Next Article in Journal
Identification of Potential Parkinson’s Disease Drugs Based on Multi-Source Data Fusion and Convolutional Neural Network
Next Article in Special Issue
Hepatoprotective Effect of Millettia dielsiana: In Vitro and In Silico Study
Previous Article in Journal
Metabolic Engineering of Shikimic Acid Biosynthesis Pathway for the Production of Shikimic Acid and Its Branched Products in Microorganisms: Advances and Prospects
Previous Article in Special Issue
Homology Modeling and Molecular Docking Approaches for the Proposal of Novel Insecticides against the African Malaria Mosquito (Anopheles gambiae)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ultrasound-Assisted Synthesis and In Silico Modeling of Methanesulfonyl-Piperazine-Based Dithiocarbamates as Potential Anticancer, Thrombolytic, and Hemolytic Structural Motifs

1
Department of Chemistry, Government College University Faisalabad, Faisalabad 38000, Pakistan
2
Department of Zoology, Government College University Faisalabad, Faisalabad 38000, Pakistan
3
Department of Biochemistry, Government College University Faisalabad, Faisalabad 38000, Pakistan
4
Department of Pharmacology, Government College University Faisalabad, Faisalabad 38000, Pakistan
5
Cornea Research Chair, Department of Optometry, College of Applied Medical Sciences, King Saud University, Riyadh 11433, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(15), 4776; https://doi.org/10.3390/molecules27154776
Submission received: 1 July 2022 / Revised: 22 July 2022 / Accepted: 25 July 2022 / Published: 26 July 2022
(This article belongs to the Special Issue In Silico Methods Applied in Drug and Pesticide Discovery)

Abstract

:
Piperazine-based dithiocarbamates serve as important scaffolds for numerous pharmacologically active drugs. The current study investigates the design and synthesis of a series of dithiocarbamates with a piperazine unit as well as their biological activities. Under ultrasound conditions, the corresponding piperazine-1-carbodithioates 5a5j were synthesized from monosubstituted piperazine 2 and N-phenylacetamides 4a4j in the presence of sodium acetate and carbon disulfide in methanol. The structures of the newly synthesized piperazines were confirmed, and their anti-lung carcinoma effects were evaluated. A cytotoxic assay was performed to assess the hemolytic and thrombolytic potential of the synthesized piperazines 5a5j. The types of substituents on the aryl ring were found to affect the anticancer activity of piperazines 5a5j. Piperazines containing 2-chlorophenyl (5b; cell viability = 25.11 ± 2.49) and 2,4-dimethylphenyl (5i; cell viability = 25.31 ± 3.62) moieties demonstrated the most potent antiproliferative activity. On the other hand, piperazines containing 3,4-dichlorophenyl (5d; 0.1%) and 3,4-dimethylphenyl (5j; 0.1%) rings demonstrated the least cytotoxicity. The piperazine with the 2,5-dimethoxyphenyl moiety (5h; 60.2%) showed the best thrombolytic effect. To determine the mode of binding, in silico modeling of the most potent piperazine (i.e., 5b) was performed, and the results were in accordance with those of antiproliferation. It exhibits a similar binding affinity to PQ10 and an efficient conformational alignment with the lipophilic site of PDE10A conserved for PQ10A.

Graphical Abstract

1. Introduction

Lung carcinoma is one of the most well-known causes of cancer death in many countries [1,2]. The annual death rate from lung carcinoma is double that of all other tumors combined [3,4]. The incomplete development of early detection methods is largely responsible for lung carcinoma’s uncontrolled spread [5]. Smoking is the leading cause of lung carcinoma worldwide [6]. Early diagnosis of other tumors, such as colon carcinoma, cervical carcinoma, and esophageal carcinoma, increases the survival rate [7]. It is vital to develop novel drug candidates for this type of carcinoma [8].
Structurally modified nitrogen heterocycles have a broad spectrum of biological activities [9]. For example, heterocycles containing the piperazine moiety have a significant role in drug discovery programs [10]. Due to their vast array of pharmaceutical applications, N-substituted piperazines have garnered attention in academia and the pharmaceutical industry. They exhibit antimicrobial [11], antimycobacterial [12], antidepressant [13], and anticancer [14,15] properties. Previous reports revealed that N-substituted piperazines are a crucial structural component of the anticancer effect [16,17]. The presence of polar nitrogen in the piperazine framework is an essential component of the skeletons of numerous biologically active drugs [18]. Figure 1 shows the structures of some common piperazine-based drugs.
Dithiocarbamates are important intermediates in the synthesis of many pharmacologically active molecules [19,20]. They are active ingredients in numerous drugs (Figure 2), as well as naturally occurring and synthesized anticancer agents [21,22]. Including a dithiocarbamate moiety in the skeleton of different organic compounds elevates the biological profile of the synthesized derivatives [23].
The incorporation of two or more biologically active molecules to create new hybrid structures with potential medicinal applications is a useful approach [24]. This strategy is adopted in an effort to enhance the biological profile of newly synthesized hybrids compared to individual molecules included in their skeletons [25,26]. In continuation of our research on hybrid molecules with cytotoxic potential, we now report the design, synthesis, and antiproliferative effect of piperazine-based acetamides against lung carcinoma (A-549).

2. Results and Discussion

2.1. Chemistry

Several piperazine-dithiocarbamate hybrids 5a5j were synthesized, as shown in Scheme 1. The treatment of excess piperazine (1) with methane sulfonyl chloride at 0 °C in dichloromethane (DCM) produced monosubstituted piperazine 2 in an 80% yield [27]. Treating aryl amines 3a3j in a basic medium with bromoacetyl bromide afforded the corresponding N-phenyl acetamides 4a4j [28,29]. Reactions of 2 and 4a4j in the presence of sodium acetate (AcONa) and carbon disulfide (CS2) in methanol (MeOH) for 12–18 h at room temperature produced 5a5j in 55–65% yields. Piperazine-1-carbodithioates 5a5j were obtained in excellent yields (80–90%) when reactions of 2 and 4a4j were carried out under ultrasonic conditions at 70 °C for 30 min.
The structures of piperazine-1-carbodithioates 5a5j were established using a variety of spectroscopic techniques, and their purity was confirmed by elemental analyses (see the experimental section for details). The NMR spectra of 5a5j are shown in Figures S1–S20. The HRMS confirms the chemical formula of each derivative’s molecular ion peak. The FTIR spectra showed the presence of absorption bands corresponding to the NH (3230–3330 cm−1), C=O (1670–1660 cm−1), C=C (1530–1520 cm−1), and C=S (1225–1220 cm−1) groups. The NMR spectra of 5a5j contained all of the expected protons and carbons. At high fields, the 1HNMR spectra of 5a5j revealed the presence of methyl protons of the sulfonyl group (2.85–2.79 ppm) and methylene protons (4.36–4.20 ppm), as well as piperazine and aryl protons. At very low fields, the 13CNMR spectra revealed the presence of C=S (197.4–196.6 ppm) and C=O (167.0–163.9 ppm) carbons.

