Home A tetranuclear Sn(IV) 4-thiazolecarboxylate complex: synthesis, structure and catalytic behavior in the bulk ring-opening polymerization of glycolide
Article Publicly Available

A tetranuclear Sn(IV) 4-thiazolecarboxylate complex: synthesis, structure and catalytic behavior in the bulk ring-opening polymerization of glycolide

  • Jian Gao , Yu Yuan , Ai-Jun Cui , Feng Tian , Ming-Yang He and Qun Chen EMAIL logo
Published/Copyright: July 2, 2016
Become an author with De Gruyter Brill

Abstract

The reaction of 4-thiazolecarboxylic acid (Htzc) and dimethyltin(IV) dichloride with NaOH in mixed MeOH-H2O solvent led to the formation of a new Sn(IV) complex (Me2Sn)4(μ3-O)2(tzc)4 (1). Its structure has been characterized by elemental analysis, IR spectroscopy and single crystal and powder X-ray diffraction. Single-crystal X-ray diffraction revealed that complex 1 crystallizes in the monoclinic P21/c space group with Z = 2 and has a tetranuclear structure with crystallographically imposed centrosymmetry. The as-synthesized complex 1 was found to be active toward the bulk solvent-free polymerization of glycolide, producing poly(glycolic acid) with a number-average molecular weight up to 55.5 kDa.

1 Introduction

Biodegradable polymers have recently attracted much interest as replacements for conventional synthetic materials [16]. The most promising candidates in this field are aliphatic polyesters such as poly(lactide) and poly(glycolide) (PGA). In particular, PGA is the simplest poly(α-hydroxyl acid) and the first polyester used in biomedical applications, as surgical sutures Dexon [7]. Ring-opening polymerization (ROP) of glycolide initiated by metal-based complexes has been proven to be the most convenient route for the synthesis of PGA [6]. The ROP process can take place in solution polymerization or in bulk solvent-free melt polymerization. In contrast to solution polymerizations, the ROP of glycolide is carried out with the bulk solvent-free melt due to the apparent advantages such as no solvent, high purity, the minimization of undesired side reactions and the large-scale production. As for the needed catalysts toward the ROP of glycolide, various complexes of main group, transition and rare earth metals were found to be active proceeding via a coordination-insertion mechanism [811]. Despite the significant progress achieved, a current trend in the ROP of glycolide is the search for alternative catalysts that could have beneficial effects.

In the past few decades, a number of N,N-, N,O-, N,P- and O,S-chelating bidentate ligands have been widely employed to construct complexes because of attractive features such as the ease of preparation and tenability of steric and electronic properties [1215]. Furthermore, these complexes have shown efficient catalytic activities toward the ROP of cyclic esters. As a N,O-chelating ligand, 4-thiazolecarboxylic acid (4-Htzc) might be a good candidate for construction of complexes with diverse structures. Very recently, Peruzzini and co-workers have initially reported two mononuclear Cu(II) and Zn(II) complexes of 4-tzc [16]. For this contribution, a new tetranuclear Sn(IV) complex (Me2Sn)4(μ3-O)2(tzc)4 (1) was synthesized by the reaction of 4-Htzc with dimethyltin(IV) dichloride in MeOH-H2O mixed solvent. Moreover, the bulk ROP of glycolide initiated by 1 was investigated for the first time. Here we would like to report these results.

