Next Article in Journal
Computational Investigations on Soundproof Applications of Foam-Formed Cellulose Materials
Next Article in Special Issue
FeNxC Based Catalysts Prepared by the Calcination of Iron-Ethylenediamine@Polyaniline as the Cathode-Catalyst of Proton Exchange Membrane Fuel Cell
Previous Article in Journal
Anionic Polymerization of β-Butyrolactone Initiated with Sodium Phenoxides. The Effect of the Initiator Basicity/Nucleophilicity on the ROP Mechanism
Previous Article in Special Issue
Preparation of Butadiene-Isoprene Copolymer with High Vinyl Contents by Al(OPhCH3)(i-Bu)2/MoO2Cl2∙TNPP
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Co-Catalysts and Polymerization Conditions on Properties of Poly(anhydride-alt-epoxide)s from ROCOP Using Salen Complexes with Different Metals

Institute for Macromolecular Studies, National Research Council, Via A. Corti, 12, 20133 Milan, Italy
*
Author to whom correspondence should be addressed.
Polymers 2019, 11(7), 1222; https://doi.org/10.3390/polym11071222
Submission received: 28 June 2019 / Revised: 15 July 2019 / Accepted: 16 July 2019 / Published: 22 July 2019
(This article belongs to the Special Issue Polymerizations Promoted by Metal Complexes)

Abstract

:
Cyclohexene oxide (CHO) and phthalic anhydride (PA) have been reacted in the presence of commercial salen–type complexes with different metals Cr (1), Al (2), and Mn (3) in combination with 4-(dimethylamino) pyridine (DMAP), bis-(triphenylphosphorydine) ammonium chloride (PPNCl) and bis-(triphenylphosphoranylidene)ammonium azide (PPNN3) as co-catalysts to obtain alternating poly(PA-alt-CHO)s by ring-opening copolymerization (ROCOP). The effect of different reaction conditions (pre-contact between catalyst and co-catalyst, polymerization time) on the productivity, molecular weight and glass transition temperature has been evaluated. By using a 24 h pre-contact, the aliphatic polyesters obtained were characterized by high molecular weight (Mn > 15 kg/mol) and glass transition temperature (Tg) up to 146 °C; the more sustainable metals Al and Mn in the presence of PPNCl give comparable results to Cr. Moreover, biodegradability data of these polyesters and the study of the microstructure reveal that the biodegradability is influenced more by the type of chain linkages rather than by the molecular weight of the polyesters.

Graphical Abstract

1. Introduction

Aliphatic polyesters (PEs) are an appealing class of polymers used in a range of applications such as biomedical devices and bulk packaging owing to their excellent properties and general biocompatibility [1,2]. They are typically synthesized by ring-opening polymerization (ROP) of lactones and lactides [3,4,5], an excellent controlled polymerization route, which gives polymers with relatively low glass transition temperatures. Alternating ring-opening copolymerization (ROCOP) of epoxides and cyclic anhydrides (Scheme 1) is becoming an attractive method for the synthesis of PEs [6,7,8,9,10,11,12,13,14,15]. It represents an alternative chain-growth route to polyesters with respect to the ring-opening polymerization (ROP) of lactones and an opportunity to broaden the range of materials produced and to overcome some of the limitations of ROP. In particular, the properties of the ensuing materials, including the thermal properties (glass transition temperature (Tg) and thermal decomposition temperature), can be tuned by changing the epoxide or cyclic anhydride. For example, low Tg values can be obtained by monomers with long side chains or using monocyclic epoxides and anhydrides [8,16]. Conversely, high Tg values are favored by rigid backbones, showing the importance of selecting more rigid monomer combinations such as bi- or tricyclic monomers and monomers bearing an aromatic group [9,10,11,12,13,14,17,18].
However, one of the major challenges faced with ROCOP is that the PEs synthesized via this technique are typically of low number average molecular weight (Mn), that renders poor mechanical and thermal properties and thus limits their applications.
In recent years, numerous organometallic catalysts have been developed for epoxide/anhydride copolymerization, including magnesium [11], aluminum [7,12,13,19,20,21,22], chromium [8,12,15,21,22,23,24,25,26], manganese [12,22,23,27,28], iron [7,29], cobalt [12,15,20,21,22,23], zinc complexes [8,9,11] and heterodinuclear polymerization catalyst [30,31,32,33], many of which showed markedly higher activity with the addition of a nucleophilic co-catalyst [18,29]. Duchateau and coworkers [34] investigated the ROCOP of cyclohexene oxide (CHO) with succinic anhydride (SA), phthalic anhydride (PA), and cyclopropane-1,2-dicarboxylic acid anhydride in bulk and in solution by using different salen or salphen metal complexes such as those of Al, Cr and Co, along with several co-catalysts. The most reactive substrate was PA, which yielded a polyester with the highest Mn value by using a combination of salphen Cr(III) complex catalyst and bis(triphenylphosphine)iminium chloride (PPNCl) as co-catalyst but no information on thermal properties has been reported. Recently, metal-free initiators for the copolymerization of epoxides with anhydrides are being explored [35,36]. A series of dinuclear complexes, in which two iron(III) amino triphenolate moieties are bridged by a phenylene backbone were synthesized by Jiang [37] for the alternating copolymerization of CHO/PA in the presence of PPNCl with good Mn value (33 kg/mol). However, research efforts in the synthesis of polyesters by ROCOP with high Tg and suitable molecular weight by using commercial catalysts are still necessary.
In this study, an investigation on the influence of co-catalysts and polymerization conditions on alternating ROCOP of CHO and PA using commercial salen complexes with different metals is reported. The polyesters were characterized by size exclusion chromatography (SEC), differential scanning calorimetry (DSC) and nuclear magnetic resonance (NMR). This research work offers access to PEs with good thermal and molecular weight properties suitable for practical applications. Moreover, the biodegradability data of selected poly(anhydride-alt-epoxide)s are reported.

