Next Article in Journal
Coordination Chemistry of Uranyl Ions with Surface-Immobilized Peptides: An XPS Study
Previous Article in Journal
In Vitro and In Silico Studies for the Identification of Potent Metabolites of Some High-Altitude Medicinal Plants from Nepal Inhibiting SARS-CoV-2 Spike Protein
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Aliphatic Polycarbonates from Diphenyl Carbonate and Diols over Zinc (II) Acetylacetonate

1
Key Laboratory for Green Process of Chemical Engineering of Xinjiang Bingtuan, School of Chemistry and Chemical Engineering, Shihezi University, Shihezi 832003, China
2
Xinjiang Tianye (Group) Coporation Limited, Shihezi 832000, China
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(24), 8958; https://doi.org/10.3390/molecules27248958
Submission received: 12 November 2022 / Revised: 12 December 2022 / Accepted: 13 December 2022 / Published: 16 December 2022
(This article belongs to the Section Macromolecular Chemistry)

Abstract

:
APCs (aliphatic polycarbonates) are one of the most important types of biodegradable polymers and widely used in the fields of solid electrolyte, biological medicine and biodegradable plastics. Zinc-based catalysts have the advantages of being low cost, being non-toxic, having high activity, and having excellent environmental and biological compatibility. Zinc (II) acetylacetonate (Zn(Acac)2) was first reported as a highly effective catalyst for the melt transesterification of biphenyl carbonate with 1,4-butanediol to synthesize poly(1,4-butylene carbonate)(PBC). It was found that the weight-average molecular weight of PBC derived from Zn(Acac)2 could achieve 143,500 g/mol with a yield of 85.6% under suitable reaction conditions. The Lewis acidity and steric hindrance of Zn2+ could obviously affect the catalytic performance of Zn-based catalysts for this reaction. The main reasons for the Zn(Acac)2 catalyst displaying a higher yield and Mw than other zinc-based catalysts should be ascribed to the presence of the interaction between acetylacetone ligand and Zn2+, which can provide this melt transesterification reaction with the appropriate Lewis acidity as well as the steric hindrance.

1. Introduction

As an environment-friendly polymer, aliphatic polycarbonates (APCs) have attracted much attention in recent years for their widespread application in the fields of waterborne polyurethane, solid electrolytes and biological medicine; with a number-average molecular weight (Mw) greater than 70 Kg/mol, they can also partly replace polyethylene as biodegradable plastics to solve the problem of white pollution [1,2,3,4]. Therefore, the effective synthesis of APCs with a high molecular weight has become more and more important. The current reported routes for APC synthesis mainly include copolymerization of CO2 with epoxides or diols [5,6], ring-opening polymerization of cyclic carbonates (ROP) [7] and condensation polymerization of dialkyl carbonate with aliphatic diols [8]. Moreover, both of the latter processes can also be considered as indirect utilization of CO2. Particularly, melt transesterification of dialkyl carbonate with aliphatic diols has established a bridge from CO2 to APCs, as dimethyl carbonate (DMC) and diphenyl carbonate (DPC) can be produced from CO2 and methanol or phenol [9].
Recently, the melt transesterification of dialkyl carbonate with diols to prepare high-molecular-weight polycarbonates has received increasing attention for synthesizing polymers with diverse structure and few catalysts. It is well known that the activity of a catalyst for this process is closely related to its acidity and chelating ability of metal specie. Therefore, numerous transition metal compounds were found to be highly active catalysts for this process, such as organotin oxides [10], titanium compounds [3,11] and even simple metal salts [12,13]. As highly effective polymerization catalysts, many zinc salts have been extensively used in the synthesis of various polymers due to their characteristics of being low cost, nontoxic and biofriendly [14,15,16]. In a previous paper, the Zn2+ cation was found to be able to activate the alcoholic hydroxyl groups by coordinating the oxygen atom in the hydroxyl group and inducing the transesterification reaction [13]. Pastusiak et al. [17] found that Zinc(II) acetylacetonate (Zn(Acac)2) could initiate the ring-opening polymerization of cyclic trimethylene carbonate resulting in high-molecular-weight poly(trimethylenecarbonate). Zn(Acac)2 was also found to be an excellent catalyst for polylactic acid synthesis [18].
Inspired by these referenced works, Zn(Acac)2 was first employed as a catalyst for the one-pot melt transesterification of DPC and aliphatic diols to synthesize high-molecular-weight APCs. Using ZnCl2 as a contrast, the reaction conditions were explored in detail with Zn(Acac)2 as a catalyst for its excellent catalytic performance. In addition, XPS was used to characterize the catalyst structure to understand the influencing factors of zinc salt in the melt transesterification reaction of DPC with diols.