2 Results and discussion

2.1 Synthesis and general characterization

Among the large variety of metal complexes investigated as initiators, tin complexes have shown excellent catalytic activity for ROP of cyclic esters to produce high molecular weight polymers with narrow molecular weight distribution [1720]. For instance, the industrially most commonly used catalyst for the preparation of PGA is stannous octanoate [19]. In this study, we selected the chelating-type ligand [4-tzc] to react with various tin salts including tin(II) dichloride, tin(II) oxalate, tin(II) trifluoromethanesulfonate, triphenyltin(IV) chloride and dimethyltin(IV) dichloride. Single crystals of 1 were obtained from the reaction of 4-Htzc and dimethyltin(IV) dichloride with NaOH in MeOH-H2O mixed solvent under ambient conditions, while only white precipitates or microcrystalline products were afforded in the other cases. Complex 1 is stable to air and moisture and insoluble in common solvents such as water, dimethyl sulphoxide, alcohol, acetonitrile, chloroform and acetone. The insolubility prevented the characterization of complex 1 by solution NMR. The elemental analysis of 1 matches well with the molecular formula as evaluated by X-ray diffraction analysis (below). In the IR spectrum of 1, the antisymmetric and symmetric carboxylate stretching vibrations are found in the range of 1590–1665 cm−1 and 1400–1440 cm−1, respectively. The absence of the characteristic band ~1680 cm−1 for the free 4-Htzc molecule indicates the complete deprotonation of the carboxyl group, which is also consistent with the crystal structure as described below.

The solid-state 13C NMR spectrum of complex 1 (Fig. 1) shows two characteristic peaks of the carboxylate carbon atoms at 161.7 and 166.1 ppm. Signals of the methyl carbon atoms are observed in the range of 1.9–19.1 ppm, while the resonances of the carbon atoms of the triazole ring appear at 124.6–156.1 ppm.

Fig. 1: The solid-state 13C NMR spectrum of complex 1.
Fig. 1:

The solid-state 13C NMR spectrum of complex 1.

2.2 Description of the crystal and molecular structure

Single-crystal X-ray diffraction analysis reveals that complex 1 crystallizes in the monoclinic space group P21/c with Z = 2. It shows a tetranuclear structure with crystallographically imposed centrosymmetry. A perspective view of the molecular structure of 1 is shown in Fig. 2a. There are two crystallographically independent Sn(IV) ions (Sn1 and Sn2). The Sn1 ion exhibits a distorted pentagonal bipyramidal geometry, which is completed by seven atoms including four oxygen atoms (O2, O3, O5 and O5#1) from two tzc anions and two bridging O2– ligands as well as one nitrogen atom (N2) from one tzc anion in the equatorial plane, and two methyl groups in apical positions. The Sn–O bond distances vary from 2.103(3) to 2.479(4) Å, and the Sn–C bond distances are 2.102(6) and 2.106(6) Å, while the Sn–N bond distance of 2.769 (6) Å is similar to those reported for other tetranuclear Sn(IV) complexes [21]. However, the Sn2 ion is five-coordinated by three oxygen atoms (O1, O3#1 and O5) from two tzc anions and one O2– ligand with the Sn–O bond distances in the range of 1.981(3)–2.207(4) Å and two carbon atoms (C11 and C12) with the Sn–C bond distances of 2.094(8) and 2.098(7) Å, respectively. The Addison parameter τ of 0.66 suggests that the coordination sphere around the Sn2 atom is properly considered as a distorted trigonal bipyramid.

Fig. 2: Views of (a) the molecular structure of 1 (symmetry code: #1, – x + 1, – y, – z + 2) and (b) the 2D supramolecular architecture (C4–H4···O2 hydrogen bonds are shown as dashed lines; other hydrogen atoms are omitted for clarity).
Fig. 2:

Views of (a) the molecular structure of 1 (symmetry code: #1, – x + 1, – y, – z + 2) and (b) the 2D supramolecular architecture (C4–H4···O2 hydrogen bonds are shown as dashed lines; other hydrogen atoms are omitted for clarity).

The tzc anions display two types of coordination modes, where one adopts a chelating/bridging coordination mode via one nitrogen atom and one carboxylate oxygen atom to connect the Sn1 and Sn2#1 atoms, while the other bridges the Sn1 and Sn2 atoms by the carboxylate group in a μ211-syn-syn-bridging fashion. The atoms O2− atom O5 are located in the midpoint of the three Sn(IV) atoms (Sn1, Sn1#1 and Sn2) and adopt a μ3-bridging mode with Sn–O–Sn angles of 104.6(1)°, 117.7(1)° and 137.4(2)°. The sum of these angles is 359.7°, indicating almost perfect planarity of these oxygen atoms. Two pairs of symmetry-related Sn(IV) atoms are linked by four tzc ligands, forming the tetranuclear molecule with the shortest Sn···Sn distance being 3.383(1) Å.