2. Materials and Methods

2.1. Materials

Phthalic anhydride (PA), cyclohexene oxide (CHO), 4-(dimethylamino) pyridine (DMAP), bis-(triphenylphosphorydine) ammonium chloride (PPNCl), (R-R)-N,N′-bis (3,5-di-tert-butylsalicylidene)-1,2-cyclohexanediaminochromium(III) chloride, (R-R)-N,N′-bis (3,5-di-tert-butylsalicylidene)-1,2-cyclohexanediaminomanganese(III) chloride were purchased from Sigma-Aldrich, Milan, Italy. (R-R)-N,N′-bis (3,5-di-tert-butylsalicylidene)-1,2-cyclohexanediaminoaluminum chloride was purchased from Strem Chemicals (Newburyport, MA, USA). Cyclohexene oxide and dichloromethane (CH2Cl2) were dried over CaH2, distilled and stored on 4 Å molecular sieves under nitrogen. Phthalic anhydride was recrystallized from dichloromethane prior to use. Bis-(triphenylphosphorydine) ammonium chloride was dissolved in dichloromethane and precipitated in diethyl ether ((C2H5)2O) twice. 4-(dimethylamino) pyridine was double recrystallized from toluene. The co-catalyst bis-(triphenylphosphorydine) ammonium azide has been synthesized according to literature procedures [24]. All manipulations were performed under an inert atmosphere or in a nitrogen-filled MBraun (M. BRAUN INERTGAS-SYSTEME GMBH (Garching, Germany)) glovebox unless stated otherwise.

2.2. Synthesis

Bulk Polymerization.
In a glove box, a 10 mL crimp cap vial equipped with a stirring bar was charged with a mixture of catalyst, co-catalyst, epoxide and anhydride with a ratio of 1:1:250:250.
Polymerization in Solvent (No Pre-contact Step).
In a glove box, a 10 mL crimp cap vial equipped with a stirring bar was charged with a mixture of catalyst, co-catalyst, epoxide, anhydride and 1 mL of toluene with a ratio of 1:1:250:250.
Polymerization in Solvent with Pre-contact step.
In a glove box, in a 10 mL crimp cap vial equipped with a stirring bar a mixture of catalyst and co-catalyst was charged in the presence of 1 mL of toluene and keep stirring for 1 h or 24 h (pre-contact step). Then, the epoxide and the anhydride were added. The ratio between catalyst: co-catalyst: epoxide: anhydride was 1:1:250:250.
Then, the vial was placed in an aluminum heating block mounted on top of a stirrer/heating plate. At the end of the polymerization the crude product was precipitated twice in methanol and collected after filtration through a 0.45 µm Nylon filter. All the analyses were performed on purified sample. Yield (%) was calculated as yield   g gCHO + gPA 100 .

2.3. Methods

The copolymers were weighed in a 5 mm NMR tube and dissolved in CHCl3. The spectra were recorded on a Bruker Avance 400 instrument (400 MHz (1H); 100.58 MHz (13C); pulse angle = 12.50 ms; acquisition time = 0.94 s; delay = 16 s). The probe head was pre-equilibrated at a fixed temperature of 35 °C.
Differential scanning calorimetry (DSC) analysis was performed on a Perkin Elmer DSC 8000 instrument using cyclic heating and cooling rates of 20 °C per minute and heated from 20 to 200 °C. The values of glass transition temperature Tg were recorded during the second thermal cycle.
Molar mass analysis was performed using about 12 mg of polymer in THF stabilized with 0.025% BHT (butylated hydroxytoluene) at 35 °C by a size exclusion chromatography (SEC) system from Waters W600 (Millford, MA), equipped with a differential refractometer Waters 410. The column set was Agilent 3 PL GEL (Polypore, Oligopore, 50 Å).
Biodegradability have been determined by respirometric biochemical oxygen demand (BOD) Oxitop method based on very accurate automatic pressure measurement in a closed bottle. When organic matter biodegrades, it demands a certain amount of oxygen. When oxygen is consumed, pressure falls and at the same time carbon dioxide is produced. The system consists of an OxiTop-C measuring head, an OxiTop Controller OC 100, capable of handling up to 120 warheads, and an inductive stirring system. Screwing in the OxiTop-C measuring head, like a “cover”, on the special dark glass bottle, it detects the pressure in the head space, using a small transducer connected to a microprocessor. The bottle was placed on a magnetic stirrer suitable for being introduced into an incubator at the set temperature. The sample volume was chosen based on the presumed BOD (biochemical oxygen demand) value, considering that a too large measuring range will lead to inaccurate results. The OxiTop® respirometric system has a special rubber housing inside where NaOH tablets that react with CO2 are placed. The removal of CO2 from the gas phase led to a decrease in the pressure of the system that was recorded. By means of suitable calculations, the OxiTop-C measuring heads converted the measurement of the pressure variation directly into mg/L of consumed O2.

