Next Article in Journal
XML-CIMT: Explainable Machine Learning (XML) Model for Predicting Chemical-Induced Mitochondrial Toxicity
Next Article in Special Issue
Semiclassical Theory of Multistage Nonequilibrium Electron Transfer in Macromolecular Compounds in Polar Media with Several Relaxation Timescales
Previous Article in Journal
The Chromatin Structure at the MECP2 Gene and In Silico Prediction of Potential Coding and Non-Coding MECP2 Splice Variants
Previous Article in Special Issue
The Novel Gallium Aminobisphenolate Initiator of the Ring-Opening Copolymerization of L-Lactide and ε-Caprolactone: A Computational Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Gallium (III) Complexes Based on Aminobisphenolate Ligands: Extremely High Active ROP-Initiators from Well-Known and Easily Accessible Compounds

1
Chemistry Department, Moscow State University, 119991 Moscow, Russia
2
N.D. Zelinsky Institute of Organic Chemistry, Russian Academy of Science, 119991 Moscow, Russia
3
Institute of General and Inorganic Chemistry, Russian Academy of Science, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(24), 15649; https://doi.org/10.3390/ijms232415649
Submission received: 8 November 2022 / Revised: 5 December 2022 / Accepted: 7 December 2022 / Published: 9 December 2022
(This article belongs to the Collection State-of-the-Art Macromolecules in Russia)

Abstract

:
We report herein the synthesis and full characterizations of the first examples of gallium complexes based on “privileged” aminobisphenolate ligands which are easily available. These complexes turned out to be extremely active in the ring-opening polymerization of ε-caprolactone even at room temperature and highly active in the ROP of L-lactide. The combination of factors such as the easy availability of these compounds and the supposedly low toxicity, together with the extremely high activity in ROP, allows us to consider these compounds as suitable for use on an industrial scale for the synthesis of biodegradable polymers for biomedical applications.

1. Introduction

Biodegradable polymers such as polylactide (PLA), poly(butylene succinate) (PBS), poly(ε-caprolactone) (PCL), polyglycolide (PGA), and poly(propylene carbonate) (PPC) during last three decades have become an attractive alternative to classic polyolefins in several fields of technology [1]. On the one hand, their physical and mechanical properties are often close to those of «classical» polymers based on alpha-olefins, on the other hand, their ability to decompose relatively quickly both in the environment and in living organisms determines two main directions of their use: these are various types of packaging and medical applications, such as, for example, suture material, etc. Among biodegradable polymers, PCL is one such material, which has received global attraction due to its biocompatibility, nontoxicity, and cost efficiency. It should also emphasize the presence of FDA approval for the internal medicinal application of PCL. These applications are controlled drug delivery systems, tissue-engineering scaffolds, the formation of artificial organs, nerve regeneration, and wound healing. Because PCL is semicrystalline as well as hydrophobic, it takes almost 3–4 years to degrade completely in vivo, making it a popular choice for long-term implants, bone tissue engineering, and slow-release medication administration [2,3,4].
Although the monomer of PCL, ε-caprolactone (ε-CL), is currently prepared from fossil resources in the industry, great success to date has been achieved in the synthesis of ε-caprolactone from renewable resources via the chemical treatment of saccharides [5]. The last makes PCL a fully sustainable polymer. In industry, PCL is made by ring-opening polymerization (ROP) of ε-caprolactone in presence of an initiator based on metal. Many main groups of metal and transition metal compounds have been reported as initiators for the ROP of ε-CL [6,7,8]. However, industrially used tin(II) bis-2-ethylhexanoic acid (tin octoate) in presence of alcohol is still the preferred catalytic system for the bulk polymerization of ε-caprolactone due to its high stability during the polymerization process and adequate activity in industry-reasonable polymerization conditions. Additionally, this initiator allows good control of polymerization and affords high-molecular-weight PCL with controlled properties [9,10]. At the same time, the use of tin octoate for PCL production for future medical applications is still in question due to the toxicity of tin compounds remaining in the polymer [11]. Thus, the synthesis of new initiators based on non-toxic metals, which produce PCL in a controlled manner with high molecular weight, and narrow molecular weight distribution due to the absence of side reactions from transesterification, remains relevant.
Amino bisphenols belong to the so-called “privileged ligands” and are easily available and inexpensive, giving them great potential as ligands for catalysts for a wide range of chemical transformations [12]. The simple synthetic approaches to these compounds allow for the introduction in the molecule of ligand different substituents. The steric and electronic properties of substituents, the nature of donor atoms in the ligand framework, and the size of the chelate formed with the metal atom determine the geometry and coordination mode adopted by metal complexes. The latter is very important for catalysis applications [13]. Aminobisphenolate complexes of main group elements (for example, Al [14,15,16,17,18,19], alkali metals [20,21,22], Ca [21], Mg [23]), transition metals (for example, Ti [19]) and lanthanides [24,25] have been reported to catalyze ROP reactions, but by now there is absolutely no information not only about catalysis but also about the synthesis of such gallium complexes in general. At the same time, gallium is a non-toxic metal if one may consider the concentrations of metal compounds that remain in the polymer during its synthesis [26], in addition, there are examples of the use of some gallium complexes in medicine [27], and gallium complexes of various structures (for example, diamido-ether dianionic ligands, (aminomethyl)phenolate monoanionic ligands, 8-quinolinolato monoanionic ligands, salan dianionic ligands, salen dianionic ligands, bis(imino)phenoxide monoanionic ligands) are active in ROP as initiators, [28,29,30,31,32,33,34,35,36,37] thus aminobisphenolate gallium derivatives should be considered as potentially promising ROP initiators and interesting targets for their synthesis and investigation of their behavior in ROP.
In order to understand more about the structural chemistry of gallium complexes and the catalytic property of 13 Group elements complexes, we began to prepare gallium complexes bearing aminebisphenolate ligands and studied their catalytic property for the polymerization of ε-CL. All of them can catalyze the polymerization of ε-CL as well as L-lactide (L-LA) in absence of internal nucleophile (alcohol) with extremely high (ε-CL) and high activity (L-LA) to prepare PCL or PLA with narrow molecular weight distribution.

