Next Article in Journal
Pharmacological Evaluation of Cannabinoid Receptor Modulators Using GRABeCB2.0 Sensor
Next Article in Special Issue
Experimental and Computational Studies on the Interaction of DNA with Hesperetin Schiff Base CuII Complexes
Previous Article in Journal
Flavonoid Oxidation Potentials and Antioxidant Activities-Theoretical Models Based on Oxidation Mechanisms and Related Changes in Electronic Structure
Previous Article in Special Issue
Transfer Hydrogenation of Vinyl Arenes and Aryl Acetylenes with Ammonia Borane Catalyzed by Schiff Base Cobalt(II) Complexes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Oxidovanadium(V) Schiff Base Complexes Derived from Chiral 3-amino-1,2-propanediol Enantiomers: Synthesis, Spectroscopic Studies, Catalytic and Biological Activity

by
Grzegorz Romanowski
1,*,
Justyna Budka
2 and
Iwona Inkielewicz-Stepniak
2,*
1
Faculty of Chemistry, University of Gdansk, Wita Stwosza 63, PL-80308 Gdansk, Poland
2
Department of Pharmaceutical Pathophysiology, Faculty of Pharmacy, Medical University of Gdansk, Dębinki 7, Building 27, PL-80211 Gdansk, Poland
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(9), 5010; https://doi.org/10.3390/ijms25095010
Submission received: 24 March 2024 / Revised: 30 April 2024 / Accepted: 1 May 2024 / Published: 3 May 2024

Abstract

:
Oxidovanadium(V) complexes, [(+)VOL1-5] and [(–)VOL1-5], with chiral tetradentate Schiff bases, which are products of monocondensation of S(‒)-3-amino-1,2-propanediol or R(+)-3-amino-1,2-propanediol with salicylaldehyde derivatives, have been synthesized. Different spectroscopic methods, viz. 1H and 51V NMR, IR, UV-Vis, and circular dichroism, as well as elemental analysis, have been used for their detailed characterization. Furthermore, the epoxidation of styrene, cyclohexene, and two monoterpenes, S(‒)-limonene and (‒)-α-pinene, using two oxidants, aqueous 30% H2O2 or tert-butyl hydroperoxide (TBHP) in decane, has been studied with catalytic amounts of all complexes. Finally, biological cytotoxicity studies have also been performed with these oxidovanadium(V) compounds for comparison with cis-dioxidomolybdenum(VI) Schiff base complexes with the same chiral ligands, as well as to determine the cytoprotection against the oxidative damage caused by 30% H2O2 in the HT-22 hippocampal neuronal cells in the range of their 10–100 μM concentration.

1. Introduction

Vanadium is an abundant element in a number of various biological systems [1]. Many examples of vanadium enzymes and compounds can be found in nature. Among them are vanadium nitrogenases found in bacteria [2,3], vanadium-dependent haloperoxidases in terrestrial fungi and lichen or marine algae [4,5,6], or natural compound amavadin found in the mushroom Amanita muscaria [7]. Catalytic activity studies of vanadium complexes have been stimulated recently by their importance in modeling the active centers of such enzymes in many biological processes [4,8] and towards diverse organic transformations [9,10]. Salicylaldimine-derived chiral Schiff bases very easily coordinate with vanadium ions and exhibit widespread applications [11,12]. Among them, chiral Schiff bases derived from N-salicyl-β-amino alcohols are considered “privileged ligands” with fine-tuned structural and electronic properties [13,14]. Multidentate unsymmetrical vanadium(V) Schiff base complexes with such chiral ligands have been widely used as catalysts in the asymmetric alkynylation of aldehydes [15], enantioselective oxidation of organic sulfides [15,16,17], stereoselective synthesis of tetrahydrofurans [18,19], epoxidation of alkenes [20,21], and oxidation of bromide [18]. It is noteworthy that nowadays, epoxidation of various olefins is a widely studied reaction in organic chemistry since epoxides have commonly been used as important intermediates for the synthesis of a wide range of commercial products, as well as useful pharmaceuticals in medicine [22]. Moreover, the formation of new carbon–carbon bonds by ring-opening reactions make optically pure epoxides especially valuable key intermediates and products in different areas of organic chemistry [23].
Vanadium complexes have extensively been reported as prospective medicinal pharmaceutics [24], mentioning their wide biological activities. Lately, their therapeutic potential against pneumonia [25] and tuberculosis bacteria [26] has been found, as well as SARS-CoV-2 [27] and HIV [28] viruses. Moreover, the interest of scientists has lately been increasingly observed in studies on various vanadium complexes possessing anticancer [29], antiparasitic [30], anti-inflammatory [31], or antidiabetic [32,33] properties, an ability that is also cytoprotective against the oxidative damage made by reactive oxygen species to hippocampal cells [34,35]. In various neurodegenerative diseases, such as stroke and brain trauma but also Alzheimer’s and Parkinson’s diseases, oxidative stress is a very consistent component. An excellent model cell for developing new innovative cytoprotective drugs and studying the consequences of endogenous oxidative stress is the hippocampal cell line HT-22 [36].
New oxidovanadium(V) complexes with chiral tetradentate Schiff base, products of monocondensation reaction of substituted salicylaldehyde derivatives with chiral amino alcohols, i.e. R(+)-3-amino-1,2-propanediol and S(‒)-3-amino-1,2-propanediol (Figure 1) are described within this paper. Their structures were confirmed by spectroscopic techniques, i.e., 1H and 51V NMR, IR, UV-Vis, and circular dichroism, as well as elemental analysis. Furthermore, their catalytic activity has been studied in the epoxidation of alkenes, i.e., styrene, cyclohexene, and naturally occurring monoterpenes, i.e., S(−)-limonene and (−)-α-pinene, during the oxidation reaction by 30% H2O2 or tert-butyl hydroperoxide. Biological cytotoxicity studies have also been performed with these oxidovanadium(V) compounds and for comparison with cis-dioxidomolybdenum(VI) complexes with the same chiral ligands. Finally, their cytoprotective properties have also been studied against the oxidative damage caused by 30% H2O2 in the HT-22 hippocampal neuronal cells in the range of their 10–100 μM concentration.

2. Results and Discussion

2.1. Infrared Spectra

In IR spectra of all chiral oxidovanadium(V) complexes, [(+)VOL1-5] and [(−)VOL1-5], strong and sharp bands appear in the 956–998 cm−1 region, which is characteristic of single V=O stretching modes. They are similar to the values found for the other oxidovanadium(V) Schiff base complexes [37]. Furthermore, characteristic imine C=N stretching vibrations at 1640–1652 cm−1 indicate the coordination of chiral Schiff base ligands to the vanadium(V) ion [38]. The deprotonation of all coordinated hydroxyl (phenolic and alcoholic) groups of Schiff base ligands is proved by the lack of any absorption bands in the range of 3200–3700 cm−1. Moreover, the coordination of all these phenolate and alkoxide oxygens to vanadium atoms is shown by strong asymmetric and symmetric ν(C-O) vibrations at 1277–1316 and 1046–1095 cm−1, respectively (Figure S1). In the case of [(+)VOL5] and [(−)VOL5], with nitro substituent in salicylaldimine fragment, asymmetric and symmetric ν(NO2) vibrations have been noticed at ca. 1560 and 1340 cm−1, respectively.

2.2. UV-Vis and Circular Dichroism Spectra

UV-Vis spectra of all oxidovanadium(V) complexes show the low-energy absorption bands belonging to electron LMCT transitions (Figure S2) from phenolate oxygen pπ orbital to an empty d orbital of vanadium atom, appearing between 339 and 359 nm [39]. The circular dichroism spectra have also revealed two bands from intraligand π‒π* transitions in 278–281, 304–322, and additional bands in the 362–390 nm regions of the same origin as it was attributed for UV-Vis spectra. Moreover, in the case of [(+)VOL1-5], the latter π‒π* transitions and LMCT transitions have the same positive Cotton effects in contrast to [(−)VOL1-5] with negative Cotton effects for these bands. Finally, it is noteworthy that all absorption bands exhibit exactly opposite Cotton effects when comparing [(+)VOL1-5] and [(−)VOL1-5] compounds, which possess asymmetric carbons in amino alcohol moiety with R and S configuration, respectively (Figure S3).

