norbornene Derivatives. Die Makromol. Chem. Rapid Commun. 1992, 13, 455–459. [Google Scholar] [CrossRef]
  43. Hausoul, P.J.C. Palladium-Catalyzed Telomerization of 1,3-Butadiene with Biomass-Based Oxygenates. Ph.D. Thesis, Utrecht University Repository, Utrecht, The Netherlands, 2013. [Google Scholar]
  44. Allman, T.; Goel, R.G. The Basicity of Phosphines. Can. J. Chem. 1982, 60, 716–722. [Google Scholar] [CrossRef]
  45. Hirsivaara, L.; Guerricabeitia, L.; Haukka, M.; Suomalainen, P.; Laitinen, R.H.; Pakkanen, T.A.; Pursiainen, J. M(CO)6 (M=Cr, Mo, W) Derivatives of (o-Anisyl)Diphenylphosphine, Bis(o-Anisyl)Phenylphosphine Tris(o-Anisyl)Phosphine and (p-Anisyl)Bis(o-Anisyl)Phosphine. Inorg. Chim. Acta 2000, 307, 48–57. [Google Scholar] [CrossRef]
  46. Amatore, C.; Jutand, A.; Meyer, G.; Atmani, H.; Khalil, F.; Chahdi, F.O. Comparative Reactivity of Palladium(0) Complexes Generated in Situ in Mixtures of Triphenylphosphine or Tri-2-Furylphosphine and Pd(Dba)2. Organometallics 1998, 17, 2958–2964. [Google Scholar] [CrossRef]
  47. Monkowius, U.; Noga, S.; Schmidbaur, H. Ligand Properties of Tri(2-Thienyl)- and Tri(2-Furyl)Phosphine and -Arsine (2-C4H3E)3P/As (E = O, S) in Gold(I) Complexes Uwe. Z. Nat. 2003, 58b, 751–758. [Google Scholar] [CrossRef]
  48. Nataro, C.; Fosbenner, S.M. Synthesis and Characterization of Transition-Metal Complexes Containing 1,1′-Bis(Diphenylphosphino)Ferrocene. J. Chem. Educ. 2009, 86, 1412. [Google Scholar] [CrossRef]
  49. Dean, P.A.W. Nuclear Magnetic Resonance Studies of the Solvation of Phosphorus(V) Selenides, 1,2-Bis(Diphenylphosphino)Ethane, and Tris(Dimethylamino)Phosphine Telluride by Sulfur Dioxide. Can. J. Chem. 1979, 57, 754–761. [Google Scholar] [CrossRef]
  50. Suslov, D.S.; Bykov, M.V.; Pakhomova, M.V.; Abramov, Z.D.; Ratovskii, G.V.; Ushakov, I.A.; Borodina, T.N.; Smirnov, V.I.; Tkach, V.S. [Pd(Acac)(PR3)(PhCN)][BF4] and [Pd(Acac)(S)2][BF4] (R = Phenyl, 2-Methoxyphenyl; S = Benzonitrile, Pyridine): Synthesis, Characterization, Reactivity and Catalytic Behavior. Crystal Structure of Pd(κ2-O,O′-Acac)(κ1-C-Acac)(P(2-MeOC6H4)3). J. Mol. Struct. 2020, 1217, 128425. [Google Scholar] [CrossRef]
  51. Suslov, D.S.; Bykov, M.V.; Abramov, Z.D.; Ushakov, I.A.; Borodina, T.N.; Smirnov, V.I.; Ratovskii, G.V.; Tkach, V.S. Cationic Palladium(II)–Acetylacetonate Complexes Containing Phosphine and Aminophosphine Ligands and Their Catalytic Activities in Telomerization of 1,3-Butadiene with Methanol. J. Organomet. Chem. 2020, 923, 121413. [Google Scholar] [CrossRef]
  52. Suslov, D.S.; Bykov, M.V.; Pakhomova, M.V.; Abramov, P.A.; Ushakov, I.A.; Tkach, V.S. Cationic Acetylacetonate Palladium Complexes/Boron Trifluoride Etherate Catalyst Systems for Hydroamination of Vinylarenes Using Arylamines. Catal. Commun. 2017, 94, 69–72. [Google Scholar] [CrossRef]
  53. Suslov, D.S.; Bykov, M.V.; Belova, M.V.; Abramov, P.A.; Tkach, V.S. Palladium(II)–Acetylacetonate Complexes Containing Phosphine and Diphosphine Ligands and Their Catalytic Activities in Telomerization of 1,3-Dienes with Diethylamine. J. Organomet. Chem. 2014, 752, 37–43. [Google Scholar] [CrossRef]
  54. Tkach, V.S.; Suslov, D.S.; Myagmarsuren, G.; Ratovskii, G.V.; Rohin, A.V.; Tuczek, F.; Shmidt, F.K. An Effective Route for the Synthesis of Cationic Palladium Complexes of General Formula [(Acac)PdL1L2]+A−. J. Organomet. Chem. 2008, 693, 2069–2073. [Google Scholar] [CrossRef]
  55. Hausoul, P.J.C.; Parvulescu, A.N.; Lutz, M.; Spek, A.L.; Bruijnincx, P.C.A.; Klein Gebbink, R.J.M.; Weckhuysen, B.M. Mechanistic Study of the Pd/TOMPP-Catalyzed Telomerization of 1,3-Butadiene with Biomass-Based Alcohols: On the Reversibility of Phosphine Alkylation. ChemCatChem 2011, 3, 845–852. [Google Scholar] [CrossRef]
