Next Article in Journal
Significantly Enhanced Electromechanical Performance of PDMS Crosslinked PVDF Hybrids
Next Article in Special Issue
Identity of Low-Molecular-Weight Species Formed in End-To-End Cyclization Reactions Performed in THF
Previous Article in Journal
A Novel In Situ Self-Assembling Fabrication Method for Bacterial Cellulose-Electrospun Nanofiber Hybrid Structures
Previous Article in Special Issue
Ring-Expansion/Contraction Radical Crossover Reactions of Cyclic Alkoxyamines: A Mechanism for Ring Expansion-Controlled Radical Polymerization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Organocatalytic Stereoselective Cyclic Polylactide Synthesis in Supercritical Carbon Dioxide under Plasticizing Conditions

1
Department of Engineering, Graduate School of Integrated Science and Technology, Shizuoka University, 3-5-1 Johoku, Hamamatsu, Shizuoka 432-8561, Japan
2
Graduate School of Science and Technology, Shizuoka University, 3-5-1 Johoku, Hamamatsu, Shizuoka 432-8561, Japan
3
Research Institute of Green Science and Technology, Shizuoka University, 3-5-1 Johoku, Hamamatsu, Shizuoka 432-8561, Japan
*
Author to whom correspondence should be addressed.
Polymers 2018, 10(7), 713; https://doi.org/10.3390/polym10070713
Submission received: 4 June 2018 / Revised: 23 June 2018 / Accepted: 26 June 2018 / Published: 28 June 2018
(This article belongs to the Special Issue Cyclic Polymers)

Abstract

:
Cyclic polylactide (cPLA) is a structural isomer of linear polylactide (PLA) although it possesses unique functionalities in comparison to its linear counterpart. Hitherto, the control of stereochemical purity in conventional cPLA synthesis has not been achieved. In this study, highly stereochemically pure cPLA was synthesized in the absence of a metal catalyst and organic solvent, which required high consumption of the residual monomer. The synthesis was conducted in supercritical carbon dioxide under CO2 plasticizing polymerization conditions in the presence of an organocatalyst and thiourea additives. In comparison with the stereocomplexes synthesized through conventional methods, cPLA from l-lactide (cPLLA) and cPLA from d-lactide (cPDLA) were synthesized with higher stereochemical purity and improved thermal stability. Moreover, the method presented herein is environmentally friendly and thus, applicable on an industrial level.