2. Results and Discussions

2.1. Selection of Catalysts

The catalytic performance of various zinc compounds for the melt transesterification of DPC with BD to synthesize PBC was evaluated at 180 °C and 200 Pa with a reaction time of 90 min; the Mw, Ð and corresponding yield are listed in Table 1. As reported in a previous paper [13], this reaction scarcely occurred in the absence of a catalyst due to the relatively low nucleophilicity of the hydroxyl group in BD, and no reaction fraction could be observed in the blank experiment. One can also see in Table 1 that the Mw of PBC obtained over Zn(Acac)2 is 69,400 g/mol at the given conditions, with a Ð and yield of 1.64 and 93.7%, respectively. As discussed in the literature [3,19,20], the presence of by-products, including tetrahydrofuran, cyclic carbonate and other volatile oligomers, would occur during the melt transesterification of dialkyl carbonates with BD, resulting in a decrease in the PBC yield. This reaction was also performed using other zinc compounds as catalysts; ZnO, ZnCO3 and Zn(OAc)2 were all found to demonstrate relatively less activity, producing PBCs with a Mw of 8300, 28,200 and 53,200 g/mol at the same conditions, respectively. Though the Mw of PBC derived with ZnCl2 possessed the maximum value of 82,300 g/mol, the PBC yield was the lowest among all the tested catalysts. Additionally, not all the zinc compounds were active in this reaction; for example, ZnSO4 was almost inert with no fraction and polymer detected under the same conditions.
Additionally, Ð seems to only depend on the Mw of PBC, and the higher the Mw value, the wider the Ð value. One can also see from Table 1 that increasing the molar ratio of Zn(Acac)2 to DPC from 0.025% to 0.1% increases the Mw of PBC from 9600 g/mol to 69,400 g/mol; continuing to increase the Zn(Acac)2 concentration to 0.2% led to an obvious decrease in the Mw and yield of PBC to 62,500 g/mol and 88.4%, respectively. Obviously, a molar ratio of Zn2+ to DPC of 0.1 mol% should be the suitable catalyst amount for this reaction considering the Mw and yield comprehensively. Therefore, Zn(Acac)2 was selected as the model catalyst for further research.

2.2. Characterization of Poly (Butylene Carbonate)

The resultant polymers obtained over Zn(Acac)2 and ZnCl2 were analyzed by FT-IR and 1H NMR, and the results are shown in Figure 1. As observed in Figure 1a, the FT-IR spectra of the two samples are very close to each other. The absorption bands appearing at 2963 cm−1 and 2875 cm−1 are attributed to the asymmetric and symmetric C-H stretching vibration of methylene, respectively. The strong absorption bands at 1744 cm−1 and 1249 cm−1 can be ascribed to the stretching and asymmetric stretching vibrations of C=O and O-C-O of the carbonate backbone, respectively [3,13]. One can also see that all the synthesized PBC samples described herein have identical 1H NMR spectra, and only two strong signals appearing at 4.12 and 1.73 ppm can be observed in Figure 1b for both samples, which are attributed to a and b protons from BD units. No remarkable feature for the end-group was detected with their chemical shift at 3.64 or 7.32–7.38 ppm, suggesting the resultant polymers bear rather high molecular weight. Additionally, no peak at 3.4–3.5 ppm can be observed in the 1H-NMR spectrum of the PBC polymer, indicating there is no ether linage (-CH2-O-CH2-) in the PBC polymer, which is not hydrolysable and decreases the mechanical properties of the polymer [3]. Obviously, all the peaks for the two polymers are well concordant with the standard spectrum of PBC, and it is consistent with what is expected for a PBC structure [3,11,12,13].

2.3. XPS of Catalyst

The electronic property of Zn2+ in ZnCl2 and Zn(Acac)2 was also further examined using XPS to understand the relationship between catalytic performance and the nature of the catalyst; the results are illustrated in Figure 2. One can see that the binding energy of Zn 2p3/2 in the Zn(Acac)2 catalyst appeared at 1021.7 eV, ascribed to the presence of bivalence of Zn(II) [21,22]. The binding energy of ZnCl2 appeared at 1022.8 eV, which was higher than that of Zn(Acac)2. That is to say, the Lewis acid strength of ZnCl2 is stronger than that of Zn(Acac)2 in view of the concept for Lewis acid.

