Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2022 Nov 10;61(46):18434–18449. doi: 10.1021/acs.inorgchem.2c02526

A Combined Experimental and Theoretical Investigation of Oxidation Catalysis by cis-[VIV(O)(Cl/F)(N4)]+ Species Mimicking the Active Center of Metal-Enzymes

Michael G Papanikolaou †,, Anastasia V Simaioforidou , Chryssoula Drouza §, Athanassios C Tsipis †,*, Haralampos N Miras ⊥,*, Anastasios D Keramidas ∥,*, Maria Louloudi †,*, Themistoklis A Kabanos †,*
PMCID: PMC9682486  PMID: 36357045

Abstract

graphic file with name ic2c02526_0022.jpg

Reaction of VIVOCl2 with the nonplanar tetradentate N4 bis-quinoline ligands yielded four oxidovanadium(IV) compounds of the general formula cis-[VIV(O)(Cl)(N4)]Cl. Sequential treatment of the two nonmethylated N4 oxidovanadium(IV) compounds with KF and NaClO4 resulted in the isolation of the species with the general formula cis-[VIV(O)(F)(N4)]ClO4. In marked contrast, the methylated N4 oxidovanadium(IV) derivatives are inert toward KF reaction due to steric hindrance, as evidenced by EPR and theoretical calculations. The oxidovanadium(IV) compounds were characterized by single-crystal X-ray structure analysis, cw EPR spectroscopy, and magnetic susceptibility. The crystallographic characterization showed that the vanadium compounds have a highly distorted octahedral coordination environment and the d(VIV–F) = 1.834(1) Å is the shortest to be reported for (oxido)(fluorido)vanadium(IV) compounds. The experimental EPR parameters of the VIVO2+ species deviate from the ones calculated by the empirical additivity relationship and can be attributed to the axial donor atom trans to the oxido group and the distorted VIV coordination environment. The vanadium compounds act as catalysts toward alkane oxidation by aqueous H2O2 with moderate ΤΟΝ up to 293 and product yields of up to 29% (based on alkane); the vanadium(IV) is oxidized to vanadium(V), and the ligands remain bound to the vanadium atom during the catalysis, as determined by 51V and 1H NMR spectroscopies. The cw X-band EPR studies proved that the mechanism of the catalytic reaction is through hydroxyl radicals. The chloride substitution reaction in the cis-[VIV(O)(Cl)(N4)]+ species by fluoride and the mechanism of the alkane oxidation were studied by DFT calculations.

Short abstract

Highly distorted N4−oxidovanadium(IV), cis-[VIV(=O)(Cl/F)(N4)]+, compounds mimicking the distorted coordination environment of the metal-ion in metal-enzymes’ active center catalyze oxidation of alkanes. The structural distortion results in significant deviation of the cw EPR parameters from the expected ones and equilibrium between six- and five- coordinate species in solution. DFT studies show that distortion, steric hindrance, and the coordinated halogen define the catalytic mechanism of oxidation of alkanes by cis-[VIV(=O)(Cl/F)(N4)]+ compounds.

Introduction

In recent years, the coordination chemistry of vanadium has drawn a lot of interest, mainly due to its biological, medicinal and catalytic applications.111 Vanadium exhibits a wide variety of oxidation states (−III to + V), with the oxidation states of +III to +V mainly found in molecular systems of biological relevance. Enzymes, such as the vanadium-dependent haloperoxidases found in algae, are able to utilize vanadium’s wide range of oxidation states in order to oxidize halides in nature.12 Moreover, vanadium plays a key role in the vanadium nitrogenase enzyme, which is a vanadium analogue of the iron–molybdenum enzyme that reduces dinitrogen to ammonia.1315 In addition, vanadium has a significant effect on cell growth, signaling processes, antitumor activity, and insulin-mimetic properties.1626

The synthesis of metal compounds, which are metal enzymes’ active site analogues, has played an important role in understanding the mechanisms of enzyme activity and in the development of small molecules with activity similar to relevant enzymes.27 The particular function of the metal-enzymes requires specific oxidation states, ligands, and coordination geometries for the metal-ion in the active site.27 The coordination geometries around the metal-ions in the enzymes’ active site, enforced by the rigidity of the protein backbones, are irregular. These enforced geometries define the activity of the enzyme.28,29 Small changes of these structural features are crucial for the specificity of the enzymes. In contrast, in the small metal compounds, there are few or no constraints dictating the geometry. Therefore, the arrangement of the ligands around the metal ions in these compounds relies on the preference of the metal ion.

Vanadium’s low-molecular-weight coordination compounds mimic the activity of enzymes, such as haloperoxidases.3033 Vanadium’s wide range of oxidation states and coordination numbers and its Lewis acid character are the key characteristics that enable the use of vanadium compounds in various catalytic reactions, mimicking haloperoxidases, such as alcohol oxidation, sulfoxidation, epoxidation, and alkane oxidation reactions.3441 In particular, the oxidation of alkanes has a high industrial significance, since it enables the functionalization of inert alkanes to more valuable and reactive organic materials such as alcohols and ketones in the presence of a suitable oxidant like H2O2 or O2, under mild conditions.4245 Some of the most commonly used ligands are nitrogen-based tetradentate pyridine or quinoline ligands, which have the ability to strongly bind and stabilize vanadium ions in the +IV or +V oxidation state. Moreover, these ligands are highly resistant to oxidation and decomposition under the catalytic conditions, and their compounds with iron(II) are some of the most efficient catalysts for alkane oxidation.46 However, in order to mimic metal-enzymes outstanding oxidative activity and synthesize efficient low-molecular-weight catalysts, it is important to understand the effect of the distortion of the coordination environments of the metal ions in the active site of the enzyme and its contribution to the catalytic action.

Herein, we describe the synthesis, physicochemical, and structural characterization and the catalytic properties in alkane oxidation reactions of various oxidovanadium(IV) compounds with the nonplanar N4 quinoline-based/amine ligands and their dimethylated analogues (Scheme 1).

Scheme 1. Drawing of the Ligands Used in This Study.

Scheme 1

The tetradentate nonplanar N4 ligands (Scheme 1) were chosen because their ligation to VIVO2+ induces a severely distorted octahedral geometry (Scheme 2), since our aim was to study the effect of structural distortions on the catalytic properties and the substitution reactions of these cis-[VIV(O)(X)(N4)]+ species. The geometric features of these cis-[VIV(O)(X)(N4)]+ (X = F, Cl) species mimic the irregularities of the enzymes’ active site, and can for the first time provide valuable information regarding their impact on the oxidative catalytic activity, the mechanism of their action (DFT calculations), and their spectroscopic properties (cw X-band EPR). Moreover, the effect of the halogen, in the cis-[VIV(O)(X)(N4)]+ (X = F, Cl) species, on their oxidative catalytic activity and the catalytic mechanism (DFT calculations) was also investigated.

Scheme 2. Highly Distorted Equatorial Plane of the Octahedral Compound cis-[VIV(O)(F)(N4)]+.

Scheme 2

Experimental Section

Synthesis of the Ligands and the Oxidovanadium(IV) Compounds

N,N′-Bis(8-quinolyl)cyclohexane-1,2-diamine, (H2bqch)

trans-Diaminocyclohexane (2.40 mL, 2.28 g, 20 mmol) and sodium metabisulfite (7.60 g, 40 mmol) were added to a suspension of 8-hydroxyquinoline (5.80 g, 40 mmol) in 200 mL of water. The mixture was heated at reflux for 10 days. Subsequently, the solution was cooled to room temperature and was made strongly alkaline (pH 13) with the addition of solid KOH. The mixture was extracted with dichloromethane (3 × 40 mL), the organic layers were combined, dried with MgSO4, and the solvent was removed under vacuum. The obtained residue was triturated with warm (40 °C) ethyl alcohol (10 mL), and the formed pale-yellow precipitate was filtered and dried under high vacuum to get 2.95 g of the desired organic molecule. Yield, 40%, based on trans-diaminocyclohexane. Anal. Calcd (%) for C24H24N4 (Mr = 368.24 g/mol): C, 78.22; H, 6.57; N, 15.21. Found (%): C, 78.20; H, 6.58; N, 15.28. Rf = 0.84 (CH3COOC2H5). Mp = 174–175 °C.

N,N′-Dimethyl-N,N′-bis(8-quinolyl)cyclohexane-1,2-diamine, (dbqch)

The dimethylated organic molecule dbqch was prepared according to Britovsek and co-workers in 61% yield.46 The purity of dbqch was confirmed with positive HR-ESI-MS, and 1H, 13C NMR. Anal. Calcd (%) for C26H28N4 (Mr = 396.28 g/mol): C, 78.74; H, 7.12; N, 14.14. Found (%): C, 78.53; H, 7.08; N, 14.28.

N,N′-Bis(8-quinolyl)ethane-1,2-diamine (H2bqen)

This organic molecule was synthesized in the same way as H2bqch, except that ethylene-1,2-diamine (1.40 mL, 1.20 g, 20 mmol) was used instead of trans-diaminocyclohexane. The final product was a yellow solid (3.77 g, 60% based on ethylene-1,2-diamine). Anal. Calcd (%) for C20H18N4 (Mr = 314.19 g/mol): C, 76.39; H, 5.77; N, 17.83. Found (%): C, 76.35; H, 5.71; N, 17.84. Rf = 0.81 (CH3COOC2H5). Mp = 160–161 °C.

N,N′-Dimethyl-N,N′-bis(8-quinolyl)ethane-1,2-diamine (dbqen)

The dimethylated organic molecule H2bqch was prepared according to Britovsek and co-workers in 69% yield.46 The purity of dbqen was confirmed with positive HR-ESI-MS, and 1H, 13C NMR. Anal. Calcd (%) for C22H22N4 (Mr = 342.22 g/mol): C, 77.15; H, 6.48; N, 16.37. Found (%): C, 76.215; H, 6.44; N, 16.18.

