Next Article in Journal
Probing the Nature of Chemical Bonds by Atomic Force Microscopy
Next Article in Special Issue
Iridaaromatics via Methoxy(alkenyl)carbeneiridium Complexes
Previous Article in Journal
Design, Synthesis and Biological Evaluation of Novel Pyrazolo[1,2,4]triazolopyrimidine Derivatives as Potential Anticancer Agents
Previous Article in Special Issue
1,1′-Biisoquinolines—Neglected Ligands in the Heterocyclic Diimine Family That Provoke Stereochemical Reflections
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Some Novel Cobalt Diphenylphosphine Complexes: Synthesis, Characterization, and Behavior in the Polymerization of 1,3-Butadiene

1
CNR—Istituto di Scienze e Tecnologie Chimiche “Giulio Natta” (SCITEC), Via Corti 12, I-20133 Milano, Italy
2
Versalis S.p.A.—Centro Ricerca Novara, Via Fauser 4, I-28100 Novara, Italy
3
Versalis S.p.A.—R&D Partner Catalysis, Piazza Boldrini 1, San Donato Milanese, I-20097 Milano, Italy
4
Dipartimento di Chimica Industriale ‘‘Toso Montanari’’, Università di Bologna, Viale Risorgimento 4, I-40136 Bologna, Italy
5
Dipartimento di Chimica e Chimica Industriale, Università di Pisa, Via Moruzzi 13, I-56124 Pisa, Italy
*
Author to whom correspondence should be addressed.
Current address: CNR—Istituto di Scienze e Tecnologie Chimiche “Giulio Natta” (SCITEC), Via Golgi 19, I-20133 Milano, Italy.
Molecules 2021, 26(13), 4067; https://doi.org/10.3390/molecules26134067
Submission received: 8 June 2021 / Revised: 23 June 2021 / Accepted: 25 June 2021 / Published: 2 July 2021
(This article belongs to the Special Issue Exclusive Feature Papers in Organometallic Chemistry)

Abstract

:
Some novel cobalt diphenylphosphine complexes were synthesized by reacting cobalt(II) chloride with (2-methoxyethyl)diphenylphosphine, (2-methoxyphenyl)diphenylphosphine, and 2-(1,1-dimethylpropyl)-6-(diphenylphosphino)pyridine. Single crystals suitable for X-ray diffraction studies were obtained for the first two complexes, and their crystal structure was determined. The novel compounds were then used in association with methylaluminoxane (MAO) for the polymerization of 1,3-butadiene, and their behavior was compared with that exhibited in the polymerization of the same monomer by the systems CoCl2(PnPrPh2)2/MAO and CoCl2(PPh3)2/MAO. Some significant differences were observed depending on the MAO/Co ratio used, and a plausible interpretation for such a different behavior is proposed.

Graphical Abstract

1. Introduction

Catalyst systems obtained by combining CoCl2(PRPh2)2 (R = methyl, ethyl, allyl, normal-propyl, iso-propyl, tert-butyl, (S)-(+)neomenthyl, cyclohexyl and phenyl) with MAO are well known to be able to give predominantly 1,2 polymers (~85%) from 1,3-butadiene [1,2,3]. The tacticity of the 1,2 polymers was found to depend on the type of phosphine bonded to the cobalt atom, the syndiotacticity degree generally increasing when increasing the steric hindrance of the phosphine ligand.
According to the diene polymerization mechanism previously proposed [4], the formation of 1,2 polymers was attributed to the presence of a phosphine ligand on the cobalt atom [5,6,7]; it is in fact worthwhile to note that naked CoCl2, in combination with MAO, provides a poly(1,3-butadiene) with a very high cis content (~97%). The formation of a certain amount of cis-1,4 units (~15%) was attributed, in our opinion, to the fact that some phosphine during the polymerization process may be removed from the cobalt atom by the large excess of MAO; therefore, some active sites such as those originating from CoCl2/MAO, specific for cis-1,4 poly(1,3-butadiene), can be formed.
To check this hypothesis, we have now synthesized some novel cobalt diphenylphosphine complexes by reacting CoCl2 with (2-methoxyethyl)diphenylphosphine, (2-methoxyphenyl) diphenylphosphine, and 2-(1,1-dimethylpropyl)-6-(diphenylphosphino)pyridine (Figure 1), that is ligands which contain within their structure heteroatoms such as O and N and have some affinity for aluminum. We have examined their behavior in the polymerization of 1,3-butadiene as the MAO/Co ratio varies, in comparison with that of the systems CoCl2(PRPh2)2/MAO (R = Ph, nPr). Furthermore, we also examined in more detail the influence of MAO/Co molar ratio on the polymerization of 1,3-butadiene with CoCl2(PRPh2)2/MAO (R = Ph, nPr), even at very high MAO/Co ratios (up to 5000). The results obtained are reported in the present paper, together with a plausible interpretation.

