Open access peer-reviewed chapter

One-Pot Synthesis of Chiral Organometallic Complexes

Written By

Mei Luo

Submitted: 15 June 2019 Reviewed: 23 September 2019 Published: 08 November 2019

DOI: 10.5772/intechopen.89865

From the Edited Volume

Synthesis Methods and Crystallization

Edited by Riadh Marzouki

Chapter metrics overview

820 Chapter Downloads

View Full Metrics

Abstract

Currently, organometallic complexes involving ligand oxazolines are typically obtained in two sequential steps, where the free ligand is given firstly from a functionalized nitrile by condensation reaction with an amino alcohol in the presence of a Lewis or Bronsted acid catalyst, followed by a further coordination with metal salts to obtain the corresponding oxazolinyl metal complexes. Usually, the yield of the two-step procedure is relatively low; considering that metal oxazoline complexes often contain Lewis acidic metals, it is possible that the two steps may be telescoped. A series of novel chiral organometallic complexes (1–23) were assembled in a single step from nitriles, chiral D/L amino alcohols, and a stoichiometric amount of metal salts (MCl2·nH2O/M(OAc)2·nH2O), with moderate to high yields (20–95%). All the crystalline compounds were fully characterized by NMR, IR, MS, and X-ray analyses.

Keywords

  • chiral organometallic complexes
  • nitriles
  • amino alcohols
  • metal salts
  • crystalline compounds

1. Introduction

Chiral oxazolines constitute an important class of “privileged” ligands in asymmetric catalysis [1, 2, 3, 4, 5, 6, 7, 8]. Organometallic complexes involving oxazoline ligands are typically obtained in two steps, where the free ligand is given firstly from a functionalized nitrile through condensation reaction with an amino alcohol in the presence of a Lewis or Bronsted acid catalyst, followed by further complexed coordination with metal salts to obtain the corresponding oxazolinyl metal complexes (Figure 1) [9, 10]. Usually, the yield of the two-step procedure is relatively low, and certain oxazolinyl organometallic complexes are difficult to obtain due to the poor coordination ability of the imine group from the oxazoline. It is conceivable that the two steps may be telescoped by using the requisite Lewis acid precursor. Herein, through the assembly of three reaction components (a nitrile, an amino alcohol, and metal salts), we first report a simple, one-step procedure for the preparation of N-containing heterocyclic zinc complexes (1–15), with the yield of certain products reaching 90% in the presence of a large amount of ZnCl2 (0.4–2.6 eq.) and certain chiral salicyloxazoline metal complexes (16–23) with yields ranging from 65 to 95% using 1.0 eq. of copper, cobalt, nickel, manganese, palladium, and platinum salts as the third component. In all the cases, the complexes were isolated, purified, and characterized. All the structures reported in this paper were confirmed by X-ray crystallography.

Figure 1.

Common method for the preparation of oxazolinyl metal complexes.

Advertisement

2. One-step multicomponent synthesis of chiral oxazolinyl-zinc complexes

The one-pot procedure was initially tested from the reaction of different 1-piperidine propionitrile derivatives with 2–3 eq. of amino alcohol refluxed in chlorobenzene for 72 h in the presence of 1–2.6 eq. of ZnCl2. After cooling to room temperature, the solvent was removed under reduced pressure, and the residue was dissolved in H2O and extracted with CH2Cl2. The combined organic extracts were evaporated to give a crude red oil, which was purified by column chromatography (petroleum ether/CH2Cl2, 4/1) to afford the title compound as crystals. During the preliminary work, it quickly became apparent that the reaction results are controlled by the amount of ZnCl2 used (Figure 2); for example, employing 1.1 eq. or 2.6 eq. of ZnCl2, the desired crystal structures of the amino-oxazolidinyl zinc complex 1 and bis-oxazolidinyl zinc complex 2 containing two monodentate ligands can be obtained from the reaction of L-leucinol or L-valinol with 3-piperidin-1-yl-propionitrile, respectively, followed by evaporation of different ratios of petroleum and dichloromethane from the mixture after column separation but with only a low yield (25%) for complex 1 and moderate yield (65%) for complex 2.

