Next Article in Journal
A Fine Analysis of Zn Species Structure and Distribution in Zn/ZSM-5 Catalysts by Linear Combination Fitting Analysis of XANES Spectra
Previous Article in Journal
Cytotoxicity Enhancement in Osteosarcoma with Multifunctional I-131 Radiotherapeutic Nanoparticles: In Vitro Three-Dimensional Spheroid Model and Release Kinetics Modeling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Primary Activity Assay of Novel Benitrobenrazide and Benserazide Derivatives

by
Karolina Juszczak
,
Wojciech Szczepankiewicz
and
Krzysztof Walczak
*
Department of Organic Chemistry, Bioorganic Chemistry and Biotechnology, Faculty of Chemistry, Silesian University of Technology, 44-100 Gliwice, Poland
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(3), 629; https://doi.org/10.3390/molecules29030629
Submission received: 2 January 2024 / Revised: 17 January 2024 / Accepted: 19 January 2024 / Published: 29 January 2024

Abstract

:
Schiff bases attract research interest due to their applications in chemical synthesis and medicinal chemistry. In recent years, benitrobenrazide and benserazide containing imine moiety have been synthesized and characterized as promising inhibitors of hexokinase 2 (HK2), an enzyme overexpressed in most cancer cells. Benserazide and benitrobenrazide possess a common structural fragment, a 2,3,4-trihydroxybenzaldehyde moiety connected through a hydrazone or hydrazine linker acylated on an N′ nitrogen atom by serine or a 4-nitrobenzoic acid fragment. To avoid the presence of a toxicophoric nitro group in the benitrobenrazide molecule, we introduced common pharmacophores such as 4-fluorophenyl or 4-aminophenyl substituents. Modification of benserazide requires the introduction of other endogenous amino acids instead of serine. Herein, we report the synthesis of benitrobenrazide and benserazide analogues and preliminary results of inhibitory activity against HK2 evoked by these structural changes. The derivatives contain a fluorine atom or amino group instead of a nitro group in BNB and exhibit the most potent inhibitory effects against HK2 at a concentration of 1 µM, with HK2 inhibition rates of 60% and 54%, respectively.

Graphical Abstract

1. Introduction

Carbonyl compounds easily react with N-centered nucleophiles, such as amines, hydrazine, and its derivatives. Schiff bases, a product of condensation of carbonyl compounds with primary amines, have a wide range of applications in pharmaceuticals [1,2,3] and in coordination chemistry as ligands and chelating agents [4,5]. The versatile pharmacophore C=N is present in biologically active compounds exhibiting antioxidant [3,6], antimicrobial [3,6,7], and anticancer [8] properties.
N-Acylhydrazides R(CO)NHNH2 are structurally like amides and a relocation of a proton can occur, thus forming an iminol form (Figure 1) [1]. Compounds containing an amide–imine bridge -C(=O)-NH-N=CH- can be considered hybrid structures of hydrazide or hydrazones [1]. These hybrid compounds have several applications in pharmacology, exhibiting antimicrobial activity [9,10,11], anticancer activity [8,9], and free radical scavenging properties [12].
Benserazide (BND) and benitrobenrazide (BNB) (Figure 2) are compounds belonging to the group of hydrazine derivatives containing common structural fragments, including a N-acyl rest of serine or 4-nitrobenzoic acid and a 2,3,4-trihydroxybenzaldehyde (pyrogallol-4-carboxaldehyde) fragment attached to an N′ nitrogen atom by a methylene or methine carbon atom. Both compounds BNB and BND have promising inhibitory activity against hexokinase 2 (HK2) [13,14,15], an enzyme involved in the phosphorylation of glucose in glycolysis, a first step of glucose metabolism [16]. An increased requirement for glucose is observed in cancer cells, especially in rapidly growing and drug-resistant tumors [17,18,19,20,21,22]. Among all human hexokinase isoenzymes, HK2 shows overexpression in cancer cells, which makes it interesting in the context of molecularly targeted therapy [23,24,25,26,27,28]. Currently, several HK2 potent inhibitors are recognized, like metformin, 2-deoxy-D-Glucose, or 3-bromopyruvate [24,29,30,31,32].
Benserazide is an FDA-approved drug for the treatment of Parkinson’s disease, but it was recently also recognized as a strong HK2 inhibitor [14]. Benitrobenrazide was identified as a potential HK2 inhibitor by structure-based virtual ligand screening [15]. According to our previous study, the key structural feature is the presence of three hydroxy groups in a benzene ring, which occupies the same HK2-binding pocket as the natural substrate, glucose [29]. The influence of other structural elements, namely -NH-NH-CH2- and -NH=N-CH-, on HK2 enzymatic activity has not yet been determined and requires investigation to explain their impact on enzymatic activity.
Herein, we present our primary attempts to explore the influence of structural features observed in benserazide and benitrobenrazide in their biological activity against HK2. The serine originally present in benserazide has been exchanged with other amino acid fragments, e.g., glycine, tyrosine, and cysteine. Application of these amino acids that are different in structure, including glycine (achiral), threonine-containing benzene ring, and cysteine (containing a thiol group instead of a hydroxyl one), in the synthesis of benserazide analogues should deliver an answer regarding the importance of serine moiety on benserazide inhibitory activity. In the BNB molecule, the 4-nitrophenyl moiety has been substituted by an alkyl chain or aromatic rings of different molecular areas (benzene, naphthalene, and anthracene) to judge their downstream influence on HK2 activity and gauge HK2 active site volume. We decided to exchange the toxicophoric nitro group present in BNB moiety by two common substituents, namely the fluorine atom and the amino group [33,34]. These substituents have the opposite effects on electron density in the benzene ring of BNB and were introduced to explain the eventual influence of electronic conditions on its activity. The fluorine atom, like the nitro group, decreases the electron density, whereas the amino group is typical of the electron-donating group. The novel analogues of benserazide and benitrobenrazide were examined for their inhibitory activity against HK2 to obtain information about the influence of structural variations on inhibition activity.