3. Results

To understand the effect of the polymerization conditions to obtain industrially processable alternating poly(anhydride-alt-epoxide)s with Mn > 11,000 g/mol and Tg > 80 °C, a series of copolymerizations of CHO with phthalic anhydride wa performed by the salen-type complexes of three different metals Cr (1), Al (2), and Mn (3) (Scheme 2) in combination with 2-dimethylaminopyridine (DMAP), and two onium salts with a bulky cation and two different anions (PPNCl and PPNN3) as co-catalysts (Scheme 3). Commercially available salen complexes were selected, the chromium complex 1 is one of those most studied in ROCOP and in particular was used by Duchateau for the ROCOP of CHO and anhydrides [34]. The aluminum and manganese catalysts 2 and 3, respectively, were selected as possible more sustainable alternatives. The DMAP and PPNCl co-catalysts investigated are also commercially available. In particular, the neutral Lewis base DMAP is one of the most studied and is taken as reference. PPNX are well-known onium salts: PPNCl is commercially available and in combination with salphen Cr complex resulted in the most effective in CHO and PA copolymerization; PPNN3 is reported as one of the most effective co-catalysts in CO2 and epoxide copolymerization [38].

ROCOP of CHO Using Complexes 13 in Presence of Different Co-Catalysts

Initially, the copolymerizations of cyclohexene oxide (CHO) with phthalic anhydride (PA) by (salen)MtCl catalysts 1, 2 and 3 and DMAP and PPNCl as co-catalysts were performed at 110 °C in bulk without pre-contact step between catalyst and co-catalyst, in polymerization conditions reported in the literature [34]. The results are shown in Table 1.
In our hands, the viscosity of the system increased rapidly especially with catalyst 1 and therefore the reactions were stopped after a short time. Yields were high, catalyst 1 resulted to be the most active one, and in general, PPNCl resulted to be the best co-catalyst. The resulting polyesters were characterized by 1H NMR, SEC and DSC. A relatively high amount of ether linkages (see 1H NMR results below) was observed under these conditions. The measured numbers average molecular weight Mn ranged from 1700 to 2500 g/mol. Interestingly, even though molar mases were low probably due to the short polymerization time and high viscosity, the Tg values were higher than 135 °C, and for copolymers by catalyst 1 Tg values of 141 °C were recorded.
Successively, with the purpose to increase the molecular weights, a series of copolymerization of CHO and PA were performed in solution by adding 1 mL of toluene in order to reduce the viscosity of the reaction medium. The copolymerizations were carried out in the presence of catalyst 1 and DMAP and as expected, the presence of a small amount of solvent decreases the viscosity of the reaction medium, and the molecular weight of CHO/PA copolymers increases up to 9500 g/mol.
Then, to make homogeneous comparisons in all copolymerizations a pre-contact step between catalysts and co-catalysts was introduced since PPNX salts are insoluble or very sparingly soluble in epoxides at room temperature. Darensbourg [39] reported that for the copolymerization of CHO and CO2 a pre-contact step seems to be a key factor in the copolymerization reaction. The catalysts and the different co-catalysts were dissolved in toluene and stirred for 1 h at room temperature (pre-contact step), upon removing the solvent under vacuum, cyclohexene oxide was added with stirring. In Table 2, Table 3 and Table 4 the results of the copolymerizations performed with catalysts 1, 2 and 3, respectively, with different co-catalysts at different reaction times, are listed.
The comparison of entries in Table 2 shows in a glance that the pre-contact step was crucial: high values of Mn for the copolymers prepared in solution and in the presence of the pre-contact step as well as increase in the Tg values were obtained. In general, good yields were obtained with the three different co-catalysts although the catalytic system 1/PPNCl gave higher productivity, especially at short polymerization time. A prolonged reaction time (24 h) does not seem to further affect the performance of catalyst 1 with the three co-catalysts.
Regarding the aluminum-based catalyst 2 (Table 3) and the manganese-based catalyst 3 (Table 4), they seemed less promising than the chromium one in terms of productivity at short polymerization time. Conversely, a prolonged polymerization time positively affected the productivity as well the molecular weight and Tgs. Molar masses, yields and Tg compared well with those obtained with the chromium based complex 1. PPNCl proved to be the best co-catalyst for the three catalytic precursors.
In addition, we investigated the effect of the pre-contact time on the CHO/PA copolymerizations. A series of additional experiments using a pre-contact time of 24 h and a polymerization time of 1 h was also conducted. The obtained results are compared in Table 5.
For catalyst 1 there was no great performance difference by varying the pre-contact time from 1 to 24 h for the same reaction time (1 h), or using a long pre-contact time (24 h) or a long polymerization time (24 h, see entries 63 vs. 37) for the three different co-catalysts. By using the catalysts 2 and 3, appreciable differences in the performances have been obtained by varying the pre-contact time from 1 to 24 h for the same reaction time (1 h), while small differences are found by exchanging 1 to 24 h of pre-contact with 24 and 1 h of polymerization time differences. It is worth noting that with 24 h of pre-contact and 1 h of polymerization catalysts 2 and 3, in terms of activity, molecular weight and thermal properties, compared very well with performances of catalyst 1. Since this effect was evident also with DMAP as co-catalyst, this indicated that the pre-contact time was important not only because of the solubility of the two phosphonium salts, but also because it facilitated the formation of the active species. This result is important since at higher polymerization time side reactions can occur [19,40,41]. Thus, copolymer microstructure has been investigated by 1H NMR.
Indeed, the samples prepared at different pre-contact and polymerization times showed some differences in the microstructure. The 1H NMR spectra of the polyesters synthesized with a pre-contact of 1 h and 24 h of polymerization always presented a broad peak between 3.6 and 3.2 ppm corresponding to CHO–CHO ether linkage, as displayed in Figure 1. It seems that the homopolymerization of epoxides could occur as a side reaction when the polymerization is complete [19,40,41].
This general behavior is well visible in Figure 2, where the CHO ether linkage percentage for the different catalytic systems with a pre-contact of 1 h and different polymerization time is reported. From this point of view, at short polymerization time PPNCl gave the best results with Al and Cr catalysts.
Since biodegradability is one of the interesting properties of these polymers, biodegradability tests were performed according to ISO standards 14,851 (Determination of the ultimate aerobic biodegradability of plastic materials in an aqueous medium) on three polyesters CHOPA 19, CHOPA 60 and CHOPA 66 obtained with catalyst 1 at different polymerization conditions by using the respirometric BOD (biochemical oxygen demand) Oxitop method. CHOPA 19 and CHOPA 60 were obtained in polymerization solution at different pre-contact (1 vs. 24 h) and polymerization times (3 vs. 1 h), showing very similar molecular weight Mn (17.5 kg/mol for CHOPA 19 and 15.7 kg/mol for CHOPA 60) and ether linkage <7%. CHOPA 66 was obtained in bulk and showed a very low molecular weight (Mn = 1.6 kg/mol) with an ether linkage >27%. Figure 3 shows the biodegradation curves (average values) as a function of time for the three CHOPA samples and for the microcrystalline cellulose used as reference material. In the curves we can distinguish a lag phase, the biodegradation phase and the plateau. For all three samples the biodegradation phase started after 30 days, that of cellulose after 8 days. The curves of CHOPA 19 and CHOPA 60 were almost overlapping, with a plateau phase close to 32%, well above that of CHOPA 66 which did not reach 12% (Figure 3).
From these data one can deduce that biodegradability was influenced more by the type of chain linkage rather than by the molecular weight of the polyesters. Therefore, the methodology used for polymerization, which influenced polymer microstructure was key in obtaining biodegradable poly(anhydride-alt-epoxide)s.

