1. Introduction
Chalcones (I) are natural products, the biosynthetic precursors of flavonoids, a large family of plant phenolic secondary metabolites [
2]. Because of the wide range of beneficial biological actions of the natural chalcones, several analogs have been syn-thesized and – among others - tested for their antioxidant, antimicrobial, antiprotozo-al, antiulcer, antihistaminic, antidiabetic, anti-inflammatory, anticancer and also neu-roprotective activities [
3,
4,
5,
6]. The molecular mechanisms of the published biologi-cal/pharmacological effects can be associated with their (a) non-covalent interactions with biological macromolecules and (b) covalent modification of preferably the soft nucleophilic thiol function(s) of amino acids, peptides, and proteins [
7,
8,
9].
The chalcone structure can be devided into three different structural units; the aromatic rings A and B and the propenone linker (
Figure 1). Modifying any of them can tune the main feature of interactions of the synthetic chalcones towards the non-covalent or the covalent pathway. In our previous studies, we have investigated how the substitution of the B-ring and the ring size of some cyclic chalcone analogs af-fect the cancer cell cytotoxic effect of more than 120 derivatives [
10,
11,
12]. While com-paring the average IC50 values of the series, the benzosuberone (II) analogs displayed the most prosperous data against P388, L1210, Molt 4/C8, and CEM cells, as well as a panel of human tumor cell lines. In particular, the (E)-2-(4-methoxyphenylmethylene)-1-benzosuberone (II c) had the most remarkable cytotoxicity, possessing 11 times the potency of the reference drug melphalan when all five screens were considered [
10,
11].
In consecutive publications, we have performed cell cycle analysis of Jurkat cells exposed to IIc and its methyl-substituted counterpart IIb. It was demonstrated that equitoxic doses of the two cyclic chalcone analogs have different effects on the cell cy-cle progression of the investigated Jurkat cells. Compound IIc showed to cause an im-mediate G1 lift and G2/M arrest, followed by hypoploidity and aneuploidy. The re-markable effect of IIb on the G1 and G2 checkpoints could not be observed [
13,
14]. TLC and HPLC analysis showed the compounds to have intrinsic reactivity towards GSH [
13,
15]. However, the two compounds had different effects on the thiol status of the cells. Compound IIc significantly increased the oxidized glutathione (GSSG) level. On the contrary, IIb increased the GSH level, indicating enhanced cellular antioxidant potency [
16].
In its reduced and oxidized forms (GSH, GSSG), glutathione is ubiquitous in mammalian cells ranging in 1–10 mM concentrations [
17]. Under physiological condi-tions, more than 98% of total GSH occurs in reduced form [
18,
19]. The GSH/GSSG re-dox system is crucially participating and has a vital role in maintaining the environ-ment of the intracellular redox system, antioxidant defense system, and cellular sig-naling processes [
20]. Furthermore, it is one of the endogenous substances involved in the metabolism of endogenous (e.g., estrogens, leukotrienes, prostaglandins) and ex-ogenous compounds (e.g., drugs, non-energy-producing xenobiotics) [
21].
Covalent bond formation of GSH with electrophilic species affects the half-cell reduction potential of the GSSG/2GSH redox system. The GSH/GSSG ratio is a critical mechanism for cell survival; in fact, it is known that it varies in association with pro-liferation, differentiation, and apoptosis [
22,
23]. In our earlier publication [
24], we re-ported on the thiol reactivity of two open-chain chalcones (Ib and IIc) with different cancer cell cytotoxicities [
25]. We could not find a direct correlation of the thiol (GSH and NAC)) reactivities and the previously published biological (cancer cell cytotoxic) effects of the two chalcones. Continuing the previous studies on the molecular mecha-nism of the cancer cell cytotoxic and cell cycle modulating effects of IIb and IIc, we report on a comparative HPLC study on their intrinsic reactivity towards GSH and NAC. Compound IIc showed IC50 values towards most investigated cancer cell lines close to two magnitudes lower than the 4-methyl analog IIb (
Table 2). Thus, the dif-ferences in the reactivities could reflect the differences in their previously published biological activities.
Table 1.
IC50 (μM) data of selected E-2-(4’-X-benzylidene)-1-benzosuberones (II).
Table 1.
IC50 (μM) data of selected E-2-(4’-X-benzylidene)-1-benzosuberones (II).