2.3. NMR Measurements

For characterization of all oxidovanadium(V) Schiff base complexes, 1H and 51V NMR spectra were recorded in DMSO-d6. Schiff bases complexes derived from S(‒)-3-amino-1,2-propanediol and R(+)-3-amino-1,2-propanediol show identical spectra; therefore, only the latter compounds have been measured. Successful condensation of all salicylaldehyde derivatives with each chiral amino alcohol has been proved by the presence of azomethine proton signals in the case of all complexes. The two-dimensional NMR experiments (COSY and NOESY) were performed to establish proximity and connection between all protons but also for the assignment and identification of all proton signals in the case of the [(+)VOL4] compound. 1H spectrum, with the help of the COSY experiment (Figure S4), unambiguously reveal a connection between methine proton (signal at 5.19 ppm) with all four methylene protons doublets of doublets. Furthermore, for the signals between two methylene protons neighboring an oxygen atom (4.89 and 5.44 ppm), as well as two methylene protons attached to a nitrogen atom (3.75 and 4.74 ppm), the cross-peaks have been found. Moreover, a cross-peak between two aromatic protons (6.71 and 7.52 ppm) has been found showing their close neighborhood. On the other hand, the information regarding the close spatial proximity between the 4.74 ppm methylene doublet of doublets and azomethine proton signal at 8.70 ppm is given by a cross-peak in NOESY spectrum [40], as well as another cross-peak between azomethine proton signal and one of the aromatic proton signal at 7.71 was useful to assign aromatic ring protons signals. During the 51V NMR experiments single signals in the range of -529.8 to -532.9 ppm have been noticed, indicating the presence of oxidovanadium(V) complexes only in the monomeric forms.

2.4. Catalytic Oxidation Studies

Oxidovanadium(V) complexes obtained in this paper, i.e., [(+)VOL1-5] and [(−)VOL1-5], have been checked as catalysts in the oxidation of alkenes, such as styrene, cyclohexene, and additionally, naturally occurring monoterpenes, i.e., S(−)-limonene and (−)-α-pinene (Figure 2). As the best solvent for these oxidation reactions, 1,2-dichloroethane (DCE) was chosen with tert-butyl hydroperoxide (TBHP) as the terminal oxidant, compared to the other common solvents like acetonitrile, toluene, ethanol, methanol, chloroform, and methylene chloride. According to our observation, for reflux conditions, the lower reaction temperature of the latter solvents is responsible for the poorer conversions of the substrates. On the other hand, when 30% H2O2 was employed to avoid a biphasic system, acetonitrile was chosen as the best solvent. The oxidation reactions with all substrates needed 5 h to reach completion, and to achieve the best conversions and selectivities, the reactions required solvents with higher reaction temperatures for their reflux conditions, as it was similarly concluded previously [41]. The effect of various reaction parameters was taken into account to achieve appropriate reaction conditions for the best conversions. For this purpose, each of the oxidants’ molar ratios to the substrate has been compared, i.e., 1:1, 2:1, 3:1, and 4:1, as well as the different amounts of catalyst used, i.e., 0.5, 1, 2, and 3 mol%, during our test reactions. Our observations have led to the conclusion that the 3:1 molar ratio of oxidant to each substrate is optimal, and 1 mol% loading of catalyst is a sufficient amount for running the oxidations in optimized reaction conditions.
The oxidation of styrene using 30% H2O2 as the oxidant aqueous or TBHP as the terminal oxidant generally can lead to five oxidation products, as was concluded by our previous observations [42], i.e., styrene oxide, phenylacetaldehyde, 1-phenylethane-1,2-diol, benzaldehyde, and benzoic acid (Figure 3). In the first step of the epoxidation reaction, styrene oxide is usually formed. Furthermore, its very fast conversion, via a nucleophilic attack of an excess of oxidant and the cleavage of hydroperoxystyrene as an intermediate product, results in the formation of benzaldehyde [43]. On the other hand, a radical mechanism can be responsible for the oxidative cleavage of the styrene side-chain double bond and the direct formation of benzaldehyde and its further oxidation to benzoic acid. The presence of water in the oxidant, such as 30% H2O2, can be blamed for the very low conversion of styrene but also for the hydrolysis of styrene oxide, resulting in the formation of 1-phenylethane-1,2-diol and the decomposition of the catalyst. Moreover, the formation of phenylacetaldehyde is also possible by way of styrene oxide isomerization.
In Table 1, the results of the oxidation of styrene by aqueous 30% H2O2 as the terminal oxidant with catalytic amounts of [(+)VOL1] and [(−)VOL1] are presented, which have shown only low conversions (25 and 27%) but with higher selectivity towards styrene oxide than benzaldehyde. The selectivity towards styrene oxide significantly decreases with TBHP, and the conversion of styrene increased to 86–96%. The formation of styrene oxide and benzaldehyde as major products was observed with both oxidants but without considerable amounts of any by-products. Our previous studies with other vanadium(V) catalysts with chiral amino alcohol-derived tridentate Schiff bases have led to similar outcomes [40,42].
Recent reports in the catalytic epoxidation of styrene by vanadium(V) complexes with Schiff base ligands, i.e., [VO2(sal-aebmz)], [VO2(sal-ambmz)], [VO2(acac-ambmz)] [44], and [VO(hap-dahp)] [45], have shown low selectivity (2–11%) and conversion from 51 to 78% for the most desired product, i.e., styrene oxide, and the formation of benzaldehyde in high amounts with 30% H2O2 as the terminal oxidant, even 90% in the case of [VO2(acac-aebmz)]. In contrast to the [(+)VOL1-5] and [(−)VOL1-5] catalysts, conversion considerably decreased to as low as 20% with TBHP but selectivity was comparable. In the case of [VO2(acac-ambmz)], the change was even up to 47% towards styrene oxide. Moreover, a comparison of dioxidovanadium(V) salicylaldehyde semicarbazone complexes with different substituents in salicyl moiety shows much lower conversions using 30% H2O2 and 70% TBHP [46]. Surprisingly, when para-substituted styrene derivatives were used as substrates, no product formation was observed. The best results for catalytic oxidation of alkenes were found for oxidovanadium(V) complexes with tridentate Schiff bases [38].
The catalytic potential of [(+)VOL1-5] and [(−)VOL1-5] complexes has also been studied for the oxidation of cyclohexene by aqueous 30% H2O2 or tert-butyl hydroperoxide (TBHP), resulting in the epoxidation products, i.e., cyclohexene oxide and cyclohexene-1,2-diol, after a continued hydrolysis reaction, as well as the formation of products of allylic oxidation, i.e., 2-cyclohexen-1-ol and 2-cyclohexen-1-one (Figure 4). In Table 2, we presented the results of the conversion of cyclohexene and the selectivity towards various oxidation products. When 30% H2O2 as the terminal oxidant was used, the [(+)VOL1] and [(−)VOL1] catalysts gave only 42 and 44% conversion, respectively, under the previously mentioned optimized reaction conditions. It is noteworthy that the cyclohexene oxidation with H2O2, reported for the oxidovanadium(V) Schiff base complex and derived from 3-hydroxy-2 naphthohydrazide [47], have shown distinctly higher conversion (92%) and selectivity towards the allylic oxidation product, cyclohexen-1-ol (55%), but low amounts of cyclohexene oxide (29%). In a comparison with the series of vanadium(V) complexes, i.e., [VO(μ2-OCH3)(L1)]2, [VO2(L2)], H2O, and [VO2(L3)] with hydrazone ONO and NNS donor Schiff base ligands, a similar outcome was achieved [48]. In contrast to H2O2, conversions were considerably higher, up to 92%, when TBHP was used with catalytic amounts of [(+)VOL1-5] and [(−)VOL1-5], and the selectivity was higher towards 2-cyclohexene-1-one formation than cyclohexene oxide. The preferential attack of the activated C‒H bond over the C=C bond may be responsible for the creation of allylic oxidation products in higher selectivity [49].
Naturally occurring mono- or bicyclic monoterpenes with a cyclohexene ring, i.e., S(−)-limonene and (−)-α-pinene, used in these catalytic studies gave analogous epoxidation and allylic oxidation products. In comparison to cyclohexene, slightly lower conversions were obtained for S(−)-limonene oxidation by 30% H2O2; the main reaction product was epoxide, up to 75%. Surprisingly, the formation of allylic oxidation products was not observed, but significant amounts of diepoxide as a by-product may be created due to additional exocyclic isopropenyl fragments, which are present in this monoterpene (Table 3). When TBHP was used as the oxygen source, lower amounts of S(−)-limonene oxide were observed, but with excellent conversions, even up to 100%. The oxidation of (−)-α-pinene has given distinctly lower conversions than in the case of S(−)-limonene (Table 4), but significantly higher selectivity was found towards (−)-α-pinene oxide (up to 83% with TBHP and 86% with 30% H2O2 in the case of [(−)VOL1]).