2.2. Antiproliferative Potential

The anticancer potential of newly synthesized piperazine-1-carbodithioates 5a5j against the human lung cancer cell line (A-549) was investigated using the MTT assay [30]. The cell viability of 5a5j varied depending on the aryl ring substituents (Table 1).
The SAR analysis of 5a5j was carried out to determine the role of various substituents on the aryl ring. Compounds 5b and 5i produced the best results, as shown in Table 1. The most potent derivative was compound 5b, which contains a chlorine substituent at the ortho-position of the aryl ring and has a cell viability of 25.11 ± 2.49. The cell viability of compound 5i, which contains two methyl groups at the ortho and para positions of the aryl ring, was comparable to that of compound 5b, at 25.31 ± 3.62. Compounds 5d, 5f, and 5h exhibited a significant effect, but it was lower than 5b and 5i. Compound 5e, with a fluorine atom at the ortho-position of the aryl ring, showed the lowest potency and highest cell viability (68.94 ± 6.64), followed by 5c (60.29 ± 5.96), which has a chlorine substituent at the para-position of the aryl ring.
The cell viability of 5b and 5i was investigated further using various concentrations (0.3–200 µg) to test the dose-response relationship. Figure 3 and Figure 4 show that 5b and 5i produced the best results at a dose of 200 µg.

2.3. Hemolytic Potential

The cytotoxic activity of 5a5j was investigated using a reported methodology [31]. The age hemolysis (%) shown in Table 1 indicated that 5a5j had a low cytotoxic profile. The age hemolysis (%) for compound 5c was 10%, demonstrating its binding affinity for hemoglobin, followed by 5h (7%). Compounds 5a (3.1%), 5b (1.27%), 5f (2.8%), 5g (1.3%), and 5i (2.8%) exhibited low toxicity. Compounds 5e (0.3%), 5d, and 5j (0.1%) had the lowest cytotoxicity when compared to ABTS (95.9%).

2.4. Thrombolytic Potential

Cerebral venous sinus thrombosis (CVST) is a common central nervous system disorder caused by thrombophilia syndrome. Due to its anticoagulant profile, heparin is regarded as an effective treatment for CVST. Many thrombolytic agents are now used to treat thrombophilia syndrome. A reported methodology was used to evaluate compounds 5a5j for thrombolysis [31]. Compound 5h, which contains a methoxy group at the ortho-position of the aryl ring, showed the highest potential (60.2%). Compounds 5a (58.9%), 5i (57.4%), and 5j (58.7%) showed moderate thrombolytic potential. Compounds 5c (47.8%), 5f (50.3%), and 5g (50.9%) had a mild effect compared to the positive control (i.e., ABTS; 80%).

2.5. Structure-Activity Relationship (SAR) Study

The effect of aryl ring substituents in compounds 5a5j on anticancer activity against A-549 was investigated. Compound 5b, with Cl at the ortho-position of the aryl ring, exhibited the lowest cell viability and thus proved to be a potent cytotoxic agent. The replacement of Cl in 5b (cell viability = 25.11 ± 2.49) with hydrogen (i.e., 5a; cell viability = 61.35 ± 2.29) or fluorine (i.e., 5e; cell viability = 68.94 ± 6.64) resulted in a significant decrease in cytotoxic potential. However, the substitution of Cl in 5b with a fluorine atom at the para-position of the aryl ring (i.e., 5f; cell viability = 41.01 ± 3.73) or a chlorine atom at the para-position (i.e., 5c; cell viability = 60.29 ± 5.96) led to a sharp decrease in anticancer activity. In contrast, the insertion of two Cl at the meta and para-positions of the aryl ring (i.e., 5d; cell viability = 38.08 ± 2.85) enhances anticancer activity. Inserting a methoxy group at the para-position of the aryl ring (i.e., 5g; cell viability = 43.36 ± 4.42), on the other hand, led to a significant change in anticancer activity. The two methoxy substituents at the ortho and meta-positions in compound 5h (cell viability = 40.25 ± 3.34) are responsible for the good antiproliferative potential when compared to 5e, which has a methoxy group at the ortho-position of the aryl ring. However, the replacement of the Cl in 5b with two methoxy groups at the ortho and para-positions of the aryl ring (i.e., 5i; cell viability = 25.31 ± 3.62) resulted in better activity when compared to the case where two methyl groups were present at the meta and para-positions of the aryl ring (i.e., 5j; cell viability = 47.94 ± 1.57).
The trend for SAR studies related to hemolysis (Table 1) suggested that the presence of the chlorine atom at the para-position of the aryl ring (i.e., 5c) showed the highest cytotoxic potential (10%) among the others. Compounds 5d and 5j exhibited the lowest cytotoxic potential (0.1%) for hemolysis, implying that the Cl and OMe substituents at the ortho and para-positions of the aryl ring have no affinity for hemoglobin. When compared to 5d and 5j, the two methoxy substituents at the ortho and meta-positions of the aryl ring (i.e., 5h; 7.1%) increased the cytotoxic activity. However, the unsubstituted phenyl ring (i.e., 5a) has 30 times more cytotoxic activity (3.1 %).
Based on the data from the SAR study of piperazine-1-carbodithioates 5a5j, it is concluded that the presence of the Cl motif at the ortho-position of the aryl ring played a significant role in the antiproliferative activity. The presence of chlorine, dimethyl, and methoxy substituents improves anticancer activity (Figure 5).

2.6. In Silico Modeling

The mode of action of compound 5b was investigated further in silico. The SwissTargetPrediction identified phosphodiesterase 10A (PDE10A) as a potential target for 5b with the highest (~1600) active similarity and 0.105 probability. The IFD method was validated by cognate re-docking, which resulted in a conformation with 0.2519 Å RMSD compared to the co-crystallized native conformation (Figure 6).
The IFD was carried out to investigate the binding of 5b with its potential therapeutic target and simulate its binding conformation into the catalytic site. The binding free energy analysis revealed that compound 5b binds at the catalytic site of PDE10A with ΔG of −8.30 kcal/mol which is comparable to −8.97 kcal/mol for the PQ-10 as a standard for PDE10A (Table 2).
The conformational analysis of compound 5b highlighted its efficiency in attaining a binding pose that established the diverse interactions with the vital residues within the catalytic pocket of PDE10A (Figure 7). Interestingly, 5b shares similar interaction patterns to PQ-10. It is efficiently penetrated and anchored into the lipophilic pocket of PDE10A within the catalytic site which is inaccessible in all other isoforms of PDE. Therefore, the comparable binding behavior and affinity may suggest that 5b exhibits a similar activity compared to PQ-10.
The conformational analysis was extended to delineate the complexation or anchoring of ligands within the binding pocket of PDE10A. The PQ-10 established H-bonds with TYR683, GLN716, VAL712, ASP664, and TYR514 to strengthen the stabilization of its confirmation within the catalytic site of PDE10A (Figure 8). This core stabilization was further supported by its extensive hydrophobic interactions with GLU711, PRO702, ILE682, PHE719, and ALA679 within the lipophobic cavity of PDE10A. Notably, 5b interacts with conserved PQ10 residues. Compound 5b’s conformation is stabilized by H-bonds with TYR683, TYR514, and SER667. The hydrophobic interactions between GLU711, MET703, PRO703, and VAL712 further stabilized this conformation. Furthermore, 5b does not interact with GLN726 conserved in other PDE isoforms, resulting in greater selectivity for PDE10A.
Compound 5b exhibits a binding affinity comparable to that of PQ10 and an efficient conformational alignment with the lipophilic site of PDE10A conserved for PQ10A. In addition, 5b establishes interactions with the conserved PQ10 residues, but with a distinct bonding pattern, implying potential differences in activity and selectivity. This study provides valuable insight into the potential therapeutic activity of 5b against PDE10A.

