Next Article in Journal
Photocatalytic Study of Cyanide Oxidation Using Titanium Dioxide (TiO2)-Activated Carbon Composites in a Continuous Flow Photo-Reactor
Previous Article in Journal
Visible-Light Radical–Radical Coupling vs. Radical Addition: Disentangling a Mechanistic Knot
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Naturally Occurring Oxazole Structural Units as Ligands of Vanadium Catalysts for Ethylene-Norbornene (Co)polymerization

by
Wioletta Ochędzan-Siodłak
1,*,
Dawid Siodłak
1,
Karolina Banaś
1,
Katarzyna Halikowska
1,
Sławomir Wierzba
2 and
Karel Doležal
3,4
1
Institute of Chemistry, University of Opole, Oleska 48, 45-052 Opole, Poland
2
Institute of Environmental Engineering and Biotechnology, University of Opole, Kard. Kominka 6a, 45-040 Opole, Poland
3
Department of Chemical Biology, Faculty of Science, Palacký University, Slechtitelu 27, CZ-78371 Olomouc-Holice, Czech Republic
4
Laboratory of Growth Regulators, Faculty of Science, Palacký University, Slechtitelu 27, CZ-78371 Olomouc-Holice, Czech Republic
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(8), 923; https://doi.org/10.3390/catal11080923
Submission received: 6 July 2021 / Revised: 26 July 2021 / Accepted: 28 July 2021 / Published: 29 July 2021
(This article belongs to the Section Catalysis in Organic and Polymer Chemistry)

Abstract

:
1,3-Oxazole and 4,5-dihydro-1,3-oxazole are common structural motifs in naturally occurring peptides. A series of vanadium complexes were synthesized using VCl3(THF)3 and methyl substituted (4,5-dihydro-1,3-oxazol-2-yl)-1,3-oxazoles as ligands and analyzed using NMR and MS methods. The complexes were found to be active catalysts both in ethylene polymerization and ethylene-norbornene copolymerization. The position of methyl substituent in the ligand has considerable impact on the performance of (co)polymerization reaction, as well as on the microstructure, and thus physical properties of the obtained copolymers.

Graphical Abstract

1. Introduction

Catalysts reactions play an irreplaceable role in the synthesis of organic compounds, which are useful as crucial chemicals, pharmaceuticals, and advanced polymers [1]. Transition metal catalysis is one of the foremost processes in production of polymer materials, mostly in olefin polymerization [2]. Polyolefins constitute more than half of global synthetic polymers, which results in a continuous search for new catalysts, composed of various transition metals and ligands. Amongst the transition metal tested, vanadium attracts considerable interest as vanadium-based catalysts display advantageous features. They enable to produce high molecular weight polyethylene, as well as copolymers with higher comonomer incorporation of alpha-olefins [3,4,5]. In particular, cyclic olefin copolymers (COCs) are a promising group of materials with specific, projectable properties, which can be also produced using vanadium catalysts [6,7]. In addition to the transition metal, ligands are also important because they influence both activity of catalysts and properties of copolymers. In particular, N-donor ligands are crucial in many recent developments [2].
In our previous works, the vanadium catalysts with 2-(1,3-oxazolin-2-yl)pyridine (Py-ox) and 2,6-bis(1,3-oxazolin-2-yl)pyridine (Py-box) ligands were investigated in ethylene polymerization and ethylene-norbornene (co)polymerization and the results were promising [8]. The Py-ox and Py-box ligands are commonly known to form stable transition metal complexes [9]. However, it is worth noticing that the pyridine-oxazole structural motif can be found in natural compounds such as geninthiocin [10], berninamycins [11], sulfomycins [12], radamycin [13], or TP-1161 [14]. Following the idea that nature is the source of promising structural units, potentially useful as ligands, we studied the vanadium catalysts with 2-hydroxyphenyloxazoline ligands, and we have proved that these catalysts are highly active in ethylene and ethylene-norbornene (co)polymerization ((7300 and 5300 kgPE/molV·h, respectively) [15]. The 2-hydroxyphenyl-oxazole/oxazoline structural motif is present in natural peptides, such as in Mycobactin T [16] or Amamistatin A [17]. Interestingly, the oxazoline residue is the key feature of the mycobactin structure, which acts as iron-chelating components and is responsible for microbial iron assimilation processes [18].
Encouraged by the previous results, we focused on new oxazole structural motifs. The oxazole rings, especially 1,3-oxazole and 4,5-dihydro-1,3-oxazole, are common structural motifs in naturally occurring peptides, because they can be easily obtained from the serine or threonine residues, when the hydroxyl group of side chain reacts with the preceding peptide group in the main chain by cyclodehydration, and further, by dehydrogenation processes [19]. The oxazole rings joined to each other can be found for example, in muscoride A [20], hennoxazole A [21], and diazonamides [22]. Moreover, the method of biosynthesis enables to obtain larger structures consisting of more than two oxazole rings, as in the cases of plantazolicins [23], ulapualide A [24], azolemycin [25], or telomestatin [26].
In order to check the potential usefulness of these structures, we have decided to study a relatively small structural motif, (4,5-dihydro-1,3-oxazol-2-yl)-1,3-oxazole (Figure 1). These heterocyclic rings, commonly described as oxazoline and oxazole, contain nitrogen atom in sp2 hybridization and electron pair in the plane of the ring, which can act as nucleophiles, and, thus, N-donors. The route of synthesis is also promising as it requires rather simple and commonly accessible substrates. The presence of methyl group can test possible differences in performance of catalysts. It also mimics the synthesis from threonine/serine side chains.