2. Results and Discussion

2.1. Chemistry of Benitrobenrazide and Benserazide Derivatives

Benitrobenrazide and benserazide contain a N-acyl fragment, 4-nitrophenyl, or a serine moiety, respectively [13,15]. We modified these regions of each molecule by incorporating aromatic rings of different molecular areas instead to probe the volume of the active site pocket of HK2. In further modifications, a nitro group suspected as a genotoxic substituent was replaced by substituents of different polarity, constituting either a fluorine atom or an amino group. The introduction of fluorine and amino substituent as pharmacophore groups into drugs molecules increases their therapeutic effectiveness [35,36]. Additionally, the imine bond in hydrazone analogue was reduced to a single carbon–nitrogen bond to identify the effect of the imine double bond (C=N) on HK2 activity.
The final benitrobenrazide analogue synthesis is depicted below (Scheme 1). Commercially available methyl propionate and benzoate 2a and 2b were used. Methyl esters of other aromatic acids 1c, 1e, and 1f were obtained in a one-pot synthesis involving the primary transformation of each appropriate carboxylic acid into its respective acyl chloride by treatment with thionyl chloride in excess, followed by esterification in the presence of methanol (MeOH) [37] (Scheme 1; Pathway A). The ester of anthracene-9-carboxylic acid 1d was synthesized by a two-step synthesis. The first step was conversion of anthracene-9-carboxylic acid into its acyl chloride in the presence of an excess of thionyl chloride and a small amount of dimethylformamide (DMF) as a catalyst. The separated acyl chloride was then subjected to esterification with ethanol (EtOH) in the presence of triethylamine, according to synthesis pathway B shown in Scheme 1.
Hydrazides 3a3f were obtained with a good yield from appropriate esters in nucleophilic substitution on a carbonyl carbon atom in the presence of an excess of hydrazine monohydrate (four equivalents) in anhydrous boiling methanol [38,39]. Conducting hydrazide synthesis 3b3c and 3f at room temperature led to a purer product and a good yield (60–77%). The preparation of hydrazide 3d needs harsher conditions, so the reaction was conducted in an excess of boiling hydrazine hydrate, with a prolonged reaction time (Scheme 1; B pathway). Hydrazides 3b3f were obtained in the form of white solids, which precipitated during the reaction process and crystallized from ethanol or aqueous ethanol. Compound 3a was purified via silica gel column chromatography using MeOH/CHCl3 (1:1 v/v) as an eluent. The hydrazide structure was confirmed by 1H NMR spectroscopy. The absence of a singlet assigned to the methyl group in the region of 3.25–3.95 ppm confirmed the conversion of the esters into hydrazides.
Compounds detailed in 3, treated with 2,3,4-trihydroxybenzaldehyde in methanol, produced the products listed in 4 at a 50–100% yield. Hydrazones 4 can adopt an E or a Z configuration at the imine double bond (-N=CH-). In the case of the E conformation, the geometrical isomer can be stabilized by the formation of an intramolecular hydrogen bond between the 2-OH hydroxyl group and the nitrogen atom of the imine [40], as is depicted in Figure 3. Similar behavior is exhibited by compound 10.
Using quantum chemical calculations based on Density Function Theory (DFT), we studied the possibility of the existence of equilibrium between conformers E and Z for 4a. For conformer E, the corresponding hydrogen bond length is 2.1 Å. Hydrogen bond formation confirms our DFT calculation of stretching vibrations for the O-H group at 3224.47 cm−1. Additionally, our calculation data show higher stability of the E conformer over the Z conformer by 9 kcal/mol when hydrogen bond formation is included in simulations [41]. We can assume that the E conformer is the predominant form of hydrazone 4a. 1H NMR spectroscopy verified the results of our calculation based on DFT. The chemical shifts of the protons of the 3-OH and 4-OH groups create singlet peaks at δ H 8.4–9.6 ppm, while the peak for the proton of the 2-OH group is shifted to a lower field and was observed in the region of δ H 11.3–12.4 ppm. We attribute this down-field shift of the 2-OH group to the formation of an intramolecular hydrogen bond.
Based on 1H and 13C NMR spectra of 4a in DMSO, we observed separate chemical signals for the imine proton (N=CH) at 8.09 and 8.17 ppm, respectively. In contrast, the protons of the ethyl group (CH2CH3) were present as two triplets in the region of 1.05–1.09 ppm and two quartets in the region of 2.19–2.23 and 2.51–2.54. In the 13C NMR spectrum, the carbon atom of the amide carbonyl groups (-C(O)NHN-) was detected from signals at 173.71 and 168.73 ppm. The signals for carbon from the ethyl groups (CH2CH3) were present at 8.47, 9.41 ppm, 25.24 and 26.97. These spectra suggest the existence of another structural feature for compounds 4, namely a keto-enol tautomerism (Scheme 2).
Quantum chemical calculations using DFT were used to quantify Gibbs free energy for the constitutional isomers of 4a, which confirms the possibility of the keto and enol forms’ occurrence. According to the calculations performed on data collected without a solvent, the keto tautomer is preferred over the enol tautomer and is thermodynamically more stable than the enol form by about 11 kcal/mol. However, based on calculations performed in the presence of DMSO, the solvation process changed the Gibbs free energy of both tautomers. Our calculations clearly indicate that, in a polar aprotic solvent like DMSO, the enol form is more stable by 5 kcal/mol when compared with the keto form due to the interaction of the OH-enol group with the oxygen atom of dimethyl sulfoxide (DMSO).
The imine bond in derivative 4f was reduced using hydrogen in the presence of Pd(OH)2 as a catalyst at elevated pressure (Scheme 3). The reaction progress was monitored by 1HNMR spectrometry. The disappearance of the signal for the 8.39 ppm imine proton region (N=CH) indicated the consumption of substrate 4f.
Continuing our study, we synthesized benserazide analogues, in which serine was replaced with a side chain moiety of another L-amino acid glycine, tyrosine, and cysteine (Scheme 4).
Commercially available L-amino acids 6ac were converted to methyl esters 7ac, using thionyl chloride and methanol. A benzyloxycarbonyl group (Cbz) was used to protect the amino group of the amino acids by reacting esters 7 with benzyl chloroformate in the presence of triethylamine (TEA) in an anhydrous methylene chloride solution [42]. This method of protection was used to facilitate the parallel deprotection of Cbz together with catalytic hydrogenation of the double bond. The reaction of N-Cbz-L-amino acid esters 8ac with 98% hydrazine hydrate under the same conditions described in Scheme 4 provided products 9 as white crystals at a 36–97% yield. Hydrazones 10a and 10c were synthesized via the reaction of hydrazide 9a and 9c with 2,3,4-trihydroxybenzaldehyde in methanol at room temperature. In the case of 10b, as the above procedure failed, we repeated the condensation in tetrahydrofuran (THF) instead of methanol in an inert atmosphere of argon at reflux, which produced product 10b at a 55% yield. All hydrazones 10ac were obtained with the E configuration, as confirmed by NMR spectroscopy. The best purification method for compounds 10 was crystallization with aqueous methanol (1:1 v/v). The last step in this synthesis was the elucidation of an efficient method for the reduction of the imine bond of hydrazone and deprotection of the Cbz group. For the corresponding hydrazone 10b, hydrogenation was performed in a Parr’s autoclave with gaseous H2 at 2.5 bar in the presence of palladium catalysts (mixture of 25% Pd/C and Pd(OH)2) at room temperature in methanolic solution. We observed only cleavage of the Cbz group.
Compound 10a was reduced under modified conditions. Instead of gaseous hydrogen, ammonium formate was used as a hydrogen donor and the same palladium catalysts were used, according to the reported protocol [43]. Product 11 was obtained as a hygroscopic powder, which rapidly liquefied after 5 min on air. For this reason, hydrazide 11 was lyophilized after the purification was complete. In the case of the cysteine analogue of benserazide 10c, hydrogenation did not occur under these conditions, or with the use of other catalysts such as Raney nickel and electrochemical reduction. The likely reason is that a compound containing sulfur in its structure causes catalyst poisoning, resulting in the loss of catalyst function [44].

2.2. Inhibitory Effect of Benserazide and Benitrobenrazide Derivatives on HK2 Enzymatic Activity

We conducted in vitro studies of representative derivatives, namely N-acylhydrazones 4a4f as benitrobenrazide derivatives and a benitrobenrazide derivative with a single carbon-nitrogen bond 5. The benserazide derivatives, hydrazones 10, 12 with imine bonds, and hydrazide 11 with a single carbon–nitrogen bond, which, like benserazide, have an L-amino acid fragment (Figure 4), were selected for in vitro experiments. The selected intermediates of hydrazide 9ac, the potential peptidomimetics, were also used to evaluate any inhibition of enzyme activity (Figure 4). The chosen compounds subjected to this enzyme activity assay were purified by preparative HPLC.
For the inhibitory activity estimation, we used a colorimetric method. The most popular HK2 enzymatic activity assays reported in the literature are tests based on spectroscopy, measuring/detecting the changes in NADPH absorbance at 320–490 nm [13,14,15,31]. An alternative assay to determine the HK2 activity by reverse-phase high-performance liquid chromatography (RP-HPLC) was reported by Guan et al. [45]. According to this method, the concentration of released ADP during the conversion of glucose into 6-glucose phosphate is measured at 254 nm. We excluded RP-HPLC as an alternative assay because tested compounds contain chromophores exhibiting absorption in UV-Vis in a region of 200–250 nm, which renders the assay results unreliable. To evaluate the potential of hexokinase 2 inhibitors, we used a commercially available Hexokinase II Inhibitor Screening Kit, which uses a spectrophotometric method by measurement of absorbance at 450 nm, where tested compounds exhibit no interference with the assay. This HK2 activity assay is based on HK2’s ability to convert glucose into glucose-6-phosphate. Glucose-6-phosphate is oxidized by glucose-6-phosphate dehydrogenase to produce NADPH, which reduces the probe, showing strong absorbance at 450 nm.
Figure 5 and Figure 6 show the results of the HK2 activity inhibition assay of synthesized compounds. Primary measurements were performed at 50 and 5 µM concentrations. We have decided that the concentration of 50 µM is the highest concentration accepted for the investigated compounds as inhibitors. Compounds 4a, 4b, 4c, 4e, and 4f, which showed the best inhibition of HK2 activity, were selected for the next assessment at a lower concentration of 1 µM. The data in Figure 5 illustrate the inhibition of HK2-mediated phosphorylation of glucose by the test compounds through changes in absorption detection at λmax 450 nm observed in kinetic mode after 5–45 min.
As shown in Figure 6, at a concentration of 50 µM, most of the compounds show significant HK2 inhibition activity, apart from hydrazides 9, which show moderate HK2 inhibition. According to the in vitro study, which considered the effect of HK2 inhibition by the tested compounds at a concentration of 5 µM, derivatives 4a4f exhibited stronger HK2 inhibition activity than the benserazide derivatives 10c, 11, and 12, which have a modified L-amino acid fragment in their structure. The maximum inhibition by benserazide derivatives 10c, 11, and 12 was approximately 50%. Hydrazides 9 have no significant effect on HK2 activity at a concentration of 5 µM, which confirms that the presence of three hydroxyl groups is required for the inhibition of HK2 activity.
A comparison of the inhibition results for derivatives 4f and 5, for which HK2 inhibition was 92% and 15%, respectively, clearly indicates that the essential feature causing the HK2 inhibitory effect of benitrobenrazide derivatives is the presence of the imine bond (-CH=N-). Referring to studies on the pocket size of the HK2 binding site, it can be assumed that a bulky group, like the anthracenyl group in the structure of the Schiff base 4d, does not fit efficiently into the active site of HK2. Compared to other Schiff bases with smaller volume substituents 4ac, compound 4d did not exhibit inhibitory activity against HK2 at a 5 µM concentration.
The most promising HK2 inhibitors assayed, 4a, 4b, 4c, 4e, and 4f, were evaluated against human HK2 at a concentration of 1 µM. The derivatives 4e and 4f, which contain a fluorine atom or amino group instead of a nitro group, exhibited the most potent inhibitory effects against HK2 at a concentration of 1 µM, with HK2 inhibition rates of 60% and 54%, respectively. The highest efficacy among the derivatives assessed against HK2 was recorded for Schiff base 4e, in which the fluorine atom acts as an electron-withdrawing substituent in the para position of the benzene ring.