3. Materials and Methods

3.1. General

Chemicals, solvents, and reagents were purchased from Merck (Gillingham, UK) and used without further purification. Melting points were recorded on a Gallenkamp instrument (Fisons; Uckfield, UK). The 1H (400 MHz) and 13C NMR (100 MHz) spectra were recorded in deuterated chloroform (CDCl3) using a Bruker DPX spectrophotometer (Bruker; Zürich, Switzerland). The chemical shifts were recorded in ppm related to either tetramethylsilane or CDCl3. Thin-layer chromatography (DCM/MeOH) was utilized in conjunction with a Spectroline E-Series UV lamp to monitor the progression of chemical reactions (Merck; Gillingham, UK). Compound 2 [27] and N-phenyl acetamides 4a4j [28,29] were produced based on reported procedures.

3.2. General Procedures for the Synthesis of Piperazine-1-carbodithioates 5a5j

3.2.1. Conventional Method

A solution of appropriate N-phenylacetamides 4a4j (0.6 mmol) in dry MeOH (0.3 mL) was added to a stirred mixture of 2 (0.10 g, 0.6 mmol), anhydrous AcONa (0.36 g, 0.60 mmol), and CS2 (0.39 g, 0.6 mmol) in dry MeOH (0.8 mL). The mixture was stirred at 20 °C for 12–18 h according to the TLC. The produced solid was collected, washed with H2O, dried, and recrystallized from EtOH to give pure 5a5j as white solids in moderate to good yields (55–65%).

3.2.2. Ultrasound-Assisted Method

The procedure was identical to that described in Section 3.2.1. with the exception that the mixture was sonicated at 70 °C for 30 min. Following crystallization, pure 5a5j was obtained in excellent yields (80–90%) as white solids.

3.2.3. 2-Oxo-2-(phenylamino)ethyl 4-(methylsulfonyl)piperazine-1-carbodithioate (5a)

Molecules 27 04776 i001
Yield 90%, m.p. 188–189 °C. IR: ν (cm−1): 3230 (NH), 1664 (C=O), 1524 (C=C), 1473 (CH2), 1222 (C=S). 1H NMR (δ): 8.79 (s, 1H, NH), 7.48 (d, 2H, J = 8.0 Hz, Ar), 7.29 (t, 2H, J = 8.0 Hz, Ar), 7.08 (t, 1H, J = 8.0 Hz, Ar), 4.48 (br s, 2H, piperazinyl), 4.22 (s, 2H, SCH2), 4.10 (br s, 2H, piperazinyl), 3.36 (t, 4H, J = 4.0 Hz, piperazinyl), 2.79 (s, 3H, Me). 13C NMR (δ): 197.4 (C=S), 166.8 (C=O), 137.8, 129.6, 124.7, 120.1, 45.7 (NCH2), 45.4 (NCH2), 40.7 (SMe), 35.5 (SCH2). HRMS: m/z [M]+ calculated for C14H19N3O3S3: 373.0589; found: 373.0582. Analysis calculated for C14H19N3O3S3: C, 45.02; H, 5.13; N, 11.25; found: C, 45.07; H, 5.17; N, 11.28%.

3.2.4. 2-((2-Chlorophenyl)amino)-2-oxoethyl 4-(methylsulfonyl)piperazine-1-carbodithioate (5b)

Molecules 27 04776 i002
Yield 82%, m.p. 180–181 °C. IR: ν (cm−1): 3330 (NH), 1662 (C=O), 1524 (C=C), 1470 (CH2), 1220 (C=S). 1H NMR (δ): 8.80 (s, 1H, NH), 8.36 (d, 1H, J = 8.0 Hz, Ar), 7.37 (d, 1H, J = 8.0 Hz, Ar), 7.28 (t, 1H, J = 8.0 Hz, Ar), 7.06 (t, 1H, J = 8.0 Hz, Ar), 4.55 (br s, 2H, piperazinyl), 4.32 (s, 2H, SCH2), 4.24 (br s, 2H, piperazinyl), 3.14 (t, 4H, J = 4.0 Hz, piperazinyl), 2.84 (s, 3H, SMe). 13C NMR (δ): 197.4 (C=S), 166.7 (C=O), 134.9, 129.5, 128.3, 125.2, 123.5, 121.8, 50.0 (NCH2), 45.3 (NCH2), 40.9 (SMe), 35.3 (SCH2). HRMS: m/z [M]+ calculated for C14H18ClN3O3S3: 407.0199; found: 407.0209. Analysis calculated for C14H18ClN3O3S3: C, 41.22; H, 4.45; N, 10.30; found: C, 41.26; H, 4.48; N, 10.35%.

3.2.5. 2-((4-Chlorophenyl)amino)-2-oxoethyl 4-(methylsulfonyl)piperazine-1-carbodithioate (5c)

Molecules 27 04776 i003
Yield 87%, m.p. 200–201 °C. IR: ν (cm−1): 3230 (NH), 1662 (C=O), 1525 (C=C), 1470 (CH2), 1225 (C=S). 1HNMR (δ): 8.89 (s, 1H, NH), 7.43 (d, 2H, J = 8.0 Hz, Ar), 7.26 (d, 2H, J = 8.0 Hz, Ar), 4.49 (br s, 2H, piperazinyl), 4.21 (s, 2H, SCH2), 4.97 (br s, 2H, piperazinyl), 3.38 (t, 4H, J = 4.0 Hz, piperazinyl), 2.84 (s, 3H, SMe). 13C NMR (δ): 197.4 (C=S), 167.0 (C=O), 136.1, 129.6, 129.3, 121.4, 45.6 (NCH2), 45.3 (NCH2), 4073 (SMe), 35.4 (SCH2). HRMS: m/z [M]+ calculated for C14H18ClN3O3S3: 407.0199; found: 407.0209. Analysis calculated for C14H18ClN3O3S3: C, 41.22; H, 4.45; N, 10.30; found: C, 41.25; H, 4.46; N, 10.33%.

