Next Article in Journal
Generational Advancements in the Transverse Shear Strength Retention of Glass Fiber-Reinforced Polymer Bars in Alkaline and Acidic Environments
Previous Article in Journal
Enhanced Synthesis of Poly(1,4-butanediol itaconate) via Box–Behnken Design Optimization
Previous Article in Special Issue
Experimental and Theoretical Insights into the Effect of Dioldibenzoate Isomers on the Performance of Polypropylene Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Syndiotactic Polyolefins by Hydrogenation of Highly Stereoregular 1,2 Polydienes: Synthesis and Structural Characterization

by
Giovanni Ricci
1,*,
Ivana Pierro
2 and
Antonella Caterina Boccia
1
1
CNR-Istituto di Scienze e Tecnologie Chimiche “Giulio Natta” (SCITEC), Via A. Corti 12, I-20133 Milano, Italy
2
Scientific Advisor, I-21052 Busto Arsizio (VA), Italy
*
Author to whom correspondence should be addressed.
Polymers 2024, 16(19), 2711; https://doi.org/10.3390/polym16192711
Submission received: 7 August 2024 / Revised: 17 September 2024 / Accepted: 20 September 2024 / Published: 25 September 2024
(This article belongs to the Special Issue Catalytic Olefin Polymerization and Polyolefin Materials)

Abstract

:
Some syndiotactic-rich polyolefins, generally difficult to synthesize through stereospecific polymerization of the corresponding monomers, were prepared by homogeneous non-catalytic hydrogenation of syndiotactic 1,2 poly(1,3-diene)s with diimide, arising from thermal decomposition of p-toluene-sulfonyl-hydrazide. All the polymers synthesized were structurally characterized by means of several analytical techniques, such as FT-IR, NMR (1H, 13C and 2D), DSC, and GPC, and herein illustrated.

1. Introduction

In the last years the stereospecific polymerization of 1,3-dienes (i.e., 1,3-butadiene and substituted 1,3-butadienes) [1] has mainly focused on catalytic systems based on the combination of aluminoxanes (e.g., methylaluminoxane (MAO), tetraisobutylaluminoxane (TIBAO)), and different transition metal (e.g., Ti, V, Cr, Co, Fe, Cu) and lanthanide (e.g., Nd, Pr, Gd, La) complexes (e.g., metallocene complexes, phosphine complexes, bypiridyl complexes, bis-imine complexes, imino-pyridine complexes, bis-imino-pyridine complexes) [2,3]. This trend has permitted the synthesis of highly stereoregular poly(1,3-diene)s with different structures (cis-1,4; trans-1,4; 1,2; 3,4; iso- and syndiotactic) depending on monomer and catalyst structure [4]. In particular, highly syndiotactic 1,2 polymers were obtained with the system CpTiCl3/MAO from (Z)-1,3-pentadiene [5,6] and 4-methyl-1,3-pentadiene [6]; highly syndiotactic 1,2 polybutadiene was obtained with chromium (e.g., CrCl2 (dmpe)2/MAO (dmpe = 1,2-bis (dimethylphosphino)ethane)) [7], cobalt (e.g., (η3-C8H13)(C4H6)Co) [8] and iron (e.g., FeCl2 (bipy)2/MAO) [9,10] catalysts; syndiotactic 1,2 polymers were also prepared from terminally substituted 1,3-butadienes (e.g., 1,3-pentadiene, 1,3-hexadiene, 5-methyl-1,3-hexadiene, 1,3-heptadiene, 1,3-octadiene) using the system CoCl2 (PiPrPh2)2/MAO [11,12,13], while isotactic (E)-1,2 poly(3-methyl-1,3-pentadiene) was synthesized with the system CoCl2 (PnPrPh2)2/MAO [14]; syndiotactic 3,4 polyisoprene [9,15,16,17] was obtained with iron- and copper-based catalysts (i.e., FeEt2 (bipy)2-MAO, FeCl2 (bipy)2/MAO and CuCl2 (bipy)/MAO (bipy = bipyridine)); syndiotactic (E)-1,2 poly(3-methyl-1,3-pentadiene) [9,15,18] was instead obtained with the systems FeEt2 (bipy)2-MAO and FeCl2 (bipy)2/MAO (bipy = bipyridine).
The results obtained allowed us to improve the knowledge on the diene polymerization mechanism [1,2,3,4], in particular highlighting the strong impact of the monomer structure (i.e., presence of substituents on the monomeric unit) and of the catalyst structure (i.e., nature of the ligand on the metal atom) on the polymerization regio- and stereo-selectivity.
Nevertheless, the extensive accessibility of all these extremely stereoregular polydienes may represent an exceptional and considerable source for highly regular olefin polymers through a simple hydrogenation reaction of the polydienes, as illustrated in Scheme 1, especially in the case where such polyolefins cannot be obtained with such a high degree of stereoregularity by simple stereospecific polymerization of the corresponding olefins with transition metal catalysts.
This hydrogenation methodology has been previously applied to prepare highly stereoregular isotactic poly((R,S)-3-methyl-1-pentene) [19,20], and syndiotactic poly(3-methyl-1-butene) [21], that were microstructurally characterized.
In this study, we now report on the preparation and structural characterization by NMR—nuclear magnetic resonance spectroscopy (1H, 13C and 2D)—of the following syndiotactic-rich polyolefins synthesized by hydrogenation of the diene polymers obtained with the catalytic system CoCl2 (PiPrPh2)2-MAO (Scheme 1): (i) syndiotactic poly(1-pentene) [H(1,2syPP)] from syndiotactic trans-1,2 poly(1,3-pentadiene) (1,2syPP) [11]; (ii) syndiotactic poly(1-hexene) [H(1,2syPHX)] from syndiotactic trans-1,2 poly(1,3-hexadiene) (1,2syPHX) [11,13]; (iii) syndiotactic poly(5-methyl-1-hexene) [H(1,2syP5MHX)] from trans-1,2 poly(5-methyl-1,3-hexadiene) (1,2syP5MHX) [12]; (iv) syndiotactic poly(1-heptene) [H(1,2syPHP)] from syndiotactic trans-1,2 poly(1,3-heptadiene) (1,2syPHP) [12]; (v) syndiotactic poly(1-octene) [H(1,2syPO)] from trans-1,2 poly(1,3-octadiene) (1,2syPO) [12].