Cis-chlorido[N,N′-Bis(8-quinolyl)cyclohexane-1,2-diamine-N,N,N,N]oxidovanadium(IV) Chloride, cis-[VIV(O)(Cl)(H2bqch)]Cl·H2O (1·H2O)

To the stirred aqueous solution (5 mL) of VIVOSO4·5H2O (172 mg, 0.68 mmol), BaCl2·2H2O (183 mg, 0.75 mmol) was added in one portion, and a white precipitate (BaSO4) was immediately formed. The mixture was stirred for 1 h and was filtered. The filtrate was evaporated to dryness under high vacuum, and the residue was dissolved in CH3CN (6 mL). A tetrahydrofuran (20 mL) solution containing the ligand H2bqch (250 mg, 0.68 mmol) was added dropwise to the stirred oxidovanadium(IV) solution. Upon addition of the ligand, the blue color of the solution changed to brown, and a light brown precipitate was formed. The solution was stirred for three additional hours, and then it was filtered and washed with diethyl ether (2 × 10 mL) and dried in vacuum to get 0.275 g of a light brown solid. Yield, 81% (based on H2bqch). Anal. Calcd (%) for C24H26Cl2N4O2V (Mr= 524.10 g/mol): C, 54.96; H, 5.00; Cl, 13.53; N, 10.69; V, 9.72. Found (%): C, 54.91; H, 4.87; Cl, 13.47; N, 10.75; V, 9.67. (High resolution electrospray ionization mass spectrometry [HR-ESI(+)-MS]: calcd for cis-[VIV(O)(Cl)(H2bqch)]Cl·H2O (C24H26Cl2N4O2V) {[M-(Cl+H2O)]+} m/z 470.1073, found 470.1075 . μeff = 1.73 μB

Cis-chlorido[N,N′-dimethyl-N,N′-bis(8-quinolyl)cyclohexane-1,2-diamine-N,N,N,N]oxidovanadium(IV) Chloride, cis-[VIV(O)(Cl)(dbqch)]Cl (2)

Compound 2 was synthesized using the same method reported for H2O with VIVOSO4·5H2O (96 mg, 0.38 mmol, 1 equiv) and dbqch (150 mg, 0.38 mmol, 1 equiv). Yield: 120 mg (59%) of a green solid. Anal. Calcd (%) for C26H28Cl2N4OV (Mr = 534.12 g/mol): C, 58.42; H, 5.28; Cl, 13.28; N, 10.49; V, 9.53. Found (%): C, 58.47; H, 4.97; Cl, 13.19; N, 10.68; V, 9.28. (High resolution electrospray ionization mass spectrometry [HR-ESI(+)-MS]: calcd for C26H28Cl2N4OV {[M-(Cl)]+} m/z 498.1386, found 498.1368. μeff = 1.71 μB

Crystals of cis-[VIV(O)(Cl)(dbqch)]BF4·2CH3CN (2’) suitable for X-ray structure analysis were prepared as follows: Compound 2 (50 mg, 0.09 mmol) was dissolved in water (10 mL) under magnetic stirring, and NH4BF4 (9.8 mg, 0.09 mmol) was added to it. Upon addition of NH4BF4, a light green precipitate was formed which was filtered and dried. Dissolution of the green solid in CH3CN and layering of diethyl ether to it resulted in the formation of crystals of cis-[VIV(O)(Cl)(dbqch)]BF4·2CH3CN.

Crystals of cis-[VIV(O)(Cl)(dbqch)]ClO4 (2’’) suitable for X-ray structure analysis were prepared using the same method reported for 2’ except that NaClO4 was used instead of NH4BF4.

Caution!Perchlorates are powerful oxidizers, they are potentially hazardous, especially in contact with reducing material, and they may explode when exposed to shock or heat.(47)

Cis-chlorido[N,N′-Bis(8-quinolyl)ethane-1,2-diamine-N,N,N,N]oxidovanadium(IV) Chloride, cis-[VIV(O)(Cl)(H2bqen)]Cl·2H2O (3·2H2O)

Compound 2H2O was prepared using the same method reported for H2O with VIVOSO4·5H2O (240 mg, 0.95 mmol, 1 equiv), BaCl2·2H2O (256 mg, 1.05 mmol) and H2bqen (300 mg, 0.95 mmol, 1 equiv). Yield, 0.368 g (85%, based on H2bqen) of a brown solid. Anal. Calcd (%) for C20H22Cl2N4O3V (Mr = 488.07 g/mol): C, 49.18; H, 4.54; Cl, 14.53; N, 11.48; V, 10.44. Found (%): C, 49.23; H, 4.56; Cl, 14.49; N, 11.24; V, 10.31 [HR-ESI(+)-MS]: calcd for cis-[VIV(O)(Cl)(H2bqen)]Cl·2H2O (C20H22Cl2N4O3V) {[M-(Cl+H2O)]+} m/z 416.0603, found 416.0600. μeff = 1.70 μB

Cis-chlorido[N,N′-dimethyl-N,N′-Bis(8-quinolyl)ethane-1,2-diamine-N,N,N,N]oxidovanadium(IV) Chloride, cis-[VIV(O)(Cl)(dbqen)]Cl·3H2O (4·3H2O)

Compound 3H2O was prepared using the same method reported for H2O with VIVOSO4·5H2O (222 mg, 0.88 mmol, 1 equiv), BaCl2·2H2O (236 mg, 0.97 mmol) and dbqen (300 mg, 0.88 mmol, 1 equiv). Yield: 383 mg (82%, based on dbqen) of a green solid. Anal. Calcd (%) for C22H28Cl2N4O4V (Mr = 534.33 g/mol): C, 49.41; H, 5.12; Cl, 13.26; N, 10.48; V, 9.53 Found (%): C, 49.61; H, 5.12; Cl, 13.23; N, 10.39, V, 9.22 (High resolution electrospray ionization mass spectrometry [HR-ESI(+)-MS]: calcd for cis-[VIV(O)(Cl)(dbqen)]Cl·3H2O (C22H28Cl2N4O4V) {[M-(Cl+3H2O)]+} m/z 444.0916, found 444.0904. μeff = 1.74 μB

Crystals of cis-[VIV(O)(Cl)(dbqen)]ClO4·2CH3CN (4’) suitable for X-ray structure analysis were prepared as follows: Compound 3H2O (50 mg, 0.09 mmol) was dissolved in water (10 mL) under magnetic stirring, and NaClO4 (11 mg, 0.09 mmol) was added to it. Upon addition of NaClO4, a light green precipitate was formed, which was filtered and dried. Dissolution of the green solid in CH3CN and layering of diethyl ether to it resulted in the formation of crystals of cis-[VIV(O)(Cl)(dbqen)]ClO4·2CH3CN.

Cis-fluorido[N,N′-Bis(8-quinolyl)cyclohexane-1,2-diamine-N,N,N,N]oxidovanadium(IV) Perchlorate, cis-[VIV(O)(F)(H2bqch)]ClO4 (5)

To a stirred solution of H2O (100 mg, 0.19 mmol) in water (20 mL) was added in one portion solid KF (12 mg, 0.21 mmol). Upon addition of KF, the light brown color of the solution changed to orange. The solution was stirred for an additional hour. Then solid NaClO4 (26 mg, 0.21 mmol) was added to it in one portion, and an orange precipitate was formed. The mixture was stirred for 3 h and filtered, washed with cold water (2 × 5 mL), and dried in vacuum to get 78 mg of the orange solid. Yield: 74% (based on H2O). Anal. Calcd (%) for C24H24ClFN4O5V (Mr = 553.63 g/mol): C, 52.02; H, 4.37; F, 3.43; N, 10.12; V, 9.20. Found (%): C, 52.01; H, 4.37; F, 3.40; N, 10.09; V, 9.19. (High resolution electrospray ionization mass spectrometry [HR-ESI(+-MS]: calcd for cis-[VIV(O)(F)(H2bqch)]ClO4· (C24H24ClFN4O5V) {[M-(ClO4)]+} m/z 454.1368, 454.1339 found; {[M-(F + H + ClO4)]+} m/z 434.1306, found 434.1285. μeff = 1.72 μB

Crystals of CH3OH (5’) suitable for X-ray structure analysis were obtained by layering diethyl ether into a concentrated methyl alcohol solution of 5.

Cis-fluorido[N,N′-Bis(8-quinolyl)ethane-1,2-diamine-N,N,N,N]oxidovanadium(IV) Perchlorate, cis-[VIV(O)(F)(H2bqen)]ClO4 (6)

Compound 6 was prepared using the same method reported for the synthesis of 5 with 2H2O (93 mg, 0.19 mmol), KF (12 mg, 0.21 mmol), and NaClO4 (26 mg, 0.22 mmol) to get 78 mg of the orange solid. Yield: 74% (based on 3·2H2O). Anal. Calcd (%) for C20H18ClFN4O5V (Mr = 499.58 g/mol): C, 48.02; H, 3.63; F, 3.80; N, 11.21; V, 10.19 Found (%): C, 47.93; H, 3.60; F, 3.75; N, 10.96; V, 10.24 (High resolution electrospray ionization mass spectrometry [HR-ESI(+)-MS]: calcd for cis-[VIV(O)(F)(H2bqen)]ClO4(C20H18ClFN4O5V) {[M-(H+F+ClO4)]+} m/z 380.0837, found 380.0814. μeff = 1.74 μB

Preparation of the compound cis-[VIV(O)(F)(H2bqen)]BF4 (6’) was performed with the same method as for the synthesis of 6 except that NH4BF4 was used instead of NaClO4. Crystals of 6’suitable for X-ray structure analysis were obtained by dissolving 6’ into methyl alcohol and layering diethyl ether into the concentrated methyl alcohol solution of cis-[VIV(O)(F)(H2bqen)]BF4.

Results and Discussion

Synthesis of the Ligands and Oxidovanadium(IV) Compounds

The synthesis of the ligands dbqch and dbqen is depicted in Scheme 3 and includes three steps: The first step involves the reaction of the diamine (1 equiv) with 8-hydroxyquinoline (2 equiv) to get the secondary amines H2bqch and H2bqen. The secondary amines were prepared by slight modification of the method of Britovsek46 and co-workers to increase their yield by 10–15%. The second step involves the deprotonation of the secondary amines with 2 equiv of n-BuLi. In the third step, the methylation by CH3I (two equivalents) of the deprotonated amines was carried out to afford the dimethylated ligands.

Scheme 3. Synthesis of the Ligands H2bqch and dbqch.

Scheme 3

The synthesis of the cis-[VIV(O)(Cl)(N4)nm/m]+ and cis-[VIV(O)(F)(N4)nm]+ compounds is shown in Schemes 4 and 5. respectively. (The indexes nm and m mean the nonmethylated and methylated ligands respectively.) The cis-[VIV(O)(F)(N4)m]+ derivatives were not synthesized, since the species cis-[VIV(O)(Cl)(N4)m]+ does not react with F due to steric hindrance (see EPR and DFT calculations for details).

Scheme 4. Synthesis of the cis-[VIV(O)(Cl)(N4)nm/m]+ Compounds (14).

Scheme 4

Scheme 5. Synthesis of cis-[VIV(O)(F)(H2bqch)]+ (5) and cis-[VIV(O)(F)(H2bqen)]+ (6) (A). The methylated cis-[VIV(O)(Cl)(N4)m]+ derivatives do not react with F (B).

Scheme 5

Crystal Structures

Crystallographic data and selective bonds and angles, for complexes 2’, 2”, 4’, 5’, and 6’, are summarized in Tables 1, 2, S1 and S2.

Table 1. Crystal Data and Details of the Structure Determination for the VIVO2+ Compounds.

parameter [VOCl(dbqch)]ClO4·2CH3CN [VOCl(dbqen)]ClO4·2CH3CN [VOF(H2bqch)]ClO4·CH3OH [VOF(H2bqen)]BF4
empirical formula C30H34Cl2N6O5.16V C26H28Cl2N6O5V C25H24FClN4O6V C20H18F5BN4OV
formula weight 682.96 626.38 581.87 487.13
temperature 100(2) K 150 K 150 K 150 K
wavelength 0.71073 0.71073 0.71073 0.71073
space group P 21/c P-1 P 21/c Fdd2
a(Å) 17.1586(14) 7.4869(5) 16.2714(10) 31.268(8)
b(Å) 12.2319(5) 12.6366(8) 15.9079(10) 9.975(3)
c(Å) 16.3466(14) 16.0650(11) 10.4324(6) 12.717(3)
α (deg) 90 106.640(3) 90 90
β (deg) 118.116(11) 97.314(4) 95.596(3) 90
γ (deg) 90 99.829(3) 90 90
vol. (Å3) 3026.0(5) 1409.66(16) 2687.5(3) 3966.7(17)
Z 4 2 4 8
ρ (g/cm–3) 1.499 1.476 1.438 1.631
Abscoeff (mm–1) 0.566 0.589 0.522 0.568
R1a 0.0950 0.0371 0.0435 0.0425
wR2b 0.2305 0.1014 0.1231 0.0893
GoF, Sc 1.040 1.045 1.041 1.124
R-Factor (%) 9.50 3.71 4.35 4.25
a

R1 = Σ||Fo| – |Fc||/Σ|Fo|.

b

w R2 = {Σ[w(Fo2Fc2)2]/Σ[w(Fo2)2]}1/2, where w = 1/[σ2(Fo2) + (aP)2 + bP], P = (Fo2 + 2Fc2)/3.

c

GoF = {Σ[w(Fo2Fc2)2]/(np)}1/2, where n = number of reflections and p is the total number of parameters refined.