3.2.6. 2-((3,4-Dichlorophenyl)amino)-2-oxoethyl 4-(methylsulfonyl)piperazine-1-carbodithioate (5d)

Molecules 27 04776 i004
Yield 86%, m.p. 162–163 °C. IR: ν (cm−1): 3235 (NH), 1667 (C=O), 1528 (C=C), 1470 (CH2), 1222 (C=S). 1H NMR (δ): 8.89 (s, 1H, NH), 7.69 (s, 1H, Ar), 7.33 (d, 1H, J = 8.0 Hz, Ar), 7.24 (d, 1H, J = 8.0 Hz, Ar), 4.48 (br s, 2H, piperazinyl), 4.20 (s, 2H, SCH2), 4.12 (br s, 2H, piperazinyl), 3.37 (t, 4H, J = 4.0 Hz, piperazinyl), 2.81 (s, 3H, SMe). 13C NMR (δ): 197.4 (C=S), 166.7 (C=O), 137.4, 132.9, 130.6, 127.7, 121.5, 119.2, 45.6 (NCH2), 45.4 (NCH2), 40.4 (SMe), 35.6 (SCH2). HRMS: m/z [M]+ calculated for C14H17Cl2N3O3S3: 440.9809; found: 440.9812. Analysis calculated for C14H17Cl2N3O3S3: C, 38.01; H, 3.87; N, 9.50; found: C, 38.05; H, 3.87; N, 9.54%.

3.2.7. 2-((2-Fluorophenyl)amino)-2-oxoethyl 4-(methylsulfonyl)piperazine-1-carbodithioate (5e)

Molecules 27 04776 i005
Yield 84%, m.p. 182–183 °C. IR: ν (cm−1): 3250 (NH), 1670 (C=O), 1530 (C=C), 1475 (CH2), 1225 (C=S). 1H NMR (δ): 8.80 (s, 1H, NH), 8.27 (d, 1H, J = 8.0 Hz, Ar), 7.08 (m, 2H, Ar), 7.04 (d, 1H, J = 8.0 Hz, Ar), 4.46 (br s, 2H, piperazinyl), 4.27 (s, 2H, SCH2), 4.10 (br s, 2H, piperazinyl), 3.37 (t, 4H, J = 4.0 Hz, piperazinyl), 2.80 (s, 3H, SMe). 13C NMR (δ): 197.4 (C=S), 166.4 (C=O), 154.2 (d, J = 243 Hz), 151.41, 126.3 (d, J = 10 Hz), 124.8 (d, J = 3 Hz), 121.8, 115.0 (d, J = 8 Hz), 45.6 (NCH2), 45.4 (NCH2), 40.8 (SMe), 35.4 (SCH2). HRMS: m/z [M]+ calculated for C14H18FN3O3S3: 391.0494; found: 391.0496. Analysis calculated for C14H18FN3O3S3: C, 42.95; H, 4.63; N, 10.73; found: C, 42.92; H, 4.67; N, 10.74%.

3.2.8. 2-((4-Fluorophenyl)amino)-2-oxoethyl 4-(methylsulfonyl)piperazine-1-carbodithioate (5f)

Molecules 27 04776 i006
Yield 83%, m.p. 185–186 °C. IR: ν (cm−1): 3330 (NH), 1660 (C=O), 1520 (C=C), 1470 (CH2), 1222 (C=S). 1H NMR (δ): 8.80 (s, 1H, NH), 7.50 (m, 2H, Ar), 7.02 (m, 2H, Ar), 4.52 (br s, 2H, piperazinyl), 4.25 (s, 2H, SCH2), 4.09 (br s, 2H, piperazinyl), 3.40 (t, 4H, J= 4.0 Hz, piperazinyl), 2.85 (s, 3H, SMe). 13C NMR (δ): 197.4 (C=S), 168.3 (C=O),159.4 (d, J = 240 Hz),133.9, 121.6, 116.5, 50.0 (NCH2), 45.7 (NCH2), 40.3 (SMe), 35.3 (SCH2). HRMS: m/z [M]+ calculated for C14H18FN3O3S3: 391.0494; found 391.0496. Analysis calculated for C14H18FN3O3S3: C, 42.95; H, 4.63; N, 10.73; found: C, 42.97; H, 4.65; N, 10.76%.

3.2.9. 2-((4-Methoxyphenyl)amino)-2-oxoethyl 4-(methylsulfonyl)piperazine-1-carbodithioate (5g)

Molecules 27 04776 i007
Yield 80%, m.p. 190 °C. IR: ν (cm−1): 3330 (NH), 1665 (C=O), 1528 (C=C), 1470 (CH2), 1222 (C=S). 1HNMR (δ): 8.63 (s, 1H, NH), 7.40 (d, 2H, J = 8.0 Hz, Ar), 6.83 (d, 2H, J = 8.0 Hz, Ar), 4.46 (br s, 2H, piperazinyl), 4.20 (s, 2H, SCH2), 4.09 (br s, 2H, piperazinyl), 3.76 (s, 3H, OMe), 3.36 (t, 4H, J = 4.0 Hz, piperazinyl), 2.80 (s, 3H, SMe). 13C NMR (δ): 197.4 (C=S), 166.1 (C=O), 157.0, 131.2, 122.3, 114.6, 56.0 (OMe), 50.2 (NCH2), 45.4 (NCH2), 40.7 (SMe), 35.8 (SCH2). HRMS: m/z [M]+ calculated for C15H21N3O4S3: 403.0694; found: 403.0697. Anal Calcd. for C15H21N3O4S3: C, 44.65; H, 5.25; N, 10.41; found: C, 44.67; H, 5.28; N, 10.45%.

3.2.10. 2-((2,5-Dimethoxyphenyl)amino)-2-oxoethyl 4-(methylsulfonyl)piperazine-1-carbodithioate (5h)

Molecules 27 04776 i008
Yield 81%, m.p. 150 °C. IR: ν (cm−1): 3270 (NH), 1668 (C=O), 1524 (C=C), 1470 (CH2), 1222 (C=S). 1H NMR (δ): 8.87 (s, 1H, NH), 8.04 (s, 1H, Ar), 6.75 (d, 1H, J = 8.0 Hz, Ar), 6.54 (d, 1H, J = 8.0 Hz, Ar), 4.36 (br s, 2H, piperazinyl), 4.26 (s, 2H, SCH2), 4.16 (br s, 2H, piperazinyl), 3.81 (s, 3H, OMe), 3.75 (s, 3H, OMe), 3.35 (t, 4H, J = 4.0 Hz, piperazinyl), 2.79 (s, 3H, SMe). 13C NMR (δ): 196.6 (C=S), 165.9 (C=O), 153.9, 142.5, 128.3, 111.2, 109.0, 106.1, 56.6 (OMe), 55.8 (OMe), 45.3 (NCH2), 45.2 (NCH2), 41.3 (SMe), 35.2 (SCH2). HRMS: m/z [M]+ calculated for C16H23N3O5S3:433.0800; found: 433.0805. Analysis calculated for C16H23N3O5S3: C, 44.32; H, 5.35; N, 9.69; found: C, 44.35; H, 5.39; N, 9.73.