Graphical Abstract

1. Introduction

Linear polylactides (PLAs) have high biodegradability and biocompatibility, with their potential applications in the medical, pharmaceutical and materials fields having drawn a significant amount of attention [1,2,3]. Cyclic polylactide (cPLA), which is a structural isomer of PLA with an unique topology, also displays favorable properties [4,5,6,7,8]. For example, a study on the effect of cPLAs on tumor cells and tumor-bearing mice demonstrated that cPLAs effectively inhibit tumor cell growth [9]. Furthermore, in a recent study, cPLA was used as a stabilizer for palladium nanoparticles in the synthesis of a cPLA-clay hybrid material, which can be a recyclable catalyst, for use in the aminocarbonylation reaction of aryl halides with various amines [10]. Despite these attractive properties, cPLAs remain underexplored in comparison to linear PLAs both with respect to their synthesis and practical applications [10].
Until now, three strategies have been reported for the synthesis of cPLAs (Scheme 1): (1) polymerization and separation; (2) polymerization and cyclization; and (3) ring-opening polymerization (ROP) and cyclization. Strategy (1) has proved to be inefficient as even when fractionated by a reverse-phase octadecylsilyl (ODS) column chromatography, a mixture of linear PLAs and cPLAs is obtained [11]. Strategy (2) requires several steps, although uniform cPLAs can be eventually separated using a gel filtration column [12,13,14,15,16]. Nevertheless, it is still inappropriate for large-scale syntheses. In contrast, strategy (3) allows for a simple and straightforward one-step synthesis of cPLAs. The latter of these strategies was reported in 2007 using an organocatalytic method [17,18,19,20,21,22,23], while researchers have used a metal complex catalytic method in 2011 [24,25,26,27,28,29].
According to these reports, cPLA was obtained with high conversion (85–97%), moderate polydispersity index (PDI = 1.22–3.60) and high number-average molecular weight (Mn = 6400–69,000). The metal complex catalysts displayed poor reactivity and required high temperatures (30–160 °C) and long reaction times (4–18.5 h). In contrast, the reaction efficiency of N-heterocyclic carbene (NHC) as an organocatalyst was extremely high and the polymerization reaction occurred within 30 s at 25 °C. However, there were several drawbacks in that the polymerization reactions were conducted in flammable or toxic organic solvents, such as THF or CH2Cl2, and required drybox or Schlenk techniques under nitrogen, thereby limiting the potential for industrialization [19].
In general, the available methodologies fail to control the stereochemistry of PLAs, which is one of the most important parameters in material development as it has a significant effect on the properties [30,31]. In fact, in most examples of cPLA syntheses, racemic lactides are employed and thus, control of the stereochemical purity is insufficiently achieved. To the best of our knowledge, the stereocomplexes of cPLA were reported in 2011 by Waymouth et al. [19]. However, the fraction of isotactic (iii) tetrads of cPLA prepared from L-lactide was 0.81–0.83, suggesting that epimerization occurred during the polymerization step. Recently, Kricheldorf et al. reported a racemization-free polymerization, in which a cyclic tin catalyst was employed under bulk polymerization conditions, but a high temperature (160 °C) was required [29].
Based on our previous findings on the synthesis of stereochemically pure linear PLA [32], in this study, we have developed a method for the synthesis of highly pure cPLA in the absence of metals, organic solvents or residual monomers. Instead, the approach employs an organocatalyst in supercritical carbon dioxide (scCO2) under CO2 plasticizing polymerization (CPP) conditions at low temperatures. The results were promising as both L- and D-cPLA were synthesized in high stereochemical purity from L- and D-lactide 1, respectively (Scheme 2), which are hereafter denoted as cPLLA and cPDLA, respectively. An improvement in the thermal properties of the stereocomplexes was also observed.

2. Materials and Methods

2.1. Materials

In this study, 1-(3,5-Bis(trifluoromethyl)phenyl)-3-cyclohexyl thiourea was synthesized according to literature procedures [33]. All other chemicals and solvents were commercially available and used as received. L-Lactide was provided by Ricoh Co., Ltd. (Tokyo, Japan), while CO2 was obtained from Air Liquide Kogyogas Ltd. (Tokyo, Japan).

2.2. Synthesis

Polymerization reactions were conducted in a scCO2 reaction system using the TVS-N2-type portable reactor (20 MPa, 260 °C) manufactured by Taiatsu Techno Corporation (Tokyo, Japan) with the JASCO PU-1586 scCO2 pump (JASCO Corporation, Tokyo, Japan). L-Lactide (1, 864 mg, 6.0 mmol, 1.0 eq) and the organocatalyst (4-(dimethylamino)pyridine (DMAP), 0.40 mmol, 6.6 mol %) were added to a 12-mL pressure-resistant reactor, before the mixture was heated to 60 °C using a water bath. The vessel was charged with scCO2 (60 °C, 10 MPa) and the solution was stirred for 5 min. The reaction mixture was allowed to stand for 5 h, before the pressure was reduced from supercritical to atmospheric, yielding cPLA as a white solid. For the thiourea-mediated cPLLA synthesis, 1-(3,5-bis(trifluoromethyl)phenyl)-3-cyclohexyl thiourea (1 mol %) was added to the mixture of lactide and DMAP. This mixture was subjected to the polymerization reaction for 2.5 h as described above.
cPLLA and cPDLA samples with different number-average molecular weights were isolated by preparative gel permeation chromatography (GPC) using a recycling high performance liquid chromatography (HPLC) system (YMC LC-Forte/R, YMC CO., Ltd., Kyoto, Japan) with a YMC-GPC T30000 (21.2 mm × 600 mm) and YMC-GPC T4000 (21.2 mm × 600 mm) columns in series (CHCl3 as the eluent, flow rate of 6.0 mL/min). For preparation of the stereocomplex of cPLA [19], cPLLA (50 mg) and cPDLA (50 mg) were dissolved separately in 5 mL of CH2Cl2. These solutions were mixed together in a 20-mL flask and stirred for 5 min at room temperature. The solvent was evaporated slowly (200 torr) at room temperature, before the residue was dried under a vacuum to remove the residual solvent. The resulting material was used for DSC measurements.