2. Results and Discussion

2.1. Synthesis and Characterization of Cobalt Complexes

The novel diphenylphosphine cobalt complexes, CoCl2(Ph2PCH2CH2OCH3)2 (1), CoCl2(Ph2P(C6H4OCH3))2 (2), and CoCl2(Ph2P(C5NH3(C5H11))2 (3) (Figure 1) were prepared according to a general experimental procedure already reported in the literature [8]. CoCl2 was reacted with an excess of phosphine ligands using ethanol as a solvent.
Crystalline products were obtained by cooling a toluene solution at low temperature and by a continuous extraction of the reaction products with boiling pentane. The structures of 1 and 2 have been determined by single-crystal X-ray diffraction (Figure 2 and Figure 3).
In the case of complex 3, the presence of highly disordered cocrystallized solvent (diethyl ether) prevented obtaining an acceptable structure, though X-ray data confirmed the formation of the compound, as shown in Figure 1.
In both complexes, the Co(II) center displayed a distorted tetrahedral geometrical coordination, being bonded to two chlorides and two phosphine ligands. The Co–Cl and Co–P bond distances (Table 1 and Table 2) were in the ranges reported for analogue CoCl2(PRPh2)2 complexes (R = ethyl [2], normal-propyl [2], iso-propyl [1], tert-butyl [3], CH2CH(OCH3)2 [9], CH2C(O)Ph [10], and Ph [11,12]), i.e., 2.21−2.24 and 2.36−2.43 Å, respectively. As for the P–C bonds, a distinction should be made according to the aliphatic/aromatic nature of the bonded carbon atom. In fact, by considering average values within each structure, P–Caliphatic bonds (varying from 1.827 [2] to 1.888 Å [3]) were systematically longer than P–Caromatic ones (ranging from 1.812 [10] to 1.833 Å, the latter value observed in 1). In particular, as already pointed out in our previous analysis on CoCl2(PRPh2)2 complexes [3], the greater the steric hindrance of the aliphatic group, the larger the difference in P–C bond lengths. For compound 1, only a little difference was observed, the average P–Caliphatic bond distance, 1.838 Å, being only slightly longer than the average P–Caromatic one, 1.833 Å. In the case of 2, only P–Caromatic bonds were present, measuring on average 1.823 Å and virtually identical to those of the CoCl2(PPh3)2 structure [11,12]. In the solid state, several short intermolecular contacts were present, including CH2-O·HCHP (in 1), Cphenyl·Hphenyl and Cphenyl·HCH2 (in 2), and Co-Cl·Hphenyl (in both structures) hydrogen bonds as well as normal van der Waals distances.

2.2. Polymerization of 1,3-Butadiene

The results concerning the polymerization of 1,3-butadiene with the catalysts obtained by combining the three novel diphenylphosphine complexes 13 with MAO are shown in Table 3; the results obtained with CoCl2(PPh3)2/MAO and CoCl2(PnPrPh2)2/MAO are reported for comparison.
Catalysts based on the new cobalt complexes 1–3 were much less active than the systems CoCl2(PnPrPh2)2/MAO and CoCl2(PPh3)2/MAO, and their activity seemed to decrease with decreasing the MAO/Co molar ratio. The polybutadienes had a molecular weight in the range 100,000–200,000 g·mol−1 and a molecular weight distribution around 2–3, values quite similar to those of the polymers obtained with CoCl2(PnPrPh2)2/MAO and CoCl2(PPh3)2/MAO.
Concerning the selectivity, all five catalytic systems, the three new ones and the two old ones, showed a very similar behavior operating at a MAO/Co molar ratio of 100 or lower, providing poly(1,3-butadiene)s with a predominantly 1,2 structure. A completely different behavior was instead observed at MAO/Co ratio of 1000: CoCl2(PPh3)2/MAO and CoCl2(PnPrPh2)2/MAO still gave predominantly 1,2 poly(1,3-butadiene)s, whereas the catalytic systems 1/MAO, 2/MAO, and 3/MAO gave highly cis-1,4 poly(1,3-butadiene)s.
Taking into consideration that the CoCl2/MAO system produces from 1,3-butadiene a polymer with a cis content of about 97%, we may formulate the following working hypothesis to justify the different behavior exhibited by the above catalysts by varying the MAO/Co molar ratio.
In the polymerization of 1,3-butadiene with cobalt phosphine complex-based catalysts, the structure of the active site, as reported in our previous papers [1,2,3,5,13], is that shown in Figure 4A, with only one phosphine ligand on the cobalt atom, the incoming monomer cis4 coordinated, and the growing chain bonded to the cobalt atom through a η3−allyl bond. Most likely, a sort of equilibrium between cobalt and aluminum (MAO) was established (Figure 4), so that the phosphine ligand, initially on the cobalt atom (Figure 4A), may migrate onto the aluminum atom (Figure 4B), causing a drastic change in the selectivity of the catalytic center from specific 1,2 to specific cis-1,4. Notably, this equilibrium was more displaced towards cobalt or aluminum according to (i) the MAO/Co molar ratio value and (ii) the affinity level of the ligand for cobalt or aluminum, strongly affected by the presence of heteroatoms on the phosphine ligand.
At low MAO/Co molar ratios (≤100), the equilibrium mentioned above clearly moved towards the cobalt atom (Figure 4A), with the phosphine mainly on the cobalt atom, and almost exclusively 1,2 units formed; some migration of the phosphine however may take place, with formation of a small amount of cis-1,4 units.
When increasing the MAO/Co molar ratio to 1000, the situation did not seem to change for the CoCl2(PPh3)2/MAO and CoCl2(PnPrPh2)2/MAO systems; that is, the phosphine ligand remained coordinated to the cobalt atom, while the same was not valid for the 1/MAO, 2/MAO, and 3/MAO systems, which provide highly cis-1,4 poly(1,3-butadienes). In this case, it is likely that the presence of nitrogen or oxygen donor atoms within the phosphine ligand structure, associated with the high concentration of MAO, caused the phosphine to migrate onto the aluminum atoms (Figure 4B), with generation of a cis-1,4 specific catalytic center, quite similar to the one obtained by reacting naked CoCl2 with MAO.
The possibility that the phosphine ligand, under certain polymerization conditions, may migrate from cobalt to aluminum, causing a drastic change in the selectivity of the catalytic center, is supported by the fact that the catalytic systems CoCl2(PPh3)2/MAO and CoCl2(PnPrPh2)2/MAO at higher MAO/Al molar ratio (up to 5000) give poly(1,3-butadiene)s with a mixed cis-1,4/1,2 structure. Evidently, in this case such a high MAO/Co molar ratio caused a migration of part of the ligand onto the aluminum atom.

