Next Article in Journal
The Influence of Solution pH on the Kinetics of Resorcinol Electrooxidation (Degradation) on Polycrystalline Platinum
Next Article in Special Issue
Highly Active Iminopyridyl Iron-Based Catalysts for the Polymerization of Isoprene
Previous Article in Journal
Different Role of Bisulfite/Sulfite in UVC-S(IV)-O2 System for Arsenite Oxidation in Water
Previous Article in Special Issue
Homo- and Copolymerization of Ethylene with Norbornene Catalyzed by Vanadium(III) Phosphine Complexes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Novel Cobalt Dichloride Complexes with Hindered Diphenylphosphine Ligands: Synthesis, Characterization, and Behavior in the Polymerization of Butadiene

1
CNR-Istituto per lo Studio delle Macromolecole (ISMAC), via A. Corti 12, 20133 Milano, Italy
2
CNR-Istituto di Scienze e Tecnologie Molecolari (ISTM), via C. Golgi 19, 20133 Milano, Italy
*
Author to whom correspondence should be addressed.
Molecules 2019, 24(12), 2308; https://doi.org/10.3390/molecules24122308
Submission received: 10 May 2019 / Revised: 13 June 2019 / Accepted: 19 June 2019 / Published: 21 June 2019
(This article belongs to the Special Issue Well-Defined Metal Complex Catalysts for Olefin Polymerization)

Abstract

:
Two novel cobalt diphenylphosphine complexes were synthesized by reacting cobalt(II) chloride with tert-butyl(diphenyl)phosphine (PtBuPh2) and (S)-(+)neomenthyldiphenylphosphine [(S)-NMDPP]. The crystal structure of the former was determined by single-crystal X-ray diffraction studies. The two complexes were then used in combination with methylaluminoxane (MAO) for the polymerization of 1,3-butadiene: crystalline highly syndiotactic 1,2 poly(1,3-butadiene)s were obtained, with a 1,2 content and a syndiotactic index (percentage of syndiotactic triads [rr]) up to 95% and 85%, respectively. The results obtained further support and confirm what was already observed in the polymerization of 1,3-butadiene with CoCl2(PRPh2)2−MAO (R = methyl, ethyl, normal-propyl, iso-propyl, and cyclohexyl): the nature of the phosphine ligand strongly affects the polymerization stereoselectivity, the polymer syndiotacticity increasing with increasing phosphine ligand steric hindrance.

Graphical Abstract

1. Introduction

Cobalt catalysts are well-known in the field of the stereospecific polymerization of 1,3-dienes, and they can perhaps be considered the most versatile catalytic systems among the various transition metal and lanthanide catalysts examined, since, depending on the catalytic formulation, they are able to provide all the possible polybutadiene stereoisomers: cis-1,4 polybutadiene, 1,2 polybutadiene, polybutadiene with a mixed cis-1,4/1,2 structure, and trans-1,4 polybutadiene [1,2,3,4,5,6].
Specifically, the catalytic systems obtained by combining CoCl2(PRPh2)2 (R = methyl, ethyl, normal-propyl, iso-propyl, and cyclohexyl) with methylaluminoxane (MAO) gave 1,2 polymers from butadiene and other terminally substituted 1,3-butadienes such as 1,3-pentadiene, 3-methyl-1,3-pentadiene, 1,3-hexadiene, 5-methyl-1,3-hexadiene, 1,3-heptadiene and 1,3-octadiene, simultaneously exhibiting a very high catalytic activity [7,8,9,10,11]. The tacticity of the 1,2 polymers obtained (i.e., the polymerization stereoselectivity) was found to depend on the type of phosphine ligand coordinated to the cobalt atom and on the monomer structure [12,13,14,15,16]: specifically, the syndiotacticity (i.e., the percentage of syndiotactic triads [rr%]) increased with increasing steric hindrance of the phosphine ligand, and the same obviously applies to the melting point of the polymers obtained.
In order to verify the above-described behavior (i.e., the increasing of syndiotacticity and melting point with increasing the ligand hindrance), and to obtain a syndiotactic 1,2polybutadiene having a melting point even closer to that of the polymer currently industrially produced by means of the Co(acac)3/AlEt3/H2O/CS2 system [17,18,19,20,21,22], we synthesized two novel cobalt complexes (1 and 2 in Figure 1) by reacting cobalt dichloride with the following highly hindered phosphine ligands: tert-butyl(diphenyl)phosphine (PtBuPh2) and (S)-(+)neomenthyldiphenylphosphine [(S)-NMDPP]. The molecular structure of complex 1 was determined by X-ray diffraction, and then both complexes 1 and 2 were used in association with MAO for the polymerization of 1,3-butadiene. The results obtained clearly confirmed and supported the trend observed in the polymerization of 1,3-butadiene with the CoCl2(PRPh2)2−MAO catalyst systems (R = methyl, ethyl, normal-propyl, iso-propyl, and cyclohexyl) [8,9], and highly syndiotactic 1,2 polybutadienes were obtained, with a 1,2 content and a syndiotactic index (percentage of syndiotactic triads [rr]) up to 95% and 85%, respectively.