2. Results and Discussion

The room temperature reaction of amino bisphenols 13 (novel ligand 3 has been prepared according to a well-known literature procedure, see for example [19]) with equimolar amounts of dimeric tris(dimethyl)amide gallium (the molar ratio 1:0.5) led to monomeric gallium monoamides 46, EtN[CH2–(3-R1–(5-R2–)C6H2–2-O)2]2Ga–NMe2 (4, R1 = tBu, R2 = Me; 5, R1 = R2= tBu; 6, R1 = PhMe2C, R2 = Me), as white powders with high yields (Scheme 1).
All of those compounds were isolated as white solids and were characterized by spectroscopic studies as well as microanalyses. 1H and 13C{1H} NMR spectra of 46 (see Supplementary Materials Figures S2–S9 for details) are indicative of a highly symmetric species in solution, with two diastereotopic signals (two doublets of AB-system) corresponding to the methylene protons of N-CH2-aryl found around 3.0–3.5 ppm in C6D6. Of interest, there are two non-equivalent methyl groups in PhMe2C-substituents (two singlets 1.76 (6H) and 1.61 (6H) ppm in 1H NMR spectrum and two signals in 13C NMR spectrum) in 6 due to hindered rotation around Ar–CMe2Ph bonds.
Unfortunately, we failed to grow crystals of 46 suitable for X-ray diffraction analysis. However, upon slow evaporation of the mother liquor remaining after recrystallization of 6, a crystal of compound 7 was obtained, which has been studied by X-ray crystallography. Compound 7 is adduct of dimethylamine with EtN[CH2-(3-R1-(5-R2-)C6H2-2-O)2]2Ga-OMe. This compound was formed by the reaction of methanol, which apparently remained in trace amounts in ligand 3 after recrystallization of that from methanol during its purification and compound 6 (Scheme 2). The molecular structure of 7 is shown in Figure 1. The methoxy group was found to be disordered over two positions with 64/36 occupancies. Selected bond lengths and angles of 7 are listed in the Figure 1 caption.
Structure 7 exhibit a monomeric structure with a five-coordinate gallium center, including two six-membered amine phenolate metal-rings. The Ga–OAr bond lengths (1.874(3), 1.878(3) Å) as well as Ga–OAlk (1.839(14) Å) are close to those found in related derivatives contained GaO3N2 fragment [31,33,37,38,39,40]. The Ga–N coordinative bond lengths (2.091(5) Å for NHC2, 2.119(4) Å for NC3) of the amine donor is close to those for related derivatives contained GaO3N2 fragment [40].
The important question is the substituents’ geometry around the Ga atom. The coordination polyhedron of the gallium atom in 7 can be described as either a distorted trigonal bipyramid with both N atoms in the apical positions and the three oxygen atoms in equatorial sites or a distorted tetragonal pyramid where N(1), N(2), O(1), O(2) atoms form pyramid base, O(3A)/O(3B) occupies the apical site. According to the approach of Addington et al. [41] and others [42], the assignment to one or another type of polyhedron can be made based on the value of the parameter τ which is applicable to five-co-ordinate structures as an index of the degree of trigonality, within the structural continuum between trigonal bipyramidal and rectangular pyramidal. For a perfectly tetragonal geometry τ = 0, while for a perfectly trigonal-bipyramidal geometry τ = 1. For 7 parameter τ has an intermediate value (0.53). Of interest, in similar Al aminobisphenolate complexes [18], where an Al atom bonded to two oxygen atoms (covalent bonds) and two nitrogen atoms (coordination bonds) parameter τ varies in the value range 0.53–0.77.
Compounds 46 were tested as catalysts towards the ROP of ε-CL and L-LA. The polymerizations of ε-CL were conducted in bulk at 80 °C. Compound 5 also was tested as the catalyst for the ROP of ε-CL at 25 °C and 100 °C; all the polymerizations of L-LA were conducted also under solvent-free conditions at 100 °C. In all studied cases, no external nucleophile such as alcohol has been used as a co-initiator. The polymerizations were monitored by taking aliquots at regular time intervals, which were analyzed using 1H NMR spectroscopy to determine the lactide conversion and by GPC (gel permeation chromatography) to determine the number of average molecular weight (Mn) and molecular weight distribution (PDIs, Mw/Mn). The polymerization results are summarized in Table 1.
Complex 5 tested was extremely active for the controlled polymerization of ε-caprolactone, as indicated by the monomer conversions and relatively narrow polydispersities of the PCL. It should be noted that the activity of initiator 5 (entry 6) is the highest (100% conversion, 15 min, 25 °C) among gallium complexes studied as initiators of the polymerization of caprolactone before. There are two comprehensive reviews of using gallium compounds for ester ROP published so far [7,8]. It is emphasized that heavier metal group 13 complexes based on gallium and indium have emerged as effective catalysts, but examples of catalysts containing these metals are still relatively rare and deserve further investigation. Nevertheless, indium compounds have been studied to a much greater extent than gallium compounds [7]. Perhaps, first of all, this is due to the fact that the Ga derivative of the salen type (O,N,N,O-type ligand with pentacoordinated Ga) is practically not active in ROP of lactide unlike the In complex of similar structure [37]. Additionally, the authors associated this poor activity with the electronic properties of the gallium atom. In addition, the gallium complex of N,O,N-type ligands with tetracoordinated Ga is practically not active in the ROP of CL [28]. However, the results of this work clearly demonstrate the particular importance of the structure of the ligand, which directly affects the activity of the initiator.
End-group analysis of the isolated PCL-200 (Table 1, entry 6) from the 1H NMR spectrum (see Supplementary Materials Figure S1 for details) indicates that the PCL chain is capped with one dimethyl amide and one hydroxyl chain end. This fact confirms that the “coordination-insertion” mechanism of ROP is implemented in this case [43]. Recently, the ability of complexes containing an amide moiety (M–NR2) to initiate ROP leading to polymers containing an amino-terminal group has also been demonstrated [44]. The Mn values of the obtained PCL samples (Table 1, entry 1–6) were lower than the expected values based on % conversion. This deviation likely implies that some transesterification reactions and/or a slow initiation step relative to propagation occurred. This trend is frequently observed in the ROP of the cyclic esters [45]. To demonstrate the capability of synthesizing PCL with a larger Mn and a well-controlled character, we accomplished ROP of ε-CL with [ε-CL]/[5] = 500 at 100 °C for 30 min (Table 1, entry 5). PCL with a large molecular weight, Mn (obsd) = 42,023 g mol−1 after the correction was obtained; its PDI in this case does not grow compared to the data obtained for a smaller ratio (Table 1, entry 2).
All synthesized complexes were also tested in ROP of L-LA (Table 1, entries 7–9) and showed slightly lower activity in this ROP compared to that for ε-CL. Previously, it was found that ε-CL usually has a higher polymerization rate in its respective homopolymerizations than L-LA [46]. The Mn values of the obtained PLA samples were very close to the expected values based on % conversion, which demonstrated the controlled character of L = LA polymerization. According to obtained NMR data, racemization did not occur during polymerization and pure poly-L-lactide is formed.