2. Materials and Methods

2.1. General Procedures and Materials

All the operations were carried out under an inert atmosphere using a dual vacuum/nitrogen line and standard Schlenk-line techniques. Toluene (Sigma-Aldrich, ≥99.7% pure) was refluxed over Na for about 8 h and then distilled and stored over molecular sieves under nitrogen. o-Xylene (Sigma-Aldrich, Milano, MI, Italy, pure anhydrous grade), p-toluenesulfonylhydrazide (TSH, Sigma-Aldrich, Milano, MI, Italy), and deuterated solvent for NMR measurements (C2D2Cl4) (Cambridge Isotope Laboratories, Inc., Tewksbury, MA, USA) were used as received. (E)-1,3-pentadiene (Sigma-Aldrich, Milano, MI, Italy, 96% pure), 1,3-hexadiene (Aldrich, Milano, MI, Italy, 99% pure, mixture of (E) and (Z) isomers), 5-methyl-1,3-hexadiene (Chemsampco, Dallas, TX, USA, 80% pure; predominantly E isomer), 1,3-heptadiene (Chemsampco, Dallas, TX, USA, 99% pure; predominantly E isomer), and 1,3-octadiene (Chemsampco, Dallas, TX, USA, 99% pure; predominantly E isomer) were refluxed over calcium hydride for about 4 h, then distilled trap-to-trap and stored under dry nitrogen. Methylaluminoxane (MAO) (10 wt% solution in toluene, Sigma-Aldrich) was used as received. CoCl2 (PiPrPh2)2 [11], syndiotactic trans-1,2 poly(1,3-pentadiene) (1,2syPP) [11], syndiotactic trans-1,2 poly(1,3-hexadiene) (1,2syPHX) [11,13], syndiotactic trans-1,2 poly(5-methyl-1,3-hexadiene) (1,2syP5MHX) [12], syndiotactic trans-1,2 poly(1,3-heptadiene) (1,2syPHP) [12], and syndiotactic trans-1,2 poly(1,3-octadiene) (1,2syPO) [12] were prepared as reported in the literature.

2.2. Hydrogenation Procedure

The poly(1,3-diene) was introduced into a flask and xylene was subsequently added. The reaction mixture thus obtained was kept under vigorous stirring at room temperature until complete solubilization of the polymer, and then TSH was added. The resulting suspension was refluxed at 120 °C for 3 days, after which it was allowed to spontaneously reach room temperature before adding again the TSH. This operation was repeated once again. Once the reaction was completed, the hydrogenated sample was hot-filtered, the volume of the solution was reduced under vacuum, and methanol was added to coagulate the polymer, which was successively recuperated by filtration. The polymer, once dried under vacuum at room temperature, was extracted with acetone with a Soxhlet for 10 h to remove any excess TSH and byproducts from TSH decomposition. The residual polymer was finally dried under vacuum, dissolved in toluene, precipitated into methanol, and dried again under vacuum at room temperature to constant weight.
The following reaction conditions were applied for the hydrogenation of each polymer:
1,2syPP_Syndiotactic trans-1,2 poly(1,3-pentadiene): polymer 1 g; xylene, 200 mL; first addition of THS, 9.0 g (5.0 × 10−2 mol); second and third addition of THS, 10.0 g (5.56 × 10−2 mol). Final yield of H(1,2syPP): 0.824 g;
1,2syPHX_Syndiotactic trans-1,2 poly(1,3-hexadiene): polymer 1.9 g; xylene, 300 mL; first addition of THS, 10.0 g (0.54 × 10−1 mol); second and third addition of THS, 20.0 g (1.07 × 10−1 mol). Final yield of H(1,2syPHX): 1.64 g;
1,2syP5MHX_Syndiotactic trans-1,2 poly(5-methyl-1,3-hexadiene): polymer 1.09 g; xylene, 250 mL; first addition of THS, 9.0 g (4.8 × 10−2 mol); second and third addition of THS, 11.0 g (5.9 × 10−2 mol). Final yield of H(1,2syP5MHX): 0.89 g;
1,2syPHP_Syndiotactic trans-1,2 poly(1,3-heptadiene): polymer 0.935 g; xylene, 80 mL; first addition of THS, 6.6 g (3.5 × 10−2 mol); second and third addition of THS, 13.0 g (6.9 × 10−2 mol). Final yield of H(1,2syPHP): 0.82 g;
1,2syPO_Syndiotactic trans-1,2 poly(1,3-octadiene): polymer 1.18 g; xylene, 250 mL; first addition of THS, 9.6 g (5.1 × 10−2 mol); second and third addition of THS, 10.0 g (5.9 × 10−2 mol). Final yield of H(1,2syPO): 0.83 g.