3. Materials and Methods

3.1. Chemistry

The 1H NMR and 13C NMR spectra were recorded using a Varian NMR system 600 spectrometer at 600 MHz in DMSO-d6, with tetramethylsilane (TMS) as the reference standard. NMR chemical shifts are reported in ppm (δ) and coupling constants (J) in Hz. Melting points were measured on a Boethius PHMK apparatus (VEB Analytik Dresden, Dresden, Germany). The progress of the reaction was monitored by thin-layer chromatography (TLC) using Merck TLC silica gel 60 F254 plates and the following developing systems: A: 5% MeOH/CHCl3, B: 10% MeOH/CHCl3, and C: 20% MeOH/CHCl3. Column chromatography was conducted using silica gel 40–60 μm 60A with methanol–chloroform mixtures as eluents. Preparative high-performance liquid chromatography was performed with the LaboACE LC-5060 system (Japan Analytical Industry Co., Ltd., Tokyo, Japan), with an ODS column (JAIGEL-ODS-AP, model SP-12-10, Japan Analytical Industry, Co., Ltd. Tokyo, Japan). The compounds were eluted with a mobile phase of MeOH at a flow rate of 9 mL/min. High-resolution mass spectroscopy (HRMS) was measured on a Waters Corporation Xevo G2 QTOF apparatus (Waters Corporation, Milford, MA, USA) using electrospray ionization (ESI).

3.1.1. General Procedure for the Synthesis of Esters 2ef

Methyl esters 2a and 2b were purchased from Merck. Methyl esters 2c and 2e2f were obtained by the well-known method for esterification of acids using thionyl chloride in methanol [37]. To a stirred and cooled (0 °C) solution of the required acid 1c, 1e1f (20 mmol) in anhydrous methanol (30 mL), thionyl chloride (1.10 equiv.) was added dropwise while stirring. The mixture was warmed up to room temperature and stirred for 24 h. After that, the excess of methanol was removed under diminished pressure and then dried under reduced pressure. In the case of 2f synthesis, after completion of the esterification reaction, the solution was neutralized by adding saturated aqueous NaHCO3 solution until no further gas evolution was observed. Solid 2f was filtered under reduced pressure.
Synthesis of ethyl anthracene-9-carboxylate 2d: to SOCl2 (7 mL), we added anthracene-9-carboxylic acid (1.80 mmol) and DMF (0.4 mL) as a catalyst. The reaction mixture was stirred at room temperature under argon atmosphere for 4h. The excess of SOCl2 was removed under reduced pressure. The residue was washed with toluene. To obtain orange solid acid chloride, EtOH (10 mL) and TEA (1.3 equiv.) were added sequentially. The reaction mixture was stirred at room temperature for 24 h. EtOH was removed under reduced pressure; then, chloroform (10 mL) was added to the residue and the organic layer was washed with water (2 × 8 mL) and dried over anhydrous Na2SO4. After filtration and evaporation under reduced pressure, the residue was purified on a silica gel packed column using (AcOEt:n-hexane 1:1, v/v) as an eluent, obtaining solid 2d at a yield of 90%.

3.1.2. General Procedure for the Synthesis of Hydrazides 3a3f

Hydrazides 3a3f were synthesized from their corresponding esters 2a2f, followed by reaction with hydrazine according to the method reported by Khan et al. [10].
Synthesis of hydrazides 3b3c and 3f: to a solution of methyl ester 2b, c, or 2f (20 mmol) in methanol (25 mL), hydrazine monohydrate (80 mmol, 4 equiv.) was added. The reaction mixture was stirred at room temperature for 72 h and then cooled to −20 °C. The formed precipitate was filtered off and dried under reduced pressure. Crystallization of crude solid from an ethanol:water (1:2, v/v) solution produced white crystals.
Benzohydrazide (3b): yield 77%. Mp 114–115 °C. (lit. Mp 110–113 °C [46]). TLC solvent system B; Rf (retention factor): 0.5. 1H NMR (600 MHz, DMSO-d6): δ 4.48 (s, NH2) 7.53–7.46 (m, 2H, C6H5), 7.49–7.51 (m, 1H, C6H5), 7.81–7.83 (m, 2H, C6H5), 9.76 (s, 1NH). 13C NMR (600 MHz, DMSO-d6): δ 132.08, 133.44, 136.18, 138.46, 171.02. HRMS (ESI-TOF): m/z calcd for (C7H8N2O + H+): 137.0709; found: 137.0724.
1-naphthohydrazide (3c): yield 69%. Mp 156–157 °C (lit. Mp 160–163 °C [38]). TLC system B; Rf: 0.6. 1H NMR (600 MHz, DMSO-d6): δ 4.59 (d, J = 3.53 Hz, NH2), 7.51–7.58 (m, 4H, C6H5), 7.96–7.98 (m, 1H, C6H5), 8.00–8.01 (m, 1H, C6H5), 8.20–8.22 (m, 1H, C6H5), 9.68 (s, 1NH) 13C NMR 150 MHz, DMSO-d6): 130.16, 130.51, 130.60, 131.40, 131.80, 133.35, 135.02, 135.18, 135.138.30, 138.55, 173.14. HRMS (ESI-TOF): m/z calcd for (C11H10N2O + H+): 187.0866; found: 187.0610.
4-aminobenzohydrazide (3f): yield 60%. Mp 224–226 °C (lit. Mp 225–227 °C [47]). TLC system C; Rf: 0.4. 1H NMR (600 MHz, DMSO-d6): δ 4.29 (s, NH2), 5.57 (s, C6H5NH2), 6.52 (d, J = 8.72, 2H, C6H5), 7.54 (d, J = 8.67, 2H, C6H5), 9.246 (s, 1NH). 13C NMR (600 MHz, DMSO-d6): 113.04, 120.40, 128.84, 151.94, 166.88. HRMS (ESI-TOF): m/z calcd for (C7H9N3O + H+): 152.0818; found: 152.0036.
Synthesis of hydrazides 3a and 3e: to a solution of methyl propionate 2a (5.50 mmol) or methyl 4-fluorobenzoate 2e (5.50 mmol) in methanol (15 mL), we added hydrazine monohydrate (22 mmol, 4 equiv.). The reaction mixture was stirred at reflux for 24 h and evaporated under diminished pressure. The oily residue 3a was purified by silica gel column chromatography and eluted with 3% MeOH/CHCl3 (v/v). Product 3e was crystallized from an ethanol:water (1:2, v/v) solution.
Propanoic acid hydrazide (3a): yield 100%. Mp 36–38 °C (lit. Mp 38–40 °C [48]). TLC system A; Rf: 0.5. 1H NMR (600 MHz, DMSO-d6): δ 0.99 (q, J = 7.62, CH3), 2.01 (t, J = 7.61, CH2), 4.13 (s, NH2), 8.90 (s, NH), 13C NMR (600 MHz, DMSO-d6): δ 10.37, 27.07, 172.86. HRMS (ESI-TOF): m/z calcd for (C3H8N2O + H+): 89.0709; found: 89.0736.
4-Fluorobenzhydrazide (3e): yield 60%. Mp 164–165 °C. (lit. Mp 160–163 °C [38]). TLC system B; Rf: 0.7. 1H NMR (600 MHz, DMSO-d6): δ 4.48 (s, 2NH2), 7.24–7.27 (m, 2H, C6H5), 7.86–7.88 (m, 2H, C6H5), 9.76 (s, 1NH). 13C NMR (600 MHz, DMSO-d6): 115.66 (d, J = 21.75), 129.96 (d, J = 8.98), 163.36, 165.01, 165.31. C7H7FN2O HRMS (ESI-TOF): m/z calcd for (C7H7FN2O + H+): 155.0615; found: 155.0623.
Synthesis of hydrazide (3d): to hydrazine monohydrate (98%, 10 mL), ethyl anthracene-9-carboxylate (2d) (1.60 mmol) was added. The reaction mixture was stirred at reflux for 72 h. The residual hydrazine was removed under reduced pressure. A crude solid was purified by flash column chromatography on silica gel, using 100% CHCl3 followed by 20% MeOH/CHCl3 (v/v) as an eluent. Additional purification by the high-performance liquid chromatography (HPLC) method using an ODS column with methanol as an eluent was performed.
Anthracene-9-carbohydrazide (3d): yield 60%. Td (thermal decomposition temperature) 242 °C. TLC system A; Rf: 0.7. 1H NMR (600 MHz, DMSO-d6): δ 4.81 (d, J = 3.10, NH2), 7.52–7.57 (m, 4H, C6H5), 7.98–8.00 (m, 2H, C6H5), 8.10–8.12 (m, 2H, C6H5), 8.65 (s, 1H, C6H5), 9.83 (s, 1NH). 13C NMR 600 MHz, DMSO-d6): 125.92, 126.02, 126.76, 127.83, 128.33, 128.77, 131.10, 132.51, 167.96. HRMS (ESI-TOF): m/z calcd for (C15H12N2O + H+): 237.1028; found: 237.1031.