2.4. Effect of Reaction Conditions

In order to obtain the optimum conditions and further understand the relationship between the structure and catalytic performance of the catalyst, the melt transesterification of DPC and BD was performed under various reaction parameters with the ZnCl2 catalytic system as the control experiment. The effect of polymerization temperature was first examined in the range of 160–210 °C. The results, shown in Figure 3a, indicate that the Mw values of PBC over Zn(Acac)2 and ZnCl2 sharply increased when raising the polymerization temperature from 160 to 190 °C, which can often be ascribed to the acceleration of the diffusion-limited polycondensation kinetics due to the decrease in polymer viscosity at a higher temperature [23]. Then, the Mw gradually decreased as the temperature continuously increased. As for the ZnCl2 catalyst, the optimum temperature for the highest Mw values of 102,400 g/mol can be observed at 190 °C, while that for Zn(Acac)2 is 200 °C. Clearly, the Mw for ZnCl2 at a lower temperature is much higher than that of Zn(Acac)2, indicating that strong Lewis acidity seems to be a positive factor for the improvement of polymerization rate. Simultaneously, it is well known that there would be a completion effect between polymerization and decomposition during the whole process because the melt transesterification of DPC and BD is a typical reverse reaction. Therefore, pure ZnCl2 more easily expedites the decomposition and depolymerization of the obtained PBC polymer, which often can be explained by the fact that strong Lewis acidity is prone to attack the carbonyl oxygen atoms in APCs and impede the growing of the polymer chain to lower the Mw and yield [24]. This is also in accordance with the results shown in Figure 3b, in which the yield for the ZnCl2 catalyst sharply decreased with the rise of temperature. The excellent catalytic performance of Zn(Acac)2 may be explained by the fact that the coordination bond formed between the acetylacetone ligand and Zn2+ not only can decrease the Lewis acidity of the central Zn2+ but can also make Zn2+ attack carbonyl oxygen atoms and hinder undesirable side reactions [19]. Hence, the connection of -OH and the -OC(O)OC6H5 end-group while removing the generated phenol at reduced pressure to increase the polymer molecular chain proceeded smoothly. Obviously, 200 °C should be selected as the suitable polymerization temperature for the Zn(Acac)2 catalyst considering its Mw and yield comprehensively.
Figure 4 shows the dependence of the Mw and yield of PBC versus reaction time over different catalysts. As shown in Figure 4a, under a short reaction time of 30 min, the Mw was only 67,400 g/mol over Zn(Acac)2. As the reaction proceeded, the Mw of PBC rapidly increased to 143,500 g/mol with increasing reaction time to 120 min. However, when the time was beyond 120 min, the value for Mw showed no significant improvement. Likewise, the Mw values for ZnCl2 also increased as the reaction time was prolonged, and the Mw reached the maximum value of 122,500 g/mol at 60 min; then, the Mw of PBC would rapidly decline to 12,000 g/mol with further increases of time to 150 min. The reason for this phenomenon might be that this polymerization process very easily proceeds at the beginning, but with an increase in molecular weight, the viscosity of the reaction system becomes high, which causes a negative influence to further polymerization [13,19]. Therefore, excessive reaction time could enhance its reverse reaction and lead to the decrease in Mw. Meanwhile, one can also see in Figure 4b that the yield of PBC continuously decreases with the prolongation of reaction time, which can be reasoned by the presence of a side reaction and sublimation of oligomer as reported in previous works [20]. Considering the above results, a temperature of 200 °C and a reaction time of 120 min were selected as the optimum reaction conditions for realizing the highest Mw with satisfactory yield for Zn(Acac)2.

2.5. Catalytic Activity towards Other Diols

To evaluate the potential and general application range of the Zn(Acac)2 catalyst, the catalytic melt transesterification of DPC with a verity of aliphatic diols, including 1,3-propanediol (PPD), 1,5-pentanediol (PD) and 1,6-hexanediol (HD) to synthesize corresponding aliphatic polycarbonates, poly(trimethylene carbonate) (PTMC), poly(pentamethylene carbonate)(PPMC), poly(hexamethylene carbonate) (PHC) and poly(hexamethylene)-co-poly(butylene carbonate) (PHBC) were investigated. As shown in Table 2, other common aliphatic diols, including PD, HD and their mixtures can also undergo efficient transesterification with DPC to high-molecular-weight APCs with high yields under the optimized reaction conditions. The inferior catalytic performance of PPD was thought to be related to its low boiling point or the poor thermal stability of the corresponding PTMC polymer [13].