As is shown in Fig. 2b, adjacent tetranuclear units are connected by weak C4–H4···O2i hydrogen bond interactions (H···O/C···O distances: 2.616/3.492 Å, angle: 157°, i = x, – y + 1/2, z – 1/2) between thiazole rings and oxygen atoms of the μ211-syn-syn-bridging carboxylate group to generate a 2D supramolecular layer along the bc plane. Furthermore, the uncoordinated carboxylate oxygen atoms are H-bonded to neighboring layers via C3–H3···O4ii hydrogen bonds g (H···O/C···O distances: 2.38/3.016 Å, angle: 125°, i = x – 1, y, z – 1), resulting in a 3D supramolecular architecture.

2.3 Powder X-ray diffraction and thermal stability analyses

The crystalline phase purity of the bulk samples of 1 was confirmed by powder X-ray diffraction (PXRD) (see Fig. 3). The experimental PXRD pattern matches with the simulated one obtained from single-crystal diffraction data, indicative of pure products. To estimate the thermal stability of 1, a thermogravimetric analysis (TGA) of a crystalline sample was performed from room temperature to 800°C under nitrogen atmosphere. As depicted in Fig. 4, complex 1 is thermally stable to ca. 260°C where the weight loss starts to end at ca. 450°C. Furthermore, to verify the stability of complex 1 at the reaction temperature of ROP, PXRD experiments for the as-synthesized sample of 1 were carried out at different temperatures (150°C, 200°C and 230°C). The results indicate that even the PXRD pattern at 230°C is still in agreement with the experimental pattern at room temperature.

Fig. 3: XRPD patterns for 1 as simulated (black) from the single-crystal data and as-synthesized 1 at 25°C (red), 150°C (blue), 200°C (magenta) and 230°C (green).
Fig. 3:

XRPD patterns for 1 as simulated (black) from the single-crystal data and as-synthesized 1 at 25°C (red), 150°C (blue), 200°C (magenta) and 230°C (green).

Fig. 4: TGA curve of complex 1.
Fig. 4:

TGA curve of complex 1.

2.4 Ring-opening polymerization of glycolide

The solid-state bulk polymerization is the key for the fabrication of high molecular weight PGA without any organic solvents, where the melt/solid polycondensation of glycolide generally occurs in the temperature of 180–230°C [22]. Such a reaction process is very fast, and PGA samples can be analyzed by isothermal differential scanning calorimeter (DSC). In this work, the bulk solvent-free melt ROP of glycolide initiated by the as-synthesized complex 1 was studied via the DSC method. The data such as the reaction rate constant, activation energy and reaction order were determined in a similar way as described in our previous work [23]. The representative polymerization data are summarized in Table 1. It was found that complex 1 could initiate the ROP of glycolide with kinetic reaction orders ranging from 0.32 to 0.48 in the range 200–220°C. The dependence of the reaction rate constant on the reaction temperature revealed that the bulk polymerization is relatively faster at higher temperature, which is also consistent with the enthalpy changes. Furthermore, by investigating the relationship between conversion and reaction time at 210°C (Fig. 5), the conversion of glycolide could be completed after 3.5 min. Because the bulk polymerization can afford high molecular weight polyesters, the polymerization of glycolide initiated by complex 1 was carried out at 210°C. It was found that the molecular weight (Mn) of the resultant PGA is 55.5 kDa with the molecular weight distribution (Mw/Mn) of 1.65. On the basis of the above experimental results and previous literature [2429], one possible reaction mechanism for the present ROP of glycolide may be the coordination-insertion mechanism. The monomer glycolide is coordinated to the Sn(IV) center in the first stage, forming an alkoxyl tin(IV) species [6]. Then, the chain propagation process occurs through a coordination-insertion process.