3.1.3. General Procedure for the Synthesis of Hydrazones (4a4f) and Hydrazide (5)

Synthesis of hydrazones 4b4c and 4e4f: an appropriate hydrazide 3b,c,e,f (4.40 mmol) and 2,3,4-trihydroxybenzaldehyde (4.40 mmol) were dissolved in anhydrous methanol (15 mL), and the reaction mixture was stirred at room temperature for 24 h. After the completion of the reaction, the formed solid was filtered and purified by recrystallization from EtOH:H2O (1:1, v/v) [13].
(E)-N′-(2,3,4-trihydroxybenzylidene)benzohydrazide (4b): yield 85%. Td 187–189 °C. TLC system B; Rf: 0.4. 1H NMR (600 MHz, DMSO-d6): δ 6.41 (d, J = 8.39 Hz, 1H, C6H5), 6.80 (d, J = 8.48 Hz, 1H, C6H5), 7.53–7.56 (m, 2H, C6H5), 7.59–7.62 (m, 1H, C6H5) 7.93–7.94 (m, 2H, C6H5)), 8.48 (s, N=CH), 9.46 (s, OH), 11.56 (s, OH), 11.97 (s, NH). 13C NMR (600 MHz, DMSO-d6): δ 108.10, 111.28, 121.63, 127.98, 128.96, 132.27, 133.16, 133.38, 147.97, 149.21, 150.63, 162.94. HRMS (ESI-TOF): m/z calcd for (C14H12N2O4 + H+): 273.0870; found: 273.0882.
(E)-N′-(2,3,4-trihydroxybenzylidene)-1-naphthohydrazide (4c): yield 69%. Td 109–110 °C. TLC system C; Rf: 0.4. 1H NMR (600 MHz, DMSO-d6): δ 6.38 (d, J = 8.42, 1H, C6H5), 6.77 (d, J = 8.50, 1H, C6H5), 7.58–7.61 (m, 3H, C6H5), 7.76 (dd, J = 1.12, J = 7.03 1H, C6H5), 8.00–8.02 (m, 1H, C6H5), 8.09 (d, J = 8.27, 1H, C6H5), 8.23 (dd, J = 1.51, 8.15 1H, C6H5), 8.35 (s, N=CH), 8.49 (s, OH), 9.48 (s, OH), 11.47 (s, OH), 12.07 (s, NH) 13C NMR (600 MHz, DMSO-d6): δ 108.14, 111.23, 121.60, 125.42, 125.59, 126.44, 126.90, 127.56, 128.81, 130.41, 131.08, 132.77, 133.16, 133.62, 147.99,149.26, 150.51, 164.49. HRMS (ESI-TOF): m/z calcd for C18H14N2O4 + H+): 323.1026; found: 323.1033.
(E-4-fluoro-N′-(2,3,4-trihydroxybenzylidene)benzohydrazide (4e): yield 100%. Td 217–219 °C. TLC system B; Rf: 0.3. 1H NMR (600 MHz, DMSO-d6): 6.41(d, J = 8.42, 1H, C6H5), 6.80 (d, J = 8.49, 1H, C6H5) 7.38 (t, J = 8.84 2H, C6H5), 7.99–8.03 (m, 2H, C6H5), 8.47 (s, N=CH) 11.98 (s, NH) 13C NMR (600 MHz, DMSO-d6): δ 108.11, 111.24, 115.95 (d, JC-F = 21.87), 121.66, 130.70 (d, JC-F = 9.12), 133.13, 147.93, 149.21, 150.75, 161.99, 163.80, 165.46. HRMS (ESI-TOF): m/z calcd for (C14H11FN2O4 + H+): 291.0781; found: 291,0875.
(E)-4-amino-N′-(2,3,4-trihydroxybenzylidene)benzohydrazide (4f): yield 72%. Mp > 250 °C. TLC system C; Rf: 0.5. 1H NMR (600 MHz, DMSO-d6): δ 5.79 (s, NH2) 6.38 (d, J = 8.39, 1H, C6H5), 6.60 (d, J = 8.63, 2H, C6H5) 6.72 (d, J = 8.48 Hz, 1H, C6H5), 7.66 (d, J = 8.41, 2H, C6H5), 8.39 (s, N=CH), 8.44 (s, OH), 9.36 (s, OH), 11.55 (s, OH). 11.78 (s, NH) 13C NMR (600 MHz, DMSO-d6): δ 107.93, 111.47, 113.11, 119.41, 121.41, 129.68, 133.13, 147.79, 148.79, 149.01, 152.81, 162.85. HRMS (ESI-TOF): m/z calcd for (C14H13N3O4 + H+): 288.0979; found: 288.0988.
Synthesis of hydrazones 4a, 4d: to a solution of anthracene-9-carbohydrazide 3d (1.14 mmol) or propanoic acid hydrazide 3a (1.14 mmol) in methanol (15 mL), 2,3,4-trihydroxybenzaldehyde (1.26 mmol, 1.1 equiv.) was added. The reaction mixture was stirred at reflux for 24 h. After consumption of the substrate, the solvent was removed under reduced pressure. The crude solid was crystallized from a solution of ethanol:water (1:1 v/v).
(E)-N′-(2,3,4-trihydroxybenzylidene)propionohydrazide (4a): yield 89%. Td 184–186 °C. TLC system B; Rf: 0.5. 1H NMR (600 MHz, DMSO-d6): δ 1.08 (t, J = 7.57, CH3), 2.21 (q, J = 7.55, CH2) 6.37 (d, J = 8.41, 1H, C6H5), 6.73 (d, J = 8.46, 1H, C6H5), 8.17 (s, N=CH), 8.44 (s, OH), 9.40 (s, OH), 11.07 (s, OH), 11.39 (s, NH). 13C NMR (150 MHz, DMSO-d6): 9.41, 26.97, 107.40, 110.66, 120.86, 132.54, 147.20, 148.14, 148.38, 168.70 (HRMS (ESI-TOF): m/z calcd for (C10H12N2O4 + H+): 225.0869; found: 225.0874.
(E)-N′-(2,3,4-trihydroxybenzylidene)anthracene-9-carbohydrazide (4d). yield 50%. Td 167–169 °C. TLC system C; Rf: 0.5. 1H NMR (600 MHz, DMSO-d6): δ 6.43 (d, J = 8.37, 1H, C6H5), 6.82 (d, J = 8.48, 1H, C6H5) 7.54–7.64 (m, 4H, C6H5), 8.01–8.02 (m, 2H, C6H5), 8.16–8.19 (m, 2H, C6H5), 8.33 (s, N=CH), 8.57 (s, OH), 8.76 (s, 1H, C6H5), 9.57 (s, OH), 11.43 (s, OH), 12.33 (s, NH). 13C NMR 600 MHz, DMSO-d6): δ 108.26, 111.25, 121.61, 125.39, 126.22, 127.32, 127.48, 128.37, 128.99, 129.11, 131.09, 133.22, 148.05, 149.40, 150.59, 164.30. HRMS (ESI-TOF): m/z calcd for (C22H16N2O4 + H+): 373.1188; found: 373.1185.
Synthesis of hydrazide 5: (E)-4-amino-N′-(2,3,4-trihydroxybenzylidene)benzohydrazide 4f (3 mmol), anhydrous methanol (10 mL), and 25% of the weight of hydrazone 4f 20% Pd(OH)2/C were added to a reaction vessel. The vessel was placed in a Parr hydrogenator and the reaction mixture was treated with hydrogen at 2.2 bar at room temperature for 6 h. The catalyst was filtered from the solution and the reaction mixture was concentrated under reduced pressure. The crude solid was crystallized from an ethanol solution.
4-amino-N′-(2,3,4-trihydroxybenzyl)benzohydrazide (5): yield 40%. Td 188–190 °C. TLC system B; Rf: 0.2.1H NMR (600 MHz, DMSO-d6): δ 3.80 (d, J = 3.80, N-CH2), 5.18 (d, J = 5.07, NH-CH2), 5.63 (s, NH2) 6.20 (d, J = 8.11, 1H, C6H5), 6.42 (d, J = 8.18, 1H, C6H5) 6.53 (d, J = 8.54, 2H, C6H5), 7.54 (d, J = 8.50, 2H, C6H5) 8.12 (s, OH), 8.74 (s, OH), 9.20 (s, OH), 9.74 (d, J = 3.27, NH) 13C NMR (600 MHz, DMSO-d6): δ 52.50, 106.56, 113.00, 115.65, 119.58, 119.85, 129.07, 133.56, 145.81, 145.98, 152.25, 166.39. HRMS (ESI-TOF): m/z calcd for (C14H15N3O4 + Na+); 312.0955; found: 312.0955.

3.1.4. General Procedure for the Synthesis of Amino Acid Methyl Ester Hydrochloride 7a7c

Methyl esters 7a7c were obtained by the well-known amino acid esterification method, which uses thionyl chloride in methanol [25]. To a stirred and cooled (0 °C) solution of required L-amino acid 6 (30 mmol) in anhydrous methanol (40 mL), SOCl2 (1.10 equiv.) was added dropwise. The reaction mixture was warmed up to room temperature and stirred for 24 h. After completion of the reaction, excess methanol was removed and dried under reduced pressure.

3.1.5. General Procedure for the Synthesis of N-Benzyloxycarbonyl-L-amino Acid Methyl Esters 8a8c

An amino group of derivatives 7a7c was protected using benzyl chloroformate [25]. To a solution of amino acid methyl ester hydrochloride salt 7a7c (25 mmol) in dichloromethane (30 mL), triethylamine (2.5 equiv.) was added. After 10 min, benzyl chloroformate (1.2 equiv.) was added dropwise to the reaction mixture at 0 °C. The reaction mixture was stirred at room temperature for 24 h; then, water was added to solubilize all salts. The organic layer was washed with water (2 × 30 mL) and dried over anhydrous Na2SO4. After filtration, the organic layer was concentrated under reduced pressure. The residue was purified on a silica gel column using a mixture of AcOEt:n-hexane, 1:1 v/v, obtaining colorless oil 8a8c.

