Next Article in Journal
Effect of Vegetable Oil on the Properties of Asphalt Binder Modified with High Density Polyethylene
Next Article in Special Issue
Characterization of Fungal Foams from Edible Mushrooms Using Different Agricultural Wastes as Substrates for Packaging Material
Previous Article in Journal
Wood Surface Finishing with Transparent Lacquers Intended for Indoor Use, and the Colour Resistance of These Surfaces during Accelerated Aging
Previous Article in Special Issue
Preparation and Performance of Biodegradable Poly(butylene adipate-co-terephthalate) Composites Reinforced with Novel AgSnO2 Microparticles for Application in Food Packaging
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Copolymerization of Carbon Dioxide with 1,2-Butylene Oxide and Terpolymerization with Various Epoxides for Tailorable Properties

College of Chemistry and Chemical Engineering, Yantai University, Yantai 264005, China
*
Authors to whom correspondence should be addressed.
Polymers 2023, 15(3), 748; https://doi.org/10.3390/polym15030748
Submission received: 11 January 2023 / Revised: 27 January 2023 / Accepted: 30 January 2023 / Published: 1 February 2023
(This article belongs to the Special Issue Polymers: Environmental Aspects)

Abstract

:
The copolymerization of carbon dioxide (CO2) with epoxides demonstrates promise as a new synthetic method for low-carbon polymer materials, such as aliphatic polycarbonate materials. In this study, a binary Schiff base cobalt system was successfully used to catalyze the copolymerization of 1,2-butylene oxide (BO) and CO2 and its terpolymerization with other epoxides such as propylene oxide (PO) and cyclohexene oxide (CHO). 1H nuclear magnetic resonance (1H NMR), diffusion-ordered spectroscopy (DOSY), gel permeation chromatography (GPC), and differential scanning calorimetry (DSC) confirmed the successful synthesis of the alternating terpolymer. In addition, the effects of the polymerization reaction conditions and copolymerization monomer composition on the polymer structure and properties were examined systematically. By regulating the epoxide feed ratio, polycarbonates with an adjustable glass transition temperature (Tg) (11.2–67.8 °C) and hydrophilicity (water contact angle: 85.2–95.2°) were prepared. Thus, this ternary polymerization method provides an effective method of modulating the surface hydrophobicity of CO2-based polymers and their biodegradation properties.

1. Introduction

CO2 is an abundant, non-toxic, and non-flammable source of waste gases [1,2,3,4]. One of the most promising pathways for CO2 utilization involves the use of epoxides and CO2 to prepare new aliphatic polycarbonate materials, which not only uses CO2 as a carbon resource, but also provides a potentially sustainable method to produce valuable carbon-neutral polycarbonate materials [5,6,7,8].
Previously, monomers such as propylene oxide (PO) and cyclohexene oxide (CHO) have been used most frequently for polymerization to produce aliphatic polycarbonates [9,10]. Poly(propylene carbonate) (PPC), a copolymer of CO2 and PO, is the most promising species for industrialization because of its good biodegradation properties. However, previous studies have shown that it is difficult to use PPC alone as a film material due to its amorphous nature and low glass transition temperature of ~35 °C, making it extremely brittle at low temperatures, as well as its poor dimensional stability at high temperatures [11]. CHO with CO2 can produce poly(cyclohexane carbonate) (PCHC) with a Tg of 115 °C [12]. However, the mechanical fragility of PCHC severely limits its practical applications. Commonly, the addition of a third monomer for terpolymerization decreases the brittleness of PCHC [12,13,14]. In addition, changes in thermal properties can also be achieved by combining different monomers into random copolymers. Typically, random copolymers have only one glass transition temperature, which lies between the Tg values of each homopolymer and is determined by the concentration of each component. Recently, Darensbourg and his colleagues have demonstrated an efficient one-pot strategy to synthesize CO2-based triblock copolymers that can be 3D-printed via direct ink writing techniques to produce porous structures that can be modified or crosslinked. ABA-type triblock copolymers were synthesized using 1,2-butylene oxide (BO) as a low Tg block and vinyl cyclohexene oxide (VCHO) as a high Tg block, yielding high thermal stability, tunable thermal transition, and good mechanical properties [15].
Similar to PO and CHO, BO is another important epoxy monomer that is widely used in industry, and the polyether prepared by its ring-opening homopolymerization exhibits hydrophobicity and a low Tg, demonstrating good potential in polyurethane and lubrication applications. However, traditional BO manufacturing involves the chlorohydrin method, which consumes a large amount of chlorine gas and generates a great deal of waste water and residue containing organic chlorides, causing severe pollution. With the improvement of epoxidation technology in recent years, the unique process route of BO generation by 1-butene catalytic epoxidation (CHP or HPPO method) has become a popular study issue [16]. The route adopts 1-butene as a raw material, cumyl hydroperoxide or hydrogen peroxide as an oxidant, and direct epoxidation under the action of a silica–titanium molecular sieve catalyst to generate BO. Unlike the chlorohydrin technique, the CHP or HPPO route does not use chlorine gas, which avoids environmental pollution and provides the possibility of the green and low-cost production of BO. Its growth is also fostered in the new materials industry. BO-derived polyether materials have attracted considerable interest due to their unique features [17,18,19,20,21,22]. Coates et al. investigated the ring-opening copolymerization of a series of epoxides and cyclic anhydrides using [FSalph]AlCl and bis(triphenylphosphine)iminium chloride (PPNCl) as the catalysts [23]. The Tg values of the polyesters obtained from BO and the six cyclic anhydrides were less than those of the others, indicating that the introduction of methyl substituents can regulate the Tg of polyesters with firmly coupled ortho ester groups [24].
Previous studies have investigated the copolymerization of BO with CO2 for the preparation of polycarbonate. However, most of these studies have focused on the use of BO as a reaction substrate to verify the versatility of its catalyst [25,26,27]. Systematic studies on the copolymerization of BO with CO2 and its polymer properties are still lacking. Considering the positive effect of BO introduction on the thermal, mechanical, and biodegradation properties of the CO2-based polycarbonate, it is important to investigate the copolymerization of CO2 with BO. However, the copolymerization of BO with CO2 to obtain high-molecular-weight polycarbonate is still a challenge [28,29]. A suitable, efficient catalyst system is essential to achieve high-molecular-weight polycarbonates by the copolymerization of BO and CO2. Various catalyst systems, including metalloporphyrins [30], zinc glutarate [31], double metal cyanide (DMC) [32], Schiff base complexes [33], and organoboron [34,35], have been used to catalyze the copolymerization of epoxides and CO2. Among them, the Schiff base cobalt/co-catalyst system is widely used because of its high activity. In this study, (R,R)-SalenCoCl and bis(triphenylphosphine)ammonium chloride (PPNCl) were selected as the catalyst and co-catalyst, respectively, to catalyze the copolymerization of CO2 with BO to prepare high-molecular-weight poly(butylene carbonate) (PBC). In addition, terpolymerization reactions of BO/PO/CO2 and BO/CHO/CO2 involving BO were investigated (Scheme 1), and the effects of structural changes on the glass transition temperature, thermal stability, and surface characteristics of polycarbonates were also examined.