2. Results and Discussion

2.1. Synthesis and Characterization of Cobalt Complexes

The two novel cobalt complexes were prepared according to a general experimental procedure already reported in the literature [23]. CoCl2 was dissolved in ethanol and reacted with an excess of phosphine. The solutions so obtained were kept under stirring at room temperature for one day, then the solvent was removed under vacuum. The residues were washed with cold pentane, then dried again under vacuum. Crystalline products were obtained by continuous extraction of the residues with boiling pentane. Single crystals suitable for X-ray diffraction studies were obtained for complex 1, allowing its molecular structure to be determined (Figure 2). Selected bond lengths and angles are reported in Table 1.
The molecular structure of 1 can be conveniently discussed together with those of the previously reported related CoCl2(PRPh2)2 complexes with R = ethyl (Et), normal-propyl (nPr), and iso-propyl (iPr) [8,9]. In fact, in these compounds, the phosphorus atoms bear the same aromatic groups (two phenyl rings) but different aliphatic substituents, going from small (Et, nPr) to larger (iPr, tBu) groups. As expected, a clear differentiation between bonds connecting phosphorus and aliphatic (P−Caliph) or aromatic carbons (P−Car) was found, the former being on average longer than the latter. Moreover, they both depended on the bulkiness of the aliphatic substituents, increasing from the complexes with R = Et and nPr to 1. Greater variation was clearly observed for the P−Caliph distances, which measured on average 1.829(2), 1.827(2), 1.850(2), and 1.888(1) Å for R = Et, nPr, iPr, and tBu, respectively. On the other hand, the average P–Car bond lengths showed a modest increase from 1.814(3) (both Et and nPr) and 1.816(2) (iPr) to 1.826(1) Å for tBu. The greater steric hindrance of the tBu group also implies larger Co−P and Co−Cl distances in 1 than those observed in the other three complexes. In particular, the average Co–P bond lengths, measuring 2.370(1), 2.380(1), and 2.363(1) Å in the complexes with R = Et, nPr, and iPr, respectively, became as large as 2.433(1) Å in 1.
Note that the different dimensions of the aliphatic substituents could explain the different conformations observed in complexes with R = tBu and iPr with respect to those of R = nPr and Et derivatives. While in the former the aliphatic groups were in trans along the P−Co−P−R sequence, in the latter they were in gauche along the same sequence. This involves a different arrangement of the phenyl rings, which were both gauche in the complexes with R = tBu and iPr, granting a more efficient accommodation of the entire phosphinic group.
The observed difference in the conformations of CoCl2(PRPh2)2 complexes may in turn account for the significantly larger P–Co–P angle for the derivatives with bulkier aliphatic groups. This angle in fact decreased from 113.69(1)° (R = tBu) to 110.64(3)° (iPr), 104.85(4)° (nPr), and 102.89(3)° (Et). Widening the P−Co−P angle allows for a partial reciprocal distancing of the phenyl rings of different phosphines, which in R = tBu and iPr derivatives happened to be in pairs in a face-to-face repulsive arrangement, only partially relieved by a partial rotation around the P−Car bonds. The other bond angles were instead almost unchanged in the series of examined structures.