  56. De Pater, J.J.M.; Tromp, D.S.; Tooke, D.M.; Spek, A.L.; Deelman, B.-J.; van Koten, G.; Elsevier, C.J. Palladium(0)-Alkene Bis(Triarylphosphine) Complexes as Catalyst Precursors for the Methoxycarbonylation of Styrene. Organometallics 2005, 24, 6411–6419. [Google Scholar] [CrossRef]
  57. Loos, M.; Gerber, C.; Corona, F.; Hollender, J.; Singer, H. Accelerated Isotope Fine Structure Calculation Using Pruned Transition Trees. Anal. Chem. 2015, 87, 5738–5744. [Google Scholar] [CrossRef]
  58. Bader, R.F.W. A Quantum Theory of Molecular Structure and Its Applications. Chem. Rev. 1991, 91, 893–928. [Google Scholar] [CrossRef]
  59. Katkova, S.A.; Mikherdov, A.S.; Kinzhalov, M.A.; Novikov, A.S.; Zolotarev, A.A.; Boyarskiy, V.P.; Kukushkin, V.Y. (Isocyano Group π-Hole)···[d-MII] Interactions of (Isocyanide)[MII] Complexes, in Which Positively Charged Metal Centers (d8-M = Pt, Pd) Act as Nucleophiles. Chem. Eur. J. 2019, 25, 8590–8598. [Google Scholar] [CrossRef]
  60. Abramov, P.A.; Novikov, A.S.; Sokolov, M.N. Interactions of Aromatic Rings in the Crystal Structures of Hybrid Polyoxometalates and Ru Clusters. CrystEngComm 2021, 23, 6409–6417. [Google Scholar] [CrossRef]
  61. Baykov, S.V.; Filimonov, S.I.; Rozhkov, A.V.; Novikov, A.S.; Ananyev, I.V.; Ivanov, D.M.; Kukushkin, V.Y. Reverse Sandwich Structures from Interplay between Lone Pair−π-Hole Atom-Directed C⋯dz2[M] and Halogen Bond Interactions. Cryst. Growth Des. 2020, 20, 995–1008. [Google Scholar] [CrossRef]
  62. Baykov, S.V.; Mikherdov, A.S.; Novikov, A.S.; Geyl, K.K.; Tarasenko, M.V.; Gureev, M.A.; Boyarskiy, V.P. π–π Noncovalent Interaction Involving 1,2,4- and 1,3,4-Oxadiazole Systems: The Combined Experimental, Theoretical, and Database Study. Molecules 2021, 26, 5672. [Google Scholar] [CrossRef]
  63. Ivanov, D.M.; Kirina, Y.V.; Novikov, A.S.; Starova, G.L.; Kukushkin, V.Y. Efficient π-Stacking with Benzene Provides 2D Assembly of Trans-[PtCl2(p-CF3C6H4CN)2]. J. Mol. Struct. 2016, 1104, 19–23. [Google Scholar] [CrossRef]
  64. Lukoyanov, A.N.; Fomenko, I.S.; Gongola, M.I.; Shul’pina, L.S.; Ikonnikov, N.S.; Shul’pin, G.B.; Ketkov, S.Y.; Fukin, G.K.; Rumyantcev, R.V.; Novikov, A.S.; et al. Novel Oxidovanadium Complexes with Redox-Active R-Mian and R-Bian Ligands: Synthesis, Structure, Redox and Catalytic Properties. Molecules 2021, 26, 5706. [Google Scholar] [CrossRef]
  65. Mukhacheva, A.A.; Komarov, V.Y.; Kokovkin, V.V.; Novikov, A.S.; Abramov, P.A.; Sokolov, M.N. Unusual π–π Interactions Directed by the [{(C 6H6)Ru} 2W8O 30(OH)2 ] 6− Hybrid Anion. CrystEngComm 2021, 23, 4125–4135. [Google Scholar] [CrossRef]
  66. Rozhkov, A.V.; Krykova, M.A.; Ivanov, D.M.; Novikov, A.S.; Sinelshchikova, A.A.; Volostnykh, M.V.; Konovalov, M.A.; Grigoriev, M.S.; Gorbunova, Y.G.; Kukushkin, V.Y. Reverse Arene Sandwich Structures Based upon π-Hole⋯[MII] (d 8 M=Pt, Pd) Interactions, Where Positively Charged Metal Centers Play the Role of a Nucleophile. Angew. Chem. Int. Ed. 2019, 58, 4164–4168. [Google Scholar] [CrossRef]
  67. Rozhkov, A.V.; Novikov, A.S.; Ivanov, D.M.; Bolotin, D.S.; Bokach, N.A.; Kukushkin, V.Y. Structure-Directing Weak Interactions with 1,4-Diiodotetrafluorobenzene Convert One-Dimensional Arrays of [MII(acac)2] Species into Three-Dimensional Networks. Cryst. Growth Des. 2018, 18, 3626–3636. [Google Scholar] [CrossRef]
  68. Espinosa, E.; Alkorta, I.; Elguero, J.; Molins, E. From Weak to Strong Interactions: A Comprehensive Analysis of the Topological and Energetic Properties of the Electron Density Distribution Involving X–H⋯F–Y Systems. J. Chem. Phys. 2002, 117, 5529–5542. [Google Scholar] [CrossRef]