Compound |
P388 |
L1210 |
Molt 4/C8 |
CEM |
Human tumor cells |
IIa |
12.7 |
106.0 |
42.7 |
28.9 |
18.6 |
IIb |
11.8 |
25.0 |
21.3 |
11.4 |
11.2 |
IIc |
1.6 |
0.34 |
0.47 |
0.35 |
0.27 |
Similar to the previous publication [
24], the reactions were studied under three conditions with different pH: (a) pH 8.0/7.4, (b) pH 6.3/6.8, and (c) pH 3.2/3.7. The first pH values indicate the pH of the aqueous solution of the thiols before starting the incubations. The second pH values indicate the virtual pH of the incubation mixtures, which contained 75.5% v/v methanol (MeOH). The basic pH was selected because such conditions mimic that of the GST-catalyzed reactions; the ionization of the GSH thiol-moiety is increased due to its interaction with the basic imidazole N-atom in the active site of the enzyme [
26]. The slightly acidic pH (pH 6.3) resembles the slightly acidic pH of the cancer cells [
27]. The strong acid conditions (pH 3.2) were selected to compare the reactivity of the protonated and the ionized forms of the thiols function of the two compounds. The pKa of GSH and NAC was reported to be 8.83 and 9.52, respectively. Accordingly, the thiol function of both compounds exists exclusively in the protonated (neutral) form [
28].
Figure 1.
Structure and numbering of 4-X-chalcones (I) and (E)-2-(4’-X-phenylmethylene-1-benzosuberones (II).
Figure 1.
Structure and numbering of 4-X-chalcones (I) and (E)-2-(4’-X-phenylmethylene-1-benzosuberones (II).
The thiol additions to enones are reported to be reversible, resulting in an equilib-rium mixture's formation. To qualitatively characterize the progress of the reactions, the composition of the incubation mixtures was analyzed at the 15, 45, 75, 105, 135, 165, 195, 225, 255, 285, and 315 min time points by HPLC-UV-VIS. Furthermore, den-sity functional theory (DFT) calculations and machine learning (ML) protocols were used to analyze the stability and regioselectivity of chalcone analogs on a structural basis. In the analyses, methanethiol (CH3SH) and its deprotonated form (CH3S-) were used as model thiols.
3. Discussion
Our experiment showed that both cyclic chalcone analogs (
IIb and
IIc) have intrinsic reactivity with GSH and NAC under all three experimental conditions. The results strengthen the results of our previous studies obtained by TLC analysis of similar incubations with GSH of the two compounds [
13]. Considering the pKa values of GSH (8.83) and NAC (9.52) thiols, it can be seen that the fraction of the stronger nucleophile thiolate form of GSH is higher than that of NAC under each experimental condition. Under the slightly basic conditions (pH 8.0/7.4), the rate of reduction of the HPLC peak area of the starting chalcones showed a linear decrease. Since the area is based on the absorbance (logarithmic function of the concentration) of the compounds, the reactions follow pseudo-first-order kinetics. In the case of both thiols, relatively high amounts of (Z)-chalcone isomers could be detected in the incubations (
Table 2 and
Table 3). Since the reaction mixtures were incubated in the dark, the corresponding retro-Michael reactions are the only source of the (Z)-isomer formations. Accordingly, the progression curves of the incubations (
Figure 2 and
Figure 3) reflect the disappearance of the starting compounds due to the net change of the reversible reactions. Similar levels of the respective (Z)-isomers could be detected in the incubations performed under slightly acidic (pH 6.3/6.7) conditions (
Table 2 and
Table 3). On the contrary, the reactions of the respective open-chain chalcones (
Ib and
Ic) performed under identical conditions did not result in a detectable level of (
Z)-isomers in the GSH or the NAC incubations [
24].
The results obtained in the pH 6.3/6.7 incubations are similar to those of the pH 8.0/7.4 ones (
Table 2 and
Table 3). Under such conditions, however, the composition of both incubations represents equilibrium mixtures. Similar to the pH 8.0/7.4 incubations, the conversion of
IIb is somewhat higher in the case of both thiols. The observation further strengthens the previously suggested view that the different reactivities can be (at least partly) the result of the different stability of the thiol adducts [
24,
30]. Similar to the results obtained under identical conditions with the respective open-chain chalcones (
I), the 4-methyl-substituted derivative (
IIb) forms the more stable adducts.