2.2. Polymerization of 1,3-Butadiene

The results obtained in the polymerization of 1,3-butadiene with the two new cobalt complexes based catalysts are shown in Table 2; the results obtained with CoCl2(PRPh2)2−MAO (R = iso-propyl, cyclohexyl) [8,9], which were found to be the systems providing 1,2 poly(1,3-butadiene) with the highest syndiotactic content, are added for comparison.
With a MAO/Co molar ratio in the range 10−100, the CoCl2(PtBuPh2)2−MAO and CoCl2[PPh2(NMDPP)]2−MAO systems both gave polybutadienes with a predominantly 1,2 structure (up to 94.5%) and a syndiotactic content up to 85.5%. The 13C-NMR spectrum of the polybutadiene obtained with the system CoCl2(PtBuPh2)2−MAO at MAO/Co = 25, 0 °C, and in heptane as solvent (Table 2, run 4) is shown in Figure 3; the 13C NMR spectra of the other polybutadienes of Table 2 (Figures S1–S6), together with the FTIR spectra of the same polymers (Figures S7–S13) are reported in the Supplementary Materials. The 1,2 content and the syndiotacticity degree seemed to be only slightly affected by the MAO/Co molar ratio (in the range examined) (cfr. runs 1−2), whereas they seemed to be more influenced by the nature of the solvent (cfr. runs 2−3 and 5−6) and the polymerization temperature in particular (cfr. runs 3–4 and 6–7). This behavior was already observed for the analogous systems CoCl2(PRPh2)2−MAO (R = methyl, ethyl, normal-propyl, iso-propyl, and cyclohexyl), and a plausible interpretation for this behavior has already been given [8,9].
The syndiotactic degrees of the polybutadienes obtained with the two systems CoCl2(PtBuPh2)2−MAO and CoCl2[PPh2(NMDPP)]2−MAO were higher than those exhibited by CoCl2(PiPrPh2)2−MAO and CoCl2(PCyPh2)2−MAO (cfr. runs 1,8 and 5,9), which, as mentioned above, in the series CoCl2(PRPh2)2−MAO (R = methyl, ethyl, normal-propyl, iso-propyl, and cyclohexyl) were those giving the highest syndiotactic content. Therefore, the results obtained with the two novel systems described in the present paper confirm once more that the steric hindrance of the ligand exerts a strong influence on the mutual orientation of the allylic unit of the growing chain and of the new incoming monomer that is responsible for the polymerization stereoselectivity.
All the obtained polybutadienes were semicrystalline (the X-ray powder spectra are reported in the Supplementary Materials, Figures S14−S18), with a melting point in the range from 141 to 161 °C, and the crystallinity degree essentially depending on the polymer syndiotacticity.

3. Materials and Methods

3.1. General Procedure and Materials

Tert-butyl(diphenyl)phosphine (PtBuPh2) (Aldrich, St. Louis, MO, USA, 97%), (S)-(+)neomenthyldiphenylphosphine [(S)-NMDPP] (Strem Chemicals, Newburyport, MA, USA, 98%), anhydrous cobalt dichloride (Aldrich, 99.9%), and MAO (Aldrich, 10 wt.% solution in toluene) were used as received. Ethyl alcohol (Aldrich, ≥99.8%) was degassed under vacuum, then by bubbling dry dinitrogen and kept over molecular sieves; pentane (Aldrich, ≥99.5%) was refluxed over Na/K alloy for ca. 8 h, then distilled and stored over molecular sieves under dry dinitrogen; toluene (Aldrich, ≥99.5% pure) was refluxed over Na for ca. 8h, then distilled and stored over molecular sieves under dry dinitrogen. 1,3-Butadiene (Aldrich, ≥99.5%) was evaporated from the container prior to each run, dried by passing through a column packed with molecular sieves and condensed into the reactor which had been precooled to −20 °C. All the phosphine cobalt complexes were synthesized as indicated below, following a general procedure already reported in the literature [23].