  69. Johnson, E.R.; Keinan, S.; Mori-Sánchez, P.; Contreras-García, J.; Cohen, A.J.; Yang, W. Revealing Noncovalent Interactions. J. Am. Chem. Soc. 2010, 132, 6498–6506. [Google Scholar] [CrossRef] [PubMed]
  70. Contreras-García, J.; Johnson, E.R.; Keinan, S.; Chaudret, R.; Piquemal, J.-P.; Beratan, D.N.; Yang, W. NCIPLOT: A Program for Plotting Noncovalent Interaction Regions. J. Chem. Theory Comput. 2011, 7, 625–632. [Google Scholar] [CrossRef] [PubMed]
  71. Bondi, A. Van Der Waals Volumes and Radii of Metals in Covalent Compounds. J. Phys. Chem. 1966, 70, 3006–3007. [Google Scholar] [CrossRef]
  72. Alvarez, S. A Cartography of the van Der Waals Territories. Dalton Trans. 2013, 42, 8617. [Google Scholar] [CrossRef]
  73. Espinosa, E.; Molins, E.; Lecomte, C. Hydrogen Bond Strengths Revealed by Topological Analyses of Experimentally Observed Electron Densities. Chem. Phys. Lett. 1998, 285, 170–173. [Google Scholar] [CrossRef]
  74. Faßbach, T.A.; Vorholt, A.J.; Leitner, W. The Telomerization of 1,3-Dienes—A Reaction Grows Up. ChemCatChem 2019, 11, 1153–1166. [Google Scholar] [CrossRef]
  75. Behr, A.; Becker, M.; Beckmann, T.; Johnen, L.; Leschinski, J.; Reyer, S. Telomerization: Advances and Applications of a Versatile Reaction. Angew. Chem. Int. Ed. Engl. 2009, 48, 3598–3614. [Google Scholar] [CrossRef]
  76. Van Leeuwen, P.W.N.M.; Clément, N.D.; Tschan, M.J.-L. New Processes for the Selective Production of 1-Octene. Coord. Chem. Rev. 2011, 255, 1499–1517. [Google Scholar] [CrossRef]
  77. Benvenuti, F.; Carlini, C.; Marchionna, M.; Patrini, R.; Raspolli Galletti, A.M.; Sbrana, G. Homogeneous Telomerization of 1,3-Butadiene with Alcohols in the Presence of Palladium Catalysts Modified by Hybrid Chelate Ligands. J. Mol. Catal. A Chem. 1999, 140, 139–155. [Google Scholar] [CrossRef]
  78. Patrini, R.; Lami, M.; Marchionna, M.; Benvenuti, F.; Raspolli Galletti, A.M.; Sbrana, G. Selective Synthesis of Octadienyl and Butenyl Ethers via Reaction of 1,3-Butadiene with Alcohols Catalyzed by Homogeneous Palladium Complexes. J. Mol. Catal. A Chem. 1998, 129, 179–189. [Google Scholar] [CrossRef]
  79. Takahashi, S.; Shibano, T.; Hagihara, N. The Dimerization of Butadiene by Palladium Complex Catalysts. Tetrahedron Lett. 1967, 8, 2451–2453. [Google Scholar] [CrossRef]
  80. Vollmüller, F.; Mägerlein, W.; Klein, S.; Krause, J.; Beller, M. Palladium-Catalyzed Reactions for the Synthesis of Fine Chemicals, Highly Efficient Palladium-Catalyzed Telomerization of Butadiene with Methanol. Adv. Synth. Catal. 2001, 343, 29–33. [Google Scholar] [CrossRef]
  81. Mesnager, J.; Kuntz, E.; Pinel, C. Isolated-Palladium Complexes for Catalyzed Telomerization of Butadiene with Methanol in the Presence of Water. J. Organomet. Chem. 2009, 694, 2513–2518. [Google Scholar] [CrossRef]
  82. Hausoul, P.J.C.; Parvulescu, A.N.; Lutz, M.; Spek, A.L.; Bruijnincx, P.C.A.; Weckhuysen, B.M.; Klein Gebbink, R.J.M. Facile Access to Key Reactive Intermediates in the Pd/PR3-Catalyzed Telomerization of 1,3-Butadiene. Angew. Chem. Int. Ed. 2010, 49, 7972–7975. [Google Scholar] [CrossRef]
  83. Vollmüller, F.; Krause, J.; Klein, S.; Mägerlein, W.; Beller, M. Palladium-Catalyzed Reactions for the Synthesis of Fine Chemicals, 14([≠]) Control of Chemo- and Regioselectivity in the Palladium-Catalyzed Telomerization of Butadiene with Methanol—Catalysis and Mechanism. Eur. J. Inorg. Chem. 2000, 8, 1825–1832. [Google Scholar] [CrossRef]