3.1.6. General Procedure for the Synthesis of N-Benzyloxycarbonyl-L-amino Acid Hydrazides 9a9c

To a solution of N-Cbz-amino acid methyl esters 8a8c (16.5 mmol) in anhydrous methanol (30 mL), hydrazine hydrate (98% 4.0 equiv.) was added. The reaction mixture was stirred for 24 h at room temperature. The insoluble product formed was filtered off and was recrystallized from a solution of ethanol: H2O, 1:1 v/v [25].
N-benzyloxycarbonyl-L-glycine hydrazide (9a): yield 97%. Mp 114-115 °C (lit. Mp 112–114 °C [49]). TLC system B; Rf: 0.8. 1H NMR (600 MHz, DMSO-d6): δ 3.55 (d, J = 6.16, CH2), 4.17(s, NH2), 5.00 (s, CH2C6H5), 7.28–7.39 (m, 5H, C6H5), 9.02 (s, 1NH), 13C NMR (600 MHz, DMSO-d6): δ 42.65, 65.91, 128.13, 128.23, 128.77, 137.44, 156.86, 168.93. HRMS (ESI-TOF): m/z calcd for (C10H13O3N3 + H+): 224.1030; found: 224.1046.
N-benzyloxycarbonyl-L-tyrosine hydrazide (9b): yield 36%. Mp 218–219 °C. (lit. Mp 219–221 °C [50]). TLC system B; Rf: 0.8. 1H NMR (600 MHz, DMSO-d6): δ 2.62–2.66 (m, HCH), 2.77–2.81 (m, HCH), 4.08–4.12 (m, CH), 4.21 (s, NH2) 4.91–4.96 (m, CH2C6H5), 5.04 (s, OH), 6.65 (d, J = 8.36, 2H, C6H5), 7.04 (d, J = 8.30, 2H, C6H5), 7.24–7.35 (m, 3H, C6H5), 7.44 (d, J = 8.71, 2H, C6H5), 9.17 (s, 1NH), 9.18 (s, NHCbz) 13C NMR (600 MHz, DMSO-d6): δ 37.48, 55.72, 65.57, 115.30, 127.86, 128.07, 128.49, 128.70, 128.76, 130.52, 137.52, 156.13, 156.19, 171.29. (ESI-TOF): m/z calcd for (C11H15N3O3S + H+): 330.1454; found: 330.1441.
N-benzyloxycarbonyl-L-cysteine hydrazide (9c): yield 67%. Mp 138–139 °C (lit. Mp 141–143 °C [51]). TLC system B; Rf: 0.7. 1H NMR (600 MHz, DMSO-d6): δ 1.23 (s, SH), 2.84–2.88 (m, CH), 3.03–3.06 (m, CH), 4.26 (s, NH2), 5.00–5.04 (m, CH2C6H5), 7.31–7.36 (m, 4H, C6H5), 7.56 (d, J = 8.53, 1H, C6H5) 9.30 (s, NH, NHCbz) 13C NMR (600 MHz, DMSO-d6): δ 53.00, 56.47, 66.00, 128.16, 128.23, 128.75, 137.33, 156.24, 169.60. HRMS (ESI-TOF): m/z calcd. for (C7H8ON2 − H): 268.0761; found: 268.0763.

3.1.7. General Procedure for the Synthesis of Hydrazones of N-Benzyloxycarbonyl-amino Acids 10a10c

Synthesis of (E)-benzyl(2-oxo-2-(2-(2,3,4-trihydroxybenzylidene)hydrazinyl)ethyl) carbamate: 2,3,4-trihydroxybenzaldehyd (11.7 mmol; 1.20 equiv.) was added to a stirring solution of N-Cbz-L-glycine hydrazide 9a (9.8 mmol) in methanol (25 mL). The mixture solution was stirred at room temperature for 24 h. After completion of the reaction, the obtained solid was filtered off. The product was obtained as a solid and purified by recrystallization from the solution of methanol: H2O (1:1 v/v).
Synthesis of (E)-benzyl(3-(4-hydroxyphenyl)-1-oxo-1-(2-(2,3,4-trihydroxybenzylidene)- hydrazinyl)propan-2-yl)carbamate 10b: 2,3,4-trihydroxybenzaldehyd (1.20 mmol; 1.20 equiv.) was added to a stirring solution of N-benzyloxycarbonyl-L-tyrosine hydrazide 9b (1.00 mmol) in THF (10 mL). The resulting solution was stirred at reflux for 72 h in an inert atmosphere of argon. The reaction mixture was concentrated under reduced pressure. The residue was crystallized from a 1:1 (v/v) solution of methanol:H2O at a 55% yield.
Synthesis of (E)-benzyl(3-mercapto-1-oxo-1-(2-(2,3,4-trihydroxybenzylidene)hydrazinyl)propan-2-yl)carbamate 10c: 2,3,4-trihydroxybenzaldehyd (16.92 mmol; 1.20 equiv.) was added to a stirring solution of N-Cbz-L-cysteine hydrazide 9c (14.10 mmol) in methanol (30 mL). The mixture solution was stirred at room temperature for 24 h and concentrated under reduced pressure. The residue was crystallized from an ethanol:H2O, 1:1 v/v solution.
(E)-benzyl(2-oxo-2-(2-(2,3,4-trihydroxybenzylidene)hydrazinyl)ethyl)carbamate (10a): yield 97%. Mp 213–215 °C (Mp 211–212 °C [52]). TLC system B; Rf: 0.3. 1H NMR (600 MHz, DMSO-d6): δ 3.75 (d, J = 4.41, CH2), 5.06 (s, CH2C6H5), 6.38 (d, J = 7.92, 1H, C6H5) 6.77 (d, J = 7.87, 1H, C6H5), 7.32–7.37 (m, 4H, C6H5), 7.60 (s, 1H, C6H5), 8.25 (s, N=CH), 8.47 (s, OH) 9.45, (s, OH), 9.52 (d, J = 15.50, NHCbz), 11.32 (s, OH), 11.54 (s, NH) 13C NMR (600 MHz, DMSO-d6): δ 43.09 66.03, 108.04, 111.16, 121.48, 128.11, 128.19, 128.26, 128.79, 133.11, 137.41, 147.80, 149.12, 149.62, 157.00, 165.65. HRMS (ESI-TOF): m/z calcd for (C17H17N3O6 + H+): 360.1190; found: 360.1197.
(E)-benzyl(3-(4-hydroxyphenyl)-1-oxo-1-(2-(2,3,4-trihydroxybenzylidene)hydrazinyl)propan-2-yl)carbamate (10b): yield 55%. Mp 106–108 °C. TLC system B; Rf: 0.3. 1H NMR (600 MHz, DMSO-d6): δ 2.72–2.746 (m, HCH), 2.88–2.91 (m, HCH), 4.20–4.21 (m, CH), 4.94–5.00 (m, CH2C6H5), 6.66 (d, J = 8.16, 1H, C6H5), 6.38 (d, J = 8.40, 1H, C6H5), 6.77 (d, J = 8.47, 1H, C6H5), 7.08 (d, J = 8.00, 2H, C6H5), 8.24 (s, N=CH), 8.46 (s, OH), 9.21 (s, NHCbz) 9.46 (s, OH), 11.27 (s, OH) 13C NMR (600 MHz, DMSO-d6): δ 36.96, 56.11, 65.76, 108.05, 111.18, 115.37, 121.45, 127.87, 127.94, 128.15, 128.70 128.73, 130.56, 133.10, 137.40, 147.81, 149.14, 149.86, 156.29, 168.02. HRMS (ESI-TOF): m/z calcd for (C24H23N3O7 + H+); 466.1609; found: 466.1610.
(E)-benzyl(3-mercapto-1-oxo-1-(2-(2,3,4-trihydroxybenzylidene)hydrazinyl)propan-2-yl)carbamate (10c): yield 49%. Mp 134–135 °C. TLC system B; Rf: 0.2. 1H NMR (600 MHz, DMSO-d6): δ 1.24 (s, SH), 2.96–2.99 (m, HCH), 3.18-3.21 (m, HCH), 4.39 (dd, J = 8.54, 14.07 CH), 5.05 (s, CH2C6H5), 6.37 (d, J = 8.41, 1H, C6H5), 6.75 (d, J = 8.41, 1H, C6H5), 7.30–7.37 (m, 4H, C6H5), 7.84 (d, J = 8.21, 1H, C6H5), 8.31 (s, N=CH), 8.48 (s, OH), 9.14 (s, NHCbz), 9.48 (s, OH), 11.25 (s, OH), 11.78 (s, NH). 13C NMR (600 MHz, DMSO-d6): δ 53.50, 65.97, 66.18, 108.10, 111.13, 121.56, 128.21, 128.29, 128.78, 133.10, 137.21, 147.88, 149.26, 150.40, 156.40, 166.49. HRMS (ESI-TOF): m/z calcd for (C18H19N3O6S − H): 404.0921; found: 404.0916.

