Next Article in Journal
Pseudomonas sp. COW3 Produces New Bananamide-Type Cyclic Lipopeptides with Antimicrobial Activity against Pythium myriotylum and Pyricularia oryzae
Previous Article in Journal
The Influence of Inulin on the Retention of Polyphenolic Compounds during the Drying of Blackcurrant Juice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Application of Diethylzinc/Propyl Gallate Catalytic System for Ring-Opening Copolymerization of rac-Lactide and ε-Caprolactone

1
Department of Biomaterials Chemistry, Chair of Analytical Chemistry and Biomaterials, Faculty of Pharmacy, Medical University of Warsaw, 1 Banacha St., Warsaw 02-097, Poland
2
Department of Environmental Health Sciences, Faculty of Pharmacy, Medical University of Warsaw, 1 Banacha St., Warsaw 02-097, Poland
*
Author to whom correspondence should be addressed.
Molecules 2019, 24(22), 4168; https://doi.org/10.3390/molecules24224168
Submission received: 14 October 2019 / Revised: 15 November 2019 / Accepted: 15 November 2019 / Published: 17 November 2019

Abstract

:
Biodegradable polyesters gain significant attention because of their wide potential biomedical applications. The ring-opening polymerization method is widely used to obtain such polymers, due to high yields and advantageous properties of the obtained material. The preparation of new, effective, and bio-safe catalytic systems for the synthesis of biomedical polymers is one of the main directions of the research in modern medical chemistry. The new diethylzinc/propyl gallate catalytic system was first used in the copolymerization of ε-caprolactone and rac-lactide. In this paper, the activity of the new zinc-based catalytic system in the copolymerization of cyclic esters depending on the reaction conditions was described. The microstructure analysis of the obtained copolyesters and their toxicity studies were performed. Resulted copolyesters were characterized by low toxicity, moderate dispersity (1.19–1.71), varying randomness degree (0.18–0.83), and average molar mass (5300–9800 Da).

1. Introduction

In recent years, biodegradable and bioresorbable homo- and copolyesters based on cyclic monomers: rac-lactide (rac-LA), L,L-lactide (LLA), ε-caprolactone (CL), glycolide (GL), and trimethylene carbonate (TMC) were widely tested for their potential use in biomedical applications [1,2]. These applications include sutures, drug delivery systems (e.g., drug nano- or microcarriers or macromolecular prodrugs), implants [3] and tissue engineering [4].
Although Poly(ε-caprolactone) (PCL) exhibits favorable biocompatibility and mechanical properties, it biodegrades in vivo very slowly—from a few months to several years [5]. Polylactide (PLA) displays variable biodegradation time, ranging from several weeks to about two years [6]. The exact biodegradation rate of PLA depends on the polymer’s average molecular weight and its dispersity (Đ), crystallinity, microstructure, etc. Co- or terpolymers of CL, rac-LA, LLA, GL, or TMC characterized by various microstructures allow obtaining the assumed time of polymer biodegradation and high controlled release of active substances from the polymeric carrier [7].
Biomedical polyesters can be mainly obtained by two methods: ring-opening polymerization (ROP) or polycondensation. The ROP process exhibits numerous advantages over classic polycondensation, such as higher product yield, proper average molecular weight, and its Đ [8]. Most of the biomedical polyesters are synthesized in the presence of the tin (II) 2-ethylhexanoate (SnOct2) initiator, which was approved by the United States Food and Drugs Administration (FDA) as a food additive [5]. However, tin derivatives are virtually irremovable from the final biomaterial. This may present a serious health risk due to their toxicity, especially toward juveniles [9,10,11,12]. Furthermore, possible, industrial application of large quantities of this compound also raises environmental concerns [13]. Several alternative ROP initiators/catalysts for the polyester synthesis were investigated over the years, such as ionic [14], coordinating [15], and enzymatic [16]. A very interesting group of coordinating initiators/catalysts are zinc [17], calcium [9] and zirconium (IV) acetylacetonates [18], which are characterized by high activity and biocompatibility. Some of the other examples of ROP catalysts are group 3 metal complexes [19,20], bismuth (III) analogs of SnOct2 [21], organic complexes of aluminum [22,23], alkoxyl titanium [24], or alkyl/alkoxyl tin compounds [25]. Metal-organic ROP catalysts seem to be the most favorable due to their selectivity. They allow obtaining—depending on the reaction conditions—polymeric products with variable microstructure. Since the microstructure of the polymer determines its hydrolytic stability, it is possible to synthesize polyesters with varying biodegradation rates by modifying polymerization conditions, such as reaction time, temperature, or used catalyst [7,26].
Zinc salts, such as lactate [13], aceturate, or L-prolinate [27] were found to be active catalysts for ROP of LLA. Organic zinc complexes were also successfully used as initiators/catalysts for the homopolymerization of LA [28,29,30,31], CL [32,33] or copolymerization of CL, and LA [34,35,36]. Zinc-based catalysts may be a viable alternative to SnOct2 due to low toxicity and programmable polymer chain microstructure. As our previous studies have shown [37,38,39], diethylzinc/propyl gallate (ZnEt2/PGA) was an efficient catalytic system for the homopolymerization of rac-LA and CL, yielding polymers with varying microstructure (depending on the reaction conditions). It is worth mentioning that propyl gallate (PGA) is used as an additive (E310) in pharmaceuticals, cosmetics, and food due to its antioxidant properties [40]. Most importantly, resulting polymers were found to be non-toxic [39]. Consequently, in this paper, the results of copolymerization of rac-LA and CL in the presence of these efficient and non-toxic ZnEt2/PGA catalytic systems have been presented.