  84. Tschan, M.J.-L.; García-Suárez, E.J.; Freixa, Z.; Launay, H.H.; Hagen, H.; Benet-Buchholz, J.; van Leeuwen, P.W.N.M.; García-Suárez, E.J.; Freixa, Z.; Launay, H.H.; et al. Efficient Bulky Phosphines for the Selective Telomerization of 1,3-Butadiene with Methanol. J. Am. Chem. Soc. 2010, 132, 6463–6473. [Google Scholar] [CrossRef]
  85. Briggs, J.R.; Hagen, H.; Julka, S.; Patton, J.T. Palladium-Catalyzed 1,3-Butadiene Telomerization with Methanol. Improved Catalyst Performance Using Bis-o-Methoxy Substituted Triarylphosphines. J. Organomet. Chem. 2011, 696, 1677–1686. [Google Scholar] [CrossRef]
  86. Tschan, M.J.-L.; López-Valbuena, J.-M.; Freixa, Z.; Launay, H.; Hagen, H.; Benet-Buchholz, J.; van Leeuwen, P.W.N.M. Large P−P Distance Diphosphines and Their Monophosphine Analogues as Ligands in the Palladium-Catalyzed Telomerization of 1,3-Butadiene and Methanol. Organometallics 2011, 30, 792–799. [Google Scholar] [CrossRef]
  87. Jackstell, R.; Harkal, S.; Jiao, H.; Spannenberg, A.; Borgmann, C.; Röttger, D.; Nierlich, F.; Elliot, M.; Niven, S.; Cavell, K.; et al. An Industrially Viable Catalyst System for Palladium-Catalyzed Telomerizations of 1,3-Butadiene with Alcohols. Chem. Eur. J. 2004, 10, 3891–3900. [Google Scholar] [CrossRef] [PubMed]
  88. Jackstell, R.; Frisch, A.; Beller, M.; Röttger, D.; Malaun, M.; Bildstein, B. Efficient Telomerization of 1,3-Butadiene with Alcohols in the Presence of in Situ Generated Palladium(0)Carbene Complexes. J. Mol. Catal. A Chem. 2002, 185, 105–112. [Google Scholar] [CrossRef]
  89. Clement, N.D.; Routaboul, L.; Grotevendt, A.; Jackstell, R.; Beller, M. Development of Palladium-Carbene Catalysts for Telomerization and Dimerization of 1,3-Dienes: From Basic Research to Industrial Applications. Chem. Eur. J. 2008, 14, 7408–7420. [Google Scholar] [CrossRef] [PubMed]
  90. Benvenuti, F.; Carlini, C.; Lami, M.; Marchionna, M.; Patrini, R.; Raspolli Galletti, A.M.; Sbrana, G. Telomerization of 1,3-Butadiene with Alcohols Catalyzed by Homogeneous Palladium(0) Complexes in the Presence of Mono- and Diphosphine Ligands. J. Mol. Catal. A Chem. 1999, 144, 27–40. [Google Scholar] [CrossRef]
  91. Tolman, C.A. Steric Effects of Phosphorus Ligands in Organometallic Chemistry and Homogeneous Catalysis. Chem. Rev. 1977, 77, 313–348. [Google Scholar] [CrossRef]
  92. Darkwa, J. Palladium Catalyzed Phenylacetylene Polymerization to Low Molecular Weight Cis-Transoidal and Trans-Cisoidal Poly(Phenylacetylene)s: A Perspective. Polym. Rev. 2017, 57, 52–64. [Google Scholar] [CrossRef]
  93. Chen, Z.-H.; Daugulis, O.; Brookhart, M. Polymerization of Terminal Acetylenes by a Bulky Monophosphine-Palladium Catalyst. Organometallics 2023, 42, 235–239. [Google Scholar] [CrossRef]
  94. Kishimoto, Y.; Eckerle, P.; Miyatake, T.; Kainosho, M.; Ono, A.; Ikariya, T.; Noyori, R. Well-Controlled Polymerization of Phenylacetylenes with Organorhodium(I) Complexes: Mechanism and Structure of the Polyenes. J. Am. Chem. Soc. 1999, 121, 12035–12044. [Google Scholar] [CrossRef]
  95. Onishi, N.; Shiotsuki, M.; Masuda, T.; Sano, N.; Sanda, F. Polymerization of Phenylacetylenes Using Rhodium Catalysts Coordinated by Norbornadiene Linked to a Phosphino or Amino Group. Organometallics 2013, 32, 846–853. [Google Scholar] [CrossRef]
  96. Shiotsuki, M.; Sanda, F.; Masuda, T. Polymerization of Substituted Acetylenes and Features of the Formed Polymers. Polym. Chem. 2011, 2, 1044–1058. [Google Scholar] [CrossRef]
  97. Huber, J.; Mecking, S. Aqueous Poly(Arylacetylene) Dispersions. Macromolecules 2010, 43, 8718–8723. [Google Scholar] [CrossRef]