3. Materials and Methods

All reactions with air- and/or water-sensitive compounds were performed under a dry, oxygen-free argon atmosphere using standard Schlenk techniques. Solvents were dried by standard methods and distilled prior to use: toluene, n-hexane were refluxed under Na and distilled; ether, THF were stored under KOH, refluxed under Na/benzophenone and then distilled; methanol was refluxed under Mg and then distilled off. Starting materials were synthesized according to the literature procedures: 6,6′-((ethylazanediyl)bis(methylene))bis(2-(tert-butyl)-4-methylphenol) (1) [47], 6,6′-((ethylazanediyl)bis(methylene))bis(2,4-di-tert-butylphenol) (2) [48], [Ga(NMe2)3]2 [49]. C6D6 (dried over sodium), CDCl3 (dried with CaH2) and DMSO-d6 (dried over CaH2, distilled under reduced pressure, and stored under an argon atmosphere over molecular sieves) obtained from Deutero GmbH. 1H (400.13 MHz), 13C (100.61 MHz), were recorded on a Bruker Avance 400 or Agilent 400-MR spectrometers at room temperature (if otherwise stated). 1H and 13C chemical shifts are reported in ppm relative to Me4Si as external standard. Elemental analysis was performed using EuroEA-3000 (EuroVector, Pavia, Italy) instrument.

3.1. Synthesis of Ligand 3

In a flask, 4-methyl-2-(2-phenylpropan-2-yl)phenol (16.49 g, 73.1 mmol), 36% aqueous formaldehyde (5.97 mL, 70.20 mmol), 70% aqueous EtNH2 (2.50 mL, 39.0 mmol) and 15 mL of distilled water were placed. Solution was refluxed for 24 h, and then water was separated by decantation, and the mixture was dissolved in CH2Cl2 and washed with water. The organic layer was dried over MgSO4. Then, the volatile materials were removed under reduced pressure. The crude product was purified by recrystallization from methanol to afford the pure product. Compound 3 (12.3 g, 61%) was obtained as a white solid, m.p. 116–118 °C (from MeOH). 1H NMR (400 MHz, C6D6): δ (ppm) 7.28 (m, 4H; ArH), 7.22 (d, J = 1.6 Hz, 2H; ArH), 7.11 (m, 4H; ArH), 7.03 (m, 2H; ArH), 6.73 (d, J = 1.4 Hz, 2H; ArH), 3.39 (s, 4H; NCH2Ar), 2.26 (s, 6H; CH3Ar), 2.21 (q, J = 7.1 Hz, 2H; CH2CH3), 1.68 (s, 12H; (CH3)2PhCAr), 0.67 (t, J = 7.1 Hz, 3H, CH2CH3); 13C NMR (100 MHz, CDCl3): δ (ppm) 153.15, 150.59, 136.21, 129.95, 129.07, 128.29, 127.36, 126.82, 126.50, 124.82 (Ar), 55.15 (NCH2Ar), 46.92 (CH2CH3), 42.45 ((CH3)2PhCAr), 30.22 ((CH3)2PhCAr), 21.49 (CH3Ar), 11.38 (CH2CH3); Anal. Calcd for C36H43NO2: C, 82.87; H, 8.31; N, 2.68; found: C, 82.95; H, 8.36; N, 2.76%.

3.2. General Synthesis of Ga Complexes (46)

A solution of the ligand (2.03 eqv.) in dry toluene was added dropwise to solution of [Ga(NMe2)3]2 (1 eqv.) in toluene with stirring at −35 °C. The resulting reaction mixture was stirred for 48 h. Then the volatile materials were removed under reduced pressure. The residue was recrystallized from toluene/n-hexane.

3.3. Synthesis of Complex 4

Prepared from 6,6′-((ethylazanediyl)bis(methylene))bis(2-(tert-butyl)-4-methylphenol) (1) (0.49 g, 1.24 mmol) and [Ga(NMe2)3]2 (0.27 g, 0.61 mmol). Yield: 0.59 g, 96%, white powder. 1H NMR (400 MHz, C6D6): δ (ppm) 7.27 (d, J = 1.9 Hz, 2H; ArH), 6.43 (d, J = 1.9 Hz, 2H; ArH), 3.33 (d, J = 12.5 Hz, 2H; NCH2Ar), 3.12 (d, J = 12.9 Hz, 2H; NCH2Ar), 2.78 (s, 6H; N(CH3)2)), 2.42 (q. J = 7.0 Hz, 2H; CH2CH3), 2.26 (s, 6H; CH3Ar), 1.69 (s, 18H; C(CH3)3), 0.62 (t, J = 7.0 Hz, 3H, CH2CH3); 13C NMR (100 MHz, C6D6): δ (ppm) 159.07, 140.36, 129.52, 129.47, 126.63, 121.46 (Ar), 56.50 (NCH2Ar), 49.22 (CH2CH3), 42.61 (N(CH3)2), 35.69 (C(CH3)3), 30.40 (C(CH3)3), 21.37 (CH3Ar), 6.95 (CH2CH3); Anal. Calcd for C28H43GaN2O2: C, 66.02; H, 8.51; N, 5.50; found: C, 66.24; H, 8.60; N, 5.57%.