2. Results and Discussion

The copolymerization of rac-LA and CL (Figure 1) was carried out for different monomer molar ratios, time, and temperature. The catalytic system was synthesized by the reaction of PGA with 3 molar equivalent of ZnEt2. PGA has been used as a bio-safe co-initiator of the ROP process. However, it is worth to mention that we were unable to establish the exact structure of the applied catalytic system. The attempts of its crystallization with numerous solvents led to the precipitating of the amorphous solid. Based on the reaction stoichiometry, however, we assume that the present phenolic O-Zn-Et groups may act as ROP initiators.
In our previous studies, gallic acid (GA) and PGA were used as co-initiators of the homopolymerization of rac-LA [38] and CL [37]. In those studies, toluene, tetrahydrofuran, or dichloromethane were investigated as a reaction medium. It was found that GA exhibited a tendency to promote macrolactonization during the ROP process. In addition, toluene has proved to be the optimal reaction medium of the ROP process [38]. As in vitro degradation of rac-LA is much faster than LLA [41], the former monomer was used to reduce polymer degradation time. Considering these facts, in current work, we decided to use ZnEt2/PGA catalytic system and toluene as a reaction medium in the copolymerization process of rac-LA and CL.
The effect of the temperature and reaction time on the yield of the product, the products average molecular weight, as well as microstructure of the synthesized materials were investigated (Table 1).

2.1. Polymers Characterization

It is well known that the polyester microstructure influences the rate of the polymer biodegradation and kinetics of drug release. The microstructure of the polymer chain depends on the intra- or intermolecular transesterification and stereoselectivity process. Transesterification is a well-known phenomenon causing polyester sequence redistribution. In short, during the ROP of LA, lactyl units (L) should appear in pairs in the final polymer chain. However, transesterification may lead to the appearance of abnormal sequences, defined as type-II transesterification [42].
The microstructure of the obtained copolymers was established by 13C NMR studies. Type-II transesterification ratios, average lengths of monomer blocks, and randomness degrees were measured by NMR signal area calculations [43,44].
As is shown in Table 1, the yield of the ROP process depended on the monomer/catalytic system’s ratio, temperature, and reaction time.
In most cases, the CL conversion was faster and higher than that of LA. The extension of the reaction time up to 16 h at 80 °C did not significantly affect it. However, it was necessary to carry out polymerization for up to 48 h at 60 °C in order to achieve satisfying conversion. The average length of lactydyl units was decreased with increasing of the reaction time, that indicates the occurrence of transesterification.
Increasing of the ROP temperature generally increased the product yield. Interestingly, it seemed that the combination of a high catalyst concentration (6/100 and 8/100), high temperature (80 °C), and long reaction time (24 and 48 h) negatively affected the reaction yield. It is possible that a high concentration of the active centers on the catalyst led to the degradation of already formed polyester chains. Oligomers could have been removed from the final product during the purification step (during precipitation from cold methanol).
Generally, Mn increased with the increasing of the reaction time. A positive correlation between Mn and the reaction temperature was also observed. Đ was found to increase with increasing the reaction time, temperature, and lactydyl unit concentration in the obtained polyester.
Type-II transesterification was present in most obtained copolymers, as a signal from the CapLCap triad (lactyl-caproyl-lactyl units) around 170.8 ppm with varying intensity was observed (see Figure 2). The occurrence of this phenomenon is unsurprising, as Zn initiators tend to promote it in higher temperatures [42]. The Zn/monomer feed ratio affected the mentioned parameter, namely 2/100 ratio was highly favored; 1/100 ratio was favored moderately, while 8/100 and 6/100 ratios was inhibited. As was expected, the increase in the polymerization temperature and reaction time increased the TII ratio. Some occurring deviations might have resulted from the 13C NMR measurements low accuracy. The degree of randomness increased (or supposedly reached plateau) with reaction time increase. A high catalyst load seemed to reduce this parameter.
Reactivity ratios: r1 (M1 = rac-LA) and r2 (M2 = CL) were roughly estimated based on the monomer feed in the reaction and composition of the resulted copolymers [41,45,46]. It was found that r1 was higher than r2; these values were in the range of 3.13–20.6 and 1.12–11.2, respectively. Both r1 and r2 were higher than 1, which indicated non-azeotropic, non-ideal copolymerization.
In summary, the 2/100 Zn/monomer feed ratio was found to be unfavorable due to its tendency to promote exceptionally high TII: from 39% for entry 17 up to 100% for entry 18. The 6/100 and 8/100 monomer feed ratios were advantageous as a satisfactory compromise between high yield and low TII. The 60 °C reaction temperature was beneficial considering low TII, but also negatively affected yield, e.g., 20% yield for entry 19 and 25% yield for entry 24. However, this effect can be largely negated by extending the reaction time to 48 h: 46% yield for entry 20, 52% yield for entry 25, and 46% yield for entry 30. In connection with the above, the optimal reaction time is 16 h (48 h for the reactions performed at 60 °C). When the ROP process has been carried out in longer reaction time, TII and Đ values were higher.
Considering optimal reaction conditions, entries 21 (CL/rac-LA molar ratio 1/1, Zn/monomers molar ratio 6/100, T = 80 °C, reaction time: 16 h, 81% yield, Mn = 8400 Da, Đ = 1.38, TII = 6.5%), 26 (CL/rac-LA molar ratio 2/1, Zn/monomers molar ratio 8/100, T = 80°C, reaction time: 16 h, 83% yield, Mn = 8600 Da, Đ = 1.38, TII = 6.1%) and 31 (CL/rac-LA molar ratio 1/2, Zn/monomers molar ratio 8/100, T = 80°C, reaction time: 16 h, 73% yield, Mn = 8000 Da, Đ = 1.54, TII = 3.0%) were favorable, as they combined high products yield and relatively low TII data.
The thermal properties of the obtained copolymers were also investigated by DSC. The results are shown in Figure 3. The melting temperature (Tm) and glass transition temperature (Tg) of the selected copolymers and references were determined and listed in Table 2. PCL as the reference was showing Tm = 69.4 °C and Tg = −60.0, whereas for PLA Tg = 53.4 °C. As is evidenced in Figure 3, Tm was not detected for PLA as the reference. The Tm value for the synthesized copolymers entitled as entry 9 (Figure 3) was detected at 51.6 °C. However, for the synthesized samples, the entitled as entries 18 and 32, next to the Tm values of 52.5 °C and 56.4 °C, respectively, extended peaks have emerged, centered as 164.9 and 181.8 °C, that are characteristic for poly(d,l-lactide) (PDLA) [47].
Significant changes in Tg values were, however, observed for the synthesized copolymers compared to the references (Table 2). They were −5.32 °C, −9.74 °C, and −32.4 °C for entries 9, 18, and 32, respectively. Higher Tg is generally a favorable property since compounds with high Tg have a reduced ability to recrystallize at a given temperature, compared to those that have a lower Tg. These results were attributed to the copolymerization process, which decreases the mobility of copolymer chains, a fact supported in the literature [48].