3. Materials and Methods

3.1. General Procedure and Materials

2-(1,1-Dimethylpropyl)-6-(diphenylphosphino)pyridine (Aldrich, Merck KGaA (Darmstadt, Germany 98% pure (HPLC)), (2-methoxyphenyl)diphenylphosphine (Aldrich (Merck KGaA (Darmstadt, Germany) 98% pure (HPLC)), anhydrous cobalt dichloride (Aldrich, Merck KGaA (Darmstadt, Germany) 99,9% pure), and methylaluminoxane (MAO) (Aldrich, Merck KGaA (Darmstadt, Germany), 10 wt% solution in toluene) were used as received. (2-Methoxyethyl)diphenylphosphine [14], CoCl2(PnPrPh2)2 [2], and CoCl2(PPh3)2 [8] were prepared according to that reported in the literature. Ethanol (Aldrich, Merck KGaA (Darmstadt, Germany)) was degassed under vacuum, then by bubbling dry dinitrogen and kept over molecular sieves; pentane (Aldrich, Merck KGaA (Darmstadt, Germany), 99% pure) was refluxed over Na/K alloy for 8 h, then distilled and stored over molecular sieves under dry dinitrogen; toluene (Aldrich, Merck KGaA (Darmstadt, Germany), 99,8% pure) was refluxed over Na for 8 h, then distilled and stored over molecular sieve under dry dinitrogen. 1,3-Butadiene (Aldrich, Merck KGaA (Darmstadt, Germany), >99,5% pure) was evaporated from the container before each run, dried by passing through a column packed with molecular sieves, and condensed into the reactor, which had been precooled to −20 °C. The two novel phosphine cobalt complexes were synthesized as indicated below, following a general procedure already reported in the literature [8].

3.2. Synthesis of Cobalt Phosphine Complexes

3.2.1. [(2-Methoxyethyl)diphenylphosphine] Cobalt Dichloride (1), Toluene as Medium

(2-Methoxyethyl)diphenylphosphine (0.605 g, 2.48 × 10−3 mol) was dissolved in toluene (20 mL), and CoCl2 (0.161 g, 1.24 × 10−3 mol) was successively added. A blue suspension was gradually formed; after 24 h it was filtered, then washed with toluene (2 × 5 mL) and heptane (2 × 10 mL). The blue solid isolate was then dried in vacuum at room temperature. The blue supernatant solution was removed, concentrated, and cooled at −18 °C, causing the precipitation of a crystalline product suitable for X-ray crystallographic studies. Yield: 0.552 g (72.0% based on CoCl2).
Anal. Calcd. for C30H34Cl2CoO2P2: C, 58.27; H, 5.54; Cl, 11,47; Co, 9.53; Found: C, 58.00; H, 5.19; Cl, 11,84; Co, 9.12.
Spectroscopic data: IR (KBr) ν (cm−1) 3051 w, 2885 w, 1584 w, 1572 w, 1485 m, 1433 s, 1386 m, 1100 vs, 953 s, 741 vs, 690 vs (see Supplementary Materials, Figures S1 and S1A).

3.2.2. [(2-Methoxyphenyl)diphenylphosphine] Cobalt Dichloride (2), Tetrahydrofurane/Ethanol as Medium

(2-Methoxyphenyl)diphenylphosphine (1.00 g, 3.42 × 10−3 mol) was dissolved in tetrahydrofurane (30 mL) and successively added to a solution of CoCl2 (0.204 g, 1.57 × 10− 3 mol) in ethanol (30 mL). A blue suspension was gradually formed; after 20 h it was filtered, washed with ethanol (2 × 10 mL) and pentane (2 × 10 mL), and then dried in vacuum at room temperature. The isolated blue solid was transferred on the filter of a Soxhlet for solids and extracted continuously with boiling diethylether. The extraction was practically complete in two days; at the end, a microcrystalline blue powder was formed on the bottom of the extraction Schlenk-tube. The blue supernatant solution was removed, concentrated, and cooled at − 30 °C, causing the precipitation of a crystalline product. Further crops of crystals were obtained by repeating this workup operation several times. Yield: 0.897 g (79.6% based on CoCl2).
Anal. Calcd. for C38H34Cl2CoO2P2: C, 63.88; H, 4.80; Cl, 9.92; Co, 8.25; Found: C, 64.0; H, 4.7; Cl, 9.8; Co, 8.4.
Spectroscopic data: IR (KBr) ν (cm−1). 1475mw, 1464 mw, 1435 m, 1101 m, 906 m, 757 s, 746 s, 693 s, 505 s, 492s (see Supplementary Materials, Figure S2).