2. Materials and Methods

2.1. Materials

All water- and oxygen-sensitive compounds were handled in an argon-protected glove box. PO (98%), CHO (98%), and 1,2-butylene oxide (BO, 98%) were purchased from Energy Chemical (Shanghai, China) or Aldrich (Milwaukee, USA) and used under argon after refluxing with calcium hydride. Dichloromethane (DCM), ethanol, and high-performance liquid chromatography-grade tetrahydrofuran (THF) were purchased from Energy Chemical (Shanghai, China) and used as received. Bis(triphenylphosphine)iminium chloride (PPNCl, 98%) was purchased from Energy Chemical (Shanghai, China) and used after drying under a vacuum. Carbon dioxide (99.999%) was purchased from Yantai Feiyuan Special Gases Co., Ltd (Yantai, Shandong, China). and used as received. SalenCo(III)Cl was prepared according to the method reported by Jacobsen et al. [36].

2.2. Characterization

All crude products were sampled into 5-mm NMR tubes and dissolved in CDCl3. 1H NMR and diffusion-ordered spectroscopy (DOSY) were recorded on a JEOL-400YH (Tokyo, Japan) instrument at a frequency of 400 MHz. The number-average relative molecular masses and relative molecular mass distributions were determined on a Waters410 gel permeation chromatography (GPC) instrument (Milford, MA, USA), using tetrahydrofuran and polystyrene as the eluent and standard, respectively. Differential scanning calorimetry measurements (DSC) of the polymer samples were conducted on a NETZSCH DSC 200F3 instrument (Selb, Germany). Samples were prepared in aluminum pans. All samples were analyzed using the following temperature program: −50 to 150 °C at 10 °C min−1, 150 to −50 °C at 10 °C min−1, and then −50 to 150 °C at 10 °C min−1. Data were processed using the StarE software. All of the reported Tg values were observed at the second heating cycle. The thermal decomposition temperature (Td) of the polymers was determined on a NETZSCH STA 449 F5 Jupiter system (Selb, Germany) under nitrogen. For contact angle tests, the slides were washed in acetone for 2 h by sonication, and were then placed in a solution of the polymer in DCM (5 mg mL−1) and incubated overnight at room temperature. After the solvent was evaporated under air, the slides were washed with deionized water to remove unadhered particles. To estimate the contact angle, the fixation drop method was used. A drop of deionized water was added to the surface for measuring the angle between the solid and liquid phases. All samples were measured six times to reduce systematic errors.

2.3. Copolymerization Procedure

In a glove box, the required catalyst and epoxide monomers (viz. BO, PO, or CHO) were charged into a pre-dried 10 mL autoclave equipped with a magnetic stir bar. Next, the autoclave was removed from the glove box, pressurized with CO2 at the preset required temperature for the required time. Polymerization was terminated by cooling the autoclave to room temperature. After releasing CO2, a small aliquot of the copolymerization mixture was removed for 1H NMR spectroscopy. The remaining crude mixture was dissolved in CH2Cl2 and then precipitated in ethanol three times to remove the residual epoxide and cyclic carbonate, affording a purple solid product. The produced polymer was used for further characterization without separating the catalyst.