Table 2. Interatomic Distances (Å) and Angles (deg) Relevant to the VIV Coordination Sphere.

parameter [VOCl(dbqch)]ClO4·2CH3CN [VOCl(dbqen)]ClO4·2CH3CN [VOF(H2bqch)]ClO4·CH3OH [VOF(H2bqen)]BF4
V(1) - Xa 2.178(2) 2.3265(7) 1.834(1) 1.730(1)b
V(1) - N(1) 2.102(4) 2.098(1) 2.132(2) 2.107(2)
V(1) - N(2) 2.337(6) 2.366(2) 2.291(2) 2.245(2)
V(1) - N(3) 2.184(4) 2.198(2) 2.173(2) 2.245(2)
V(1) - N(4) 2.104(4) 2.094(1) 2.104(2) 2.107(2)
V(1) - O(1)a 1.643(4) 1.609(1) 1.626(2) 1.626(2)
X - V(1) - N(1) 91.70(1) 88.76(5) 88.36(7) 98.08(7)
X - V(1) - N(2) 88.76(1) 89.02(4) 84.82(7) 88.13(7)
X - V(1) - N(3) 163.22(1) 165.55(5) 155.46(7) 161.70(7)
X - V(1) - N(4) 92.34(1) 94.19(5) 89.50(7) 91.70(7)
X - V(1) - O(1) 102.61(1) 102.30(5) 107.25(7) 107.92(7)
N(1) - V(1) - N(2) 74.37(2) 75.81(6) 73.73(7) 77.02(7)
N(1) - V(1) - N(3) 95.62(2) 93.44(6) 102.48(7) 89.96(7)
N(1) - V(1) - N(4) 169.33(2) 164.54(6) 171.93(8) 163.37(8)
N(2) - V(1) - N(3) 78.75(2) 77.70(6) 77.40(7) 77.64(7)
N(2) - V(1) - N(4) 95.85(2) 89.04(6) 98.33(7) 89.96(7)
N(3) - V(1) - N(4) 78.00(2) 80.05(6) 76.68(7) 77.02(7)
O(1) - V(1) - N(1) 92.85(2) 95.38(7) 91.87(8) 91.70(8)
O(1) - V(1) - N(2) 163.28(2) 165.62(7) 161.19(8) 161.70(8)
O(1) - V(1) - N(3) 92.11(2) 91.73(7) 94.48(8) 88.13(8)
O(1) - V(1) - N(4) 95.89(2) 98.81(7) 96.19(8) 98.08(8)
a

X corresponds to Cl(1) for [VOCl(dbqch)]ClO4·2CH3CN and [VOCl(dbqen)]ClO4·2CH3CN. X corresponds to F(1) for [VOF(H2bqch)]ClO4·CH3OH and to F(1) or O(1) for [VOF(H2bqen)]BF4.

b

In this structure, there is a disorder between oxygen and fluorine atoms, and thus, the reported d(V–F) in Table 2 is a mean value of the d(V–F) and d(V = O).

A perspective view of the structure of the cation of 2”, cis-[VIV(O)(Cl)(dbqch)]+ with the atomic numbering scheme used is shown in Figure 1A. The structure of [VIV(O)(Cl)(dbqch)]+ reveals that the ligand adopts a cis-α topology around the vanadium(IV) center with the two quinoline rings trans to each other and the two N–CH3 groups in an anticonformation. The vanadium(IV) atom in [VIV(O)(Cl)(dbqch)]+ is bonded to a tetradentate (Nq,Na,Na,Nq) dbqch ligand, and an oxido and chlorido ligands. The donor atoms surrounding the vanadium(IV) atom are disposed in a severely distorted octahedral geometry where the two quinoline nitrogens N(1) and N(4), the amine nitrogen atom N(3) and the chloride ion occupy the equatorial plane, while the amine nitrogen atom N(2) and the oxido ligand occupy the axial positions. The dbqch ligand forms three five-membered fused chelate rings. The vanadium(IV) bond distances of the trans-quinoline nitrogens [VIV–N(1) = 2.102(4) and VIV–N(4) = 2.104(4) Å] are noticeably shorter than those of the amine nitrogens N(2) and N(3) [2.337(4) and 2.184(4) Å]. The two VIV–Namine bond lengths are substantially different due to the strong trans influence of the oxido ligand. The long V(1)–N(2) bond length [2.337(4) Å] shifts the equatorial N(3) amine donor atom 0.198 (5) Å for 2” [0.245 (3) Å for 2’] under the equatorial plane defined by the two quinoline N and the Cl donor atoms. Consequently, the quinoline ring, parallel to equatorial plane, tilts, forming a 17.7(1) o for 2” [16.3(1) o for 2’] angle with the equatorial plane.

Figure 1.

Figure 1

ORTEP plot of the cations of 2’ (A) and 4’ (B) (the dimethylated-chlorido derivatives), with 50% thermal ellipsoids. Hydrogen atoms are omitted for clarity.

The d(VIV=O) of 1.643(4) Å lies in the upper limit of the range observed for oxidovanadium(IV) complexes (1.56–1.66 Å),4852 while the d(VIV–Cl) of 2.178(2) Å lies in the expected range.50,52,53 The structure of 4’ (Figure 1B) has very similar structural features with those of 2’: therefore, it will not be discussed.

The molecular structures of the cis-[VIV(O)(F)(H2bqch)]ClO4·CH3OH (5’) and cis-[VIV(O)(F)(H2bqen)]BF4 (6’) are depicted in Figure 2A and 2B, respectively. In cis-[VIV(O)(F)(H2bqch)]ClO4·CH3OH (5’), the vanadium(IV) atom is coordinated to two quinoline N atoms, two secondary amine N atoms, a fluorine atom, and an oxygen atom. The vanadium adopts a highly distorted octahedral geometry and is displaced above the mean equatorial plane, defined by the two quinoline N atoms, one secondary amine N, and a fluorine atom, by 0.256 Å toward the oxido ligand. The long V(1)–N(2) bond length [2.292(2) Å] shifts the equatorial N(3) amine donor atom 0.465 (2) Å under the equatorial plane defined by the two quinoline N and the F donor atoms (Scheme 2). The quinoline ring, parallel to the equatorial plane, tilts, forming a 16.50(6)° angle with the equatorial plane. The fluorine atom is coordinated in a cis position to the oxido ligand and the d(VIV–F) of 1.834(1) Å has been, to our knowledge, the shortest observed for oxyfluoride vanadium(IV) compounds.5457 Compound cis-[VIV(O)(F)(H2bqen)]BF4 (6’) has similar structural characteristics: therefore, it will not be discussed.

Figure 2.

Figure 2

ORTEP plot of the cations of 5’ (A) and 6’ (B) (the amine-fluorido derivatives), with 50% thermal ellipsoids. Hydrogen atoms are omitted for clarity.

IR Spectroscopy

The IR spectra of the six oxidovanadium(IV) compounds exhibit a very strong and sharp band in the range 962–977 cm–1, which was assigned to ν[VIV(O)] (see Figures S1–S6). The IR spectra of 5 and 6 (the fluorido compounds) reveal a moderate sharp band at 563 and 557 cm–1 respectively (see Figures S5–S6), which is missing in the spectra of the chlorido compounds 14 (see Figures S1–S4), and this band was assigned to ν(VIV–F).

Catalytic Evaluation

The present oxidovanadium(IV) compounds 16 were used as catalysts for the cyclohexane oxidation with H2O2 (30% w/w) at room temperature (25 ± 0.5 °C). In catalytic reactions, the used molar ratio was [catalyst: H2O2: cyclohexane] = [1:1000:2500 μmoles] in the presence or not of 100 μmoles of HCl. Catalytic data including product yields (%), TON, and TOF are given in Table 3.

Table 3. Catalytic Oxidation of Cyclohexane by Oxidovanadium(IV) Complexes 16 in the Presence of H2O2.

VIVO-compounds products yield (%)c TONd TOF (h–1)e
1a cyclohexanol 17.7 241 40.2
cyclohexanone 6.4
1b cyclohexanol 15.2 203 33.8
cyclohexanone 5.1
2a cyclohexanol 5.0 77 12.8
cyclohexanone 2.7
2b cyclohexanol 14.3 195 32.5
cyclohexanone 5.2
3a cyclohexanol 15.2 213 35.5
cyclohexanone 6.1
3b cyclohexanol 13.2 198 33
cyclohexanone 6.6
4a cyclohexanol 6.0 102 17
cyclohexanone 4.2
4b cyclohexanol 10.8 173 28.8
cyclohexanone 6.5
5a cyclohexanol 9.5 111 18.5
cyclohexanone 1.6
5b cyclohexanol 21.6 241 40.2
cyclohexanone 2.5
6a cyclohexanol 18.6 293 48.8
cyclohexanone 10.7
6b cyclohexanol 15.7 234 39
cyclohexanone 7.7
a

Conditions: molar ratio of [catalyst:H2O2:substrate]= [1:1000:2500] in 1 mL of CH3CN (1 equiv = 1 μmol).

b

Conditions: molar ratio of [catalyst:HCl:H2O2:substrate]= [1:100:1000:2500] in 1 mL of CH3CN (1 equiv = 1 μmol).

c

Yields based on the starting substrate and products formed. The mass balance is 98–100%. The reaction time was 6 h.

d

TON: total turnover number, moles of products formed per mole of catalyst.

e

TOF: turnover frequency which is calculated by the expression [products]/[catalyst] × time (h–1).

According to Table 3, the oxidovanadium(IV) compounds 16 are able to oxidize cyclohexane with hydrogen peroxide at room temperature. More specifically, oxidation of cyclohexane catalyzed by 1 and 3 produced cyclohexanol and cyclohexanone with 17.7, 6.4% and 15.2, 6.1% yields, respectively resulting in 24.1 and 21.3% total yields (Figure 3). The addition of 100 μmoles of HCl in the catalytic reaction led to reduced total yields, i.e., 20.3 and 19.8% for 1 and 3, respectively. The corresponding methylated oxidovanadium(IV) compounds 2 and 4 without HCl provided total yields for cyclohexane oxidation 7.7 and 10.2%, respectively, which increased in the presence of 100 μmoles of HCl, 19.5% and 17.3%, respectively (Figure 3). The nonmethylated fluorido compounds 5 and 6 gave total yields 11.1% and 29.3%, respectively, which were higher than their chlorido analogues (Table 3). Cyclohexane oxidation catalyzed by 5 and 6 was not affected by HCl as a promoter; its presence resulted in even lower yields (24.1% and 23.4% respectively). TONs achieved by the catalysts 16 ranged from 77 to 293 and are visualized in Figure 4.

Figure 3.