Figure 2.

Effect of the reaction stoichiometry of (metal precursor) ZnCl2.

The nature of the side chain (R1) influenced the reaction outcome. Using L-phenylalaninol with 1.6 eq. of ZnCl2 or 1.5 eq. of ZnCl2 with 1-morpholinepropionitrile (X = O) and D-phenylglycinol, both led to the cleavage of the propionitrile, providing asymmetric diamine complexes 3 and 4 at very good yields of 86% and 90%, respectively. Interestingly, using 1-(2-cyanoethyl)-4-methylpiperazine (Z = NMe) as a precursor with 2.5 eq. of ZnCl2 led only to the formation of the zwitterionic piperazine complex 5, irrespective of the amino alcohol used. The results again prove the effects of different amounts of metal salts on the reaction.

From the crystal structures of complexes 2–5, we conclude that the propionitrile precursors are unstable and decompose into acetonitrile or the parent cyclic amines to afford complexes 2 and 3–5, respectively, in the presence of a large amount of zinc chloride. For this reason, a number of nitrile precursors with additional N-donor were selected to be more robust against degradation under the reaction conditions. Consequently, a number of aromatic nitrile precursors containing additional N-donors were applied widely in these three-component reactions. In the process of selecting these reactions, the appropriate amount of ZnCl2 was carefully optimized to ensure specific results. Complex 6, listed in Figure 3, contains two monodentate ligands coordinated via the oxazoline nitrogen and was afforded from the use of 3-aminobenzonitrile and D-leucinol in the presence of 0.44 eq. of ZnCl2. Similarly, the bis-chelated complex 7 and the mono-chelated complex 8 were afforded to the corresponding yields of 80 and 78% from the use of 2-cyanopyridine with L-phenylalaninol and D-valinol, respectively, in the presence of 1.2 eq. of ZnCl2.

Figure 3.

Zinc complexes 6–8 derived from 3-aminobenzonitrile and 2-cyanopyridine.

The formation of complexes 9 and 10 were studied using different amino alcohols, as seen in Figure 4. C2-symmetrical bis-oxazolines formed seven-membered chelate rings which derived from 1,2-dicyanobenzene, affording a 1:1 adduct with zinc dichloride. Indeed, the addition of isophthalonitrile with D-phenylglycinol (0.56 eq.) provided the predicted mono-chelated complex 9 [11] at a good yield (68%). However, a combination of a slight excess of L-valinol (0.72 eq.) caused the addition of three amino alcohols to give complex 10 with a yield of 66%.

Figure 4.

Complexes 9–10 derived from isophthalonitrile.

Surprisingly, the combination of L-leucinol and L-phenylglycinol with tetracyanoethylene in the presence of 0.42 eq. of ZnCl2, respectively, provided neutral bis[bis(oxazoline)]zinc (II) complexes 11 and 12 with corresponding yields of 88 and 86% (Figure 5). The crystal structures of these methylene-bis(oxazoline) indicate that the tricyanomethane was formed as an intermediate from a disproportionation-rearrangement of the tetracyanoethylene precursor, although the precise mechanism of this pathway is unclear. In 2016, Kögel et al. reported the synthesis of complex 12 by a different route [12]. Interestingly, complex 12 was reported to exhibit an intense cotton effect as a result of exciton coupling. Indeed, the X-ray crystal structures of complexes 11 and 12 have been proven that due to the their coordination environments, isobutyl-substituted complex 11 has shown a fairly symmetrical tetrahedral comformation, while complex 12 is in highly distorted. This maybe the result of the favorable intramolecular π-interaction between one of the phenyl groups with the semicorrin structure of the adjacent ligand within 3.5 Å, effectively leading to the two chiral chromophores’ close proximity to convenience exciton coupling [13].

Figure 5.

Neutral zinc complexes derived from tetracyanoethylene.