2.3. Characterization

The conversion by polymerization was determined through 1H NMR analysis. 1H NMR (300 MHz) spectra were recorded using the JEOL JNM-AL spectrometer (JEOL, Tokyo, Japan) at ambient temperature. The spectra were recorded using CDCl3 as a solvent and tetramethylsilane (TMS) as an internal standard (δ = 0 ppm). The number- and weight-average molecular weights (Mn, Mw) and polydispersity index (PDI = Mw/Mn) of the products were determined by gel permeation chromatography (GPC) analysis using the SHIMADZU LC-5A pump (SHIMADZU, Kyoto, Japan) and JAI R1-3H RI detector (Japan Analytical Industry Co., Ltd., Tokyo, Japan) with GPC column GPC K-806L (Shodex, Tokyo, Japan, 8.0 mm × 300 mm, flow rate 0.8 mL/min), using polystyrene as a standard and CHCl3 as an eluent. LAsoft CDS-Lite software (LAsoft, Ltd., Chiba, Japan) was used for the molecular weight calculations.
The topology of the polymer product being cyclic rather than linear was confirmed by 1H NMR and matrix assisted laser desorption/ionization time-of-flight mass spectrometry (MALDI-TOF-MS) (Supplementary Materials, Figures S1 and S2). The mass spectra were recorded using ultrafleXtreme mass spectrometry (MALDI-TOF-MS, Bruker Japan, Yokohama, Japan). For MALDI-TOF-MS measurements, the cPLA was dissolved in CHCl3. Trans-2-[3-(4-tert-butylphenyl)-2-methyl-2-propenylidene]malononitrile (DCTB) was used as the matrix and silver trifluoroacetate was added as a cation source.
In order to hydrolyze the polymer materials, an aqueous solution of NaOH (1 M, 10 mL) was added to the polymer. After 4 h under reflux conditions, the mixture was cooled down to the ambient temperature and an aqueous solution of HCl (2 M, 5.5 mL) was added. The resulting lactic acid was analyzed by chiral HPLC. The HPLC analysis was conducted under the following conditions: SUMICHIRAL OA5000 column (Sumika Chemical Analysis Service, Ltd., 4.6 mm × 150 mm, flow rate 1.0 mL/min) with 2 mM CuSO4 aq/2-propanol = 95:5; UV detection at 254 nm (L-lactide: retention time = 8.5 min; D-lactide: retention time = 10.3 min) (Supplementary Materials, Figure S3).
Differential scanning calorimetry (DSC) analysis was performed using SHIMADZU DSC-60 (SHIMADZU, Kyoto, Japan). For DSC measurements, a 5-mg sample was heated from 25 to 250 °C at 10 °C/min rate under an argon flow (50 mL/min). The data collection interval was 1.0 s.