3.4. Synthesis of Complex 5

Prepared from 6,6′-((ethylazanediyl)bis(methylene))bis(2,4-di-tert-butylphenol) (2) (0.76 g, 1.57 mmol) and [Ga(NMe2)3]2 (0.32 g, 0.79 mmol). Yield: 0.85 g, 92%, white powder. 1H NMR (400 MHz, C6D6): δ (ppm) 7.59 (d, J = 2.4 Hz, 2H; ArH), 6.69 (d, J = 2.4 Hz, 2H; ArH), 3.40 (d, J = 12.9 Hz, 2H; NCH2Ar), 3.16 (d, J = 12.9 Hz, 2H; NCH2Ar), 2.78 (s, 6H; N(CH3)2)), 2.42 (q. J = 7.0 Hz, 2H; CH2CH3), 1.71 (s, 18H; C(CH3)3), 1.39 (s, 18H; C(CH3)3), 0.61 (t, J = 7.0 Hz, 3H, CH2CH3); 13C NMR (100 MHz, C6D6): δ (ppm) 158.87, 140.14, 139.79, 128.69, 125.54, 120.81 (Ar), 57.06 (NCH2Ar), 49.20 (CH2CH3), 42.58 (N(CH3)2), 36.04, 34.67 (C(CH3)3), 32.40, 30.44 (C(CH3)3), 6.92 (CH2CH3); Anal. Calcd for C34H55GaN2O2: C, 68.80; H, 9.34; N, 4.72; found: C, 67.62; H, 9.26; N, 4.64%.

3.5. Synthesis of Complex 6

Prepared from 6,6′-((ethylazanediyl)bis(methylene))bis(4-methyl-2-(2-phenylpropan-2-yl)phenol) (3) (0.81 g, 1.56 mmol) and [Ga(NMe2)3]2 (0.31 g, 0.78 mmol). Yield: 1.12 g, 94%, white powder. 1H NMR (400 MHz, C6D6): δ (ppm) 7.26 (m, 6H; ArH), 7.03 (m, 6H; ArH), 6.42 (s, 2H; ArH), 3.26 (d, J = 13.2 Hz, 2H; NCH2Ar), 3.03 (d, J = 13.2 Hz, 2H; NCH2Ar), 2.91 (q. J = 7.0 Hz, 2H; CH2CH3), 2.50 (s, 6H; N(CH3)2)), 2.28 (s, 6H; CH3Ar), 1.76 (s, 6H; (CH3)2PhCAr), 1.61 (s, 6H; (CH3)2PhCAr), 0.61 (t, J = 7.0 Hz, 3H, CH2CH3); 13C NMR (100 MHz, C6D6): δ (ppm) 159.07, 152.86, 137.53, 129.67, 128.91, 128.47, 128.45, 126.63, 126.16, 121.29 (Ar), 57.19 (NCH2Ar), 48.39 (CH2CH3), 43.62 (N(CH3)2), 34.32 ((CH3)2PhCAr), 28.36 (CH3Ar), 21.42 ((CH3)2PhCAr), 6.49 (CH2CH3); Anal. Calcd for C38H47GaN2O2: C, 72.04; H, 7.48; N, 4.42; found: C, 72.15; H, 7.54; N, 4.46%.

3.6. Typical Polymerization Procedure in Bulk

All manipulations were performed under inert atmosphere. To the initiator 4 (0.0477 g, 0.098 mmol) ε-caprolactone (2.2257 g, 19.50 mmol) was added. The reaction mixture was heated at 80 °C for 15 min. The reaction was terminated by addition of MeOH (1.0 mL), evaporated, and purified by reprecipitation using CH2Cl2 as solvent and methanol as a non-solvent. The polymer obtained was dried in vacuum.

3.7. Single Crystal X-ray Diffraction Studies

Crystal data for 7: C46H59GaN2O3, Fw = 757.67, monoclinic, a = 13.1068(6), b = 22.1139(11), c = 14.1810(8) Å, β = 95.663(2)°, V = 4090.2(4) Å3, space group P21/n, Z = 4, Dc = 1.230 g/cm3, F(000) = 1616, μ(MoKα) = 0.714 mm−1, colourless block with dimensions ca. 0.15 × 0.10 × 0.05. Total of 44,297 reflections (8026 unique, Rint = 0.169) was measured on a Bruker SMART APEX II diffractometer (graphite monochromatized MoKα radiation, λ = 0.71073 Å) using ω-scan mode at 150 K. Absorption correction based on measurements of equivalent reflections were applied [50]. The structure was solved by direct methods and refined by full-matrix least-squares on F2 with anisotropic thermal parameters for all non-hydrogen atoms [51]. Amino hydrogen atom H2 was found from different Fourier syntheses and refined isotropically. All other H atoms were placed in calculated positions and refined using a riding model. Methoxy group was disordered over two positions with occupancy ratio of 0.714(7)/0.286(7). Solvent toluene molecule was found to be disordered over three sites. The final residuals were: R1 = 0.0715, wR2 = 0.1489 for 4509 reflections with I > 2σ(I) and 0.1489, 0.1774 for all data and 515 parameters. GoF = 1.004, maximum Δρ = 0.745 e × Å−3.
X-ray diffraction studies were performed at the Centre of Shared Equipment of IGIC RAS.
The crystallographic data for 7 have been deposited with the Cambridge Crystallographic Data Centre as supplementary publications under the CCDC number 2195341.