2. Results and Discussion

2.1. Synthesis and Analysis of the Catalysts

The ligands synthesis was a considerable modification of the procedures described in literature [9,27]. In a three-step procedure, methyl 2-methyl-1,3-oxazole-5-carboxylate (L1, L2) or 5-methyl-1,3-oxazole-5-carboxylate (L3) reacts with the appropriate amino alcohol, 2-aminoethanol (L1, L3) or (±)-2-amino-1-propanol (L2) to form the intermediate N-(2-hydroxyethyl)-carboxyamide, which is then converted with SOCl2 to the hydrochloride. Finally, N-(2-chloroethyl)-carboxyamide reacts with NaH to the final product (L1–L3). The detailed procedures are given in Supplementary Materials (Figures S1–S13).
The vanadium complexes, denoted as L1-V, L2-V, and L3-V, were synthesized using VCl3(THF)3 precursor and the proper ligand L1–3 in methylene dichloride solution using Schlenk system. The complexation reaction were confirmed by both the NMR and MS spectra of complexes (Figures S14–S19). There is a considerable shift in position of the signals, in particular for 4,5-dihydro-1,3-oxazole (oxazoline) ring. In 13C NMR spectra Δδ of C2 carbon is shifted 4 ppm downfield whereas C5 (C-O) and C4 (C-N) methylene carbons are shifted about 23 and 13 ppm up-field, respectively. In the 1H NMR spectra of complexes the shift of signals of methylene group of about Δδ 0.5 ppm downfield is observed. Additionally, the multiplicity of these signals are diffused due to partial presence of paramagnetic oxidation state of vanadium element. The MS spectra of complexes show the ligand fragment at m/z = 152 (L1), 166 (L2), and 152 (L3) (Figures S6, S11, S13, S15, S17, and S19). There are also characteristic groups of ions at m/z = 323, 325, and 327, regardless of the complex determined, which correspond to the formula C7H8Cl3N2O3V (LVCl3O). In the case of complex with the ligand L2 this can be explained by the loss of additional methyl group. This indicates the presence of ligand and chlorine atoms, and suggests that the structure of complexes resembles that determined for metal-organic framework stabilized vanadium catalyst for olefin polymerization [28]. Efforts to obtain crystal form of the complexes suitable for X-ray analysis were not successful.

2.2. Polymerization and Copolymerization

The obtained vanadium complexes, in the presence of activator AlEt2Cl and ethyl trichloroacetate re-activator, were found to be active as catalysts both in ethylene homo-polymerization and ethylene-norbornene co-polymerization (Table 1). In the ethylene polymerization, the most effective turned out to be the catalyst L3-V resulting up to 4800 kgPE/molV·h. Interestingly, in the L3 ligand, the methyl group is at the position 5 of the 1,3-oxazole ring, which imposes a steric hindrance between the heterocyclic rings and prevents their co-planarity. In contrast, the lowest activity reveals the catalyst L2-V with substituted oxazoline ring (3150 kgPE/molV·h).
The same catalyst systems, at similar reaction conditions, were applied in copolymerization of ethylene with norbornene (NB). The activity of copolymerization reaction considerably decreases as compared to the ethylene polymerization. The concentration of norbornene was relatively low in the reaction medium (0.5–1.5 mol/dm3). Nevertheless, the NB incorporation in the range from 11 to 32 mol% was obtained, and increases with the increase in norbornene concentration. Again, differences amongst the catalysts activity are observed. The catalyst L2-V produces the copolymer with the highest NB content (32.4 mol%), even at the lowest NB concentration in the reaction medium. The catalysts L1-V and L3-V reveal similar level of NB incorporation, with the lowest incorporation 11.4 mol% at the NB concentration 0.5 mol/dm3 for the catalyst L3-V.