3. Results and Discussion

The synthesis of cPLA via an organocatalytic ROP was examined according to previous findings on the synthesis of stereochemically pure linear PLA (Table 1) [32]. Notably, 1H NMR analysis of the polymerization reaction mixture showed that the remaining unreacted monomers were less than 5% in any of the polymerization reactions. The monomer conversion rate and weight of the recovered polymer were both higher than 95%. In the conventional solution polymerization method, the reaction was slow and resulted in the formation of amorphous cPLLA with low stereochemical purity, which was determined by HPLC analysis after hydrolysis of cPLLA to lactic acid (Table 1, entry 1). In contrast, the reaction proceeded smoothly in scCO2 and crystalline cPLLA was obtained in a high enantiomeric excess (ee) of 90.5% (Table 1, entry 2). The success of the polymerization reaction under scCO2 conditions was attributed to the high concentration conditions similar to those in bulk polymerization and uniform conditions similar to those in solution polymerization. We refer to this process as a CO2 plasticizing polymerization (CPP) method [32]. Under CPP conditions, epimerization was suppressed since the reactive zwitterionic intermediate (3) was less likely to be solvated by scCO2 (dielectric constant (εr) = 1.15 at 10 MPa, 60 °C) as this has a lower εr than pentane (εr = 1.84) and hexane (εr = 1.89) [34,35]. Thus, this allows for the ROP to be preferred over the epimerization, which would form the meso-lactide 1 (Scheme 3). It was assumed that the stereochemical purity of cPLLA could be improved by increasing the polymerization consumption rate of the monomer. Thus, the additives that selectively activated the carbonyl group of L-lactide were investigated (Figure 1). In particular, 1-(3,5-bis(trifluoromethyl)phenyl)-3-cyclohexylthiourea accelerated the ROP reaction and improved the stereochemical purity, giving an ee of 93.5% (Table 1, entry 3). Given the success of the reaction using L-lactide, the same reaction conditions were applied to the ROP of D-lactide, with cPDLA being obtained in high stereoselectivity (entries 4 and 5). The PDI of the cPLA formed herein was 1.20–1.60, which was comparable to that obtained via the previously reported NHC catalytic method (1.3–1.4) [19].
In the cPLA synthesis, since a cyclic structure was formed via an intramolecular cyclization, the PDI could not be improved unless the linear intermediate of the same molecular weight is selectively cyclized.
Similar to PLA, cPLA has been known to form a stereocomplex of cPLA (sc-cPLA) [19,36,37]. In this work, cPDLA was synthesized according to the same procedure as that of cPLLA, while differential scanning calorimetry (DSC) was applied to create sc-cPLA (Figure 2). Interestingly, although the molecular weight was lower than the one reported in the literature (Table 2, entries 1 and 2 vs. 3), the measured melting point was higher than the previously reported value [19], which indicates stronger intermolecular forces. The melting point of sc-cPLA with the highest stereochemical purity was 212 °C. As the stereochemical purity decreased, the melting point decreased to 207 °C (Table 2, entries 1 vs. 2). Therefore, we can conclude that the higher stereochemical purity of the D- and L-forms led to stronger intermolecular interactions between the two forms in the stereocomplex and hence, a higher melting point (Figure 3). In addition, similar to the case of the stereocomplex of cPLA, the melting point of cPLA itself also depended on the stereochemical purity.
Subsequently, the effect of the ring size on the thermal stability was investigated (Table 3 and Figure 4). Three stereocomplex samples with different number-average molecular weights were prepared using cPLLA and cPDLA, which are shown in Table 2, entry 2. These were obtained and purified by a preparative GPC. Sample 1, which had very different molecular weights of cPLLA and cPDLA, had a melting point of 190 °C (Table 3, entry 1). In Sample 2, a more moderate difference in the molecular weight distribution led to an increase in the melting point to 200 °C (Table 3, entry 2). Finally, the smallest difference in the molecular weights in sample 3 gave the highest melting point (Table 3, entry 3). These results suggested that the degree of formation of the stereocomplex depended not only on the stereochemical purity, but also on the ring size (and similarities thereof) because the overlapping portion between cPLLA and cPDLA was limited due to a different radius of the curvature.

4. Conclusions

In this study, highly stereochemically pure cPLA was synthesized in the absence of a metal catalyst or organic solvent, which required high consumption of the residual monomer. The synthesis was conducted in scCO2 under CO2 plasticizing polymerization conditions in the presence of an organocatalyst and thiourea additives. In this approach, the epimerization of the obtained cPLA was suppressed and the stereochemical purity was higher than those obtained from conventional methods, although the control of stereochemical purity in conventional cPLA synthesis has not yet been achieved. In comparison to stereocomplexes synthesized through the conventional methods, cPLA from L-lactide (cPLLA) and cPLA from D-lactide (cPDLA) stereocomplexes were synthesized with higher stereochemical purity and improved thermal stability. Based on these results, we concluded that the method presented herein could meet the increasing demands of medical and material applications in terms of safety and high functionality in the large-scale industrial manufacturing of cPLA and in academic research.