Comparing the compositions of the 315 min incubation mixtures of the two series (
I and
II), it can be seen that the conversions of the 4-CH
3- and 4-OCH
3-substituted chalcones (
Ib and
Ic) are much higher than those of the
IIb and
IIc (
Table 6).
13C NMR shifts, indicating the electron density around the particular nucleus, of the β-carbon atom of
IIb (138.0 ppm) and
IIc (137.7 ppm), were reported to be similar. A similar slight (0.3 ppm) difference was observed in the case of the respective open-chain chalcones
Ib and
Ic [
31]. Since the nature of thiols and the aromatic substituents are the same, the ring structure can explain the observed differences in reactivities of the two series.
Amslinger et al. investigated the thiol reactivity of chalcones with various substituents in their α-position. The kinetics of thiol reactivities of the derivatives were correlated with some of their biological effects directly connected to their Michael acceptor ability [
32,
33]. For example, α-methyl substitution of 2',3,4,4’-tetramethoxychalcone (TMC) decreased, α-cyano substitution substantially increased the thiol reactivity of the nonsubstituted TMC [
34]. Based on these earlier observations, it is reasonable to suppose that the reduced reactivity of the benzosuberone derivatives
IIb and
IIc is the consequence of added effects of the α-alkyl substitution and the conformational strain caused by the cyclic structure of the starting enone and the reaction intermediate. Further research is needed to characterize the electronic and stereochemical effects of ring numerically.
As a result of the addition reactions, the formation of four diastereomeric adducts is possible. Because of the inherent chirality of GSH and NAC, two
cis adducts and two
trans adducts are expected to be formed. Earlier, Armstrong et al. reported on the stereochemistry of the GSTM 4-4-catalyzed reaction of GSH and the open-chain chalcone analog (
E)-(4’-X-phenyl)-3-butene-2-ones (PBO). In the reactions, a higher amount of the more polar adducts were formed [
35]. Based on the results of HPLC separation of the diastereomeric pairs of the PBO-GSH [
35] and the
I-GSH [
24] adducts, we can presume that the two separated peaks formed in the present reactions correspond to the diastereomeric
cis and
trans adducts.
The ratio of the area of the two separated peaks in the GSH incubates (315 min timepoint) was close to the unity for
IIb and
IIc under both pH (pH 8.0/7.4 and 6.3/6.8) conditions (
Table 2). Similar to our previous results, higher peak areas of the least polar adducts were observed in each case. On the contrary, HPLC analysis of the reactions of
IIb and
IIc with NAC showed different (1.8-8.57 times) excess of the least polar diastereomer (
Table 3). Similar to the previous results obtained with the open-chain chalcones (
Ib and
Ic) [
24], the observed diastereoselectivity was affected by the nature of the 4-substituent and the pH. Thus, the methyl–substituted
IIb showed higher diastereoselectivity at both pH values. Diastereoselectivity was increased as the pH was reduced (
Table 3). It is worth mentioning, however, that the observed diastereoselectivities do not reflect the diastereoselectivity of the addition reactions. Under both conditions (pH 8.0/7.4 and 6.3/6.8), the peak areas of the (
Z)-isomers and the adducts are comparable (
Table 3). Since the retro-Michael reaction is the only source of formation of the (
Z)-isomers, the observed ratios reflect the actual balance of the kinetic and thermodynamic controls.
Under the acid conditions (pH 3.2/3.8), the formation of the respective conjugates is exclusively due to the nucleophilic addition of the protonated thiol forms onto the polarized carbon-carbon double bounds. In comparison of the respective compositions of the GSH incubates with those of the previously reported (open chain) chalcones (
Ib and
Ic) [
24], the derivatives with the same substituent showed similar GSH reactivities (
Table 2). However, different results were obtained in the case of the reactions with NAC. The 315-minute percent conversion was found to be higher for
IIb (23.7%) and
IIc (12.1%) than those of the corresponding open-chain chalcones
Ib and
Ic (10.9% and 1.5%, respectively) (
Table 5). However, no
II-NAC adducts could be identified in the HPLC-UV chromatograms. Instead, several small, unidentified peaks appeared (
Figures S15 and S16). HPLC-MS analysis could identify the expected conjugates. The structural characterization of the other products is out of the scope of the present work.