4. Conclusions

In conclusion, we have prepared the first examples of aminobisphenolate-based gallium complexes. These complexes are contained Ga-NMe2 fragments with the coordination number of the gallium atom equal to 4. We have demonstrated that these gallium complexes can be highly effective initiators for the production of PCL and PLA with relatively narrow dispersities and controllable molecular weights. This highlights the importance of the combination of metal/ligand in the rational design of initiators for the controlled ROP of cyclic esters.

Supplementary Materials

The supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms232415649/s1.

Author Contributions

Conceptualization, M.P.E. and S.S.K.; methodology, B.N.M.; software, B.N.M.; investigation, B.N.M., L.F.H. and A.V.C.; data curation, S.S.K.; writing—original draft preparation, B.N.M. and S.S.K.; writing—review and editing, S.S.K.; funding acquisition, M.P.E. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Russian Science Foundation (grant no. 20-13-00391). The synthesis of ligands was partially sponsored by the Moscow University Development Program (partially performed by L.F.H). X-ray diffraction studies were performed A.V.C. at the Centre of Shared Equipment of IGIC RAS.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schneiderman, D.K.; Hillmyer, M.A. 50th Anniversary Perspective: There Is a Great Future in Sustainable Polymers. Macromolecules 2017, 50, 3733–3749. [Google Scholar]
  2. Malikmammadov, E.; Tanir, T.E.; Kiziltay, A.; Hasirci, V.; Hasirci, N. PCL and PCL-based materials in biomedical applications. J. Biomater. Sci. Polym. Ed. 2018, 29, 863–893. [Google Scholar] [CrossRef] [PubMed]
  3. Dhanasekaran, N.P.D.; Muthuvelu, K.S.; Arumugasamy, S.K. Recent Advancement in Biomedical Applications of Polycaprolactone and Polycaprolactone-Based Materials. In Encyclopedia of Materials: Plastics and Polymers; Hashmi, M.S.J., Ed.; Elsevier: Oxford, UK, 2022; pp. 795–809. [Google Scholar]
  4. Mandal, P.; Shunmugam, R. Polycaprolactone: A biodegradable polymer with its application in the field of self-assembly study. J. Macromol. Sci. Part A 2021, 58, 111–129. [Google Scholar]
  5. Rani, G.U.; Sharma, S. Biopolymers, Bioplastics and Biodegradability: An Introduction. In Encyclopedia of Materials: Plastics and Polymers; Hashmi, M.S.J., Ed.; Elsevier: Oxford, UK, 2022; pp. 474–486. [Google Scholar]
  6. Labet, M.; Thielemans, W. Synthesis of polycaprolactone: A review. Chem. Soc. Rev. 2009, 38, 3484–3504. [Google Scholar] [CrossRef] [PubMed]
  7. Gao, J.; Zhu, D.; Zhang, W.; Solan, G.A.; Ma, Y.; Sun, W.-H. Recent progress in the application of group 1, 2 & 13 metal complexes as catalysts for the ring opening polymerization of cyclic esters. Inorg. Chem. Front. 2019, 6, 2619–2652. [Google Scholar]
  8. Dagorne, S.; Normand, M.; Kirillov, E.; Carpentier, J.-F. Gallium and indium complexes for ring-opening polymerization of cyclic ethers, esters and carbonates. Coord. Chem. Rev. 2013, 257, 1869–1886. [Google Scholar] [CrossRef] [Green Version]
  9. Kowalski, A.; Duda, A.; Penczek, S. Kinetics and mechanism of cyclic esters polymerization initiated with tin(II) octoate, 1. Polymerization of ε-caprolactone. Macromol. Rapid Commun. 1998, 19, 567–572. [Google Scholar]
  10. Punyodom, W.; Limwanich, W.; Meepowpan, P.; Thapsukhon, B. Ring-opening polymerization of ε-caprolactone initiated by tin(II) octoate/n-hexanol: DSC isoconversional kinetics analysis and polymer synthesis. Des. Monomers Polym. 2021, 24, 89–97. [Google Scholar] [CrossRef]
  11. Giram, P.S.; Garnaik, B. Evaluation of biocompatibility of synthesized low molecular weight PLGA copolymers using zinc L-proline through green route for biomedical application. Polym. Adv. Technol. 2021, 32, 4502–4515. [Google Scholar]
  12. Wichmann, O.; Sillanpää, R.; Lehtonen, A. Structural properties and applications of multidentate [O,N,O,X′] aminobisphenolate metal complexes. Coord. Chem. Rev. 2012, 256, 371–392. [Google Scholar] [CrossRef]
  13. Akintayo, D.C.; Munzeiwa, W.A.; Jonnalagadda, S.B.; Omondi, B. Zn(II) pyridinyl amine complexes, synthesis and crystal structure studies: A comparative study of the effect of nuclearity and benzoate type on the ring-opening polymerization of cyclic esters. Polyhedron 2022, 213, 115589. [Google Scholar] [CrossRef]
  14. Cross, E.D.; Tennekone, G.K.; Decken, A.; Shaver, M.P. Aluminum amine-(bis)phenolate complexes for ring-opening polymerization of rac-lactide and ε-caprolactone. Green Mater. 2013, 1, 79–86. [Google Scholar] [CrossRef] [Green Version]