2.3. Polymer Characterization

Infrared spectroscopy (FTIR) spectra were recorded at room temperature in the 4000–600 cm−1 range using a Perkin Elmer Spectrum Two spectrometer (Perkin Elmer Italia S.p.A. 20900 Monza, Italia).
13C and 1H NMR spectra were recorded on Bruker Avance 400 spectrometer. The spectra were obtained in C2D2Cl4 at 103 °C (hexamethyldisiloxane, HMDS, as internal standard). The concentration of polymer solutions was about 10 wt%. 13C parameters were: spectral width 17 kHz; 90° pulse 11.0 μs PL1 −5.0 dB, with a delay of 16 s.
Two-dimensional heteronuclear 1H-13C experiments were recorded on a Bruker DRX 600 MHz spectrometer, (14.1 T), thermostated at 330 K. The g-HSQC experiment, (gradient-Heteronuclear Single Quantum Correlation), was performed by applying a coupling constant 1JCH = 140 Hz; data matrix 2 K × 512; number of scans:128; 7.47 μs as 90° pulse. The g-HMBC experiments (gradient-heteronuclear multiple bond correlation) were performed by applying a delay of 50 ms for the evolution of long-range coupling; data matrix 2 K × 512; number 150 of scans 128; D1 2.00 s. Data were zero-filled and weighted with a sine bell function before Fourier transformation. The microstructure of the resultant polymers was determined by 1H and 13C NMR, according to the literature data.
The molecular weight averages (Mw) and the molecular weight distribution (Mw/Mn) were determined by a high-temperature Waters GPCV2000 size exclusion chromatography (SEC) system using two online detectors: a differential viscometer and a refractometer. The experimental conditions consisted of three PL Gel Olexis columns, o-DCB (ortho dichlorobenzene) as the mobile phase, 0.8 mL min−1 flow rate, and 145 °C temperature. Universal calibration of the SEC system was performed using eighteen narrow Mw/Mn polystyrene standards with molar weights ranging from 162 to 5.6 × 106 g mol−1. For the analysis, about 12 mg of the polymer was dissolved in 5 mL of o-DCB with 0.05% of BHT as the antioxidant.

3. Results and Discussion

Data concerning the preparation of syndiotactic 1,2 poly(1,3-diene)s, such as (1,2syPP) [11], (1,2syPHX) [11,13], (1,2syP5MHX) [12], (1,2syPHP) [12] and (1,2syPO) [12], are summarized in Table 1.
The non-catalytic hydrogenation with diimide [22,23,24,25,26] of the above-reported polymers herein illustrated (Scheme 1) was carried out at 120 °C in o-xylene in homogeneous conditions. Practically, the diimide molecule (N2H2), formed in situ through the thermal decomposition of p-toluenesulfonyl hydrazide (TSH), is capable to release the hydrogen molecule responsible of the hydrogenation reaction of the olefinic double bond.
Hereafter it is reported the adopted nomenclature referring to the hydrogenated polymers: [H(1,2syPP)] is the syndiotactic poly(1-pentene), [H(1,2syPHX)] is the syndiotactic poly(1-hexene), [H(1,2syP5MHX)] is the syndiotactic poly(5-methyl-1-hexene), [H(1,2syPHP)] is the syndiotactic poly(1-heptene)}, and [H(1,2syPO)] is the syndiotactic poly(1-octene).
Data concerning the syndiotactic degree, which essentially remains that of the starting 1,2 poly(1,3-diene), molecular weight and molecular weight distribution for all the hydrogenated polymers are summarized in Table 2.

3.1. Syndiotactic Poly(1-pentene)_H(1,2syPP)

Syndiotactic-rich poly(pentene) was prepared via hydrogenation of trans-1,2 poly(1,3-pentadiene), synthesized with the catalyst system CoCl2 (PiPrPh2)2/MAO as reported in Section 2.2, and illustrated in Scheme 2.
NMR data, such as 1H and 13C NMR spectra (Figure 1), confirmed the completeness of the hydrogenation diene due to the disappearance of the characteristic olefinic signals in the spectra.
Figure 1 shows the comparison of NMR data of syndiotactic trans-1,2 poly(1,3-pentadiene) and of the hydrogenated product.
The structure and tacticity of the obtained syndiotactic poly(1-pentene) was analyzed by means of 1H and 13C NMR and verified according to literature data [27,28,29].