3.2. Synthesis of Cobalt Phosphine Complexes

3.2.1. CoCl2(PtBuPh2) (1)

Cobalt dichloride (CoCl2) anhydrous (1.24 g, 9.6 mmol) and 40 mL of ethanol were placed in a 250 mL flask: the blue solution obtained was kept under stirring, at 22 °C, for about 2 hours. Subsequently, tert-butyldiphenylphosphine (PtBuPh2) (5.1 g, 21 mmol) dissolved in ethanol (30 mL) was added: everything was kept under stirring, at room temperature, for 24 hours and subsequently the solvent was removed almost completely under vacuum. Then, pentane (40 mL) was added and the obtained suspension was kept under stirring at room temperature for 2 hours: at the end, the light blue/blue suspension obtained was filtered, the residue obtained was washed further with cold pentane (2 × 10 mL) and then dried under vacuum at room temperature. The blue solid obtained was then continuously extracted with boiling pentane; crystals of CoCl2(PtBuPh2)2 were formed directly on the bottom of the Schlenk tube during the extraction and further crops of crystals were obtained by cooling the supernatant pentane solution at −30 °C.
Yield, 4.71 g, 80% conversion based on the loaded cobalt dichloride. Elementary analysis [found (calculated) for C32H38Cl2CoP2, m.w. 614,43]: C: 62.50% (62.55%); H: 6.20% (6.23%); Cl: 11.50% (11.54%); P: 10.00% (10.08%); Co: 9.50% (9.59%).
The crystallographic data obtained are shown in Table 3. The FTIR spectrum of 1 is shown in Figure S19 of the Supplementary Materials.

3.2.2. CoCl2[PPh2(NMDPP)]2 (2)

Cobalt dichloride (CoCl2) anhydrous (0.182 g, 1.4 mmol) and 20 mL of ethanol were placed in a 100 mL flask: the blue solution obtained was kept under stirring at room temperature for 1 hour. Subsequently, (S)-(+)neomenthyl-diphenylphosphine [(S)-NMDPP] (1.0 g, 3.08 mmol) dissolved in ethanol (30 mL) was added: the solution obtained was brought to 60 °C and kept under stirring at this temperature for 24 hours. The solvent was removed under vacuum, then pentane (30 mL) was added and the suspension was kept under stirring at room temperature for 2 hours. At the end, the light blue/blue suspension obtained was filtered and the residue obtained was further washed with pentane (2 × 10 mL) and then dried under vacuum at room temperature. Yield, 0.950 g of a light blue powder, 87% conversion with respect to the loaded cobalt dichloride.
Elementary analysis [found (calculated) for C44H58Cl2CoP2, m.w. 778,72]: C: 67.60% (67.86%); H: 7.40% (7.51%); Cl: 8.90% (9.11%); P: 7.90% (7.96%); Co: 7.50% (7.57%).
The FTIR spectrum of 2 is shown in Figure S20 of the Supplementary Materials.

3.3. X-ray Crystallographic Studies

A summary of the experimental details concerning the X-ray diffraction study of 1 is reported in Table 3. The crystals used for data collection were entirely covered with perfluorinated oil to reduce crystal decay. X-ray data were collected on a Bruker Smart Apex CCD area detector (Bruker AXS Inc., Madison, WI, USA) equipped with fine-focus sealed tube operating at 50 kV and 30 mA, using graphite-monochromated Mo Kα radiation (λ = 0.71073 Å). Data reduction was made using SAINT programs [24]; absorption corrections based on multiscan were obtained by SADABS [24]. The structures were solved by SHELXS-97 [25] and refined on F2 by full-matrix least-squares using SHELXL-14 [26]. The program ORTEP-III [27] was used for molecular graphics.

3.4. Polymerization

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

3.5. Polymer Characterization

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.%. The polymer microstructure was determined as reported in the literature [28,29]. Differential scanning calorimetry (DSC) scans were carried out on a Perkin Elmer Pyris 8000 (Waltham, MA, USA). Typically, ca. 10 mg of polymer was analyzed in each run, while the scan speed was 20 °C/min under a dinitrogen atmosphere. Wide-angle X-ray diffraction (XRD) experiments were performed at 25 °C under nitrogen flux, using a Siemens D-500 diffractometer equipped with Soller slits (2°) placed before samples, 0.3° aperture and divergence windows, and a VORTEX detector with extreme energy resolution specific for thinner films. Cu Kα radiation with 40 kV × 40 mA power used was adopted, and each spectrum was carried out with steps of 0.05° 2θ, and 6 s measure 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 average molecular weight (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, 0.8 mL/min flow rate, and 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

We synthesized and characterized two novel cobalt diphenylphosphine complexes which, in combination with MAO, gave from 1,3-butadiene highly syndiotactic 1,2 polymers. The results obtained were as expected on the basis of the mechanistic hypotheses previously formulated [12,13], confirming their validity.