2.5. Biological Studies

Oxidovanadium(V) complexes [(+)VOL1], [(+)VOL3], [(+)VOL4], [(−)VOL3], and [(−)VOL4] have been studied to establish their cytotoxic and cytoprotective abilities on the hippocampal neuronal cell line HT-22. Moreover, for comparison, the cis-dioxidomolybdenum(VI) Schiff base complexes derived from the same chiral tetradentate amino alcohols [50], i.e., [(+)MoO2(HL1)(CH3OH)] and [(+)MoO2(HL3)], were also used in this study.
The concentration-dependent effect of the oxidovanadium(V) and dioxidomolybdenum(VI) compounds, as well as their cytotoxicity, on the viability of the hippocampal neuronal cell line HT-22 in the range of 10–100 μM concentration, has been investigated, as well as their cytoprotective ability against 500 µM H2O2-induced oxidative damage [51].
The MTT assay results of oxidovanadium(V) and cis-dioxidomolybdenum(VI) tetradentate Schiff base complexes were compared and showed different influences in their cytotoxicity on the HT-22 cell line (Figure 5). In the case of [(+)MoO2(HL1)(CH3OH)] and [(+)MoO2(HL3)], their 10 and 25 μM concentrations have not exhibited cytotoxic effects on the viability of tested cells. Furthermore, it has been observed that a 50 μM concentration reduces the viability of cells below 70 and 80%, respectively, and their 100 μM concentrations lead to a larger decrement, even to 55% in the case of [(+)MoO2(HL1)(CH3OH)]. The MTT assay for [(+)VOL3], [(+)VOL4], [(−)VOL3], and [(−)VOL4] have disclosed no significant impact on the viability of HT-22 cell line only at a 10 μM dose, although in case of [(+)VOL1], there was decrease below 80%. Higher concentrations of vanadium(V) compounds caused a further decrease in viability, even below 30% for 100 μM doses, which revealed their higher cytotoxic effect in comparison to cis-dioxidomolybdenum(VI) Schiff base complexes.
Cytoprotective activity experiments of all complexes were performed with a 500 µM dose of 30% H2O2 to induce a cellular injury in the hippocampal cells. After 24 h, exposure of HT-22 cells to such an amount of H2O2 results in cellular death of about 24% due to oxidative stress (Figure 6). [(+)MoO2(HL1)(CH3OH)] and [(+)MoO2(HL3)], which exhibited no significant cytotoxicity at 10 and 25 μM concentrations, have been tested only in these doses in experiments revealing their cytoprotective properties. These cis-dioxidomolybdenum(VI) complexes have shown strong cytoprotective properties, causing the highest increase, up to 62%, in the viability of HT-22 cells (10 μM dose) and 55% (25 μM dose) for [(+)MoO2(HL3)]. Furthermore, the addition of [(+)MoO2(HL1)(CH3OH)] with 500 µM H2O2 has a weaker influence on the viability of hippocampal cells, i.e., 57% at a 10 μM dose and 49% at a 25 μM dose.
In the screening of cytotoxic effects of [(+)VOL1], [(+)VOL3], [(+)VOL4], [(−)VOL3], and [(−)VOL4], no considerable effect on the viability of HT-22 cells at only a 10 μM concentration was observed. It is noteworthy that only [(−)VOL3] did not distincly decrease the viability at a 25 μM dose, just slightly below 80%. On the contrary, potassium bisperoxo(1,10-phenanthroline)oxovanadate complex (BpV), which is a specific inhibitor of protein tyrosine phosphatases, has shown much higher cytotoxicity after 24 h with below 60 and even 20% cell viability at 3 and 30 μM concentrations, respectively [52]. All investigated oxidovanadium(V) complexes exhibited a cytoprotective effect against oxidative damage only at 10 μM doses. Unexpectedly, in contrast to cis-dioxidomolybdenum(VI) complexes, the weakest cytoprotective properties have been found in the case of [(+)VOL3] and [(−)VOL3] without any significant differences between these oxidovanadium(V) complexes with Schiff base possessing opposite chirality. Moreover, [(−)VOL3] at a 25 μM concentration did not show any cytoprotective activity. The highest cytoprotection at the mitochondrial level in the MTT test occurred for [(+)VOL4], with up to 60% of HT-22 cell viability for a 10 μM dose. The higher cytotoxicity of vanadium(V) complexes is very likely caused by the formation of peroxidovanadium(V) species after the addition of 30% H2O2, in comparison to cis-dioxidomolybdenum(VI) Schiff base complexes derived from the same chiral tetradentate amino alcohols and can be the reason for the stronger harmful effect on hippocampal cells [53]. Such properties of various vanadium compounds may emphasize their importance as promising drug candidates in chemotherapy and targeted cancer therapies, with possible relatively strong cytotoxicity and antimetastatic activity [54].
Morphological changes and cell survival in HT-22 cells have been observed on microscopic images comparing control cells (Figure 7a) with the same cells after treatment at 30% H2O2 (500 µM), which manifest themselves in irregular shapes of the hippocampal cells and shrink (Figure 7b). This is also true for the same cells pre-treated with cis-dioxidomolybdenum(VI) (Figure 7c) and oxidovanadium(V) (Figure 7d) complexes, which show the best cytoprotective ability.

3. Materials and Methods

3.1. Measurements

All reagents for synthesis of the complexes, i.e., chiral amino alcohols, salicylaldehyde derivatives, and vanadium(V) oxytripropoxide were purchased from Aldrich company. The other chemicals were obtained from local sources and used without further purification. Carlo Erba MOD 1106 instrument was used to perform elemental analyses. UV-Vis spectra were conducted using a Perkin-Elmer LAMBDA 18 spectrophotometer, and CD spectra were conducted using a Jasco J-815 spectropolarimeter. IR spectra were recorded on Bruker IFS 66 as KBr pellets. Bruker AVANCE III 700 MHz spectrometer was employed for all NMR spectra measurements using TMS as a reference and DMSO-d6 as a solvent. The catalytic reactions progress was monitored on a Shimadzu GC-2025 gas chromatograph with an FID detector, and Zebron ZB-5 capillary column and Shimadzu GCMS-QP2010 SE equipment were used for the confirmation of the identities of all oxidation products. The absorbance of the cells during the MTT assay was measured at 570 nm and 660 nm (reference value) by an ASYS Hitech GmbH microplate reader. Microscopic images were taken at 1 × 4 magnification with an Olympus microscope.