2.2. Toxicity Studies

In the Spirotox test, it was found that none of the tested samples had been toxic to the protozoan S. Ambiguum. Some of the samples exhibited low toxicity toward luminescent bacteria Allivibrio Fischeri (The percent of a toxic effect (PE) from 21 to 34, Table 3). However, regarding the standard deviation, only entries 5 and 9 could be unambiguously considered slightly toxic. The toxic effect of these samples could have been caused by insufficient removal of catalytic system traces from the final polymer. In the umu-test with S. Typhimurium, all of the tested samples showed no cyto- and genotoxic potential with or without metabolic activation (IR < 1.5, Table 4).

3. Materials and Methods

3.1. Materials

All chemicals were stored in an inert atmosphere of dry argon. Chloroform-d for NMR (99.8 atom % D, with 0.1 v/v% TMS, stabilized with silver), and toluene (99.85%, Extra Dry, over Molecular Sieves, AcroSeal®) were purchased from Acros Organics (Geel, Belgium) and used as received. Rac-LA (rac-lactide, 3,6-dimethyl-1,4-dioxane-2,5-dione), ZnEt2 solution (15% diethylzinc in toluene), PGA (propyl gallate, 98%+), poly(d,l-lactide) (viscosity 0.68 dL/g), and polycaprolactone (average Mw ca. 14,000; average Mn ca. 10,000) were purchased from Sigma-Aldrich Co. (Poznan, Poland) and used as received. CL (ε-caprolactone, 6-caprolactone, 99%+) was purchased from Sigma-Aldrich Co. (Poznan, Poland) and stored over 5 A° molecular sieves. Dichloromethane (pure, 99%) and methanol (pure, 99.9%) were purchased from Chempur (Piekary Śląskie, Poland) and distilled before use.

3.2. Methods

3.2.1. Synthesis of the Catalytic System

The catalytic system was prepared using air-free techniques. PGA (128 mg, 0.603 mmol) and magnetic stir bar were placed in a 25 mL one-necked round-bottom flask. The high vacuum was applied for 1 h. After that, the flask was evacuated and backfilled with argon three times, and 8.37 mL of dry toluene was added. The flask was sealed, and the reaction was stirred overnight. Next, the reaction mixture was cooled to 0 °C, and 1.63 mL of 15% ZnEt2 (1.81 mmol, three equivalents) solution was added. The reaction was stirred further for 2 h with cooling, after that, the mixture was allowed to warm to room temperature. The prepared solution contains ca. 0.18 mmol of Zn in 1 mL.

3.2.2. Polymerization Procedure

Copolymerization was carried out in a vacuum-dried 10 mL glass tubes with the joint. A mixture of monomers in appropriate ratios (18 mmol total) was placed in a reaction vessel, which was evacuated and backfilled with argon three times. After that, 5 ml of dry toluene and 1 mL of the catalytic system (0.18 mmol of Zn) was added. Glass tube was sealed, shaken, placed in the preheated oil bath with thermostat and kept at the appropriate temperature and for the required amount of time. Then, the reaction mixture was removed from the oil bath, allowed to cool to room temperature, dissolved in dichloromethane, and washed two times with 5% HCl solution and one time with distilled water. The product was precipitated by adding a concentrated dichloromethane solution of polymer to cold (2–8 °C) methanol (ca. 10 times volume of dichloromethane solution). The solvent mixture was decanted, the precipitate was washed with cold methanol, dried in the air overnight, and in vacuum for ca. 48 h.

3.2.3. Measurements

The polymerization products were characterized in a deuterium chloroform solution by means of 1H- (300 MHz) and 13C-NMR (75 MHz) spectroscopy (Varian, LabX, Midland, ON, Canada).
Relative average molecular mass and molecular mass distribution were determined by Gel permeation chromatography (GPC). GPC instrument (GPC Max + TDA 305, Viscotek, Malvern, UK) was equipped with Jordi DVB Mixed Bed columns (one guard and two analytical) at 30 °C in CH2Cl2 (HPLC grade, Sigma-Aldrich, Poznań, Poland), at a flow rate of 1 mL/min with RI detection and calibration based on narrow PS standards (ReadyCal Set, Fluka, Poznan, Poland). The results were processed with OmniSEC software (version 4.7). Mn values were not corrected.
A differential scanning calorimetry technique (DSC, TA Instruments, New Castle, USA) was used to analyze the thermal transitions of the polymers. The DSC data were obtained between −80 and 200 °C using the Q200 apparatus. The sample was heated and cooled at a rate of 10 °C min−1. An empty Tzero aluminum pan was used as the reference.