Supplementary Materials

Tables of atomic coordinates, anisotropic thermal parameters, bond lengths and angles of 1 may be obtained free of charge from The Director CCDC, 12 Union Road, Cambridge CB2 1 EZ, UK, on quoting the deposition numbers CCDC 1908992 the names of the authors and the journal citation (http://www.ccdc.cam.ac.uk). The following are available online: Figures S1−S6: 13C NMR spectra (olefinic region) of the obtained poly(1,3-butadiene)s, Figures S7−S13: FTIR spectra of the obtained poly(1,3-butadiene)s, Figures S14−S18: X-ray powder spectra of some selected poly(1,3-butadiene)s, Figure S19: FTIR spectrum of 1, Figure S20: FTIR spectrum of 2.

Author Contributions

Polymers and organometallic synthesis and characterization, G.L., I.P., and G.Z.; X-ray studies, A.F.; technical teaching and supervision, G.L.; project administration, conceptualization, supervision, and original draft preparation, G.R.

Funding

This research received no external funding.

Acknowledgments

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Porri, L.; Giarrusso, A. Conjugated Diene Polymerization. In Comprehensive Polymer Science; Eastmond, G., Edwith, A., Russo, S., Sigwalt, P., Eds.; Pergamon Press Ltd.: Oxford, UK, 1989; Volume 4, p. 53. [Google Scholar]
  2. Thiele, S.K.H.; Wilson, D.R. Alternate Transition Metal Complex Based Diene Polymerization. J. Macromol. Sci. Part C 2003, 43, 581–628. [Google Scholar] [CrossRef]
  3. Osakada, K.; Takeuchi, D. Coordination Polymerization of Dienes, Allenes, and Methylcycloalkanes. Adv. Polym. Sci. 2004, 171, 137–194. [Google Scholar]
  4. Ricci, G.; Masi, F.; Boglia, A.; Sommazzi, A.; Ricci, M. Cobalt phosphine complexes for the polymerization of 1,3-dienes. Influence of the phosphine ligand on the polymerization chemo- and stereoselectivity. In Advances in Organometallic Chemistry Research; Yamamoto, K., Ed.; Nova Science Publisher, Inc.: Hauppauge, NY, USA, 2007; pp. 1–36. [Google Scholar]
  5. Ricci, G.; Leone, G.; Boglia, A.; Masi, F.; Sommazzi, A. Cobalt compounds based catalysts for the stereospecific polymerization of olefins, cycloolefins and conjugated diolefins. In Cobalt: Characteristics, Compounds, and Applications; Lucas, J., Ed.; Nova Science Publisher, Inc.: Hauppauge, NY, USA, 2011; pp. 39–81. [Google Scholar]
  6. Friebe, L.; Nuyken, O.; Obrecht, W. Neodymium-Based Ziegler/Natta Catalysts and their Application in Diene Polymerization. Adv. Polym. Sci. 2006, 204, 1–154. [Google Scholar]
  7. Takeuchi, M.; Shiono, T.; Soga, K. Polymerization of 1,3-butadiene with the catalyst system composed of a cobalt compound and methylaluminoxane. Polym. Int. 1992, 29, 209–212. [Google Scholar] [CrossRef]
  8. 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 associated with MAO. Macromolecules 2005, 38, 1064–1070. [Google Scholar] [CrossRef]
  9. 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]
  10. Ricci, G.; Leone, G.; Boglia, A.; Bertini, F.; Boccia, A.C.; Zetta, L. Synthesis and Characterization of Isotactic 1,2-Poly(E-3-methyl-1,3-pentadiene). Some Remarks about the Influence of Monomer Structure on Polymerization Stereoselectivity. Macromolecules 2009, 42, 3048–3056. [Google Scholar] [CrossRef]
  11. Boccia, A.C.; Leone, G.; Boglia, A.; Ricci, G. Novel Stereoregular cis-1,4 and trans-1,2 Poly(diene)s: Synthesis, Characterization, and Mechanistic Considerations. Polymer 2013, 54, 3492–3503. [Google Scholar] [CrossRef]