  98. Li, K.; Mohlala, M.S.; Segapelo, T.V.; Shumbula, P.M.; Guzei, I.A.; Darkwa, J. Bis(Pyrazole)- and Bis(Pyrazolyl)-Palladium Complexes as Phenylacetylene Polymerization Catalysts. Polyhedron 2008, 27, 1017–1023. [Google Scholar] [CrossRef]
  99. Suslov, D.S.; Pakhomova, M.V.; Bykov, M.V.; Ushakov, I.A.; Tkach, V.S. Polymerization of Phenylacetylene by Cationic Acetylacetonate Palladium Complexes. Catal. Commun. 2019, 119, 16–21. [Google Scholar] [CrossRef]
  100. Shiotsuki, M.; Takahashi, K.; Rodriguez Castanon, J.; Sanda, F. Synthesis of Block Copolymers Using End-Functionalized Polyacetylenes as Macroinitiators. Polym. Chem. 2018, 9, 3855–3863. [Google Scholar] [CrossRef]
  101. Li, K.; Wei, G.; Darkwa, J.; Pollack, S.K. Polymerization of Phenylacetylene Catalyzed by Diphosphinopalladium(II) Complexes. Macromolecules 2002, 35, 4573–4576. [Google Scholar] [CrossRef]
  102. Simionescu, C.I.; Percec, V.; Dumitrescu, S. Polymerization of Acetylenic Derivatives. XXX. Isomers of Polyphenylacetylene. J. Polym. Sci. Polym. Chem. Ed. 1977, 15, 2497–2509. [Google Scholar] [CrossRef]
  103. Castanon, J.R.; Sano, N.; Shiotsuki, M.; Sanda, F. New Approach to the Polymerization of Disubstituted Acetylenes by Bulky Monophosphine-Ligated Palladium Catalysts. ACS Macro Lett. 2014, 3, 51–54. [Google Scholar] [CrossRef]
  104. Rodriguez-Castanon, J.; Murayama, Y.; Sano, N.; Sanda, F. Polymerization of a Disubstituted Acetylene Using Palladium Catalysts. Chem. Lett. 2015, 44, 1200–1201. [Google Scholar] [CrossRef]
  105. Li, M.; Chen, C. Polymerization of Disubstituted Acetylenes by Monodentate NHC-Pd Catalysts. Polym. Chem. 2015, 6, 7127–7132. [Google Scholar] [CrossRef]
  106. Zou, W.; Pang, W.; Chen, C. Redox Control in Palladium Catalyzed Norbornene and Alkyne Polymerization. Inorg. Chem. Front. 2017, 4, 795–800. [Google Scholar] [CrossRef]
  107. Jesus Rodriguez, C.; Kuwata, K.; Shiotsuki, M.; Sanda, F. Synthesis of End-Functionalized Polyacetylenes That Contain Polar Groups by Employing Well-Defined Palladium Catalysts. Chem. Eur. J. 2012, 18, 14085–14093. [Google Scholar] [CrossRef] [PubMed]
  108. Pelagatti, P.; Carcelli, M.; Pelizzi, C.; Costa, M. Polymerisation of Phenylacetylene in Water Catalysed by Pd(NN′O)Cl Complexes. Inorg. Chim. Acta 2003, 342, 323–326. [Google Scholar] [CrossRef]
  109. Yoshida, I.; Kobayashi, H.; Ueno, K. Differential Thermal Analysis of Some Divalent Metal Chelates of 1,5-Dialkylpentane-2,4-Diones. J. Inorg. Nucl. Chem. 1973, 35, 4061–4070. [Google Scholar] [CrossRef]
  110. Klosin, J.; Abboud, K.A.; Jones, W.M. Bis(Triphenylphosphine)Palladium Cycloheptadienynylium Tetrafluoroborate: A Palladium Complex of Tropyne. Organometallics 1996, 15, 2465–2468. [Google Scholar] [CrossRef]
  111. Sheldrick, G.M. SHELXT—Integrated Space-Group and Crystal-Structure Determination. Acta Crystallogr. A 2015, 71, 3–8. [Google Scholar] [CrossRef]
  112. Sheldrick, G.M. Crystal Structure Refinement with SHELXL. Acta Crystallogr. C 2015, 71, 3–8. [Google Scholar] [CrossRef]
  113. Hübschle, C.B.; Sheldrick, G.M.; Dittrich, B. ShelXle: A Qt Graphical User Interface for SHELXL. J. Appl. Crystallogr. 2011, 44, 1281–1284. [Google Scholar] [CrossRef]
  114. Frisch, M.J.; Frisch, G.W.; Trucks, H.B.; Schlegel, G.E.; Scuseria, M.A.; Robb, J.R.; Cheeseman, G.; Scalmani, V.; Barone, B.; Mennucci, G.A.; et al. Gaussian 09, Revision C.01; Gaussian, Inc.: Wallingford, CT, USA, 2010. [Google Scholar]
  115. Lu, T.; Chen, F. Multiwfn: A Multifunctional Wavefunction Analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef]
Scheme 1. Examples of the reported cationic η5-cyclopentadienyl Pd-complexes [17,26,27,28,29,30,31,32,33,34,35,36].