In further study, 2-hydroxy-6-methylnicotinonitrile was employed as a precursor to test the applicability of the one-pot methodology in assembling complex multinuclear structures. In the presence of different amounts of ZnCl2 (1.72, 1.31, and 1.54 eq.), the corresponding condensation products with valinol, leucinol, or phenylalaninol furnished the binuclear zwitterionic complex 13 and highly symmetrical tetramers 14 and 15 at yields of 86, 80, and 82%, accordingly (Figure 6). Presumably, the formation of higher aggregates is prevented by the sterically demanding isopropyl substituents. A six-membered N,O-chelate ligand is complexed at each zinc metal center and connected to another metal center with a bridging donor ligand from the pendant pyridine. With each zinc atom located at a corner of a square grid, the planar N,O,N-ligands are oriented perpendicularly to one another with diagonal Zn···Zn distances of ca. 6 Å.

Figure 6.

Multinuclear zinc complexes 13–15.

The crystal structures of all the complexes (Figures 721) are determined and reported by X-ray diffraction, elemental analysis, and IR. In all the cases, a distorted tetrahedral geometry is found at zinc(II), and the C〓N double-bond character of the oxazolinyl ligand is largely retained in the metal complexes (Table 1).

Figure 7.

The crystal structure of complex 1.

Figure 8.

The crystal structure of complex 2.

Figure 9.

The crystal structure of complex 3.

Figure 10.

The crystal structure of complex 4.

Figure 11.

The crystal structure of complex 5.

Figure 12.

The crystal structure of complex 6.

Figure 13.

The crystal structure of complex 7.

Figure 14.

The crystal structure of complex 8.

Figure 15.

The crystal structure of complex 9.

Figure 16.

The crystal structure of complex 10.

Figure 17.

The crystal structure of complex 11.

Figure 18.

The crystal structure of complex 12.

Figure 19.

The crystal structure of complex 13.

Figure 20.

The crystal structure of complex 14.

Figure 21.

The crystal structure of complex 15.

ZnCl2 (%) Products Yield (%)a
114.2 1 25
259.8 2 65
152.8 3 86
144.6 4 90
245.8–256.0 5 56
44.1 6 90
121.8 7 80
122.9 8 85
56.1 9 86
72.6 10 90
42.2 11 88
42.2 12 86
172.1 13 86
130.7 14 80
153.6 15 82`

Table 1.

One-pot synthesis of zinc complexes (1–15).

Isolated yield from silica gel.


Advertisement

3. One-step templated synthesis of chiral organometallic salicyloxazoline complexes

Chiral oxazolinyl organometallic complexes are very important catalysts in organic chemistry [14, 15, 16, 17, 18, 19, 20, 21, 22]. Several organometallic complexes containing 2-(2′-hydroxyphenyl) oxazolines are reported in the literature [23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42]. The general approach to the synthesis of metal complexes begins with ligand synthesis, followed by ligand reaction with metal salts to afford organometallic complexes [43]. In 2017, our research group first reported a one-pot multicomponent synthesis of chiral oxazolinyl-zinc complexes [44] in the presence of a large amount of ZnCl2 (0.4–2.6 eq.); the yields of certain products reached 90%. Herein, the chiral salicyloxazoline metal complexes 16–23 can be obtained by using 1.0 eq. of metal salts such as copper, cobalt, nickel, manganese, palladium, and platinum salts as the third component. The structures of these complexes were characterized by X-ray crystallography. The results prove that organometallic complexes can be assembled with two reactants and different amounts of metal salts.

Chiral bis(oxazoline) copper complex 16, nickel complex 17, cobalt complex 18, and palladium complex 19 were generated as crystals with the chemical formula ML2 (L = 2-(4-R1–4,5-dihydrooxazol-2-yl)phenol, R1: d-Ph, M: Cu, Ni, Co; R1: l-CH2Ph; M: Pd). The syntheses of these complexes can be summarized as follows: A mixture of 2-hydroxybenzonitrile and d-phenylglycinol or l-phenylalaninol in 50 mL of chlorobenzene was refluxed for 72 h with 1.0 eq. of each of the above appropriate metal salts. After removal of the chlorobenzene, single crystals of chiral bis(oxazolinyl) metal complexes 16–19 were present after natural evaporation of the recrystallization or chromatographic solvent with petroleum and dichloromethane (Figure 22).