  12. 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]
  13. 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]
  14. Ricci, G.; Leone, G. Recent progresses in the polymerization of butadiene over the last decade. Polyolefins J. 2014, 1, 43–60. [Google Scholar]
  15. 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-VCH Verlag GmbH: Weinheim, Germany, 2017; Volume 1, pp. 251–273. [Google Scholar]
  16. 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]
  17. Furukawa, J.; Haga, K.; Kobayashi, E.; Iseda, Y.; Yoshimoto, T.; Sakamoto, K. cis-Vinyl-1:1 Polymer of Butadiene. Polym. J. 1971, 2, 371–378. [Google Scholar] [CrossRef] [Green Version]
  18. Ashitaka, H.; Ishikawa, H.; Ueno, H.J. Syndiotactic 1,2-Polybutadiene with Co-CS2 catalyst System. I. Pteparation, Properties, and Application of Highly Crystalline Syndiotactic 1,2-Polybutadiene. Polym. Sci. Polym. Chem. Ed. 1983, 21, 1853–1860. [Google Scholar] [CrossRef]
  19. Ashitaka, H.; Jinda, K.; Ueno, H. Syndiotactic 1,2-Polybutadiene with Co-CS2 Catalyst System. II Catalysts for Stereospecific Polymerization of Butadiene to Syndiotactic 1,2-Polybutadiene. J. Polym. Sci. Polym. Chem. Ed. 1983, 21, 1951–1972. [Google Scholar] [CrossRef]
  20. Ashitaka, H.; Inaishi, K.; Ueno, H. Synditactic 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. Ashitaka, H.; Jinda, K.; Ueno, H. Syndiotactic 1,2-Polybutadiene with Co-CS2 Catalyst System. IV. Mechanism of Syndiotactic Polymerization of Butadiene with Cobalt Compounds-Organoaluminum-CS2. J. Polym. Sci. Polym. Chem. Ed. 1983, 21, 1989–1995. [Google Scholar] [CrossRef]
  22. Ricci, G.; Italia, S.; Porri, L. Polymerization of butadiene to 1,2-syndiotactic polymer with (η3-C8H13)(C4H6)Co. Some observations on the factors that determine the stereospecificity. Polym. Commun. 1988, 29, 305–307. [Google Scholar]
  23. Chatt, J.; Shaw, B.L. Alkyls and aryls of transition metals. Part IV. Cobalt(II) and iron(II) derivatives. J. Chem. Soc. 1961, 285–290. [Google Scholar] [CrossRef]
  24. Bruker. Smart, Saint and Sadabs; Bruker AXS Inc.: Madison, WI, USA, 1997. [Google Scholar]
  25. Sheldrick, G.M. A short history of SHELX. Acta Cryst. Sect. A Found. Crystallogr. 2008, 64, 112–122. [Google Scholar] [CrossRef]
  26. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  27. Burnett, M.N.; Johnson, C.K. ORTEP-III: Oak Ridge Thermal Ellipsoid Plot Program for Crystal Structure Illustrations, Oak Ridge National Laboratory Report ORNL-6895. Acta Crystallogr. Sect. A Found. Crystallogr. 1996, 52, C93. [Google Scholar] [CrossRef]
  28. Mochel, V.D. Carbon-13 NMR of polybutadiene. J. Polym. Sci. Part A-1 Polym. Chem. 1972, 10, 1009–1018. [Google Scholar] [CrossRef]
  29. Elgert, K.F.; Quack, G.; Stutzel, B. Zur struktur des polybutadiens, 2. Das 13C-NMR-Spektrum des 1,2-polybutadiens. Makromol. Chem. 1974, 175, 1955–1960. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compounds are available from the authors.
Figure 1. Bis-diphenylphosphine cobalt complexes used in this work.
Figure 1. Bis-diphenylphosphine cobalt complexes used in this work.
Molecules 24 02308 g001
Figure 2. Molecular structure of 1 with thermal ellipsoids drawn at 50% probability level.
Figure 2. Molecular structure of 1 with thermal ellipsoids drawn at 50% probability level.