Table 1:

DSC data for the polymerization of glycolide initiated by complex 1.a

Temperature (°C)ln(1–α)bnbKb (S−1 × 10−3)Eab (kJ mol−1)
200–1.240.383.03814.57 ± 1.46
205–1.340.333.169
210–1.330.403.357
215–1.380.373.731
220–1.330.484.394

aEach reaction was performed in the melt at a different reaction temperature with a molar ratio of glycolide to Sn(IV) complex of 4000:1.

bThe data were determined by DSC measurement.

Fig. 5: Plot of conversion of glycolide versus reaction time with 1 as an initiator at 210°C.
Fig. 5:

Plot of conversion of glycolide versus reaction time with 1 as an initiator at 210°C.

In summary, a new tetranuclear Sn(IV) 4-thiazolecarboxylate complex was synthesized and structurally characterized. The complex can promote the bulk ROP of glycolide to afford the PGA with high molecular weight. The further work on the design and synthesis of more highly active, low cost and less toxic complexes as initiators toward ROP of glycolide is proceeding in our laboratory.

3 Experimental section

All reagents and solvents for synthesis were commercially available and used without further purification. The Fourier transform (FT)-IR spectra (KBr pellets) were recorded on a Nicolet ESP 460 FT-IR spectrometer in the range of 4000–400 cm−1. Elemental analyses were performed with a PE-2400II (Perkin-Elmer) instrument. The solid-state 13C NMR spectrum was obtained on a Bruker ADVANCE III 400WB spectrometer operated at 9.4 T with a frequency of 100.6 MHz. The calculated PXRD patterns were obtained from the single-crystal diffraction data using the PLATON software [30]. TGA experiments were carried out on a Dupont thermal analyzer from room temperature to 800°C (heating rate of 10°C min−1, nitrogen stream).

3.1 Synthesis of (Me2Sn)4(μ3-O)2(tzc)4 (1)

A mixture of dimethyltin(IV) dichloride (43.9 mg, 0.2 mmol), Htzc (25.8 mg, 0.2 mmol) and NaOH (8.0 mg, 0.2 mmol) was dissolved in MeOH-H2O (v/v = 4:1, 6 mL) mixed solvent with stirring for 30 min. The resulting solution was filtrated and then left to stand at room temperature. Colorless needle-like crystals of 1 were isolated in ca. 42% yield (23.9 mg, based on Htzc). – Anal. for C24H32N4O10S4Sn4 (%): calcd. C 29.29, H 2.83, N 4.92; found, C 30.08, H 2.81, N 4.83. – IR (cm−1, KBr pellet): v = 3138 (m), 3119 (w), 3076 (w), 3042 (m), 2917 (w), 1665 (s), 1593 (vs), 1505 (m), 1492 (m), 1436 (s), 1419 (s), 1379 (m), 1331 (s), 1308 (vs), 1220 (m), 1196 (w), 1095 (w), 946 (m), 883 (m), 1331 (s), 1308 (vs), 1220 (m), 1196 (w), 1095 (w), 946 (m), 883 (m), 856 (m), 843 (m), 780 (s), 755 (m), 663 (s), 623 (m), 583 (m), 566 (m), 520 (m), 487 (s), 431 (m). – 13C NMR (100.6 MHz, 9.4 T, solid state): δ = 1.9, 3.4, 18.5, 19.1 (all methyl carbon atoms), 124.6, 131.1, 151.5, 151.5, 156.1, 156.1 (all carbon atoms on the thiazole rings), 161.7, 166.1 (carboxylate carbon atoms).

3.2 Glycolide polymerization procedure

The isothermal experiments of glycolide polymerization were performed on a Perkin-Elmer DSC under a nitrogen atmosphere (50 mL min−1), and the DSC data were analyzed by Pyris Kinetic Analysis software. Indium (156.6°C) was employed as a standard sample for the calibration of temperature and heat. A mixture of glycolide (10 g) and complex 1 (24.6 mg) was grinded for 30 min at room temperature, and a sample (5 mg) was taken for DSC measurements. The experiments were run in aluminum pans. The samples were heated at a heating rate of 600°C min−1 to the temperature in the range of 200–220°C, and then kept for 15 min.