  15. Alcazar-Roman, L.M.; O’Keefe, B.J.; Hillmyer, M.A.; Tolman, W.B. Electronic influence of ligand substituents on the rate of polymerization of ε-caprolactone by single-site aluminium alkoxide catalysts. Dalton Trans. 2003, 15, 3082–3087. [Google Scholar]
  16. Chen, C.-T.; Huang, C.-A.; Huang, B.-H. Aluminium metal complexes supported by amine bis-phenolate ligands as catalysts for ring-opening polymerization of ε-caprolactone. Dalton Trans. 2003, 19, 3799–3803. [Google Scholar] [CrossRef]
  17. Chen, C.-T.; Huang, C.-A.; Huang, B.-H. Aluminum Complexes Supported by Tridentate Aminophenoxide Ligand as Efficient Catalysts for Ring-Opening Polymerization of ε-Caprolactone. Macromolecules 2004, 37, 7968–7973. [Google Scholar] [CrossRef]
  18. Phomphrai, K.; Chumsaeng, P.; Sangtrirutnugul, P.; Kongsaeree, P.; Pohmakotr, M. Reverse orders of reactivities in the polymerization of cyclic esters using N2O2 aluminium alkoxide complexes. Dalton Trans. 2010, 39, 1865–1871. [Google Scholar]
  19. Kuchuk, E.A.; Zaitsev, K.V.; Mamedova, F.A.; Churakov, A.V.; Zaitseva, G.S.; Lemenovsky, D.A.; Karlov, S.S. Synthesis, structure, and catalytic activity of new aluminum and titanium complexes based on aminobisphenolate ligands containing bulky substituents. Russ. Chem. Bull. 2016, 65, 1743–1749. [Google Scholar]
  20. Li, X.; Jia, Z.; Pan, X.; Wu, J. Isoselective Ring-Opening Polymerization of rac-Lactide Catalyzed by Sodium/potassium Tetradentate Aminobisphenolate Ion-paired Complexes. Chem. Asian J. 2019, 14, 662–669. [Google Scholar] [CrossRef]
  21. Devaine-Pressing, K.; Oldenburg, F.J.; Menzel, J.P.; Springer, M.; Dawe, L.N.; Kozak, C.M. Lithium, sodium, potassium and calcium amine-bis(phenolate) complexes in the ring-opening polymerization of rac-lactide. Dalton Trans. 2020, 49, 1531–1544. [Google Scholar] [CrossRef]
  22. Yao, C.; Yang, Y.; Xu, S.; Ma, H. Potassium complexes supported by monoanionic tetradentate amino-phenolate ligands: Synthesis, structure and catalysis in the ring-opening polymerization of rac-lactide. Dalton Trans. 2017, 46, 6087–6097. [Google Scholar]
  23. Devaine-Pressing, K.; Lehr, J.H.; Pratt, M.E.; Dawe, L.N.; Sarjeant, A.A.; Kozak, C.M. Magnesium amino-bis(phenolato) complexes for the ring-opening polymerization of rac-lactide. Dalton Trans. 2015, 44, 12365–12375. [Google Scholar] [CrossRef] [Green Version]
  24. Yuan, F.; Zhou, Y.; Li, L.; Zhu, X. Synthesis and structures of amine bis(phenolate) lanthanide thiolates and their application in the polymerization of ε-caprolactone. Inorg. Chim. Acta 2013, 408, 33–38. [Google Scholar] [CrossRef]
  25. Li, M.; Zhang, J.; Chen, J.; Yao, Y.; Luo, Y. Rare-earth metal derivatives supported by aminophenoxy ligand: Synthesis, characterization and catalytic performance in lactide polymerization. Appl. Organomet. Chem. 2020, 34, e5296. [Google Scholar] [CrossRef]
  26. White, S.J.O.; Shine, J.P. Exposure Potential and Health Impacts of Indium and Gallium, Metals Critical to Emerging Electronics and Energy Technologies. Curr. Environ. Health Rep. 2016, 3, 459–467. [Google Scholar] [PubMed]
  27. Chitambar, C.R. Medical Applications and Toxicities of Gallium Compounds. Int. J. Environ. Res. Public Health 2010, 7, 2337–2361. [Google Scholar] [PubMed] [Green Version]
  28. Hild, F.; Neehaul, N.; Bier, F.; Wirsum, M.; Gourlaouen, C.; Dagorne, S. Synthesis and Structural Characterization of Various N,O,N-Chelated Aluminum and Gallium Complexes for the Efficient ROP of Cyclic Esters and Carbonates: How Do Aluminum and Gallium Derivatives Compare ? Organometallics 2013, 32, 587–598. [Google Scholar] [CrossRef]
  29. Dodonov, V.A.; Morozov, A.G.; Rumyantsev, R.V.; Fukin, G.K.; Skatova, A.A.; Roesky, P.W.; Fedushkin, I.L. Synthesis and ε-Caprolactone Polymerization Activity of Electron-Deficient Gallium and Aluminum Species Containing a Charged Redox-Active dpp-Bian Ligand. Inorg. Chem. 2019, 58, 16559–16573. [Google Scholar] [CrossRef]
  30. Basiak, D.; Dobrzycki, Ł.; Socha, P.; Rzepiński, P.; Plichta, A.; Bujnowski, K.; Synoradzki, L.; Orłowska, N.; Ziemkowska, W. Aminophenolates of aluminium, gallium and zinc: Synthesis, characterization and polymerization activity. Appl. Organomet. Chem. 2017, 31, e3748. [Google Scholar] [CrossRef]
  31. Bakewell, C.; White, A.J.P.; Long, N.J.; Williams, C.K. 8-Quinolinolato Gallium Complexes: Iso-selective Initiators for rac-Lactide Polymerization. Inorg. Chem. 2013, 52, 12561–12567. [Google Scholar]