3. Results and Discussion

3.1. Copolymerization of BO/CO2

The copolymerization of CO2 and BO was successfully conducted by using an appropriate reaction temperature and an appropriate ratio of BO/cat/PPNCl, and the copolymerized product exhibited excellent reactivity and selectivity (Table 1). For example, PBC was prepared by the copolymerization of CO2 and BO at 25 °C for 12 h at a BO/cat/PPNCl mole ratio of 1000/1/1. The crude product (Figure 1A) and the purified product precipitated in ethanol (Figure 1B) were characterized by 1H NMR. The chemical shifts at 4.9 ppm and 4.2 ppm corresponded to the methine and methylene protons next to the carbonate linkage in the polycarbonate (Figure 1A). In addition, the methylene signals corresponding to the five-membered cyclic carbonate were observed at 4.5 ppm and 4.3 ppm. The peaks at 2.4 ppm, 2.7 ppm, and 2.8 ppm corresponded to the methylene and methine proton peaks of unreacted BO. In the 1H NMR spectra of the purified product, peaks corresponding to cyclic carbonate (h, g) and the raw material (e, f) disappeared completely, indicative of the synthesis of a pure polycarbonate (Figure 1B).
Copolymerization was considerably affected by the reaction temperature, reaction time, catalyst concentration, and CO2 pressure [37]. First, the effect of the reaction time on polymerization was examined. A series of polymerization reactions with different reaction times were conducted at 25 °C with a BO/cat/PPNCl feed mole ratio of 1000:1:1 (Table 1, entries 1–6). As expected, the conversion of the monomer increased gradually with time from 66.2% to 95.2%. Further studies revealed that the selectivity of the polymerized products decreased to 77.6% with increasing reaction time to 24 h, accompanied by a decrease in molecular weight to 40 kg mol−1. In general, it has been hypothesized that the metal-catalyzed copolymerization of CO2 and epoxides involves a coordination–insertion mechanism [38,39,40]. An assumption similar to that described above has also been proposed for binary SalenCo(III)X/nucleophilic co-catalyst systems, where two reaction routes are involved in the formation of cyclic carbonates: back-biting of the metal-bound carboxylate or the depolymerization of the propagating polycarbonates chain (unzipping) [41]. In the initial stage of the reaction, polycarbonates were formed selectively by the careful selection of the co-catalyst and reaction temperature. The prolonged reaction caused an increase in system viscosity and a decrease in mass and heat transfer, enabling the unzipping of the polymer. This result indicates that the selectivity of the polymerized product decreases with increasing reaction time, mainly due to the depolymerization of the resultant linear polycarbonate with a completely alternating arrangement, which produces more thermodynamically stable cyclic carbonate.
In addition, the effect of the catalyst concentration on the copolymerization of CO2 with BO at a reaction time of 12 h was evaluated (Table 1). With the decrease in the mole ratio to 500:1:1 (Table 1, entry 7 vs. 3), the BO conversion was significantly improved from 86.6% to >99%, and the selectivity of the polymerization product was also up to 99%. With the further reduction of the catalyst dosage with a BO/cat/PPNCl feed mole ratio of 2000:1:1, conversion of 67.8% with a TOF value of 113 h−1 was obtained (Table 1, entry 8), but the molecular weight was only 15.0 kg mol−1. When the mole ratio increased from 5000:1:1 and 10,000:1:1 (Table 1, entries 9 and 10), the purified polymer was barely obtained by the dissolution/precipitation process due to its low molecular weight. The 1H NMR spectrum also showed a decrease in conversion to 22.4% and 7.5%.
For the binary systems of SalenCoCl and PPNCl, the polymerization selectivity is highly sensitive to the reaction temperature. As can be observed from Table 1, the polymer selectivity was reduced to 46.9% at a reaction temperature of 40 °C (Table 1, entry 11). The further increase in the temperature decreased the selectivity to 39.7% (Table 1, entry 12). At 80 °C (Table 1, entry 13), the pressure of the reaction system dropped rapidly and then ceased dropping after a short time (e.g., 1 h). Accordingly, the reaction system reacted extremely rapidly at high temperatures, albeit with extremely poor selectivity. This result was consistent with the previously proposed mechanism: increasing the temperature led to a higher tendency for the reaction to produce a more thermodynamically stable cyclic carbonate. This result may be related to the fact that the counter cation is not linked intramolecularly to the cobalt complex. Therefore, the counter cation and dissociated anion can be far from the cobalt center, particularly at high reaction temperatures and/or in a highly diluted solution [42].
In addition, the effect of pressure on the reaction process was evaluated. After 12 h of reaction, the molecular weight of the polycarbonate gradually increased and then decreased at CO2 pressures of 1 MPa, 2 MPa, 3 MPa, and 4 MPa (Table 1, entries 14, 15, 3, 16, respectively). A molecular weight of 41.7 kg mol−1 was achieved at 3 MPa, and the molecular weight distribution did not change significantly under these conditions. Another effect of the pressure on polymerization is that, at low CO2 pressures, the carbonate unit content of PBC decreases, indicating that a small amount of polyether is produced, which is consistent with that reported in previous studies [43].