4. Conclusions

In conclusion, we examined the influence of different polymerization conditions on the copolymerization of cyclohexene oxide with phthalic anhydride by commercial salen complexes 13 in combination with DMAP, PPNCl and PPNN3. The pre-contact step between catalyst and co-catalyst is a key factor in obtaining polymers with molecular weight > 15 kg/mol and glass transition temperature up to 140 °C in high yields. In general, prolonged polymerization times lead to the formation of large quantities of CHO–CHO ether linkage, which lowers the poly(CHO-alt-PA)s biodegradability.
Finally, since with a long pre-contact time, but at short polymerization time, the catalysts 2 and 3 with more sustainable Al and Mn metals in the presence of PPNCl give comparable results to those from the benchmark Cr catalyst, 2 and 3 may be a more sustainable alternative.

Author Contributions

Conceptualization, L.B.; methodology, L.B.; polymer synthesis, M.P., N.G.G., microstructural and chain end analysis, S.L., writing and review, L.B. and I.T.

Funding

This research was funded by CARIPLO FOUNDATION, grant number 2016-0643.

Acknowledgments

The authors thank Paola Rizzarelli (IPCB-CNR) and Marco Rapisarda (IPCB-CNR) for the biodegradability test.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rabnawaz, M.; Wyman, I.; Auras, R.; Cheng, S. A roadmap towards green packaging: The current status and future outlook for polyesters in the packaging industry. Green Chem. 2017, 19, 4737–4753. [Google Scholar] [CrossRef]
  2. Hillmyer, M.A.; Tolman, W.B. Aliphatic Polyester Block Polymers: Renewable, Degradable, and Sustainable. Accounts Chem. Res. 2014, 47, 2390–2396. [Google Scholar] [CrossRef] [PubMed]
  3. Ajellal, N.; Carpentier, J.F.; Guillaume, C.; Guillaume, S.M.; Helou, M.; Poirier, V.; Sarazin, Y.; Trifonov, A. Metal-catalyzed immortal ring-opening polymerization of lactones, lactides and cyclic carbonates. Dalton Trans. 2010, 39, 8363–8376. [Google Scholar] [CrossRef] [PubMed]
  4. Mecerreyes, D.; Jerome, R.; Dubois, P. Novel Macromolecular Architectures Based on Aliphatic Polyesters: Relevance of the “Coordination-Insertion” Ring-Opening Polymerization. In Macromolecular Architectures. Advances in Polymer Science; Springer: Berlin/Heidelberg, Germany, 1999; Volume 147, pp. 1–59. [Google Scholar]
  5. Dijkstra, P.J.; Du, H.; Feijen, J. Single site catalysts for stereoselective ring-opening polymerization of lactides. Polym. Chem. 2011, 2, 520–527. [Google Scholar] [CrossRef]
  6. Longo, J.M.; Sanford, M.J.; Coates, G.W. Ring-Opening Copolymerization of Epoxides and Cyclic Anhydrides with Discrete Metal Complexes: Structure–Property Relationships. Chem. Rev. 2016, 116, 15167–15197. [Google Scholar] [CrossRef] [PubMed]
  7. Sanford, M.J.; Carrodeguas, L.P.; Van Zee, N.J.; Kleij, A.W.; Coates, G.W. Alternating Copolymerization of Propylene Oxide and Cyclohexene Oxide with Tricyclic Anhydrides: Access to Partially Renewable Aliphatic Polyesters with High Glass Transition Temperatures. Macromolecules 2016, 49, 6394–6400. [Google Scholar] [CrossRef]
  8. DiCiccio, A.M.; Coates, G.W. Ring-Opening Copolymerization of Maleic Anhydride with Epoxides: A Chain-Growth Approach to Unsaturated Polyesters. J. Am. Chem. Soc. 2011, 133, 10724–10727. [Google Scholar] [CrossRef]
  9. Jeske, R.C.; DiCiccio, A.M.; Coates, G.W. Alternating Copolymerization of Epoxides and Cyclic Anhydrides: An Improved Route to Aliphatic Polyesters. J. Am. Chem. Soc. 2007, 129, 11330–11331. [Google Scholar] [CrossRef]