3.2.11. 2-((2,4-Dimethylphenyl)amino)-2-oxoethyl 4-(methylsulfonyl)piperazine-1-carbodithioate (5i)

Molecules 27 04776 i009
Yield 87%, m.p. 186–187 °C. IR: ν (cm−1): 3330 (NH), 1665 (C=O), 1525 (C=C), 1475 (CH2), 1220 (C=S). 1H NMR (δ): 8.30 (s, 1H, NH), 7.69 (d, 1H, J = 8.0 Hz, Ar), 6.99 (s, 1H, Ar), 6.96 (d, 1H, J = 8.0 Hz, Ar), 4.47 (br s, 2H, piperazinyl), 4.27 (s, 2H, SCH2), 4.13 (br s, 2H, piperazinyl), 3.35 (t, 4H, J = 4.0 Hz, piperazinyl), 2.80 (s, 3H, SMe), 2.26 (s, 3H, Me), 2.23 (s, 3H, Me). 13C NMR (δ): 197.4 (C=S), 166.6 (C=O), 135.0, 133.2, 131.3, 129.2, 127.3, 122.8, 45.8 (NCH2) 45.4 (NCH2), 40.6 (SMe), 35.4 (SCH2), 21.0 (Me), 18.4 (Me). HRMS: m/z [M]+ calculated for C16H23N3O3S3: 401.0902; found: 401.0906. Analysis calculated for C16H23N3O3S3: C, 47.86; H, 5.77; N, 10.46; found: C, 47.90; H, 5.79; N, 10.48%.

3.2.12. 2-((3,4-Dimethylphenyl)amino)-2-oxoethyl-4-(methylsulfonyl)piperazine-1-carbodithioate (5j)

Molecules 27 04776 i010
Yield 82%, m.p. 170–171 °C. IR: ν (cm−1): 3235 (NH), 1660 (C=O), 1525 (C=C), 1470 (CH2), 1220 (C=S). 1H NMR (δ): 8.65 (s, 1H, NH), 7.28 (s, 1H, Ar), 7.26 (d, 1H, J = 8.0 Hz, Ar), 7.08 (d, 1H, J = 8.0 Hz, Ar), 4.53 (br s, 2H, piperazinyl), 4.26 (s, 2H, SCH2), 4.15 (br s, 2H, piperazinyl), 3.39 (t, 4H, J = 4.0 Hz, piperazinyl), 2.84 (s, 3H, SMe), 2.26 (s, 3H, Me), 2.23 (s, 3H, Me). 13C NMR (δ): 197.4 (C=S), 163.9 (C=O), 137.9, 135.9, 133.2, 129.9, 121.9, 117.1, 50.0 (NCH2), 49.6 (NCH2), 40.9 (SMe), 35.2 (SCH2) 19.9 (Me), 19.2 (Me). HRMS: m/z [M]+ calculated for C16H23N3O3S3: 401.0902; found: 401.0906. Analysis calculated for C16H23N3O3S3: C, 47.86; H, 5.77; N, 10.46; found: C, 47.89; H, 5.77; N, 10.50%.

3.3. Experimental Procedures for Antiproliferative, Hemolytic, and Thrombolytic Activity

3.3.1. Culture and Treatment of Cell Lines

The activity of piperazine-1-carbodithioates 5a5j against human liver carcinoma (A-549) was investigated. Dulbecco’s Modified Eagles Medium (DMEM) was used for the cultural growth of A-549. The DMEM was enriched with fetal bovine serum (10%), streptomycin (100 units/mL), and penicillin (100 μg/mL). The inoculation was carried out in a moistened atmosphere containing carbon dioxide (5%). Compounds 5a5j were dissolved in dimethyl sulfoxide (DMSO). The cell cultures (0.05%) were added to DMSO and employed as a negative control [32,33].

3.3.2. Evaluation of Cell Viability

A-549 cells (100 µL) were cultured overnight in 96-well plates, treated with compounds 5a5j, and incubated (48 h at 37 °C). The MTT reagent (10 µL, 5 mg/mL) was added to each plate and incubated (4 h at 37 °C). In the final step, each cell plate was diluted with DMSO (150 µL), and the absorbance was estimated at 490 nm using a microplate-reader to calculate the cell viability percentage.

3.3.3. Hemolytic Potential

A blood sample (5 mL) from albino mice was centrifuged (1000 rpm for 5 min). The pellet of red blood cells (RBCs) was isolated and washed four times with chilled phosphate buffer saline (pH 7.4). A solution of 5a5j (20 μL; 10 mg/mL) was added to the RBCs pellet (180 μL). The sample was incubated (30 min at 37 °C); removed from the incubator; cooled in an ice bath for 5 min; and centrifuged (13,000 rpm for 5 min). Ice-cold phosphate buffer saline was added to the supernatant (100 µL) in each tube. The DMSO was designated as a negative control, while 2,2′-azino-bis (3-ethylbenzothiazoline-6-sulfonic acid (ABTS)) served as a positive control. The absorbance of the sample was measured at 517 nm, and the age hemolysis (%) was calculated using Equation (1) [34].
age   hemolysis   % = Absorbance   of   sample Absorbance   of   negative   control Absorbance   of   positive   control × 100

3.3.4. Thrombolytic Potential

A blood sample (1 mL) was taken from albino mice and transferred to various rinsed and weighed Eppendorf tubes. The Eppendorf tube sample was incubated (1 h at 37 °C) to form clots. The serum was discarded, and the Eppendorf tubes were weighed to calculate the initial weight of the clot. A solution (100 μL) of 5a5j in DMSO was added separately to the Eppendorf tubes and incubated (3 h at 37 °C). The ABTS was used as a positive control, while water was used as a negative control. The serum was removed again, and the Eppendorf tubes were weighed to measure the clot lysis (%) using Equation (2) [35].
Clot   lysis   % = Initial   clot   weight Final   clot   weight Initial   weight   of   the   clot × 100

3.4. Computational Study

Compound 5b was investigated for its in silico-modeled anticancer activity and potential mechanism of action. A potential target of compound 5b was predicted using the SwissTargetPrediction tool. It employs a ligand-based approach and derives the probability of prediction using 2D and 3D similarity to chemical compounds with known targets [36]. Compound 5b was molecularly docked on a predicted target to investigate its binding affinity, mode of action, and interactions using the induced fit docking (IFD) protocol in the Molecular Operating Environment (MOE) 2015.10. The IFD protocol was validated by cognate redocking of the co-crystalized ligand within the active site, with the root-mean-square deviation (RMSD) serving as the measure of validation. The PubChem CID: 11704101 (PQ-10) is an inhibitor of PDE10A in lung carcinoma and is modeled as a standard molecule in this simulation. The 3D X-ray crystalized structure of phosphodiesterase 10A (PDB ID: 3HR1; 1.53 Å) was retrieved from the RSCB Protein Data Bank (http://www.rscb.org; accessed on 2 February 2022). The inherent structural problems of macromolecules were corrected in the QuickPrep module of MOE. The structure was further protonated and minimized under the Amber 10:EHT forcefield to optimize the molecular mechanics’ refinements and tether restraints in docked poses. The Site Finder module defined the binding pocket in the vicinity of the co-crystallized ligand at the active site. The Dock module was used to dock Compound 5b using the triangular matcher placement method and the London dG scoring function. The docked poses were further refined using the IFD method and the GBVI/WSA dG scoring function. A pose with the lowest binding free energy (ΔG) was investigated further for potential binding affinity, binding pose, and interactions using Discovery Studio Visualizer v17.2 (San Diego, CA, USA).