3.2.3. [2-(1,1-dimethylpropyl)-6-(diphenylphosphino)pyridine] Cobalt Dichloride (3)

2-(1,1-Dimethylpropyl)-6-(diphenylphosphino)pyridine (0.997 g, 2.99 × 10−3 mol) was dissolved in ethanol (30 mL) and successively added to a solution of CoCl2 (0.180 g, 1.38 × 10−3 mol) in ethanol (20 mL). A deep blue solution was gradually formed; after 48 h the solvent was removed under vacuum, obtaining a blue oil product, which was washed several times at low temperature with pentane (4 × 20 mL) and then dried under vacuum at room temperature. A blue microcrystalline powder was at the end obtained: yield 0.685 g (62.3% based on CoCl2). Single crystals of the complex were obtained by dissolving the complex in diethylether and by recrystallizing at low temperature.
Anal. Calcd. for C44H48Cl2CoN2P2: Co, 7.40; Cl, 8.90; P, 7.78; C, 66.34; H, 6.07; N, 3.52. Found: Co, 7.6; Cl, 9.0; P, 7.7; C, 66.2; H, 6.2; N, 3.6.
Spectroscopic data: IR (KBr) ν (cm−1) 1599 s, 1435 ms, 1260 m, 1095 m, 808 m, 742 s, 692 s, 504 m (see Supplementary Materials, Figure S3).

3.3. X-ray Crystallographic Studies

A summary of the experimental details concerning the single crystal X-ray diffraction studies on complexes 1 and 2 is reported in Table 4.
The crystals used for data collection were entirely covered with perfluorinated oil to reduce crystal decay. Data were recorded on Bruker Photon 100 (1) or APEX II (2) area detector diffractometers (Bruker AXS Inc., Madison, WI, USA) using Mo–Kα radiation. Data were corrected for Lorentz polarization and absorption effects by SADABS [15]. The structures were solved by direct methods and refined by full-matrix least-squares based on all data using F2 [16]. Hydrogen atoms were fixed at calculated positions and refined by a riding model.

3.4. Polymerization

A standard procedure is reported. 1,3-Butadiene was condensed into a 25 mL dried glass reactor kept at −20 °C, then toluene was added, and the solution obtained was brought to the desired polymerization temperature. MAO and the cobalt compound were then added, as toluene solutions, in that order. The polymerization was terminated with methanol containing a small amount of hydrochloric acid, and the polymer was coagulated and repeatedly washed with methanol, then dried in vacuum at room temperature to constant weight.

3.5. Polymer Characterization

The polymer microstructure was determined through FT-IR and NMR (1H and 13C) analyses (see Supplementary Materials, Figures S4–S15) as reported in the literature [17,18,19,20,21,22,23].13C NMR and 1H NMR measurements were performed with a Bruker AM 400 instrument (Bruker Italia Srl, Milano, Italy). The spectra were obtained in C2D2Cl4 at 103 °C (hexamethyldisiloxane, HMDS, as internal standard). The concentration of polymer solutions was about 10 wt.%. Wide-angle X-ray diffraction (XRD) experiments (see Supplementary Materials, Figures S16 and S17) were performed at 25 °C under nitrogen flux, using a Siemens D-500 diffractometer equipped with Soller slits (2°) placed before sample, 0.3° aperture and divergence windows, and a VORTEX detector with extreme energy resolution specific for thinner films. CuKα radiation with power use of 40 KV × 40 mA was adopted, and each spectrum was measured with steps of 0.05° 2θ and 6s measurement time. FTIR spectra were acquired using a Perkin-Elmer (Waltham, MA, USA) Spectrum Two in attenuated total reflectance (ATR) mode in the spectral range of 4000–500 cm−1. The molecular weight average (Mw) and the molecular weight distribution (Mw/Mn) were obtained by a high-temperature Waters GPCV2000 (Milford, MA, USA) size-exclusion chromatography (SEC) system equipped with a refractometer detector. The experimental conditions consisted of three PL Gel Olexis columns, ortho-dichlorobenzene (o-DCB) as the mobile phase, a 0.8 mL/min flow rate, and a 145 °C temperature. The calibration of the SEC system was constructed using eighteen narrow Mw/Mn poly(styrene) standards with Mws ranging from 162 to 5.6 × 106 g mol−1. For SEC analysis, about 12 mg of polymer was dissolved in 5 mL of o-DCB with 0.05% of BHT as antioxidant.