To obtain physicochemical properties insights into different reactivities of chalcones (
I) [
24] and their seven-membered cyclic analogs (
II), HOMO and LUMO molecular orbital energy and some electrophilic reactivity parameters of
Ia,
IIa, - and as model thiols –
CH3SH and
CH3S- were calculated (
Table 5). According to the Hard and Soft, Acids and Bases (HSAB) theory [
36], nucleophilic-electrophilic reactions occur preferably between electrophiles and nucleophiles of similar hardness or softness. In the case of the α,ß-unsaturated ketone, the carbonyl oxygen atom withdraws electrons from the C
2=C
10 bond – generating an electron deficiency at C
10 – the most likely site to receive nucleophilic attacks. In methanethiol, the nucleophilic attacks can occur at the sulfur atom. In compounds
Ia and
IIa, the carbonyl O has a high negative charge density, indicating its Lewis base behavior. On the other hand, regions of lower charge density, which appear in blue, indicate the Lewis acid behavior of the molecules.
The LUMO energy showed that
Ia (-35.98 kcal/mol) is more acidic than
IIa (-28.44 kcal/mol. The LUMO energy of
CH3SH is (-2.979 kcal/mol), which increases to (77.99 kcal/mol) in its deprotonated form (
CH3S-). These characters are also reflected by all the other determined parameters (
Table 5). Therefore, molecular orbital calculations provided data to support the experimental findings. The equilibrium (close-to-equilibrium) compositions of
Ib and
Ic show a higher product ratio than the cyclic chalcone analog
IIb and
IIc.
4. Materials and Methods
4.1. Chemicals and Reagents
Chalcones
IIb and
IIc were synthesized as previously published [
10]. Their structures were characterized by IR and NMR spectroscopy [
37]. The purity and structures of the investigated samples were verified by HPLC-MS (
Figures S1 and S2). Reduced
l-glutathione, N-acetyl l-cysteine, HPLC, and MS-grade methanol solvent were obtained from Sigma-Aldrich (Budapest, Hungary). Trifluoroacetic acid HiperSolve CHROMANORM and formic acid were obtained from VWR (Budapest, Hungary) and Fischer Chemicals, respectively. Deionized water for use in HPLC and HPLC-MS measurements was purified by Millipore Direct-Q
TM at the Institute of Pharmaceutical Chemistry (University of Pécs). Mobile phases used for HPLC measurements were degassed by an ultrasonic water bath before use.
4.2. Preparation of Solutions
The thiol solutions (reduced glutathione (GSH) and N-acetylcysteine (NAC)) preparation were as follows: 2.0 × 10−1 mol·L−1 (0.3 mmol) of the respective thiol was dissolved in water, and the pH was set to either 3.2, 6.3, or 8.0 using 1M NaOH solution to a final volume of 1.5 mL (solution-1). The chalcone solution consisted of 6.5 × 10−3 mol·L−1 (0.03 mmol) chalcone analog dissolved in 4.6 mL HPLC-grade methanol (solution-2). Solution-1 and solution-2 were mixed to give a final volume of 6.1 mL. The molar ratio of thiol to chalcone in the mixture was 10:1. The mixture was kept in the dark, 37 °C water bath for 315 minutes. The first sample was taken at 15 minutes, and onward samples were taken at every 30 minutes time points (11 samples in total).
To evaluate the initial (0 min) peak area of chalcones
1 and
2, solution-2 was prepared without any change, while solution-1 was prepared without the thiol component. Before mixing, the solutions were pre-incubated at 37 °C for 30 min to mimic the incubation conditions. To compare the products of the previously proven light-initiated
E/Z isomerization of the parent compounds [
27] with those of the non-light (retro-Michael addition)-initiated isomerization, solution-2 of the respective chalcones were prepared and exposed to the unscattered laboratory light for 1 week. The solutions were analyzed by HPLC-UV-VIS and HPLC-MS. (
Figures S17 and S18).
4.3. RP-HPLC-UV-VIS Measurements
UV–VIS detector coupled Agilent 1100 HPLC system analyzed the samples at 260 nm wavelength. The separation system was a reversed-phase chromatographic system, and the column Zorbax Eclipse XBD-C8 (150 mm × 4.6 mm, particle size 5 µm; Agilent Technologies, Waldbronn, Germany) was used. The oven temperature was set to 25 °C to avoid room temperature fluctuations. The injection volume was 10 µL. At a 1.2 mL/min flow rate, gradient elution was performed by (A) water and 0.1% trifluoroacetic acid and (B) methanol and 0.1% trifluoroacetic acid. The elution profile consisted of 8 min of 40% isocratically, an increase to 60% B in 4 min, and a further linear increase of eluent B to 90% in 3 minutes. The elution gradient remained constant for 5 min period. Then it was linearly decreased to the initial 40% in 2 min, followed by a 3 min constant of 40% of eluent B for equilibration of the column.