3.2.4. Spectroscopic Data

Typical NMR shifts ranges were as follows: 1H NMR δ 5.19–5.10 (m, C(O)CH(CH3)O), 4.11 (br s, CH2CH2OC(O)CH(CH3)), 4.04 (t, CH2CH2OC(O)CH2CH2), 2.37 (br s, CH(CH3)OC(O)CH2CH2CH2), 2.28 (t, CH2CH2OC(O)CH2CH2CH2), 1.64–1.36 (br m, CH(CH3), C(O)CH2CH2CH2CH2CH2O) ppm; 13C NMR δ ca. 172.8–172.0 (C(O)CH2CH2CH2CH2CH2O), 171.0–170.8 (CH2CH2OC(O)CH(CH3)OC(O)CH2CH2), 170.3–168.7 (OC(O)CH(CH3)O), 69.3–67.8 (OC(O)CH(CH3)O), 64.8–64.4 (C(O)CH2CH2CH2CH2CH2O), 33.5–32.9 (C(O)CH2CH2CH2CH2CH2O), 27.9–27.5 (C(O)CH2CH2CH2CH2CH2O), 25.0–24.5 (C(O)CH2CH2CH2CH2CH2O), 24.2–23.7 (C(O)CH2CH2CH2CH2CH2O), 16.7–16.3 (OC(O)CH(CH3)O) ppm.

3.3. Toxicity Studies

3.3.1. Microtox and Spirotox Tests

Five milligrams of the copolymer was placed in the glass tube with 5 mL of Tyrode’s solution (Spirotox) or 2% NaCl (Microtox). The tubes were incubated at 37 °C for 24 h with shaking.
Microtox: A short-term bioassay with the luminescent bacteria Allivibrio fischeri (previously known as Vibrio fischeri). The procedure was based on the International Organization for Standardization (ISO) standard [49]. Shortly, the tested and the control (2% NaCl) samples were incubated with the bacteria at 15 °C for 15 min, and the luminescence was measured in the Microtox M500 luminometer. Then the percent of inhibition of the luminescence was calculated in comparison to the control.
Spirotox: A short-term bioassay with the ciliated protozoan Spirostomum ambiguum. The test was performed according to the ISO 11348-3:2007 standard protocol [50]. Shortly, the tested and the control (Tyrod solution) samples were incubated with the protozoans at 25 °C for 24 h, and the sublethal (deformations) and lethal effects are observed with the dissection microscope (magnification 10×). Then the percent of affected protozoans was calculated for the sample in comparison to the control.

3.3.2. Umu—Test

Umu—a test that detects the induction of the SOS system in the strain S. typhimurium TA1535/pSK1002. SOS system is the bacterial response to the DNA-damaging agents. The test strain is genetically modified—the umuC gene activity is linked to the synthesis of β-galactosidase, while other DNA regions responsible for this enzyme synthesis were deleted. Therefore β-galactosidase activity strictly depends on the SOS system induction level and the genotoxic activity of the tested compound [51]. The enzyme converts colorless substrate (ortho-nitrophenyl-β-galactoside) into the yellow product, which can be quantified colorimetrically at 420 nm. Additionally, the bacteria growth (G) is evaluated by measurement of an optical density to determine the cytotoxicity of tested samples. The genotoxic potential of the sample is presented as the Induction Ratio (IR)—the β-galactosidase activity ratio of the tested sample in comparison to the negative control. Samples with IR ≥ 1.5 are considered as genotoxic.
In the present study, the umu-test was carried out in the micro-plate variant according to the ISO 13829 guideline, with and without metabolic activation (S9 liver fraction from male Sprague-Dawley rats treated five days before the isolation with a single dose of 500 mg/kg body weight of Aroclor 1254 in soya oil) [49]. Deionized sterile water was used as a negative control, 2-aminoanthracene and 4-nitroquinoline N-oxide were used as positive controls, and phosphate-buffered saline (PBS from Gibco, Thermo Fisher Scientific, Darmstadt, Germany) as solvent control. All tested samples were incubated in PBS - 1 mg/mL for 24 h, 37 °C, with shaking. Before the assay, all extracts were sterilized by filtration (0.20 µm). All samples were tested in two-fold dilution series (four concentrations, the highest concentration of 0.66 mg/mL).

4. Conclusions

In summary, the developed, non-toxic catalytic system enables the synthesis of rac-LA/CL copolymers. The obtained products were characterized by a wide range of Mn (from 5400 to 9800 Da), adequate for biomedical applications. It is worth to note that the copolyesters were found to be no cyto- nor genotoxic, and thus, they can be used in the drug formulation technology. Depending on the reaction conditions (monomer feed ratio, reaction temperature and time, catalyst feed), the produced copolyesters vary in their microstructure, thus can be used to obtain various drug delivery systems (middle- and short term), characterized by different drug release kinetics. We have also found that the optimal conditions for the copolymerization process were: 6/100 or 8/100 Zn/monomer ratio, 16 h reaction time, and 80°C reaction temperature. The copolymeric products obtained under these conditions were characterized by a high reaction yield and low type-II transesterification ratio.

Author Contributions

The contributions of the respective authors are as follows: R.W. and M.S. gave the concept of the article, interpreted the results and wrote the article, made discussion and conclusions. R.W. also performed chemical synthesis. E.O. has interpreted the results and wrote the article, made discussion and conclusions. R.F. performed and interpreted biological studies. All authors have contributed substantially to the work reported.

Funding

This work was supported by the research program Project Young Researcher FW23/PM1/17 of the Medical University of Warsaw.