4. Conclusions

Three novel cobalt diphenyl phosphine complexes were synthesized, and the crystal structure of two of them was determined by single crystal X-ray diffraction. The behavior of these complexes in combination with MAO in the polymerization of 1,3-butadiene was examined, and it was found to be strongly affected by the MAO/Co ratio, giving predominantly 1,2 polymers at low MAO/Co molar ratios (up to 100) and essentially cis-1,4 polymers at higher Al/Co molar ratios (1000). The different behavior by varying the MAO/Co ratios was attributed to the presence of donor heteroatoms within the ligand structure, making easier the displacement of the phosphine ligand from the cobalt atom, resulting in a drastic change in selectivity of the catalytic center, from 1,2 specific to cis-1,4 specific.
The possibility of modifying the catalytic selectivity during the polymerization process simply by varying the MAO/Co ratio could be interesting since it could permit the preparation of poly(1,3-butadiene)s consisting of polymeric blocks with different stereoregularity, having elastomeric or thermoplastic features depending on the block microstructure. Examples of this type, in which the catalytic selectivity can be adjusted by varying the aluminum-alkyl ratio, have already been reported in the literature, for example in the case of iron [24] and neodymium-based catalysts [25,26].

Supplementary Materials

Tables of atomic coordinates, anisotropic thermal parameters, bond lengths and angles of 1 and 2 may be obtained free of charge from The Director CCDC, 12 Union Road, Cambridge CB2 1 EZ, UK, on quoting the deposition number CCDC 1905516 (compound 1) and CCDC 2074828 (compound 2), the names of the authors and the journal citation (fax: +44-1223-336-033; e-mail: [email protected]; web site: http://www.ccdc.cam.ac.uk). FT-IR spectra of the cobalt complexes 1-3; FT-IR spectra, NMR (1H and 13C) spectra, and XRD spectra of some of the polybutadienes of Table 3. Figure S1. FT-IR spectrum of (2-methoxyethyl)diphenylphosphine; Figure S1A. FT-IR spectrum of [(2-methoxyethyl)diphenylphosphine]2 cobalt dichloride (1), Figure S2. FT-IR spectrum of [(2-methoxyphenyl)diphenylphosphine]2 cobalt dichloride (2); Figure S3. FT-IR spectrum of [2-(1,1-dimethylpropyl)-6-(diphenylphosphino)pyridine]2 cobalt dichloride (3); Figure S4. FT-IR spectrum of the polybutadiene of Table 3, run 1; Figure S5. FT-IR spectrum of the polybutadiene of Table 3, run 2; Figure S6. FT-IR spectrum of the polybutadiene of Table 3, run 3; Figure S7. FT-IR spectrum of the polybutadiene of Table 3, run 4; Figure S8. FT-IR spectrum of the polybutadiene of Table 3, run 5; Figure S9. FT-IR spectrum of the polybutadiene of Table 3, run 6; Figure S10. FT-IR spectrum of the polybutadiene of Table 3, run 7; Figure S11. FT-IR spectrum of the polybutadiene of Table 3, run 8; Figure S12. NMR spectra (1H down, 13C up) of the polybutadiene of Table 3, run 2; Figure S13. NMR spectra (1H down, 13C up) of the polybutadiene of Table 3, run 3; Figure S14. NMR spectra (1H down, 13C up) of the polybutadiene of Table 3, run 5; Figure S15. NMR spectra (1H down, 13C up) of the polybutadiene of Table 3, run 6; Figure S16. XRD spectrum of the polybutadiene of Table 3, run 5; Figure S17. XRD spectrum of the polybutadiene of Table 3, run 6.

Author Contributions

G.L., G.Z., and B.P. planned and carried out the polymerization experiments and part of the organometallic syntheses. A.S., F.M., M.G., and G.P. performed the synthesis of the ligands and of the organometallic complexes and their characterization, and contributed to preparing the manuscript. A.F. and S.Z. determined the X-ray molecular structures of the complexes and contributed to writing the paper. G.R. was the principal investigator, conceived and designed the experiments, interpreted the results, and wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Authors wish to thank Fulvia Greco and Daniele Piovani for their support in NMR and GPC analyses of the polymers.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the cobalt compounds and of the polymers are not available from the authors.