  10. Harrold, N.D.; Li, Y.; Chisholm, M.H. Studies of Ring-Opening Reactions of Styrene Oxide by Chromium Tetraphenylporphyrin Initiators. Mechanistic and Stereochemical Considerations. Macromolecules 2013, 46, 692–698. [Google Scholar] [CrossRef]
  11. Saini, P.K.; Romain, C.; Zhu, Y.; Williams, C.K. Di-magnesium and zinc catalysts for the copolymerization of phthalic anhydride and cyclohexene oxide. Polym. Chem. 2014, 5, 6068–6075. [Google Scholar] [CrossRef] [Green Version]
  12. Paoniasari, A.; Duchateau, R.; Nejad, E.H.; Koning, C.E. Semi-aromatic polyesters by alternating ring-opening copolymerisation of styrene oxide and anhydrides. Polym. Chem. 2012, 3, 1308–1313. [Google Scholar]
  13. Aida, T.; Inoue, S. Catalytic Reaction on Both Sides of a Metalloporphyrin Plane Alternating Copolymerization of Phthalic Anhydride and Epoxypropane with an Aluminum Porphyrin-Quaternary Salt System. J. Am. Chem. Soc. 1985, 107, 1358–1364. [Google Scholar] [CrossRef]
  14. Si, G.; Li, B.; Dong, J.; Zhang, L.; Han, B.; Duan, Z.; Liu, B. Novel chromium complexes with a [OSSO]-type bis(phenolato) dianionic ligand mediate the alternating ring-opening copolymerization of epoxides and phthalic anhydride. Polym. Chem. 2015, 6, 6372–6377. [Google Scholar] [CrossRef]
  15. Winkler, M.; Romain, C.; Meier, M.A.R.; Williams, C.K. Renewable polycarbonates and polyesters from 1,4-cyclohexadiene. Green Chem. 2015, 17, 300–306. [Google Scholar] [CrossRef]
  16. Biermann, U.; Sehlinger, A.; Meier, M.A.R.; Metzger, J.O. Catalytic Copolymerization of Methyl 9,10-Epoxystearate and Cyclic Anhydrides under Neat Conditions. Eur. J. Lipid Sci. Technol. 2016, 118, 104–110. [Google Scholar] [CrossRef]
  17. Carrodeguas, L.P.; Martín, C.; Kleij, A.W. Semiaromatic Polyesters Derived from Renewable Terpene Oxides with High Glass Transitions. Macromolecules 2017, 50, 5337–5345. [Google Scholar] [CrossRef]
  18. Han, B.; Zhang, L.; Liu, B.; Dong, X.; Kim, I.; Duan, Z.; Theato, P. Controllable Synthesis of Stereoregular Polyesters by Organocatalytic Alternating Copolymerizations of Cyclohexene Oxide and Norbornene Anhydrides. Macromolecules 2015, 48, 3431–3437. [Google Scholar] [CrossRef]
  19. Van Zee, N.J.; Sanford, M.J.; Coates, G.W. Electronic Effects of Aluminum Complexes in the Copolymerization of Propylene Oxide with Tricyclic Anhydrides: Access to Well-Defined, Functionalizable Aliphatic Polyesters. J. Am. Chem. Soc. 2016, 138, 2755–2761. [Google Scholar] [CrossRef]
  20. Ren, W.-M.; Lu, X.-B.; Li, J.; Liu, Y. Asymmetric Alternating Copolymerization of Meso-epoxides and Cyclic Anhydrides: Efficient Access to Enantiopure Polyesters. J. Am. Chem. Soc. 2016, 138, 11493–11496. [Google Scholar]
  21. Bernard, A.; Chatterjee, C.; Chisholm, M.H. The influence of the metal (Al, Cr and Co) and the substituents of the porphyrin in controlling the reactions involved in the copolymerization of propylene oxide and cyclic anhydrides by porphyrin metal(III) complexes. Polymer 2013, 54, 2639–2646. [Google Scholar] [CrossRef]
  22. Liu, Y.; Guo, J.-Z.; Lu, H.W.; Wang, H.B.; Lu, X.B. Making Various Degradable Polymers from Epoxides Using a Versatile Dinuclear Chromium Catalyst. Macromolecules 2018, 51, 771–778. [Google Scholar] [CrossRef]
  23. Robert, C.; De Montigny, F.; Thomas, C.M. Tandem synthesis of alternating polyesters from renewable resources. Nat. Commun. 2011, 2, 586. [Google Scholar] [CrossRef] [PubMed]
  24. Poland, R.R.; Escobedo, C.; Darensbourg, D.J. Kinetic Studies of the Alternating Copolymerization of Cyclic Acid Anhydrides and Epoxides, and the Terpolymerization of Cyclic Acid Anhydrides, Epoxides, and CO2 Catalyzed by (salen)Cr III Cl. Macromolecules 2012, 45, 2242–2248. [Google Scholar]