Molecular weights (Mn and Mw) and molecular weight dispersities (Mw/Mn) were measured by gel permeation chromatography. The measurements were performed at 40°C on a Waters 1525 binary system equipped with a Waters 2414 Refractive Index detector and a Waters 2487 dual λ absorption (UV, λabs = 220 nm) detector. In the case of the analyses performed using a solution of sodium trifluoroacetate (0.68 g, 5 mmol) in 1,1,1,3,3,3-hexafluoro-2-propanol (1000 mL) as eluent at a flow rate of 0.6 mL min−1, a system of four Styragel HR columns (7.8 × 300 mm; range 103–106 Å) was employed. The molecular weights were calculated with respect to poly(methyl methacrylate) standards (Mn ranging from 7000 to 200000).

3.3 X-ray structure determination

The single-crystal X-ray diffraction measurement was performed on a Bruker Apex II CCD diffractometer at ambient temperature with MoKα radiation (λ = 0.71073 Å). A semiempirical absorption correction was applied using Sadabs [31], and the program Saint was used for integration of the diffraction profiles [32]. The structure was solved by Direct Methods using Shelxs of the Shelxtl program package and refined anisotropically for all non-H atoms by full-matrix least squares on F2 with Shelxl [3336]. In general, hydrogen atoms were located geometrically and allowed to ride during the subsequent refinement. Crystallographic data and numbers pertinent to data collection and structure refinement are summarized in Table 2. Selected bond lengths and angles are listed in Table 3.

Table 2:

Crystal structure data for 1.

FormulaC24H32N4O10S4Sn4
Mr1139.53
Crystal size, mm30.22 × 0.20 × 0.20
Crystal systemMonoclinic
Space groupP21/c
a, Å11.299(2)
b, Å20.880(2)
c, Å8.576(2)
β, deg104.11(1)
V, Å31962.5(3)
Z2
Dcalcd, g cm−31.9
μ(MoKα), cm−12.8
F(000), e1096
hkl range–13 ≤ h ≤ +12
–24 ≤ k ≤ +24
–10 ≤ l ≤ +10
Refl. measured10696
Refl. unique/Rint3453/0.040
Param. refined/restraints212/0
Ra/Rwb0.0258/0.1023
GoF (F2)c1.014
Δρfin (max/min), e Å−31.09/–1.32

aR1 = Σ||Fo|–|Fc||/Σ|Fo|.

bwR2 = [Σw(Fo2Fc2)2w(Fo2)2]1/2, w = [σ2(Fo2)+(AP)2+BP]−1, where P = (Max(Fo2, 0)+2Fc2)/3.

cGoF = S = [Σw(Fo2Fc2)2/(nobsnparam)]1/2.

Table 3:

Selected bond lengths (Å) and angles (deg) for 1 with estimated standard deviations in parentheses.a

Distances
Sn1–O22.379(4)Sn1–O32.479(4)
Sn1–O52.103(3)Sn1–O5#12.172(3)
Sn1–N22.769(1)Sn1–C92.106(6)
Sn1–C102.102(6)Sn2–O12.176(4)
Sn2–O3#22.207(4)Sn2–O51.981(3)
Sn2–C112.098(7)Sn2–C122.094(8)
Angles
O2–Sn1–O3131.6(1)O2–Sn1–O585.8(2)
O2–Sn1–O5#1160.9(1)O2–Sn1–N270.8(1)
O2–Sn1–C987.8(2)O2–Sn1–C1084.2(2)
O3–Sn1–O5142.5(1)O3–Sn1–O5#167.1(2)
O3–Sn1–N260.8(1)O3–Sn1–C986.1(2)
O3–Sn1–C1086.5(2)O5–Sn1–O5#175.3(2)
O5–Sn1–N2156.5(1)O5–Sn1–C997.8(2)
O5–Sn1–C1098.6(2)O5#1–Sn1–N2128.0(1)
O5#1–Sn1–C997.9(2)O5#1–Sn1–C1095.2(2)
N2–Sn1–C979.6(2)N2–Sn1–C1081.5(2)
C9–Sn1–C10161.1(3)O1–Sn2–O3#1169.6(2)
O1–Sn2–O594.2(2)O1–Sn2–C1189.0(2)
O1–Sn2–C1292.0(3)O3#1–Sn2–O576.0(1)
O3#1–Sn2–C1192.2(2)O3#1–Sn2–C1295.1(3)
O5–Sn2–C11116.8(3)O5–Sn2–C12113.3(3)
C11–Sn2–C12129.6(4)Sn1–O5–Sn2137.4(2)
Sn1–O5–Sn1#1104.6(1)Sn2–O5–Sn1#1117.7(1)

aSymmetry code: #1: –x + 1, –y, –z + 2.