2.3. Polymer Properties

The polyethylene samples obtained using the studied vanadium complexes were subjected to thermal analysis DSC, to determined melting point (Tm) and crystallinity degree (X). After the second heating cycle, the Tm of polyethylene was similar (136.2 °C, 138.3 °C, and 135.3 °C), as was their crystallinity (52.3%, 54.6%, 63.5%) (Figures S21, S25 and S29). Based on the results of average molecular weight (Mw) and the molecular weight distribution (Mw/Mn) determined by the GPC method, it was found that polyethylene samples are characterized by high homogeneity (Mw/Mn~2.0) and the Mw does not exceed 900 × 103 g/mol. It is worth noting that the PE obtained using the catalyst L2-V was characterized by slightly higher values of all the parameters determined (Table 1).
The obtained copolymers have lower molecular weight than polyethylene (Mw, 180–410 × 103 g/mol). The molecular weight distribution is typical for vanadium catalysts (Mw/Mn, 2.2–2.8). The copolymers Mw decreases with increase in incorporation of norbornene units in copolymer chain (Table 1). DSC analysis (Figures S22–S24, S26–S28, S30–S32) shows that copolymers are amorphous materials with a glass transition temperature (Tg) that increases with increasing NB content. Differences amongst the copolymers produced using the studied catalysts are observed (Figure 2). Interestingly, the DSC curves of the copolymers prepared by the complex L1-V show both the glass transition temperature (Tg) and the melting point (Tm). Low melting peaks observed indicates relatively long ethylene sequences. The melting points (128 °C) of these copolymers are lower than those of polyethylene. This can be attributed to the long PE sequences in the copolymer containing small amounts of isolated NB units [29]. In contrast, the copolymers obtained using the catalysts L2-V and L3-V do not reveal melting points (Tm). This is evidence of the considerable homogeneity of copolymer, its amorphous nature.
The microstructure of the copolymers was further determined by analyzing the 13C NMR spectra (Figure 3, Figures S33–S41) according to literature [30,31]. Generally, the spectra show the signals assigned to the isolated NB units (45.02 ppm C2/C3, 39.50 ppm C1/C4, 30.89 ppm C7, and 28.33 ppm C5/C6). The signals 45.20 (C2/C3 racemo) and 45.72 ppm (C2/C3 meso), 39.54 ppm (C4 meso), 39.89 ppm (C1 meso), 31.04 ppm (C7 meso), 30.99 ppm (C7 racemo) are assigned to NB in the syndiotactic and alternating isotactic sequences, respectively. This shows that the copolymers have a random tacticity. The signals at 27.76–28.78 ppm are assigned to the methylene sequences. Increasing the comonomer incorporation reduces the number of isolated NB units.
In summary, differences amongst the copolymers produced using the studied catalysts are observed. The catalyst L2-V gives the higher incorporation of NB units than the catalysts L1-V and L3-V, which results in higher amount of alternated units. The copolymers obtained using the catalysts L1-V and L3-V, reveal similar NB incorporation and rather similar microstructure. However, the ethylene sequences of the copolymer obtained using the catalyst L1-V are longer and more organized, which is shown by DSC method.
The presented data show the influence of catalyst structure on the performance of polyreaction and the properties of polymer product. Substituent (methyl group) at the position 4 of oxazoline ligand reduces the catalyst activity (L2-V), what was observed previously [15]. The substituent at the position 5 of oxazole ring (L3-V), which imposes a steric hindrance and prevents co-planarity of both heterocyclic rings, increases the catalyst activity, but it does not influence the norbornene incorporation.

3. Materials and Methods

3.1. Materials and Ligand Synthesis

The substrates for syntheses and the methods of purification of solvents, as well as the detailed method of ligands synthesis are given in Supplementary Materials.