  25. Bao, Y.Y.; Liu, Y.; Ren, W.M.; Lu, X.B.; Liu, J. Binuclear chromium–salan complex catalyzed alternating copolymerization of epoxides and cyclic anhydrides. Polym. Chem. 2013, 4, 1439–1444. [Google Scholar]
  26. Huijser, S.; Hosseininejad, E.; Sablong, R.; De Jong, C.; Koning, C.E.; Duchateau, R. Ring-Opening Co- and Terpolymerization of an Alicyclic Oxirane with Carboxylic Acid Anhydrides and CO2in the Presence of Chromium Porphyrinato and Salen Catalysts. Macromolecules 2011, 44, 1132–1139. [Google Scholar] [CrossRef]
  27. Liu, D.F.; Zhu, L.Q.; Wu, J.; Wu, L.Y.; Lü, X.Q. Ring-opening copolymerization of epoxides and anhydrides using manganese( III ) asymmetrical Schiff base complexes as catalysts. RSC Adv. 2015, 5, 3854–3859. [Google Scholar] [CrossRef]
  28. Robert, C.; Ohkawara, T.; Nozaki, K. Manganese-Corrole Complexes as Versatile Catalysts for the Ring-Opening Homo- and Co-Polymerization of Epoxide. Chem. Eur. J. 2014, 20, 4789–4795. [Google Scholar] [CrossRef] [PubMed]
  29. Mundil, R.; Hošt’álek, Z.; Šeděnková, I.; Merna, J. Alternating ring-opening copolymerization of cyclohexene oxide with phthalic anhydride catalyzed by iron(III) salen complexes. Macromol. Res. 2015, 23, 161–166. [Google Scholar] [CrossRef]
  30. Garden, J.A.; Saini, P.K.; Williams, C.K. Greater than the Sum of Its Parts: A Heterodinuclear Polymerization Catalyst. J. Am. Chem. Soc. 2015, 137, 15078–15081. [Google Scholar] [CrossRef] [Green Version]
  31. Zhu, Y.; Romain, C.; Williams, C.K. Selective Polymerization Catalysis: Controlling the Metal Chain End Group to Prepare Block Copolyesters. J. Am. Chem. Soc. 2015, 137, 12179–12182. [Google Scholar] [CrossRef] [Green Version]
  32. Thevenon, A.; Garden, J.A.; White, A.J.P.; Williams, C.K. Dinuclear Zinc Salen Catalysts for the Ring Opening Copolymerization of Epoxides and Carbon Dioxide or Anhydrides. Inorg. Chem. 2015, 54, 11906–11915. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Yu, C.Y.; Chuang, H.J.; Ko, B.T. Bimetallic bis(benzotriazole iminophenolate) cobalt, nickel and zinc complexes as versatile catalysts for coupling of carbon dioxide with epoxides and copolymerization of phthalic anhydride with cyclohexene oxide. Catal. Sci. Technol. 2016, 6, 1779–1791. [Google Scholar] [CrossRef]
  34. Nejad, E.H.; Van Melis, C.G.W.; Vermeer, T.J.; Koning, C.E.; Duchateau, R. Alternating Ring-Opening Polymerization of Cyclohexene Oxide and Anhydrides: Effect of Catalyst, Cocatalyst, and Anhydride Structure. Macromolecules 2012, 45, 1770–1776. [Google Scholar] [CrossRef]
  35. Hošťálek, Z.; Trhlíková, O.; Walterová, Z.; Martinez, T.; Peruch, F.; Cramail, H.; Merna, J. Alternating copolymerization of epoxides with anhydrides initiated by organic bases. Eur. Polym. J. 2017, 88, 433–447. [Google Scholar] [CrossRef]
  36. Kummari, A.; Pappuru, S.; Chakraborty, D.; Kanji, A. Fully alternating and regioselective ring-opening copolymerization of phthalic anhydride with epoxides using highly active metal-free Lewis pairs as a catalyst. Polym. Chem. 2018, 9, 4052–4062. [Google Scholar] [CrossRef]
  37. Shi, Z.; Jiang, Q.; Song, Z.; Wang, Z.; Gao, C. Dinuclear iron(iii) complexes bearing phenylene-bridged bis(amino triphenolate) ligands as catalysts for the copolymerization of cyclohexene oxide with carbon dioxide or phthalic anhydride. Polym. Chem. 2018, 9, 4733–4743. [Google Scholar] [CrossRef]
  38. Darensbourg, D.J. Making Plastics from Carbon Dioxide: Salen Metal Complexes as Catalysts for the Production of Polycarbonates from Epoxides and CO2. Chem. Rev. 2007, 107, 2388–2410. [Google Scholar] [CrossRef]