Acknowledgments

The authors would like to thank Andrzej Plichta (Department of Chemistry and Polymer Technology, Warsaw Institute of Technology) for GPC analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dobrzyński, P.; Kasperczyk, J.; Jelonek, K.; Ryba, M.; Walski, M.; Bero, M. Application of the lithium and magnesium initiators for the synthesis of glycolide, lactide, and epsilon-caprolactone copolymers biocompatible with brain tissue. J. Biomed. Mater. Res. A 2006, 79, 865–873. [Google Scholar] [CrossRef] [PubMed]
  2. Dobrzynski, P.; Kasperczyk, J. Synthesis of biodegradable copolymers with low-toxicity zirconium compounds. IV. Copolymerization of glycolide with trimethylene carbonate and 2,2-dimethyltrimethylene carbonate: Microstructure analysis of copolymer chains by high-resolution nuclear magnetic resonance spectroscopy. J. Polym. Sci. Part Polym. Chem. 2006, 44, 98–114. [Google Scholar]
  3. Seyednejad, H.; Ghassemi, A.H.; van Nostrum, C.F.; Vermonden, T.; Hennink, W.E. Functional aliphatic polyesters for biomedical and pharmaceutical applications. J. Control. Release 2011, 152, 168–176. [Google Scholar] [CrossRef] [PubMed]
  4. Kenar, H.; Ozdogan, C.Y.; Dumlu, C.; Doger, E.; Kose, G.T.; Hasirci, V. Microfibrous scaffolds from poly(l-lactide-co-ε-caprolactone) blended with xeno-free collagen/hyaluronic acid for improvement of vascularization in tissue engineering applications. Mater. Sci. Eng. C 2019, 97, 31–44. [Google Scholar] [CrossRef] [PubMed]
  5. Labet, M.; Thielemans, W. Synthesis of polycaprolactone: A review. Chem. Soc. Rev. 2009, 38, 3484–3504. [Google Scholar] [CrossRef] [PubMed]
  6. Ahmed, J.; Varshney, S.K. Polylactides—Chemistry, Properties and Green Packaging Technology: A Review. Int. J. Food Prop. 2011, 14, 37–58. [Google Scholar] [CrossRef]
  7. Orchel, A.; Jelonek, K.; Kasperczyk, J.; Dobrzynski, P.; Marcinkowski, A.; Pamula, E.; Orchel, J.; Bielecki, I.; Kulczycka, A. The Influence of Chain Microstructure of Biodegradable Copolyesters Obtained with Low-Toxic Zirconium Initiator to In Vitro Biocompatibility. BioMed. Res. Int. 2013, 2013, 12. [Google Scholar] [CrossRef]
  8. Dechy-Cabaret, O.; Martin-Vaca, B.; Bourissou, D. Controlled Ring-Opening Polymerization of Lactide and Glycolide. Chem. Rev. 2004, 104, 6147–6176. [Google Scholar] [CrossRef]
  9. Dobrzyński, P.; Kasperczyk, J.; Bero, M. Application of Calcium Acetylacetonate to the Polymerization of Glycolide and Copolymerization of Glycolide with ε-Caprolactone and l-Lactide. Macromolecules 1999, 32, 4735–4737. [Google Scholar] [CrossRef]
  10. De Mattos, J.; Dantas, F.; Bezerra, R.; Bernardo-Filho, M.; Cabral-Neto, J.; Lage, C.; Leitão, A.; Caldeira-de-Araújo, A. Damage induced by stannous chloride in plasmid DNA. Toxicol. Lett. 2000, 116, 159–163. [Google Scholar] [CrossRef]
  11. Salánki, Y.; D’eri, Y.; Platokhin, A.; Sh.-Rózsa, K. The neurotoxicity of environmental pollutants: The effects of tin (Sn2+) on acetylcholine-induced currents in greater pond snail neurons. Neurosci. Behav. Physiol. 2000, 30, 63–73. [Google Scholar]
  12. Lewis, R.J. Sax’s Dangerous Properties of Industrial Materials, 8th ed.; Van Nostrand Reinhold: New York, NY, USA, 1992. [Google Scholar]
  13. Kricheldorf, H.R.; Kreiser-Saunders, I.; Damrau, D.-O. Resorbable initiators for polymerizations of lactones. Macromol. Symp. 1999, 144, 269–276. [Google Scholar] [CrossRef]
  14. Guillerm, B.; Lemaur, V.; Cornil, J.; Lazzaroni, R.; Dubois, P.; Coulembier, O. Ammonium betaines: Efficient ionic nucleophilic catalysts for the ring-opening polymerization of L-lactide and cyclic carbonates. Chem. Commun. 2014, 50, 10098–10101. [Google Scholar] [CrossRef] [PubMed]
  15. Chisholm, M.H. Concerning the ring-opening polymerization of lactide and cyclic esters by coordination metal catalysts. Pure Appl. Chem. 2010, 82, 1647–1662. [Google Scholar] [CrossRef]
  16. Piotrowska, U.; Sobczak, M. Enzymatic Polymerization of Cyclic Monomers in Ionic Liquids as a Prospective Synthesis Method for Polyesters Used in Drug Delivery Systems. Molecules 2014, 20, 1–23. [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 Polym. Chem. 2011, 49, 2504–2512. [Google Scholar] [CrossRef]
  18. Dobrzynski, P. Synthesis of biodegradable copolymers with low-toxicity zirconium compounds. II. Copolymerization of glycolide with ϵ-caprolactone initiated by zirconium(IV) acetylacetonate and zirconium(IV) chloride. J. Polym. Sci. Part Polym. Chem. 2002, 40, 1379–1394. [Google Scholar] [CrossRef]