3. Experimental Section

3.1. Materials

Commercial DPC, purchased from Guanghua Scitech Co., Ltd., Shenzhen, China, was purified by recrystallization in absolute ethyl alcohol. 1,4-butanediol (BD, 98%) was dehydrated by distillation over calcium hydride under dry nitrogen gas. ZnCl2, Zn(OAc)2, Zn(NO3)2 and ZnSO4 were obtained from Chengdu Kelong Chemical Reagent Co., Chengdu, China, and were dehydrated as described in the literature [11] before use. Other chemicals and catalysts were used without any further purification and treatment.

3.2. Synthesis of APCs

PBC polymer was synthesized by a one-pot melt polymerization method [13]. Typically, DPC (21.41 g, 0.1 mol), BD (9.01 g, 0.1 mol) and a certain amount of catalyst were charged into a 150 mL three-necked flask equipped with a mechanical stirrer, reflux condenser and thermometer. The reaction mixture was heated to 120 °C under stirring for a certain time until it became homogeneous under a nitrogen atmosphere. Then, a lower pressure (ca 200 Pa) was applied slowly over a period of ca 20 min to carry out the melt transesterification reaction at the given temperature. After a certain time, the pressure was returned to atmospheric pressure, and the volatile by-products could be removed through the reflux condenser. The resulting PBC polymer was separated by dissolving in CH2Cl2 and precipitating with ethanol, and then dried under vacuum at 50 °C for 12 h.

3.3. Characterization

The chemical structures of the resulting PBCs were identified by 1H-NMR. The 1H-NMR spectra were acquired in CDCl3 at 25 °C with a Bruker DRX-300 NMR (Brucker, Romanshorn, Switzerland) spectrometer. The weight-average molecular weight (Mw) and dispersity (Ð) of the obtained APCs were determined by gel permeation chromatography (GPC). The GPC measurements were carried out at 30 °C on a Waters 515 HPLC system (Waters, Milford, MA, USA) equipped with a 2690D separation module and a 2410 refractive index detector. Tetrahydrofuran (THF) was used as the eluent at a flow rate of 0.5 mL/min. Polystyrene with a narrow molecular weight distribution was used as the standard for calibration. All the given data in this study were collected by averaging the scores on at least three tests.
Fourier transform infrared spectroscopy (FT-IR) was carried out on a Nicolet-38 FT-IR spectrometer (Thermo Electron, Boston, MA, USA) in the range of 400–4000 cm−1 using the KBr pellet technique. The TGA experiments were carried out using a Q600 SDT thermal analysis machine (TA instrument, Waltham, MA, USA) under a flow of N2 in a temperature range from 50 °C to 550 °C with a heating rate of 10 °C/min. The binding energy values and the atomic surface concentration of the corresponding elements of the samples were analyzed by X-ray photoelectron spectroscopy and performed on an ESCA LAB 250 photoelectron spectroscope at 3.0 × 10−10 mbar with a hemispherical analyzer and monochromatic Mg Kα radiation (E = 1253.6 eV). All the binding energies were referenced to the C1s peak at 284.5 eV of the surface adventitious carbon.

4. Conclusions

The catalytic properties of zinc (II) acetylacetonate in melt transesterification of diphenyl carbonate with aliphatic diols were investigated. The Mw, yield and Ð of the obtained PBCs are influenced by the catalyst used, reaction temperature and time. Lewis acidity is found to be dominant for the polymerization rate at lower temperature, and increasing steric hindrance seems to give a positive effect on the improvement of yield. Therefore, the rise in Mw and yield for Zn(Acac)2 compared with ZnCl2 is mainly due to the decrease in Lewis acidity and the increase in steric hindrance.