4. Conclusions

The structures of a new class of N-methylsulfonyl-piperazine-based acetamides have been determined. Their anticancer potential against human lung carcinoma (A-549) was investigated. The types of substituents on the aryl ring within the synthesized piperazines affect the anticancer activity. For example, piperazines containing 2-chlorophenyl and 2,4-dichlorophenyl moieties showed the most potent antiproliferative activity. Piperazines with 2,4-dichlorophenyl and 3,4-dichlorophenyl rings showed the lowest cytotoxicity. The cytotoxic effect of the synthesized piperazines was investigated through hemolysis and thrombolysis. The piperazine with a 2,5-dimethoxylphenyl moiety had the greatest thrombolytic effect. The anticancer activity of the most potent piperazine was investigated further using in silico studies to delineate its potential mechanism of action. More structural modifications to piperazine aryl rings are required to create more potent and selective antiproliferative and cytotoxic candidates.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27154776/s1, Figure S1: 1H NMR spectrum of 5a, Figure S2: 13C NMR spectrum of 5a, Figure S3: 1H NMR spectrum of 5b, Figure S4: 13C NMR spectrum of 5b, Figure S5: 1H NMR spectrum of 5c, Figure S6: 13C NMR spectrum of 5c, Figure S7: 1H NMR spectrum of 5d, Figure S8: 13C NMR spectrum of 5d, Figure S9: 1H NMR spectrum of 5e, Figure S10: 13C NMR spectrum of 5e, Figure S11: 1H NMR spectrum of 5f, Figure S12: 13C NMR spectrum of 5f, Figure S13: 1H NMR spectrum of 5g, Figure S14: 13C NMR spectrum of 5g, Figure S15: 1H NMR spectrum of 5h, Figure S16: 13C NMR spectrum of 5h, Figure S17: 1H NMR spectrum of 5i, Figure S18: 13C NMR spectrum of 5i, Figure S19: 1H NMR spectrum of 5j, and Figure S20: 13C NMR spectrum of 5j.

Author Contributions

Conceptualization and experimental design: F.H., A.F.Z., A.R., A.M., R.N., Z.R., K.G.A., and A.I.; experimental work: F.H., A.F.Z., A.R., A.M., R.N., Z.R., K.G.A., and A.I.; data analysis: F.H., A.F.Z., A.R., A.M., R.N., Z.R., K.G.A., A.I., and G.A.E.-H.; writing—original draft preparation: A.F.Z., A.I., and G.A.E.-H.; writing—review and editing: A.F.Z., A.I., and G.A.E.-H. All authors have read and agreed to the published version of the manuscript.

Funding

The authors extend their appreciation to the Deanship of Scientific Research, King Saud University for funding through the Vice Deanship of Scientific Research Chairs, Research Chair of Cornea.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

F.H. thanks the Higher Education Commission, Pakistan for an indigenous fellowship.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

Sample Availability

Samples of compounds 5 are available from the authors.