  19. Amgoune, A.; Thomas, C.M.; Roisnel, T.; Carpentier, J.-F. Ring-Opening Polymerization of Lactide with Group 3 Metal Complexes Supported by Dianionic Alkoxy-Amino-Bisphenolate Ligands: Combining High Activity, Productivity, and Selectivity. Chem. Eur. J. 2006, 12, 169–179. [Google Scholar] [CrossRef]
  20. Abderramane, A.; Thomas, C.M.; Carpentier, J.-F. Controlled ring-opening polymerization of lactide by group 3 metal complexes. Pure Appl. Chem. 2007, 79, 2013. [Google Scholar]
  21. Kricheldorf, H.R.; Bornhorst, K.; Hachmann-Thiessen, H. Bismuth(III) n-Hexanoate and Tin(II) 2-Ethylhexanoate Initiated Copolymerizations of ε-Caprolactone and l-Lactide. Macromolecules 2005, 38, 5017–5024. [Google Scholar] [CrossRef]
  22. Castro-Osma, J.A.; Alonso-Moreno, C.; Márquez-Segovia, I.; Otero, A.; Lara-Sánchez, A.; Fernández-Baeza, J.; Rodríguez, A.M.; Sánchez-Barba, L.F.; García-Martínez, J.C. Synthesis, structural characterization and catalytic evaluation of the ring-opening polymerization of discrete five-coordinate alkyl aluminium complexes. Dalton Trans. 2013, 42, 9325–9337. [Google Scholar] [CrossRef] [PubMed]
  23. Bero, M.; Kasperczyk, J.; Adamus, G. Coordination polymerization of lactides, 3. Copolymerization of L,L-lactide and ε-caprolactone in the presence of initiators containing Zn and Al. Makromol. Chem. 1993, 194, 907–912. [Google Scholar] [CrossRef]
  24. Kim, Y.; Jnaneshwara, G.K.; Verkade, J.G. Titanium Alkoxides as Initiators for the Controlled Polymerization of Lactide. Inorg. Chem. 2003, 42, 1437–1447. [Google Scholar] [CrossRef]
  25. Stassin, F.; Jérôme, R. Polymerization of (L,L)-lactide and copolymerization with ϵ-caprolactone initiated by dibutyltin dimethoxide in supercritical carbon dioxide. J. Polym. Sci. Part Polym. Chem. 2005, 43, 2777–2789. [Google Scholar] [CrossRef]
  26. Kasperczyk, J.; Jelonek, K.; Dobrzyñski, P.; Jarz, B. The influence of copolymer chain microstructure on cyclosporine a (CyA) and Sirolimus prolonged and sustained release from PLA/TMC and PLA/PCL matrices. J. Control. Release 2006, 116, e5–e6. [Google Scholar] [CrossRef] [PubMed]
  27. Kricheldorf, H.R.; Damrau, D.-O. Polylactones, 43. Polymerization of L-lactide catalyzed by zinc amino acid salts. Macromol. Chem. Phys. 1998, 199, 1747–1752. [Google Scholar] [CrossRef]
  28. González, D.M.; Cisterna, J.; Brito, I.; Roisnel, T.; Hamon, J.-R.; Manzur, C. Binuclear Schiff-base zinc(II) complexes: Synthesis, crystal structures and reactivity toward ring opening polymerization of rac-lactide. Polyhedron 2019, 162, 91–99. [Google Scholar] [CrossRef]
  29. McKeown, P.; McCormick, S.N.; Mahon, M.F.; Jones, M.D. Highly active Mg(ii) and Zn(ii) complexes for the ring opening polymerisation of lactide. Polym. Chem. 2018, 9, 5339–5347. [Google Scholar] [CrossRef]
  30. Munzeiwa, W.A.; Nyamori, V.O.; Omondi, B. N,O-Amino-phenolate Mg(II) and Zn(II) Schiff base complexes: Synthesis and application in ring-opening polymerization of ε-caprolactone and lactides. Inorg. Chim. Acta 2019, 487, 264–274. [Google Scholar] [CrossRef]
  31. Wang, B.; Wei, Y.; Li, Z.-J.; Pan, L.; Li, Y.-S. From Zn(C6F5)2 to ZnEt2-based Lewis Pairs: Significantly Improved Catalytic Activity and Monomer Adaptability for the Ring-opening Polymerization of Lactones. ChemCatChem 2018, 10, 5287–5296. [Google Scholar] [CrossRef]
  32. Hu, Q.; Jie, S.; Braunstein, P.; Li, B.-G. Highly active tridentate amino-phenol zinc complexes for the catalytic ring-opening polymerization of ε-caprolactone. J. Organomet. Chem. 2019, 882, 1–9. [Google Scholar] [CrossRef]
  33. Posada, F.A.; Macías, A.M.; Movilla, S.; Miscione, P.G.; Pérez, D.L.; Hurtado, J.J. Polymers of ε-Caprolactone Using New Copper(II) and Zinc(II) Complexes as Initiators: Synthesis, Characterization and X-Ray Crystal Structures. Polymers 2018, 10, 1239. [Google Scholar] [CrossRef] [PubMed]
  34. Honrado, M.; Otero, A.; Fernández-Baeza, J.; Sánchez-Barba, L.F.; Garcés, A.; Lara-Sánchez, A.; Rodríguez, A.M. Copolymerization of Cyclic Esters Controlled by Chiral NNO-Scorpionate Zinc Initiators. Organometallics 2016, 35, 189–197. [Google Scholar] [CrossRef]
  35. Darensbourg, D.J.; Karroonnirun, O. Ring-Opening Polymerization of l-Lactide and ε-Caprolactone Utilizing Biocompatible Zinc Catalysts. Random Copolymerization of l-Lactide and ε-Caprolactone. Macromolecules 2010, 43, 8880–8886. [Google Scholar] [CrossRef]