Scheme 1. Examples of the reported cationic η5-cyclopentadienyl Pd-complexes [17,26,27,28,29,30,31,32,33,34,35,36].
Molecules 28 04141 sch001
Scheme 2. Synthesis of new cationic η5-cyclopentadienyl Pd-complexes (110, isolated yields are given in parentheses) with phosphine (L) or diphosphine ligands (L^L).
Scheme 2. Synthesis of new cationic η5-cyclopentadienyl Pd-complexes (110, isolated yields are given in parentheses) with phosphine (L) or diphosphine ligands (L^L).
Molecules 28 04141 sch002
Scheme 3. Plausible route for the synthesis of complex 3. Anion [BF4] is omitted for clarity.
Scheme 3. Plausible route for the synthesis of complex 3. Anion [BF4] is omitted for clarity.
Molecules 28 04141 sch003
Scheme 4. Conformations of ML2 planes with respect to cyclopentadienyl rings in CpML2 complexes: (a) the eclipsed conformation; (b) the staggered conformation.
Scheme 4. Conformations of ML2 planes with respect to cyclopentadienyl rings in CpML2 complexes: (a) the eclipsed conformation; (b) the staggered conformation.
Molecules 28 04141 sch004
Figure 1. Molecular structure of 1∙C6H5CH3 (a), 7∙MeCN (b), and 8∙C6H5CH3 (c) with atoms represented by displacement ellipsoids of 50% probability. Anion [BF4], solvent molecules, and hydrogen atoms are omitted for clarity. Details of the structure refinement are given in Table S1. Selected bond distances (Å) and angles (°): 1: Pd1–P1 = 2.2880(6), Pd1–P2 = 2.2685(6), Pd1–C2 = 2.268(2), Pd1–C1 = 2.372(2), Pd1–C5 = 2.330(2), Pd1–C4 = 2.268(2), Pd1–C3 = 2.329(2), ∠P1–Pd–P2 = 97.87(2); 7: Pd1–P1(a) = 2.244, Pd1–C17(a) = 2.317, Pd1–C16 = 2.254, Pd1–C15 = 2.312, ∠P1–Pd–P1a = 94.99; 8: Pd1–P1 = 2.263(1), Pd1–P2 = 2.258(1), Pd1–C22 = 2.271(4), Pd1–C26 = 2.335(3), Pd1–C25 = 2.270(4), Pd1–C24 = 2.349(4), Pd1–C23 = 2.346(4), ∠P1–Pd–P2 = 96.12(4).
Figure 1. Molecular structure of 1∙C6H5CH3 (a), 7∙MeCN (b), and 8∙C6H5CH3 (c) with atoms represented by displacement ellipsoids of 50% probability. Anion [BF4], solvent molecules, and hydrogen atoms are omitted for clarity. Details of the structure refinement are given in Table S1. Selected bond distances (Å) and angles (°): 1: Pd1–P1 = 2.2880(6), Pd1–P2 = 2.2685(6), Pd1–C2 = 2.268(2), Pd1–C1 = 2.372(2), Pd1–C5 = 2.330(2), Pd1–C4 = 2.268(2), Pd1–C3 = 2.329(2), ∠P1–Pd–P2 = 97.87(2); 7: Pd1–P1(a) = 2.244, Pd1–C17(a) = 2.317, Pd1–C16 = 2.254, Pd1–C15 = 2.312, ∠P1–Pd–P1a = 94.99; 8: Pd1–P1 = 2.263(1), Pd1–P2 = 2.258(1), Pd1–C22 = 2.271(4), Pd1–C26 = 2.335(3), Pd1–C25 = 2.270(4), Pd1–C24 = 2.349(4), Pd1–C23 = 2.346(4), ∠P1–Pd–P2 = 96.12(4).
Molecules 28 04141 g001
Figure 2. C–H…π interactions in the crystal packing of 1.
Figure 2. C–H…π interactions in the crystal packing of 1.
Molecules 28 04141 g002
Figure 3. C–H…π interactions in the crystal packing of complexes 7 (a) and 8 (b).
Figure 3. C–H…π interactions in the crystal packing of complexes 7 (a) and 8 (b).
Molecules 28 04141 g003
Figure 4. Contour line diagram of the Laplacian of electron density distribution ∇2ρ(r), bond paths, and selected zero-flux surfaces (left panel), visualization of electron localization function (ELF, center panel), and reduced density gradient (RDG, right panel) analyses for C–H⋯π interactions in the X-ray structure 7. Bond critical points (3, −1) are shown in blue, nuclear critical points (3, −3)—in pale brown, ring critical points (3, +1)—in orange, and cage critical points (3, +3)—in light green. Bond paths are shown as pale brown lines, length units are in Å, and the color scales for the ELF and RDG maps are presented in a.u.