Figure 22.

Templated synthesis of complexes 16–19.

In Figures 23 and 24, refluxing a mixture of 2-cyanophenol and d-phenylglycinol in chlorobenzene for 72 h with 1.0 eq. of cobalt chloride hexahydrate or 1.0 eq. of cobalt acetate tetrahydrate, respectively, afforded complexes 20 and 21. Further, through slow evaporation from a 1:1 mixture of ethanol and chloroform, crystals of complex 20 were obtained. However, the crystals of complex 21 were present after column separation with a 4:1 solution of petroleum ether and dichloromethane, followed by evaporation of the volatile components.

Figure 23.

Solvent effects on the formation of complexes 18 and 20.

Figure 24.

One-pot synthesis of tri(oxazoline) metal complexes 21 and 22.

Notably, the product complexes 18 and 20 were obtained using CoCl2·6H2O as a reagent with different solvents in the workup procedure. As seen in Figure 23, when a nonpolar solvent, such as petroleum ether or n-hexane, was used in the recrystallization medium, crystals of complex 18 were obtained. However, if the recrystallization was carried out with a mixture of two polar solvents, such as ethanol and chloroform, crystals of complex 20 were obtained.

Similarly, in the synthesis of chiral oxazoline manganese complex 22 by the title method, 2-hydroxybenzonitrile and d-phenylglycinol were refluxed with 1.0 eq. of manganese acetate tetrahydrate in chlorobenzene for 60 h (Figure 24). After removal of the chlorobenzene and slow evaporation with a mixture of absolute ethanol and chloroform, crystals of complex 22 were obtained.

Interestingly, in Figure 25, when employed by 1.0 eq. of PtCl2 in the reaction of 2-hydroxybenzonitrile with D-phenylglycinol in chlorobenzene, the crystal structure of the resulting Pt complex was different from the aforementioned complexes 16–22 (Figures 2632); a complex containing one unit of (R)-2-(4-phenyl-4,5-dihydrooxazol-2-yl)phenol and one unit of D-phenylglycinol was obtained after column chromatography with petroleum ether and dichloromethane (4:1) followed by crystallization via slow evaporation (Figures 25 and 33).

Figure 25.

One-pot synthesis of oxazoline platinum complex 23.

Figure 26.

The crystal structure of complex 16.

Figure 27.

The crystal structure of complex 17.

Figure 28.

The crystal structure of complex 18.

Figure 29.

The crystal structure of complex 19.

Figure 30.

The crystal structure of complex 20.

Figure 31.

The crystal structure of complex 21.

Figure 32.

The crystal structure of complex 22.

Figure 33.

The crystal structure of complex 23.

The proposed mechanism shows that excess metal salts can activate the reaction of 2-hydroxybenzonitrile with D-phenylglycinol in chlorobenzene to form ligand intermediates and then directly furnish the corresponding organometallic complexes via a one-step procedure (Table 2).

Metal salts The amount of metal salts Products (%) Yield (%)
Cu(OAc)2⋅H2O 55.7 16 65
CuCl2⋅2H2O 53.2 16 85
Ni(OAc)2⋅4H2O 51.0 17 92
NiCl2⋅6H2O 53.0 17 95
CoCl2⋅6H2O 44.3 18, 20 72, 85
PdCl2 49.8 19 86
Co(OAc)2⋅4H2O 42.3 21 70
Mn(OAc)2⋅4H2O 52.6 22 80
PtCl2 33.7 23 82

Table 2.

One-pot synthesis of salicyloxazoline complexes.

Advertisement

4. Conclusions

One-pot synthesis of oxazolinyl-zinc(II) complexes 1–23 at yields 25–95% was firstly demonstrated by assembling three-component reactions between metal salts, amino alcohols, and a variety of nitrile precursors. From the crystal structures of the complexes 1–23, the reaction product is highly dependent on the presence of ligands, the amount of metal salts, and the nature of the substituent at the stereogenic center, giving a variety of coordination modes, such as mono- and bis-chelate complexes, mononuclear and mutinuclear complexes, etc.