3.2. Synthesis of Oxidovanadium(V) Complexes

All vanadium(V) complexes were synthesized employing the same procedure. To a solution of 1 mmol of chiral amino alcohol, i.e., R(+)-3-amino-1,2-propanediol or S(−)-3-amino-1,2-propanediol in MeOH (10 mL), 1 mmol of one of the following aromatic o-hydroxyaldehyde, i.e., 3-methoxysalicylaldehyde, 5-methoxysalicylaldehyde, 5-methylsalicylaldehyde, 5-bromosalicylaldehyde, or 5-nitrosalicylaldehyde in 10 mL of MeOH, was added. Then, reaction mixture was stirred and heated under reflux for 1 h. After addition of vanadium(V) oxytripropoxide (1 mmol) in MeOH (10 mL), reactants in solution were stirred under reflux for next 2 h. Obtained precipitates were filtered off and washed several times with cold MeOH. All complexes are crystalline solids (Figure S5).
[(+)VOL1]: Yield 83%, brown. Anal. Calc. for C11H12NO5V: C, 45.7; H, 4.2; N, 4.8. Found: C, 45.8; H, 4.1; N, 4.8%. IR (KBr, cm−1): 1642 (νC=N); 1598 (νC=C); 1286, 1082 (νC-O); 968 (νV=O). UV-Vis spectrum in DMSO [λmax (nm), ɛ (M−1 cm−1)]: 351 (3200). CD spectrum in DMSO [λmax (nm), Δɛ (M−1 cm−1)]: 280 (−0.40), 305 (4.19), 374 (1.97). 1H NMR (DMSO-d6, ppm): 8.69 (1H, s) (azomethine); 7.16 (1H, d, 3J = 8.4 Hz), 7.13 (1H, d, 3J = 7.8 Hz), 6.87 (1H, t, 3J = 8.4 Hz) (aromatic); 5.16 (1H, m) (methine); 5.42 (1H, dd, 3J = 11.2 Hz, 4J = 2.8 Hz), 4.86 (1H, dd, 3J = 11.2 Hz, 4J = 5.5 Hz), 4.72 (1H, dd, 3J = 13.1 Hz, 4J = 5.5 Hz), 3.73 (1H, dd, 3J = 13.1 Hz, 4J = 5.5 Hz) (methylene), 3.82 (3H, s) (methoxy). 51V NMR (DMSO-d6, ppm): −531.8.
[(+)VOL2]: Yield 78%, brown. Anal. Calc. for C11H12NO5V: C, 45.7; H, 4.2; N, 4.8. Found: C, 45.7; H, 4.3; N, 4.7%. IR (KBr, cm−1): 1650 (νC=N); 1612 (νC=C); 1293, 1095 (νC-O); 980 (νV=O). UV-Vis spectrum in DMSO [λmax (nm), ɛ (M−1 cm−1)]: 359 (3540). CD spectrum in DMSO [λmax (nm), Δɛ (M−1 cm−1)]: 280 (−0.40), 310 (3.86), 379 (1.53). 1H NMR (DMSO-d6, ppm): 8.68 (1H, s) (azomethine); 7.18 (1H, d, 4J = 2.8 Hz), 7.15 (1H, dd, 3J = 8.8 Hz, 4J = 2.8 Hz), 6.84 (1H, d, 3J = 8.8 Hz) (aromatic); 5.18 (1H, m) (methine); 5.40 (1H, dd, 3J = 11.2 Hz, 4J = 2.8 Hz), 4.84 (1H, dd, 3J = 11.2 Hz, 4J = 5.5 Hz), 4.71 (1H, dd, 3J = 13.1 Hz, 4J = 5.5 Hz), 3.72 (1H, dd, 3J = 13.1 Hz, 4J = 5.5 Hz) (methylene), 3.81 (3H, s) (methoxy). 51V NMR (DMSO-d6, ppm): −532.9.
[(+)VOL3]: Yield 85%, dark red. Anal. Calc. for C11H12NO4V: C, 48.4; H, 4.4; N, 5.1. Found: C, 48.2; H, 4.5; N, 5.1%. IR (KBr, cm−1): 1640 (νC=N); 1608 (νC=C); 1286, 1058 (νC-O); 979 (νV=O). UV-Vis spectrum in DMSO [λmax (nm), ɛ (M−1 cm−1)]: 339 (3480). CD spectrum in DMSO [λmax (nm), Δɛ (M−1 cm−1)]: 279 (−2.10), 309 (3.33), 370 (2.89). 1H NMR (DMSO-d6, ppm): 8.69 (1H, s) (azomethine); 7.38 (1H, s), 7.31 (1H, d, 3J = 9.2 Hz), 6.84 (1H, d, 3J = 8.3 Hz) (aromatic); 5.15 (1H, m) (methine); 5.35 (1H, dd, 3J = 11.0 Hz, 4J = 2.7 Hz), 4.86 (1H, dd, 3J = 11.0 Hz, 4J = 5.5 Hz), 4.71 (1H, dd, 3J = 13.0 Hz, 4J = 5.5 Hz), 3.81 (1H, dd, 3J = 13.0 Hz, 4J = 5.5 Hz) (methylene). 2.31 (3H, d, 3J = 5.3) (methyl). 51V NMR (DMSO-d6, ppm): −532.5.
[(+)VOL4]: Yield 80%, orange. Anal. Calc. for C10H9BrNO4V: C, 35.5; H, 2.7; N, 4.1. Found: C, 35.6; H, 2.7; N, 4.0%. IR (KBr, cm−1): 1646 (νC=N); 1587 (νC=C); 1280, 1048 (νC-O); 980 (νV=O). UV-Vis spectrum in DMSO [λmax (nm), ɛ (M−1 cm−1)]: 340 (2860). CD spectrum in DMSO [λmax (nm), Δɛ (M−1 cm−1)]: 281 (−1.47), 322 (1.79), 367 (2.67). 1H NMR (DMSO-d6, ppm): 8.70 (1H, s) (azomethine); 7.71 (1H, d, 3J = 2.6 Hz), 7.52 (1H, dd, 3J = 8.9 Hz, 4J = 2.6 Hz), 6.71 (1H, d, 3J = 8.9 Hz) (aromatic); 5.19 (1H, m) (methine); 5.44 (1H, dd, 3J = 11.2 Hz, 4J = 2.8 Hz), 4.89 (1H, dd, 3J = 11.2 Hz, 4J = 5.5 Hz), 4.74 (1H, dd, 3J = 13.1 Hz, 4J = 5.5 Hz), 3.75 (1H, dd, 3J = 13.1 Hz, 4J = 5.5 Hz) (methylene). 51V NMR (DMSO-d6, ppm): −531.7.
[(+)VOL5]: Yield 88%, brown. Anal. Calc. for C10H9N2O6V: C, 39.5; H, 3.0; N, 9.2. Found: C, 39.4; H, 3.1; N, 9.3%. IR (KBr, cm−1): 1641 (νC=N); 1607 (νC=C); 1558, 1339 (νNO2); 1316, 1049 (νC-O); 976 (νV=O). UV-Vis spectrum in DMSO [λmax (nm), ɛ (M−1 cm−1)]: 355 (12060). CD spectrum in DMSO [λmax (nm), Δɛ (M−1 cm−1)]: 279 (-2.03), 304 (3.16), 390 (2.74). 1H NMR (DMSO-d6, ppm): 8.68 (1H, s) (azomethine); 8.46 (1H, dd, 3J = 8.4 Hz, 4J = 2.9 Hz), 8.28 (1H, dd, 3J = 8.4 Hz, 4J = 2.9 Hz), 6.96 (1H, d, 3J = 8.4 Hz) (aromatic); 5.36 (1H, m) (methine); 5.48 (1H, dd, 3J = 11.2 Hz, 4J = 2.8 Hz), 4.92 (1H, dd, 3J = 11.2 Hz, 4J = 5.5 Hz), 4.75 (1H, dd, 3J = 13.1 Hz, 4J = 5.5 Hz), 3.77 (1H, dd, 3J = 13.1 Hz, 4J = 5.5 Hz) (methylene). 51V NMR (DMSO-d6, ppm): −529.8.
[(−)VOL1]: Yield 87%, brown. Anal. Calc. for C11H12NO5V: C, 45.7; H, 4.2; N, 4.8. Found: C, 45.6; H, 4.2; N, 4.8%. IR (KBr, cm−1): 1643 (νC=N); 1599 (νC=C); 1277, 1081 (νC-O); 956 (νV=O). UV-Vis spectrum in DMSO [λmax (nm), ɛ (M−1 cm−1)]: 350 (3360). CD spectrum in DMSO [λmax (nm), Δɛ (M−1 cm−1)]: 278 (1.53), 307 (−6.70), 362 (−2.27).
[(−)VOL2]: Yield 82%, brown. Anal. Calc. for C11H12NO5V: C, 45.7; H, 4.2; N, 4.8. Found: C, 45.6; H, 4.3; N, 4.8%. IR (KBr, cm−1): 1650 (νC=N); 1612 (νC=C); 1278, 1084 (νC-O); 998 (νV=O). UV-Vis spectrum in DMSO [λmax (nm), ɛ (M−1 cm−1)]: 358 (3420). CD spectrum in DMSO [λmax (nm), Δɛ (M−1 cm−1)]: 280 (0.98), 310 (−3.72), 379 (−1.43).
[(−)VOL3]: Yield 86%, dark red. Anal. Calc. for C11H12NO4V: C, 48.4; H, 4.4; N, 5.1. Found: C, 48.3; H, 4.4; N, 5.0%. IR (KBr, cm−1): 1641 (νC=N); 1584 (νC=C); 1286, 1058 (νC-O); 978 (νV=O). UV-Vis spectrum in DMSO [λmax (nm), ɛ (M−1 cm−1)]: 340 (3560). CD spectrum in DMSO [λmax (nm), Δɛ (M−1 cm−1)]: 278 (2.24), 308 (−4.14), 365 (−2.58).
[(−)VOL4]: Yield 84%, orange. Anal. Calc. for C10H9BrNO4V: C, 35.5; H, 2.7; N, 4.1. Found: C, 35.4; H, 2.8; N, 4.0%. IR (KBr, cm−1): 1647 (νC=N); 1589 (νC=C); 1280, 1047 (νC-O); 981 (νV=O). UV-Vis spectrum in DMSO [λmax (nm), ɛ (M−1 cm−1)]: 342 (2940). CD spectrum in DMSO [λmax (nm), Δɛ (M−1 cm−1)]: 280 (1.34), 318 (−2.18), 366 (−2.48).
[(−)VOL5]: Yield 81%, brown. Anal. Calc. for C10H9N2O6V: C, 39.5; H, 3.0; N, 9.2. Found: C, 39.6; H, 3.1; N, 9.2%. IR (KBr, cm−1): 1652 (νC=N); 1607 (νC=C); 1560, 1336 (νNO2); 1305, 1046 (νC-O); 956 (νV=O). UV-Vis spectrum in DMSO [λmax (nm), ɛ (M−1 cm−1)]: 355 (12330). CD spectrum in DMSO [λmax (nm), Δɛ (M−1 cm−1)]: 278 (1.76), 304 (−2.58), 390 (−2.13).