  32. Beament, J.; Mahon, M.F.; Buchard, A.; Jones, M.D. Salan group 13 complexes—Structural study and lactide polymerisation. New J. Chem. 2017, 41, 2198–2203. [Google Scholar] [CrossRef]
  33. Specklin, D.; Fliedel, C.; Hild, F.; Mameri, S.; Karmazin, L.; Bailly, C.; Dagorne, S. Mononuclear salen-gallium complexes for iso-selective ring-opening polymerization (ROP) of rac-lactide. Dalton Trans. 2017, 46, 12824–12834. [Google Scholar] [PubMed]
  34. Horeglad, P.; Cybularczyk, M.; Trzaskowski, B.; Żukowska, G.Z.; Dranka, M.; Zachara, J. Dialkylgallium Alkoxides Stabilized with N-Heterocyclic Carbenes: Opportunities and Limitations for the Controlled and Stereoselective Polymerization of rac-Lactide. Organometallics 2015, 34, 3480–3496. [Google Scholar]
  35. Ghosh, S.; Gowda, R.R.; Jagan, R.; Chakraborty, D. Gallium and indium complexes containing the bis(imino)phenoxide ligand: Synthesis, structural characterization and polymerization studies. Dalton Trans. 2015, 44, 10410–10422. [Google Scholar] [CrossRef] [PubMed]
  36. Cybularczyk-Cecotka, M.; Zaremba, R.; Hurko, A.; Plichta, A.; Dranka, M.; Horeglad, P. Dialkylgallium alkoxides—A tool for facile and stereoselective synthesis of PLA–drug conjugates. New J. Chem. 2017, 41, 14851–14854. [Google Scholar] [CrossRef] [Green Version]
  37. Kremer, A.B.; Andrews, R.J.; Milner, M.J.; Zhang, X.R.; Ebrahimi, T.; Patrick, B.O.; Diaconescu, P.L.; Mehrkhodavandi, P. A Comparison of Gallium and Indium Alkoxide Complexes as Catalysts for Ring-Opening Polymerization of Lactide. Inorg. Chem. 2017, 56, 1375–1385. [Google Scholar] [CrossRef] [Green Version]
  38. Motekaitis, R.J.; Martell, A.E.; Koch, S.A.; Hwang, J.; Quarless, D.A.; Welch, M.J. The Gallium(III) and Indium(III) Complexes of Tris(2-mercaptobenzyl)amine and Tris(2-hydroxybenzyl)amine. Inorg. Chem. 1998, 37, 5902–5911. [Google Scholar] [CrossRef]
  39. Schmidbaur, H.; Lettenbauer, J.; Kumberger, O.; Lachmann, J.; Müller, G. Modellsysteme für die Gallium-Extraktion, II/Model Systems for Gallium Extraction, II. Zeitschrift Naturforschung B 1991, 46, 1065–1076. [Google Scholar] [CrossRef]
  40. Lanznaster, M.; Hratchian, H.P.; Heeg, M.J.; Hryhorczuk, L.M.; McGarvey, B.R.; Schlegel, H.B.; Verani, C.N. Structural and Electronic Behavior of Unprecedented Five-Coordinate Iron(III) and Gallium(III) Complexes with a New Phenol-Rich Electroactive Ligand. Inorg. Chem. 2006, 45, 955–957. [Google Scholar]
  41. Addison, A.W.; Rao, T.N.; Reedijk, J.; van Rijn, J.; Verschoor, G.C. Synthesis, structure, and spectroscopic properties of copper(II) compounds containing nitrogen–sulphur donor ligands; the crystal and molecular structure of aqua[1,7-bis(N-methylbenzimidazol-2′-yl)-2,6-dithiaheptane]copper(II) perchlorate. Dalton Trans. 1984, 7, 1349–1356. [Google Scholar] [CrossRef]
  42. Ershova, I.V.; Bogomyakov, A.S.; Fukin, G.K.; Piskunov, A.V. Features of Magnetic Behavior in the Row of Pentacoordinated Bis-o-Iminobenzosemiquinonato Metal (Al, Ga, In) Complexes. Ber. Dtsch. Chem. Ges. 2019, 2019, 938–948. [Google Scholar]
  43. Ajellal, N.; Carpentier, J.-F.; Guillaume, C.; Guillaume, S.M.; Helou, M.; Poirier, V.; Sarazin, Y.; Trifonov, A. Metal-catalyzed immortal ring-opening polymerization of lactones, lactides and cyclic carbonates. Dalton Trans. 2010, 39, 8363–8376. [Google Scholar] [CrossRef] [PubMed]
  44. Li, C.-Y.; Liu, D.-C.; Ko, B.-T. Synthesis, characterization and reactivity of single-site aluminium amides bearing benzotriazole phenoxide ligands: Catalysis for ring-opening polymerization of lactide and carbon dioxide/propylene oxide coupling. Dalton Trans. 2013, 42, 11488–11496. [Google Scholar] [PubMed]
  45. Ikpo, N.; Saunders, L.N.; Walsh, J.L.; Smith, J.M.B.; Dawe, L.N.; Kerton, F.M. Zinc Complexes of Piperazinyl-Derived Aminephenolate Ligands: Synthesis, Characterization and Ring–Opening Polymerization Activity. Ber. Dtsch. Chem. Ges. 2011, 2011, 5347–5359. [Google Scholar] [CrossRef]
  46. Stirling, E.; Champouret, Y.; Visseaux, M. Catalytic metal-based systems for controlled statistical copolymerisation of lactide with a lactone. Polym. Chem. 2018, 9, 2517–2531. [Google Scholar] [CrossRef]
  47. Zaitsev, K.V.; Kuchuk, E.A.; Churakov, A.V.; Navasardyan, M.A.; Egorov, M.P.; Zaitseva, G.S.; Karlov, S.S. Synthesis and structural characterization of low-valent group 14 metal complexes based on aminobisphenol ligands. Inorg. Chim. Acta 2017, 461, 213–220. [Google Scholar] [CrossRef]