Figure 3

Distribution of oxidation products catalyzed by the oxidovanadium(IV) compounds 16 in the presence of H2O2. Conditions: molar ratio of [catalyst:H2O2:substrate]= [1:1000:2500] in 1 mL CH3CN (1 equiv = 1 μmol) or molar ratio of [catalyst:HCl:H2O2:substrate] = [1:100:1000:2500] in 1 mL CH3CN (1 equiv = 1 μmol). Yields based on the starting substrate and products formed. The reaction time was 6 h. See Table 3 for further details on reaction conditions.

Figure 4.

Figure 4

Turnover frequency of the oxidation of cyclohexane catalyzed oxidovanadium(IV) compounds 16. Conditions: molar ratio of [catalyst:H2O2:substrate] = [1:1000:2500] in 1 mL CH3CN (1 equiv = 1 μmol) or molar ratio of [catalyst:HCl:H2O2:substrate] = [1:100:1000:2500] in 1 mL of CH3CN (1 equiv = 1 μmol). Yields based on the starting substrate and products formed. The reaction time was 6 h. See Table 3 for further details on reaction conditions.

Based on our catalytic data, the addition of HCl to the cyclohexane oxidation, catalyzed by the methylated oxidovanadium(IV) compounds 2 and 4, increases the yield of oxidation products. For the nonmethylated compounds 1, 3, 5, and 6, the addition of HCl decreases the catalytic activity. Analogous negative effect on catalytic cyclohexane oxidation was observed when HCl was replaced by 2-pyrazine carboxylic acid (PCA) or HNO3 (data not shown). The use of PCA as promoter in alkane oxidation catalyzed by oxidovanadium(IV) compounds is well-known due to its assistance for H+ migration from a coordinated H2O2 to the oxido-ligand.45 Here, the observed chemical behavior of 1, 3, 5, and 6 reveals that the two −NH– groups in conjunction with the oxido-ligand are able to manage the hydrogen peroxide deprotonation which is coordinated to vanadium center toward homolytic O–O bond cleavage and generation of OH radicals. Cyclohexane oxidation most probably occurs via these OH radicals which abstract a cyclohexane hydrogen atom to form cyclohexyl radicals. The alkyl radicals in oxygenated organic solvents readily form alkyl hydroperoxides (cyclohexyl hydroperoxide in our case) as primary intermediate oxidation product which transformed to cyclohexanol and cyclohexanone.5860

Magnetism and X-Band Continuous-Wave (cw) EPR Spectra of 16

The magnetic moments of compounds 16, at 298 K, have magnetic moments in the range of 1.70–1.74 μB, in accord with the spin-only value expected for d1, S = 1/2 systems. These μeff values constitute clear evidence that the oxidation of vanadium in 16 is IV.

The X-band cw EPR parameters, of the frozen (120 K) solutions (DMSO) of the oxidovanadium(IV) compounds 16 are depicted in Table 4 and were calculated from the simulation of their experimental EPR spectra. The spectra of the compounds cis-[VIV(O)(Cl)(H2bqen)]Cl·2H2O (2H2O), cis-[VIV(O)(F)(H2bqch)]ClO4 (5) (Figure S7), and cis-[VIV(O)(F)(H2bqen)]ClO4 (6) (Figure S8) were successfully simulated assuming one VIV species in solution, while the spectra of cis-[VIV(O)(Cl)(H2bqch)]Cl·H2O (H2O), cis-[VIV(O)(Cl)(dmbqch)]Cl (2) (Figure 5), and cis-[VIV(O)(Cl)(dmbqen)]Cl·3H2O (3H2O) (Figure S8) were successfully simulated assuming the coexistence of another species (B) in equilibrium with A (Scheme 6). The spectra of 14 were simulated considering axial symmetry, while those of 5 and 6 were simulated with respect to rhombic symmetry. For the simulations of the spectra of 5 and 6, the superhyperfine coupling of the electron spin with the neighboring F was also included in the Hamiltonian. The A|| values of 14 are ∼177 × 10–4 and ∼155 × 10–4 cm–1 for the species A and B, respectively, and the Az values of 5 and 6 are ∼179 × 10–4 cm–1.

Table 4. Cw X-Band EPR Parameters of the DMSO Frozen Solutions of Oxidovanadium(IV) Compounds 16a.

compound gx,gy(g) gz(g||) Ax,Ay(A) × 10–4 (cm–1) Az(A||) × 10–4 (cm–1) Ax(F),Ay(F)(A⊥(F))× 10–4 (cm–1) Az(F)(A||(F)) × 10–4 (cm–1) equatorial coordination environment
1 (A, 79%) 1.971 1.929 –65.9 –177.6     N3Cl
1 (B) 1.976 1.947 –54.9 –155.6     N2Cl
2 (A, 35%) 1.970 1.928 –65.7 –177.5     N3Cl
2 (B) 1.974 1.952 –53.4 –156.9     N2Cl
3 (A) 1.970 1.928 –65.8 –177.4     N3Cl
4 (A, 24%) 1.971 1.928 –66.0 –177.4     N3Cl
4 (B) 1.975 1.953 –53.8 –157.2     N2Cl
5 (A) 1.976, 1.965 1.927 –65.5, −65.6 –178.8 15.5, 20.8 10.4 N3F
6 (A) 1.976, 1.964 1.928 –65.6, −65.6 –179.2 13.5, 18.2 9.7 N3F
a

The A and B correspond to the two identified isomers depicted in Scheme 6.

Figure 5.

Figure 5

X-band cw EPR spectrum of a frozen solution of the compound cis-[VIV(O)(Cl)(dmbqch)]Cl (2, species A and B) in DMSO (1.00 mM) at 120 K and its simulated spectrum.

Scheme 6. Equilibrium of the Five- (A) cis-[VIV(O)(Cl)N4]+ Oxidovanadium(IV) Compounds (16) in Solution (DMSO) and Six-Coordinate (B).

Scheme 6

The structure B is identical with that determined from the single-crystal X-ray structure analysis of 2 and 4.

Theoretical calculations support the presence of two possible minimum energy structures for the compounds 16 in solution, (a) the six-coordinate distorted octahedral structure (Scheme 6B) found in the single crystal structures of 2, 4, 5, and 6 and (b) a five-coordinate species (Scheme 6A) formed from the dissociation of the axial to oxido group amine nitrogen atom, exhibiting a highly distorted trigonal bipyramidal structure (Scheme 7) with a trigonality index τ = 0.51 [τ = (a–b)/60 = 0.51; a = N(3)–V–Cl = 134.0°, b = N(1)–V–N(4) = 164.7°).61 Other possible structures of 16 in solution, such as five-coordinate species formed from the dissociation of Cl, result in high energy species based on theoretical calculations (vide infra). In addition, conductivity measurements of the DMSO and CH3CN solutions of 16 gave values 55–65 cm–1 mol–1 Ω–1 (DMSO) and 130–150 cm–1 mol–1 Ω–1 (CH3CN), as expected for 1:1 electrolytes. Thus, it is clear from the conductivity measurements that the Cl or F donor atoms do not dissociate in solution. DFT calculations of the EPR parameters of 1 (structure A) and 1 (structure B) at the BHandHLYP/6-311g (d,p) level of theory predict Az = −173.8 × 10–4, Ay = −75.2 × 10–4, Ax = −71.6 × 10–4 and Az = −152.8 × 10–4, Ay = −56.9 × 10–4, Ax = −54.1 × 10–4 for the five-coordinate (structure A) and six-coordinate (structure B) respectively. The theoretically predicted Az values are ∼2.5% lower than the experimental, due to the accuracy of the method used, and this deviation is similar to the deviation reported for the Gaussian calculations of charged vanadium complexes at the same level of theory.62

Scheme 7. Equilibrium Geometry of the (A) [VIV(=O)(Cl)(N4H)]2+, τ = 0.51, d(V···N2) = 3.303 Å) and (B) [VIV(=O)(F)(N4H)]2+, τ = 0.70, d(V···N2)= 3.341 Å, in Acetonitrile Solution Optimized at the PBE0/Def2-TZVP(V)υ6-31+G(d)(E)/PCM Level of Theory and Selected Bond Distances in Å.

Scheme 7

The A|| or Az parameters depend on the donor atoms in the equatorial plane of the vanadium(IV) compounds and can be calculated from the empirical additivity relationship (eq 4).63,64

graphic file with name ic2c02526_m001.jpg 4

A||,i is the contribution of each donor atom to A||.

The donor atoms in the equatorial plane of 14 consist of a Cl, two quinoline N (Nq), and one aromatic amine N (NArNH2) atoms. The A||,i contributions Nq and NArNH2 have not been determined previously. The A||,i value of other aromatic heterocyclic N donor atoms, such as imidazole, pyridine, etc., and NRNH2 were used instead for the contribution of Nq and NArNH2 respectively.65 The calculated A|| value using eq 4 is approximately −165 × 10–4 cm–1. However, the experimental and the calculated values of A|| are significantly lower for the octahedral species B and significantly higher than the five-coordinate species A.

The dramatic decrease of the experimental A|| values of species B, compared with the values calculated from the additivity relationship, is attributed to the coordination of the amine nitrogen in the axial position trans to the oxido group (Scheme 6).66 Tolis et al. have also suggested that axial donor atoms induce a radial expansion of the vanadium dxy orbital, resulting in a reduced electron density on the VIV and decrease of Az.67

On the other hand, the much higher Az (−177.5 × 10–4 cm–1) experimental values of species A in comparison to the predicted Az values for 16 from the additivity relationship are attributed to the distortion in the equatorial plane by the elongation of V–N(1) (2.307 Å), due to the tension in the N(3)---N(1) eight-membered ring (Scheme 7). Apparently, the weakening of the bonding at the equatorial plane results in an increase of Az values.

The equilibrium between species A and B is shifted toward B, when the amine hydrogen atoms of the ligands (H2bqch, H2bqen, in compounds 1, 3) are replaced with the bulky methyl groups (dbqch, dbqen in compounds 2, 4). In addition, theoretical calculations revealed that 3(A) is thermodynamically more stable than 3(B), whereas 4(B) is thermodynamically more stable than 4(A). In addition, from the quantities of B in the solution being 21% and 0% for the compounds 1 (the cyclohexane derivative) and 3 (the ethylenediamine derivative), respectively, it is reasonable to conclude that cyclohexane-1,2-diamine chelate ring is more rigid than the 1,2-ethylenediamine one. The chelate ring defined by the vanadium(IV) atom and the two amine nitrogen atoms is stretched due to the elongation of the bond VIV–Nam.axial. Moreover, the attachment of the methyl groups to the amine nitrogen atoms increases the steric interactions between Cl and the −CH3 group (Scheme 8) forcing equatorial Namine to remain ligated to vanadium nucleus, forming the six-coordinate species B (Scheme 6). Dissociation of the equatorial Namine atom results in the formation of an eight-membered chelate ring (Scheme 7) similar to the chelate rings for other VIV compounds reported and characterized by crystallography.68

Scheme 8. (A) Possible Mechanism with Which the Steric Hindrance of the Bulky Methyl Group Forces the VIV Compounds to Acquire the Six-Coordinate Structure in Solution. (B) In the Absence of Steric Hindrance, the Compounds Adopt the Five-Coordinate Structure in Solution.