3.2. Terpolymerization of BO/CO2 and Other Epoxides

One of the advantages of epoxide/CO2 copolymerization is that polymer properties can be modified not only by varying the monomer but also by copolymerizing with another epoxide. It should be noted that although CO2/PO and CO2/CHO copolymers have been extensively studied, the lower glass transition temperature of PPC, the high brittleness of PCHC, and other inherent defects have seriously limited their application. The research on adjusting the Tg of polymers and improving their mechanical properties by introducing a third monomer has received much attention. The main difficulty in terpolymerization is the difference in reactivity between monomers, making it difficult to achieve effective control over the composition and structure of the polymer. Table 2 summarizes the results of the terpolymerization of BO/PO/CO2 and BO/CHO/CO2 catalyzed by the (R,R)-SalenCoCl and PPNCl binary system. To further clarify the terpolymer structure, and to exclude the possibility of the mechanical mixture of the two polymers (i.e., PBC and PPC), DOSY NMR analysis of the samples was conducted. The DOSY method is a form of two-dimensional nuclear magnetic resonance that is used to measure the diffusion coefficients of molecules. It is more suitable for distinguishing a mixture of two or three components with different self-diffusion factors. As shown in Figure 2, the DOSY spectra revealed only a single signal with a low diffusion coefficient, corresponding to the 1H NMR signals of PBC-co-PPC (Table 2, entry 18) and PBC-co-PCHC (Table 2, entry 21).
At a BO to PO feed ratio of 9:1 (Table 2, entry 17), the ratio of the PBC to PPC linkage in the terpolymer was 1:0.47. The polymer selectivity was ~96.0%, and the molecular weight was 58.8 kg mol−1. With the increase in the PO feed ratio, terpolymerization was more likely to occur, leading to an increase in the PPC linkage of the polymer chains. In addition, the selectivity of the polymer was decreased, its TOF value was not significantly altered, and its molecular weight was increased. The PBC to PPC linkage ratio in the terpolymer was 1:1.41 with a BO feed ratio of 5:5 (Table 2, entry 19). Accordingly, the polymer selectivity was decreased to 91.1%, and the molecular weight was 68.6 kg mol−1. In addition, the terpolymerization of CO2, BO, and CHO was investigated. At a BO to CHO feed ratio of 9:1 (Table 2, entry 20), the PBC to PCHC linkage of the polymer was 1:0.37, while the selectivity of the polymer was 97.1%, and the molecular weight was 32.2 kg mol−1. The molecular weight of the product increased first and then decreased with the increasing CHO feed ratio; at a BO to CHO feed ratio of 5:5, the molecular weight reached 48.4 kg mol−1 (Table 2, entry 22).

3.3. Thermal Analysis of PBC, PBC-co-PCHC, and PBC-co-PPC

DSC was employed to measure the Tg of the resultant copolymer and terpolymer. As shown in Figure 3, the Tg of PBC was 11.2 °C (Table 2, entry 3). All of the terpolymers in Table 2 with different molar ratios of BO (PO or CHO) exhibited a single Tg (Figure 3). At a BO:PO feed ratio of 9:1, Tg was 14.3 °C. When the PO feed ratio was increased to 1:1, the Tg of the resulting polymer increased to 20.1 °C. This result was related to the increase in the PO content of the polymer, which stiffened the chain segments; thus, their glass transition temperatures increased. In general, however, the introduction of BO further lowers the glass transition temperature of PPC, which is unfavorable for its use as a raw material for disposable film packaging due to its low glass transition temperature. In contrast, it is more practical to introduce BO into the copolymerization of CHO and CO2, resulting in the modulation of the material’s mechanical properties and its glass transition temperature. Therefore, the terpolymerization of BO, CHO, and CO2 was also investigated with a further increase in the Tg of the polymer. At a BO:CHO ratio of 9:1, a polymer with a molecular weight of 32.2 kg mol−1 and a Tg of 32.4 °C was produced. With the increase in the CHO content of the feed ratio, the chain rigidity and Tg increased, and at a feed ratio of 1:1, Tg reached 67.8 °C. This result was attributed to the presence of longer side chains of BO, resulting in stronger plasticizing effects and regioirregular microstructures. With the addition of PO/CHO, the otherwise flexible PBC chain segments slowly hardened, making the polycarbonate more rigid and increasing its Tg.
In addition, TGA was employed to analyze the thermal stability of the obtained polycarbonates. As shown in Figure 4, the Td of PBC-co-PPC at 5 wt% loss was 241 °C, which was less than those of PBC at 247 °C and PBC-co-PCHC at 255 °C.
It is important to note that in some systems, the decomposition of polymers (unzipping) may occur at very low temperatures due to the presence of the catalyst, but the resultant cPC cannot be volatilized at such low temperatures, making the change invisible on the TGA curve. To determine whether this is the case for our system, a control experiment was performed. The 1H NMR spectra before and after holding the polymer at a constant temperature of 240 °C for 10 min (inset in Figure 4) revealed low polymer degradation (the heated spectra revealed insignificant peaks of cyclic carbonate at 4.5 ppm and 4.3 ppm), but this change was almost negligible (the percentage of integrated area was greater than 99%). The results indicate that the unzipping reaction of the polycarbonate at low temperatures is insignificant during the extremely short TGA test.