  39. Darensbourg, D.J.; Ulusoy, M.; Karroonnirum, O.; Poland, R.R.; Reibenspies, J.H.; Çetinkaya, B. Highly Selective and Reactive (salan)CrCl Catalyst for the Copolymerization and Block Copolymerization of Epoxides with Carbon Dioxide. Macromolecules 2009, 42, 6992–6998. [Google Scholar] [CrossRef]
  40. Fieser, M.E.; Sanford, M.J.; Mitchell, L.A.; Dunbar, C.R.; Mandal, M.; Van Zee, N.J.; Urness, D.M.; Cramer, C.J.; Coates, G.W.; Tolman, W.B. Mechanistic Insights into the Alternating Copolymerization of Epoxides and Cyclic Anhydrides Using a (Salph)AICI and Iminium Salt Catalytic System. J. Am. Chem. Soc. 2017, 139, 15222–15231. [Google Scholar] [CrossRef] [PubMed]
  41. Van Zee, N.J.; Coates, G.W. Alternating Copolymerization of Propylene Oxide with Biorenewable Terpene-Based Cyclic Anhydrides: A Sustainable Route to Aliphatic Polyesters with High Glass Transition Temperatures. Angew. Chem. Int. Ed. 2015, 54, 2665–2668. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Ring opening copolymerization (ROCOP).
Scheme 1. Ring opening copolymerization (ROCOP).
Polymers 11 01222 sch001
Scheme 2. General structure of (salen)MtCl catalysts utilized for the copolymerization reactions.
Scheme 2. General structure of (salen)MtCl catalysts utilized for the copolymerization reactions.
Polymers 11 01222 sch002
Scheme 3. Structures of co-catalysts utilized for the copolymerization reactions.
Scheme 3. Structures of co-catalysts utilized for the copolymerization reactions.
Polymers 11 01222 sch003
Figure 1. 1H NMR spectra of poly(CHO-alt-PA) obtained with catalyst 1 and PPNCl as co-catalyst with a pre-contact of 1 h and polymerization time of: (a) 24 h (entry CHOPA 18) and (b) 1 h (entry CHOPA 51).
Figure 1. 1H NMR spectra of poly(CHO-alt-PA) obtained with catalyst 1 and PPNCl as co-catalyst with a pre-contact of 1 h and polymerization time of: (a) 24 h (entry CHOPA 18) and (b) 1 h (entry CHOPA 51).
Polymers 11 01222 g001
Figure 2. CHO–CHO ether linkage percentage for the different catalytic systems with a polymerization time/pre-contact time of 1 h/24 h (a) and of 24 h/1 h (b).
Figure 2. CHO–CHO ether linkage percentage for the different catalytic systems with a polymerization time/pre-contact time of 1 h/24 h (a) and of 24 h/1 h (b).
Polymers 11 01222 g002
Figure 3. Biodegradation curves for CHO/PA polyesters.
Figure 3. Biodegradation curves for CHO/PA polyesters.
Polymers 11 01222 g003
Table 1. Cyclohexene oxide (CHO)/phthalic anhydride (PA) bulk copolymerization a.
Table 1. Cyclohexene oxide (CHO)/phthalic anhydride (PA) bulk copolymerization a.
EntryCatalyst/Co-CatalystTime (h)Yield (%)Conversion (%) bEther Linkages (mol%) bMn (kg/mol)ĐTg (°C)
CHOPA
CHOPA 661/DMAP0.3749263271.73.2141
CHOPA 691/PPNCl0.3828778161.83.0141
CHOPA 682/DMAP1.1717866182.32.5136
CHOPA 712/PPNCl1.1576453201.92.9137
CHOPA 673/DMAP1.0768471152.42.5139
CHOPA 703/PPNCl1.0818678162.52.0139
a Polymerization conditions: temperature = 110 °C, CHO:PA:cat:cocat = 250:250:1:1, catalyst (20 µmol) and co-catalyst (20 µmol), oxirane (5 mmol) and anhydride (5 mmol). b Calculated from 1H NMR spectra of purified samples.
Table 2. CHO/PA copolymerization catalyzed by complex 1 and different co-catalysts a.
Table 2. CHO/PA copolymerization catalyzed by complex 1 and different co-catalysts a.
EntryCo-CatalystTime (h)Yield (%)Mn (kg/mol)ĐTg (°C)
CHOPA 48DMAP17615.51.2147
CHOPA 19DMAP39117.51.2146
CHOPA 20DMAP248215.41.3146
CHOPA 51PPNCl19015.71.1146
CHOPA 17PPNCl39016.11.2145
CHOPA 18PPNCl249216.81.3145
CHOPA 54PPNN318414.81.1146
CHOPA 36PPNN338814.21.1147