3.2. Catalysts Preparation

Argon atmosphere was applied for all steps of synthesis. The ligands, 4-(4,5-dihydro-1,3-oxazol-2-yl)-2-methyl-1,3-oxazole (L1), 4-(4-methyl-4,5-dihydro-1,3-oxazol-2-yl)-2-methyl-1,3-oxazole (L2), and 4-(4,5-dihydro-1,3-oxazol-2-yl)-5-methyl-1,3-oxazole (L3), were dried under vacuum during 2.5 h prior to the synthesis. The proper amount of L1–3 (0.60 mmol) was dissolved in dichloromethane (5 mL), freshly distilled over P2O5. At the same time, the solid VCl3(THF)3 (0.54 mmol) was dissolved in dichloromethane. In both cases homogenous, transparent solution was obtained. To vigorously stirred VCl3(THF)3 solution at −10 °C temperature (ice-acetone bath), the proper ligand L1-3 solution was added dropwise. The change of color of solution was observed during reaction progress (Figure S20). The reaction was carried out overnight. The solvent was removed using argon-vacuum and resulting solid was washed hexane (3 × 20 mL) freshly distilled over sodium in argon atmosphere. The solid complex was dried under vacuum to remove solvent.
L1-V complex. Yield 82%. 1H NMR (400MHz DMSO-d 6) δ (ppm): 8.49 (1H, s, C-H), 3.70 (2H, t, O-CH2), 3.54 (2H, t, N-CH2), 2.46 (3H, s, CH3). 13C NMR (400MHz DMSO-d6) δ (ppm): 161.80 (O-C=N), 160.96 (O-C=N), 142.18 (C-H), 136.33 (C), 46.42 (OCH2), 43.18 (NCH2), 13.88 (CH3) (Figure S14).
L2-V complex. Yield 88%. 1H NMR (400 MHz DMSO-d6) δ (ppm): 8.18 (1H, s, C-H), 4.23 (1H, t, O-C-Hcis), 3.72 (1H. ss. Cl-CH2), 3.67 (1H. t. O-C-Htrans), 2.46 (3H, s, CH3) 1.76 (3H, d, CH3). 13C NMR (400 MHz DMSO-d6) δ (ppm): 161.87 (N=C-O), 160.21 (O-C=N), 142.43 (O-C), 130.53 (N-C), 48.48 (O-CH2), 46.27 (N-CH2), 25.77 (CH3), 21.81 (CH3) (Figure S16).
L3-V complex. Yield 87%. 1H NMR (400 MHz DMSO-d6) δ (ppm): 8.37 (1H, s, C-H), 3.70 (2H, t, O-CH2), 3.54 (2H, t, N-CH2), 1.76 (3H, s, CH3). 13C NMR (400 MHz DMSO-d6) δ (ppm): 161.76 (O-C=N), 152.96 (C-H), 152.47 (O-C), 124.54 (N-C), 43.42 (O-CH2), (covered DMSO N-CH2), 25.62 (CH3) (Figure S18).

3.3. Polymerization Procedure

The (co)polymerizations were performed in a glass reactor (500 mL) in nitrogen atmosphere. Hexane (total 150 mL), norbornene (0.5, 1.0 and 1.50 mol/dm3) or not (as hexane solution), AlEt2Cl, (6.0 × 10–3 mol Al), ethyl trichloroacetate (ETA) (3.5 × 10–4 mol), vanadium catalysts (L-V, 7.1 × 10–6 mol Mt) as toluene solution (2 mL), and ethylene were successively added. The reactions were performed at 60 min, ethylene pressure 0.5 MPa, and temperature 30 °C. To terminate the reactions, the ethylene feeding was closed, pressure reduced to 0.1 MPa, and acidified methanol was added. The polymer product was separated by filtration, washed first by hexane and then by methanol, and dried to constant mass at 50 °C. The reactions were repeated to ensure reproducibility of the results.