3.1.8. General Procedure for the Synthesis of Substituted Amino Acid Hydrazide 11 and Substituted Amino Acid Hydrazone 12

Synthesis of 2-amino-N′-(2,3,4-trihydroxybenzyl)acetohydrazide 11: ammonium formate (1 mmol, 1 equiv.), 40% (w/v) hydrazone 10a, and 10% PdOH2 (145 mg) were added to a stirring solution of N-(N-Cbz-L-glycine)-2,3,4-trihydroxybenzaldehyde hydrazone 10a (1 mmol) in methanol (10 mL). The reaction mixture was stirred at 50 °C under an argon atmosphere for 12 h then was cooled down to room temperature, filtered to remove the catalyst, and concentrated under reduced pressure. The residue was purified by crystallization from diethyl ether and the pure powder was lyophilized. The hygroscopic powder was obtained in 48% yield.
Synthesis of (E)-2-amino-3-(4-hydroxyphenyl)-N′-(2,3,4-trihydroxybenzylidene)propanehydrazide 12: hydrazone 10b (0.5 mmol), anhydrous methanol (30 mL), and 25% (w/v) hydrazone 10b were added to a mixture of 10% Pd/C and Pd(OH)2. The reaction vessel was placed in a Parr shaker hydrogenator at pressures of up to 2.5 bar at room temperature for 6 h. The solid catalyst was filtered off and the reaction solution was concentrated under reduced pressure. The product was purified by reverse-phase preparative HPLC on an ODS column using 70% MeOH:H2O (v/v).
2-Amino-N′-(2,3,4-trihydroxybenzyl)acetohydrazide (11): yield 48%. Hygroscopic powder. TLC system B; Rf: 0.1. 1H NMR (600 MHz, DMSO-d6): δ 1.99 (s, NH2), 3.06 (s, CH2), 3.49 (s, NH-CH2), 3.60 (s, NH-CH2), 6.12 (d, J = 8.06, 1H, C6H5), 6.28 (d, J = 8.09, 1H, C6H5), 8.86 (s, NH). 13C NMR (600 MHz, DMSO-d6): δ 43.17, 44.11, 106.75, 115.31, 119.81, 133.78, 145.00, 145.16, 172.52. HRMS (ESI-TOF): m/z calcd for (C9H13N3O4 + H+) 228.0978; found: 228.0331.
(E)-2-Amino-3-(4-hydroxyphenyl)-N′-(2,3,4-trihydroxybenzylidene)propanehydrazide (12): yield 20%. Td 176–178 °C TLC system C; Rf: 0.2. 1H NMR (600 MHz, DMSO-d6): δ 2.58–2.61 (m, HCH), 2.76–2.79 (m, HCH), 3.65–3.69 (m, CH), 4.41 (d, J = 7.43, NH), 4.51 (d, J = 7.37, NH), 6.18–6.21 (m, 1H, C6H5), 6.36–6.40 (m, 1H, C6H5), 6.62–6.68 (m, 2H, C6H5), 6.91–7.07 (m, 2H, C6H5), 8.04 (s, N=CH), 8.31 (s, NH) 13C NMR (600 MHz, DMSO-d6): δ 29.46, 55.21, 106.66, 115.39, 119.68, 128.29, 128.45, 130.57, 130.65, 133.46, 145.63, 146.01, 156.26, 172.87 HRMS (ESI-TOF): m/z calcd for (C16H17N3O5 + H+): 332.1241; found: 332.1253.
1H and 13C NMR spectra of synthesized compounds are presenting in Supplementary Materials.

3.2. Computational Methods

Calculations were performed by the Orca 4.2.1 package on the DFT level (B3LYP/def2-SVP) [53]. The accuracy of the optimization process was determined using the Hessian eigenvalue analysis. All the calculated Hessian eigenvalues were positive for the compounds evaluated. In the case of calculations in a solvent environment, the CPCM continuous solvation model for DMSO was used.

3.3. Hexokinase Activity Assay

For studying the potential of HK2 inhibitors, a commercially available assay test (ab211114) Hexokinase II Inhibitor Screening (colorimetric) was used. The in vitro hexokinase activity assay was conducted according to the manufacturer’s instructions. Briefly, the enzyme and substrate solution were prepared. Next, the enzyme solution was added to the wells containing sample compounds and incubated for 5 min at 25 °C. Then, the substrate solution was added to the wells and the absorbance was measured at 450 nm every 5 min for 45 min using the Thermo Scientific™ Varioskan™ LUX multimode microplate reader (ThermoFisher Inc.Waltham, MA, USA). The test compounds were dissolved in DMSO at 50 µM, 5 µM, and 1 µM concentrations, with the final concentration of solvent not exceeding 1% by volume.

4. Conclusions

In our research, we synthesized benitrobenrazide and benserazide analogues. We identified that some of these compounds, namely compounds 4e and 4f, represent a promising class of HK2 inhibitors, inhibiting HK2 at a concentration of 5 µM by 98% and 82%, respectively. At the lower concentration of 1 µM, 4e and 4f inhibited HK2 by 60% and 54%, respectively. We have confirmed that the presence of a bulky anthracenyl group in 4d has no significant effect on HK2 enzymatic activity. The exchanging of serine by glycine or threonine in benserazide analogues has a minor effect on their inhibition activity against HK2. Compounds 11 and 12 reduce the enzymatic activity of HK2 by approx. 40% in comparison with the negative control. The presented findings suggest that the imine scaffold -CH=N-, in the structure of the potent HK2 inhibitors, helps to enhance and regulate their biological activities. The -CH=N- core is responsible for the possible binding of various groups with nucleophilic and electrophilic properties and, thus, can interact with targeted enzymes and inhibit their enzymatic activity.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29030629/s1, Figure S1: Hydrogen bonding interaction in compound 4a; 1H and 13C NMR spectra of synthesized compounds. Figures S2–S27: 1H and 13C NMR spectrum of synthesized compounds.