CHOPA 37PPNN3247915.21.1146
a Polymerization in solution and with 1 h pre-contact step: solvent = toluene = 1 mL, temperature = 110 °C, CHO:PA:cat:cocat = 250:250:1:1.
Table 3. CHO/PA copolymerization catalyzed by complex 2 and different co-catalysts a.
Table 3. CHO/PA copolymerization catalyzed by complex 2 and different co-catalysts a.
EntryCo-CatalystTime (h)Yield (%)Mn (kg/mol) aĐTg (°C)
CHOPA 50DMAP1164.51.36129
CHOPA 23DMAP36610.41.47136
CHOPA 24DMAP248514.11.24143
CHOPA 53PPNCl1409.41.10133
CHOPA 21PPNCl38416.01.16144
CHOPA 22PPNCl249515.81.25145
CHOPA 56PPNN31418.81.10140
CHOPA 38PPNN337612.81.17144
CHOPA 39PPNN324789.31.24144
a Polymerization in solution and with 1 h pre-contact step: solvent = toluene = 1 mL, temperature = 110 °C, CHO:PA:cat:cocat = 250:250:1:1.
Table 4. CHO/PA copolymerization catalyzed by complex 3 and different co-catalysts a.
Table 4. CHO/PA copolymerization catalyzed by complex 3 and different co-catalysts a.
EntryCo-CatalystTime (h)Yield (%)Mn (kg/mol)ĐTg (°C)
CHOPA 49DMAP1144.51.31144
CHOPA 27DMAP37412.01.17/
CHOPA 28DMAP249614.01.27/
CHOPA 52PPNCl1357.81.21137
CHOPA 25PPNCl38614.91.15144
CHOPA 26PPNCl249215.11.26144
CHOPA 55PPNN31397.51.19136
CHOPA 40PPNN336812.61.20138
CHOPA 41PPNN3248115.71.22144
a Polymerization in solution and with 1 h pre-contact step: solvent = toluene = 1 mL, temperature = 110 °C CHO:PA:cat:cocat = 250:250:1:1.
Table 5. CHO/PA copolymerization with different pre-contact times a.
Table 5. CHO/PA copolymerization with different pre-contact times a.
EntryCatalyst/Co-CatalystPolym.
Time (h)
Pre-Contact Time (h)Yield (%)Mn (kg/mol)ĐTg
(°C)
CHOPA 481/DMAP117615.51.20147
CHOPA 571/DMAP1246416.31.20147
CHOPA 201/DMAP2418215.41.30146
CHOPA 511/PPNCl119015.71.10146
CHOPA 601/PPNCl1249015.71.13146
CHOPA 181/PPNCl2419216.81.30145
CHOPA 541/PPNN3118414.81.10146
CHOPA 631/PPNN31248915.51.10143
CHOPA 371/PPNN32417915.21.10146
CHOPA 502/DMAP11164.51.36129
CHOPA 592/DMAP1248014.21.13145
CHOPA 242/DMAP2418514.11.24143
CHOPA 532/PPNCl11409.41.10133
CHOPA 622/PPNCl1249216.31.11146
CHOPA 222/PPNCl2419515.81.25145
CHOPA 562/PPNN311418.81.10140
CHOPA 652/PPNN31249316.21.20148
CHOPA 392/PPNN3241789.31.24144
CHOPA 493/DMAP11144.51.31144
CHOPA 583/DMAP1248415.41.13147
CHOPA 283/DMAP2419614.01.27144
CHOPA 523/PPNCl11357.81.21137
CHOPA 613/PPNCl1249814.81.09146
CHOPA 263/PPNCl2419215.11.26144
CHOPA 553/PPNN311397.51.19136
CHOPA 643/PPNN31249714.71.20145
CHOPA 413/PPNN32418115.71.22144
a Polymerization conditions: solvent = toluene = 1 mL, temperature = 110 °C, CHO:PA:cat:cocat = 250:250:1:1.

Share and Cite

MDPI and ACS Style

Proverbio, M.; Galotto Galotto, N.; Losio, S.; Tritto, I.; Boggioni, L. Influence of Co-Catalysts and Polymerization Conditions on Properties of Poly(anhydride-alt-epoxide)s from ROCOP Using Salen Complexes with Different Metals. Polymers 2019, 11, 1222. https://doi.org/10.3390/polym11071222

AMA Style

Proverbio M, Galotto Galotto N, Losio S, Tritto I, Boggioni L. Influence of Co-Catalysts and Polymerization Conditions on Properties of Poly(anhydride-alt-epoxide)s from ROCOP Using Salen Complexes with Different Metals. Polymers. 2019; 11(7):1222. https://doi.org/10.3390/polym11071222

Chicago/Turabian Style

Proverbio, Matteo, Nella Galotto Galotto, Simona Losio, Incoronata Tritto, and Laura Boggioni. 2019. "Influence of Co-Catalysts and Polymerization Conditions on Properties of Poly(anhydride-alt-epoxide)s from ROCOP Using Salen Complexes with Different Metals" Polymers 11, no. 7: 1222. https://doi.org/10.3390/polym11071222

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