Supplementary Materials

The experimental and 1H NMR spectroscopic and analytic data are available online at https://www.mdpi.com/2073-4360/10/7/713/s1.

Author Contributions

All authors contributed equally. N.M. conceived and designed the experiments; S.Y. and Y.N. performed the experiments and analyzed the data; K.S. and T.N. contributed reagents/materials/analysis tools; M. wrote the paper.

Funding

This work was funded in part by the Grant-in-Aid for Scientific Research on Innovative Areas (No. 24105511, 26105722) from the Ministry of Education, Culture, Sports, Science and Technology (MEXT).

Acknowledgments

We gratefully acknowledge the excellent technical assistance for MALDI-TOF-MS measurements T.N. and N.S. from Bruker Japan K.K. We express our appreciation to Ricoh Co., Ltd. for providing highly pure l-lactide.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Thakur, V.K.; Thakur, M.K. Handbook of Polymers for Pharmaceutical Technologies: Biodegradable Polymers; Scrivener Publishing LLC: New York, NY, USA, 2015; Volume 3, pp. 1–583. [Google Scholar] [CrossRef]
  2. Rieger, B.; Kunkel, A.; Coates, G.W.; Reichardt, R.; Dinjus, E.; Zevaco, T.A. Synthetic Biodegradable Polymers; Springer: Berlin/Heidelberg, Germany, 2012; Volume 245, pp. 1–364. [Google Scholar]
  3. Lendlein, A.; Sisson, A. Handbook of Biodegradable Polymers: Isolation, Synthesis, Characterization and Applications; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2011; pp. 1–405. [Google Scholar]
  4. Nomura, K.; Ohara, H. Synthesis and function of cyclic poly(lactic acid) and oligo(lactic acid). Mini-Rev. Org. Chem. 2017, 14, 35–43. [Google Scholar] [CrossRef]
  5. Tezuka, Y. Topological Polymer Chemistry: Progress of Cyclic Polymers in Syntheses, Properties and Functions; World Scientific Publishing Co., Pte., Ltd.: Singapore, 2013; pp. 1–352. [Google Scholar]
  6. Yamamoto, T.; Tezuka, Y. Cyclic and multicyclic topological polymers. In Complex Macromolecular Architectures: Synthesis, Characterization, and Self-Assembly; Hadjichristidis, N., Hirao, A., Tezuka, Y., DuPrez, F., Eds.; John Wiley & Sons (Asia) Pte. Ltd.: Singapore, 2011; pp. 3–19. [Google Scholar]
  7. Endo, K. Synthesis and properties of cyclic polymers. In New Frontiers in Polymer Synthesis; Kobayashi, S., Ed.; Springer: Berlin/Heidelberg, Germany, 2008; Volume 217, pp. 121–183. [Google Scholar]
  8. Roovers, J. Organic cyclic polymers. In Cyclic Polymers, 2nd ed.; Semlyen, J.A., Ed.; Kluwer Academic Publishers: Dordrecht, The Netherlands, 2000; pp. 347–384. [Google Scholar]
  9. Takada, S.; Nagato, Y.; Yamamura, M. Effect of cyclic polylactates on tumor cells and tumor bearing mice. Biochem. Mol. Biol. Int. 1997, 43, 9–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Prasad, A.V.; Biying, A.O.; Ling, W.Y.; Stubbs, L.P.; Zhu, Y. Synthesis and new application of green and recyclable cyclic poly(l-lactide)-clay hybrid. J. Polym. Sci. Part A Polym. Chem. 2013, 51, 4167–4174. [Google Scholar] [CrossRef]
  11. Kato, H.; Naganushi, Y. Preparation of l-Lactic Acid Condensation Products for Cancer Treatment. JPH06336427A, 6 December 1994. [Google Scholar]