3.2. Syndiotactic Poly(1-hexene)_H(1,2syPHX)

Syndiotactic-rich poly(1-hexene) [H(1,2syPHX)] has been prepared by hydrogenation of syndiotactic trans-1,2 poly(1,3-hexadiene), synthesized with the catalyst system CoCl2 (PiPrPh2)2/MAO, as shown in Scheme 3.
The hydrogenation product (i.e., syndiotactic poly(1-hexene)) features are listed in Table 2.
Comparing the 1H and 13C NMR spectra (Figure 2) of the starting syndiotactic trans-1,2 poly(1,3-hexadiene) with those of the corresponding hydrogenated product, it was possible to establish the successful complete hydrogenation reaction.
The structure and tacticity of the resulting syndiotactic poly(1-hexene) was examined through 1H and 13C NMR, as shown in Figure S3 [28,29,30,31].
The microstructure of the hydrogenated polymer was investigated through two-dimensional NMR experiments, such as HSQC (heteronuclear single quantum coherence), which is used to determine the proton–carbon single-bond correlations. As for the polypropylene, the simplest polymer, the syndiotactic poly(1-hexene) obtained after the hydrogenation process of the trans-1,2 poly(1,3-hexadiene) has the two protons on the carbon C1 magnetically equivalent, as they experienced the same chemical environment. The NMR equivalence of the C1 protons was confirmed by the presence in the HSQC spectrum (Figure S4) of a single cross peak (at δC = 39.66 ppm) associated to the single bond correlation of the two protons (δC = 1.037 ppm) linked to the C1 carbon atom. Practically, the saturated polymer retains the microstructure of the native polymer even after the hydrogenation process. The correctness of the protons and carbons assignment was also confirmed trough the HMBC experiments (Figure S4) which provided correlations between protons and carbons that are two or three bonds away.

3.3. Syndiotactic Poly(5-methyl-1-hexene) [H(1,2syP5MHX)]

The syndiotactic-rich poly(5-methyl-1-hexene) was obtained by hydrogenation of the syndiotactic trans-1,2 poly(5-methyl-1,3-hexadiene) obtained by polymerizing 5-methyl-1,3-hexadiene with the system CoCl2 (PiPrPh2)2-MAO (Scheme 4).
The hydrogenation product (i.e., syndiotactic poly(5-methyl-1-hexene) features are listed in Table 2.
The completion of the hydrogenation reaction was confirmed by comparing of the FT-IR spectra (Figure S1) and the NMR (1H and 13C; Figure 3) spectra of the starting syndiotactic trans-1,2 poly(1,3-diene) to those of the corresponding hydrogenated product.
In the FT-IR spectrum of the syndiotactic poly(5-methyl-1-hexene) (Figure S1), the strong band at 965 cm−1 observed in the spectrum of trans-1,2 poly(5-methyl-1,3-hexadiene), indicative of the presence of a trans double bond, is completely absent.
The structure and tacticity of the resulting syndiotactic 1,2 poly(5-methyl-1-hexene) were determined through NMR (1H and 13C). Figure 3 shows the 13C NMR spectra of the syndiotactic trans-1,2 poly(5-methyl-1,3-hexadiene) and its saturated polymer together with the peak attribution.

3.4. Syndiotactic Poly(1-heptene)_H(1,2syPHP)

Syndiotactic-rich poly(1-heptene) was obtained by hydrogenation of syndiotactic trans-1,2 poly(1,3-heptadiene) synthesized by polymerizing 1,3-heptadiene with the system CoCl2 (PiPrPh2)2-MAO (Scheme 5).
The hydrogenation product (i.e., syndiotactic poly(1-heptene) features are listed in Table 2.
The complete hydrogenation of the diene polymer was evident by comparing the 1H NMR spectra (Figure 4) of the starting syndiotactic trans-1,2 poly(1,3-heptadiene) with the spectra of the resulting hydrogenated product. The signals in the olefinic region (from 5.2 to 5.4 ppm), detected in the 1H NMR spectrum of the diene polymer and due to the olefinic hydrogen atoms, are not observed in the 1H NMR spectrum of the hydrogenated polymer, according indeed with the complete hydrogenation of the diene polymer.
The structure and tacticity of the resulting syndiotactic poly(heptene) were examined by 1H and 13C NMR [28,31,32]. Figure 4 shows the 13C NMR spectra of the syndiotactic trans-1,2 poly(heptadiene) and its saturated polymer together with the peak attribution.
The microstructure of the hydrogenated syndiotactic poly(1-heptene) was investigated through two-dimensional HSQC experiments (Figure S5), observing that the two protons on the carbon C1 were magnetically equivalent meaning they experienced the same chemical environment. The NMR equivalence of the C1 protons was confirmed by the presence in the HSQC spectrum of a single cross peak associated to the single bond correlation of the two protons (at δC = 1.02 ppm) linked to the C1 carbon atom (δC = 39.72 ppm). As for the H(1,2syPHX) polymer, the correctness of the protons and carbons assignment was confirmed with the HMBC experiments (Figure S5).