3.4. Contact Angle Measurements

The hydrophobic and hydrophilic performance of polymers is well known to be an important factor for their biodegradability. Hydrophilicity promotes enzymatic degradation due to the presence of abundant surface water molecules. Therefore, an appropriate balance of hydrophobic and hydrophilic characteristics provides an opportunity to regulate the degradation rate of polymers [44,45]. Polybutylene oxide is more hydrophobic than polypropylene oxide. Thus, the introduction of the BO unit into PPC may enhance the hydrophobicity of the polymer. Thus far, only a few studies have investigated the surface properties of CO2-based polycarbonates [46,47]. Herein, the contact angles of PBC and terpolymers were measured. The contact angles of pure PBC and PPC were 95.2° and 79.5°, respectively. The replacement of the methyl group in PO with an ethyl or cyclohexane ring (in BO and CHO) resulted in the enhanced hydrophobicity of the polymer, i.e., the contact angle increased with the increase in the PBC content (Figure 5). The same result was observed for the terpolymers of CHO, BO, and CO2: with the increase in PBC content, the contact angle increased. Thus, this terpolymerization method offers a modification method for the surface characteristics of CO2-based polymers. The effect of surface hydrophilicity on the degradation rate is still being investigated.

4. Conclusions

In summary, the alternating copolymerization of CO2 with BO and the terpolymerization of CO2, BO, and PO or CHO were successfully realized by using a binary SalenCoCl catalyst. In addition, the effects of the polymerization reaction conditions and copolymerization monomer composition on the polymer structure and properties were investigated. The thermal properties of polycarbonates such as Tg and Td can be easily adjusted by alternating the epoxide feed ratio. In addition to modulating the thermal properties, a third monomer, such as PO or CHO, also can alter the hydrophobicity of CO2/BO polycarbonates, providing an opportunity to modulate the degradation rates.