  36. D’Auria, I.; Lamberti, M.; Mazzeo, M.; Milione, S.; Roviello, G.; Pellecchia, C. Coordination Chemistry and Reactivity of Zinc Complexes Supported by a Phosphido Pincer Ligand. Chem. Eur. J. 2012, 18, 2349–2360. [Google Scholar] [CrossRef]
  37. Żółtowska, K.; Sobczak, M.; Olędzka, E. Novel Zinc-Catalytic Systems for Ring-Opening Polymerization of ε-Caprolactone. Molecules 2015, 20, 2816–2827. [Google Scholar] [CrossRef] [Green Version]
  38. Żółtowska, K.; Piotrowska, U.; Oledzka, E.; Sobczak, M. Efficient Diethylzinc/Gallic Acid and Diethylzinc/Gallic Acid Ester Catalytic Systems for the Ring-Opening Polymerization of rac-Lactide. Molecules 2015, 20, 21909–21923. [Google Scholar] [CrossRef] [Green Version]
  39. Żółtowska, K.; Oledzka, E.; Kuras, M.; Skrzypczak, A.; Nałęcz-Jawecki, G.; Sobczak, M. Cyto- and genotoxicity evaluation of the biomedical polyesters obtained in the presence of new zinc catalytic systems. Int. J. Polym. Mater. Polym. Biomater. 2017, 66, 768–772. [Google Scholar] [CrossRef]
  40. Gálico, D.A.; Nova, C.V.; Guerra, R.B.; Bannach, G. Thermal and spectroscopic studies of the antioxidant food additive propyl gallate. Food Chem. 2015, 182, 89–94. [Google Scholar] [CrossRef]
  41. Pappalardo, D.; Annunziata, L.; Pellecchia, C. Living Ring-Opening Homo- and Copolymerization of ε-Caprolactone and l- and d,l-Lactides by Dimethyl(salicylaldiminato)aluminum Compounds. Macromolecules 2009, 42, 6056–6062. [Google Scholar] [CrossRef]
  42. Kasperczyk, J.; Bero, M. Coordination polymerization of lactides, 4. The role of transesterification in the copolymerization of L,L-lactide and ε-caprolactone. Makromol. Chem. 1993, 194, 913–925. [Google Scholar] [CrossRef]
  43. Bero, M.; Kasperczyk, J. Coordination polymerization of lactides, 5. Influence of lactide structure on the transesterification processes in the copolymerization with ε-caprolactone. Macromol. Chem. Phys. 1996, 197, 3251–3258. [Google Scholar] [CrossRef]
  44. Vanhoorne, P.; Dubois, P.; Jerome, R.; Teyssie, P. Macromolecular engineering of polylactones and polylactides. 7. Structural analysis of copolyesters of ε-caprolactone and L- or D,L-lactide initiated by triisopropoxyaluminum. Macromolecules 1992, 25, 37–44. [Google Scholar] [CrossRef]
  45. Zambelli, A.; Caprio, M.; Grassi, A.; Bowen, D.E. Syndiotactic styrene-butadiene block copolymers synthesized with CpTiX3/MAO (Cp = C5H5, X = Cl, F; Cp = C5Me5, X = Me) and TiXn/MAO (n = 3, X = acac; n = 4, X = O-tert-Bu). Macromol. Chem. Phys. 2000, 201, 393–400. [Google Scholar] [CrossRef]
  46. Cuomo, C.; Serra, M.C.; Maupoey, M.G.; Grassi, A. Copolymerization of Styrene with Butadiene and Isoprene Catalyzed by the Monocyclopentadienyl Titanium Complex Ti(η5-C5H5)(η2-MBMP)Cl. Macromolecules 2007, 40, 7089–7097. [Google Scholar] [CrossRef]
  47. Sarasua, J.R.; López-Rodríguez, N.; Zuza, E.; Petisco, S.; Castro, B.; del Olmo, M.; Palomares, T.; Alonso-Varona, A. Crystallinity assessment and in vitro cytotoxicity of polylactide scaffolds for biomedical applications. J. Mater. Sci. Mater. Med. 2011, 22, 2513–2523. [Google Scholar] [CrossRef]
  48. Lin, L.; Xu, Y.; Wang, S.; Xiao, M.; Meng, Y. Ring-opening polymerization of l-lactide and ε-caprolactone catalyzed by versatile tri-zinc complex: Synthesis of biodegradable polyester with gradient sequence structure. Eur. Polym. J. 2016, 74, 109–119. [Google Scholar] [CrossRef]
  49. International Organization for Standardization. Available online: https://www.iso.org/standard/73342.html (accessed on 6 April 2019).
  50. Nałęcz-Jawecki, G. Spirotox—Spirostomum ambiguum acute toxicity test—10 years of experience. Environ. Toxicol. 2004, 19, 359–364. [Google Scholar] [CrossRef]
  51. Oda, Y.; Nakamura, S.; Oki, I.; Kato, T.; Shinagawa, H. Evaluation of the new system (umu-test) for the detection of environmental mutagens and carcinogens. Mutat. Res. Mutagen. Relat. Subj. 1985, 147, 219–229. [Google Scholar] [CrossRef]
Sample Availability: Samples of the all compounds are available from the authors.
Figure 1. The scheme of copolymerization of rac-LA and CL.
Figure 1. The scheme of copolymerization of rac-LA and CL.
Molecules 24 04168 g001
Figure 2. Typical 13C NMR spectrum of the obtained copolymers (carbonyl region).
Figure 2. Typical 13C NMR spectrum of the obtained copolymers (carbonyl region).
Molecules 24 04168 g002
Figure 3. DSC curves of the selected samples and references.
Figure 3. DSC curves of the selected samples and references.
Molecules 24 04168 g003
Table 1. Copolymerization conditions of CL and rac-LA.