3.5. Syndiotactic Poly(1-octene)_[H(1,2syPO)]

Syndiotactic trans-1,2 poly(1,3-octadiene), obtained by polymerizing 1,3-octadiene with CoCl2 (PiPrPh2)2-MAO, was treated with tosylhydrazide providing syndiotactic poly(1-octene) (Scheme 6).
The hydrogenation product (i.e., syndiotactic poly(1-octene) features are listed in Table 2.
The comparison between the two 1H NMR spectra (Figure 5) of the starting syndiotactic trans-1,2 poly(1,3-diene) and of the resulting hydrogenated product highlights the complete hydrogenation of the starting polydiene.
In the FT-IR spectrum of the syndiotactic poly(octene) (Figure S2), the strong band at 967 cm−1 indicative of the presence of a trans double bond is not detectable, while an intense band at 735 cm−1, attributed to the vibration of a –CH2– unit, typical of saturated polyolefins, was observed.
The peaks in the olefinic region (from 5.2 to 5.4 ppm), detected in the 1H NMR spectrum (Figure 5) of the diene polymer and due to the olefinic hydrogen atoms, are not visible in the 1H NMR spectrum of the hydrogenated polymer, confirming that the hydrogenation reaction was completed. The structure and tacticity of the syndiotactic poly(octene) has been examined through of 1H and 13C NMR [28,31]. Figure 5 shows the 13C NMR spectra of the syndiotactic trans-1,2 poly(octadiene) and its saturated polymer together with the peak attribution.

4. Conclusions

Several syndiotactic-rich polyolefins were prepared by hydrogenation of highly stereoregular syndiotactic 1,2 poly(1,3-diene)s, and their microstructure was determined through NMR analysis. Such highly stereoregular polyolefins could be quite useful as polymer models for the microstructural characterization of analogous polymers, even with lower stereoregularity, which can be obtained by stereospecific polymerization of the corresponding monomers.
The polyolefins obtained could be of potential interest for applications in the elastomeric and/or thermoplastic fields, although it is difficult to imagine an industrial application, given the current high cost of the monomers used for the preparation of the polydienes, from which polyolefins are obtained through hydrogenation. However, in the near future, these niche polymers may be produced at more accessible costs opening the scenario on appealing industry applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/polym16192711/s1. Figure S1. FTIR spectra of syndiotactc trans-1,2 poly(5-methy-1,3-hexadiene) (top, (1,2syP5MHX)) and its hydrogenated product syndiotactic poly(5-methyl-1-hexene) (bottom, [H(1,2syP5MHX)]. Figure S2. FTIR spectra of syndiotactic trans-1,2 poly(1,3-octadiene) (1,2syPO) (top) and its hydrogenated product syndiotactic poly(1-octene) ([H(1,2syPO)], bottom). Figure S3: 13C NMR spectrum of [H(1,2syPHX)] sample @600 MHz in TCE, showing the assignment of tacticity at triad level. Figure S4: HSQC on the left) and HMBC (on the right) experiments of [H(1,2syPHX)] sample @600 MHz in TCE. Figure S5: HSQC on the left) and HMBC (on the right) experiments of [H(1,2syPHP)] sample @600 MHz in TCE.

Author Contributions

Conceptualization, G.R.; investigation G.R., I.P. and A.C.B., writing—original draft preparation, G.R., I.P. and A.C.B.; writing—review and editing, A.C.B. and G.R.; supervision, G.R. 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.

Data Availability Statement

The original contributions presented in the study are included in the article/supplementary material, further inquiries can be directed to the corresponding author.

Acknowledgments

What is reported in this paper is taken from the PhD thesis of Ivana Pierro, titled “Polyolefins by hydrogenation of stereoregular poly(1,3-diene)s and chain-walking polymerization: synthesis, structure and mechanical properties”.