Author Contributions

Draft preparation, investigation, and writing, S.T.; reviewing and editing, H.S., R.Q., H.T., M.S. and Y.Q.; resources, supervision, and project administration, Y.Q. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (No. m52073244), Taishan Scholar Program (TSQN201909086), and Central Government Special Funds Supporting the Development of Local Science and Technology (No. YDZX20203700001726).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cooper, A.I. Polymer synthesis and processing using supercritical carbon dioxide. J. Mater. Chem. 2000, 10, 207–234. [Google Scholar] [CrossRef]
  2. Arakawa, H.; Aresta, M.; Armor, J.N.; Barteau, M.A.; Beckman, E.J.; Bell, A.T.; Bercaw, J.E.; Creutz, C.; Dinjus, E.; Dixon, D.A.; et al. Catalysis Research of Relevance to Carbon Management:  Progress, Challenges, and Opportunities. Chem. Rev. 2001, 101, 953–996. [Google Scholar] [CrossRef] [PubMed]
  3. Sakakura, T.; Kohno, K. The synthesis of organic carbonates from carbon dioxide. Chem. Commun. 2009, 9, 1312–1330. [Google Scholar] [CrossRef] [PubMed]
  4. Schäffner, B.; Schäffner, F.; Verevkin, S.P.; Börner, A. Organic Carbonates as Solvents in Synthesis and Catalysis. Chem. Rev. 2010, 110, 4554–4581. [Google Scholar] [CrossRef] [PubMed]
  5. Darensbourg, D.J.; Holtcamp, M.W. Catalysts for the reactions of epoxides and carbon dioxide. Coord. Chem. Rev. 1996, 153, 155–174. [Google Scholar] [CrossRef]
  6. Hauenstein, O.; Agarwal, S.; Greiner, A. Bio-based polycarbonate as synthetic toolbox. Nat. Commun. 2016, 7, 11862. [Google Scholar] [CrossRef]
  7. Qin, Y.S.; Wang, X.H. Carbon dioxide-based copolymers: Environmental benefits of PPC, an industrially viable catalyst. Biotechnol. J. 2010, 5, 1164–1180. [Google Scholar] [CrossRef]
  8. Zhu, Y.; Romain, C.; Williams, C.K. Sustainable polymers from renewable resources. Nature 2016, 540, 354–362. [Google Scholar] [CrossRef]
  9. Darensbourg, D.J.; Wilson, S.J. What’s new with CO2? Recent advances in its copolymerization with oxiranes. Green Chem. 2012, 14, 2665–2671. [Google Scholar] [CrossRef]
  10. Lu, X.B.; Ren, W.M.; Wu, G.P. CO2 Copolymers from Epoxides: Catalyst Activity, Product Selectivity, and Stereochemistry Control. Acc. Chem. Res. 2012, 45, 1721–1735. [Google Scholar] [CrossRef]
  11. Allen, S.D.; Byrne, C.M.; Coates, G.W. Carbon Dioxide as a Renewable C1 Feedstock: Synthesis and Characterization of Polycarbonates from the Alternating Copolymerization of Epoxides and CO2. In Feedstocks for the Future; ACS Symposium Series; ACS: Washington, DC, USA, 2006; Volume 921, pp. 116–129. [Google Scholar]
  12. Koning, C.; Wildeson, J.; Parton, R.; Plum, B.; Steeman, P.; Darensbourg, D.J. Synthesis and physical characterization of poly(cyclohexane carbonate), synthesized from CO2 and cyclohexene oxide. Polymer 2001, 42, 3995–4004. [Google Scholar] [CrossRef]
  13. Hsu, T.J.; Tan, C.S. Block copolymerization of carbon dioxide with butylene oxide, propylene oxide and 4-vinyl-1-cyclohexene-1,2-epoxide in based poly(cyclohexene carbonate). J. Chin. Inst. Chem. Eng. 2003, 34, 335–344. [Google Scholar]
  14. Thorat, S.D.; Phillips, P.J.; Semenov, V.; Gakh, A. Physical properties of aliphatic polycarbonates made from CO2 and epoxides. J. Appl. Polym. Sci. 2003, 89, 1163–1176. [Google Scholar] [CrossRef]
  15. Wei, P.; Bhat, G.A.; Cipriani, C.E.; Mohammad, H.; Schoonover, K.; Pentzer, E.B.; Darensbourg, D.J. 3D Printed CO2-Based Triblock Copolymers and Post-Printing Modification. Angew. Chem. Int. Ed. 2022, 61, e202208355. [Google Scholar] [CrossRef]
  16. Khatib, S.J.; Oyama, S.T. Direct Oxidation of Propylene to Propylene Oxide with Molecular Oxygen: A Review. Catal. Rev. 2015, 57, 306–344. [Google Scholar] [CrossRef]
  17. Allgaier, J.; Willbold, S.; Chang, T. Synthesis of Hydrophobic Poly(alkylene oxide)s and Amphiphilic Poly(alkylene oxide) Block Copolymers. Macromolecules 2007, 40, 518–525. [Google Scholar] [CrossRef]
  18. Misaka, H.; Tamura, E.; Makiguchi, K.; Kamoshida, K.; Sakai, R.; Satoh, T.; Kakuchi, T. Synthesis of end-functionalized polyethers by phosphazene base-catalyzed ring-opening polymerization of 1,2-butylene oxide and glycidyl ether. J. Polym. Sci. A Polym. Chem. 2012, 50, 1941–1952. [Google Scholar] [CrossRef]
  19. Isono, T.; Kamoshida, K.; Satoh, Y.; Takaoka, T.; Sato, S.-i.; Satoh, T.; Kakuchi, T. Synthesis of Star- and Figure-Eight-Shaped Polyethers by tBu-P4-Catalyzed Ring-Opening Polymerization of Butylene Oxide. Macromolecules 2013, 46, 3841–3849. [Google Scholar] [CrossRef]
  20. Zhao, J.; Alamri, H.; Hadjichristidis, N. A facile metal-free “grafting-from” route from acrylamide-based substrate toward complex macromolecular combs. Chem. Commun. 2013, 49, 7079–7081. [Google Scholar] [CrossRef]
  21. Zhao, J.; Pahovnik, D.; Gnanou, Y.; Hadjichristidis, N. A “Catalyst Switch” Strategy for the Sequential Metal-Free Polymerization of Epoxides and Cyclic Esters/Carbonate. Macromolecules 2014, 47, 3814–3822. [Google Scholar] [CrossRef]