Author Contributions

Conceptualization, K.J. and K.W.; methodology, K.J. and K.W.; formal analysis, software, W.S.; investigation, K.J.; resources, K.J. and K.W.; writing—original draft preparation, K.J.; writing—review and editing, K.W.; supervision, K.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by STU internal fund, grant number 04/020/BKM22/1061.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and supplementary materials.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Raczuk, E.; Dmochowska, B.; Samaszko-Fiertek, J.; Madaj, J. Different Schiff Bases-Structure, Importance and Classification. Molecules 2022, 27, 787. [Google Scholar] [CrossRef]
  2. Uddin, N.; Rashid, F.; Ali, S.; Tirmizi, S.A.; Ahmad, I.; Zaib, S. Synthesis, characterization, and anticancer activity of Schiff bases. J. Biomol. Struct. Dyn. 2020, 38, 3246–3259. [Google Scholar] [CrossRef]
  3. Kajal, A.; Bala, S.; Kamboj, S.; Sharma, N.; Saini, V. Schiff Bases: A Versatile Pharmacophore. J. Catal. 2013, 893512. [Google Scholar] [CrossRef]
  4. Mounika, K.; Anupama, B.; Pragathi, J.; Gyanakumari, C. Synthesis, characterization and biological activity of a Schiff base derived from 3-ethoxy salicylaldehyde and 2-amino benzoic acid and its transition metal complexes. J. Sci. Res. 2010, 2, 513–524. [Google Scholar] [CrossRef]
  5. Kargar, H.; Fallah-Mehrjardi, M.; Behjatmanesh-Ardakani, R.; Tahir, M.N.; Ashfaqc, M.; Munawar, K.S. Synthesis, crystal structure determination, Hirshfeld surface analysis, spectral characterization, theoretical and computational studies of titanium(IV) Schiff base complex. J. Coord. Chem. 2021, 74, 2682–2700. [Google Scholar] [CrossRef]
  6. Awantu, A.F.; Fongang, Y.S.F.; Ayimele, G.A.; Nantia, E.A.; Fokou, P.V.T.; Boyom, F.F.; Ngwang, C.K.; Bruno, N.; Lenta, B.N.; Ngouela, S.A. Novel Hydralazine Schiff Base Derivatives and Their Antimicrobial, Antioxidant and Antiplasmodial Properties. Int. J. Org. Chem. 2020, 10, 1–16. [Google Scholar] [CrossRef]
  7. Taha, M.; Ismail, N.H.; Imran, S.; Anouar, H.; Selvaraj, M.; Jamil, W.; Ali, M.; Kashif, S.M.; Rahim, F.; Khan, K.M.; et al. Synthesis and molecular modelling studies of phenyl linked oxadiazole-phenylhydrazone hybrids as potent antileishmanial agents. Eur. J. Med. Chem. 2017, 126, 1021–1033. [Google Scholar] [CrossRef] [PubMed]
  8. Li, L.-Y.; Peng, J.-D.; Zhou, W.; Qiao, H.; Deng, X.; Li, Z.-H.; Li, J.-D.; Fu, Y.-D.; Li, S.; Sun, K.; et al. Potent hydrazone derivatives targeting esophageal cancer cells. Eur. J. Med. Chem. 2018, 148, 359–371. [Google Scholar] [CrossRef]
  9. Küçükgüzel, S.G.; Mazi, A.; Sahin, F.; Öztürk, S.; Stables, S. Synthesis and biological activities of diflunisal hydrazide-hydrazones. Eur. J. Med. Chem. 2003, 38, 1005–1013. [Google Scholar] [CrossRef]
  10. Khan, K.M.; Rasheed, M.; Ullah, Z.; Hayat, S.; Kaukab, F.; Choudhary, M.I.; Ur-Rahman, A.; Perveen, S. Synthesis and in vitro leishmanicidal activity of some hydrazides and their analogues. Bioorg. Med. Chem. 2003, 11, 1381–1387. [Google Scholar] [CrossRef]
  11. Durcik, M.; Tammela, P.; Barančoková, M.; Tomašič, T.; Ilaš, J.; Kikelj, D.; Zidar, N. Synthesis and Evaluation of N-Phenylpyrrolamides as DNA Gyrase B Inhibitors. Chem. Med. Chem. 2018, 13, 186–198. [Google Scholar] [CrossRef] [PubMed]
  12. Kareem, H.S.; Ariffin, A.; Nordin, N.; Heidelberg, T.; Azlina Abdul-Aziz, A.; Kong, K.W.; Yehye, W.A. Correlation of antioxidant activities with theoretical studies for new hydrazone compounds bearing a 3,4,5-trimethoxy benzyl moiety. Eur. J. Med. Chem. 2015, 103, 497–505. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, Y.; Li, M.; Zhang, Y.; Wu, C.; Yang, K.; Gao, S.; Zheng, M.; Li, X.; Li, H.; Chen, L. Structure based discovery of novel hexokinase 2 inhibitors. Bioorg. Chem. 2020, 96, 103609. [Google Scholar] [CrossRef] [PubMed]
  14. Wei, L.; Mengzhu, Z.; Shuangping, W.; Gao, S.; Yang, M.; Li, Z.; Min, Q.; Sun, W.; Chen, L.; Xiang, G.; et al. Benserazide, a dopa decarboxylase inhibitor, suppresses tumor growth by targeting hexokinase 2. J. Exp. Clin. Cancer Res. 2017, 36, 58–70. [Google Scholar] [CrossRef]
  15. Zheng, M.; Wu, C.; Yang, K.; Yang, Y.; Liu, Y.; Gao, S.; Wang, Q.; Li, C.; Chen, L.; Li, H. Novel selective hexokinase 2 inhibitor Benitrobenrazide blocks cancer cells growth by targeting glycolysis. Pharmacol. Res. 2021, 164, 105367. [Google Scholar] [CrossRef] [PubMed]
  16. Pelicano, H.; Martin, D.S.; Xu, R.-H.; Huang, P. Glycolysis inhibition for anticancer treatment. Oncogene 2006, 25, 4633–4646. [Google Scholar] [CrossRef] [PubMed]
  17. Hay, N. Aerobic glycolysis Oxidative phosphorylation Reprogramming glucose metabolism in cancer: Can it be exploited for cancer therapy? Nat. Publ. Gr. 2016, 16, 635. [Google Scholar] [CrossRef]
  18. Heiden, M.G.V.; Cantley, L.C.; Thompson, C.B. Understanding the Warburg Effect: The Metabolic Requirements of Cell Proliferation. Science 2009, 324, 1029. [Google Scholar] [CrossRef]
  19. Semenza, G.L. HIF-1: Upstream and downstream of cancer metabolism. Curr. Opin. Genet. Dev. 2010, 20, 51–56. [Google Scholar] [CrossRef]
  20. Wu, J.; Hu, L.; Wu, F.; Zou, J.; He, T. Poor prognosis of hexokinase 2 overexpression in solid tumors of digestive system: A meta-analysis. Oncotarget 2017, 8, 32332. [Google Scholar] [CrossRef]
  21. Ciscato, F.; Ferrone, L.; Masgras, I.; Laquatra, C.; Rasola, A. Hexokinase 2 in Cancer: A Prima Donna Playing Multiple Characters. Int. J. Mol. Sci. 2021, 22, 4716. [Google Scholar] [CrossRef] [PubMed]
  22. Counihan, J.L.; Grossman, E.A.; Nomura, D.K. Cancer Metabolism: Current Understanding and Therapies. Chem. Rev. 2018, 118, 6893–6923. [Google Scholar] [CrossRef] [PubMed]
  23. Wilson, J.E. Isozymes of mammalian hexokinase: Structure, subcellular localization, and metabolic function. J. Exp. Biol. 2003, 206, 2049–2057. [Google Scholar] [CrossRef] [PubMed]
  24. Garcia, S.N.; Guedes, R.C.; Marques, M.M. Unlocking the Potential of HK2 in Cancer Metabolism and Therapeutics. Curr. Med. Chem. 2019, 26, 7285–7322. [Google Scholar] [CrossRef]
  25. Tsai, J.H.; Wilson, J.E. Functional organization of mammalian hexokinases: Characterization of the rat type III isozyme and its chimeric forms, constructed with the N- and C-terminal halves of the type I and type II isozymes. Arch. Biochem. Biophys. 1997, 338, 183–192. [Google Scholar] [CrossRef]
  26. Zhang, F.; Angelova, A.; Garamus, V.M.; Angelov, B.; Tu, S.; Kong, L.; Zhang, X.; Li, N.; Zou, A. Mitochondrial Voltage-Dependent Anion Channel 1−Hexokinase-II Complex-Targeted Strategy for Melanoma Inhibition Using Designed Multiblock Peptide Amphiphiles. ACS Appl. Mater. Interfaces 2021, 13, 35281–35293. [Google Scholar] [CrossRef] [PubMed]
  27. Lin, H.; Zeng, J.; Xie, R.; Schulz, M.J.; Tedesco, R.; Qu, J.; Erhard, K.F.; Mack, J.F.; Raha, K.; Rendina, A.R.; et al. Discovery of a Novel 2,6-Disubstituted Glucosamine Series of Potent and Selective Hexokinase 2 Inhibitors. ACS Med. Chem. Lett. 2016, 7, 217–222. [Google Scholar] [CrossRef]
  28. Tanbin, S.; Fuad, F.A.A.; Hamid, A.A.A. Virtual Screening for Potential Inhibitors of Human Hexokinase II for the Development of Anti-Dengue Therapeutics. BioTech 2021, 10, 1. [Google Scholar] [CrossRef]
  29. Juszczak, K.; Kubicka, A.; Kitel, R.; Zawadzki, S.; Marczak, A.; Dzido, G.; Walczak, K.; Łabieniec-Watała, M.; Matczak, K.; Tomczyk, M.D. Hexokinase 2 Inhibition and Biological Effects of BNBZ and Its Derivatives: The Influence of the Number and Arrangement of Hydroxyl Groups. Int. J. Mol. Sci. 2022, 23, 2616. [Google Scholar] [CrossRef]
  30. Shan, W.; Zhou, Y.; Tam, K.Y. The development of small-molecule inhibitors targeting hexokinase 2. Drug Discov. Today 2022, 27, 2574–2585. [Google Scholar] [CrossRef]
  31. Agnihotri, S.; Mansouri, S.; Burrell, K.; Li, M.; Mamatjan, Y.; Liu, J.; Nejad, R.; Kumar, S.; Jalali, S.; Singh, S.K.; et al. Ketoconazole and Posaconazole Selectively Target HK2-expressing Glioblastoma Cells. Clin. Cancer Res. 2019, 25, 844–855. [Google Scholar] [CrossRef] [PubMed]
  32. Salani, B.; Del Rio, A.; Marini, C.; Sambuceti, G.; Cordera, R.; Maggi, D. Metformin, cancer, and glucose metabolism. Endocr.-Relat. Cancer 2014, 21, 461–471. [Google Scholar] [CrossRef] [PubMed]
  33. Nepali, K.; Lee, H.Y.; Liou, J.P. Nitro-Group-Containing Drugs. J. Med. Chem. 2019, 62, 2851–2893. [Google Scholar] [CrossRef] [PubMed]
  34. Nishiwaki, N. A Walk through Recent Nitro Chemistry Advances. Molecules 2020, 25, 3680. [Google Scholar] [CrossRef] [PubMed]