3.3. Catalytic Activity

All oxidovanadium(V) complexes have been used as catalysts in the oxidation of cyclohexene, styrene, S(−)-limonene, and (−)-α-pinene using as an oxidant aqueous 30% H2O2 or 5.5 M solution of tert-butyl hydroperoxide (TBHP) in decane. All reactions were optimized with different reaction condition parameters. The optimal reaction temperature was 80 °C, and molar ratio of catalyst, substrate, and oxidant was 0.01:1:2. The best solvent for catalytic reactions with TBHP was 1,2-dichloroethane (DCE), and acetonitrile for reaction performed with H2O2 as the terminal oxidant.

3.4. Biological Activity

3.4.1. The Cell Culture and Treatments

The mouse hippocampal neuronal cell line HT-22 was cultured in standard conditions according to protocol described by our team previously [34,35]. Working solutions at 10, 25, 50, and 100 µM concentrations were prepared ex tempore in serum-free medium DMEM. The measurements of the viability and cytoprotective action of investigated compounds on HT-22 cells after 24 h incubation were tested by mitochondrial dehydrogenase activity (MTT) assay.

3.4.2. MTT Assay

HT-22 cells were passaged at 8 × 103 cells in well of a 96-well plate and incubated for 24 h in standard conditions. HT-22 cells were treated later for the next 24 h with selected concentrations of the investigated compounds, which were not cytotoxic. In the cytoprotective activity experiments, vanadium(V) and cis-dioxidomolybdenum(VI) complexes were added to the HT-22 cells for 1 h and then treated for 24 h with 500 µM H2O2. MTT assay was performed according to previously described protocol by our team [34,35].

3.4.3. Microscopy

HT-22 cells were seeded in 12-well plates. The plate was cultivated for 24 h under standard conditions. HT-22 cells were treated with investigated complexes for 1 h and then, after addition of 500 µM H2O2, incubated for 24 h.

4. Conclusions

Within this paper, we present the synthesis of a series of oxidovanadium(V) complexes derived from chiral tetradentate Schiff bases, [(+)VOL1-5] and [(−)VOL1-5], products of a single condensation of salicylaldehyde derivatives with R(+)-3-amino-1,2-propanediol or S(−)-3-amino-1,2-propanediol, which have been further characterized by one- and two-dimensional NMR spectroscopy, as well as IR, UV-Vis, and CD techniques.
The catalytic potential of [(+)VOL1-5] and [(−)VOL1-5] complexes have also been studied in the epoxidation of styrene, cyclohexene, S(−)-limonene, and (−)-α-pinene in the presence of 30% H2O2 or tert-butyl hydroperoxide as the terminal oxidant. The conversion of all these olefins was distinctly lower when 30% H2O2 was used, probably due to its strong oxidizing nature and the presence of water, usually responsible for the hydrolysis of epoxides and the decomposition of catalyst. Generally, the catalytic amounts of all vanadium(V) complexes revealed high conversions and selectivities towards corresponding epoxides, especially in the case of monoterpenes in a non-aqueous environment when tert-butyl hydroperoxide (TBHP) as the oxidant. In some cases, the highest selectivities showed catalysts with electron-donating substituents in salicyladimine moiety.
The biological cytotoxicity of oxidovanadium(V) and cis-dioxidomolybdenum(VI) complexes derived from the same chiral Schiff bases, as well as cytoprotection against the oxidative damage caused by H2O2 towards the hippocampal neuronal cell line HT-22, based on an assessment of mitochondrial activity in MTT tests, has also been studied. The comparison of the cytotoxic and cytoprotective results between cis-dioxidomolybdenum(VI) and oxidovanadium(V) complexes has shown significantly higher cytotoxic activity for the latter compounds. The best cytoprotective activity has shown cis-dioxidomolybdenum(VI) complexes with the 5-methyl substituent in salicylaldimine fragment and, in the case of oxidovanadium(V) compounds, with the 5-bromo electron donating group.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms25095010/s1.

Author Contributions

Conceptualization, G.R. and I.I.-S.; Investigation, G.R. and J.B.; Methodology, G.R., J.B. and I.I.-S.; Writing—original draft, G.R.; Writing—review and editing, I.I.-S. All authors have read and agreed to the published version of the manuscript.