Conflicts of Interest

The authors declare no conflicts 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, Part II, pp. 53–108. [Google Scholar]
  2. 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]
  3. Ricci, G.; Pampaloni, G.; Sommazzi, A.; Masi, F. Dienes polymerization: Where we are and what lies ahead. Macromolecules 2021, 54, 5879–5914. [Google Scholar] [CrossRef]
  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.; Italia, S.; Porri, L. Polymerization of (Z)-1,3-pentadiene with CpTiCl3/MAO. Effect of temperature on polymer structure and mechanistic implications. Macromolecules 1994, 27, 868–869. [Google Scholar] [CrossRef]
  6. Ricci, G.; Italia, S.; Giarrusso, A.; Porri, L. Polymerization of 1,3-dienes with the soluble catalyst system methylaluminoxanes-[CpTiCl3]. Influence of monomer structure on polymerization stereospecificity. J. Organomet. Chem. 1993, 451, 67–72. [Google Scholar] [CrossRef]
  7. Ricci, G.; Battistella, M.; Porri, L. Chemoselectivity and stereospecificity of Cr(II) catalysts for 1,3-diene polymerization. Macromolecules 2001, 34, 5766–5769. [Google Scholar] [CrossRef]
  8. Ricci, G.; Italia, S.; Porri, L. Polymerization of butadiene to 1,2-syndiotactic polymer with (η3-C8H13)(η4-C4H6)Co. Some observations on the factors that determine the stereospecificity. Polymer Commun. 1988, 29, 305–307. [Google Scholar]
  9. Ricci, G.; Morganti, D.; Sommazzi, A.; Santi, R.; Masi, F. Polymerization of 1,3-dienes with iron complexes based catalysts. Influence of the ligand on catalyst activity and stereospecificity. J. Mol. Cat. A Chem. 2003, 204/205, 287–293. [Google Scholar] [CrossRef]
  10. Ricci, G.; Leone, G.; Zanchin, G.; Palucci, B.; Boccia, A.C.; Sommazzi, A.; Masi, F.; Zacchini, S.; Guelfi, M.; Pampaloni, G. Highly Stereoregular 1,3-Butadiene and Isoprene Polymers through Monoalkyl−N−Aryl Substituted Iminopyridine Iron Complex-Based Catalysts: Synthesis and Characterization. Macromolecules 2021, 54, 9947–9959. [Google Scholar] [CrossRef]
  11. 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]
  12. 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]
  13. Ricci, G.; Boglia, A.; Motta, T.; Bertini, F.; Boccia, A.C.; Zetta, L.; Alberti, E.; Famulari, A.; Arosio, P.; Meille, S.V. Synthesis and Structural Characterization of Syndiotactic trans-1,2 and cis-1,2 Polyhexadienes. J. Polym. Sci Part A Polym. Chem. 2007, 45, 5339–5353. [Google Scholar] [CrossRef]
  14. 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]
  15. Bazzini, C.; Giarrusso, A.; Porri, L. Diethylbis(2,2-bipyridine)iron/MAO. A Very Active and Stereospecific Catalyst for 1,3-Diene Polymerization. Macromol. Rapid Commun. 2002, 23, 922–927. [Google Scholar] [CrossRef]
  16. Pirozzi, B.; Napolitano, R.; Petraccone, V.; Esposito, S. Determination of the Crystal Structure of Syndiotactic 3,4-Poly(2-methyl-1,3-butadiene) by Molecular Mechanics and X-Ray Diffraction. Macromol. Chem. Phys. 2004, 205, 1343–1350. [Google Scholar] [CrossRef]
  17. Ricci, G.; Leone, G.; Zanchin, G.; Masi, F.; Guelfi, M.; Pampaloni, G. Dichloro(2,2′-bipyridine)copper/MAO: An Active and Stereospecific Catalyst for 1,3-Diene Polymerization. Molecules 2023, 28, 374. [Google Scholar] [CrossRef]
  18. Ricci, G.; Bertini, F.; Boccia, A.C.; Zetta, L.; Alberti, E.; Pirozzi, B.; Giarrusso, A.; Porri, L. Synthesis and Characterization of Syndiotactic 1,2 Poly(3-Methyl-1,3-Pentadiene). Macromolecules 2007, 40, 7238–7243. [Google Scholar] [CrossRef]
  19. De Rosa, C.; Auriemma, F.; Santillo, C.; Di Girolamo, R.; Leone, G.; Ricci, G. Chirality, entropy and crystallization in polymers: Isotactic poly(3-methyl-1-pentene) as an example of influence of chirality and entropy on the ctystal structure. CrystEngComm 2015, 17, 6006–6013. [Google Scholar] [CrossRef]
  20. De Rosa, C.; Auriemma, F.; Santillo, C.; Di Girolamo, R.; Leone, G.; Boccia, A.C.; Ricci, G. Crystal structure of isotactic poly((R,S)-3-methyl-1-pentene). Macromolecules 2015, 48, 5251–5266. [Google Scholar] [CrossRef]
  21. De Rosa, C.; Malafronte, A.; Scoti, M.; Auriemma, F.; Pierro, I.; Leone, G.; Ricci, G. Synthesis and Structure of Syndiotactic Poly(3-Methyl- 1-Butene): A Case of 3/1 Helical Conformation for Syndiotactic Polymers. Macromolecules 2018, 51, 8574–8584. [Google Scholar] [CrossRef]
  22. Samran, J.; Phinyocheep, P.; Daniel, P.; Kittipoom, S. Hydrogenation of unsaturated rubbers using diimide as a reducing agent. J. App. Polym. Sci. 2005, 95, 16–27. [Google Scholar] [CrossRef]