Figure 4. Contour line diagram of the Laplacian of electron density distribution ∇2ρ(r), bond paths, and selected zero-flux surfaces (left panel), visualization of electron localization function (ELF, center panel), and reduced density gradient (RDG, right panel) analyses for C–H⋯π interactions in the X-ray structure 7. Bond critical points (3, −1) are shown in blue, nuclear critical points (3, −3)—in pale brown, ring critical points (3, +1)—in orange, and cage critical points (3, +3)—in light green. Bond paths are shown as pale brown lines, length units are in Å, and the color scales for the ELF and RDG maps are presented in a.u.
Molecules 28 04141 g004
Scheme 5. Telomerization of 1,3-butadiene with methanol.
Scheme 5. Telomerization of 1,3-butadiene with methanol.
Molecules 28 04141 sch005
Scheme 6. Polymerization of phenylacetylene with Pd catalysts 110.
Scheme 6. Polymerization of phenylacetylene with Pd catalysts 110.
Molecules 28 04141 sch006
Table 1. 31P{1H} chemical shifts of the cationic phosphine-Pd-Cp complexes and coordination-induced shifts Δ for the complexes.
Table 1. 31P{1H} chemical shifts of the cationic phosphine-Pd-Cp complexes and coordination-induced shifts Δ for the complexes.
Free Ligandδ 31P{1H} Chemical Shift, ppmComplexδ 31P{1H} Chemical
Shift, ppm
Δ(31P{1H})
complex − δligand), ppm
PPh3−6.3 [44]133.840.1
P(p-Tol)3−8.3 [44]232.941.2
TOMPP−37.9 [45]38.246.1
TFP−75.2 [46]4−24.051.2
TTP−45.6 [47]5−1.843.8
dppf−16.8 [48]638.655.4
dppp−17.7 [49]716.734.4
dppb−16.3 [49]833.149.4
dppent−16.5 [49]919.135.6
dpphex−16.4 [49]1028.645.0
Table 2. Selected 1H chemical shift values (δ, ppm) for the cationic palladium complexes (400 MHz CD3CN, 25 °C).
Table 2. Selected 1H chemical shift values (δ, ppm) for the cationic palladium complexes (400 MHz CD3CN, 25 °C).
ComplexCpPhosphine Ligand
1a5.51 (t, J = 2.1 Hz, 5H)7.45–7.38 (m, 6H, HPh), 7.38 –7.28 (m, 24H, HPh)
25.49 (t, J = 2.1 Hz, 5H)7.25–7.16 (m, 12H, HPh), 7.16–7.08 (m, 12H, HPh), 2.35 (s, 18H, CH3)
35.12 (t, J = 2.2 Hz, 5H)7.61–7.05 (m, 12H, HPh), 7.02–6.73 (m, 12H, HPh), 3.75–2.84 (m, br., 18H, CH3O)
46.04 (t, J = 2.3 Hz, 5H)7.79–7.74 (m, 6H, HFur5) (Fur = furyl), 6.84–6.78 (m, 6H, HFur3), 6.58–6.48 (m, 6H, HFur4)
55.80 (t, J = 2.2 Hz, 5H)7.88–7.82 (m, 6H, HThi5) (Thi = thienyl), 7.38–7.32 (m, 6H, HThi3), 7.14 (t, J = 4.3 Γц, 6H, HThi4)
65.45 (t, J = 2.1 Hz, 5H)7.70–7.60 (m, 12H, HPh), 7.59–7.49 (m, 8H, HPh,meta), 4.59–4.52 (m, 4H, HCp′(dppf),β), 4.39–4.34 (m, 4H, HCp′(dppf),α).
75.51 (t, J = 2.1 Hz, 5H)7.61–7.42 (m, 20H, HPh), 2.78–2.57 (m, CH2, 4H), 2.17 (s, CH2).
85.39 (t, J = 2.1 Hz, 5H)7.68–7.49 (m, 20H, HPh), 2.60 (s, br, 4H, CH2), 1.83–1.63 (m, 4H, CH2).
95.72 (t, J = 2.0 Hz, 0.7H, minor isomer), 5.44 (t, J = 2.0 Hz, 4H, major isomer),7.61–7.26 (m, 20H, HPh), 2.54 (tt, J = 9.6, 4.8 Γц, 4H, CH2), 2.30–2.10 (m, 2H, CH2) 1.84–1.69 (m, 4H, CH2).
105.78 (t, J = 2.0 Hz, 3H, major isomer), 5.76–5.69 (m, 1.5H, minor isomer), 7.72–7.21 (m, 20H, HPh), 2.34–2.06 (m, 3H, CH2), 1.99–1.46 (m, 3H, CH2), 1.41–1.21 (m, 6H, CH2).
a 400 MHz, CDCl3, 25 °C.
Table 4. Influence of the nature of catalysts 110 on conversion and selectivity in telomerization of 1,3-butadiene with methanol.
Table 4. Influence of the nature of catalysts 110 on conversion and selectivity in telomerization of 1,3-butadiene with methanol.