  12. Osaka, I.; Watanabe, M.; Takama, M.; Murakami, M.; Arakawa, R. Characterization of linear and cyclic polylactic acids and their solvolysis products by electrospray ionization mass spectrometry. J. Mass Spectrom. 2006, 41, 1369–1377. [Google Scholar] [CrossRef] [PubMed]
  13. Osaka, I.; Yoshimoto, A.; Watanabe, M.; Takama, M.; Murakami, M.; Kawasaki, H.; Arakawa, R. Quantitative determination of cyclic polylactic acid oligomers in serum by direct injection liquid chromatography tandem mass spectrometry. J. Chromatogr. B 2008, 870, 247–250. [Google Scholar] [CrossRef] [PubMed]
  14. Sugai, N.; Yamamoto, T.; Tezuka, Y. Synthesis of orientationally isomeric cyclic stereoblock polylactides with head-to-head and head-to-tail linkages of the enantiomeric segments. ACS Macro Lett. 2012, 1, 902–906. [Google Scholar] [CrossRef]
  15. Josse, T.; De Winter, J.; Dubois, P.; Coulembier, O.; Gerbaux, P.; Memboeuf, A. A tandem mass spectrometry-based method to assess the architectural purity of synthetic polymers: A case of a cyclic polylactide obtained by click chemistry. Polym. Chem. 2015, 6, 64–69. [Google Scholar] [CrossRef]
  16. Stanford, M.J.; Pflughaupt, R.L.; Dove, A.P. Synthesis of stereoregular cyclic poly(lactide)s via “thiol-ene” click chemistry. Macromolecules 2010, 43, 6538–6541. [Google Scholar] [CrossRef]
  17. Culkin, D.A.; Jeong, W.H.; Csihony, S.; Gomez, E.D.; Balsara, N.R.; Hedrick, J.L.; Waymouth, R.M. Zwitterionic polymerization of lactide to cyclic poly(lactide) by using N-heterocyclic carbene organocatalysts. Angew. Chem. Int. Ed. 2007, 46, 2627–2630. [Google Scholar] [CrossRef] [PubMed]
  18. Jeong, W.; Shin, E.J.; Culkin, D.A.; Hedrick, J.L.; Waymouth, R.M. Zwitterionic polymerization: A kinetic strategy for the controlled synthesis of cyclic polylactide. J. Am. Chem. Soc. 2009, 131, 4884–4891. [Google Scholar] [CrossRef] [PubMed]
  19. Shin, E.J.; Jones, A.E.; Waymouth, R.M. Stereocomplexation in cyclic and linear polylactide blends. Macromolecules 2012, 45, 595–598. [Google Scholar] [CrossRef]
  20. Prasad, A.V.; Stubbs, L.P.; Zhun, M.; Zhu, Y.H. Zwitterionic ring opening polymerization of lactide by metal free catalysts: Production of cyclic polymers. J. Appl. Polym. Sci. 2012, 123, 1568–1575. [Google Scholar] [CrossRef]
  21. Kricheldorf, H.R.; Lomadze, N.; Schwarz, G. Cyclic polylactides by imidazole-catalyzed polymerization of l-lactide. Macromolecules 2008, 41, 7812–7816. [Google Scholar] [CrossRef]
  22. Brown, H.A.; De Crisci, A.G.; Hedrick, J.L.; Waymouth, R.M. Amidine-mediated zwitterionic polymerization of lactide. ACS Macro Lett. 2012, 1, 1113–1115. [Google Scholar] [CrossRef]
  23. Zhang, X.Y.; Waymouth, R.M. Zwitterionic ring opening polymerization with isothioureas. ACS Macro Lett. 2014, 3, 1024–1028. [Google Scholar] [CrossRef]
  24. Yao, L.; Wang, L.; Pan, X.; Tang, N.; Wu, J. Synthesis, characterization and catalytic activity of salen-(sodium)2 and (salen)2-lanthanum-sodium complexes. Inorg. Chim. Acta 2011, 373, 219–225. [Google Scholar] [CrossRef]
  25. Weil, J.; Mathers, R.T.; Getzler, Y.D.Y.L. Lactide cyclopolymerization by an alumatrane-inspired catalyst. Macromolecules 2012, 45, 1118–1121. [Google Scholar] [CrossRef]