Funding

This scientific work was supported by the grant ST-54 from the Medical University of Gdansk and the Polish Ministry of Science and Higher Education (Grant No. 531-T050-D840-24).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tracey, A.S.; Willsky, G.R.; Takeuchi, E.S. Vanadium: Chemistry, Biochemistry, Pharmacology and Practical Applications; Chapter 10; CRC Press: Boca Raton, FL, USA, 2007. [Google Scholar]
  2. Eady, R.R. Current status of structure function relationships of vanadium nitrogenase. Coord. Chem. Rev. 2003, 237, 23–30. [Google Scholar] [CrossRef]
  3. Antipov, A.N. Vanadium in Live Organisms; Kretsinger, R.H., Uversky, V.N., Permyakov, E.A., Eds.; Encyclopedia of Metalloproteins; Springer: New York, NY, USA, 2013. [Google Scholar]
  4. Carter-Franklin, J.N.; Parrish, J.D.; Tschirret-Guth, R.A.; Little, R.D.; Butler, A. Vanadium haloperoxidase-catalyzed bromination and cyclization of terpenes. J. Am. Chem. Soc. 2003, 125, 3688–3689. [Google Scholar] [CrossRef] [PubMed]
  5. Rehder, D.; Santoni, G.; Licini, G.M.; Schulzke, C.; Meier, B. The medicinal and catalytic potential of model complexes of vanadate-dependent haloperoxidases. Coord. Chem. Rev. 2003, 237, 53–63. [Google Scholar] [CrossRef]
  6. Littlechild, J.A.; Garcia-Rodriguez, E. Structural studies on the dodecameric vanadium bromoperoxidase from Corallina species. Coord. Chem. Rev. 2003, 237, 65–76. [Google Scholar] [CrossRef]
  7. Rehder, D. Bioinorganic Vanadium Chemistry; John Wiley & Sons Ltd.: Chichester, UK, 2008. [Google Scholar]
  8. Weyand, M.; Hecht, H.J.; Kiess, M.; Liaud, M.F.; Vilter, H.; Schomburg, D. X-ray structure determination of a vanadium-dependent haloperoxidase from Ascophyllum nodosum at 2.0 Å resolution. J. Mol. Biol. 1999, 293, 595–611. [Google Scholar] [CrossRef] [PubMed]
  9. Ligtenbarg, A.G.J.; Hage, R.; Feringa, B.L. Catalytic oxidations by vanadium complexes. Coord. Chem. Rev. 2003, 237, 89–101. [Google Scholar] [CrossRef]
  10. Bolm, C. Vanadium-catalyzed asymmetric oxidations. Coord. Chem. Rev. 2003, 237, 245–256. [Google Scholar] [CrossRef]
  11. Hashmi, K.; Gupta, S.; Siddique, A.; Khan, T.; Joshi, S. Medicinal applications of vanadium complexes with Schiff bases. J. Trace Elem. Med. Biol. 2023, 79, 127245. [Google Scholar] [CrossRef] [PubMed]
  12. Langeslay, R.R.; Kaphan, D.M.; Marshall, C.L.; Stair, P.C.; Sattelberger, A.P.; Delferro, M. Catalytic applications of vanadium: A mechanistic perspective. Chem. Rev. 2019, 119, 2128–2191. [Google Scholar] [CrossRef]
  13. Jos, S.; Suja, N.R. Chiral Schiff base ligands of salicylaldehyde: A versatile tool for medical applications and organic synthesis-A review. Inorg. Chim. Acta 2023, 547, 121323. [Google Scholar] [CrossRef]
  14. Pradeep, C.P.; Das, S.K. Coordination and supramolecular aspects of the metal complexes of chiral N-salicyl-β-amino alcohol Schiff base ligands: Towards understanding the roles of weak interactions in their catalytic reactions. Coord. Chem. Rev. 2013, 257, 1699–1715. [Google Scholar] [CrossRef]
  15. Hsieh, S.-H.; Kuo, Y.-P.; Gau, H.-M. Synthesis, characterization, and structures of oxovanadium(V) complexes of Schiff bases of β-amino alcohols as tunable catalysts for the asymmetric oxidation of organic sulfides and asymmetric alkynylation of aldehydes. Dalton Trans. 2007, 97–106. [Google Scholar] [CrossRef] [PubMed]
  16. Zeng, Q.; Wang, H.; Weng, W.; Lin, W.; Gao, Y.; Huang, X.; Zhao, Y. Substituent effects and mechanism elucidation of enantioselective sulfoxidation catalyzed by vanadium Schiff base complexes. New J. Chem. 2005, 29, 1125–1127. [Google Scholar] [CrossRef]
  17. Wojaczyńska, E.; Wojaczyński, J. Enantioselective synthesis of sulfoxides: 2000-2009. Chem. Rev. 2010, 110, 4303–4356. [Google Scholar] [CrossRef]
  18. Hartung, J. Stereoselective syntheses of functionalized cyclic ethers via (Schiff-base) vanadium(V)-catalyzed oxidations. Pure Appl. Chem. 2005, 77, 1559–1574. [Google Scholar] [CrossRef]
  19. Hartung, J.; Drees, S.; Greb, M.; Schmidt, P.; Svoboda, I.; Fuess, H.; Murso, A.; Stalke, D. (Schiff-base)vanadium(V) complex-catalyzed oxidations of substituted bis(homoallylic) alcohols—Stereoselective synthesis of functionalized tetrahydrofurans. Eur. J. Org. Chem. 2003, 2388–2408. [Google Scholar] [CrossRef]
  20. Cordelle, C.; Agustin, D.; Daran, J.-C.; Poli, R. Oxo-bridged bis oxo-vanadium(V) complexes with tridentate Schiff base ligands (VOL)2O (L = SAE, SAMP, SAP): Synthesis, structure and epoxidation catalysis under solvent-free. Inorg. Chim. Acta 2010, 364, 144–149. [Google Scholar] [CrossRef]
  21. Romanowski, G.; Kira, J.; Wera, M. Vanadium(V) complexes with chiral tridentate Schiff base ligands derived from 1S,2R(+)-2-amino-1,2-diphenylethanol and with acetohydroxamate co-ligand: Synthesis, characterization and catalytic activity in the oxidation of prochiral sulfides and olefins. J. Mol. Catal. A Chem. 2014, 381, 148–160. [Google Scholar] [CrossRef]
  22. Moschona, F.; Savvopoulou, I.; Tsitopoulou, M.; Tataraki, D.; Rassias, G. Epoxide syntheses and ring-opening reactions in drug development. Catalysts 2020, 10, 1117. [Google Scholar] [CrossRef]
  23. Meninno, S.; Lattanzi, A. Epoxides: Small Rings to Play with under Asymmetric Organocatalysis. ACS Org. Inorg. Au 2022, 2, 289–305. [Google Scholar] [CrossRef]
  24. Costa Pessoa, J.; Etcheverry, S.; Gambino, D. Vanadium compounds in medicine. Coord. Chem. Rev. 2015, 301–302, 24–48. [Google Scholar] [CrossRef]
  25. Tsave, O.; Petanidis, S.; Kioseoglou, E.; Yavropoulou, M.P.; Yovos, J.G.; Anestakis, D.; Tsepa, A.; Salifoglou, A. Role of vanadium in cellular and molecular immunology: Association with immune-related inflammation and pharmacotoxicology mechanisms. Oxid. Med. Cell. Longev. 2016, 2016, 4013639. [Google Scholar] [CrossRef] [PubMed]
  26. Correia, I.; Adão, P.; Roy, S.; Wahba, M.; Matos, C.; Maurya, M.R.; Marques, F.; Pavan, F.R.; Leite, C.Q.F.; Avecilla, F.; et al. Hydroxyquinoline derived vanadium(IV and V) and copper(II) complexes as potential anti-tuberculosis and anti-tumor agents. J. Inorg. Biochem. 2014, 141, 83–93. [Google Scholar] [CrossRef]