3.4. Instruments

NMR spectra (13C and 1H) were obtained using spectrometer 400 MHz (Bruker Ultrashield, Billerica, MA, USA). Dimethyl sulfoxide-d6 (ligands and complexes) or o-dichlorobenzene-d4 (polymers and copolymers) were used as solvents. Total norbornene (NB) incorporation in copolymers was calculated using equation: NB mol% = (1/3(2IC7 + IC1/C4 + IC2/C3)/ICH2) × 100% where: ICH2, IC7, IC1/C4, IC2/C3 are total area of the 13C NMR signal at 26–31, 31–34, 36–42, and 43–50 ppm.
EI + MS spectra were obtained using mass spectrometer Finnigan Polaris Qequipped with a Direct Insertion Probe. Xcalibur data system was directly coupled to the spectrometer. The range of heating in an ion source, 50 °C to 500 °C. The mass monitoring interval, 50 to 1000 amu. Cyclical scans used to collect the spectra, 1.0 s. The electron energy, 70 eV.
Melting temperatures (Tm), crystallinity degree (X), and glass transition temperatures (Tg) of the polymers samples were recorded by differential scanning calorimetry (DSC) The calorimeter (2010 TA) with an automated sampler was used. The cycle heat-cool-heat, heating rate of 10 °C/min, and nitrogen atmosphere were applied to collect data. The crystallinity degree of polyethylene was calculated using the following equation: X = (DHf/DHt,c) × 100% (DHf is enthalpy of fusion of the polyethylene sample and DHt,c is enthalpy of fusion of standard, 290 J/g).
Molecular weight (Mw) and molecular weight distribution (Mw/Mn) of polymer samples were measured using gel permeation chromatography system (PL-GPC 220) equipped with refractive index and viscosity detector. Run conditions: temperature 160 °C, solvent 1,2,4-trichlorobenzene (TCB), flow rate 1 mL/min. The polymers were analyzed on a set of gel columns (Olexis). polyethylene and polystyrene standards with narrow Mw/Mn were applied for calibration.

4. Conclusions

The vanadium catalyst with oxazole-oxazoline ligands revealed activity both in ethylene polymerization and ethylene-norbornene copolymerization. The work shows that the search of structural units in natural compounds and their further adaptation as ligands for transition metal catalysts can be promising procedure. The presented relatively simple bidentate oxazole-oxazoline ligands can be further developed both to bidentate oxazole-oxazole ligands, but also tridentate oxazole ligands. Both these structure are present in natural compounds. It would be also worth to check these ligands using various transition metals. Furthermore, the route of synthesis from simple serine/threonine peptide substrates is potentially possible, which should fulfill requirements of green chemistry.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/catal11080923/s1. Nomenclature of ligands, Materials, Ligand synthesis, Figures S1–S5: NMR for L1 synthesis, Figure S6: MS for L1, Figures S7–S10: NMR for L2 synthesis, Figure S11: MS for L2, Figures S12 and S13: NMR and MS spectra for L3, Figures S14–S19: NMR and MS spectra for vanadium complexes, Figure S20: synthesis of complexes, Figures S21–S32: DSC thermograms of polymer samples, Figures S33–S41: 13C NMR spectra of (co)polymer samples.

Author Contributions

Conceptualization, W.O.-S.; Ligand synthesis, D.S. and K.B.; Catalysts synthesis, W.O.-S., K.H.; Polymerization, W.O.-S. and K.H; Data analysis, W.O.-S., D.S., K.D., and S.W.; MS measurement, K.D.; Writing, W.O.-S. and D.S.; Consultation, S.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financed by the University of Opole (505-50/9-45-00-00/402-10-03-01-0-0 /21-85-001/000-6-85-0/00-000-1/01-085-001) and the Ministry of Education, Youth and Sports of the Czech Republic, an ERDF project entitled “Development of Pre-Applied Research in Nanotechnology and Biotechnology” (No. CZ.02.1.01/0.0/0.0/17_048/0007323).