Scheme 8

The EPR parameters calculated from the simulation of the experimental spectra reveal that compounds 5 and 6 acquire the structure A in DMSO. This might be attributed to the stronger trans effect of F than Cl on N(3) (Scheme 7). On the basis of the additivity relationship,63,64,69,70 and considering Az contribution for F, either −40.1 × 10–4 cm–1 {cis-[VIV(O)(F)(4,4′-dtbipy)2]BF4}55 or −41.8 × 10–4 cm–1 {[VIV(O)(F)2(DMSO)3]},71 one would expect lower experimental A|| values for 5 and 6 in comparison to those of 14 (Table 4). In marked contrast, the Az values of 5 and 6 were slightly higher. Theoretical calculation of 6(B) (Scheme 7) at BHandHLYP/6-311g (d,p) level gives a value for Az (−178.3 × 10–4 cm–1) which is very close to the experimental one. The higher experimental Az values of 5 and 6 than 14 and the higher Az calculated values using eq 4, are attributed to higher trigonality index of 5 and 6 (∼0.70) than 14 (∼0.51).61,66 The stronger trans effect of F than Cl causes lengthening of the V–N(3) bond (Scheme 7), increasing the tension in the chelate rings. The energy of the compounds 5 and 6, decreases by adopting trigonal bipyramidal structure in solution. The increase of the trigonality index in 5A and 6A increases the distance between the vanadium atom and N(2), resulting only in five-coordinate species in the solutions of 5 and 6.(61) The failure to synthesize the V–F compounds with the sterically hindered dbqch and dbqen ligands (the dimethylated molecules) is attributed to the high energy, required for these ligands, to adopt trigonal bipyramidal structure in solution (Scheme 8). The sterically hindered dbqch and dbqen ligands in 2, and 4, force N(2) close to the vanadium atom (vide supra), taking octahedral or distorted square pyramidal structures only. The low spin - 19F superhyperfine coupling constant of 5 and 6 (∼15 × 10–4 cm–1) than cis-[VIV(O)(F)(4,4′-dtbipy)2]BF4 (41 × 10–4 cm–1),44 indicate that the VIV–F interactions in 5 and 6 have a much smaller covalent character than the VIV–F bond in cis-[VIV(O)(F)(4,4′-dtbipy)2]BF4.

The X-band cw EPR spectra of the frozen solution of the compounds 16 in CH3CN gave a broad unresolved peak centered at g = 1.982 (Figure 6). This spectrum improves with the addition in CH3CN of solvents with high dielectric constants such as H2O, DMSO etc. In contrast, the X-band EPR spectra of the CH3CN solutions at room temperature of 16 gave well resolved octaplets of both isomers confirming that A and B are present in CH3CN solutions (Figure S9).

Figure 6.

Figure 6

X-band cw EPR spectra of a frozen (120 K) solution (CH3CN, 1.25 mM) of the compounds cis-[VIV(O)(Cl)(dbqen)]Cl·3H2O (A) (3H2O) and cis-[VIV(O)(F)(H2bqch)]ClO4 (B) (5) with various quantities of aqueous HCl. The EPR parameters of 4·3H2O and 5 are the same with those of both compounds in frozen DMSO solution.

Addition of aqueous HCl into the CH3CN solution of 3H2O (Figure 6I) and 5 (Figure 6II) results in well-resolved spectra that contain both species A and B. Increasing the quantity of aqueous HCl into the CH3CN solution of 16 the equilibrium is shifted toward A, and this is in line with the theoretical calculations (vide infra). Extrapolation of the quantities of A vs the quantity of aqueous HCl in CH3CN shows that both A and B are present in pure CH3CN. The X-band cw EPR spectra of the CH3CN solutions of 16 gave well resolved octuplets of both isomers confirming that A and B are present in CH3CN solutions (Figure S9).

Speciation of the Catalytic Reaction Mixtures with 51V NMR Spectroscopy and 5,5-Dimethylpyrollidone Oxide (DMPO) Trap EPR Experiments

The 51V and 1H NMR spectra of 3 in solution (CH3CN, 5.0 mM) after the addition of H2O2 (5.0 M, 30%) are shown in Figures S10 and S11 respectively. The 51V NMR spectrum of 3 shows the presence of two peaks at −606 ppm and −674 ppm assigned to the monoperoxido and bisperoxido VV species respectively, which are absent from the 51V NMR spectrum of an CH3CN solution of 3. Thus, it is obvious that the VIV of compound 3 is oxidized to VV upon addition of H2O2. Addition of aqueous HCl (50 mM) to the CH3CN solution of 3+H2O2 results in the appearance of a new peak at −569 ppm (Figure S10) assigned to the dioxido VV and its formation is due to the partial decomposition of peroxido VV complexes. The 1H NMR of the ligand at the same conditions and the spike experiments show that the solutions of 3 + H2O2 and 3 + H2O2 + aqueous HCl do not contain free ligand (Figure S11). Apparently, the VV peroxido and dioxido species retain the ligands attached to the metal ion.

The cw X-band EPR spectra of 3 in solution (CH3CN, 1.0 mM) + H2O2 (10 mM, 30%) + DMPO (1.0 mM) vs time are shown in Figure S12. After the addition of H2O2 into the CH3CN solution of 3 + DMPO at zero time the EPR spectrum shows a strong peak at g = 2.0153 assigned to the radical of DMPO adduct with various radicals that might be formed in solution including superperoxide and hydroxide radicals. This peak after ∼30 min turned to an 9-fold peak at g = 2.0044 and AN ∼ 7 G and AH ∼ 4G identified as 5,5-dimethyl-pyrrolidone-(2)-oxyl- (DMPOX) the oxidation product of DMPO-·OH as assigned previously.7274 In conclusion, the VIV of the catalysts is oxidized to VV upon addition of H2O2 and the ligands remain bound to the vanadium atom under the conditions of catalysis, whereas, the mechanism of the catalytic reaction is through hydroxyl radicals.

Mechanistic Details for the Reactivity of the cis-[VIV(O)(Cl)(N4)]+ Compounds, with F

The substitution reaction of chloride by fluoride in the cis-[VIV(O)(Cl)(N4)]+ (N4 = H2bqen, H2bqch, dbqen, dbqch) compounds was modeled through DFT methods. The dissociative (D), associative (A) and concerted interchange (both the dissociative Id and associative Ia variations) mechanisms were explored for this reaction. A representative geometric and energetic profile for the ligand substitution reaction of the octahedral cis-[VIV(O)(Cl)(H2bqen)]+ is shown in Figure 7. The equilibrium geometries of the cis-[VIV(O)(Cl)(N4)]+ (N4 = H2bqen, H2bqch, dbqen, dbqch) compounds between the 5-coordinate [VIV(O)(N4)]2+ and the 7-coordinate [VIV(O)(Cl)(F)(N4)] transition states in acetonitrile solutions, optimized at the PBE0/Def2-TZVP(V)υ6-31+G(d)(E)/PCM level of theory along with selected structural parameters, are given in Figures S13, S14. The optimized structural parameters of the cis-[VIV(O)(Cl)(N4)]+ (N4 = dbqen, dbqch) and cis-[VIV(O)(F)(H2bqch)]+ compounds are in line with those derived from the X-ray structural analysis.

Figure 7.

Figure 7

Geometric and energetic profile of the substitution of Cl by F in the octahedral cis-[VIV(O)(Cl)(H2bqen)]+ complex, following the D and Id pathways calculated by the PBE0/Def2-TZVP(V)Inline graphic6-31+G(d)(E)/PCM computational protocol in acetonitrile solutions.

Substitution of Cl by F ligand is not reasonable to follow the Ia mechanism, since all attempts to identify a 7-coordinate transition state or intermediate in these reactions were not successful. The dissociative mechanism (Figure 7) is not favored since the dissociation of Cl needs relatively high activation energy ∼32.5 kcal/mol for the formation of the 5-coordinate intermediate. It is more likely that the substitution reaction follows the concerted dissociative interchange Id pathway. This pathway is “free” of any activation barrier, since the formation of the 7-coordinate transition state releases energy 11.8 kcal/mol with the concerted dissociation of Cl demanding only 13.7 kcal/mol of energy. The methyl substituents on the amine N atoms of the N4 ligand hinder the approach of the incoming F ligand to attack the vanadium(IV) atom of the compounds to form the 7-coordinate transition state. This fact is in line with the experimental data (vide supra) and explains why our efforts to prepare the cis-[VIV(O)(F)(N4,dm)]+ derivatives starting from cis-[VIV(O)(Cl)(N4,dm)]+ have failed (N4,dm = the dimethylated derivatives).

Mechanistic Studies of cis-[VIV(O)(Cl/F)(N4)]+ Catalysts through DFT Computations

The oxidation of alkanes catalyzed by vanadium-based catalytic systems proceeds via hydroxyl radicals (OH), generated upon metal catalyzed decomposition of H2O2, which abstract hydrogen atoms from alkanes (RH) to form alkyl radicals (R).75 The energetic profiles for pathways (A and B) that generate OH radicals upon homolytic cleavage of the HO–OH bond catalyzed by the cis-[VIV(O)(X)(N4)]+ (X = Cl, F; N4 = H2bqen, H2bqch) complexes are shown in Figure 8.

Figure 8.

Figure 8

Energetic profiles for the reaction pathways A and B that generate hydroxyl OH radicals upon homolytic cleavage of the HO–OH bond catalyzed by the cis-[VIV(O)(X)(N4)]+ (X = Cl, F; N4 = H2bqen, H2bqch) compounds calculated by the PBE0/Def2-TZVP(V)υ6-31+G(d)(E)/PCM computational protocol in acetonitrile solutions.

The first step in both pathways, A and B, involves the nucleophilic attack on the vanadium metal center by the H2O2, which is assisted by a hydrogen bond formation O···H–N between the distal O atom of the coordinated H2O2 and the H atom of the secondary amino moiety. In the methylated catalysts the presence of the methyl groups in the inner coordination sphere of the catalysts hinders the nucleophilic attack on the vanadium atom by the H2O2, and this is in line with the low yields being 7.7 and 10.2% for 2 and 4 catalysts, respectively. According to NBO analysis, the vanadium central atom acquires positive natural atomic charge ranging from 0.604 up to 0.943 |e|. The natural atomic charge on the vanadium metal center is higher in the fluorido- cis-[VIV(O)(F)(N4)]+ than in the chlorido- cis-[VIV(O)(Cl)(N4)]+ compounds. Therefore, the cis-[VIV(O)(F)(N4)]+ compounds are more susceptible to nucleophilic attack by H2O2. Interestingly the N donor atoms of the groups -NH- and -N(CH3)- acquire higher negative natural atomic charges (negative natural atomic charges in the range of −0.507 up to −0.682 |e|) than the two quinoline N donor atoms (−0.429 up to −0.457 |e|) of the N4 ligand and the X (−0.339 up to −0.541 |e|) and O (−0.417 up to −0.501 |e|) donor atoms of the catalysts.

In the reaction pathway A, the second step involves the homolytic cleavage of the O–O bond in the [VV(O)(H2O2)(X)(N4)]2+ (X = F, Cl) species generating directly OH radicals and the (oxido)(hydroxido) [VV(O)(OH)(X)(N4)]+ species. The estimated energy barriers for the generation of the OH radicals are 19.1, 16.2, 25.7, and 22.4 kcal/mol for the [VV(O)(H2O2)(Cl)(H2bqen)]2+, [VV(O)(H2O2)(F)(H2bqen)]2+, [VV(O)(H2O2)(Cl)(H2bqch)]2+, and [VV(O)(H2O2)(F)(H2bqch)]2+ species, respectively. According to the estimated energy barriers for the [VV(O)(H2O2)(X)(H2bqen)]2+ and [VV(O)(H2O2)(X)(H2bqch)]2+ species, the catalytic efficacy of the former should be higher than the latter. This is in line with the experimental catalytic activity of the fluorine vanadium compounds, 5 and 6. The chloride (X = Cl) exhibits the same catalytic activity, and this can be interpreted if we assume that Cl compounds follow an alternative mechanism, pathway B (vide infra). The resulting from the homolytic cleavage of the O–O bond [VV(O)(OH)(X)(N4)]+ species reacts with protons and yields the (oxido)(aquo) [ViV(O)(OH2)(X)(N4)]+ species (Figure 8) which releases directly the aquo ligand to regenerate the cis-[VIV(O)(X)(N4)]+ catalysts.