  26. Piedra-Arroni, E.; Ladavière, C.; Amgoune, A.; Bourissou, D. Ring-opening polymerization with Zn(C6F5)2-based lewis pairs: Original and efficient approach to cyclic polyesters. J. Am. Chem. Soc. 2013, 135, 13306–13309. [Google Scholar] [CrossRef] [PubMed]
  27. Piromjitpong, P.; Ratanapanee, P.; Thumrongpatanaraks, W.; Kongsaeree, P.; Phomphrai, K. Synthesis of cyclic polylactide catalysed by bis(salicylaldiminato)tin(II) complexes. Dalton Trans. 2012, 41, 12704–12710. [Google Scholar] [CrossRef] [PubMed]
  28. Wongmahasirikun, P.; Prom-On, P.; Sangtrirutnugul, P.; Kongsaeree, P.; Phomphrai, K. Synthesis of cyclic polyesters: Effects of alkoxy side chains in salicylaldiminato tin(II) complexes. Dalton Trans. 2015, 44, 12357–12364. [Google Scholar] [CrossRef] [PubMed]
  29. Kricheldorf, H.R.; Weidner, S.M.; Scheliga, F. Cyclic poly(l-lactide)s via ring-expansion polymerizations catalysed by 2,2-dibutyl-2-stanna-1,3-dithiolane. Polym. Chem. 2017, 8, 1589–1596. [Google Scholar] [CrossRef]
  30. Farah, S.; Anderson, D.G.; Langer, R. Physical and mechanical properties of PLA, and their functions in widespread applications—A comprehensive review. Adv. Drug Deliv. Rev. 2016, 107, 367–392. [Google Scholar] [CrossRef] [PubMed]
  31. Martin, O.; Avérous, L. Poly(lactic acid): Plasticization and properties of biodegradable multiphase systems. Polymer 2001, 42, 6209–6219. [Google Scholar] [CrossRef]
  32. Mase, N.; Nakaya, N. Organocatalytic synthesis of highly pure linear- and cyclic-polylactide in supercritical carbon dioxide. In Proceedings of the Symposium on Molecular Chirality Asia 2016, Osaka, Japan, 22 April 2016. No. PB-33. [Google Scholar]
  33. Pratt, R.C.; Lohmeijer, B.G.G.; Long, D.A.; Lundberg, P.N.P.; Dove, A.P.; Li, H.; Wade, C.G.; Waymouth, R.M.; Hedrick, J.L. Exploration, optimization, and application of supramolecular thiourea-amine catalysts for the synthesis of lactide (co)polymers. Macromolecules 2006, 39, 7863–7871. [Google Scholar] [CrossRef]
  34. Wesch, A.; Dahmen, N.; Ebert, K.H. Measuring the static dielectric constants of pure carbon dioxide and carbon dioxide mixed with ethanol and toluene at elevated pressures. Ber. Bunsengesellsch. Phys. Chem. 1996, 100, 1368–1371. [Google Scholar] [CrossRef]
  35. Moriyoshi, T.; Kita, T.; Uosaki, Y. Static relative permittivity of carbon dioxide and nitrous oxide up to 30 MPa. Ber. Bunsengesellsch. Phys. Chem. 1993, 97, 589–596. [Google Scholar] [CrossRef]
  36. Tsuji, H. Poly(lactide) stereocomplexes: Formation, structure, properties, degradation, and applications. Macromol. Biosci. 2005, 5, 569–597. [Google Scholar] [CrossRef] [PubMed]
  37. Ikada, Y.; Jamshidi, K.; Tsuji, H.; Hyon, S.H. Stereocomplex formation between enantiomeric poly(lactides). Macromolecules 1987, 20, 904–906. [Google Scholar] [CrossRef]
Scheme 1. Synthetic strategies of cyclic polylactide (cPLA).
Scheme 1. Synthetic strategies of cyclic polylactide (cPLA).
Polymers 10 00713 sch001
Scheme 2. ROP of lactide, which creates PLA and cPLA.
Scheme 2. ROP of lactide, which creates PLA and cPLA.
Polymers 10 00713 sch002