Acknowledgments

Hana Omamikova performed MS measurements, Klaudia Kołaczkiewicz and Justyna Błaszkiewicz participated in synthesis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dawooda, K.M.; Nomura, K. Recent Developments in z-selective olefin metathesis reactions by molybdenum, tungsten, ruthenium, and vanadium catalysts. Adv. Synth. Catal. 2021, 363, 1970–1997. [Google Scholar] [CrossRef]
  2. Yuan, S.-F.; Yan, Y.; Solan, G.A.; Ma, Y.; Sun, W.-H. Recent advancements in N-ligated group 4 molecular catalysts for the (co)polymerization of ethylene. Coord. Chem. Rev. 2020, 411, 213254. [Google Scholar] [CrossRef]
  3. Phillips, A.M.F.; Suo, H.; da Silva, M.F.C.G.; Pombeiro, A.J.L.; Sun, W.-H. Recent developments in vanadium-catalyzed olefin coordination polymerization. Coord. Chem. Rev. 2020, 416, 213332. [Google Scholar] [CrossRef]
  4. Ishikura, H.; Neven, R.; Lange, T.; Galetová, A.; Blom, B.; Romano, D. Developments in vanadium-catalysed polymerisation reactions: A review. Inorg. Chim. Acta 2021, 515, 120047. [Google Scholar] [CrossRef]
  5. Spronck, M.; Klein, A.; Blom, B.; Romano, D. Synthesis of disentangled ultra-high molecular weight polyethylene using vanadium(V)-based catalysts. Z. Anorg. Allg. Chem. 2018, 644, 993–998. [Google Scholar] [CrossRef]
  6. Bihun-Kisiel, A.; Ochędzan-Siodłak, W. Vanadium catalysts for ethylene-norbornene copolymerization. Polimery 2020, 65, 11–12. [Google Scholar] [CrossRef]
  7. Shakeri, S.E.; Mortazavi, S.M.M.; Ahmadjo, S.; Zohuri, G.H. Comparison of mono and dinuclear α–diimine Ni based catalysts for synthesis of polynorbornene and its microstructure study. J. Macromol. Sci. Part A 2020, 57, 837–843. [Google Scholar] [CrossRef]
  8. Ochędzan-Siodłak, W.; Bihun-Kisiel, A.; Siodłak, D.; Poliwoda, A.; Dziuk, B. Titanium and vanadium catalysts with oxazoline ligands for ethylenenorbornene (co)polymerization. Eur. Polym. J. 2018, 106, 148–155. [Google Scholar] [CrossRef]
  9. Guo, J.; Liu, H.; Bi, J.; Zhang, C.; Zhang, H.; Bai, C.; Hua, Y.; Zhang, X. Pyridine–oxazoline and quinoline–oxazoline ligated cobalt complexes: Synthesis, characterization, and 1,3-butadiene polymerization behaviors. Inorg. Chim. Acta 2015, 435, 305–312. [Google Scholar] [CrossRef]
  10. Yun, B.-S.; Hidaka, T.; Furihata, K.; Seto, H. Microbial metabolites with tipA promoter inducing activity-II: Geninthiocin, a novel thiopeptide produced by Streptomyces sp. DD84. J. Antibiot. 1994, 47, 969–975. [Google Scholar] [CrossRef] [Green Version]
  11. Kodani, S.; Ninomiya, A. Isolation of new thiopeptide berninamycin E from Streptomyces atroolivaceus. Asian J. Chem. 2013, 25, 490–492. [Google Scholar] [CrossRef]
  12. Vijaya Kumar, E.K.S.; Kenia, J.; Mukhopadhyay, T.; Nadkarni, S.R. Methylsulfomycin I, a new cyclic peptide antibiotic from a streptomyces sp. HIL Y-9420704. J. Nat. Prod. 1999, 62, 1562–1564. [Google Scholar] [CrossRef] [PubMed]
  13. Castro Rodríguez, J.; Holgado, G.G.; Santamaría Sánchez, R.I.; Cañedo, L.M. Radamycin, a novel thiopeptide produced by Streptomyces sp. RSP9. II. Physico-chemical properties and structure determination. J. Antibiot. 2002, 55, 391–395. [Google Scholar] [CrossRef] [Green Version]
  14. Engelhardt, K.; Degnes, K.F.; Kemmler, M.; Bredholt, H.; Fjærvik, E.; Klinkenberg, G.; Sletta, H.; Ellingsen, T.E.; Zotchev, S.B. Production of a new thiopeptide antibiotic, TP-1161, by a marine Nocardiopsis species. Appl. Environ. Microbiol. 2010, 76, 4969–4976. [Google Scholar] [CrossRef] [Green Version]
  15. Ochędzan-Siodłak, W.; Siodłak, D.; Piontek, A.; Doležal, K. Titanium and vanadium catalysts with 2-hydroxyphenyloxazoline and oxazine ligands for ethylene-norbornene (co)polymerization. Catalysts 2019, 9, 1041. [Google Scholar] [CrossRef] [Green Version]