  23. Mango, L.A.; Lenz, R.W. Hydrogenation of unsaturated polymers with diimide. Makromol. Chem. 1973, 163, 13. [Google Scholar] [CrossRef]
  24. Harwood, H.J.; Russel, D.B.; Verthe, J.J.A.; Zymonas, J. Diimide as a reagent for the hydrogenation of unsaturated polymers. Makromol. Chem. 1973, 163, 1–12. [Google Scholar] [CrossRef]
  25. Santin, C.K.; Jacobi, M.M.; Schuster, R.H. Influence of 1,2 units content on the hydrogenation of polydienes by TSH. J. Appl. Polym. Sci. 2011, 119, 1195–1203. [Google Scholar] [CrossRef]
  26. Singha, N.K.; Bhattacharjee, S.; Sivaram, S. Hydrogenation of Diene Elastomers, Their Properties and Applications: A Critical Review. Rubber Chem. Technol. 1997, 70, 309–367. [Google Scholar] [CrossRef]
  27. Sozzani, P.; Simonutti, R.; Galimberti, M. MAS NMR Characterization of Syndiotactic Polypropylene: Crystal Structure and Amorphous Phase Conformation. Macromolecules 1993, 26, 5782–5789. [Google Scholar] [CrossRef]
  28. Asakura, T.; Demura, M.; Nishiyama, Y. Carbon-13 NMR Spectral Assignment of Five Polyolefins Determined from the Chemical Shift Calculation and the Polymerization Mechanism. Macromolecules 1991, 24, 2334–2340. [Google Scholar] [CrossRef]
  29. Saito, J.; Suzuki, Y.; Makio, H.; Tanaka, H.; Onda, M.; Fujita, T. Polymerization of Higher R-Olefins with a Bis(Phenoxyimine)Ti Complex/iBu3Al/Ph3CB(C6F5)4: Formation of Stereo- and Regioirregular High Molecular Weight Polymers with High Efficiency. Macromolecules 2006, 39, 4023–4031. [Google Scholar] [CrossRef]
  30. Janas, Z.; Godbole, D.; Nerkowski, T.; Szezegot, K. Zirconium and hafnium complexes of the thio(bisphenolato) ligand: Synthesis, structural characterization and testing as 1-hexene polymerization catalyst. Dalton Trans. 2009, 8846–8863. [Google Scholar] [CrossRef]
  31. Lukesova, L.; Ward, B.D.; Bellemin-Laponnaz, S.; Wadepohl, H.; Gade, L.H. High tacticity control in organolanthanide polymerization catalysis: Formation of isotactic poly(α-alkenes) with a chiral C3-symmetric thulium complex. Dalton Trans. 2007, 920–922. [Google Scholar] [CrossRef]
  32. Delfini, M.; Di Cocco, M.E.; Paci, M.; Aglietto, M.; Carlini, C.; Crisci, L.; Ruggeri, G. 13C n.m.r. study of poly[(S)-5-metrhyl-1-heptene] and poly(1-heptene) prepared by Ziegler-Natta catalyst. Polymer 1985, 26, 1459–1462. [Google Scholar] [CrossRef]
Scheme 1. Polyolefins by hydrogenation of highly stereoregular 1,2 (3,4) polydienes.
Scheme 1. Polyolefins by hydrogenation of highly stereoregular 1,2 (3,4) polydienes.
Polymers 16 02711 sch001
Scheme 2. Polymerization of trans-1,3-pentadiene with CoCl2 (PiPrPh2)2/MAO and successive hydrogenation of the obtained syndiotactic trans-1,2 (polypentadiene) to give fully saturated syndiotactic poly(pentene).
Scheme 2. Polymerization of trans-1,3-pentadiene with CoCl2 (PiPrPh2)2/MAO and successive hydrogenation of the obtained syndiotactic trans-1,2 (polypentadiene) to give fully saturated syndiotactic poly(pentene).
Polymers 16 02711 sch002
Figure 1. 1H (left) and 13C NMR (right) spectra of syndiotactic trans 1,2 poly(1,3-pentadiene) (top, (1,2syPP)) and syndiotactic poly(1-pentene) (bottom, [H(1,2syPP)]).
Figure 1. 1H (left) and 13C NMR (right) spectra of syndiotactic trans 1,2 poly(1,3-pentadiene) (top, (1,2syPP)) and syndiotactic poly(1-pentene) (bottom, [H(1,2syPP)]).
Polymers 16 02711 g001
Scheme 3. Polymerization of trans-1,3-hexadiene with CoCl2 (PiPrPh2)2/MAO and successive hydrogenation of the obtained syndiotactic trans-1,2 (poly(1,3-hexadiene)) (1,2syPHX) to give fully saturated syndiotactic poly(1-hexene) [H(1,2syPHX)].
Scheme 3. Polymerization of trans-1,3-hexadiene with CoCl2 (PiPrPh2)2/MAO and successive hydrogenation of the obtained syndiotactic trans-1,2 (poly(1,3-hexadiene)) (1,2syPHX) to give fully saturated syndiotactic poly(1-hexene) [H(1,2syPHX)].
Polymers 16 02711 sch003
Figure 2. 1H (left) and 13C (right) NMR spectra of syndiotactic trans 1,2 poly(1,3-hexadiene) (top (1,2syPHX)) and the hydrogenated syndiotactic poly(1-hexene) (bottom, H(1,2syPHX)).
Figure 2. 1H (left) and 13C (right) NMR spectra of syndiotactic trans 1,2 poly(1,3-hexadiene) (top (1,2syPHX)) and the hydrogenated syndiotactic poly(1-hexene) (bottom, H(1,2syPHX)).
Polymers 16 02711 g002
Scheme 4. Polymerization of 5-methyl-1,3-hexadiene with CoCl2 (PiPrPh2)2/MAO to syndiotactic trans-1,2 poly(5-methyl-1,3-hexadiene) (1,2syP5MHX) and successive hydrogenation giving fully saturated syndiotactic poly(5-methyl-1-hexene) [H(1,2syP5MHX)].