  27. Semiz, S. Vanadium as potential therapeutic agent for COVID-19: A focus on its antiviral, antiinflamatory, and antihyperglycemic effects. J. Trace Elem. Med. Biol. 2022, 69, 126887. [Google Scholar] [CrossRef] [PubMed]
  28. Nareetsile, F.; Matshwele, J.T.; Ndlovu, S.; Ngaski, M. Transition metal complexes with HIV/AIDS inhibitory properties. Chem. Rev. Lett. 2020, 3, 140–160. [Google Scholar]
  29. Kioseoglou, E.; Petanidis, S.; Gabriel, C.; Salifoglou, A. The chemistry and biology of vanadium compounds in cancer therapeutics. Coord. Chem. Rev. 2015, 301–302, 87–105. [Google Scholar] [CrossRef]
  30. Gambino, D. Potentiality of vanadium compounds as anti-parasitic agents. Coord. Chem. Rev. 2011, 255, 2193–2203. [Google Scholar] [CrossRef]
  31. Kim, A.T.; Kim, D.O. Anti-inflammatory effects of vanadium-binding protein from Halocynthia roretzi in LPS-stimulated RAW264. 7 macrophages through NF-κB and MAPK pathways. Int. J. Biol. Macromol. 2019, 133, 732–738. [Google Scholar] [CrossRef]
  32. Treviño, S.; Díaz, A.; Sánchez-Lara, E.; Sanchez-Gaytan, B.L.; Perez-Aguilar, J.M.; González-Vergara, E. Vanadium in biological action: Chemical, pharmacological aspects, and metabolic implications in Diabetes Mellitus. Biol. Trace Elem. Res. 2019, 188, 68–98. [Google Scholar] [CrossRef]
  33. Jakusch, T.; Kiss, T. In vitro study of the antidiabetic behavior of vanadium compounds. Coord. Chem. Rev. 2017, 351, 118–126. [Google Scholar] [CrossRef]
  34. Wyrzykowski, D.; Inkielewicz-Stępniak, I.; Pranczk, J.; Żamojć, K.; Zięba, P.; Tesmar, A.; Jacewicz, D.; Ossowski, T.; Chmurzyński, L. Physicochemical properties of ternary oxovanadium(IV) complexes with oxydiacetate and 1,10-phenanthroline or 2,2′-bipyridine. Cytoprotective activity in hippocampal neuronal HT22 cells. Biometals 2015, 28, 307–320. [Google Scholar] [CrossRef] [PubMed]
  35. Tesmar, A.; Inkielewicz-Stępniak, I.; Sikorski, A.; Wyrzykowski, D.; Jacewicz, D.; Zięba, P.; Pranczk, J.; Ossowski, T.; Chmurzyński, L. Structure, physicochemical and biological properties of newcomplex salt of aqua-(nitrilotriacetato-N,O,O′,O″)-oxidovanadium(IV) anion with 1,10-phenanthrolinium cation. J. Inorg. Biochem. 2015, 152, 53–61. [Google Scholar] [CrossRef] [PubMed]
  36. Pfeiffer, A.; Jaeckel, M.; Lewerenz, J.; Noack, R.; Pouya, A.; Schacht, T.; Hoffmann, C.; Winter, J.; Schweiger, S.; Schäfer, M.K.E.; et al. Mitochondrial function and energy metabolism in neuronal HT22 cells resistant to oxidative stress. Br. J. Pharmacol. 2014, 171, 2147–2158. [Google Scholar] [CrossRef]
  37. Romanowski, G.; Lis, T. Chiral oxidovanadium(V) complexes with tridentate Schiff bases derived from S(+)-2-amino-1-propanol: Synthesis, structure, characterization and catalytic activity. Inorg. Chim. Acta 2013, 394, 627–634. [Google Scholar] [CrossRef]
  38. Romanowski, G.; Kira, J. Oxidovanadium(V) complexes with chiral tridentate Schiff bases derived from R(−)-phenylglycinol: Synthesis, spectroscopic characterization and catalytic activity in the oxidation of sulfides and styrene. Polyhedron 2013, 53, 172–178. [Google Scholar] [CrossRef]
  39. Bhattacharya, S.; Gosh, T. Synthesis, spectral and electrochemical studies of alkoxo-bonded mixed-ligand oxovanadium(IV) and oxovanadium(V) complexes incorporating tridentate ONO donor azophenolalcoholate/aldiminealcoholates. Transit. Met. Chem. 2002, 27, 89–94. [Google Scholar] [CrossRef]
  40. Romanowski, G. Synthesis, characterization and catalytic activity in the oxidation of sulfides and styrene of vanadium(V) complexes with tridentate Schiff base ligands. J. Mol. Catal. A Chem. 2013, 368–369, 137–144. [Google Scholar] [CrossRef]
  41. Rezaeifard, A.; Jafarpour, M.; Raissi, H.; Alipour, M.; Stoeckli-Evans, H. Economical oxygenation of olefins and sulfides catalyzed by new molybdenum(VI) tridentate Schiff base complexes: Synthesis and crystal structure. Z. Anorg Allg. Chem. 2012, 638, 1023–1030. [Google Scholar] [CrossRef]
  42. Romanowski, G.; Kira, J.; Wera, M. Five- and six-coordinate vanadium(V) complexes with tridentate Schiff base ligands derived from S(+)-isoleucinol: Synthesis, characterization and catalytic activity in the oxidation of sulfides and olefins. Polyhedron 2014, 67, 529–539. [Google Scholar] [CrossRef]
  43. Hulea, V.; Dumitriu, E. Styrene oxidation with H2O2 over Ti-containing molecular sieves with MFI, BEA and MCM-41 topologies. Appl. Catal. A Gen. 2004, 277, 99–106. [Google Scholar] [CrossRef]
  44. Maurya, M.R.; Kumar, A.; Ebel, M.; Rehder, D. Synthesis, characterization, reactivity, and catalytic potential of model vanadium(IV,V) complexes with benzimidazole-derived ONN donor ligands. Inorg. Chem. 2006, 45, 5924–5937. [Google Scholar] [CrossRef] [PubMed]
  45. Maurya, M.R.; Bisht, M.; Chaudhary, N.; Avecilla, F.; Kumar, U.; Hsu, H.-F. Synthesis, structural characterization, encapsulation in zeolite Y and catalytic activity of an oxidovanadium(V) complex with a tribasic pentadentate ligand. Polyhedron 2013, 54, 180–188. [Google Scholar] [CrossRef]
  46. McCaffrey, V.P.; Conover, O.Q.; Bernard, M.A.; Yarranton, J.T.; Lessnau, N.R.; Hempfling, J.P. Substituent effects in dioxovanadium(V) schiff-base complexes: Tuning the outcomes of oxidation reactions. Polyhedron 2021, 205, 115268. [Google Scholar] [CrossRef]
  47. Monfared, H.H.; Bikas, R.; Mayer, P. Homogeneous green catalysts for olefin oxidation by mono oxovanadium(V) complexes of hydrazone Schiff base ligands. Inorg. Chim. Acta 2010, 363, 2574–2583. [Google Scholar] [CrossRef]
  48. Monfared, H.H.; Kheirabadi, S.; Lalami, N.A.; Mayer, P. Dioxo- and oxovanadium(V) complexes of biomimetic hydrazone ONO and NNS donor ligands: Synthesis, crystal structure and catalytic reactivity. Polyhedron 2011, 30, 1375–1384. [Google Scholar] [CrossRef]
  49. Koola, J.D.; Kochi, J.K. Cobalt-catalyzed epoxidation of olefins. Dual pathways for oxygen atom transfer. J. Org. Chem. 1987, 52, 4545–4553. [Google Scholar] [CrossRef]
  50. Karman, M.; Romanowski, G. Chiral cis-dioxidomolybdenum(VI) complexes with Schiff bases possessing two alkoxide groups: Synthesis, structure, spectroscopic studies and their catalytic activity in sulfoxidation and epoxidation. Polyhedron 2020, 187, 114653. [Google Scholar] [CrossRef]