  16. Wu, J.; Mu, R.; Liu, Z.-J.; Lu, S.-C.; Liu, G. Scalable total synthesis of a mycobactin T analogue utilizing a novel synthetic and protection strategy. Org. Chem. Front. 2019, 6, 2467–2470. [Google Scholar] [CrossRef]
  17. Yokokawa, F.; Izumi, K.; Omata, J.; Shioiri, T. Total synthesis of amamistatin A, an antiproliferative linear peptide from an actinomycete. Tetrahedron 2000, 56, 3027–3034. [Google Scholar] [CrossRef]
  18. Miller, M.J.; Zhu, H.; Xu, Y.; Wu, C.; Walz, A.J.; Vergne, A.; Roosenberg, J.M.; Moraski, G.; Minnick, A.A.; McKee-Dolence, J.; et al. Utilization of microbial iron assimilation processes for the development of new antibiotics and inspiration for the design of new anticancer agents. Biometals 2009, 22, 61–75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Mhlongo, J.T.; Brasil, E.; de la Torre, B.G.; Albericio, F. Naturally occurring oxazole-containing peptides. Mar. Drugs 2020, 18, 203. [Google Scholar] [CrossRef]
  20. Nagatsu, A.; Kajitani, H.; Sakakibara, J. Muscoride A: A new oxazole peptide alkaloid from freshwater cyanobacterium Nostoc muscorum. Tetrahedron Lett. 1995, 36, 4097–4100. [Google Scholar] [CrossRef]
  21. Smith, T.E.; Kuo, W.; Balskus, E.P.; Bock, V.D.; Roizen, J.L.; Theberge, A.B.; Carroll, K.A.; Kurihara, T.; Wessler, J.D. Total synthesis of (-)-hennoxazole A. J. Org. Chem. 2008, 73, 142–150. [Google Scholar] [CrossRef]
  22. Fernández, R.; Martín, M.J.; Rodríguez-Acebes, R.; Reyes, F.; Francesch, A.; Cuevas, C. Diazonamides C–E, new cytotoxic metabolites from the ascidian Diazona sp. Tetrahedron Lett. 2008, 49, 2283–2285. [Google Scholar] [CrossRef]
  23. Scholz, R.; Molohon, K.J.; Nachtigall, J.; Vater, J.; Markley, A.L.; Süssmuth, R.D.; Mitchell, D.A.; Borriss, R. Plantazolicin, a novel microcin B17/streptolysin S-like natural product from Bacillus amyloliquefaciens FZB42. J. Bacteriol. 2011, 193, 215–224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Pattenden, G.; Ashweek, N.J.; Baker-Glenn, C.A.G.; Kempson, J.; Walker, G.M.; Yee, J.G.K. Total synthesis of (-)-ulapualide A, a novel tris-oxazole macrolide from marine nudibranchs, based on some biosynthesis speculation. Org. Biomol. Chem. 2008, 6, 1478–1497. [Google Scholar] [CrossRef]
  25. Liu, N.; Song, L.; Liu, M.; Shang, F.; Anderson, Z.; Fox, D.J.; Challis, G.L.; Huang, Y. Unique post-translational oxime formation in the biosynthesis of the azolemycin complex of novel ribosomal peptides from Streptomyces sp. FXJ1.264. Chem. Sci. 2016, 7, 482. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Shin-ya, K.; Wierzba, K.; Matsuo, K.-i.; Ohtani, T.; Yamada, Y.; Furihata, K.; Hayakawa, Y.; Seto, H. Telomestatin, a novel telomerase inhibitor from Streptomyces anulatus. J. Am. Chem. Soc. 2001, 123, 1262–1263. [Google Scholar] [CrossRef] [PubMed]
  27. Sarada, G.; Sim, B.; Cho, W.; Yoon, J.; Gal, Y.S.; Kim, J.J.; Jin, S.H. New sky-blue and bluishegreen emitting Ir(III) complexes containing an azoline ancillary ligand for highly efficient PhOLEDs. Dye. Pigment. 2016, 131, 60–68. [Google Scholar] [CrossRef]
  28. Comito, R.J.; Wu, Z.; Zhang, G.; Lawrence, J.A., III; Korzyński, M.D.; Kehl, J.A.; Miller, J.T.; Dincă, M. Stabilized vanadium catalyst for olefin polymerization by site isolation in a metal–organic framework. Angew. Chem. Int. Ed. 2018, 57, 8135–8139. [Google Scholar] [CrossRef] [PubMed]
  29. Pei, L.; Tang, Y.; Gao, H. Homo- and copolymerization of ethylene and norbornene with anilido–imine chromium catalysts. Polymers 2016, 8, 69. [Google Scholar] [CrossRef] [Green Version]