Investigations into other oxazolinyl organometallic complexes and the catalytic properties of these complexes as chiral ligands are currently ongoing. These complexes exhibit bioactivities as anticancer reagents, and the future use of these complexes in medical fields is currently being developed.

Advertisement

Acknowledgments

This work was supported by the Hefei University of Technology and the University of Science and Technology of China. The authors also thank prof. King Kuok (Mimi) from Imperial College London, prof. Peter J. Stang from the University of Utah and prof. K.H. Lee from the University of Carolina at Chapel Hill for providing supports and help.

Advertisement

Additional information is provided in the article as follows:

  1. The characterization spectra of compounds 115 [44] are available free of charge via the Internet at https://ccj.springeropen.com/. The crystallographic information of compounds 115 are available from the Cambridge Crystallographic Data Center (CCDC) as supplementary publications CCDC 853709–853,710, 931,745–931,746, 931,745–931,748, 931,751–931,753, 931,756, 1,014,806–1,014,807, and 1,540,756, deposit@ccdc.cam.ac.uk or http://www. ccdc.cam.ac.uk.

  2. Supporting information including the NMR spectra for compounds 1–15 [44] are available free of charge via the Internet at https://ccj.springeropen.com/articles/10.1186/s13065-017-0305-1.

  3. Supporting information including the NMR spectra for compounds 16–23 [45] are available free of charge via the Internet at: Crystallographic Data Center (CCDC) as supplementary publications CCDC 853709–853710, 931745–931746, 931745–931748, 931751–931753, 931756, 1014806–1014807 and 1540756.

  4. The crystallographic information of compounds 1–23 are available from the Cambridge Crystallographic Data Center (CCDC) as supplementary publications CCDC 853709–853,710, 931,745–931,746, 931,745–931,748, 931,751–931,753, 931,756, 1,014,806–1,014,807, 1,540,756, and CCDC 1035470–1,035,473, 1,035,475–1,035,476, 1,035,626, and 1,835,179, deposit@ccdc.cam.ac.uk or http://www.ccdc.cam.ac.uk.

Advertisement

Conflicts of interest

The authors declare that they have no competing interests.