Scheme 3. Competitive reaction between polymerization (and hence, cyclization) and epimerization.
Scheme 3. Competitive reaction between polymerization (and hence, cyclization) and epimerization.
Polymers 10 00713 sch003
Figure 1. Activation of L-lactide monomer by a thiourea additive.
Figure 1. Activation of L-lactide monomer by a thiourea additive.
Polymers 10 00713 g001
Figure 2. DSC measurements of the stereocomplex of cPLA (Table 2, entry 1).
Figure 2. DSC measurements of the stereocomplex of cPLA (Table 2, entry 1).
Polymers 10 00713 g002
Figure 3. Effect of stereochemical purity on the intermolecular interactions in the stereocomplex formation.
Figure 3. Effect of stereochemical purity on the intermolecular interactions in the stereocomplex formation.
Polymers 10 00713 g003
Figure 4. GPC traces of cPLA with different ring sizes and images of sc-cPLA.
Figure 4. GPC traces of cPLA with different ring sizes and images of sc-cPLA.
Polymers 10 00713 g004
Table 1. Organocatalytic cPLA synthesis in supercritical carbon dioxide (scCO2).
Table 1. Organocatalytic cPLA synthesis in supercritical carbon dioxide (scCO2).
Polymers 10 00713 i001
EntrySolventTime (h)Mn (a)PDI (b)ee (%) (c)
1CHCl312012,7003.1139.0
2scCO2555001.4090.5
3 (d)scCO22.511,0001.6093.5
4 (e)scCO2560001.3191.0
5 (d,e)scCO22.568001.2097.0
(a) Determined by GPC analysis using a polystyrene standard. (b) Polydispersity index. (c) Determined by HPLC analysis after hydrolysis of the polymer to lactic acid. (d) 1-(3,5-Bis(trifluoromethyl)phenyl)-3-cyclohexyl thiourea (1 mol%) was added. (e) d-Lactide (ent-1) was used.
Table 2. DSC measurements of the stereocomplex of cPLA.
Table 2. DSC measurements of the stereocomplex of cPLA.
EntrycPLAMnPDIEe (%)Tm (°C)Tm (sc-cPLA) (°C)
1cPDLA68001.2097.0152212
cPLLA11,0001.6093.5149
2cPDLA60001.3191.0143207
cPLLA55001.4090.5145
3 (a)cPDLA26,0001.40(0.81) (b)132179
cPLLA30,0001.30(0.83) (b)135
(a) Data reported in reference 19. (b) Fraction of isotactic tetrads.
Table 3. Stereocomplex of PLA with different ring sizes.
Table 3. Stereocomplex of PLA with different ring sizes.
EntrycPLAMnPDITm (sc-cPLA)
1cPDLA79001.18190
cPLLA26001.20
2cPDLA51001.23200
cPLLA74001.23
3cPDLA60001.31207
cPLLA55001.40

Share and Cite

MDPI and ACS Style

Mase, N.; Moniruzzaman; Yamamoto, S.; Nakaya, Y.; Sato, K.; Narumi, T. Organocatalytic Stereoselective Cyclic Polylactide Synthesis in Supercritical Carbon Dioxide under Plasticizing Conditions. Polymers 2018, 10, 713. https://doi.org/10.3390/polym10070713

AMA Style

Mase N, Moniruzzaman, Yamamoto S, Nakaya Y, Sato K, Narumi T. Organocatalytic Stereoselective Cyclic Polylactide Synthesis in Supercritical Carbon Dioxide under Plasticizing Conditions. Polymers. 2018; 10(7):713. https://doi.org/10.3390/polym10070713

Chicago/Turabian Style

Mase, Nobuyuki, Moniruzzaman, Shoji Yamamoto, Yoshitaka Nakaya, Kohei Sato, and Tetsuo Narumi. 2018. "Organocatalytic Stereoselective Cyclic Polylactide Synthesis in Supercritical Carbon Dioxide under Plasticizing Conditions" Polymers 10, no. 7: 713. https://doi.org/10.3390/polym10070713

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