CCDC 1441860 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif.

Award Identifier / Grant number: 21201026

Award Identifier / Grant number: BK20131142

Funding statement: We gratefully acknowledge financial support by the National Natural Science Foundation of China (21201026), and a project funded by the Priority Academic Program Development of Jiangsu Higher Education Institutions (PAPD) and the Nature Science Foundation of Jiangsu Province (BK20131142).

Acknowledgments

We gratefully acknowledge financial support by the National Natural Science Foundation of China (21201026), and a project funded by the Priority Academic Program Development of Jiangsu Higher Education Institutions (PAPD) and the Nature Science Foundation of Jiangsu Province (BK20131142).

References

[1] A. J. Ragauskas, C. K. Williams, B. H. Davison, T. Techaplinski, Science2006, 311, 484.10.1126/science.1114736Search in Google Scholar PubMed

[2] C. K. Williams, M. A. Hillmyer, Polym. Rev.2008, 48, 1.10.1080/15583720701834133Search in Google Scholar

[3] R. H. Platel, L. M. Hodgson, C. K. Williams, Polym. Rev.2008, 48, 11.10.1080/15583720701834166Search in Google Scholar

[4] S. Inkinen, M. Hakkarainen, A. C. Albertsson, A. Sodergard, Biomacromolecules2011, 12, 523.10.1021/bm101302tSearch in Google Scholar PubMed

[5] S. Mecking, Angew. Chem. Int. Ed.2004, 43, 1078.10.1002/anie.200301655Search in Google Scholar PubMed

[6] O. Dechy, B. Martin, D. Bourissou, Chem. Rev.2004, 104, 6147.10.1021/cr040002sSearch in Google Scholar PubMed

[7] E. E. Schmitt, R. A. Polistina, US 3097033, 1967.Search in Google Scholar

[8] A. K. Sutar, T. Maharana, S. Dutta, C. T. Chen, C. C. Lin, Chem. Soc. Rev.2010, 39, 1724.10.1039/B912806ASearch in Google Scholar PubMed

[9] T. K. Saha, V. Ramkumar, D. Chakraborty, Inorg. Chem.2011, 50, 2720.10.1021/ic1025262Search in Google Scholar PubMed

[10] A. D. Schwarz, K. R. Herbert, C. Paniagua, P. Mountford, Organometallics2010, 29, 4171.10.1021/om100508jSearch in Google Scholar

[11] M. Labet, W. Thielemans, Chem. Soc. Rev.2009, 38, 3484.10.1039/b820162pSearch in Google Scholar PubMed

[12] F. T. Edelmann, Chem. Soc. Rev.2009, 38, 2253.10.1039/b800100fSearch in Google Scholar

[13] K. Yang, H. Fu, Q. J. Song, M. Zhang, Y. Q. Ding, Organometallics2011, 30, 77.10.1021/om100714gSearch in Google Scholar

[14] P. W. Roesky, M. T. Gamer, M. Puchner, A. Greiner, Chem. Eur. J.2002, 8, 5265.10.1002/1521-3765(20021115)8:22<5265::AID-CHEM5265>3.0.CO;2-VSearch in Google Scholar

[15] F. D. Monica, E. Luciano, G. Roviello, A. Grassi, S. Milione, C. Capacchione, Macromolecules2014, 47, 2830.10.1021/ma5003358Search in Google Scholar

[16] A. Rossin, B. D. Credico, G. Giambastiani, L. Gonsalvi, M. Peruzzini, G. Reginato, Eru. J. Inorg. Chem.2011, 4, 539.10.1002/ejic.201000928Search in Google Scholar