  22. Liu, Y.; Wei, W.; Xiong, H. Gradient and block side-chain liquid crystalline polyethers. Polym. Chem. 2015, 6, 583–590. [Google Scholar] [CrossRef]
  23. Sanford, M.J.; Peña Carrodeguas, L.; 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]
  24. Yu, X.; Jia, J.; Xu, S.; Lao, K.U.; Sanford, M.J.; Ramakrishnan, R.K.; Nazarenko, S.I.; Hoye, T.R.; Coates, G.W.; DiStasio, R.A. Unraveling substituent effects on the glass transition temperatures of biorenewable polyesters. Nat. Commun. 2018, 9, 2880. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, X.H.; Wei, R.J.; Zhang, Y.Y.; Du, B.Y.; Fan, Z.Q. Carbon Dioxide/Epoxide Copolymerization via a Nanosized Zinc–Cobalt(III) Double Metal Cyanide Complex: Substituent Effects of Epoxides on Polycarbonate Selectivity, Regioselectivity and Glass Transition Temperatures. Macromolecules 2015, 48, 536–544. [Google Scholar] [CrossRef]
  26. Scharfenberg, M.; Seiwert, J.; Scherger, M.; Preis, J.; Susewind, M.; Frey, H. Multiarm Polycarbonate Star Polymers with a Hyperbranched Polyether Core from CO2 and Common Epoxides. Macromolecules 2017, 50, 6577–6585. [Google Scholar] [CrossRef]
  27. Jia, M.; Zhang, D.; de Kort, G.W.; Wilsens, C.H.R.M.; Rastogi, S.; Hadjichristidis, N.; Gnanou, Y.; Feng, X. All-Polycarbonate Thermoplastic Elastomers Based on Triblock Copolymers Derived from Triethylborane-Mediated Sequential Copolymerization of CO2 with Various Epoxides. Macromolecules 2020, 53, 5297–5307. [Google Scholar] [CrossRef]
  28. Takeda, N.; Inoue, S. Polymerization of 1,2-epoxypropane and copolymerization with carbon dioxide catalyzed by metalloporphyrins. Makromol. Chem. 1978, 179, 1377–1381. [Google Scholar] [CrossRef]
  29. Ren, W.-M.; Zhang, X.; Liu, Y.; Li, J.F.; Wang, H.; Lu, X.B. Highly Active, Bifunctional Co(III)-Salen Catalyst for Alternating Copolymerization of CO2 with Cyclohexene Oxide and Terpolymerization with Aliphatic Epoxides. Macromolecules 2010, 43, 1396–1402. [Google Scholar] [CrossRef]
  30. Cao, H.; Qin, Y.S.; Zhuo, C.W.; Wang, X.H.; Wang, F.S. Homogeneous Metallic Oligomer Catalyst with Multisite Intramolecular Cooperativity for the Synthesis of CO2-Based Polymers. ACS Catal. 2019, 9, 8669–8676. [Google Scholar] [CrossRef]
  31. Ree, M.; Hwang, Y.; Kim, J.S.; Kim, H.; Kim, G.; Kim, H. New findings in the catalytic activity of zinc glutarate and its application in the chemical fixation of CO2 into polycarbonates and their derivatives. Catal. Today 2006, 115, 134–145. [Google Scholar] [CrossRef]
  32. Liu, S.J.; Qin, Y.S.; Chen, X.F.; Wang, X.H.; Wang, F.S. One-pot controllable synthesis of oligo(carbonate-ether) triol using a Zn-Co-DMC catalyst: The special role of trimesic acid as an initiation-transfer agent. Polym. Chem. 2014, 5, 6171–6179. [Google Scholar] [CrossRef]
  33. Lu, X.B.; Wang, Y. Highly Active, Binary Catalyst Systems for the Alternating Copolymerization of CO2 and Epoxides under Mild Conditions. Angew. Chem. Int. Ed. 2004, 43, 3574–3577. [Google Scholar] [CrossRef]
  34. Yang, G.W.; Zhang, Y.Y.; Xie, R.; Wu, G.P. Scalable Bifunctional Organoboron Catalysts for Copolymerization of CO2 and Epoxides with Unprecedented Efficiency. J. Am. Chem. Soc. 2020, 142, 12245–12255. [Google Scholar] [CrossRef]
  35. Zhang, D.; Boopathi, S.K.; Hadjichristidis, N.; Gnanou, Y.; Feng, X.S. Metal-Free Alternating Copolymerization of CO2 with Epoxides: Fulfilling “Green” Synthesis and Activity. J. Am. Chem. Soc. 2016, 138, 11117–11120. [Google Scholar] [CrossRef]
  36. Nielsen, L.P.C.; Stevenson, C.P.; Blackmond, D.G.; Jacobsen, E.N. Mechanistic Investigation Leads to a Synthetic Improvement in the Hydrolytic Kinetic Resolution of Terminal Epoxides. J. Am. Chem. Soc. 2004, 126, 1360–1362. [Google Scholar] [CrossRef]
  37. Darensbourg, D.J.; Bottarelli, P.; Andreatta, J.R. Inquiry into the Formation of Cyclic Carbonates during the (Salen)CrX Catalyzed CO2/Cyclohexene Oxide Copolymerization Process in the Presence of Ionic Initiators. Macromolecules 2007, 40, 7727–7729. [Google Scholar] [CrossRef]
  38. Coates, G.W.; Moore, D.R. Discrete Metal-Based Catalysts for the Copolymerization of CO2 and Epoxides: Discovery, Reactivity, Optimization, and Mechanism. Angew. Chem. Int. Ed. 2004, 43, 6618–6639. [Google Scholar] [CrossRef]
  39. Lu, X.B.; Darensbourg, D.J. Cobalt catalysts for the coupling of CO2 and epoxides to provide polycarbonates and cyclic carbonates. Chem. Soc. Rev. 2012, 41, 1462–1484. [Google Scholar] [CrossRef]
  40. Decortes, A.; Castilla, A.M.; Kleij, A.W. Salen-Complex-Mediated Formation of Cyclic Carbonates by Cycloaddition of CO2 to Epoxides. Angew. Chem. Int. Ed. 2010, 49, 9822–9837. [Google Scholar] [CrossRef]
  41. Ren, W.M.; Wu, G.P.; Lin, F.; Jiang, J.Y.; Liu, C.; Luo, Y.; Lu, X.-B. Role of the co-catalyst in the asymmetric coupling of racemic epoxides with CO2 using multichiral Co(III) complexes: Product selectivity and enantioselectivity. Chem. Sci. 2012, 3, 2094–2102. [Google Scholar] [CrossRef]