References

  1. Ricci, G.; Forni, A.; Boglia, A.; Motta, T.; Zannoni, G.; Canetti, M.; Bertini, F. Synthesis and X-ray Structure of CoCl2(PiPrPh2)2. A New Highly Active and Stereospecific Catalyst for 1,2 Polymerization of Conjugated Dienes When Used in Association with MAO. Macromol 2005, 38, 1064–1070. [Google Scholar] [CrossRef]
  2. Ricci, G.; Forni, A.; Boglia, A.; Sommazzi, A.; Masi, F. Synthesis, structure and butadiene polymerization behavior of CoCl2(PRxPh3−x)2 (R=methyl, ethyl, propyl, allyl, isopropyl, cyclohexyl; x=1, 2). Influence of the phosphorous ligand on polymerization stereoselectivity. J. Organomet. Chem. 2005, 690, 1845–1854. [Google Scholar] [CrossRef]
  3. Ricci, G.; Leone, G.; Pierro, I.; Zanchin, G.; Forni, A. Novel Cobalt Dichloride Complexes with Hindered Diphenylphosphine Ligands: Synthesis, Characterization, and Behavior in the Polymerization of Butadiene. Molecules 2019, 24, 2308. [Google Scholar] [CrossRef] [Green Version]
  4. Porri, L.; Giarrusso, A.; Ricci, G. Recent views on the mechanism of diolefin polymerization with transition metal initiator systems. Prog. Polym. Sci. 1991, 16, 405–441. [Google Scholar] [CrossRef]
  5. Ricci, G.; Sommazzi, A.; Masi, F.; Ricci, M.; Boglia, A.; Leone, G. Well-defined transition metal complexes with phosphorus and nitrogen ligands for 1,3-dienes polymerization. Coord. Chem. Rev. 2010, 254, 661–676. [Google Scholar] [CrossRef]
  6. Ricci, G.; Leone, G. Recent progresses in the polymerization of butadiene over the last decade. Polyolefins J. 2014, 1, 43–60. [Google Scholar]
  7. Ricci, G.; Leone, G. Polymerization of 1,3-butadiene with organometallic complexes-based catalysts. In Applied Homogeneous Catalysis with Organometallic Compounds: A Comprehensive Handbook in Four Volumes, 3rd ed.; Cornils, B., Hermann, W.A., Beller, M., Paciello, R., Eds.; Wiley-VC: Weinheim, Germany, 2017; Volume 1, pp. 251–273. [Google Scholar]
  8. Chatt, J.; Shaw, B.L. 51. Alkyls and aryls of transition metals. Part IV. Cobalt(II) and iron(II) derivatives. J. Chem. Soc. 1961, 285–290. [Google Scholar] [CrossRef]
  9. Warad, I.; Barakat, A. Synthesis, physicochemical analysis of two new hemilabile ether-phosphine ligands and their first stable bis-ether-phosphine/cobalt(II) tetrahedral complexes. J. Mol. Struct. 2017, 1134, 17–24. [Google Scholar] [CrossRef]
  10. Braunstein, P.; Kelly, D.G.; Dusausoy, Y.; Bayeul, D.; Lanfranchi, M.; Tiripicchio, A. Complexes of functional phosphines. 23. PPh3 and β-ketophosphine complexes of cobalt(II). Crystal and molecular structures of CoCl2[Ph2PCH2C(O)Ph]2 and Co2(CO)6[Ph2PCH2C(O)Ph]2 and of the trinuclear, mixed-metal complex [(Ph3P)2Ag(μ-Cl)2Co(μ-Cl)2Ag(PPh3)2]·0.5 OEt2. Inorg. Chem. 1994, 33, 233–242. [Google Scholar] [CrossRef]
  11. Carlin, R.L.; Chirico, R.D.; Sinn, E.; Mennenga, G.; De Jongh, L.J. Magnetic ordering in cobalt chloride-triphenylphosphine (CoCl2·2P(C6H5)3) and cobalt bromide-triphenylphosphine (CoBr2·2P(C6H5)3). Inorg. Chem. 1982, 21, 2218–2222. [Google Scholar] [CrossRef]
  12. Yang, F.; Zhou, Q.; Zhang, Y.; Zeng, G.; Li, G.; Shi, Z.; Wang, B.; Feng, S. Inspiration from old molecules: Field-induced slow magnetic relaxation in three air-stable tetrahedral cobalt(ii) compounds. Chem. Commun. 2013, 49, 5289–5291. [Google Scholar] [CrossRef] [PubMed]
  13. Ricci, G.; Boccia, A.C.; Leone, G.; Forni, A. Novel Allyl Cobalt Phosphine Complexes: Synthesis, Characterization and Behavior in the Polymerization of Allene and 1,3-Dienes. Catalysts 2017, 7, 381. [Google Scholar] [CrossRef] [Green Version]
  14. McEwen, W.E.; Janes, A.B.; Knapczyk, J.W.; Kyllingstad, V.L.; Shiau, W.-I.; Shore, S.; Smith, J.H. Role of through space 2p-3d overlap in the alkylation of phosphines. J. Am. Chem. Soc. 1978, 100, 7304–7311. [Google Scholar] [CrossRef]
  15. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. Sect. A Found. Crystallogr. 2008, 64, 112–122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. 2015, 71, 3–8. [Google Scholar]