Author Contributions

Conceptualization, X.S., Z.W. (Zhong Wei) and Z.W. (Ziqing Wang); methodology, J.F. (Jun Feng), J.L.,J.F. (Jian Feng) and Z.W. (Zhong Wei); software, J.F. (Jun Feng); validation, J.F. (Jun Feng) and Z.W. (Zhong Wei); formal analysis, X.S.; investigation, J.F. (Jun Feng) and J.L.; resources, X.S. and Z.W. (Zhong Wei); data curation, J.F. (Jian Feng) and Z.W. (Zhong Wei); writing—original draft, J.F. (Jun Feng) and Z.W.; writing—review and editing, Z.W. (Zhong Wei) and X.S.; visualization, J.L., supervision, Z.W. (Ziqing Wang) and X.S.; project administration, Z.W. (Ziqing Wang) and J.F. (Jun Feng); funding acquisition, Z.W. (Ziqing Wang) and X.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Key industrial innovation development plans in XinJiang production and construction corps (2021DB004), the significant science and technology project of XinJiang production and construction corps (2022AA002) and the significant science and technology project of Shihezi (2022JB02).

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.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Liu, Y.; Ren, W.M.; Zhang, W.P.; Zhao, R.R.; Lu, X.B. Crystalline CO2-based polycarbonates prepared from racemic catalyst through intramolecularly interlocked assembly. Nat. Commun. 2015, 6, 8594. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Wu, D.D.; Li, W.; Zhao, Y.; Deng, Y.J.; Zhang, H.L.; Zhang, H.X.; Dong, L.S. Thermal, mechanical and rheological properties of biodegragable poly(propylene carbonate) and poly(butylene carbonate) blends. Chin. J. Poly Sci. 2015, 33, 444–455. [Google Scholar] [CrossRef]
  3. Zhu, W.X.; Huang, X.; Li, C.C.; Xiao, Y.N.; Zhang, D.; Guan, G.H. High molecular weight aliphatic polycarbonates by melt polycondensation of dimethyl carbonate and aliphatic diols: Synthesis and characterization. Polym. Int. 2011, 60, 1060–1067. [Google Scholar] [CrossRef]
  4. Wang, L.P.; Xiao, B.; Wang, G.Y.; Wu, J.Q. Synthesis of polycarbonate diol catalyzed by metal-organic-framework Zn4O[CO2-C6H4CO2]3. Sci. China Chem. 2011, 54, 1468–1473. [Google Scholar] [CrossRef]
  5. Qin, Y.S.; Sheng, X.F.; Liu, S.J.; Ren, G.J.; Wang, X.H.; Wang, F.S. Recent advances in carbon dioxide based copolymers. J. CO2 Util. 2015, 11, 3–9. [Google Scholar] [CrossRef]
  6. Tamura, M.; Nakagawa, Y.; Tomishige, K. Direct CO2 transformation to aliphatic polycarbonates. Asian J. Org. Chem. 2022, 11, e202200445. [Google Scholar] [CrossRef]
  7. Guillaume, S.M.; Carpentier, J.F. Recent advances in metallo/organo-catalyzed immortal ring-opening polymerization of cycliccarbonates. Catal. Sci. Technol. 2012, 2, 898–906. [Google Scholar] [CrossRef]
  8. Park, J.H.; Jeon, J.Y.; Le, J.J.; Jang, Y.; Varghese, J.K.; Lee, B.Y. Preparation of high-molecular-weight aliphatic polycarbonates by condensation polymerization of diols and dimethyl carbonate. Macromolecules 2013, 46, 3301–3308. [Google Scholar] [CrossRef]
  9. Li, Z.H.; Qin, Z.F. Synthesis of diphenyl carbonate from phenol and carbon dioxide in carbon tetrachloride with zinc halides as catalyst. J. Mol. Catal. A Chem. 2007, 264, 255–259. [Google Scholar] [CrossRef]
  10. He, X.L.; Li, Z.H.; Su, K.M.; Cheng, B.W.; Ming, J. Study on the reaction between bisphenol A and dimethyl carbonate over organotin oxide. Catal. Commun. 2013, 33, 20–23. [Google Scholar] [CrossRef]