  42. Ren, W.M.; Liu, Z.W.; Wen, Y.Q.; Zhang, R.; Lu, X.B. Mechanistic Aspects of the Copolymerization of CO2 with Epoxides Using a Thermally Stable Single-Site Cobalt(III) Catalyst. J. Am. Chem. Soc. 2009, 131, 11509–11518. [Google Scholar] [CrossRef] [PubMed]
  43. Darensbourg, D.J.; Mackiewicz, R.M.; Phelps, A.L.; Billodeaux, D.R. Copolymerization of CO2 and Epoxides Catalyzed by Metal Salen Complexes. Acc. Chem. Res. 2004, 37, 836–844. [Google Scholar] [CrossRef] [PubMed]
  44. Nagata, M.; Kiyotsukuri, T. Biodegradability of copolyesteramides from hexamethylene adipate and hexamethyleneadipamide. Eur. Polym. J. 1994, 30, 1277–1281. [Google Scholar] [CrossRef]
  45. Nagata, M. Enzymatic degradation of aliphatic polyesters copolymerized with various diamines. Macromol. Rapid Commun. 1996, 17, 583–587. [Google Scholar] [CrossRef]
  46. Jia, M.; Zhang, D.; Gnanou, Y.; Feng, X.S. Surfactant-Emulating Amphiphilic Polycarbonates and Other Functional Polycarbonates through Metal-Free Copolymerization of CO2 with Ethylene Oxide. ACS Sustainable Chem. Eng. 2021, 9, 10370–10380. [Google Scholar] [CrossRef]
  47. Song, B.; Qin, A.J.; Tang, B.Z. Syntheses, properties, and applications of CO2-based functional polymers. Cell Rep. Phys. Sci. 2022, 3, 100719. [Google Scholar] [CrossRef]
Scheme 1. Copolymerization of CO2 and epoxides (BO, PO, and CHO).
Scheme 1. Copolymerization of CO2 and epoxides (BO, PO, and CHO).
Polymers 15 00748 sch001
Figure 1. 1H NMR spectra of crude products (A) and purified products (B) of PBC (Table 1, entry 3).
Figure 1. 1H NMR spectra of crude products (A) and purified products (B) of PBC (Table 1, entry 3).
Polymers 15 00748 g001
Figure 2. DOSY spectra of (A) PBC-co-PPC and (B) PBC-co-PCHC.
Figure 2. DOSY spectra of (A) PBC-co-PPC and (B) PBC-co-PCHC.
Polymers 15 00748 g002
Figure 3. DSC analysis of the obtained polycarbonates.
Figure 3. DSC analysis of the obtained polycarbonates.
Polymers 15 00748 g003
Figure 4. TGA curves of the obtained polycarbonates and 1H NMR of pure PBC (A) after heating for 10 min (B).
Figure 4. TGA curves of the obtained polycarbonates and 1H NMR of pure PBC (A) after heating for 10 min (B).
Polymers 15 00748 g004
Figure 5. Water contact angles of CO2-based polycarbonates.
Figure 5. Water contact angles of CO2-based polycarbonates.
Polymers 15 00748 g005
Table 1. Copolymerization of CO2 and BO.
Table 1. Copolymerization of CO2 and BO.
Polymers 15 00748 i001
Entrycat:cocat:BOTime
(h)
Conv.
(%) 2
CU
(%) 3
Selectivity
(%) 4
TOF
(h−1) 5
Mn
(kg mol−1) 6
Đ 6
1 11:1:1000466.2>99>9916619.51.31
2 11:1:1000875.6>99>999528.01.34
3 11:1:10001286.6>9998.17241.61.27
4 11:1:10001692.6>9988.55738.21.28
5 11:1:10002093.7>9980.04636.91.30
6 11:1:10002495.2>9977.64021.71.35
7 11:1:50012>99>99>994236.31.26
8 11:1:20001267.8>99>9911315.01.27
9 11:1:50001222.4>99>9993//
10 11:1:10,000127.5>9994.363//
11 71:1:10001293.3>9946.97831.11.12
12 81:1:100012>99>9939.78324.71.46
13 91:1:10001283.6>9912.57020.51.20
14 101:1:10001270.998.0>995926.01.28
15 111:1:10001272.5>99>996031.01.27
16 121:1:10001284.7>99>997031.11.30
1 The copolymerization was carried out at 25 °C and 3 MPa unless noted otherwise. 2 Conversion of BO determined by 1H NMR analysis of the crude reaction mixture. 3 The content of carbonate unit. 4 Percentage of polymer formed vs. cyclic carbonate as determined by 1H NMR. 5 TOF = turnover frequency, mol of epoxide consumed × mol of SalenCoCl−1 × h−1. 6 Determined by GPC in THF, at 30 °C, calibrated with polystyrene standards. 7 The reaction temperature was 40 °C. 8 The reaction temperature was 60 °C. 9 The reaction temperature was 80 °C. 10 The reaction pressure was 1 MPa. 11 The reaction pressure was 2 MPa. 12 The reaction pressure was 4 MPa.
Table 2. Terpolymerization of CO2, BO, and other epoxides.
Table 2. Terpolymerization of CO2, BO, and other epoxides.
EntryEpoxideBO:epo 1PBC:PPC/PCHCSelectivity
(%)
Mn
(kg mol−1) 2
Đ 2Tg
(°C) 3
3\\\98.141.61.2711.2
17PO9:11:0.4795.658.81.1214.3
18PO7:31:1.1595.262.71.0815.9
19PO5:51:1.4191.168.61.0720.1
20CHO9:11:0.3797.132.21.1432.4
21CHO7:31:1.0298.050.71.1242.6
22CHO5:51:2.2297.148.41.1167.8
1 Mole ratio of epoxide, [epoxide]:SalenCoCl:PPNCl = 1000:1:1. 2 Determined by GPC in THF, at 30 °C, calibrated with polystyrene standards. 3 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

Tang, S.; Suo, H.; Qu, R.; Tang, H.; Sun, M.; Qin, Y. Copolymerization of Carbon Dioxide with 1,2-Butylene Oxide and Terpolymerization with Various Epoxides for Tailorable Properties. Polymers 2023, 15, 748. https://doi.org/10.3390/polym15030748

AMA Style

Tang S, Suo H, Qu R, Tang H, Sun M, Qin Y. Copolymerization of Carbon Dioxide with 1,2-Butylene Oxide and Terpolymerization with Various Epoxides for Tailorable Properties. Polymers. 2023; 15(3):748. https://doi.org/10.3390/polym15030748

Chicago/Turabian Style

Tang, Shuo, Hongyi Suo, Rui Qu, Hao Tang, Miao Sun, and Yusheng Qin. 2023. "Copolymerization of Carbon Dioxide with 1,2-Butylene Oxide and Terpolymerization with Various Epoxides for Tailorable Properties" Polymers 15, no. 3: 748. https://doi.org/10.3390/polym15030748

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