  17. Morero, D.; Santambrogio, A.; Porri, L.; Ciampelli, F. Infrared analysis of polybutadienes obtained by stereospecific synthesis. Stereoregul. Polym. Stereospecif. Polym. 1967, 41, 693–695. [Google Scholar] [CrossRef]
  18. Tanaka, Y.; Takeuchi, Y.; Kobayashi, M.; Tadokoro, H. Characterization of diene polymers. I. Infrared and NMR studies: Nonadditive behavior of characteristic infrared bands. J. Polym. Sci. Part A-2 Polym. Phys. 1971, 9, 43–57. [Google Scholar] [CrossRef]
  19. Zymonas, J.; Santee, E.R.; Harwood, H.J. 1H Nuclear Magnetic Resonance Studies of 1,2 Polybutadienes. Macromol 1973, 6, 129–133. [Google Scholar] [CrossRef]
  20. Ashitaka, H.; Inaishi, K.; Ueno, H. Syndiotactic 1,2-polybutadiene with Co-CS2 catalyst system. III. 1H- and 13C-NMR study of highly syndiotactic 1,2-polybutadiene. J. Polym. Sci. Polym. Chem. Ed. 1983, 21, 1973–1988. [Google Scholar] [CrossRef]
  21. Kumar, D.; Rao, M.R.; Rao, K.V.C. Carbon-13 NMR of 1,2-polybutadienes: Configurational sequencing. J. Polym. Sci. Polym. Chem. Ed. 1983, 21, 365–374. [Google Scholar] [CrossRef]
  22. Mochel, V.D. Carbon-13 NMR of polybutadiene. J. Polym. Sci. Part A-1 Polym. Chem. 1972, 10, 1009–1018. [Google Scholar] [CrossRef]
  23. Elgert, K.-F.; Quack, G.; Stützel, B. Zur struktur des polybutadiens, 2. Das 13C-NMR-Spektrum des 1,2-polybutadiens. Die Makromol. Chem. 1974, 175, 1955–1960. [Google Scholar] [CrossRef]
  24. Zheng, W.; Wang, F.; Bi, J.; Zhang, H.; Zhang, C.; Hu, Y.; Bai, C.; Zhang, X. Synthesis and characterization of soft-hard stereoblock polybutadiene with Fe(2-EHA)3 /Al(i-Bu)3 /DEP catalyst system. J. Polym. Sci. Part A Polym. Chem. 2015, 53, 1182–1188. [Google Scholar] [CrossRef]
  25. Tanaka, R.; Shinto, Y.; Nakayama, Y.; Shiono, T. Synthesis of Stereodiblock Polybutadiene Using Cp*Nd(BH4)2(thf)2 as a Catalyst. Catalysts 2017, 7, 284. [Google Scholar] [CrossRef] [Green Version]
  26. Tanaka, R.; Yuuya, K.; Sato, H.; Eberhardt, P.; Nakayama, Y.; Shiono, T. Synthesis of stereodiblock polyisoprene consisting of cis-1,4 and trans-1,4 sequences by using a neodymium catalyst: Change of the stereospecificity triggered by an aluminum compound. Polym. Chem. 2016, 7, 1239–1243. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Chemical structures of CoCl2(Ph2PCH2CH2OCH3)2 (1), CoCl2(Ph2P(C6H4OCH3)) 2 (2), and CoCl2(Ph2P(C5NH3(C5H11))2 (3).
Figure 1. Chemical structures of CoCl2(Ph2PCH2CH2OCH3)2 (1), CoCl2(Ph2P(C6H4OCH3)) 2 (2), and CoCl2(Ph2P(C5NH3(C5H11))2 (3).
Molecules 26 04067 g001
Figure 2. Molecular structure of 1 with key atoms labeled. Thermal ellipsoids are drawn at the 50% probability level.
Figure 2. Molecular structure of 1 with key atoms labeled. Thermal ellipsoids are drawn at the 50% probability level.
Molecules 26 04067 g002
Figure 3. Molecular structure of 2 with key atoms labeled. Thermal ellipsoids are drawn at the 50% probability level.
Figure 3. Molecular structure of 2 with key atoms labeled. Thermal ellipsoids are drawn at the 50% probability level.
Molecules 26 04067 g003
Figure 4. Possible structure of the active sites with the phosphine ligand on the cobalt atom (A) and on MAO (B).
Figure 4. Possible structure of the active sites with the phosphine ligand on the cobalt atom (A) and on MAO (B).
Molecules 26 04067 g004
Table 1. Selected bond lengths and angles for 1.
Table 1. Selected bond lengths and angles for 1.
Bond Lengths (Å)
Co–Cl12.229(3)Co–Cl22.220(3)P1–C101.831(11)P2–C301.834(11)
Co–P12.361(3)Co–P22.356(3)C1–C21.514(15)C21–C221.526(15)
P1–C11.843(10)P2–C211.834(10)C2–O11.393(14)C22–O21.415(13)
P1–C41.821(12)P2–C241.848(12)O1–C31.408(14)O2–C231.431(14)
Bond Angles (°)
Cl1–Co–Cl2114.72(12)P1–Co–P2107.54(11)C10–P1–Co122.1(4)C30–P2–Co112.5(4)
Co–P1–C1109.7(4)Co–P2–C21118.1(4)P1–C1–C2111.4(8)P2–C21–C22113.0(7)
C1–P1–C4105.2(5)C21–P2–C24104.7(5)C1–C2–O1106.0(10)C21–C22–O2106.0(8)
C4–P1–C10105.1(6)C24–P2–C30103.0(5)C2–O1–C3112.5(10)C22–O2–C23111.3(8)