References

  1. 1. Desimoni G, Faita G, Jorgensen KA. Update 1 of C-2-symmetric chiral bis(oxazoline) ligands in asymmetric catalysis. Chemical Reviews. 2011;111:284-437. DOI: 10.1021/cr100161x
  2. 2. Hargaden GC, Guiry PJ. Recent applications of oxazoline-containing ligands in asymmetric catalysis. Chemical Reviews. 2009;109:2505-2550. DOI: 10.1021/cr800400z
  3. 3. O’Reilly S, Guiry PJ. Recent applications of C-1-symmetric bis(oxazoline)-containing ligands in asymmetric catalysis. Synthesis-Stuttgart. 2014;46:722-739. DOI: 10.1055/s-0033-1340829
  4. 4. Gotoh R, Yamanaka M. Chiral Zn(II)-bisamidine complex as a Lewis-Bronsted combined acid catalyst: Application to asymmetric Mukaiyama aldol reactions of alpha-ketoesters. Molecules. 2012;17:9010-9022. DOI: 10.3390/molecules17089010
  5. 5. Makino K, Ogawa I, Hamada Y. Synthesis of new chiral bis-oxazoline ligand with zinc triflate-selective chelating ability and its applications. Heterocycles. 2005;66:433-440. DOI: 10.3987/COM-05-S(K)49
  6. 6. Abbina S, Du G. Zinc-catalyzed highly isoselective ring opening polymerization of rac-lactide. ACS Macro Letters. 2014;3:689-692. DOI: 10.1021/mz5002959
  7. 7. Abbina S, Du G. Chiral amido-oxazolinate zinc complexes for asymmetric alternating copolymerization of CO2 and cyclohexene oxide. Organometallics. 2012;31:7394-7403. DOI: 10.1021/om3006992
  8. 8. Bleith T, Deng Q-H, Wadepohl H, Gade LH. Radical changes in Lewis acid catalysis: Matching metal and substrate. Angewandte Chemie, International Edition. 2016;55:7852-7856. DOI: 10.1002/anie.201603072
  9. 9. Le Roux E, Merle N, Tornroos KW. Synthesis and characterisation of trigonal C-2-chiral di- and tetra-substituted bis(oxazoline) alkylzinc complexes and their reactivity towards protic reagents. Dalton Transactions. 2011;40:1768-1777. DOI: 10.1039/C0DT01252A
  10. 10. Witte H, Seeliger W. Formation of cyclic imidic esters by reaction of nitriles with amino alcohols. Liebigs Annalen der Chemie. 1974:996-1009. DOI: 10.1002/jlac.197419740615
  11. 11. Bolm C, Weickhardt K, Zehnder M, Ranff T. Synthesis of optically active bis(2-oxazolines): Crystal structure of a 1,2-bis(2-oxazolinyl)benzene ZnCl2 complex. Chemische Berichte. 1991;124:1173-1180. DOI: 10.1002/cber.19911240532
  12. 12. Kögel JF, Kusaka S, Sakamoto R, Iwashima T, Tsuchiya M, Toyoda R, et al. Heteroleptic [bis(oxazoline)](dipyrrinato)zinc(ii) complexes: Bright and circularly polarized luminescence from an originally achiral dipyrrinato ligand. Angewandte Chemie, International Edition. 2016;55:1377-1381. DOI: 10.1002/anie.201603072
  13. 13. Telfer SG, McLean TM, Waterland MR. Exciton coupling in coordination compounds. Dalton Transactions. 2011;40:3097-3108. DOI: 10.1039/C0DT01226B
  14. 14. Rechavi D, Lemaire M. Enantioselective catalysis using heterogeneous bis(oxazoline) ligands: Which factors influence the enantioselectivity? Chemical Reviews. 2002;102:3467-3494. DOI: 10.1021/cr020008m
  15. 15. Johnson JS, Evans DA. Enantioselective catalysis using heterogeneous bis(oxazoline) ligands: Which factors influence the enantioselectivity? Accounts of Chemical Research. 2000;33:325-335. DOI: 10.1021/ar960062n
  16. 16. Pfaltz A. Chiral semicorrins and related nitrogen heterocycles as ligands in asymmetric catalysis. Accounts of Chemical Research. 1993;26:339-345. DOI: 10.1021/ar00030a007
  17. 17. Ghosh AK, Mathivanan P, Cappiello J. C2-symmetric chiral bis(oxazoline)–metal complexes in catalytic asymmetric synthesis. Tetrahedron: Asymmetry. 1998;9:1-45. DOI: 10.1016/S0957-4166(97)00593-4