  48. Safaei, E.; Rasouli, M.; Weyhermüller, T.; Bill, E. Synthesis and characterization of binuclear [ONXO]-type amine-bis(phenolate) copper(II) complexes. Inorg. Chim. Acta 2011, 375, 158–165. [Google Scholar] [CrossRef]
  49. Waggoner, K.M.; Olmstead, M.M.; Power, P.P. Structural and spectroscopic characterization of the compounds [Al(NMe2)3]2, [Ga(NMe2)3]2, [(Me2N)2Al{μ-N(H)1-Ad}]2 (1-Ad = 1-adamantanyl) and [{Me(μ-NPh2)Al}2NPh(μ-C6H4)]. Polyhedron 1990, 9, 257–263. [Google Scholar] [CrossRef]
  50. Sheldrick, G.M. SADABS. In Program for Scaling and Correction of Area Detector Data; University of Göttingen: Göttingen, Germany, 1997. [Google Scholar]
  51. Sheldrick, G. A short history of SHELX. Acta Cryst. 2008, 64, 112–122. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of complexes 46.
Scheme 1. Synthesis of complexes 46.
Ijms 23 15649 sch001
Scheme 2. The plausible synthetic way for the formation of complex 7.
Scheme 2. The plausible synthetic way for the formation of complex 7.
Ijms 23 15649 sch002
Figure 1. Molecular structure of 7. Hydrogen atoms (except hydrogen atom on N(2)) and solvate molecule of toluene omitted for clarity. Selected bond Lengths (Å) and angles (deg) for 7: Ga(1)-O(3A) 1.839(14), Ga(1)-O(1) 1.874(3), Ga(1)-O(2) 1.878(3), Ga(1)-O(3B) 1.89(2), Ga(1)-N(2) 2.091(5), Ga(1)-N(1) 2.119(4), O(3A)-Ga(1)-O(1) 118.7(4), O(3A)-Ga(1)-O(2) 108.6(4), O(1)-Ga(1)-O(2) 132.65(15), O(1)-Ga(1)-O(3B) 108.9(6), O(2)-Ga(1)-O(3B) 118.3(6), O(3A)-Ga(1)-N(2) 98.7(5), O(1)-Ga(1)-N(2) 88.54(18), O(2)-Ga(1)-N(2) 82.62(17), O(3B)-Ga(1)-N(2) 98.6(5), O(3A)-Ga(1)-N(1) 95.9(5), O(1)-Ga(1)-N(1) 89.12(15), O(2)-Ga(1)-N(1) 87.71(14), O(3B)-Ga(1)-N(1) 96.7(5), N(2)-Ga(1)-N(1) 164.46(17).
Figure 1. Molecular structure of 7. Hydrogen atoms (except hydrogen atom on N(2)) and solvate molecule of toluene omitted for clarity. Selected bond Lengths (Å) and angles (deg) for 7: Ga(1)-O(3A) 1.839(14), Ga(1)-O(1) 1.874(3), Ga(1)-O(2) 1.878(3), Ga(1)-O(3B) 1.89(2), Ga(1)-N(2) 2.091(5), Ga(1)-N(1) 2.119(4), O(3A)-Ga(1)-O(1) 118.7(4), O(3A)-Ga(1)-O(2) 108.6(4), O(1)-Ga(1)-O(2) 132.65(15), O(1)-Ga(1)-O(3B) 108.9(6), O(2)-Ga(1)-O(3B) 118.3(6), O(3A)-Ga(1)-N(2) 98.7(5), O(1)-Ga(1)-N(2) 88.54(18), O(2)-Ga(1)-N(2) 82.62(17), O(3B)-Ga(1)-N(2) 98.6(5), O(3A)-Ga(1)-N(1) 95.9(5), O(1)-Ga(1)-N(1) 89.12(15), O(2)-Ga(1)-N(1) 87.71(14), O(3B)-Ga(1)-N(1) 96.7(5), N(2)-Ga(1)-N(1) 164.46(17).
Ijms 23 15649 g001
Table 1. Polymerization in bulk of ε-caprolactone and L-lactide.
Table 1. Polymerization in bulk of ε-caprolactone and L-lactide.
EntryCatalyst, [cat]t, [min]Conversion, [%]Mn a (calc), [g/mol]Mn b (exp), [g/mol]PDI
ε-caprolactone (80 °C, [M]0/[cat] = 200:1)
141510022,80091721.80
251510022,80015,4071.51
361510022,80028,5381.47
ε-caprolactone (100 °C, [M]0/[cat] = 500:1)
451586---
53010072,00042,0231.79
ε-caprolactone (25 °C, [M]0/[cat] = 200:1)
651510022,80012,2701.38
L-lactide (100 °C, [M]0/[cat] = 200:1)
741575---
83010028,80028,6201.36
951579---
103010028,80021,5951.17
1161571---
123010028,80021,8341.26
a Calculated according to the monomer conversion: Mn(calc) = Mw(ε-CL) × ([ε-CL]0/[cat]) × (conversion) or Mn(calc) = Mw(LA) × ([LA]0/[cat]) × (conversion). b The molecular weights and the PDI of the polymers were determined by GPC relative to polystyrene standards and multiplied by a correction factor of 0.58 for polylactide and 0.56 polycaprolactone.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mankaev, B.N.; Hasanova, L.F.; Churakov, A.V.; Egorov, M.P.; Karlov, S.S. Gallium (III) Complexes Based on Aminobisphenolate Ligands: Extremely High Active ROP-Initiators from Well-Known and Easily Accessible Compounds. Int. J. Mol. Sci. 2022, 23, 15649. https://doi.org/10.3390/ijms232415649

AMA Style

Mankaev BN, Hasanova LF, Churakov AV, Egorov MP, Karlov SS. Gallium (III) Complexes Based on Aminobisphenolate Ligands: Extremely High Active ROP-Initiators from Well-Known and Easily Accessible Compounds. International Journal of Molecular Sciences. 2022; 23(24):15649. https://doi.org/10.3390/ijms232415649

Chicago/Turabian Style

Mankaev, Badma N., Leyla F. Hasanova, Andrei V. Churakov, Mikhail P. Egorov, and Sergey S. Karlov. 2022. "Gallium (III) Complexes Based on Aminobisphenolate Ligands: Extremely High Active ROP-Initiators from Well-Known and Easily Accessible Compounds" International Journal of Molecular Sciences 23, no. 24: 15649. https://doi.org/10.3390/ijms232415649

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