References

  1. Alberg, A.J.; Samet, J.M. Epidemiology of lung cancer. Chest 2003, 123, 21S–49S. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Spiro, S.G.; Silvestri, G.A. One hundred years of lung cancer. Am. J. Respir. Crit. Care Med. 2005, 172, 523–529. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Jemal, A.; Murray, T.; Ward, E.; Sammuels, A.; Tewari, R.C.; Ghafoor, A.; Feuer, E.J.; Thun, M.J. Cancer Statistics, 2005. CA Cancer J. Clin. 2005, 55, 10–30. [Google Scholar] [CrossRef] [PubMed]
  4. Travis, W.D. Pathology of lung cancer. Clin. Chest Med. 2002, 23, 65–81. [Google Scholar] [CrossRef]
  5. Parkin, D.M.; Bray, F.I.; Devesa, S.S. Cancer burden in the year 2000. The global picture. Eur. J. Cancer 2001, 37, S4–S66. [Google Scholar] [CrossRef]
  6. Jackman, D.M.; Johnson, B.E. Small-cell lung cancer. Lancet 2005, 366, 1385–1396. [Google Scholar] [CrossRef]
  7. Wistuba, I.I.; Gazdar, A.F. Lung cancer preneoplasia. Annu. Rev. Pathol. Mech. Dis. 2006, 1, 331–348. [Google Scholar] [CrossRef]
  8. Minna, J.D.; Roth, J.A.; Gazdar, A.F. Focus on lung cancer. Cancer Cell 2002, 1, 49–52. [Google Scholar] [CrossRef] [Green Version]
  9. Khan, I.; Ibrar, A.; Abbas, N.; Saeed, A. Recent advances in the structural library of functionalized quinazoline and quinazolinone scaffolds: Synthetic approaches and multifarious applications. Eur. J. Med. Chem. 2014, 76, 193–244. [Google Scholar] [CrossRef]
  10. Jida, M.; Soveidan, M.; Willand, N.; Agbossou-Niedercorn, F.; Pelinski, L.; Laconde, G.; Poulain, R.D.; Deprez, B. A facile and rapid synthesis of N-benzyl-2-substituted piperazines. Tetrahedron Lett. 2011, 52, 1705–1708. [Google Scholar] [CrossRef]
  11. Koparde, S.; Hosamani, K.M.; Kulkarni, V.; Joshi, S.D. Synthesis of coumarin-piperazine derivatives as potent anti-microbial and anti-inflammatory agents, and molecular docking studies. Chem. Data Collect. 2018, 15–16, 197–206. [Google Scholar] [CrossRef]
  12. Hafeez, F.; Zahoor, A.F.; Ahmad, S.; Ahmad, M.; Faiz, S. Recent progress in the synthesis of diclofenac based NSAIDs analogs/derivatives. Synth. Commun. 2018, 49, 325–350. [Google Scholar] [CrossRef]
  13. Kaya, B.; Yurttas, L.; Saǧlik, B.N.; Levent, S.; Özkay, Y.; Kaplancikli, Z.A. Novel 1-(2-pyrimidin-2-yl)piperazine derivatives as selective monoamine oxidase (MAO)-A inhibitors. J. Enzyme Inhib. Med. Chem. 2017, 32, 193–202. [Google Scholar] [CrossRef] [Green Version]
  14. Gou, C.-C.; Tong, R.-B.; Li, K.-L. Chloroalkyl piperazine and nitrogen mustard porphyrins: Synthesis and anticancer activity. Bioorg. Med. Chem. 2004, 12, 2469–2475. [Google Scholar] [CrossRef]
  15. Akhtar, R.; Zahoor, A.F.; Rasul, A.; Ahmad, M.; Anjum, M.N.; Ajmal, M.; Raza, Z. Design, synthesis, in-silico study and anticancer potential of novel n-4-piperazinyl-ciprofloxacin-aniline hybrids. Pak. J. Pharm. Sci. 2019, 32, 2215–2222. [Google Scholar]
  16. Zahoor, A.F.; Yousaf, M.; Siddique, R.; Ahmad, S.; Naqvi, S.A.R.; Rizvi, S.M.A. Synthetic strategies toward the synthesis of enoxacin-, levofloxacin-, and gatifloxacin-based compounds: A review. Synth. Commun. 2017, 47, 1021–1039. [Google Scholar] [CrossRef]
  17. Akhtar, R.; Yousaf, M.; Naqvi, S.A.R.; Irfan, M.; Zahoor, A.F.; Hussain, A.I.; Chatha, S.A.S. Synthesis of ciprofloxacin-based compounds: A review. Synth. Commun. 2016, 46, 1849–1879. [Google Scholar] [CrossRef]
  18. Todorovic, A.; Haskell-Leuvano, C. A review of melanocortin receptor small molecule ligands. Peptides 2005, 26, 2026–2036. [Google Scholar] [CrossRef]
  19. De Freitas Oliveira, J.W.; Rocha, H.A.O.; de Medeiros, W.M.T.Q.; Silva, M.S. Application of dithiocarbamates as potential new antitrypanosomatids-drugs: Approach chemistry, functional and biological. Molecules 2019, 24, 2806. [Google Scholar] [CrossRef] [Green Version]
  20. Pang, H.; Chen, D.; Cui, Q.C.; Dou, Q.P. Sodium diethyldithiocarbamate, an AIDS progression inhibitor and a copper-binding compound, has proteasome-inhibitory and apoptosis-inducing activities in cancer cells. Int. J. Mol. Med. 2007, 19, 809–816. [Google Scholar] [CrossRef] [Green Version]
  21. Amir, M.K.; Rehman, Z.; Hayat, F.; Khan, S.Z.; Hogarth, G.; Kondratyuk, T.; Pezzuto, J.M.; Tahir, M.N. Monofunctional platinum(ii) dithiocarbamate complexes: Synthesis, characterization and anticancer activity. RSC Adv. 2016, 6, 110517–110524. [Google Scholar] [CrossRef]
  22. Fu, D.-J.; Li, J.-H.; Yang, J.-J.; Li, P.; Zhang, Y.-B.; Liu, S.; Li, Z.-R.; Zhang, S.-Y. Discovery of novel chalcone-dithiocarbamates as ROS-mediated apoptosis inducers by inhibiting catalase. Bioorg. Chem. 2019, 86, 375–385. [Google Scholar] [CrossRef]
  23. Buac, D.; Schmitt, S.; Ventro, G.; Kona, F.R.; Dou, Q.P. Dithiocarbamate-based coordination compounds as potent proteasome inhibitors in human cancer cells. Mini Rev. Med. Chem. 2012, 12, 1193–1201. [Google Scholar] [CrossRef] [Green Version]
  24. Viegas-Junior, C.; Danuello, A.; da Silva Bolzani, V.; Barreiro, E.J.; Fraga, C.A.M. Molecular hybridization: A useful tool in the design of new drug prototypes. Curr. Med. Chem. 2007, 12, 1829–1852. [Google Scholar] [CrossRef]
  25. Kouznetsov, V.V.; Gómez-Barrio, A. Recent developments in the design and synthesis of hybrid molecules based on aminoquinoline ring and their antiplasmodial evaluation. Eur. J. Med. Chem. 2009, 44, 3091–3113. [Google Scholar] [CrossRef]
  26. Geltik, M.; Grütter, C.; Simrad, J.R.; Klüter, S.; Rabiller, M.; Rode, H.B.; Robubi, A.; Rauh, D. Hybrid compound design to overcome the gatekeeper T338M mutation in cSrc. J. Med. Chem. 2009, 52, 3915–3926. [Google Scholar] [CrossRef]
  27. Henderson, B.J.; Carper, D.J.; Gonzǎlez-Cestari, T.F.; Yi, B.; Mohasenan, K.; Pavlovicz, R.E.; Dalefield, M.L.; Coleman, R.S.; Li, C.; McKay, D.B. Structure–activity relationship studies of sulfonylpiperazine analogues as novel negative allosteric modulators of human neuronal nicotinic receptors. J. Med. Chem. 2011, 54, 8681–8692. [Google Scholar] [CrossRef] [Green Version]
  28. Cormier, R.; Burdab, W.N.; Harringtonb, L.; Edlingera, J.; Kodigepallib, K.M.; Thomas, J.; Kepolka, R.; Roma, G.; Anderson, B.E.; Turosa, E.; et al. Studies on the antimicrobial properties of N-acylated ciprofloxacins. Bioorg. Med. Chem. Lett. 2012, 22, 6513–6520. [Google Scholar] [CrossRef] [Green Version]
  29. Faiz, S.; Zahoor, A.F.; Ajmal, M.; Kamal, S.; Ahmad, S.; Abdelgawed, A.M.; Elnaggar, M.E. Design, synthesis, antimicrobial evaluation, and laccase catalysis effect of novel benzofuran–oxadiazole and benzofuran–triazole hybrids. J. Heterocycl. Chem. 2019, 56, 2839–2852. [Google Scholar] [CrossRef]
  30. Hafeez, F.; Zahoor, A.F.; Rasool, A.; Ahmad, S.; Mansha, A. Synthesis and anticancer evaluation of 2-oxo-2-(arylamino) ethyl 4-phenylpiperazine-1-carbodithioates. Pak. J. Pharm. Sci. 2021, 34, 353–357. [Google Scholar]