  51. Sayre, L.M.; Perry, G.; Smith, M.A. Oxidative stress and neurotoxicity. Chem. Res. Toxicol. 2008, 21, 172–188. [Google Scholar] [CrossRef] [PubMed]
  52. Tian, X.-L.; Jiang, S.-Y.; Zhang, X.-L.; Yang, J.; Cui, J.-H.; Liu, X.-L.; Gong, K.-R.; Yan, S.-C.; Zhang, C.-Y.; Shao, G. Potassium bisperoxo (1,10-phenanthroline) oxovanadate suppresses proliferation of hippocampal neuronal cell lines by increasing DNA methyltransferases. Neural Regen Res. 2019, 14, 826–833. [Google Scholar]
  53. Kothandan, S.; Sheela, A. DNA interaction and cytotoxic studies on mono/di-oxo and peroxo- vanadium (V) complexes—A review. Mini-Rev. Med. Chem. 2021, 21, 1909–1924. [Google Scholar] [CrossRef]
  54. Crans, D.C.; Yang, L.; Haase, A.; Yang, X. Metallo-Drugs: Development and Action of Anticancer Agents; Sigel, A., Sigel, H., Freisinger, E., Sigel, R.K.O., Eds.; Chapter 9; De Gruyter: Berlin, Germany, 2018; pp. 251–274. [Google Scholar]
Figure 1. Structural formulae of oxidovanadium(V) complexes.
Figure 1. Structural formulae of oxidovanadium(V) complexes.
Ijms 25 05010 g001
Figure 2. Olefinic and monoterpene substrates chosen for catalytic studies.
Figure 2. Olefinic and monoterpene substrates chosen for catalytic studies.
Ijms 25 05010 g002
Figure 3. Possible products of catalytic oxidation of styrene.
Figure 3. Possible products of catalytic oxidation of styrene.
Ijms 25 05010 g003
Figure 4. Allylic oxidation and epoxidation products of cyclohexene.
Figure 4. Allylic oxidation and epoxidation products of cyclohexene.
Ijms 25 05010 g004
Figure 5. MTT assay after 24 h exposure to test compounds showing the viability of HT-22 cell line. Data are expressed as mean ± SD values from three experiments. * p < 0.05, *** p < 0.001, as compared to control (untreated) cells.
Figure 5. MTT assay after 24 h exposure to test compounds showing the viability of HT-22 cell line. Data are expressed as mean ± SD values from three experiments. * p < 0.05, *** p < 0.001, as compared to control (untreated) cells.
Ijms 25 05010 g005
Figure 6. MTT assay after 24 h exposure to test compounds and 500 µM H2O2 showing the viability of HT-22 cell line. Data are expressed as mean ±SD values from three experiments. * p < 0.05, ** p < 0.01, *** p < 0.001, as compared to control (untreated) cells and to 500 µM H2O2.
Figure 6. MTT assay after 24 h exposure to test compounds and 500 µM H2O2 showing the viability of HT-22 cell line. Data are expressed as mean ±SD values from three experiments. * p < 0.05, ** p < 0.01, *** p < 0.001, as compared to control (untreated) cells and to 500 µM H2O2.
Ijms 25 05010 g006
Figure 7. HT-22 cell morphology: (a) control cells and after exposure to (b) 500 µM of 30% H2O2, (c) 10 µM [(+)MoO2(HL3)] and 500 µM of 30% H2O2, and (d) 10 µM of [(+)VOL4] and 500 µM of 30% H2O2.
Figure 7. HT-22 cell morphology: (a) control cells and after exposure to (b) 500 µM of 30% H2O2, (c) 10 µM [(+)MoO2(HL3)] and 500 µM of 30% H2O2, and (d) 10 µM of [(+)VOL4] and 500 µM of 30% H2O2.
Ijms 25 05010 g007
Table 1. Styrene oxidation with oxidovanadium(V) Schiff base complexes as catalysts.
Table 1. Styrene oxidation with oxidovanadium(V) Schiff base complexes as catalysts.
EntryCatalystConversion (%)OxidantStyrene Oxide (%)Benzaldehyde (%)
1[(+)VOL1]25H2O25842
2[(+)VOL1]95TBHP4654
3[(+)VOL2]91TBHP3763
4[(+)VOL3]89TBHP3565
5[(+)VOL4]90TBHP4357
6[(+)VOL5]86TBHP3169
7[(−)VOL1]27H2O25644
8[(−)VOL1]96TBHP4852
9[(−)VOL2]92TBHP5248
10[(−)VOL3]93TBHP3862
11[(−)VOL4]95TBHP4258
12[(−)VOL5]92TBHP4951
Table 2. Cyclohexene oxidation with oxidovanadium(V) Schiff base complexes as catalysts.
Table 2. Cyclohexene oxidation with oxidovanadium(V) Schiff base complexes as catalysts.
EntryCatalystConversion (%)OxidantEpoxide (%)Alcohol (%)Ketone (%)
1[(+)VOL1]42H2O2442927
2[(+)VOL1]90TBHP223742
3[(+)VOL2]87TBHP263044
4[(+)VOL3]90TBHP213742
5[(+)VOL4]87TBHP224038
6[(+)VOL5]84TBHP332740
7[(−)VOL1]44H2O2413227
8[(−)VOL1]82TBHP362143
9[(−)VOL2]79TBHP332939
10[(−)VOL3]92TBHP302248
11[(−)VOL4]87TBHP322444
12[(−)VOL5]84TBHP282646
Table 3. The oxidation of S(−)-limonene with oxidovanadium(V) Schiff base complexes as catalysts.
Table 3. The oxidation of S(−)-limonene with oxidovanadium(V) Schiff base complexes as catalysts.
EntryCatalystConversion (%)OxidantEpoxide (%)Diepoxide (%)
1[(+)VOL1]34H2O27525
2[(+)VOL1]87TBHP7030
3[(+)VOL2]93TBHP7228
4[(+)VOL3]88TBHP7030
5[(+)VOL4]92TBHP8416
6[(+)VOL5]75TBHP7327
7[(−)VOL1]37H2O27129
8[(−)VOL1]100TBHP5743
9[(−)VOL2]99TBHP5941
10[(−)VOL3]99TBHP5644
11[(−)VOL4]87TBHP6139
12[(−)VOL5]92TBHP7525
Table 4. The oxidation of (−)-α-pinene with oxidovanadium(V) Schiff base complexes as catalysts.
Table 4. The oxidation of (−)-α-pinene with oxidovanadium(V) Schiff base complexes as catalysts.
EntryCatalystConversion (%)Oxidant(−)-α-pinene Oxide (%)
1[(+)VOL1]26H2O279
2[(+)VOL1]77TBHP79
3[(+)VOL2]62TBHP77
4[(+)VOL3]83TBHP72
5[(+)VOL4]79TBHP73
6[(+)VOL5]76TBHP70
7[(−)VOL1]32H2O286
8[(−)VOL1]69TBHP83
9[(−)VOL2]62TBHP79
10[(−)VOL3]87TBHP76
11[(−)VOL4]70TBHP73
12[(−)VOL5]62TBHP68
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Romanowski, G.; Budka, J.; Inkielewicz-Stepniak, I. Oxidovanadium(V) Schiff Base Complexes Derived from Chiral 3-amino-1,2-propanediol Enantiomers: Synthesis, Spectroscopic Studies, Catalytic and Biological Activity. Int. J. Mol. Sci. 2024, 25, 5010. https://doi.org/10.3390/ijms25095010

AMA Style

Romanowski G, Budka J, Inkielewicz-Stepniak I. Oxidovanadium(V) Schiff Base Complexes Derived from Chiral 3-amino-1,2-propanediol Enantiomers: Synthesis, Spectroscopic Studies, Catalytic and Biological Activity. International Journal of Molecular Sciences. 2024; 25(9):5010. https://doi.org/10.3390/ijms25095010

Chicago/Turabian Style

Romanowski, Grzegorz, Justyna Budka, and Iwona Inkielewicz-Stepniak. 2024. "Oxidovanadium(V) Schiff Base Complexes Derived from Chiral 3-amino-1,2-propanediol Enantiomers: Synthesis, Spectroscopic Studies, Catalytic and Biological Activity" International Journal of Molecular Sciences 25, no. 9: 5010. https://doi.org/10.3390/ijms25095010

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