Molecules 24 02308 g002
Figure 3. 13C-NMR spectrum (olefinic region) of polybutadiene of Table 2, run 2.
Figure 3. 13C-NMR spectrum (olefinic region) of polybutadiene of Table 2, run 2.
Molecules 24 02308 g003
Table 1. Selected bond lengths (Å) and angles (°) for CoCl2(PtBuPh2)2 (1).
Table 1. Selected bond lengths (Å) and angles (°) for CoCl2(PtBuPh2)2 (1).
Bond LengthsBond Angles
Co–Cl12.2415(4)Cl1–Co–Cl2112.41(2)
Co–Cl22.2298(4)P1–Co–P2113.69(1)
Co–P12.4368(4)C1–P1–C5104.17(6)
Co–P22.4300(4)C1–P1–C11105.68(6)
P1–C11.8870(13)C17–P2–C2 106.34(6)
P2–C171.8886(14)C17–P2–C27103.12(6)
P1–C51.8258(13)C5–P1–C11108.39(6)
P1–C111.8317(14)C21–P2–C27107.45(6)
P2–C211.8271(14)
P2–C271.8195(13)
Table 2. Polymerization of 1,3-butadiene with cobalt catalytic systems.1.
Table 2. Polymerization of 1,3-butadiene with cobalt catalytic systems.1.
RunCatalystPolymerizationPolymer
Co-complex (μmol)Al/CosolventT
(°C)
Time
(min)
Yield
(%)
1,22
(%)
rr3
(%)
Mw4
(×103)
Mw/Mn4m.p.5
(°C)
1CoCl2(PtBuPh2)2 (10)100toluene201010085.479.5175.62.4140.9
2CoCl2(PtBuPh2)2 (30)10toluene203010085.680.6197.02.2146.3
3CoCl2(PtBuPh2)2 (30)10heptane206010090.882.3186.72.1153.5
4CoCl2(PtBuPh2)2 (30)25heptane06010094.284.1225.61.9158.2
5CoCl2[PPh2(NMDPP)]2 (10)100toluene201510086.081.3203.42.3141.8
6CoCl2[PPh2(NMDPP)]2 (20)25heptane206010088.383.7196.02.1154.7
7CoCl2[PPh2(NMDPP)]2 (30)25heptane0107294.585.5237.22.0160.8
8CoCl2(PiPrPh2)2 (5)100toluene20510085.474.0169.02.2126.0
9CoCl2(PCyPh2)2 (5)100toluene2057584.569.0172.02.4109.0
1 Polymerization conditions: solvent, total volume 16 mL; butadiene, 2 mL; 2 percentage of 1,2 units, determined by 1H- and 13C-NMR; the remaining units are essentially cis-1,4, the trans-1,4 units being almost negligible; 3 percentage of syndiotactic triads, determined by 13C-NMR; 4 average molecular weight (Mw, in g mol−1) and molecular weight distribution (Mw/Mn) by SEC; 5 melting point by DSC.
Table 3. Crystallographic data, data collection details, and results of refinement for complex 1.
Table 3. Crystallographic data, data collection details, and results of refinement for complex 1.
formula, Mr C32H38Cl2CoP2, 614.39
crystal systemMonoclinic
space group, Z P21/n, 4
Dcalc, g cm−31.347
a, Å11.9782(5)
b, Å15.4120(6)
c, Å16.7851(7)
β, °102.098(1)
V, Å33029.8(2)
crystal size, mm0.35 × 0.20 × 0.08
color, habitlight blue, plate
μ, mm−10.868
radiationMoKα
T, K153(2)
2θmax, °63.69
h, k, l ranges −17→17; −22→22; −24→24
intensity decay, %0.00
adsorption correctionmulti-scan
Tmin, Tmax0.682, 0.746
measured reflections60180
Rint0.0320
independent reflections9958
reflections with I > 2σ(I)8422
no. of parameters 340
R, wR [F2 > 2σ( F2)] 0.0353, 0.0900
goodness of fit1.048
Δρ max, Δρ min (eÅ−3) 1.069, −0.231

Share and Cite

MDPI and ACS Style

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. https://doi.org/10.3390/molecules24122308

AMA Style

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(12):2308. https://doi.org/10.3390/molecules24122308

Chicago/Turabian Style

Ricci, Giovanni, Giuseppe Leone, Ivana Pierro, Giorgia Zanchin, and Alessandra Forni. 2019. "Novel Cobalt Dichloride Complexes with Hindered Diphenylphosphine Ligands: Synthesis, Characterization, and Behavior in the Polymerization of Butadiene" Molecules 24, no. 12: 2308. https://doi.org/10.3390/molecules24122308

Article Metrics

Back to TopTop