  31. Hafeez, F.; Mansha, A.; Zahoor, A.F.; Ali, K.G.; Khan, S.G.; Naqvi, S.A.R. Facile green approach towards the synthesis of some phenyl piperazine based dithiocarbamates as potent hemolytic and thrombolytic agents. Pak. J. Pharm. Sci. 2021, 34, 1885–1890. [Google Scholar] [PubMed]
  32. Rasul, A.; Di, J.; Millimouno, F.M.; Malhi, M.; Tsuji, I.; Ali, M.; Li, J.; Li, X. Reactive oxygen species mediate isoalantolactone-induced apoptosis in human prostate cancer cells. Molecules 2013, 18, 9382–9396. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Shahzadi, I.; Zahoor, A.F.; Rasul, A.; Rasool, N.; Raza, Z.; Faisal, S.; Parveen, B.; Kamal, S.; Zia-ur-Rehman, M.; Zahid, F.M. Synthesis, anticancer, and computational studies of 1, 3, 4-oxadiazole-purine derivatives. J. Heterocycl. Chem. 2020, 57, 2782–2794. [Google Scholar] [CrossRef]
  34. Riaz, M.; Rasool, N.; Bukhari, I.H.; Shahid, M.; Zubair, M.; Rizwan, K.; Rashid, U. In vitro antimicrobial, antioxidant, cytotoxicity and GC-MS analysis of Mazus goodenifolius. Molecules 2012, 17, 14275–14287. [Google Scholar] [CrossRef] [Green Version]
  35. Batool, M.; Tajammal, A.; Farhat, F.; Verpoort, F.; Khattak, Z.A.K.; Mehr-un-Nisa, M.S.; Ahmad, H.A.; Munawar, M.A.; Zia-ur-Rehman, M.; Basra, M.A.R. Molecular docking, computational, and antithrombotic studies of novel 1,3,4-oxadiazole derivatives. Int. J. Mol. Sci. 2018, 19, 3606. [Google Scholar] [CrossRef] [Green Version]
  36. Gfeller, D.; Grosdider, A.; Wirth, M.; Diana, A.; Michielin, O.; Zoete, V. SwissTargetPrediction: A web server for target prediction of bioactive small molecules. Nucleic Acids Res. 2014, 42, W32–W38. [Google Scholar] [CrossRef]
Figure 1. Structures of some common piperazine-based drugs.
Figure 1. Structures of some common piperazine-based drugs.
Molecules 27 04776 g001
Figure 2. Structures of dithiocarbamate based drugs.
Figure 2. Structures of dithiocarbamate based drugs.
Molecules 27 04776 g002
Scheme 1. Synthesis of piperazine-1-carbodithioates 5a5j. Reagents and conditions: (a): methanesulfonyl chloride, DCM, 0 °C, (b): BrCH2COBr, pyridine, DCM, and (c): AcONa, CS2, MeOH, 70 °C, 30 min.
Scheme 1. Synthesis of piperazine-1-carbodithioates 5a5j. Reagents and conditions: (a): methanesulfonyl chloride, DCM, 0 °C, (b): BrCH2COBr, pyridine, DCM, and (c): AcONa, CS2, MeOH, 70 °C, 30 min.
Molecules 27 04776 sch001
Figure 3. Dose-response of compound 5b.
Figure 3. Dose-response of compound 5b.
Molecules 27 04776 g003
Figure 4. Dose-response of compound 5i.
Figure 4. Dose-response of compound 5i.
Molecules 27 04776 g004
Figure 5. SAR summary.
Figure 5. SAR summary.
Molecules 27 04776 g005
Figure 6. Validation of docking protocol by cognate redocking of the co-crystallized ligand of 3HR1; co-crystallized conformation (pink) compared to the re-docked conformation (purple) of native ligand in three-dimensional space.
Figure 6. Validation of docking protocol by cognate redocking of the co-crystallized ligand of 3HR1; co-crystallized conformation (pink) compared to the re-docked conformation (purple) of native ligand in three-dimensional space.
Molecules 27 04776 g006
Figure 7. Conformational analysis of ligands docked within the binding pocket of PDE10A. The simulated best binding mode with the lowest ΔG of (a): PQ-10 and (b) compound 5b within the 3D space of the PDE10A active site.
Figure 7. Conformational analysis of ligands docked within the binding pocket of PDE10A. The simulated best binding mode with the lowest ΔG of (a): PQ-10 and (b) compound 5b within the 3D space of the PDE10A active site.
Molecules 27 04776 g007
Figure 8. The Interaction behavior of ligands within the active site of PDE10A; (a) a 2D representation of PQ10 and (b) compound 5b interacting with the PDE10A catalytic residues colored by type of interactions.
Figure 8. The Interaction behavior of ligands within the active site of PDE10A; (a) a 2D representation of PQ10 and (b) compound 5b interacting with the PDE10A catalytic residues colored by type of interactions.
Molecules 27 04776 g008
Table 1. The anti-lung cancer activity, hemolytic, and thrombolytic potential of piperazine-1-carbodithioates 5a5j according to Scheme 1.
Table 1. The anti-lung cancer activity, hemolytic, and thrombolytic potential of piperazine-1-carbodithioates 5a5j according to Scheme 1.
PiperazineArCell Viability A-549 aHemolysis (%)Thrombolysis (%)
5aPhenyl61.35 ± 2.293.158.9
5b2-Chlorophenyl25.11 ± 2.491.2754.4
5c4-Chlorophenyl60.29 ± 5.961047.8
5d3,4-Dichlorophenyl38.08 ± 2.850.156.03
5e2-Fluorophenyl68.94 ± 6.640.353.5
5f4-Fluorophenyl41.01 ± 3.732.850.3
5g4-Methoxyphenyl43.36 ± 4.421.350.9
5h2,5-Dimethoxyphenyl40.25 ± 3.347.160.2
5i2,4-Dimethylphenyl25.31 ± 3.622.857.4
5j3,4-dimethylphenyl47.94 ± 1.570.158.7
Standard 100 ± 0
DMSO 0.010.57
ABTS 95.980
a Cell viability: (Mean ± SD) in triplicate with a concentration of 100 μg/mL. DMSO and ABTS were used as negative and positive controls, respectively.
Table 2. The IFD parameters of 5b and PQ-10 simulated in the active site of PDE10A.
Table 2. The IFD parameters of 5b and PQ-10 simulated in the active site of PDE10A.
LigandΔG (kcal/mol)Interacting ResiduesInteraction Type
5b−8.30TYR514, PHE686, TYR683, VAL712, PRO702, MET703, GLU711, SER667H–bonding, sulfur–X, π–σ, π–π stacked, amide–π stacked, alkyl, π–alkyl
PQ-10−8.97GLU711, PRO702, TYR683, VAL712, GLN716, ILE682, ALA679, TYR514, ASP664, PHE716, MET703H–bonding, sulfur–X, π–σ, π–π stacked, amide–π stacked, alkyl, π–alkyl
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hafeez, F.; Zahoor, A.F.; Rasul, A.; Mansha, A.; Noreen, R.; Raza, Z.; Ali, K.G.; Irfan, A.; El-Hiti, G.A. Ultrasound-Assisted Synthesis and In Silico Modeling of Methanesulfonyl-Piperazine-Based Dithiocarbamates as Potential Anticancer, Thrombolytic, and Hemolytic Structural Motifs. Molecules 2022, 27, 4776. https://doi.org/10.3390/molecules27154776

AMA Style

Hafeez F, Zahoor AF, Rasul A, Mansha A, Noreen R, Raza Z, Ali KG, Irfan A, El-Hiti GA. Ultrasound-Assisted Synthesis and In Silico Modeling of Methanesulfonyl-Piperazine-Based Dithiocarbamates as Potential Anticancer, Thrombolytic, and Hemolytic Structural Motifs. Molecules. 2022; 27(15):4776. https://doi.org/10.3390/molecules27154776

Chicago/Turabian Style

Hafeez, Freeha, Ameer Fawad Zahoor, Azhar Rasul, Asim Mansha, Razia Noreen, Zohaib Raza, Kulsoom Ghulam Ali, Ali Irfan, and Gamal A. El-Hiti. 2022. "Ultrasound-Assisted Synthesis and In Silico Modeling of Methanesulfonyl-Piperazine-Based Dithiocarbamates as Potential Anticancer, Thrombolytic, and Hemolytic Structural Motifs" Molecules 27, no. 15: 4776. https://doi.org/10.3390/molecules27154776

Article Metrics

Back to TopTop