In the reaction pathway B (Figure 8), the second step involves coordination of H2O2 nucleophile to the vanadium atom promoting the dissociation of the leaving ligand X (X = F, Cl) yielding the transient cis-[VV(O)(O2H2)(H2bqen)]3+ and cis-[VV(O)(O2H2)(H2bqch)]3+ species. The third step along the reaction pathway B involves the homolytic cleavage of the O–O bond in the cis-[VV(O)(O2H2)(N4)]3+ (N4 = H2bqen, H2bqch) species yielding the cis-[VV(O)(OH)(H2bqen)]2+ cation. The homolytic cleavage of the HO–OH bond, of the vanadium coordinated H2O2, demands very low energy (around 4.5 kcal/mol), while the energy of the homolytic cleavage of the “free” H2O2 is 44.1 kcal/mol at the PBE0/6-31+G(d)(E)/PCM level of theory. Next, the cis-[VIV(O)(OH)(N4)]+ species reacts with protons and Cl with concomitant release of the aquo ligand to regenerate the [VIV(O)(X)(N4)]+ catalysts.

The breaking of the V–F and V–Cl bonds demands an energy barrier 48 and 22 kcal/mol, respectively. The high energy barriers for breaking the V–F bonds are not in favor of the reaction pathway B for the fluorine VIV species. Pathway A predicts the catalytic activity of flourido- vanadium complexes. Unlikely, pathway B predicts chloride- complexes to have higher activity than the flourido- complexes and 5 and 6 to exhibit the same activity. On the other hand, the chloride complexes 1 and 3 follow pathway B, predicting both compounds to exhibit similar catalytic activity in line with the experiment. The energy barriers for the [VV(O)(H2O2)(H2bqen)]3+ and [VV(O)(H2O2)(H2bqch)]3+ species are almost the same.

Apparently, the oxidation of alkanes catalyzed by vanadium-based catalytic systems proceeds through pathway A for the fluoride/vanadium and through pathway B for the chloride/vanadium compounds. The reaction pathway is controlled by the strength of the V–X (X = F, Cl) bond. The lower activity of the N–CH3 than N–H vanadium compounds is attributed to the steric hindrance caused by the methyl groups, hindering H2O2 approach to vanadium at the first step of the reaction.

Mechanistic Details for the Catalytic Oxidation Reactions with Addition of HCl

The catalytic activity of the oxidovanadium(IV) compounds toward cyclohexane oxidation in the presence of HCl was also investigated by DFT calculations on the protonated at the -N(CH3)- moieties of the [VIV(O)(Cl)(N4H+)]2+ (N4H+ = dbqenH+, dbqchH+) compounds. The equilibrium geometries of all species and products involved in the reaction pathways that yield the hydroxyl OH catalyzed by the protonated methylated [VIV(O)(X)(N4H)]2+ (X = Cl, F; N4H+ = dbqenH+, dbqchH+) catalysts optimized at the PBE0/Def2-TZVP(V)υ6-31+G(d)(E)/PCM level of theory in acetonitrile solutions, along with selected structural parameters are given in Figure S15.

The protonation of -N(CH3)- induces remarkable changes on the inner coordination sphere of the vanadium (Figure 9), more specifically: (i) its coordination number changes from six to five, in line with the EPR experiment, and thus, leaving an open site for the coordination of H2O2 to vanadium atom. ιi) the amine hydrogen atom of [-NH(CH3)-]+ assists the H2O2 attack to the vanadium atom by the formation of a hydrogen bond between the distal O atom of the coordinated O2H2 and the amine hydrogen atom of the [NH(CH3)]+ moiety.

Figure 9.

Figure 9

Energetic profile for the reaction pathway that generates hydroxyl OH radicals catalyzed by the [VIV(O)(Cl)(N4H)]2+ (N4H+ = dbqenH+, dbqchH+) compounds calculated by the PBE0/Def2-TZVP(V)υ6-31+G(d)(E)/PCM computational protocol in acetonitrile solutions.

The formation of the [VV(O)(H2O2)(Cl)(dbqenH+)]3+ species is slightly exothermic (ΔH = −4.3 kcal/mol), while the formation of the [VV(O)(H2O2)(Cl)(dbqchH+)]3+ species is slightly endothermic (ΔH = 3.4 kcal/mol).

The second step involves the homolytic cleavage of the O–O bond in the [VV(O)(H2O2)(Cl)(N4H+)]3+ (N4H+ = dbqenH+, dbqchH+) species generating directly OH radicals and [VV(O)(OH)(Cl)(N4H+)]2+ species, through a strongly exothermic dissociation process (ΔH ∼ −128 up to −131 kcal/mol). Next, the [VV(O)(OH)(Cl)(N4H+)]2+ cation reacts with protons (HCl) and yields the (oxido)(aquo) [VIV(O)(OH2)(Cl)(N4H+)]2+ species (Figure 9) which releases the aquo ligand and the catalysts [VIV(O)(Cl)(N4H+)]2+. The transformation of [VV(O)(OH)(Cl)(N4H+)]2+ cation to [VIV(O)(Cl)(N4H+)]2+ species in the acidic media is strongly exothermic (ΔH ∼ −143 up to −156 kcal/mol).

The pathway in Figure 9 agrees with the experimental efficiency toward the catalytic oxidation of alkanes observed for all the complexes 16 in the presence of HCl.

Conclusions

In summary, a series of four oxidovanadium(IV) compounds of the general formula cis-[VIV(O)(Cl)(N4)]Cl was prepared by reacting VIVOCl2 with the nonplanar tetradentate N4 bis-quinoline ligands. Sequential treatment of the two nonmethylated N4 oxidovanadium(IV) compounds with KF and NaClO4 resulted in the isolation of the species with the general formula cis-[VIV(O)(F)(N4)]ClO4. The oxidovanadium(IV) compounds were physicochemically and structurally characterized.

Their catalytic oxidation reactions of the highly distorted octahedral VIVO2+ compounds, mimicking the irregular geometries of the coordination environment of the metal ions in proteins, with the nonplanar N4 tetradentate amine ligands were examined. The distortion of the coordination sphere of the VIVO2+ cation induced by the N4 ligands was further enforced by partially replacing ligand’s H- with bulky cyclohexyl- and/or methyl- groups and by introducing F- or Cl- coligands in the VIVO2+ coordination sphere.

The experimental EPR parameters of these distorted VIVO2+ compounds deviate from those calculated from the empirical additivity relationship. The deviation has been assigned either to the coordination of the axial nitrogen donor atom or the trigonal distortion of the VIV coordination environment. cw X-band EPR speciation studies in frozen polar solvents reveal that the introduction of the hindered cyclohexyl- and methyl- groups causes retention in solution of the octahedral solid-state crystal structure, whereas, ligands without steric hindrance allow dissociation of one of the ligand’s amine donor atom from the six-coordinate sphere of VIV ion in solution, resulting in five-coordinate structures. Based on the equilibrium between six- and five-coordinate species, we concluded that the steric hindrance in the VIVO2+ compounds is increasing according to the following series, -HNCH2CH2NH- > -HNC6H10NH- > -(CH3)NCH2CH2N(CH3)- > -(CH3)NC6H10N(CH3)-. cw X-band EPR spectra of the VIVO2+ compounds in frozen CH3CN show that 16 five- or both five and six-coordinate structures, however addition of aqueous HCl into their CH3CN solution results in the full dissociation of the equatorial amine group and the formation of only five-coordinate species. The sterically hindered compounds 2 and 4, containing the dimethylated ligands, inhibit the approach of the nucleophiles (F, H2O2) to the vanadium nucleus, resulting in unsuccessful replacement of Cl ligand by the F and lower oxidative catalytic activity compared with the less sterically hindered 1 and 3, which contain the nonmethylated ligands.

The variation of the oxidative catalytic activities between the chloride and fluoride VIV compounds is attributed to two different mechanisms of catalytic action controlled by the V-X (X = F, Cl) bond strengths (V–F is stronger than V–Cl). The generation of OH radical for the cis-[VIV(O)(Cl)(N4)]+ species takes place via the dissociation of Cl, while for the cis-[VIV(O)(F)(N4)]+ species via the formation of seven-coordinate [VIV(O)(F)(H2O2)(N4)]+ cation. The distortion of the coordination environment of the VIV ion, mimicking the active site of metal-proteins, can be used as a highly desirable methodology allowing for the modification of the functionality of the metal compounds such as in the case of oxidative catalysis.

The vanadium(IV) of the compounds 16 is oxidized to vanadium(V) upon addition of H2O2 and the ligands remain bound to the vanadium atom under the conditions of catalysis, as it was evidenced with 51V and 1H NMR spectroscopies. cw X-band EPR trap studies proved that the mechanism of the catalytic reaction is through hydroxyl radicals.

Suitable ligands that introduce the desirable amount of distortion on the metal ion’s coordination environment can result in a fruitful design approach for the development of effective catalysts tailored for specific applications.

Acknowledgments

The research work was funded by the European Regional Development Fund and the Republic of Cyprus through the Research and Innovation Foundation (Project: EXCELLENCE/1216/0515). H.N.M. thanks the University of Glasgow for supporting this work. We thank Prof. Sproules for his help with EPR discussion.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.2c02526.

  • UV–vis-NIR, FT-IR, NMR spectra, and crystallographic data (PDF)

Author Contributions

Conceptualization, T.A.K., A.D.K. and H.N.M.; synthesis of the ligands and of the vanadium(IV) compounds M.G.P.; solid-state IR M.G.P.; H.N.M.; crystallography, A.D.K. and H.N.M.; catalysis, A.V.S. and M.L.; EPR, C.D., and A.D.K.; DFT calculations, A.C.T.; writing-original draft preparation, T.A.K., H.N.M., A.D.K., M.L., A.C.T., and M.G.P.; writing—review and editing, T.A.K., H.N.M., and A.D.K.; supervision of all contributions, T.A.K., H.N.M., and A.D.K. All authors have read and agreed to the published version of the manuscript.

The authors declare no competing financial interest.