  30. Boggioni, L.; Losio, S.; Tritto, I. Microstructure of Copolymers of Norbornene Based on Assignments of 13C NMR Spectra: Evolution of a Methodology. Polymers 2018, 10, 647. [Google Scholar] [CrossRef] [Green Version]
  31. Leone, G.; Mauri, M.; Losio, S.; Bertini, F.; Ricci, G.; Porri, L. Copolymerization of ethylene with a-olefins and cyclic olefins catalyzed by a Ti(IV) diisopropoxy complex bearing a tridentate [O,S,O]-type bis(phenolato) ligand. Polym. Chem. 2014, 5, 3412–3423. [Google Scholar] [CrossRef]
Figure 1. The structure of ligands: (L1) 4-(4,5-dihydro-1,3-oxazol-2-yl)-2-methyl-1,3-oxazole, (L2) 4-(4-methyl-4,5-dihydro-1,3-oxazol-2-yl)-2-methyl-1,3-oxazole, (L3) 4-(4,5-dihydro-1,3-oxazol-2-yl)-5-methyl-1,3-oxazole.
Figure 1. The structure of ligands: (L1) 4-(4,5-dihydro-1,3-oxazol-2-yl)-2-methyl-1,3-oxazole, (L2) 4-(4-methyl-4,5-dihydro-1,3-oxazol-2-yl)-2-methyl-1,3-oxazole, (L3) 4-(4,5-dihydro-1,3-oxazol-2-yl)-5-methyl-1,3-oxazole.
Catalysts 11 00923 g001
Figure 2. DSC thermograms of selected copolymer samples obtained using the studied catalysts (Table 1, items 4, 8, 12, Figures S24, S28 and S32).
Figure 2. DSC thermograms of selected copolymer samples obtained using the studied catalysts (Table 1, items 4, 8, 12, Figures S24, S28 and S32).
Catalysts 11 00923 g002
Figure 3. 13C NMR spectra of ethylene-norbornene copolymers in o-dichlorobenzene-d4 (Table 1, items 3, 7, 11, Figures S34, S37 and S40).
Figure 3. 13C NMR spectra of ethylene-norbornene copolymers in o-dichlorobenzene-d4 (Table 1, items 3, 7, 11, Figures S34, S37 and S40).
Catalysts 11 00923 g003
Table 1. Selected parameters of (co)polymerization performed using studied vanadium catalysts.
Table 1. Selected parameters of (co)polymerization performed using studied vanadium catalysts.
CatalystItemNB YieldActivityNBTgTmMw·10−3Mw/Mn
(mol/dm3)(g)(kg/molV/h)(mol%)(°C)(°C)(g/mol)
L1-V1-29.804198--136.26302.0
20.512.88181417.311.7128.73602.7
31.015.05212021.428.3128.52402.4
41.513.45189526.346.4128.62102.5
L2-V5-22.363150--138.38902.1
60.58.80124025.035.5-3302.2
71.09.83138526.642.7-2302.4
81.513.23186432.461.7-1802.8
L3-V9-34.104804--135.37202.1
100.512.11170611.49.8-4102.5
111.017.38244821.730.8-2802.6
121.519.35272624.142.0-2202.4
Reaction conditions: catalyst concentration (7.1 × 10–6 mol), activator (AlEt2Cl) (6.0 × 10–3 mol), ethyl trichloroacetate (ETA) (3.5 × 10–4 mol), temperature (30 °C), pressure (0.5 MPa), (co)polymerization time (60 min), solvent (hexane 150 mL). Activity (kg polymer per mol V per hour). NB content (mol%) estimated by 13C NMR. Mw and Mw/Mn determined by GPC.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ochędzan-Siodłak, W.; Siodłak, D.; Banaś, K.; Halikowska, K.; Wierzba, S.; Doležal, K. Naturally Occurring Oxazole Structural Units as Ligands of Vanadium Catalysts for Ethylene-Norbornene (Co)polymerization. Catalysts 2021, 11, 923. https://doi.org/10.3390/catal11080923

AMA Style

Ochędzan-Siodłak W, Siodłak D, Banaś K, Halikowska K, Wierzba S, Doležal K. Naturally Occurring Oxazole Structural Units as Ligands of Vanadium Catalysts for Ethylene-Norbornene (Co)polymerization. Catalysts. 2021; 11(8):923. https://doi.org/10.3390/catal11080923

Chicago/Turabian Style

Ochędzan-Siodłak, Wioletta, Dawid Siodłak, Karolina Banaś, Katarzyna Halikowska, Sławomir Wierzba, and Karel Doležal. 2021. "Naturally Occurring Oxazole Structural Units as Ligands of Vanadium Catalysts for Ethylene-Norbornene (Co)polymerization" Catalysts 11, no. 8: 923. https://doi.org/10.3390/catal11080923

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