[17] S. Nagase, Acc. Chem. Res.1995, 28, 469.10.1021/ar00059a005Search in Google Scholar

[18] J. L. Gorczynski, J. Chen, C. L. Fraser, J. Am. Chem. Soc.2005, 127, 14956.10.1021/ja0530927Search in Google Scholar

[19] A. Kowalski, A. Duda, S. Penczek, Macromolecules2000, 33, 7359.10.1021/ma000125oSearch in Google Scholar

[20] L. M. Pitet, S. B. Hait, T. J. Lanyk, D. M. Knauss, Macromolecules2007, 40, 2327.10.1021/ma0618068Search in Google Scholar

[21] M. N. Xanthopoulou, S. K. Hadjikakou, N. Hadjiliadis, M. Kubicki, S. Skoulika, T. Bakas, M. Baril, I. S. Butler, Inorg. Chem.2007, 46, 1187.10.1021/ic061601fSearch in Google Scholar

[22] K. Takahashi, I. Taniguchi, M. Miyamoto, Y. Kimura, Polymer2000, 41, 8725.10.1016/S0032-3861(00)00282-2Search in Google Scholar

[23] S.-C. Chen, A.-Q. Dai, K.-L. Huang, Z.-H. Zhang, A.-J. Cui, M.-Y. He, Q. Chen, Dalton Trans.2016, 45, 3577.10.1039/C5DT04687DSearch in Google Scholar

[24] H. R. Kricheldorf, M. Berl, N. Scharnagl, Macromolecules1988, 21, 286.10.1021/ma00180a002Search in Google Scholar

[25] M. Nishiura, Z. Hou, T. A. Koizumi, T. Imamoto, Y. Wakatsuki, Macromolecules1999, 32, 8245.10.1021/ma990101lSearch in Google Scholar

[26] H. R. Kricheldorf, I. Kreiser-Saunders, Polymer2000, 41, 3957.10.1016/S0032-3861(99)00606-0Search in Google Scholar

[27] C. K. Williams, L. E. Breyfogle, S. K. Choi, W. Nam, V. G. Young, Jr., M. A. Hillmyer, W. B. Tolman, J. Am. Chem. Soc.2003, 125, 11350.10.1021/ja0359512Search in Google Scholar PubMed

[28] D. Pappalardo, L. Annunziata, C. Pellecchia, Macromolecules2009, 42, 6056.10.1021/ma9010439Search in Google Scholar

[29] D. J. Darensbourg, O. Karroonnirun, Macromolecules2010, 43, 8880.10.1021/ma101784ySearch in Google Scholar

[30] A. L. Spek, PLATON, A Multipurpose Crystallographic Tool, Utrecht University, Utrecht (The Netherlands) 2002.Search in Google Scholar

[31] G. M. Sheldrick, SADABS, Program for Empirical Absorption Correction of Area Detector Data, University of Göttingen (Germany) 2002.Search in Google Scholar

[32] SAINT, Software Reference Manual, Bruker Analytical X-ray Instruments Inc., Madison, WI (USA) 1998.Search in Google Scholar

[33] G. M. Sheldrick, SHELXTL NT (Version 5.1), Bruker Analytical X-ray Instruments Inc., Madison, WI (USA) 2001.Search in Google Scholar

[34] G. M. Sheldrick, SHELXS/L-97, Programs for Crystal Structure Determination, University of Göttingen, Göttingen (Germany) 1997.Search in Google Scholar

[35] G. M. Sheldrick, Acta Crystallogr.1990, A46, 467.10.1107/S0108767390000277Search in Google Scholar

[36] G. M. Sheldrick, Acta Crystallogr.2008, A64, 112.10.1107/S0108767307043930Search in Google Scholar PubMed

Received: 2015-12-13
Accepted: 2016-2-18
Published Online: 2016-7-2
Published in Print: 2016-8-1

©2016 by De Gruyter

Downloaded on 14.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/znb-2015-0211/html
Scroll to top button