Supplementary Material

ic2c02526_si_001.pdf (1.2MB, pdf)

References

  1. Vanadium Compounds: Chemistry, Biochemistry, and Therapeutic Applications; Tracey A. S., Crans D. C., Eds.; American Chemical Society, Washington, DC, 1998. [Google Scholar]
  2. Tracey A. S.; Willsky G. R.; Takeuchi E.. Vanadium chemistry, biochemistry, pharmacology and practical applications; CRC Press: Boca Raton, FL, 2007. [Google Scholar]
  3. Rehder D.Bioinorganic vanadium chemistry; John Wiley & Sons: Chichester, England; Hoboken, NJ, 2008. [Google Scholar]
  4. Fraústo da Silva J. J. R.; Williams R. J. P.. The biological chemistry of the elements: the inorganic chemistry of life; Oxford University Press: Oxford, 2009. [Google Scholar]
  5. Sutradhar M.; Da Silva J. A. L.; Pombeiro A. J. L.. Introduction: Vanadium, Its Compounds and Applications. In Vanadium Catalysis; Sutradhar M., Pombeiro A. J. L., da Silva J. A. L., Eds.; Royal Society of Chemistry: London, 2021; Chapter 1, pp 1–11. [Google Scholar]
  6. Rehder D. Implications of vanadium in technical applications and pharmaceutical issues. Inorg. Chim. Acta 2017, 455, 378–389. 10.1016/j.ica.2016.06.021. [DOI] [Google Scholar]
  7. Shul’pin G. B.; Kozlov Y. N.; Shul’pina L. S. Metal complexes containing redox-active ligands in oxidation of hydrocarbons and alcohols: A review. Catalysts 2019, 9 (12), 1046. 10.3390/catal9121046. [DOI] [Google Scholar]
  8. Dawood K. M.; Nomura K. Recent Developments in Z-Selective Olefin Metathesis Reactions by Molybdenum, Tungsten, Ruthenium, and Vanadium Catalysts. Adv. Synth. Catal. 2021, 363 (8), 1970–1997. 10.1002/adsc.202001117. [DOI] [Google Scholar]
  9. Aureliano M.; Gumerova N. I.; Sciortino G.; Garribba E.; Rompel A.; Crans D. C. Polyoxovanadates with emerging biomedical activities. Coord. Chem. Rev. 2021, 447, 214143. 10.1016/j.ccr.2021.214143. [DOI] [Google Scholar]
  10. Rehder D. The potentiality of vanadium in medicinal applications. Inorg. Chim. Acta 2020, 504, 119445. 10.1016/j.ica.2020.119445. [DOI] [PubMed] [Google Scholar]
  11. Selvaraj S.; Krishnan U. M. Vanadium-Flavonoid Complexes: A Promising Class of Molecules for Therapeutic Applications. J. Med. Chem. 2021, 64 (17), 12435–12452. 10.1021/acs.jmedchem.1c00405. [DOI] [PubMed] [Google Scholar]
  12. Leblanc C.; Vilter H.; Fournier J. B.; Delage L.; Potin P.; Rebuffet E.; Michel G.; Solari P. L.; Feiters M. C.; Czjzek M. Vanadium haloperoxidases: From the discovery 30 years ago to X-ray crystallographic and V K-edge absorption spectroscopic studies. Coord. Chem. Rev. 2015, 301–302, 134–146. 10.1016/j.ccr.2015.02.013. [DOI] [Google Scholar]
  13. Sippel D.; Einsle O. The structure of vanadium nitrogenase reveals an unusual bridging ligand. Nat. Chem. Biol. 2017, 13 (9), 956–960. 10.1038/nchembio.2428. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Sippel D.; Rohde M.; Netzer J.; Trncik C.; Gies J.; Grunau K.; Djurdjevic I.; Decamps L.; Andrade S. L. A.; Einsle O. A bound reaction intermediate sheds light on the mechanism of nitrogenase. Science 2018, 359 (6383), 1484–1489. 10.1126/science.aar2765. [DOI] [PubMed] [Google Scholar]
  15. Pan H. R.; Hsu H. F.. Vanadium Catalysis Relevant to Nitrogenase. In Vanadium Catalysis; Sutradhar M., Pombeiro A. J. L., da Silva J. A. L., Eds.; Royal Society of Chemistry: London, 2021; Chapter 25, pp 564–576. [Google Scholar]
  16. Kostova I. Titanium and vanadium complexes as anticancer agents. Anticancer Agents Med. Chem. 2009, 9 (8), 827–42. 10.2174/187152009789124646. [DOI] [PubMed] [Google Scholar]
  17. Lampronti I.; Bianchi N.; Borgatti M.; Fabbri E.; Vizziello L.; Khan M. T.; Ather A.; Brezena D.; Tahir M. M.; Gambari R. Effects of vanadium complexes on cell growth of human leukemia cells and protein-DNA interactions. Oncol. Rep. 2005, 14 (1), 9–15. [PubMed] [Google Scholar]
  18. Thompson K. H.; McNeill J. H.; Orvig C. Vanadium Compounds as Insulin Mimics. Chem. Rev. 1999, 99 (9), 2561–2572. 10.1021/cr980427c. [DOI] [PubMed] [Google Scholar]
  19. Amante C.; De Sousa-Coelho A. L.; Aureliano M. Vanadium and Melanoma: A Systematic Review. Metals 2021, 11 (5), 828. 10.3390/met11050828. [DOI] [Google Scholar]
  20. Jakusch T.; Kiss T. In vitro study of the antidiabetic behavior of vanadium compounds. Coord. Chem. Rev. 2017, 351, 118–126. 10.1016/j.ccr.2017.04.007. [DOI] [Google Scholar]
  21. Kowalski S.; Wyrzykowski D.; Inkielewicz-Stepniak I. Molecular and Cellular Mechanisms of Cytotoxic Activity of Vanadium Compounds against Cancer Cells. Molecules 2020, 25 (7), 1757. 10.3390/molecules25071757. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Pessoa J. C.; Correia I. Misinterpretations in Evaluating Interactions of Vanadium Complexes with Proteins and Other Biological Targets. Inorganics 2021, 9 (2), 17. 10.3390/inorganics9020017. [DOI] [Google Scholar]
  23. Scibior A.; Pietrzyk L.; Plewa Z.; Skiba A. Vanadium: Risks and possible benefits in the light of a comprehensive overview of its pharmacotoxicological mechanisms and multi-applications with a summary of further research trends. J. Trace Elem. Med. Biol. 2020, 61, 126508. 10.1016/j.jtemb.2020.126508. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Trevino S.; Diaz A.; Sanchez-Lara E.; Sanchez-Gaytan B. L.; Perez-Aguilar J. M.; Gonzalez-Vergara E. Vanadium in Biological Action: Chemical, Pharmacological Aspects, and Metabolic Implications in Diabetes Mellitus. Biol. Trace Elem. Res. 2019, 188 (1), 68–98. 10.1007/s12011-018-1540-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Crans D.; Henry L.; Cardiff G.; Posner B.. Developing Vanadium as an Antidiabetic or Anticancer Drug: A Clinical and Historical Perspective. In Medicine: Therapeutic Use and Toxicity of Metal Ions in the Clinic; Carver P. L., Ed.; De Gruyter GmbH: Berlin, 2019; Vol. 19, pp 203–230. [DOI] [PubMed] [Google Scholar]
  26. Crans D. C.; Yang L.; Haase A.; Yang X.. Health Benefits of Vanadium and Its Potential as an Anticancer Agent. In Metallo-Drugs: Development and Action of Anticancer Agents; Sigel A., Sigel H., Freisinger E., Sigel R. K. O., Eds.; De Gruyter GmbH: Berlin, 2018; Vol. 18. [DOI] [PubMed] [Google Scholar]
  27. Bouwman E.; Reedijk J.. Bioinorganic Catalysis, 2nd ed.; Marcel Dekker, Inc.: New York, 1999. [Google Scholar]
  28. Eom H.; Song W. J. Emergence of metal selectivity and promiscuity in metalloenzymes. J. Biol. Inorg. Chem. 2019, 24 (4), 517–531. 10.1007/s00775-019-01667-0. [DOI] [PubMed] [Google Scholar]
  29. McLauchlan C. C.; Murakami H. A.; Wallace C. A.; Crans D. C. Coordination environment changes of the vanadium in vanadium-dependent haloperoxidase enzymes. J. Inorg. Biochem. 2018, 186, 267–279. 10.1016/j.jinorgbio.2018.06.011. [DOI] [PubMed] [Google Scholar]
  30. Chen Z. Recent development of biomimetic halogenation inspired by vanadium dependent haloperoxidase. Coord. Chem. Rev. 2022, 457, 214404. 10.1016/j.ccr.2021.214404. [DOI] [Google Scholar]
  31. Costa Pessoa J.; Garribba E.; Santos M. F. A.; Santos-Silva T. Vanadium and proteins: Uptake, transport, structure, activity and function. Coord. Chem. Rev. 2015, 301–302, 49–86. 10.1016/j.ccr.2015.03.016. [DOI] [Google Scholar]
  32. Mubarak M. Q. E.; de Visser S. P. Second-Coordination Sphere Effect on the Reactivity of Vanadium–Peroxo Complexes: A Computational Study. Inorg. Chem. 2019, 58 (23), 15741–15750. 10.1021/acs.inorgchem.9b01778. [DOI] [PubMed] [Google Scholar]
  33. Rehder D.Vanadium in Catalytically Proceeding Natural Processes. In Vanadium Catalysis; Sutradhar M., Pombeiro A. J. L., da Silva J. A. L., Eds.; The Royal Society of Chemistry: 2021; Chapter 23, pp 535–547. [Google Scholar]
  34. 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 (13), 3688–3689. 10.1021/ja029271v. [DOI] [PubMed] [Google Scholar]
  35. Conte V.; Coletti A.; Floris B.; Licini G.; Zonta C. Mechanistic aspects of vanadium catalysed oxidations with peroxides. Coord. Chem. Rev. 2011, 255 (19), 2165–2177. 10.1016/j.ccr.2011.03.006. [DOI] [Google Scholar]
  36. da Silva J. A.L.; da Silva J. J.R. F.; Pombeiro A. J.L. Oxovanadium complexes in catalytic oxidations. Coord. Chem. Rev. 2011, 255 (19-20), 2232–2248. 10.1016/j.ccr.2011.05.009. [DOI] [Google Scholar]
  37. Hirao T. Vanadium in Modern Organic Synthesis. Chem. Rev. 1997, 97 (8), 2707–2724. 10.1021/cr960014g. [DOI] [PubMed] [Google Scholar]
  38. Kirihara M. Aerobic oxidation of organic compounds catalyzed by vanadium compounds. Coord. Chem. Rev. 2011, 255 (19), 2281–2302. 10.1016/j.ccr.2011.04.001. [DOI] [Google Scholar]
  39. 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 (4), 2128–2191. 10.1021/acs.chemrev.8b00245. [DOI] [PubMed] [Google Scholar]
  40. Ligtenbarg A. G. J.; Hage R.; Feringa B. L. Catalytic oxidations by vanadium complexes. Coord. Chem. Rev. 2003, 237 (1), 89–101. 10.1016/S0010-8545(02)00308-9. [DOI] [Google Scholar]
  41. 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 (1), 53–63. 10.1016/S0010-8545(02)00300-4. [DOI] [Google Scholar]
  42. Beller M. B. C.Transition metals for organic synthesis: building blocks and fine chemicals; Wiley-VCH: Weinheim; New York, 1998. [Google Scholar]