Table 1. Copolymerization conditions of CL and rac-LA.
EntryMolar Ratio CL/rac-LAMolar Ratio Zn/MonomersReaction Time [h]Temp. [°C]Conv.LAConv.CL[L] aYield b [%]lCap clLL dR eMnfĐfTII [%]
11/11/10016800.880.920.77642.324.830.4590001.4231
21/11/10024800.850.960.60832.242.650.4797001.5447
31/11/10048800.820.970.58871.851.630.7490001.6174
42/11/10016800.760.960.50653.131.820.5592001.3539
52/11/10024800.740.970.39932.581.090.7680001.4667
62/11/10048800.750.970.43853.031.340.6592001.5360
71/21/10016800.890.920.77681.764.450.4986001.6038
81/21/10024800.850.930.68901.683.340.4779001.5140
91/21/10048800.860.940.73851.783.820.4986001.7023
101/12/10016800.800.920.65642.002.160.6754001.3260
111/12/10024800.800.950.58582.321.870.6454001.2440
121/12/10048800.880.920.58632.411.820.6653001.3845
132/12/10016800.670.960.42672.211.090.8060001.3989
142/12/10024800.730.960.46692.611.180.7967001.3569
152/12/10048800.760.980.44622.841.240.7267001.4560
161/22/10016800.880.920.78671.703.960.5871001.5246
171/22/10024800.820.950.79641.854.140.5671001.4239
181/22/10048800.840.870.77661.272.670.8377001.56100
191/16/10024600.610.750.512011.683.100.3377001.227.4
201/16/10048600.900.830.64465.197.510.1884001.2911
211/16/10016800.500.550.70817.157.430.2384001.386.5
221/16/10024800.550.650.727910.072.540.6985001.4312
231/16/10048800.890.860.62524.403.290.4096001.5128
242/18/10024600.830.290.79258.1212.940.1979001.191.7
252/18/10048600.860.960.48529.893.540.2784001.269.6
262/18/10016800.780.930.49837.764.050.2486001.386.1
272/18/10024800.850.960.49597.233.210.3192001.3614
282/18/10048800.840.960.49328.822.950.3398001.405.3
291/28/10024600.850.600.82445.9622.040.1373001.320
301/28/10048600.900.870.80466.0115.960.1680001.4015
311/28/10016800.770.900.81736.3915.970.1780001.543.0
321/28/10024800.880.900.80423.9713.310.1986001.5939
331/28/10048800.870.900.75344.677.310.2893001.7112
a Molar fraction of lactyl units in the polymer (determined by 1H NMR); b Isolated yield; c Average length of caproyl blocks; d Average length of lactydyl blocks; e Randomness degree; f Determined by GPC.
Table 2. Differential scanning calorimetry (DSC) results of the selected copolymers and homopolymers as references.
Table 2. Differential scanning calorimetry (DSC) results of the selected copolymers and homopolymers as references.
EntryTm1 [°C] aTm2 [°C]Tg [°C] b
PCL69.4-−60.0
PLA--53.4
951.6-−3.8
1852.5164.9−10.0
3256.4181.8−32.5
a Melting temperature; b Glass transition temperature.
Table 3. The cytotoxicity results of the synthesized copolymers.
Table 3. The cytotoxicity results of the synthesized copolymers.
EntrySpirotox
24 h-PE 1
Microtox
15 min-PE 1
1013 ± 12
2025 ± 6
3010 ± 6
4016 ± 3
5034 ± 3
6018 ± 4
7010 ± 6
8021 ± 3
9022 ± 2
1 Percent of toxic effect.
Table 4. The results of the umu-test for the highest concentrations of the tested extracts (0.66 mg/mL).
Table 4. The results of the umu-test for the highest concentrations of the tested extracts (0.66 mg/mL).
Entry−S9 a+S9 b
G c ± SDIR d ± SDG c ± SDIR d ± SD
11.02 ± 0.020.86 ± 0.110.91 ± 0.021.00 ± 0.02
21.00 ± 0.010.77 ± 0.110.90 ± 0.041.00 ± 0.31
31.05 ± 0.030.73 ± 0.080.91 ± 0.131.01 ± 0.15
41.08 ± 0.020.77 ± 0.081.03 ± 0.190.79 ± 0.07
51.00 ± 0.060.75 ± 0.090.87 ± 0.051.06 ± 0.02
60.97 ± 0.070.90 ± 0.141.05 ± 0.150.81 ± 0.09
71.13 ± 0.050.85 ± 0.061.04 ± 0.061.02 ± 0.08
81.02 ± 0.130.88 ± 0.281.01 ± 0.160.75 ± 0.09
90.99 ± 0.020.84 ± 0.131.07 ± 0.100.77 ± 0.09
Negative Control1.01 ± 0.090.91 ± 0.171.00 ± 0.100.99 ± 0.14
Solvent Control0.92 ± 0.120.83 ± 0.120.93 ± 0.070.84 ± 0.07
a Without metabolic activation, b With metabolic activation, c Growth, d Induction Ratio.

Share and Cite

MDPI and ACS Style

Wyrębiak, R.; Oledzka, E.; Figat, R.; Sobczak, M. Application of Diethylzinc/Propyl Gallate Catalytic System for Ring-Opening Copolymerization of rac-Lactide and ε-Caprolactone. Molecules 2019, 24, 4168. https://doi.org/10.3390/molecules24224168

AMA Style

Wyrębiak R, Oledzka E, Figat R, Sobczak M. Application of Diethylzinc/Propyl Gallate Catalytic System for Ring-Opening Copolymerization of rac-Lactide and ε-Caprolactone. Molecules. 2019; 24(22):4168. https://doi.org/10.3390/molecules24224168

Chicago/Turabian Style

Wyrębiak, Rafał, Ewa Oledzka, Ramona Figat, and Marcin Sobczak. 2019. "Application of Diethylzinc/Propyl Gallate Catalytic System for Ring-Opening Copolymerization of rac-Lactide and ε-Caprolactone" Molecules 24, no. 22: 4168. https://doi.org/10.3390/molecules24224168

Article Metrics

Back to TopTop