  11. Liu, F.; Yang, X.G.; Li, J.G.; Liu, S.Y.; Yao, J.; Chen, T.; Wang, G.Y. Synthesis of poly (butylene carbonate) by melt transesterification of diphenyl carbonate and 1,4-butanediol. Acta Polym. Sin. 2014, 5, 628–635. [Google Scholar]
  12. Wang, Z.Q.; Yang, X.G.; Liu, S.Y.; Zhang, H.; Wang, G.Y. Magnesium acetate used as an effective catalyst for synthesizing aliphatic polycarbonates via melt transesterification process. Chem. Res. Chin. Univ. 2016, 32, 512–516. [Google Scholar] [CrossRef]
  13. Wang, Z.Q.; Yang, X.G.; Liu, S.Y.; Hu, J.; Zhang, H.; Wang, G.Y. One-pot synthesis of high-molecular-weight aliphatic polycarbonates via melt transesterification of diphenyl carbonate and diols using Zn(OAc)2 as a catalyst. RSC Adv. 2015, 5, 87311–87319. [Google Scholar] [CrossRef]
  14. Liu, M.S.; Li, X.; Lin, X.L.; Liang, L.; Gao, X.X.; Sun, J.M. Facile synthesis of [urea-Zn]I2 eutectic-based ionic liquid for efficient conversion of carbon dioxide to cyclic carbonates. J. Mol. Catal. A Chem. 2016, 412, 20–26. [Google Scholar] [CrossRef]
  15. Yang, R.L.; Xu, G.Q.; Lv, C.D.; Dong, B.Z.; Zhou, L.; Wang, Q.G. Zn(HMDS)2 as a versatile transesterification catalyst for polyesters synthesis and degradation toward a circular materials economy approach. ACS Sustain. Chem. Eng. 2020, 8, 18347–18353. [Google Scholar] [CrossRef]
  16. Esposito, R.; Melchiorre, M.; Annunziata, A.; Cucciolito, M.E.; Ruffo, F. Emerging catalysis in biomass valorisation: Simple Zn(II) catalysts for fatty acids esterification and transesterification. ChemCatChem 2020, 12, 5858–5879. [Google Scholar] [CrossRef]
  17. Pastusiak, M.; Dobrzynski, P.; Kaczmarczyk, B.; Kasperczyk, J. Polymerization mechanism of trimethylene carbonate carried out with Zinc(II) acetylacetonate monohydrate. J. Polym. Sci. Part A Polym. Chem. 2011, 49, 2504–2512. [Google Scholar] [CrossRef]
  18. Pastusik, M.; Dobrzynski, P.; Kaczmarczyk, B.; Kasperczyk, J.; Smola, A. The polymerization mechanism of lactide initiated with Zinc (II) acetylacetonate monohydrate. Polymer 2011, 52, 5255–5261. [Google Scholar] [CrossRef]
  19. Wang, Z.Q.; Yang, X.G.; Liu, S.Y.; Hu, J.; Wang, G.Y. Synthesis of aliphatic polycarbonates by transesterification route with poly (vinylpyrrolidone)-immobilized zinc halide as catalysts. Acta Polym. Sin. 2016, 12, 1654–1661. [Google Scholar]
  20. He, Q.; Zhang, Q.; Liao, S.R.; Zhao, C.S.; Xie, X.Y. Understanding cyclic by-products and ether linkage formation pathways in the transesterification synthesis of aliphatic polycarbonates. Eur. Polym. J. 2017, 97, 253–262. [Google Scholar]
  21. Jeon, H.S.; Timoshenko, J.; Scholten, F.; Sinev, I.; Herzog, A.; Haase, F.T.; Cuenya, B.R. Operando insight into the correlation between the structure and composition of CuZn nanoparticles and their selectivity for the electrochemical CO2 reduction. J. Am. Chem. Soc. 2019, 141, 19879–19887. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Schmidt, S.A.; Kumar, N.; Shchukarev, A.; Eränena, K.; Mikkola, J.P.; Murzina, D.Y.; Salmi, T. Preparation and characterization of neat and ZnCl2 modified zeolites and alumina for methyl chloride synthesis. Appl. Catal. A Gen. 2013, 468, 120–134. [Google Scholar] [CrossRef]
  23. Wang, J.T.; Zhu, Q.; Lu, X.L.; Meng, Y.Z. ZnGA-MMT catalyzed the copolymerization of carbon dioxide with propylene oxide. Eur. Polym. J. 2005, 41, 1108–1114. [Google Scholar] [CrossRef]
  24. Jung, J.H.; Ree, M.; Kim, H. Acid- and base-catalyzed hydrolyses of aliphatic polycarbonates and polyesters. Catal. Today 2006, 115, 283–287. [Google Scholar] [CrossRef]