Entry [a]PdConv. BD % [b]Selectivity, mol%Chemo,mol% [c]n/iso [d]E/Z [e]TON [f]
OCT3-MOD1-MOD
1167.413.55.081.586.516108100
4266.813.05.681.587.115108000
5325.94.73.392.095.328183100
648.74.84.490.895.221121050
7518.05.04.990.195.018102150
860.3
970.2
1080.3
1190.3
12100.2
13 [g]149.212.84.981.586.4171211,800
14 [h]149.316.64.577.882.3171123,700
15 [i]13.511.44.569.974.415104200
[a] Reaction conditions: t = 70 °C, 0.0083 mol% of Pd, [BD]0/[Pd]0 = 12,000, [MeOH]0/[Bd]0 = 1, nPd = 3.25 μmol, reaction time—2 h. [b] 1,3-Butadiene conversion. [c] Chemoselectivity = (1-MOD + 3-MOD)(1-MOD + 3-MOD + OCT + VCH)−1. [d] Regioselectivity = (1-MOD)(3-MOD)−1. [e] Stereoselectivity = (trans-1-MOD)(cis-1-MOD)−1. [f] In units of (mol of BD) (mol of Pd)−1. [g] nPd = 1.63 μmol, [BD]0/[Pd]0 = 24,000. [h] nPd = 1.63 μmol, [BD]0/[Pd]0 = 48,000, reaction time—5 h. [i] nPd = 0.33 μmol, [BD]0/[Pd]0 = 120,000, reaction time—8 h.
Table 5. Polymerization of PA with Pd catalysts 110.
Table 5. Polymerization of PA with Pd catalysts 110.
Entry [a]ComplexTime
(h)
Solventt
(°C)
Yield
(%)
TON [b]Activity [c]Sel [d] (%)Cis [e] (%)Polymer
Mw
(/103) [f]
Mw/
Mn[f]
1124250.331363902.61.6
2224250.22965942.21.5
33242526.22621113789417.22.1
442425000
552425000
6624251.31355929219.02.1
772425000
882425000
992425000
10102425000
11 [h]32442.71460>999021.42.1
12 [h]3242553.42671140799415.52.0
13 [h]3245091.3457194047987.91.6
14 [h]3246084.8424180038982.01.5
15 [h]31252.211121999n.d.16.52.1
16 [h]32255.2261320999417.91.9
17 [h]33257.0351120999517.31.9
18 [h]342512.3621580999419.12.0
19 [h]352525.21262570989516.22.0
20 [h]3242553.92701150799413.42.0
21 [h]3482557.0285605679616.42.0
22 [h]35pentane250.4240958714.92.4
23 [h]35benzene252.01020098908.51.9
24 [h]35THF2587.04358890949223.22.6
25 [h]35MeCN254.02041096902.31.4
26 [h]35CH2Cl22519.3971980949016.12.2
27 [h]35DCE [g]2525.61282610908718.32.2
[a] Reaction conditions: 0.1 mol% of Pd, [PA]0/[Pd]0 = 1000, nPd = 9.1 μmol, VPA = 1.0 mL. [b] In units of (mol of PA) (mol of Pd)−1. [c] In units of (g of polymer) (mol of Pd)−1 h−1. [d] Selectivity to polymers, estimated via GPC using peak-area ratio (response factors were determined using polymer obtained in entry 1 and oligomer obtained using BF3∙OEt2 as initiator). [e] The percentages of cis sequences determined as prescribed in ref [102]. [f] Estimated via GPC in THF (polystyrene standard). [g] DCE—1,2-dichloroethane. [h] Reaction conditions: 0.2 mol% of Pd, [PA]0/[Pd]0 = 500, nPd = 18.2 μmol, VPA = 1.0 mL; Vsolv. = 1 mL.
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

Suslov, D.S.; Bykov, M.V.; Pakhomova, M.V.; Orlov, T.S.; Abramov, Z.D.; Suchkova, A.V.; Ushakov, I.A.; Abramov, P.A.; Novikov, A.S. Novel Route to Cationic Palladium(II)–Cyclopentadienyl Complexes Containing Phosphine Ligands and Their Catalytic Activities. Molecules 2023, 28, 4141. https://doi.org/10.3390/molecules28104141

AMA Style

Suslov DS, Bykov MV, Pakhomova MV, Orlov TS, Abramov ZD, Suchkova AV, Ushakov IA, Abramov PA, Novikov AS. Novel Route to Cationic Palladium(II)–Cyclopentadienyl Complexes Containing Phosphine Ligands and Their Catalytic Activities. Molecules. 2023; 28(10):4141. https://doi.org/10.3390/molecules28104141

Chicago/Turabian Style

Suslov, Dmitry S., Mikhail V. Bykov, Marina V. Pakhomova, Timur S. Orlov, Zorikto D. Abramov, Anastasia V. Suchkova, Igor A. Ushakov, Pavel A. Abramov, and Alexander S. Novikov. 2023. "Novel Route to Cationic Palladium(II)–Cyclopentadienyl Complexes Containing Phosphine Ligands and Their Catalytic Activities" Molecules 28, no. 10: 4141. https://doi.org/10.3390/molecules28104141

Article Metrics

Back to TopTop