  18. 18. Bolm C, Luong TKK, Schlingloft G. Enantioselective metal-catalyzed Baeyer-Villiger oxidation of cyclobutanones. Synlett. 1997:10;1151-1152
  19. 19. Bolm C, Bienewald F, Schlingloft G. Copper- and vanadium-catalyzed asymmetric oxidations. Journal of Molecular Catalysis A: Chemical. 1997;117:347-350. DOI: 10.1016/S1381-1169(96)00359-7
  20. 20. Bolm C, Bienewald F, Harms K. Syntheses and vanadium complex of salen-like bissulfoximines. Synlett. 1996;8:775-776. DOI: 10.1055/s-1996-5512
  21. 21. Bolm C, Bienewald F. Asymmetric sulfide oxidation with vanadium catalysts and H2O2. Angewandte Chemie, International Edition in English. 1995;34:2640-2642. DOI: 10.1002/ange.19951072317
  22. 22. Moreno RM, Bueno A, Moyano A. 4-Ferrocenyl-1,3-oxazoline derivatives as ligands for catalytic asymmetric allylation reactions. Journal of Organometallic Chemistry. 2003;671:187-188. DOI: 10.1016/S0022-328X(03)00095-0
  23. 23. Braunstein P, Naud F. Hemilability of hybrid ligands and the coordination chemistry of oxazoline-based systems. Angewandte Chemie, International Edition. 2001;40:680-699. DOI: 10.1002/1521-3773(20010216)40:4<680::AIDANIE6800>3.0.CO;2-0
  24. 24. Bolm C. Bis(4,5-dihydrooxazolyl) derivatives in asymmetric catalysis. Angewandte Chemie, International Edition in English. 1991;30:542-543. DOI: 10.1002/anie.199105421
  25. 25. Gómez M, Muller G, Rocamora M. Coordination chemistry of oxazoline ligands. Coordination Chemistry Reviews. 1999;193-195:769-835. DOI: 10.1016/S0010-8545(99)00086-7
  26. 26. Cozzi PG, Floriani C, Chiesi-Villa A, Rizzoli C. Oxazoline early transition metal complexes: Functionalizable achiral titanium(IV), titanium(III), zirconium(IV), vanadium(III), and chiral zirconium(IV) bis(oxazoline) complexes. Inorganic Chemistry. 1995;34:2921-2930. DOI: 10.1021/ic00115a020
  27. 27. Cozzi PG, Gallo E, Floriani C, Chiesi-Villa A, Rizzoli C. (Hydroxyphenyl)oxazoline: A novel and remarkably facile entry into the area of chiral cationic alkylzirconium complexes which serve as polymerization catalysts. Organometallics. 1995;14:4994-4996. DOI: 10.1021/om00011a010
  28. 28. Gant TG, Meyers AI. The chemistry of 2-oxazolines (1985–present), (hydroxyphenyl)oxazoline: A novel and remarkably facile entry into the area of chiral cationic alkylzirconium complexes which serve as polymerization catalysts. Tetrahedron. 1994;50:2297-2360. DOI: 10.1016/S0040-4020(01)86953-2
  29. 29. Bolm C, Schlingloff G. Metal-catalyzed enantiospecific aerobic oxidation of cyclobutanones. Chemical Communications. 1995:26;1247-1248
  30. 30. Bolm C, Schlingloff G, Weickhardt K. Optically-active lactones from a Baeyer-Villiger-type metal-catalyzed oxidation with molecular-oxygen. Angewandte Chemie, International Edition in English. 1994;33:1848-1849. DOI: 10.1002/anie.199418481
  31. 31. Peng Y, Feng X, Yu K, Li Z, Jiang Y, Yeung C-H. Synthesis and crystal structure of bis-[(4S,5S)-4,5-dihydro-4,5-diphenyl-2-(2′-oxidophenyl-chi O)oxazole-chi N] copper(II) and its application in the asymmetric Baeyer-Villiger reaction. Journal of Organometallic Chemistry. 2001;619:204. DOI: 10.1016/S0022-328X(00)00687-2
  32. 32. Bolm C, Luong TK, Harms K. Bis[2-(oxazolinyl)phenolato]oxovanadium(IV) complexes: Syntheses, crystal structures and catalyses. Chemische Berichte/Recueil. 1997;130:887-890. DOI: 10.1002/cber.19971300712
  33. 33. Hwang JJH, Abu-Omar MM. New vanadium oxazoline catalysts for epoxidation of allylic alcohols. Tetrahedron Letters. 1999;40:8313-8316. DOI: 10.1016/S0040-4039(99)01711-6