Table 2. Selected bond lengths and angles for 2.
Table 2. Selected bond lengths and angles for 2.
Bond Lengths (Å)
Co–Cl12.2218(6)Co–Cl22.2395(5)
Co–P12.4064(6)Co–P22.4115(6)
P1–C11.821(2)P2–C191.820(2)
P1–C71.821(2)P2–C251.829(2)
P1–C131.829(2)P2–C311.817(2)
Bond Angles (°)
Cl1–Co–Cl2124.01(2)P1–Co–P2111.02(2)
Co–P1–C1109.25(7)Co–P2–C19119.67(7)
C1–P1–C7104.87(9)C19–P2–C25104.18(9)
C7–P1–C13101.57(9)C25–P2–C31105.08(9)
C13–P1–Co114.16(7)C31–P2–Co110.80(7)
Table 3. Polymerization of 1,3-butadiene with cobalt catalytic systems. 1
Table 3. Polymerization of 1,3-butadiene with cobalt catalytic systems. 1
RunCo-Complex (μmol)Al/CoTime (min)Yield (%)N 2 (min−1)cis-1,4 3 (%)1,2 3 (%) [rr] 4 (%)
11 (10)10003088.5413087.312.7
21 (10)1003061.8288419.780.351
31 (30)206084.966118.981.150
42 (10)10006092.9216879.419.6
52 (10)1006078.2182519.480.666
62 (30)206050.439221.778.365
73 (10)10001090.112,61494.75.3
83 (10)1001081.711,43836.363.746
9CoCl2(PnPrPh2)2 (3)30003087.313,58048.251.846
10CoCl2(PnPrPh2)2 (5)1600458.841,16033.466.646
11CoCl2(PnPrPh2)2 (5)1000488.962,23028.771.346
12CoCl2(PnPrPh2)2 (10)500290.763,49026.074.043
13CoCl2(PnPrPh2)2 (10)300293.565,45021.878.245
14CoCl2(PnPrPh2)2 (10)100294.866,36021.378.744
15CoCl2(PPh3)2 (1)500013530.8319454.145.966
16CoCl2(PPh3)2 (5)1000780.632,24026.074.064
17CoCl2(PPh3)2 (5)500672.433,78726.473.664
18CoCl2(PPh3)2 (5)100610046,66724.076.065
19CoCl2(PPh3)2 (5)50797.238,88023.776.366
20CoCl2(PPh3)2 (10)20790.318,06023.476.665
1 Polymerization conditions: toluene, total volume 16 mL; 1,3-butadiene, 2 mL; MAO; 22 °C. The polymer molecular weights are in the range 100,000–200,000 g·mol−1, with a Mw/Mn in the range 2–3, 2 N = moles of butadiene polymerized/moles of Co × minutes, 3 determined by FTIR and 1H NMR, 4 percentage of syndiotactic triads, determined by 13C NMR.
Table 4. Crystal data, data collection, and refinement details for 1 and 2.
Table 4. Crystal data, data collection, and refinement details for 1 and 2.
12
Crystal data
Chemical formulaC30H34Cl2CoO2P2C38H34Cl2CoO2P2
Mr618.34714.42
Crystal systemtriclinictriclinic
Space groupP 1 (No. 2)P 1 (No. 2)
Temperature [K]100(2)120(2)
a [Å]9.922(4)9.3758(7)
b [Å]9.927(4)12.3136(9)
c [Å]16.872(6)15.7602(12)
α [°]101.510(9)99.7760(10)
β [°]90.902(9)104.3110(10)
γ [°]110.168(7)100.9440(10)
V3]1522.1(10)1685.5(2)
Z22
μ(MoKα) [mm−1]0.8690.796
Crystal size [mm]0.19 × 0.14 × 0.110.15 × 0.12 × 0.10
Crystal color, habitblue, platelight blue, prism
Data collection
Tmin, Tmax0.509, 0.7460.708, 0.746
No. of measured reflns10,33633,407
No. of independent refls527210,240
No. of observed refls [I > 2σ(I)]36407397
Rint0.08270.0467
Rσ0.13070.0551
(sin θ/λ)max−1]0.5950.715
Refinement
R[F2 > 2σ(F2)]0.12210.0463
wR(F2)0.30660.1138
S1.1451.007
No. of reflections527210,240
No. of parameters336408
No. of restraints420
Δρmax, Δρmin (e Å−3)2.130, −0.9261.080, −0.295
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ricci, G.; Leone, G.; Zanchin, G.; Palucci, B.; Forni, A.; Sommazzi, A.; Masi, F.; Zacchini, S.; Guelfi, M.; Pampaloni, G. Some Novel Cobalt Diphenylphosphine Complexes: Synthesis, Characterization, and Behavior in the Polymerization of 1,3-Butadiene. Molecules 2021, 26, 4067. https://doi.org/10.3390/molecules26134067

AMA Style

Ricci G, Leone G, Zanchin G, Palucci B, Forni A, Sommazzi A, Masi F, Zacchini S, Guelfi M, Pampaloni G. Some Novel Cobalt Diphenylphosphine Complexes: Synthesis, Characterization, and Behavior in the Polymerization of 1,3-Butadiene. Molecules. 2021; 26(13):4067. https://doi.org/10.3390/molecules26134067

Chicago/Turabian Style

Ricci, Giovanni, Giuseppe Leone, Giorgia Zanchin, Benedetta Palucci, Alessandra Forni, Anna Sommazzi, Francesco Masi, Stefano Zacchini, Massimo Guelfi, and Guido Pampaloni. 2021. "Some Novel Cobalt Diphenylphosphine Complexes: Synthesis, Characterization, and Behavior in the Polymerization of 1,3-Butadiene" Molecules 26, no. 13: 4067. https://doi.org/10.3390/molecules26134067

Article Metrics

Back to TopTop