  43. Shilov A. E.; Shul’pin G. B. Activation of C–H Bonds by Metal Complexes. Chem. Rev. 1997, 97 (8), 2879–2932. 10.1021/cr9411886. [DOI] [PubMed] [Google Scholar]
  44. Shul’pin G. B. Metal-catalyzed hydrocarbon oxygenations in solutions: the dramatic role of additives: a review. J. Mol. Catal. A: Chem. 2002, 189 (1), 39–66. 10.1016/S1381-1169(02)00196-6. [DOI] [Google Scholar]
  45. Shul’pin B. G. Hydrocarbon Oxygenations with Peroxides Catalyzed by Metal Compounds. Mini-Reviews in Organic Chemistry 2009, 6 (2), 95–104. 10.2174/157019309788167738. [DOI] [Google Scholar]
  46. England J.; Britovsek G. J.; Rabadia N.; White A. J. Ligand topology variations and the importance of ligand field strength in non-heme iron catalyzed oxidations of alkanes. Inorg. Chem. 2007, 46 (9), 3752–67. 10.1021/ic070062r. [DOI] [PubMed] [Google Scholar]
  47. Sax N. I.; Lewis R. J.. Dangerous properties of industrial materials; Van Nostrand Reinhold: New York, 1989; Vol. 3. [Google Scholar]
  48. Fomenko I. S.; Gushchin A. L.; Abramov P. A.; Sokolov M. N.; Shulpina L. S.; Ikonnikov N. S.; Kuznetsov M. L.; Pombeiro A. J. L.; Kozlov Y. N.; Shul’pin G. B. New Oxidovanadium(IV) Complexes with 2,2′-bipyridine and 1,10-phenathroline Ligands: Synthesis, Structure and High Catalytic Activity in Oxidations of Alkanes and Alcohols with Peroxides. Catalysts 2019, 9 (3), 217. 10.3390/catal9030217. [DOI] [Google Scholar]
  49. Homden D. M.; Redshaw C.; Hughes D. L. Vanadium Complexes Possessing N2O2S2-Based Ligands: Highly Active Procatalysts for the Homopolymerization of Ethylene and Copolymerization of Ethylene/1-hexene. Inorg. Chem. 2007, 46 (25), 10827–10839. 10.1021/ic701461b. [DOI] [PubMed] [Google Scholar]
  50. Ma J.; Zhao K.-Q.; Walton M.; Wright J. A.; Hughes D. L.; Elsegood M. R. J.; Michiue K.; Sun X.; Redshaw C. Tri- and tetra-dentate imine vanadyl complexes: synthesis, structure and ethylene polymerization/ring opening polymerization capability. Dalton Trans 2014, 43 (44), 16698–16706. 10.1039/C4DT01448K. [DOI] [PubMed] [Google Scholar]
  51. Wen D.; Zhou J.; Zou H.-H. A series of new oxo-vanadium(IV) complexes with octahedral coordinated vanadium centers. J. Coord. Chem. 2019, 72 (5–7), 1064–1074. 10.1080/00958972.2019.1596264. [DOI] [Google Scholar]
  52. Xu S.-y.; Chen X.-m.; Huang L.-c.; Li F.; Gao W. Vanadium chlorides supported by BIAN (BIAN = bis(arylimo)-acenaphthene) ligands: Synthesis, characterization, and catalysis on ethylene polymerization. Polyhedron 2019, 164, 146–151. 10.1016/j.poly.2019.01.041. [DOI] [Google Scholar]
  53. Fomenko I. S.; Vincendeau S.; Manoury E.; Poli R.; Abramov P. A.; Nadolinny V. A.; Sokolov M. N.; Gushchin A. L. An oxidovanadium(IV) complex with 4,4′-di-tert-butyl-2,2′-bipyridine ligand: Synthesis, structure and catalyzed cyclooctene epoxidation. Polyhedron 2020, 177, 114305. 10.1016/j.poly.2019.114305. [DOI] [Google Scholar]
  54. Chang Y.-P.; Furness L.; Levason W.; Reid G.; Zhang W. Complexes of vanadium(IV) oxide difluoride with neutral N- and O-donor ligands. J. Fluorine Chem. 2016, 191, 149–160. 10.1016/j.jfluchem.2016.09.020. [DOI] [Google Scholar]
  55. Passadis S. S.; Tsiafoulis C.; Drouza C.; Tsipis A. C.; Miras H. N.; Keramidas A. D.; Kabanos T. A. Synthesis, Bonding, and Reactivity of Vanadium(IV) Oxido-Fluorido Compounds with Neutral Chelate Ligands of the General Formula cis-[VIV(=O)(F)(LN-N)2]+. Inorg. Chem. 2016, 55 (4), 1364–1366. 10.1021/acs.inorgchem.5b02895. [DOI] [PubMed] [Google Scholar]
  56. Stasch A.; Schormann M.; Prust J.; Roesky H. W.; Schmidt H.-G.; Noltemeyer M. Acetylacetonatodifluorooxometalates of vanadium and molybdenum: syntheses and crystal structures. J. Chem. Soc., Dalton Trans. 2001, 13, 1945–1947. 10.1039/b100740h. [DOI] [Google Scholar]
  57. Stephens N. F.; Buck M.; Lightfoot P. Polar ordering of polar octahedra in [C2N2H10][VOF4(H2O)]. J. Mater. Chem. 2005, 15 (40), 4298–4300. 10.1039/b508959j. [DOI] [Google Scholar]
  58. Gupta S.; Kirillova M. V.; Guedes da Silva M. F.; Pombeiro A. J. L. Highly efficient divanadium(V) pre-catalyst for mild oxidation of liquid and gaseous alkanes. Appl. Catal. A-Gen. 2013, 460–461, 82–89. 10.1016/j.apcata.2013.03.034. [DOI] [Google Scholar]
  59. Sutradhar M.; Martins L. M. D. R. S.; Guedes da Silva M. F. C.; Pombeiro A. J. L. Vanadium complexes: Recent progress in oxidation catalysis. Coord. Chem. Rev. 2015, 301–302, 200–239. 10.1016/j.ccr.2015.01.020. [DOI] [Google Scholar]
  60. Gupta S.; Kirillova M. V.; Guedes da Silva M. F. C.; Pombeiro A. J. L.; Kirillov A. M. Alkali Metal Directed Assembly of Heterometallic Vv/M (M = Na, K, Cs) Coordination Polymers: Structures, Topological Analysis, and Oxidation Catalytic Properties. Inorg. Chem. 2013, 52 (15), 8601–8611. 10.1021/ic400743h. [DOI] [PubMed] [Google Scholar]
  61. Addison A. W.; Rao T. N.; Reedijk J.; van Rijn J.; Verschoor G. C. Synthesis, structure, and spectroscopic properties of copper(II) compounds containing nitrogen–sulphur donor ligands; the crystal and molecular structure of aqua[1,7-bis(N-methylbenzimidazol-2′-yl)-2,6-dithiaheptane]copper(II) perchlorate. J. Chem. Soc., Dalton Trans. 1984, 7, 1349–1356. 10.1039/DT9840001349. [DOI] [Google Scholar]
  62. Micera G.; Garribba E. On the prediction of 51V hyperfine coupling constants in VIVO complexes through DFT methods. Dalton Trans 2009, 11, 1914–1918. 10.1039/b819620f. [DOI] [PubMed] [Google Scholar]
  63. Chasteen D. N.Biological Magnetic Resonance; Reuben J., Ed.; Plenum Press: New York, 1981; Vol. 3, p 53. [Google Scholar]
  64. Smith T. S. Ii; LoBrutto R.; Pecoraro V. L. Paramagnetic spectroscopy of vanadyl complexes and its applications to biological systems. Coord. Chem. Rev. 2002, 228 (1), 1–18. 10.1016/S0010-8545(01)00437-4. [DOI] [Google Scholar]
  65. Smith T. S. Ii; Root C. A.; Kampf J. W.; Rasmussen P. G.; Pecoraro V. L. Reevaluation of the additivity relationship for vanadyl-imidazole complexes: Correlation of the EPR hyperfine constant with ring orientation. J. Am. Chem. Soc. 2000, 122 (5), 767–775. 10.1021/ja991552o. [DOI] [Google Scholar]
  66. Gorelsky S.; Micera G.; Garribba E. The Equilibrium Between the Octahedral and Square Pyramidal Form and the Influence of an Axial Ligand on the Molecular Properties of VIVO Complexes: A Spectroscopic and DFT Study. Chem.—Eur. J. 2010, 16 (27), 8167–8180. 10.1002/chem.201000679. [DOI] [PubMed] [Google Scholar]
  67. Tolis E. J.; Soulti K. D.; Raptopoulou C. P.; Terzis A.; Deligiannakis Y.; Kabanos T. A. Structural, electron paramagnetic resonance and electron spin echo envelope modulation studies of oxovanadium()–amidate compounds containing monoanionic axial ligands: effect on the V-hyperfine coupling constants. Chem. Commun. 2000, 7, 601–602. 10.1039/b000399i. [DOI] [Google Scholar]
  68. Crans D. C.; Keramidas A. D.; Amin S. S.; Anderson O. P.; Miller S. M. Six-co-ordinated vanadium-(IV) and -(V) complexes of benzimidazole and pyridyl containing ligands. J. Chem. Soc., Dalton Trans. 1997, 16, 2799–2812. 10.1039/a700710h. [DOI] [Google Scholar]
  69. Lodyga-Chruscinska E.; Micera G.; Garribba E. Complex formation in aqueous solution and in the solid state of the potent insulin-enhancing V(IV)O2+ compounds formed by picolinate and quinolinate derivatives. Inorg. Chem. 2011, 50 (3), 883–899. 10.1021/ic101475x. [DOI] [PubMed] [Google Scholar]
  70. Mukherjee T.; Costa Pessoa J.; Kumar A.; Sarkar A. R. Oxidovanadium(IV) schiff base complex derived from vitamin B6: Synthesis, characterization, and insulin enhancing properties. Inorg. Chem. 2011, 50 (10), 4349–4361. 10.1021/ic102412s. [DOI] [PubMed] [Google Scholar]
  71. Nicolaou M.; Drouza C.; Keramidas A. D. Controlled one pot synthesis of polyoxofluorovanadate molecular hybrids exhibiting peroxidase like activity. New J. Chem. 2019, 43 (45), 17595–17602. 10.1039/C9NJ01999E. [DOI] [Google Scholar]
  72. Mao G. D.; Thomas P. D.; Poznansky M. J. Oxidation of spin trap 5,5-dimethyl-1-pyrroline-1-oxide in an electron paramagnetic resonance study of the reaction of methemoglobin with hydrogen peroxide. Free Radic Biol. Med. 1994, 16 (4), 493–500. 10.1016/0891-5849(94)90127-9. [DOI] [PubMed] [Google Scholar]
  73. Rosen G. M.; Rauckman E. J. Spin trapping of the primary radical involved in the activation of the carcinogen N-hydroxy-2-acetylaminofluorene by cumene hydroperoxide-hematin. Mol. Pharmacol. 1980, 17 (2), 233–238. [PubMed] [Google Scholar]
  74. Stan S. D.; Daeschel M. A. 5,5-Dimethyl-2-pyrrolidone-N-oxyl Formation in Electron Spin Resonance Studies of Electrolyzed NaCl Solution Using 5,5-Dimethyl-1-pyrroline-N-oxide as a Spin Trapping Agent. J. Agric. Food. Chem. 2005, 53 (12), 4906–4910. 10.1021/jf047918k. [DOI] [PubMed] [Google Scholar]
  75. Sanz J.; Lombraña J. I.; De Luis A. M.; Ortueta M.; Varona F. Microwave and Fenton’s reagent oxidation of wastewater. Environ. Chem. Lett. 2003, 1 (1), 45–50. 10.1007/s10311-002-0007-2. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ic2c02526_si_001.pdf (1.2MB, pdf)

Articles from Inorganic Chemistry are provided here courtesy of American Chemical Society

RESOURCES