4.4. HPLC-MS Measurements
HPLC ESI-MS analyses were performed on an Ultimate 3000 liquid chromatograph (Dionex, Sunnyvale, CA, USA) coupled with a Thermo Q Exactive Focus quadrupole-Orbitrap hybrid mass spectrometer (Thermo Fisher Scientific, Waltham, MA, USA). The scan monitored m/z values ranging from 100 to 1000 Da. Data acquisition was carried out using Q Exactive Focus 2.1 and Xcalibur 4.2 software (Thermo Fisher Scientific). Analysis of compounds and adducts was performed in HESI positive and negative ionization modes with the following parameters: spray voltage, 3500 V; vaporizer temperature, 300 °C; capillary temperature, 350 °C; spray and auxiliary gas flows, 30 and 10 arbitrary units, respectively; resolution, 35,000 at 200 m/z; and fragmentation, 20 eV.
HPLC separation was performed on an Accucore C18 column (150 mm × 2.1 mm, particle size 2.6 µm), and an Accucore C18 guard column (5 mm × 2.1 mm, particle size 2.6 µm) was also used. The injection volume was 5 µL; the flow rate was 0.4 mL/min. Data analysis and evaluations were performed using Xcalibur 4.2 and FreeStyle 1.7 software. A binary gradient of eluents was used, consisting of mobile phases A and B.
The gradient parameters in chalcones were (A) water and 0.1% formic acid and (B) methanol and 0.1% formic acid. The gradient elution was as follows: isocratic elution for 1 min to 20% eluent B, continued by a linear gradient to 100% in 9 min, followed by an isocratic plateau for 2 min. Then, the column was equilibrated back to 20% in 0.5 min and continued isocratically for 2.5 min. The sampler was at room temperature, and the column oven was at 40 °C.
The parameters of the gradient in the case of adducts were (A) water and 0.1% formic acid and (B) methanol and 0.1% formic acid. The gradient elution was as follows: isocratic elution for 1 min to 10% eluent B, continued by a linear gradient to 95% in 13 min, followed by an isocratic plateau for 3 min. Finally, the column was equilibrated to 10% in 0.1 min and continued isocratically for 2.9 min. The sampler was at room temperature, and the column oven was at 40 °C. The diode array detector was also set at 260 nm wavelength alongside MS analysis.
4.5. Molecular modeling analysis
The structures
Ia, IIa, CH3SH, and
CH3S- were constructed using the Gaussview 6.0 software. Theoretical calculations were performed by DFT [
38,
39], implemented in the G16 [
40] software package. The molecules were optimized using the hybrid exchange and correlation functional with long-range correction, M06-2X [
41], combined with the basis set 6-311++G(d,p) in the gas phase. Frontier molecular orbitals (FMO) [
42] were obtained. Molecular electrostatic potential maps contributed to the global electrophilicity analysis through their electronic isodensity surfaces. MEP [
43] maps provide a visual representation of the electrostatic potential on the surface of a molecule, which can reveal regions of high and low electron density. The electrostatic potential V(
r) [
44] at point
r is defined as.
where Z
Ais the charge of nuclei
at point
and
is the charge density at point
r. The local electrophilicity of the molecules was determined by the Fukui function [
45,
46], and then it was possible to predict the molecular site selectivity.
where
is the number of electrons in the system, and the constant term
in the partial derivative is external potential. Multiwfn 3.6 program [
47] was used to calculate the Fukui. In addition, the
pySiRC [
48] – a machine-learning computational platform, a machine-learning computational platform, was used to simulate oxidation reactions facilitated by free-radical compounds. To imitate the oxidation impact induced by a radical attack, the hydroxyl radical (˙OH) was chosen as the archetype system of degradation reactions. The reaction rate constant of the oxidative attack caused by the hydroxyl radical on chalcones compounds was predicted using the XGBoost ML algorithm and the MACCS fingerprint was employed as a structural descriptor.