  35. Wang, Q.; Jianlin Han, J.; Sorochinsky, A.; Landa, A.; Soloshonok, V.A. The Latest FDA-Approved Pharmaceuticals ContainingFragments of Tailor-Made Amino Acids and Fluorine. Pharmaceuticals 2022, 15, 999. [Google Scholar] [CrossRef] [PubMed]
  36. Al-Harthy, T.; Zoghaib, W.; Abdel-Jalil, R. Importance of Fluorine in Benzazole Compounds. Molecules 2020, 25, 4677. [Google Scholar] [CrossRef] [PubMed]
  37. Hosangadi, B.D.; Dave, R.H. An efficient general method for esterification of aromatic carboxylic acids. Tetrahedron Lett. 1996, 37, 6375–6378. [Google Scholar] [CrossRef]
  38. Li, X.; Zhao, Z.; Li, G.; Shi, P. Design and Synthesis of Novel Molecular Tweezer Anion Receptors based on Diphenic Acid Carbonyl Thiosemicarbazide. J. Chem. Res. 2010, 34, 410–413. [Google Scholar] [CrossRef]
  39. Rohane, S.H.; Chauhan, A.J.; Fuloria, N.K.; Fuloria, S. Synthesis and invitro antimycobacterial potential of novel hydrazones of eugenol. Arab. J. Chem. 2020, 13, 4495–4504. [Google Scholar] [CrossRef]
  40. Gloaguen, E.; Brenner, V.; Alauddin, M.; Tardivel, B.; Mons, M.; Zehnacker-Rentien, A.; Declerck, V.; Aitkenet, D.J. Direct spectroscopic evidence of hyperconjugation unveils the conformational landscape of hydrazides. Angew. Chem. Int. Ed. Engl. 2014, 53, 13756–13759. [Google Scholar] [CrossRef]
  41. Özen, A.S.; De Proft, F.; Aviyente, V.; Geerlings, P. Interpretation of hydrogen bonding in the weak and strong regions using conceptual DFT descriptors. J. Phys. Chem. A 2006, 110, 5860–5868. [Google Scholar] [CrossRef]
  42. Green, T.W.; Wuts, P.G.M. Protective Groups in Organic Synthesis; Wiley-Interscience: New York, NY, USA, 1999. [Google Scholar]
  43. Kozioł, A.; Lendzion-Paluch, A.; Manikowski, A. A fast and effective hydrogenation process of protected pentasaccharide: A key step in the synthesis of fondaparinux sodium. Org. Process Res. Dev. 2013, 17, 869–875. [Google Scholar] [CrossRef]
  44. Bartholomew, C.H. Mechanisms of catalyst deactivation. Appl. Catal. A Gen. 2001, 212, 17–60. [Google Scholar] [CrossRef]
  45. Guan, Y.; Wang, J.; Sun, J. A method for determination of hexokinase activity by RP-HPLC. J. Nat. Sci. 2011, 16, 535–540. [Google Scholar] [CrossRef]
  46. Rodrigues, D.A.; Guerra, F.S.; Sagrillo, F.S.; Pinheiro, P.S.M.; Alves, M.A.; Thota, S.; Chaves, L.S.; Sant’Anna, C.M.R.; Fernandes, P.D.; Fraga, C.A.M. Design, Synthesis, and Pharmacological Evaluation of First-in-Class Multitarget N-Acylhydrazone Derivatives as Selective HDAC6/8 and PI3Kα Inhibitors. ChemMedChem 2020, 15, 539–551. [Google Scholar] [CrossRef] [PubMed]
  47. Younis, A.; Awad, G.E.A. Utilization of Ultrasonic as an Approach of Green Chemistry for Synthesis of Hydrazones an Bishydrazones as Potential Antimicrobial Agents. Egypt. J. Chem. 2020, 63, 599–610. [Google Scholar] [CrossRef]
  48. Saha, A.; Kumar, R.; Kumar, R.; Devakumar, C. Development and assessment of green synthesis of hydrazides. Indian J. Chem. 2010, 49B, 526–531. [Google Scholar]
  49. de Fátima, S.; Barreto, A.; dos Santos, V.A.; Andrade, C.K.Z. Synthesis of acylhydrazino-peptomers, a new class of peptidomimetics, by consecutive Ugi and hydrazino-Ugi reactions. Beilstein J. Org. Chem. 2016, 12, 2865–2872. [Google Scholar] [CrossRef]
  50. Schröder, E.; Gibian, H. Über Peptidsynthesen, XII. Synthese von Glukagon-Teilsequenzen. Liebigs Ann. Chem. 1962, 656, 190–204. [Google Scholar] [CrossRef]
  51. Schnabel, E. Nebenreaktionen bei der Synthese von Peptiden nach dem Azidverfahren von Curtius. Liebigs Ann. Chem. 1962, 659, 168–184. [Google Scholar] [CrossRef]
  52. Bartholini, G.; Hegedues, B. Hoffmann La Roche. Ein Hydrazid und dessen Saeureadditionssalze. Patent Germany 03.12.1970 DE1941261, 1970. [Google Scholar]
  53. Neese, F. The ORCA program system. WIREs Comput. Mol. Sci. 2012, 2, 73–78. [Google Scholar] [CrossRef]
Figure 1. Tautomerism in a hydrazide molecule.
Figure 1. Tautomerism in a hydrazide molecule.
Molecules 29 00629 g001
Figure 2. The chemical structures of reported HK2 inhibitors.
Figure 2. The chemical structures of reported HK2 inhibitors.
Molecules 29 00629 g002
Scheme 1. Preparation of benitrobenrazide derivatives. Reagent and conditions: (A) a: SOCl2 (1.2 equiv.), MeOH, 0 °C to RT, 24 h; b: N2H4·H2O (98%, 4 equiv.), MeOH, RT/reflux 24/72 h.; c: (HO)3C6H2CHO; (1.0 equiv.), MeOH, RT/reflux, 24 h. (B) d: SOCl2, (53.6 equiv.), DMF, RT, Ar, 4 h; e: EtOH, TEA (1.2 equiv.), RT, 24 h; g: N2H4·H2O (98%, 100 equiv.), reflux, 72 h; g: (HO)3C6H2CHO; (1.1 equiv.), MeOH, reflux; 24 h.
Scheme 1. Preparation of benitrobenrazide derivatives. Reagent and conditions: (A) a: SOCl2 (1.2 equiv.), MeOH, 0 °C to RT, 24 h; b: N2H4·H2O (98%, 4 equiv.), MeOH, RT/reflux 24/72 h.; c: (HO)3C6H2CHO; (1.0 equiv.), MeOH, RT/reflux, 24 h. (B) d: SOCl2, (53.6 equiv.), DMF, RT, Ar, 4 h; e: EtOH, TEA (1.2 equiv.), RT, 24 h; g: N2H4·H2O (98%, 100 equiv.), reflux, 72 h; g: (HO)3C6H2CHO; (1.1 equiv.), MeOH, reflux; 24 h.
Molecules 29 00629 sch001
Figure 3. Intramolecular hydrogen bonding interaction in synthesized Schiff base 4.
Figure 3. Intramolecular hydrogen bonding interaction in synthesized Schiff base 4.
Molecules 29 00629 g003
Scheme 2. Tautomeric equilibrium in 4a benitrobenrazide.
Scheme 2. Tautomeric equilibrium in 4a benitrobenrazide.
Molecules 29 00629 sch002
Scheme 3. Reduction of hydrazone 4f.
Scheme 3. Reduction of hydrazone 4f.
Molecules 29 00629 sch003
Scheme 4. Preparation of benserazide analogues. Reagent and conditions: a: MeOH, SOCl2 (1.2 equiv.), 0 °C to RT, 24 h; b: CbzCl (1.2 equiv.), TEA (2.5 equiv.), DCM, RT, 24 h; c: 98% N2H4.H2O (4.0 equiv.), MeOH, RT, 24 h; d: (HO)3C6H2CHO, (1.2 equiv.), MeOH, RT, 24 h/(HO)3C6H2CHO; (1.2 equiv.), THF, reflux, 72 h; e: HCOONH4, 25% Pd(OH)2/C, MeOH, 50 °C, Ar, 12 h; f: H2, 25% mixture Pd(OH)2/C and Pd/C, 2.5 bar, MeOH, RT, 6 h.
Scheme 4. Preparation of benserazide analogues. Reagent and conditions: a: MeOH, SOCl2 (1.2 equiv.), 0 °C to RT, 24 h; b: CbzCl (1.2 equiv.), TEA (2.5 equiv.), DCM, RT, 24 h; c: 98% N2H4.H2O (4.0 equiv.), MeOH, RT, 24 h; d: (HO)3C6H2CHO, (1.2 equiv.), MeOH, RT, 24 h/(HO)3C6H2CHO; (1.2 equiv.), THF, reflux, 72 h; e: HCOONH4, 25% Pd(OH)2/C, MeOH, 50 °C, Ar, 12 h; f: H2, 25% mixture Pd(OH)2/C and Pd/C, 2.5 bar, MeOH, RT, 6 h.
Molecules 29 00629 sch004
Figure 4. Structure of the final derivatives evaluated for inhibition of HK2.
Figure 4. Structure of the final derivatives evaluated for inhibition of HK2.
Molecules 29 00629 g004
Figure 5. Inhibition of HK2 activity, demonstrated as a reduction in rate or extent of generation of HK2-dependent absorbance (OD) at 450 nm by a synthesized compound. Ctrl- is an enzyme control without inhibitor, which shows the normal rate of phosphorylation by HK2. All experiments were performed independently in triplicate.
Figure 5. Inhibition of HK2 activity, demonstrated as a reduction in rate or extent of generation of HK2-dependent absorbance (OD) at 450 nm by a synthesized compound. Ctrl- is an enzyme control without inhibitor, which shows the normal rate of phosphorylation by HK2. All experiments were performed independently in triplicate.
Molecules 29 00629 g005
Figure 6. Inhibitory effect of synthesized compounds on HK2 enzyme activity. Bromopyruvic acid was used as a positive HK2 inhibitor control. The relative activity of the negative enzyme control (without inhibitor) was set at 100%. Results are displayed as the mean ± SD from three independent experiments.
Figure 6. Inhibitory effect of synthesized compounds on HK2 enzyme activity. Bromopyruvic acid was used as a positive HK2 inhibitor control. The relative activity of the negative enzyme control (without inhibitor) was set at 100%. Results are displayed as the mean ± SD from three independent experiments.
Molecules 29 00629 g006
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

Juszczak, K.; Szczepankiewicz, W.; Walczak, K. Synthesis and Primary Activity Assay of Novel Benitrobenrazide and Benserazide Derivatives. Molecules 2024, 29, 629. https://doi.org/10.3390/molecules29030629

AMA Style

Juszczak K, Szczepankiewicz W, Walczak K. Synthesis and Primary Activity Assay of Novel Benitrobenrazide and Benserazide Derivatives. Molecules. 2024; 29(3):629. https://doi.org/10.3390/molecules29030629

Chicago/Turabian Style

Juszczak, Karolina, Wojciech Szczepankiewicz, and Krzysztof Walczak. 2024. "Synthesis and Primary Activity Assay of Novel Benitrobenrazide and Benserazide Derivatives" Molecules 29, no. 3: 629. https://doi.org/10.3390/molecules29030629

Article Metrics

Back to TopTop