Figure 1. (a) FT-IR spectra and (b) 1H-NMR spectra of resultant copolymers obtained over ZnCl2 and Zn(Acac)2.
Figure 1. (a) FT-IR spectra and (b) 1H-NMR spectra of resultant copolymers obtained over ZnCl2 and Zn(Acac)2.
Molecules 27 08958 g001
Figure 2. XPS spectra of Zn 2p3/2 for (a) ZnCl2 and (b) Zn(Acac)2.
Figure 2. XPS spectra of Zn 2p3/2 for (a) ZnCl2 and (b) Zn(Acac)2.
Molecules 27 08958 g002
Figure 3. The effect of reaction temperature on the (a) Mw and (b) yield of PBC polymer over Zn(Acac)2 and ZnCl2. Reaction conditions: molar ratio of Zn2+ to DPC 0.10%, reaction time 90 min.
Figure 3. The effect of reaction temperature on the (a) Mw and (b) yield of PBC polymer over Zn(Acac)2 and ZnCl2. Reaction conditions: molar ratio of Zn2+ to DPC 0.10%, reaction time 90 min.
Molecules 27 08958 g003
Figure 4. The effects of reaction time on the (a) Mw and (b) yield of PBC polymer over ZnCl2 and Zn(Acac)2. Reaction conditions: molar ratio of Zn2+ to DPC 0.1%, reaction temperature 200 °C.
Figure 4. The effects of reaction time on the (a) Mw and (b) yield of PBC polymer over ZnCl2 and Zn(Acac)2. Reaction conditions: molar ratio of Zn2+ to DPC 0.1%, reaction temperature 200 °C.
Molecules 27 08958 g004
Table 1. The catalytic performance of various zinc-based catalysts for the condensation polymerization of BD and DPC to PBC a.
Table 1. The catalytic performance of various zinc-based catalysts for the condensation polymerization of BD and DPC to PBC a.
Entry CatalystMw (Kg/mol)Ð (Mw/Mn)PBC Yield b (%)
1ZnCl282.31.7269.0
2Zn(NO3)213.21.6393.2
3ZnSO4n.dn.dn.d
4ZnCO328.21.5793.5
5ZnO8.31.4696.2
6Zn(OAc)253.21.6292.5
7Zn(Acac)269.41.6493.7
8 cZn(Acac)29.61.5295.2
9 dZn(Acac)242.81.6392.8
10 eZn(Acac)262.51.6488.4
a Reaction conditions: reaction temperature 180 °C, reaction time 90 min, reaction pressure 200 Pa, the molar ratio of Zn2+ to DPC 0.1 mol %; b Yield expressed as a percentage of the theoretical value, which was calculated based on the 100% conversion of DPC to APCs; c the molar ratio of Zn2+ to DPC 0.025%; d the molar ratio of Zn2+ to DPC 0.05%; e the molar ratio of Zn2+ to DPC 0.2 mol%.
Table 2. Melt transesterification of DPC with a verity of diols in the presence of the Zn(Acac)2 catalyst a.
Table 2. Melt transesterification of DPC with a verity of diols in the presence of the Zn(Acac)2 catalyst a.
EntryDiolsProductMw (Kg/mol)Yield (%)Ð
1PPDPTMC13.242.71.36
2BDPBC143.585.61.76
3PDPPMC141.593.61.81
4HD PHC159.292.31.80
5BD + HD bPBHC152.489.61.80
a Reaction conditions: reaction temperature 200 ℃, reaction time 120 min, the molar ratio of Zn2+ to DPC 0.1 mol %; b the molar ratio of BD to HD is 1:1.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Feng, J.; Li, J.; Feng, J.; Wei, Z.; Wang, Z.; Song, X. Synthesis of Aliphatic Polycarbonates from Diphenyl Carbonate and Diols over Zinc (II) Acetylacetonate. Molecules 2022, 27, 8958. https://doi.org/10.3390/molecules27248958

AMA Style

Feng J, Li J, Feng J, Wei Z, Wang Z, Song X. Synthesis of Aliphatic Polycarbonates from Diphenyl Carbonate and Diols over Zinc (II) Acetylacetonate. Molecules. 2022; 27(24):8958. https://doi.org/10.3390/molecules27248958

Chicago/Turabian Style

Feng, Jun, Jin Li, Jian Feng, Zhong Wei, Ziqing Wang, and Xiaoling Song. 2022. "Synthesis of Aliphatic Polycarbonates from Diphenyl Carbonate and Diols over Zinc (II) Acetylacetonate" Molecules 27, no. 24: 8958. https://doi.org/10.3390/molecules27248958

Article Metrics

Back to TopTop