  34. 34. Gómez M, Jansat S, Muller G, Noguera G, Teruel H, Moliner V, et al. First dioxomolybdenum(VI) complexes containing chiral oxazoline ligands: Synthesis, characterization and catalytic activity. European Journal of Inorganic Chemistry. 2001:4;1071-1076
  35. 35. Arias J, Newlands CR, Abu-Omar MM. Kinetics and mechanisms of catalytic oxygen atom transfer with oxorhenium(V) oxazoline complexes. Inorganic Chemistry. 2001;40:2185-2192. DOI: 10.1021/ic000917z
  36. 36. Hoogenraad M, Kooijman H, Spek AL, Bouwman E, Haasnoot JG, Reedijk J. Manganese oxidation catalysts with 2-(2′-hydroxyphenyl)oxazoline ligands: Catalyst intermediates and degradation products—Crystal structure of bis[{2-(2′-oxazolinyl)phenolato}bis(1-methylimidazole)manganese(iii)] perchlorate. European Journal of Inorganic Chemistry. 2002:11;2897-2903
  37. 37. Hoogenraad M, Ramkisoensing K, Gorter S, Driessen WL, Bouwman E, Haasnoot JG, et al. Electrochemical and catalytic properties of novel manganese(III) complexes with substituted 2-(2′-hydroxyphenyl)oxazoline ligands—X-ray structures of tris[5-methyl-2-(2′-oxazolinyl)phenolato]manganese(III) and tris[5-chloro-2-(2′-oxazolinyl)phenolato]manganese(III). European Journal of Inorganic Chemistry. 2002:2;377-387
  38. 38. Hoogenraad M, Ramkisoensing K, Driessen WL, Kooijman H, Spek AL, Bouwman E, et al. Catalytic and electrochemical properties of new manganese(III) compounds of 2-(2′-hydroxyphenyl)-oxazoline (Hphox or HClphox). Molecular structures of [Mn(Clphox)2(MeOH)2](ClO4) and [Mn(phox)2(MeOH)2][Mn(phox)2(ClO4)2](H2O)2. Inorganica Chimica Acta. 2001;320:117-126. DOI: 10.1016/S0020-1693(01)00485-6
  39. 39. Hoogenraad M, Ramkisoensing K, Kooijman H, Spek AL, Bouwman E, Haasnoot JG, et al. Catalytic and electrochemical properties of new manganese(III) compounds of 2-(2′-hydroxyphenyl)-oxazoline (Hphox or HClphox). Molecular structures of [Mn(Clphox)2(MeOH)2](ClO4) and [Mn(phox)2(MeOH)2][Mn(phox)2(ClO4)2]’(H2O)2. Inorganica Chimica Acta. 1998;27:217-220. DOI: 10.1016/S0020-1693(98)00118-2
  40. 40. Miller KJ, Baag JH, Abu-Omar MM. Synthesis, characterization, and reactivity of palladium(ii) salen and oxazoline complexes. Inorganic Chemistry. 1999;38:4510-4514. DOI: 10.1021/ic981450j
  41. 41. Ðaković S, Liščić-Tumir L, Kirin SI, Vinković V, Raza Z, Sÿuste A, et al. Enantioselectivity in cyclopropanation catalyzed by Cu(I) complexes increased by π stacking of two monodentate oxazoline ligands. Journal of Molecular Catalysis A: Chemical. 1997;18:27-31. DOI: 10.1016/S1381-1169(96)00113-6
  42. 42. Abu-Omar MM, McPherson LD, Arias J, Béreau VM. Clean and efficient catalytic reduction of perchlorate. Angewandte Chemie, International Edition. 2000;39:4310-4332. DOI: 10.1002/1521-3773(20001201)39:23<4310::AID-ANIE4310>3.0.CO;2-D
  43. 43. Hu Y, Khan MA, Nicholas KM. Chiral Pd-, Cu- and Ni-chelates and their utilities as catalysts in allylic acetoxylation of alkenes. Journal of Molecular Catalysis A: Chemical. 1994;91:319-334. DOI: 10.1016/0304-5102(94)00037-9
  44. 44. Luo M, Zhang JC, Pang WM, Hii KK. One-step multicomponent synthesis of chiral oxazolinyl-zinc complexes. Chemistry Central Journal. 2017;11:81. DOI: 10.1186/s13065-017-0305-1
  45. 45. Luo M, Zhang JC, Yin H, Wang CM, Morris-Natschke S, Lee K-H. One-step templated synthesis of chiral organometallic salicyloxazoline complexes. BMC Chemistry. 2019;13:51. DOI: 10.1186/s13065-019-0565-z

Written By

Mei Luo

Submitted: 15 June 2019 Reviewed: 23 September 2019 Published: 08 November 2019