Scheme 4. Polymerization of 5-methyl-1,3-hexadiene with CoCl2 (PiPrPh2)2/MAO to syndiotactic trans-1,2 poly(5-methyl-1,3-hexadiene) (1,2syP5MHX) and successive hydrogenation giving fully saturated syndiotactic poly(5-methyl-1-hexene) [H(1,2syP5MHX)].
Polymers 16 02711 sch004
Figure 3. 1H (left) and 13C (right) NMR spectra of syndiotactic trans-1,2 poly(5-methyl-1,3-hexadiene) (top (1,2syP5MHX)) and syndiotactic poly(5-methyl-1-hexene) (bottom, H(1,2syP5MHX)).
Figure 3. 1H (left) and 13C (right) NMR spectra of syndiotactic trans-1,2 poly(5-methyl-1,3-hexadiene) (top (1,2syP5MHX)) and syndiotactic poly(5-methyl-1-hexene) (bottom, H(1,2syP5MHX)).
Polymers 16 02711 g003
Scheme 5. Polymerization of trans-1,3-heptadiene with CoCl2 (PiPrPh2)2/MAO and successive hydrogenation of the obtained syndiotactic trans-1,2 poly(1,3-heptadiene) (1,2syPHP) to give fully saturated syndiotactic poly(1-heptene) [H(1,2syPHP)].
Scheme 5. Polymerization of trans-1,3-heptadiene with CoCl2 (PiPrPh2)2/MAO and successive hydrogenation of the obtained syndiotactic trans-1,2 poly(1,3-heptadiene) (1,2syPHP) to give fully saturated syndiotactic poly(1-heptene) [H(1,2syPHP)].
Polymers 16 02711 sch005
Figure 4. 1H (left) and 13C (right) NMR spectra of syndiotactic trans-1,2 poly(1,3-heptadiene) (top (1,2syPHP)) and syndiotactic poly(1-heptene) (bottom, H(1,2syPHP)).
Figure 4. 1H (left) and 13C (right) NMR spectra of syndiotactic trans-1,2 poly(1,3-heptadiene) (top (1,2syPHP)) and syndiotactic poly(1-heptene) (bottom, H(1,2syPHP)).
Polymers 16 02711 g004
Scheme 6. Polymerization of 1,3-octadiene with CoCl2 (PiPrPh2)2/MAO and successive hydrogenation of the obtained syndiotactic trans-1,2 poly(1,3-octadiene) (1,2syPO) to give fully saturated syndiotactic poly(octene) [H(1,2syPO)].
Scheme 6. Polymerization of 1,3-octadiene with CoCl2 (PiPrPh2)2/MAO and successive hydrogenation of the obtained syndiotactic trans-1,2 poly(1,3-octadiene) (1,2syPO) to give fully saturated syndiotactic poly(octene) [H(1,2syPO)].
Polymers 16 02711 sch006
Figure 5. 1H (left) and 13C (right) NMR spectra of syndiotactic trans-1,2 poly(1,3-octadiene) (top (1,2syPO)) and syndiotactic poly(1-octene) (bottom, [H(1,2syPO)]).
Figure 5. 1H (left) and 13C (right) NMR spectra of syndiotactic trans-1,2 poly(1,3-octadiene) (top (1,2syPO)) and syndiotactic poly(1-octene) (bottom, [H(1,2syPO)]).
Polymers 16 02711 g005
Table 1. Polymerization of 1,3-dienes to 1,2 polymers with CoCl2 (PiPrPh2)2/MAO.
Table 1. Polymerization of 1,3-dienes to 1,2 polymers with CoCl2 (PiPrPh2)2/MAO.
PolymerMonomerTime
(h)
Yield
(%)
1,2 a
(%)
[rrrr] b
(%)
m.p. c
(°C)
Mw d
(kg/mol)
Mw/Mn d
1,2syPPEP1688999601681672.5
1,2syPHXHX48499958109741.6
1,2syP5MHX5MHX236749950-694.3
1,2syPHPHP336739043-292.7
1,2syPOO144769244-302.3
Polymerization conditions: monomer, 2 mL; toluene as solvent, total volume 16 mL; MAO, 2 × 10−3 mol; CoCl2 (PiPrPh2)2, 20 μmol; polymerization temperature, −20 °C. a Percentage of 1,2 units, determined by NMR analysis. The remaining units are essentially cis-1,4; b percentage of syndiotactic pentads, determined by 13C NMR analysis; c melting temperature, determined by DSC; d determined by SEC. EP = trans-1,3-pentadiene; HX = 1,3-hexadiene; 5MHX = 5-methyl-1,3-hexadiene; HP = 1,3-heptadiene; O = 1,3-octadiene.
Table 2. Syndiotactic-rich polyolefins features.
Table 2. Syndiotactic-rich polyolefins features.
Polymer[rrrr]  aMw  bMw/Mn  bTg  c
(%)(kg/mol)(°C)
H(1,2syPP)_syndiotactic poly(1-pentene)621752.6−46
H(1,2syPHX)_syndiotactic poly(1-hexene)59851.4−50
H(1,2syP5MH)_syndiotactic poly(5-methyl-1-hexene)50772.9−55
H(1,2syPHP)_syndiotactic poly(1-heptene)42382.0−56
H(1,2syPO)_syndiotactic poly(1-octene)45352.1−60
a Percentage of syndiotactic pentads, evaluated by NMR spectroscopy; b determined by SEC; c determined by DSC.
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

Ricci, G.; Pierro, I.; Boccia, A.C. Syndiotactic Polyolefins by Hydrogenation of Highly Stereoregular 1,2 Polydienes: Synthesis and Structural Characterization. Polymers 2024, 16, 2711. https://doi.org/10.3390/polym16192711

AMA Style

Ricci G, Pierro I, Boccia AC. Syndiotactic Polyolefins by Hydrogenation of Highly Stereoregular 1,2 Polydienes: Synthesis and Structural Characterization. Polymers. 2024; 16(19):2711. https://doi.org/10.3390/polym16192711

Chicago/Turabian Style

Ricci, Giovanni, Ivana Pierro, and Antonella Caterina Boccia. 2024. "Syndiotactic Polyolefins by Hydrogenation of Highly Stereoregular 1,2 Polydienes: Synthesis and Structural Characterization" Polymers 16, no. 19: 2711. https://doi.org/10.3390/polym16192711

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop