Skip to main content

Novel thiazole-based cyanoacrylamide derivatives: DNA cleavage, DNA/BSA binding properties and their anticancer behaviour against colon and breast cancer cells

Abstract

A novel series of 2-cyano-3-(pyrazol-4-yl)-N-(thiazol-2-yl)acrylamide derivatives (3af) were synthesized using Knoevenagel condensation and characterized using various spectral tools. The weak nuclease activity of compounds (3af) against pBR322 plasmid DNA was greatly enhanced by irradiation at 365 nm. Compounds 3b and 3c, incorporating thienyl and pyridyl moieties, respectively, exhibited the utmost nuclease activity in degrading pBR322 plasmid DNA through singlet oxygen and superoxide free radicals’ species. Furthermore, compounds 3b and 3c affinities towards calf thymus DNA (CT-DNA) and bovine serum albumin (BSA) were investigated using UV–Vis and fluorescence spectroscopic analysis. They revealed good binding characteristics towards CT-DNA with Kb values of 6.68 × 104 M−1 and 1.19 × 104 M−1 for 3b and 3c, respectively. In addition, compounds 3b and 3c ability to release free radicals on radiation were targeted to be used as cytotoxic compounds in vitro for colon (HCT116) and breast cancer (MDA-MB-231) cells. A significant reduction in the cell viability on illumination at 365 nm was observed, with IC50 values of 23 and 25 µM against HCT116 cells, and 30 and 9 µM against MDA-MB-231 cells for compounds 3b and 3c, respectively. In conclusion, compounds 3b and 3c exhibited remarkable DNA cleavage and cytotoxic activity on illumination at 365 nm which might be associated with free radicals’ production in addition to having a good affinity for interacting with CT-DNA and BSA.

Graphical Abstract

Peer Review reports

Introduction

Over the previous decade, there has been rapid growth in research areas that focus on the synthesis of various molecules that exhibit biological activity and can be utilized for various pharmaceutical purposes [1,2,3,4]. The promising category of heterocyclic organic compounds exhibits a broad spectrum of physical, chemical, and biological characteristics that could fulfill these intentions [5,6,7,8]. For example, nitrogen-containing heterocycles are one of the unique compounds well-known for their engagement in medicinal sciences [9, 10]. Pyrazole is a heterocyclic compound whose chemical structure comprises a five-membered ring embracing two adjacent nitrogen atoms [11] and is renowned for its uses as a pharmaceutical agent utilizing its anti-inflammatory [12], anticancer [13], antimicrobial [14], and various therapeutic effects [15]. Thiazole ring is a nitrogen/sulfur atoms five-membered ring [16], which is acquainted for its anti-inflammatory [17], antiviral [18], treatment of Alzheimer's disease [19], and antibacterial [20], besides its utilization in cancer therapy, in which some of its derivatives exhibited a potential effect, particularly in the management of colon and breast cancer [21,22,23,24,25]. Furthermore, cyanoacrylamide moiety is appreciated for its anticancer activity which has been reported for certain derivatives in the treatment of breast and colon cancer [26,27,28,29], antimicrobial [30], antiviral [31], and diversity of biological activities, for instance, DNA fragmentation [32], kinase/anhydrase inhibitors [33], and antioxidant activity [34].

Numerous scientific fields, including biochemistry, pharmaceutical chemistry, and cancer therapy, have focused on exploring the interaction between small molecules and DNA [35,36,37]. Since DNA is a key pharmacological goal and how a drug binds to DNA determines how successful it is, this is the initial step in developing a new class of therapeutic medicines. In general, a series of covalent and non-covalent interactions enhance the chemotherapeutic agent binding to DNA [38]. In malignant cells, these interactions give rise to DNA damage and hinder replication and/or transcription, which ultimately leads to cell death [39].

Serum albumins have a variety of physiological roles and represent about 55% of the proteins in the blood. They serve as a carrier protein for a wide range of endogenous and exogenous molecules, including fatty acids, hormones, and several medications. Because of its affordability, high degree of resemblance to human serum albumin, and ease of availability, BSA is considered an ideal model in the study of drug-protein interactions. Therefore, it is crucial to investigate the interaction of small molecules with BSA to provide an extra benefit for using them as possible chemotherapeutic agents [40, 41].

In continuation of our interest in the synthesis of bioactive heterocycles [42,43,44,45,46,47,48] herein, novel thiazole-related cyanoacrylamide derivatives were synthesized via Knoevenagel condensation. The cleavage activity of the synthesized compounds against pBR322 plasmid DNA was assessed using agarose gel electrophoresis with investigating the potential mechanism involved. Besides, the interaction of the potent compounds with different biological molecules such as CT-DNA and BSA using fluorescence and UV–visible spectroscopy was also investigated to understand the mode of binding and the affinity of these compounds towards those crucial biomolecules. Furthermore, the cytotoxic effect of the active cyanoacrylamide against human cancer cell lines (HCT116 and MDA-MB-231 cells) was examined to explore their ability to be used as chemotherapeutic agents.

Experimental

Materials

2-Aminothiazole was bought from HIMEDIA (India), and all pyrazole aldehydes used were obtained as previously [49]. Solvents of analytical grade were purchased from Sigma-Aldrich or Merck. Calf-thymus DNA (CT-DNA, CAS 73049-39-5), agarose (molecular biology grade, CAS 9012-36-6), doxorubicin (CAS 25316-40-9), 3-(4,5-dimethylthiazol-2-yl)-2,5-tetrazolium bromide (MTT, CAS 298-93-1), and ethidium bromide (EtBr, CAS 1239-45-8) were procured from Sigma-Aldrich (USA). Bovine serum albumin (BSA) was purchased from BioBasic Inc. (Canada). Supercoiled pBR322 DNA was obtained from SibEnzyme Ltd. (Russia). DMEM, fetal bovine serum, and penicillin/streptomycin were provided by Gibco, Thermo Fisher Scientific Inc.

Physical measurements

Melting points were measured with a Stuart melting point apparatus and are uncorrected. The IR spectra were recorded using a FTIR Bruker–vector 22 spectrophotometer as KBr pellets. The 1H NMR spectra were recorded in dimethyl sulfoxide (DMSO)–d6 as a solvent on Varian Gemini NMR spectrometer at 300 MHz or Bruker AVS NMR spectrometer at 500 MHz using TMS as the internal standard. Chemical shifts are reported as δ values in ppm. Mass spectra were recorded with a Shimadzu GCMS–QP–1000 EX mass spectrometer in EI (70 eV) model. The elemental analyses were performed at the Microanalytical centre, Cairo University. Electronic absorption spectra were recorded using a Shimadzu UV-3101 spectrophotometer in the range of 200–800 nm. Fluorescence spectra were recorded on a spectrofluorometer Jasco FP-6200, Japan.

General procedure for the synthesis of thiazole-based cyanoacrylamide derivatives

A mixture of 2-cyano-N-(thiazol-2-yl)acetamide (1) (167 mg, 10 mmol) and the corresponding aldehydes 2af (10 mmol) was heated at reflux for 3 h in absolute EtOH (10 mL) in the presence of piperidine (0.2 mL, 2 mmol) as a basic catalyst. The formed solid product was then filtered, washed with ethanol, dried, and crystallized from EtOH/dioxane (5:1, v/v) to give 2-cyanoacrylamide derivatives 3af.

2-Cyano-3-(1,3-diphenyl-1H-pyrazol-5-yl)-N-(thiazol-2-yl)acrylamide (3a)

figure b

Pale yellow solid (365 mg, 92%); Mp 265–267°C; IR (KBr): \(\overline{\nu }\) 3325 (NH), 2307 (C≡N), 1697 (C=O amide), 1636 (C=C) cm−1; 1H NMR (500 MHz, DMSO-d6; Fig. S1): δ 7.16 (d, J = 4.4 Hz, 1H, Ar–H), 7.43 (t, J = 7.4 Hz, 1H, Ar–H), 7.48 (d, J = 4.2 Hz, 1H, Ar–H), 7.52–7.59 (m, 6H, Ar–H), 7.66 (d, J = 7.9 Hz, 2H, Ar–H), 7.91 (d, J = 8.4 Hz, 2H, Ar–H), 8.23 (s, 1H, vinylic H), 9.17 (s, 1H, pyrazolyl-H5), 13.18 (s, 1H, NH) ppm. MS (EI, 70 eV): m/z (%) 397]M+[; Anal. Calcd for C22H15N5OS: C, 66.48; H, 3.80; N, 17.62; S, 8.07. Found: C, 66.31; H, 3.65; N, 17.44; S, 8.01.

2-Cyano-3-(1-phenyl-3-(thiophen-2-yl)-1H-pyrazol-4-yl)-N-(thiazol-2-yl)acrylamide (3b)

figure c

Pale yellow solid (359 mg, 89%); Mp 252–256°C; IR (KBr): \(\overline{\nu }\) 3294 (NH), 2214 (CN), 1662 (C=O amide) cm−1; 1H NMR (500 MHz, DMSO-d6; Fig. S2): δ 7.19 (t, J = 5.1 Hz, 1H, Ar–H), 7.28 (d, J = 4.5 Hz, 1H, Ar–H), 7.40–7.63 (m, 5H, Ar–H), 7.78 (d, J = 5.4 Hz, 1H, Ar–H), 7.89 (d, J = 7.9 Hz, 2H, Ar–H), 8.41 (s, 1H, vinylic CH), 9.17 (s, 1H, pyrazolyl-H5), 13.16 (s, 1H, NH) ppm; MS (EI, 70 eV): m/z (%) 403]M+[; Anal. Calcd for C20H13N5OS2: C, 59.54; H, 3.25; N, 17.36; S, 15.89. Found: C, 59.32; H, 3.07; N, 17.21; S, 15.79.

2-Cyano-3-(1-phenyl-3-(pyridin-3-yl)-1H-pyrazol-4-yl)-N-(thiazol-2-yl)acrylamide (3c)

figure d

Pale yellow solid (354 mg, 89%); Mp 250–252°C; IR (KBr): \(\overline{\nu }\) 3292 (NH), 2214 (CN), 1667 (C=O amide) cm−1; 1H NMR (300 MHz, DMSO-d6; Fig. S3): δ 7.20 (d, J = 4.0 Hz, 1H, Ar–H), 7.40–7.71 (m, 5H, Ar–H), 7.95 (d, J = 8.2 Hz, 2H, Ar–H), 8.14 (dd, J = 7.9, 2.1 Hz, 1H, Ar–H), 8.25 (s, 1H, vinylic CH), 8.75 (d, J = 4.8 Hz, 1H, Ar–H), 8.92 (s, 1H, pyridin-H3), 9.23 (s, 1H, pyrazolyl-H5), 13.11 (s, 1H, NH) ppm; 13C NMR (75 MHz, DMSO-d6; Fig. S4): δ 112.9, 115.1, 117.2, 119.7, 124.1, 127.1, 128.2, 129.5, 129.9, 130.4, 131.6, 136.5, 138.6, 141.4, 149.2, 150.2, 151.9, 157.8, 164.7 ppm; MS (EI, 70 eV): m/z (%) 398]M+[; Anal. Calcd for C21H14N6OS: C, 63.30; H, 3.54; N, 21.09; S, 8.05. Found: C, 63.17; H, 3.43; N, 21.03; S, 8.12.

2-Cyano-3-(1-phenyl-3-(p-tolyl)-1H-pyrazol-4-yl)-N-(thiazol-2-yl)acrylamide (3d)

figure e

Pale yellow solid (308 mg, 75%); Mp 255–257°C; IR (KBr): \(\overline{\nu }\) 3232 (NH), 2216 (CN), 1674 (C=O) cm−1; 1H NMR (300 MHz, DMSO-d6; Fig. S5): δ 2.41 (s, 3H, CH3), 7.18 (d, J = 4.0 Hz, 1H, Ar–H), 7.40 (d, J = 8.1 Hz, 2H, Ar–H), 7.48 (dd, J = 10.4, 5.7 Hz, 2H, Ar–H), 7.54–7.70 (m, 4H, Ar–H), 7.92 (d, J = 7.7 Hz, 2H, Ar–H), 8.26 (s, 1H, vinylic CH), 9.17 (s, 1H, pyrazolyl-H5), 13.17 (s, 1H, NH) ppm; MS (EI, 70 eV): m/z (%) 411]M+[; Anal. Calcd for C23H17N5OS: C, 67.14; H, 4.16; N, 17.02; S, 7.79. Found: C, 67.04; H, 4.09; N, 17.14; S, 7.62.

2-Cyano-3-(3-(4-methoxyphenyl)-1-phenyl-1H-pyrazol-4-yl)-N-(thiazol-2-yl) acrylamide (3e)

figure f

Pale yellow solid (342 mg, 80%); Mp 252–254°C; IR (KBr): \(\overline{\nu }\) 3236 (NH), 2214 (CN), 1667 (C=O) cm−1; 1H NMR (300 MHz, DMSO-d6; Fig. S6): δ 3.85 (s, 3H, OCH3), 7.16 (m, 3H, Ar–H), 7.48 (dd, J = 12.6, 5.8 Hz, 2H, Ar–H), 7.54–7.67 (m, 4H, Ar–H), 7.92 (d, J = 8.0 Hz, 2H, Ar–H), 8.26 (s, 1H, vinylic CH), 9.16 (s, 1H, pyrazolyl-H5), 12.81 (s, 1H, NH) ppm; MS (EI, 70 eV): m/z (%) 427]M+[; Anal. Calcd for C23H17N5OS: C, 67.14; H, 4.16; N, 17.02; S, 7.79. Found: C, 67.04; H, 4.09; N, 17.14; S, 7.62.

3-(3-(4-Chlorophenyl)-1-phenyl-1H-pyrazol-4-yl)-2-cyano-N-(thiazol-2-yl)acrylamide (3f)

figure g

Pale yellow solid (358 mg, 83%); Mp 270–272°C; IR (KBr): \(\overline{\nu }\) 3240 (NH), 2215 (CN), 1674 (C=O) cm−1; 1H NMR (300 MHz, DMSO-d6; Fig. S7): δ 7.19 (d, J = 4.0 Hz, 1H, Ar–H), 7.42–7.54 (m, 2H, Ar–H), 7.56–7.77 (m, 6H, Ar–H), 7.93 (d, J = 8.5 Hz, 2H, Ar–H), 8.25 (s, 1H, vinylic CH), 9.19 (s, 1H, pyrazolyl-H5), 13.16 (s, 1H, NH) ppm; MS (EI, 70 eV): m/z (%) 431]M+[; Anal. Calcd for C22H14ClN5OS: C, 61.18; H, 3.27; Cl, 8.21; N, 16.22; S, 7.42. Found: C, 61.11; H, 3.21; Cl, 8.11; N, 16.07; S, 7.19.

DNA cleavage studies

Using agarose gel electrophoresis, the DNA cleavage studies of the synthesized compounds were conducted in the absence and presence of irradiation at 365 nm. The compounds were solubilized in DMSO and diluted to a final concentration (200 μM) when mixed with Tris/NaCl buffer (pH 7.1) containing pBR322 plasmid DNA (0.3 μg), where the final DMSO percentage was 5%. Then, the mixture was incubated at 37°C for 30 min for dark conditions and followed by illumination for irradiated conditions using UV-A light at 365 nm for 60 min at room temperature. Subsequently, the solutions were mixed with loading buffer (10 mM Tris–HCl (pH 7.6), 0.03% bromophenol blue, 0.03% xylene cyanol FF, 60% glycerol, and 60 mM EDTA) and loaded on the agarose-gel (1%, w/v), comprised of TBE buffer (Tris–HCl, boric acid, and EDTA) treated with (0.5 μg/ml) EtBr. Samples were run in the previous system for 2 h at 75 V and the gel images were acquired by UV trans-illuminator at the end of the process. ImageJ software was used to assess the percentage of cleavage fragments [50]. To conduct mechanistic pBR322 DNA photo-cleavage investigations of the selected compounds, 200 μM of the 3b and 3c were independently mixed with pBR322 plasmid DNA in the presence and absence of several radical scavengers, such as NaN3, KI, and DMSO, as previously reported [50].

DNA binding studies

CT-DNA stock solution was prepared by dissolving CT-DNA in Tris/NaCl buffer (pH 7.1). Then, it was stored at 4°C and was manipulated within 4 days of preparation. The concentration was estimated by applying the equation \(C= \frac{{A}_{260 nm}}{\varepsilon \times \text{ b}}\) where C is the concentration of DNA solution, A is the absorbance at wavelength 260 nm, b is the path length and ε is the molar extinction coefficient of DNA that equals 6600 M−1 cm−1. The purity also was calculated using the absorbance ratio at wavelength 260 and 280 respectively, which was 1.8 indicating free DNA from protein contamination. Stock solutions of compounds 3b and 3c were prepared by dissolving in DMSO.

DNA UV–Visible absorption spectrophotometric studies

The interaction between the selected compounds (3b and 3c) and CT-DNA was assessed through the spectrophotometric titration method [50]. In brief, a fixed concentration of compounds (25 μM) was mixed with an accumulative addition of CT-DNA (0–60 μM) at room temperature with a constant DMSO percentage of 5%. After 5 min from each addition, the UV–Vis absorption spectrum (250–650 nm) was recorded. The hyper/hypochromic percentage change at the characteristic peak of 3b and 3c was assessed using the formula: \(H\%= \frac{{A}_{free}- {A}_{bound}}{{A}_{bound}}\times 100\). While the binding constant (Kb) of 3b and 3c was estimated using Wolfe-Shimmer Equation [51]:

$$\frac{[DNA]}{({\upvarepsilon }_{a}- {\upvarepsilon }_{f})}=\frac{[DNA]}{({\varepsilon }_{b}- {\varepsilon }_{f})}+ \frac{1}{{K}_{b}({\varepsilon }_{b}- {\varepsilon }_{f})}$$

where [DNA] denotes CT-DNA concentration, εa, εf and εb stands for the apparent absorption coefficient, the extinction coefficient of the unbound compound, and the extinction coefficient when fully bound to DNA respectively. Kb is estimated from [DNA] / (εa − εf) versus [DNA] plots by computation of the ratio of slope to intercept. Standard Gibbs free energy (ΔG°) was calculated using the equation ΔG° =  − RTlnKb (where T is the temperature, 298 K; R is the universal gas constant, 8.314 J K−1 mol−1).

The Ethidium Bromide quenching assay

EtBr is a conventional intercalator whose fluorescence intensity intensifies in response to CT-DNA interaction. Solution of CT-DNA-EtBr with final concentrations of EtBr (20 µM) and CT-DNA (100 µM) in Tris–HCl buffer solution (pH = 7.1) was prepared, then compounds were solubilized in DMSO, added to the prepared solution to attain a range of concentration (0–160 µM) and (0–80 µM) for 3b and 3c, respectively. Then, the mixtures were incubated 3 min after each addition and the emission spectra (540–710 nm) were recorded using an excitation wavelength of 520 nm.

BSA binding studies

BSA was dissolved in Tris/NaCl buffer (pH 7.1) to prepare the BSA stock solution, which was then kept at 4°C and utilized within a week after preparation [52, 53]. The concentration was evaluated by applying the equation \(C= \frac{{A}_{280 nm}}{\varepsilon \times \text{ b}}\) where C is the concentration of protein solution, A is the absorbance at wavelength 280 nm, b is the path length and ε is molar extinction coefficient of protein that equals 44,300 M−1 cm−1.

BSA UV–Visible absorption spectrophotometric studies

The interaction was evaluated through dilution of BSA to 15 μM in Tris/NaCl (pH 7.1) during the spectrophotometric titration. After that, the compounds dissolved in DMSO were added with a range of increasing concentration (0–40 μM), where the final DMSO percentage is 5%. UV–Vis absorption spectrum measurement for each concentration was assessed by Shimadzu UV-3101 PC NIR spectrophotometer after 3 min of each addition. The Kb for compounds 3b and 3c was calculated using equation [54]:

$$\frac{1}{{\text{A}}_{obs}- {\text{A}}_{0}}=\frac{1}{{A}_{c}- {A}_{0}}+ \frac{1}{{K}_{b}\left({A}_{c}- {A}_{0}\right)[\text{compound }]}$$

where A0, Aobs and Ac are absorbance of solution in the absence, various concentrations, and saturation of the compound at the characteristic peak, respectively. Where Kb is estimated from the plot of \(\frac{1}{{\text{A}}_{obs}- {\text{A}}_{0}}\) versus \(\frac{1}{[Drug]}\). Hence, Kb is evaluated from the ratio between intercept and slope.

Tryptophan quenching assay with BSA

BSA (15 µM) solution was prepared in Tris/HCl buffer solution (pH = 7.1) and mixed with serial concentrations of compounds 3b and 3c (0–40 µM), achieving 5% final concentration of DMSO. Following 3 min incubation after each addition, fluorescence quenching spectra were recorded from 300 to 450 nm using a spectrofluorometer with an excitation wavelength of 289 nm.

Cell viability assay

The cytotoxicity of 3b and 3c against MDA-MB-231 and HCT116 cell lines, with and without illumination, was assessed using the MTT test [55]. Two human cancer cell lines, including MDA-MB-231 and HCT116 cell lines, were purchased from the National Research Centre, Dokki, Cairo, Egypt. Briefly, 96-well plates were seeded with cells (5 × 103 cells/well) and left overnight. The compounds dissolved in DMSO were then applied to the cells following the serial dilution procedure in the range of 0–100 µM. The cells were then incubated for an additional 20 h. Subsequently, the cells were exposed to UV light for 20 min at a distance of 10 cm. To study the effects of the illumination, a control plate was created under the same circumstances but without illumination. The two plates were incubated for a further 20 h. After that, PBS was used to wash the cells and MTT solution (0.5 mg/ml) was added. To dissolve the generated formazan crystals, the solution was replaced with 100 µl of DMSO per well after 4 h of incubation. An ELISA reader operating at 492 nm was used to measure the absorbance after 15 min of plate shaking. By relating the absorbance of treated and negative control cells, the relative cell viability was calculated. GraphPad Prism software was used to calculate the half inhibitory concentration (IC50). Doxorubicin (0–100 μM) was utilized as a positive cytotoxicity control in the usual dark environment, while the vehicle (DMSO, 0.1%) was used as a negative control.

Results and discussion

Synthetic chemistry

2-Cyano-N-(thiazol-2-yl)acetamide (1) was obtained in high yields and purity following the method reported by us [28, 56] and others earlier [26]. Knoevenagel condensation of 2-cyano-N-(thiazol-2-yl)acetamide (1) with aromatic aldehydes 2af in the presence of piperidine as a basic catalyst affords the corresponding cyanoacrylamides incorporating thiazole and pyrazole moieties 3af in very good yields (75–92%) (Scheme 1). The constitutions of the obtained products were confirmed based on spectral data. Thus, 1H NMR of compound 3c as a representative example indicated four singlet signals at 8.25, 8.92, 9.23, and 13.11 corresponding to vinylic CH, pyridin-H3, pyrazole-H5, and NH, respectively. The aromatic protons appear at their expected positions. The 13C NMR featured 19 signals corresponding to 19 different signals.

Scheme 1.
scheme 1

Knoevenagel condensation reaction of 2-cyano-N-(thiazol-2-yl)acetamide (1) with different aldehydes 2af

DNA cleavage

According to Modi et al., pyrazole derivatives have been reported to have DNA cleavage activity, where all compounds incorporating pyrazole moiety can generate detectable plasmid DNA cleavage at 100 μM [57]. Furthermore, thiazole hybrids appeared to have promising effects on DNA cleavage activity [58]. DNA cleavage activity of the synthesized compounds bearing pyrazole and thiazole rings in combination with cyanoacrylamide moiety was estimated chemically using gel electrophoresis. When plasmid DNA is mixed with the cleavage agent and loaded for gel electrophoresis, the supercoiled form exhibits the fastest migration (Form I), followed by the nicked circular (Form II) owing to a break in one strand which moves more slowly as it becomes loose. Mutual cleavage of the two strands, on the other hand, produces a linear form (Form III), which has a moderate migration speed [59, 60].

Chemically induced DNA cleavage by gel electrophoresis

Compounds 3af (200 μM) displayed insignificant cleavage activity with pBR322 DNA (0.3 μg) in Tris/NaCl (5 mM Tris–HCl/50 mM NaCl (pH 7.1)) buffer containing 5% DMSO, as illustrated in Fig. 1 and Fig. S8.

Fig. 1
figure 1

Agarose gel electrophoresis pattern of pBR322 DNA (0.3 μg) cleavage by compounds 3af (200 μM) incubated in the dark at 37°C for 30 min. Lane 1: DNA control; Lane 2: DNA + 3a; Lane 3: DNA + 3b; Lane 4: DNA + 3c; Lane 5; DNA + 3d; Lane 6: DNA + 3e; Lane 7: DNA + 3f

Photo-induced DNA cleavage by gel electrophoresis

The DNA photocleavage activity of these compounds was examined under the same conditions, in addition to exposing them along with pBR322 DNA for an incubation time of 30 min at 37°C followed by irradiation at 365 nm for 60 min. Figure 2 and Fig. S9 depicted that compounds 3a and 3d exhibited non-detectable DNA cleavage, while the rest showed remarkable photocleavage activity uniquely 3b and 3c.

Fig. 2
figure 2

Agarose gel electrophoresis pattern of pBR322 DNA (0.3 μg) cleavage by compounds 3af (200 μM) incubated in the dark at 37°C for 30 min followed by irradiation at 365 nm for 60 min (A). Lane 1: DNA control; Lane 2: DNA + 3a; Lane 3: DNA + 3b; Lane 4: DNA + 3c; Lane 5; DNA + 3d; Lane 6: DNA + 3e; Lane 7: DNA + 3f. Histogram illustrated the DNA cleavage percentage for each compound (B)

Structure–activity relationship (SAR)

In general, the stability of aromatic compounds dictates how applied UV light will affect them, determining their ability to withstand structural changes and the production of free radicals. Mainly, the benzene ring has reputable stability owing to its high resonance energy, unlike the heterocyclic pyridine and thiophene rings which have lower resonance energies. This stability arrangement may be responsible for their susceptibility to photochemical reactions and their capability of producing free radicals culminating in their DNA cleavage potentials. Furthermore, the ring strain has been attributed to the enhancement of thiophene and pyridine's reactivity propensity toward photochemical processes, and the creation of distinct photoproducts or reaction pathways missing in less strained aromatic compounds [61,62,63,64,65,66]. As a result, it was found that the newly synthesized compounds containing heterocyclic moieties had superior photo-cleavage activity than those substituted with aromatic moieties. As illustrated in Fig. 3, thienyl (5-membered ring)-substituted derivative 3b was more effective than pyridinyl (6-membered ring)-substituted one 3c. Meanwhile, the aromatic-substituted derivatives demonstrated that the electron-withdrawing chloro-substituted derivative 3f displayed a lower activity than electron-donating methoxy-substituted compound 3e. But more effective than methyl-substituted derivative 3d. Notably, the substituted aromatic derivatives generally displayed higher activity than the unsubstituted aromatic derivative 3a.

Fig. 3
figure 3

The photocleavage activity of cyanoacrylamide derivatives structure–activity

Mechanistic pBR322 DNA photo-cleavage studies

To determine their mechanistic effect, compounds 3b and 3c were chosen for further investigation. Different free radical scavengers such as DMSO (hydroxyl radical scavenger), KI (superoxide radical scavenger), and NaN3 (singlet oxygen scavenger) were added along with the irradiated incubation of pBR322 DNA with compounds 3b and 3c. The scavenger’s addition is to elucidate the photocleavage mechanism of these compounds. The outcomes from Fig. 4 and Fig. S10 have shown total hindrance of compounds 3b and 3c activity in the presence of NaN3 and KI revealing that the mechanistic action of these compounds is through singlet oxygen and superoxide free radicals’ species. While the presence of DMSO did not influence the photocleavage activity of the compounds so hydroxyl radicals are not implicated in the compound’s activity. Consequently, 3b and 3c were selected for additional investigation utilising different conformation of DNA (CT-DNA) and BSA.

Fig. 4
figure 4

Agarose gel electrophoresis pattern of pBR322 DNA (0.3 μg) cleavage by 3b (200 μM) (A) and 3c (200 μM) (B) with different free radical scavengers incubated in the dark at 37°C for 30 min followed by irradiation at 365 nm for 60 min. DNA. Lane 1: DNA + DMSO (200 mM); Lane 2: DNA + 3b/3c + DMSO (200 mM); Lane 3: DNA + KI (200 mM); Lane 4: DNA + 3b/3c + KI (200 mM); Lane 5; DNA + NaN3 (200 mM); Lane 6: DNA + 3b/3c + NaN3 (200 mM); Lane 7: DNA + 3b/3c; Lane 8: DNA control

DNA binding studies

UV–Vis measurements

Electronic absorption titration is the technique most frequently used for investigating the interaction between a small molecule and DNA through monitoring the changes in absorption spectra (position or absorbance of the bands) resulting from the interaction between the examined compounds and DNA [67]. Compounds could interact with DNA non-covalently through different forms, including intercalation between stacked DNA base pairs, groove binding, or electrostatic contact with the negatively charged backbone of DNA [68,69,70]. As depicted in Fig. 5, UV–Vis spectra of compounds 3b and 3c exhibited distinctive peaks at 423 nm and 359 nm with a hypochromism (~ 14% and ~ 40%) after adding increasing concentrations of CT-DNA to 3b and 3c, respectively, with slight blue shift (~ 1 nm) for both compounds. Compound-DNA complex stability was measured by calculating the intrinsic binding constant Kb values using Wolfe–Shimmer equation [71] for 3b and 3c which was 6.68 × 104 M−1 and 1.19 × 104 M−1, respectively. The magnitude of the binding constant was lower than those of strong intercalators like EtBr (Kb 106 M−1) [72], but comparable to partial intercalators such as ferrocene appended naphthylamide derivatives (Kb 104 M−1) [73]. Nevertheless, UV–Vis measurement provides only preliminary information regarding the mode of binding, therefore, further measurement would be required to ensure the binding mode. The spontaneity of the interaction was also assessed by Gibbs free energy change (ΔG) which was − 6.47 and − 5.47 kcal mol−1 for compounds 3b and 3c, respectively, in which negative values suggest directing of reaction towards product formation.

Fig. 5
figure 5

Absorption spectra of 3b (A) and 3c (B) show a hypochromic effect for both compounds denoted by the arrow direction along with increasing CT-DNA concentration (0–60 µM). The incorporated graph exemplifies the plot used for binding constant (Kb) computation using Wolfe–Shimmer equation

Ethidium bromide displacement measurements

The interaction mode between compound and CT-DNA can be discriminated through fluorescence spectroscopy with the EtBr–CT-DNA system [74]. EtBr is a classic indicator of intercalation where an intense fluorescence signal is emitted from the EtBr-CT-DNA system through intercalation of EtBr between adjacent DNA base pairs with no obvious signal for unaccompanied EtBr [75, 76]. According to earlier research, substances with comparable DNA binding modes can substitute intercalated EtBr, significantly reducing EtBr's fluorescence intensity [77]. In order to demonstrate the fluorescence emission spectra of EtBr-CT-DNA at an excitation wavelength of 500 nm, fluorescence titration for each compound (3b and 3c) against EtBr-CT-DNA was carried out. This accomplished a significant reduction in fluorescence intensity (~ 33.3% and ~ 30.7%, respectively) in the presence of 3b and 3c, as illustrated in Figs. 6 and 7, respectively. The Stern–Volmer equation was used for the study of the effect of quenchers on the CT-DNA-EtBr system [78, 79].

$${\text{F}}_{0}/{\text{F}}{\text{= 1+} }{\text{K}}_{\text{sv}}\left[{\text{Q}}\right]{\text{ = 1+} }{\text{k}}_{\text{q}}{\tau }_{0}{\text{ [Q] }}$$

where F0 and F are fluorescence intensities in the absence and presence of the compound used for quenching, respectively; KSV is a linear Stern–Volmer quenching constant; [Q] is the concentration of the quenching compound and τ0 is the fluorescence lifetime of the fluorophore (10−8 s) in absence of the compound responsible for quenching. KSV for 3b and 3c is computed from the linear regression plot of F0/F against [Q]. Also, the binding constant (Kb) and the number of binding sites (n) were determined by the Scatchard method [80].

Fig. 6
figure 6

A Fluorescence emission spectra of 3b with CT-DNA-EtBr (λex = 500 nm, 3b: 0–160 µM in the direction of the arrow from lowest to highest). B The Stern–Volmer plot quenching effect of 3b on CT-DNA-EtBr at room temperature. C Scatchard method for assigning binding constant Kb for 3b with CT-DNA-EtBr at room temperature

Fig. 7
figure 7

A Fluorescence emission spectra of 3c with CT-DNA-EtBr (λex = 500 nm, 3c: 0–80 µM in the direction of the arrow from lowest to highest). B The Stern–Volmer plot quenching effect of 3c on CT-DNA-EtBr at room temperature. C Scatchard method for assigning binding constant Kb for 3c with CT-DNA-EtBr at room temperature

$${\text{log }}(({{\text{F}}_{0}}-{\text{F}})/{\text{F}})={\text{log}}{{\text{ K}}_{\text{b}}}+ {\text{n}} {\text{ log}}{\text{ [Q]}}$$

Kb has been assigned as the intercept and n as the slope in the linear plotting of log((F0–F)/F) versus log[Q]. The detailed results for 3b and 3c were computed and depicted in Table 1. It was found that Kb values obtained for 3b and 3c were very small in comparison to the reported value of a classical intercalator of EtBr–DNA which is in order of 107 M−1 [81]. Moreover, ΔG for both 3b and 3c showed negative values indicating that the interaction is spontaneous.

Table 1 Parameters of 3b and 3c interaction with CT-DNA-EtBr: quenching constants (KSV), quenching rate constant (kq), binding constants (Kb), binding site numbers (n), and binding energy (ΔG)

BSA binding studies

UV–Visible measurements

Several investigations undoubtedly focus on serum albumin owing to its prominence and abundance in blood plasma [82]. Besides its crucial role in the exogenous and endogenous molecules transportation, the binding ability of albumin to small molecules as drugs enhances solubility, and drug-life time, and reduces their toxicity, which is functional in therapeutic targets [83]. Moreover, it has a role in the regulation of redox potential and colloid osmotic pressure between blood and tissue [84]. Therefore, albumins have extensive physiological, medical, and biochemical applications [85]. BSA has opted for investigation concerning bioactive chemical binding and homology to human serum albumin [86]. UV–Vis spectrophotometric titration effectively investigated the physical changes in proteins and analysed the protein–drug interaction [87]. The interaction between compounds (3b and 3c) and BSA was examined (Fig. 8), revealing that 3b and 3c had noteworthy hyperchromicity (~ 52.2% and ~ 48.2%) at 277 nm, respectively. Furthermore, the estimated Kb values for 3b and 3c were found to be 8.81 × 103 and 3.41 × 103 M−1, respectively. The interaction capability was also predicted by computation of ΔG that was shown to be − 7.6 for 3b and − 6.8 kcal mol−1 for 3c, where negative values imply the spontaneous interaction between the compounds and BSA.

Fig. 8
figure 8

UV–Vis absorption spectra of BSA (15 µM) in the absence and presence of 3b (A) and 3c (B), showed that the increase in the concentration of compounds (0–40 µM) had a hyperchromic effect denoted by the arrow direction. The incorporated graph exemplified the plot used for binding constant (Kb) computation

Tryptophan quenching experiment

Fluorescence investigations could be used to conduct a specified analysis of whether synthetic molecules bind to biomolecules as BSA. It is widely perceived that predominantly tryptophan (Trp) is responsible for the inherited fluorescence among the other two amino acids [tyrosine (Tyr) and phenylalanine (Phe)] in the BSA [88]. The tryptophan emission spectrum is frequently altered because of protein conformational variation, subunit interactions, substrate binding, or denaturation [89]. Reduced fluorescence emission intensity is observed for 3b and 3c, which may be due to disrupted BSA structure altering the tryptophan microenvironments. Moreover, the internalization of fluorophore residues into the protein's hydrophobic core is due to protein folding or related to the hydrophobic amino acid residues around Trp, such as Leu, Tyr, and phe [90]. The serial addition of the 3b and 3c to the BSA solution revealed a reduction in the maximum peak emission of BSA, which was around 341 nm, as depicted in Figs. 9 and 10. Also, a linear plot of F0/F against the concentration of 3b and 3c was used for computing the values of KSV (9.0 × 104 and 9.9 × 104 M−1) and kq (9.0 × 1012 and 9.9 × 1012 M−1 s−1) for the titled compounds (3b and 3c, respectively). Furthermore, the Kb value (~ 0.3 × 102 and 0.34 × 102 M−1) and n value (~ 1.3) assessed from the plot of log(F0−F/F) against log(compound concentration) was nearly the same for both 3b and 3c. The negative values of ΔG also supported the UV–Vis results of free energy changes and implied the spontaneity of 3b/3c–BSA binding, where the interaction parameters were illustrated in Table 2.

Fig. 9
figure 9

A Fluorescence emission spectra of BSA (15 μM) with increasing amounts of 3b (0–40 μM) (λex = 289 nm, with the arrow direction from the lowest to the highest concentration of 3b). B The Stern–Volmer plot quenching effect of 3b on BSA at room temperature. C Scatchard method for assigning Kb for BSA with 3b at room temperature

Fig. 10
figure 10

A Fluorescence emission spectra of BSA (15 μM) with increasing amounts of 3c (0–40 μM) (λex = 289 nm, with the arrow direction from the lowest to the highest concentration of 3c). B The Stern–Volmer plot quenching effect of 3c on BSA at room temperature. C Scatchard method for assigning Kb for BSA with 3c at room temperature

Table 2 Parameters of 3b and 3c interaction with BSA: quenching constants (KSV), quenching rate constant (kq), binding constants (Kb), binding site numbers (n), and binding energy (ΔG)

Cytotoxic studies

To assess the cytotoxic impact of compounds 3b and 3c on HCT116 and MDA-MB-231 cancer cells with and without illumination, the MTT test was conducted. As shown in (Fig. 11), a slight cytotoxic effect of compounds 3b and 3c on the HCT116 cell line in the absence of illumination with IC50 values of 72 and 82 µM, respectively, compared to doxorubicin as a positive control (Fig. S11), with IC50 value of 40 µM. A significant reduction in viability was observable upon irradiation with substantial IC50 values of 23 and 25 µM for compounds 3b and 3c, respectively.

Fig. 11
figure 11

The cytotoxic effects of 3b (A) and 3c (B) on human colon cancer (HCT116) cells without/with illumination at 365 nm for 20 min at different concentrations (0, 6.25, 12.5, 25, 50 and 100 µM)

Furthermore, 3b and 3c revealed a weaker anticancer potential against MDA-MB-231 cells (Fig. 12) in the dark conditions with IC50 values of 91 and 64 µM, respectively, relative to doxorubicin with IC50 value of 64.8 µM (Fig. S11). On the other hand, treating MDA-MB-231 cells with the titled compounds under illuminating conditions (Fig. 12) revealed remarkable IC50 values of 30 and 9 µM for compounds 3b and 3c, respectively. This may be accredited to free radicals’ species generation as singlet oxygen and superoxide, which facilitate the process of apoptosis for cancer cells.

Fig. 12
figure 12

The cytotoxic effects of 3b (A) and 3c (B) on human breast cancer (MDA-MB-231) cells without/with illumination at 365 nm for 20 min at different concentrations (0, 6.25, 12.5, 25, 50 and 100 µM)

Conclusion

A novel six thiazole-related cyanoacrylamide derivatives (3af) were synthesized and characterized and their nuclease activity was assessed through DNA cleavage experiment. These derivatives exhibited insignificant chemically induced nuclease activity but on irradiation, this activity was improved mostly for the derivatives 3b and 3c through the released singlet oxygen and superoxide free radical species. Moreover, a partial intercalation binding of 3b and 3c with CT-DNA was supported by ethidium bromide displacement and UV–Vis measurements. In addition, 3b and 3c interactions with BSA, demonstrated by tryptophan quenching and UV–Vis measurements, could provide facilitated movement inside the human body. The evaluation of the cytotoxic effect against colon and breast cancer cells, before and after irradiation revealed an enhanced activity that could serve as a base for the synthesis of potential cancer therapeutic agents that specifically function with their enhanced effect under irradiation. Therefore, targeting this advantageous property toward other biological activities may reveal promising results. In general, photodynamic therapy for cancer treatment could benefit from these thiazole-related cyanoacrylamide derivatives (3b and 3c) for targeted therapy that depends on irradiation.

Availability of data and materials

Data available at request (Dr. Mohamed A. Ragheb, and Ismail A. Abdelhamid, email: mattia@cu.edu.eg; ismail_shafy@cu.edu.eg; ismail_shafy@yahoo.com). All data generated or analyzed during this study are included in this published article and its supplementary information file.

References

  1. Dong P, Rakesh KP, Manukumar HM, Mohammed YHE, Karthik CS, Sumathi S, et al. Innovative nano-carriers in anticancer drug delivery-a comprehensive review. Bioorg Chem. 2019;85:325–36.

    Article  CAS  PubMed  Google Scholar 

  2. Rakesh KP, Wang S-M, Leng J, Ravindar L, Asiri AM, Marwani HM, et al. Recent development of sulfonyl or sulfonamide hybrids as potential anticancer agents: a key review. Anticancer Agents Med Chem. 2018;18:488–505.

    Article  CAS  PubMed  Google Scholar 

  3. Fang W-Y, Ravindar L, Rakesh KP, Manukumar HM, Shantharam CS, Alharbi NS, et al. Synthetic approaches and pharmaceutical applications of chloro-containing molecules for drug discovery: a critical review. Eur J Med Chem. 2019;173:117–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Zhao C, Rakesh KP, Ravidar L, Fang W-Y, Qin H-L. Pharmaceutical and medicinal significance of sulfur (SVI)-Containing motifs for drug discovery: a critical review. Eur J Med Chem. 2019;162:679–734.

    Article  CAS  PubMed  Google Scholar 

  5. Faisal M, Saeed A, Hussain S, Dar P, Larik FA. Recent developments in synthetic chemistry and biological activities of pyrazole derivatives. J Chem Sci. 2019;131:70.

    Article  Google Scholar 

  6. Wilson Lucas S, Zijian Qin R, Rakesh KP, Sharath Kumar KS, Qin H-L. Chemical and biology of sulfur fluoride exchange (SuFEx) click chemistry for drug discovery. Bioorg Chem. 2023;130:106227.

    Article  CAS  PubMed  Google Scholar 

  7. Zhao C, Rakesh KP, Mumtaz S, Moku B, Asiri AM, Marwani HM, et al. Arylnaphthalene lactone analogues: synthesis and development as excellent biological candidates for future drug discovery. RSC Adv. 2018;8:9487–502.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Rakesh KP, Shantharam CS, Sridhara MB, Manukumar HM, Qin H-L. Benzisoxazole: a privileged scaffold for medicinal chemistry. Medchemcomm. 2017;8:2023–39.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Ardiansah B. Recent reports on pyrazole-based bioactive compounds as candidate for anticancer agents. Asian J Pharm Clin Res. 2017;10:45.

    Article  CAS  Google Scholar 

  10. Moku B, Ravindar L, Rakesh KP, Qin H-L. The significance of N-methylpicolinamides in the development of anticancer therapeutics: Synthesis and structure-activity relationship (SAR) studies. Bioorg Chem. 2019;86:513–37.

    Article  CAS  PubMed  Google Scholar 

  11. Rodhan WF, Kadhium SS, Ali ZZM, Eleiwi AG, Abbas RF, Mohamed IR, et al. Chemistry and synthesis of Bis pyrazole derivatives and their biological activity: a review. J Phys Conf Ser. 2021;1853:012059.

    Article  CAS  Google Scholar 

  12. Ebenezer O, Shapi M, Tuszynski JA. A review of the recent development in the synthesis and biological evaluations of pyrazole derivatives. Biomedicines. 2022;10:1124.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Bennani FE, Doudach L, Cherrah Y, Ramli Y, Karrouchi K, Ansar M, et al. Overview of recent developments of pyrazole derivatives as an anticancer agent in different cell line. Bioorg Chem. 2020;97:103470.

    Article  CAS  PubMed  Google Scholar 

  14. Li M-M, Huang H, Pu Y, Tian W, Deng Y, Lu J. A close look into the biological and synthetic aspects of fused pyrazole derivatives. Eur J Med Chem. 2022;243:114739.

    Article  CAS  PubMed  Google Scholar 

  15. Yu W, Zhai Z-W, Wedge DE, Duke SO, Wu H-K, Weng J-Q, et al. Synthesis and biological activity of novel 1,3,4-oxadiazole derivatives containing a pyrazole moiety. Res Chem Intermed. 2019;45:5989–6001.

    Article  CAS  Google Scholar 

  16. Niu Z-X, Wang Y-T, Zhang S-N, Li Y, Chen X-B, Wang S-Q, et al. Application and synthesis of thiazole ring in clinically approved drugs. Eur J Med Chem. 2023;250:115172.

    Article  CAS  PubMed  Google Scholar 

  17. Ankali KN, Rangaswamy J, Shalavadi M, Naik N, Krishnamurthy GN. Synthesis and molecular docking of novel 1,3-thiazole derived 1,2,3-triazoles and in vivo biological evaluation for their anti anxiety and anti inflammatory activity. J Mol Struct. 2021;1236:130357.

    Article  CAS  Google Scholar 

  18. Singh IP, Gupta S, Kumar S. Thiazole compounds as antiviral agents: an update. Med Chem. 2020;16:4–23.

    Article  CAS  PubMed  Google Scholar 

  19. Sağlık BN, Levent S, Osmaniye D, Acar Çevik U, Kaya Çavuşoğlu B, Özkay Y, et al. Design, synthesis, and biological activity evaluation of new donepezil-like compounds bearing thiazole ring for the treatment of alzheimer’s disease. Crystals. 2020;10:637.

    Article  Google Scholar 

  20. Mohanty P, Behera S, Behura R, Shubhadarshinee L, Mohapatra P, Barick AK, Jali BR. Antibacterial activity of thiazole and its derivatives: a review. Biointerface Res Appl Chem. 2021;12:2171–95.

    Article  Google Scholar 

  21. Sabry MA, Ghaly MA, Maarouf AR, El-Subbagh HI. New thiazole-based derivatives as EGFR/HER2 and DHFR inhibitors: synthesis, molecular modeling simulations and anticancer activity. Eur J Med Chem. 2022;241:114661.

    Article  CAS  PubMed  Google Scholar 

  22. Salem MG, El-Maaty DMA, El-Deen YIM, Elesawy BH, El AA, Saleh A, et al. Novel 1,3-thiazole analogues with potent activity against breast cancer: a design, synthesis, in vitro, and in silico study. Molecules. 2022;27:4898.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Petrou A, Fesatidou M, Geronikaki A. Thiazole ring—a biologically active scaffold. Molecules. 2021;26:3166.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Abu-Melha S, Edrees MM, Salem HH, Kheder NA, Gomha SM, Abdelaziz MR. Synthesis and biological evaluation of some novel thiazole-based heterocycles as potential anticancer and antimicrobial agents. Molecules. 2019;24:539.

    Article  PubMed  PubMed Central  Google Scholar 

  25. Mamidala S, Mudigunda VS, Peddi SR, Bokara KK, Manga V, Vedula RR. Design and synthesis of new thiazoles by microwave-assisted method: evaluation as an anti-breast cancer agents and molecular docking studies. Synth Commun. 2020;50:2488–501.

    Article  CAS  Google Scholar 

  26. Mohamed MF, Saddiq AA, Abdelhamid IA. Attacking the mitochondria of colorectal carcinoma by novel 2-cyanoacrylamides linked to ethyl 1,3-diphenylpyrazole-4-carboxylates moiety as a new trend for chemotherapy. Bioorg Chem. 2020;103:104195.

    Article  CAS  PubMed  Google Scholar 

  27. Mohamed MF, Samir N, Ali A, Ahmed N, Ali Y, Aref S, et al. Apoptotic induction mediated p53 mechanism and Caspase-3 activity by novel promising cyanoacrylamide derivatives in breast carcinoma. Bioorg Chem. 2017;73:43–52.

    Article  CAS  PubMed  Google Scholar 

  28. Mohamed MF, Elhakim HKA, Saddiq AA, Abdelhamid IA. A novel inhibitor, 2-cyano-3-(1-phenyl-3-(thiophen-2-yl)-pyrazol-4-yl)acrylamide linked to sulphamethoxazole, blocks anti-apoptotic proteins via molecular docking and strongly induced apoptosis of HCT116 cell line by different molecular tools. Arab J Chem. 2020;13:5978–95.

    Article  CAS  Google Scholar 

  29. Sroor FM, Mahrous KF, Shafey HI, Eid NR, Abdelhamid IA, Ibrahim NS. In silico and in vitro studies of novel cyanoacrylamides incorporating pyrazole moiety against breast and prostate carcinomas. Med Chem Res. 2023;32:1190–203.

    Article  CAS  Google Scholar 

  30. Ragab A, Fouad SA, Abu Ali OA, Ahmed EM, Ali AM, Askar AA, et al. Sulfaguanidine hybrid with some new pyridine-2-one derivatives: design, synthesis, and antimicrobial activity against multidrug-resistant bacteria as dual DNA gyrase and DHFR inhibitors. Antibiotics. 2021;10:1–31.

    Article  Google Scholar 

  31. Amblard F, Zhou S, Liu P, Yoon J, Cox B, Muzzarelli K, et al. Synthesis and antiviral evaluation of novel peptidomimetics as norovirus protease inhibitors HHS Public Access. Bioorg Med Chem Lett. 2018;28:2165–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Mohamed MF, Saddiq AA, Al-Shaikh TM, Ibrahim NS, Abdelhamid IA. Computational studies and sever apoptotic bioactivity of new heterocyclic cyanoacrylamide based p-fluorophenyl and p-phenolic compounds against liver carcinoma (Hepg2). Bioorg Chem. 2021;114:105147.

    Article  CAS  PubMed  Google Scholar 

  33. Kang SJ, Lee JW, Song J, Park J, Choi J, Suh KH, et al. Synthesis and biological activity of 2-cyanoacrylamide derivatives tethered to imidazopyridine as TAK1 inhibitors. J Enzyme Inhib Med Chem. 2020;35:1928–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Krishnan KG, Kumar CU, Lim WM, Mai CW, Thanikachalam PV, Ramalingan C. Novel cyanoacetamide integrated phenothiazines: synthesis, characterization, computational studies and in vitro antioxidant and anticancer evaluations. J Mol Struct. 2020;1199:127037.

    Article  CAS  Google Scholar 

  35. Asadi Z, Nasrollahi N, Karbalaei-Heidari H, Eigner V, Dusek M, Mobaraki N, et al. Investigation of the complex structure, comparative DNA-binding and DNA cleavage of two water-soluble mono-nuclear lanthanum(III) complexes and cytotoxic activity of chitosan-coated magnetic nanoparticles as drug delivery for the complexes. Spectrochim Acta Part A Mol Biomol Spectrosc. 2017;178:125–35.

    Article  CAS  Google Scholar 

  36. Rakesh KP, Darshini N, Manukumar HM, Vivek HK, Eissa MYH, Prasanna DS, et al. Xanthone conjugated amino acids as potential anticancer and DNA binding agents: molecular docking, cytotoxicity and SAR studies. Anticancer Agents Med Chem. 2019;18:2169–77.

    Article  Google Scholar 

  37. Sridhara MB, Rakesh KP, Manukumar HM, Shantharam CS, Vivek HK, Kumara HK, et al. Synthesis of dihydrazones as potential anticancer and DNA binding candidates: a validation by molecular docking studies. Anticancer Agents Med Chem. 2020;20:845–58.

    Article  CAS  PubMed  Google Scholar 

  38. Andrezálová L, Országhová Z. Covalent and noncovalent interactions of coordination compounds with DNA: an overview. J Inorg Biochem. 2021;225:111624.

    Article  PubMed  Google Scholar 

  39. Venugopal S, Sharma V, Mehra A, Singh I, Singh G. DNA intercalators as anticancer agents. Chem Biol Drug Des. 2022;100:580–98.

    Article  CAS  PubMed  Google Scholar 

  40. Hoogenboezem EN, Duvall CL. Harnessing albumin as a carrier for cancer therapies. Adv Drug Deliv Rev. 2018;130:73–89.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Cho H, Jeon SI, Ahn C-H, Shim MK, Kim K. Emerging albumin-binding anticancer drugs for tumor-targeted drug delivery: current understandings and clinical translation. Pharmaceutics. 2022;14:728.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Ibrahim NS, Mohamed MF, Elwahy AHM, Abdelhamid IA. Biological activities and docking studies on novel Bis 1,4-DHPS linked to arene core via ether or ester linkage. Lett Drug Des Discov. 2018;15:10.

    Article  Google Scholar 

  43. Ghozlan SAS, Abdelmoniem AM, Butenschön H, Abdelhamid IA. Discrepancies in the reactivity pattern of azaenamines towards cinnamonitriles: synthesis of novel aza-steroid analogues. Tetrahedron. 2015;71:1413–8.

    Article  CAS  Google Scholar 

  44. Fathi EM, Sroor FM, Mahrous KF, Mohamed MF, Mahmoud K, Emara M, et al. Design, synthesis, in silico and in vitro anticancer activity of novel bis-furanyl-chalcone derivatives linked through alkyl spacers. ChemistrySelect. 2021;6:6202–11.

    Article  CAS  Google Scholar 

  45. Sroor FM, Abdelmoniem AM, Abdelhamid IA. Facile synthesis, structural activity relationship, molecular modeling and in vitro biological evaluation of new urea derivatives with incorporated isoxazole and thiazole moieties as anticancer agents. ChemistrySelect. 2019;4:10113–21.

    Article  CAS  Google Scholar 

  46. Mohamed MF, Sroor FM, Ibrahim NS, Salem GS, El-Sayed HH, Mahmoud MM, et al. Novel [l,2,4]triazolo[3,4-a]isoquinoline chalcones as new chemotherapeutic agents: Block IAP tyrosine kinase domain and induce both intrinsic and extrinsic pathways of apoptosis. Invest New Drugs. 2021;39:1.

    Article  Google Scholar 

  47. Helmy MT, Sroor FM, Mahrous KF, Mahmoud K, Hassaneen HM, Saleh FM, et al. Anticancer activity of novel 3-(furan-2-yl)pyrazolyl and 3-(thiophen-2-yl)pyrazolyl hybrid chalcones: synthesis and in vitro studies. Arch Pharm. 2022;355: e2100381.

    Article  Google Scholar 

  48. Mohamed MF, Ibrahim NS, Elwahy AHM, Abdelhamid IA. Molecular studies on novel antitumor Bis 1,4-dihydropyridine derivatives against lung carcinoma and their limited side effects on normal melanocytes. Anticancer Agents Med Chem. 2018;18:2156–68.

    Article  CAS  PubMed  Google Scholar 

  49. Kolluri PK, Gurrapu N, Subhashini NJP, Putta S, Singh SS, Vani T, et al. Design, synthesis of novel (Z)-2-(3-(4-((3-benzyl-2,4-dioxothiazolidin-5-ylidene)methyl)-1-phenyl-1H-pyrazol-3-yl)phenoxy)-N-arylacetamide derivatives: evaluation of cytotoxic activity and molecular docking studies. J Mol Struct. 2020;1202:127300.

    Article  CAS  Google Scholar 

  50. Ragheb MA, Eldesouki MA, Mohamed MS. DNA binding, photo-induced DNA cleavage and cytotoxicity studies of lomefloxacin and its transition metal complexes. Spectrochim Acta Part A Mol Biomol Spectrosc. 2015;138:585–95.

    Article  CAS  Google Scholar 

  51. Tarai SK, Mandal S, Tarai A, Som I, Pan A, Bagchi A, et al. Biophysical study on DNA and BSA binding activity of Cu(II) complex: synthesis, molecular docking, cytotoxic activity, and theoretical approach. Appl Organomet Chem. 2023;37: e7164.

    Article  CAS  Google Scholar 

  52. Yu J, Liu J-Y, Xiong W-M, Zhang X-Y, Zheng Y. Binding interaction of sodium benzoate food additive with bovine serum albumin: multi-spectroscopy and molecular docking studies. BMC Chem. 2019;13:95.

    Article  PubMed  PubMed Central  Google Scholar 

  53. Feizi-Dehnayebi M, Dehghanian E, Mansouri-Torshizi H. A novel palladium(II) antitumor agent: synthesis, characterization, DFT perspective, CT-DNA and BSA interaction studies via in-vitro and in-silico approaches. Spectrochim Acta Part A Mol Biomol Spectrosc. 2021;249:119215.

    Article  CAS  Google Scholar 

  54. Bhunia A, Mistri S, Manne RK, Santra MK, Manna SC. Synthesis, crystal structure, cytotoxicity study, DNA/protein binding and molecular docking of dinuclear copper(II) complexes. Inorg Chim Acta. 2019;491:25–33.

    Article  CAS  Google Scholar 

  55. Khaled RM, Friedrich A, Ragheb MA, Abdel-Ghani NT, Mansour AM. Cytotoxicity of photoactivatable bromo tricarbonyl manganese(I) compounds against human liver carcinoma cells. Dalt Trans. 2020;49:9294–305.

    Article  CAS  Google Scholar 

  56. Ibrahim NS, Sroor FM, Mahrous KF, Elaleem JAA, Abdelhamid IA. Cytotoxic effect of new (E)-2-cyano-N-(tetrahydrobenzo[b]thiophen-2-yl)acrylamide derivatives: down-regulation of RBL2 and STAT2 and triggering of DNA damage in breast carcinoma. ChemistrySelect. 2023;8: e202301754.

    Article  CAS  Google Scholar 

  57. Modi M, Jain M. Green approach for the synthesis of 3-methyl-1-phenyl-4-((2-phenyl-1H-indol 3-yl)methylene)-1H-pyrazole-5(4H)-ones and their DNA Cleavage, antioxidant, and antimicrobial activities. J Heterocycl Chem. 2019;56:3303–12.

    Article  CAS  Google Scholar 

  58. Doğan A, Özdemir S, Yalcin M, Sari H, Nural Y. Naphthoquinone–thiazole hybrids bearing adamantane: synthesis, antimicrobial, DNA cleavage, antioxidant activity, acid dissociation constant, and drug-likeness. J Res Pharm. 2021;25(3):292–304.

    Google Scholar 

  59. Bengi K, Maddikayala S, Pulimamidi SR. DNA binding, cleavage, docking, biological and kinetic studies of Cr(III), Fe(III), Co(II) and Cu(II) complexes with ortho-vanillin Schiff base derivative. Appl Organomet Chem. 2022;36: e6451.

    Article  CAS  Google Scholar 

  60. Göktürk T, Topkaya C, Sakallı Çetin E, Güp R. New trinuclear nickel(II) complexes as potential topoisomerase I/IIα inhibitors: in vitro DNA binding, cleavage and cytotoxicity against human cancer cell lines. Chem Pap. 2022;76:2093–109.

    Article  Google Scholar 

  61. Sharma AK, Ryan DL, Marr NL, Wadsley MP, Cheatham SF. Rearrangement products in aqueous photolysis of thifensulfuron methyl. J Photochem Photobiol A Chem. 2017;346:401–10.

    Article  CAS  Google Scholar 

  62. Vaganova E, Eliaz D, Shimanovich U, Leitus G, Aqad E, Lokshin V, et al. Light-induced reactions within poly(4-vinyl pyridine)/pyridine gels: the 1,6-polyazaacetylene oligomers formation. Molecules. 2021;26:6925.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Zhang L, Yan W, Kohtani S, Fukuyoshi S, Hu M, Nagao S, et al. Promotive effects of marine-derived dimethyl sulfoxide on the photodegradation of phenanthrene in the atmosphere. Sci Total Environ. 2024;926:171938.

    Article  CAS  PubMed  Google Scholar 

  64. Zakharov AV, Lvov AG, Rostovtseva IA, Metelitsa AV, Chernyshev AV, Krayushkin MM, et al. Photocyclization of diarylethenes: the effect of imidazole on the oxidative photodegradation process. Photochem Photobiol Sci. 2019;18:1101–9.

    Article  CAS  PubMed  Google Scholar 

  65. Gritzapis PS, Varras PC, Andreou N-P, Katsani KR, Dafnopoulos K, Psomas G, et al. p -Pyridinyl oxime carbamates: synthesis, DNA binding, DNA photocleaving activity and theoretical photodegradation studies. Beilstein J Org Chem. 2020;16:337–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Zhussupova GE, Zhussupova AI. Impact of “organic chemistry” course on subsequent courses, studying biologically active compounds. Proced Soc Behav Sci. 2015;191:1247–54.

    Article  Google Scholar 

  67. Jomová K, Hudecova L, Lauro P, Simunkova M, Alwasel SH, Alhazza IM, et al. A switch between antioxidant and prooxidant properties of the phenolic compounds myricetin, morin, 3′,4′-dihydroxyflavone, taxifolin and 4-hydroxy-coumarin in the presence of copper(II) ions: a spectroscopic, absorption titration and DNA damage study. Molecules. 2019;24:4335.

    Article  PubMed  PubMed Central  Google Scholar 

  68. Kumar N, Kaushal R, Awasthi P. Non-covalent binding studies of transition metal complexes with DNA: a review. J Mol Struct. 2023;1288:135751.

    Article  CAS  Google Scholar 

  69. Sánchez-González Á, Castro TG, Melle-Franco M, Gil A. From groove binding to intercalation: unravelling the weak interactions and other factors modulating the modes of interaction between methylated phenanthroline-based drugs and duplex DNA. Phys Chem Chem Phys. 2021;23:26680–95.

    Article  PubMed  Google Scholar 

  70. El-Adl K, El-Helby A-GA, Sakr H, Elwan A. Design, synthesis, molecular docking and anti-proliferative evaluations of [1,2,4]triazolo[4,3-a]quinoxaline derivatives as DNA intercalators and Topoisomerase II inhibitors. Bioorg Chem. 2020;105:104399.

    Article  CAS  PubMed  Google Scholar 

  71. Rajeshwari K, Anantha Lakshmi PV, Archana J, Sumakanth M. Ternary Cobalt(II), Nickel(II), and Copper(II) complexes containing metformin and ethylenediamine: synthesis, characterization, thermal, in vitro DNA binding, in silico molecular docking, and in vivo antihyperglycemic studies. Appl Organomet Chem. 2021;35: e6100.

    Article  CAS  Google Scholar 

  72. Çeşme M, Özaltay A. DNA-binding studies of ofloxacin using a series of spectroscopic, electrochemical techniques and in silico approaches. ChemistrySelect. 2022;7: e202202278.

    Article  Google Scholar 

  73. Jia D-G, Zheng J-A, Fan Y-R, Yu J-Q, Wu X-L, Wang B-J, et al. Ferrocene appended naphthalimide derivatives: synthesis, DNA binding, and in vitro cytotoxic activity. J Organomet Chem. 2019;888:16–23.

    Article  CAS  Google Scholar 

  74. Zhang Y-P, Li Y, Xu G-C, Li J-Y, Luo H-Y, Li J-Y, et al. Synthesis, crystal structure, DNA/bovine serum albumin binding and antitumor activity of two transition metal complexes with 4-acylpyrazolone derivative. Appl Organomet Chem. 2019;33: e4668.

    Article  Google Scholar 

  75. Wani TA, Alsaif N, Bakheit AH, Zargar S, Al-Mehizia AA, Khan AA. Interaction of an abiraterone with calf thymus DNA: investigation with spectroscopic technique and modelling studies. Bioorg Chem. 2020;100:103957.

    Article  CAS  PubMed  Google Scholar 

  76. Thakor KP, Lunagariya MV, Bhatt BS, Patel MN. Fluorescence and absorption studies of DNA–Pd(II) complex interaction: synthesis, spectroanalytical investigations and biological activities. Luminescence. 2019;34:113–24.

    Article  CAS  PubMed  Google Scholar 

  77. Shabaninejad Z, Nikkhah M, Nabavizadeh SM. DNA binding properties and cytotoxic effects of two double rollover cycloplatinated (II) complexes on cancer cell lines. J Inorg Biochem. 2023;243:112194.

    Article  CAS  PubMed  Google Scholar 

  78. Shivakumara N, Murali KP. Synthesis, spectral characterization and DNA interactions of 5-(4-substituted phenyl)-1,3,4-thiadiazol-2-amine scaffolds. J Mol Struct. 2020;1199:126999.

    Article  Google Scholar 

  79. Lavanya K, Babu PV, Bodapati ATS, Reddy RS, Madku SR, Sahoo BK. Binding of dicoumarol analog with DNA and its antioxidant studies: a biophysical insight by in-vitro and in-silico approaches. Int J Biol Macromol. 2023;244:125301.

    Article  CAS  PubMed  Google Scholar 

  80. Naskar R, Ghosh P, Mandal S, Jana S, Murmu N, Mondal TK. Palladium(II) complex bearing benzothiazole based O, N, S donor pincer ligand: study of in-vitro cytotoxicity, interaction with CT-DNA and BSA protein. J Chem Sci. 2022;134:103.

    Article  CAS  Google Scholar 

  81. Babu LT, Paira P. 9-Arylacenaphtho[1,2- b ]quinoxalines via Suzuki coupling reaction as cancer therapeutic and cellular imaging agents. New J Chem. 2021;45:20447–58.

    Article  CAS  Google Scholar 

  82. Siddiqui S, Ameen F, ur Rehman S, Sarwar T, Tabish M. Studying the interaction of drug/ligand with serum albumin. J Mol Liq. 2021;336:116200.

    Article  CAS  Google Scholar 

  83. Jahanban-Esfahlan A, Ostadrahimi A, Jahanban-Esfahlan R, Roufegarinejad L, Tabibiazar M, Amarowicz R. Recent developments in the detection of bovine serum albumin. Int J Biol Macromol. 2019;138:602–17.

    Article  CAS  PubMed  Google Scholar 

  84. Rabbani G, Ahn SN. Structure, enzymatic activities, glycation and therapeutic potential of human serum albumin: a natural cargo. Int J Biol Macromol. 2019;123:979–90.

    Article  CAS  PubMed  Google Scholar 

  85. Ragheb MA, Abdelwahab RE, Darweesh AF, Soliman MH, Elwahy AHM, Abdelhamid IA. Hantzsch-Like synthesis, DNA photocleavage, DNA/BSA binding, and molecular docking studies of Bis(sulfanediyl)bis(tetrahydro-5-deazaflavin) analogs linked to naphthalene core. Chem Biodivers. 2022;19: e202100958.

    Article  CAS  PubMed  Google Scholar 

  86. Ragheb MA, Omar RS, Soliman MH, Elwahy AHM, Abdelhamid IA. Synthesis, characterization, DNA photocleavage, in silico and in vitro DNA/BSA binding properties of novel hexahydroquinolines. J Mol Struct. 2022;1267:133628.

    Article  CAS  Google Scholar 

  87. Douadi K, Chafaa S, Douadi T, Al-Noaimi M, Kaabi I. Azoimine quinoline derivatives: synthesis, classical and electrochemical evaluation of antioxidant, anti-inflammatory, antimicrobial activities and the DNA / BSA binding. J Mol Struct. 2020;1217:128305.

    Article  CAS  Google Scholar 

  88. Verma R, Pyreddy S, Redmond CE, Qazi F, Khalid A, O’Brien-Simpson NM, et al. Detection and identification of amino acids and proteins using their intrinsic fluorescence in the visible light spectrum. Anal Chim Acta. 2023;1282:341925.

    Article  CAS  PubMed  Google Scholar 

  89. Zhao Z, Shi T, Chu Y, Cao Y, Cheng S, Na R, et al. Comparison of the interactions of flupyrimin and nitenpyram with serum albumins via multiple analysis methods. Chemosphere. 2022;289:133139.

    Article  CAS  PubMed  Google Scholar 

  90. Elshami FI, Ramadan AEM, Ibrahim MM, El-Mehasseb IM, Al-Juaid S, Shaban SY. Metformin containing nickel (II) complexes: synthesis, structural characterization, binding and kinetic interactions with BSA, antibacterial and in-vitro cytotoxicity studies. Appl Organomet Chem. 2020;34: e5437.

    Article  CAS  Google Scholar 

Download references

Acknowledgements

This paper is based upon work supported by Science, Technology & Innovation Funding Authority (STDF) under Grant 48864.

Funding

Open access funding provided by The Science, Technology & Innovation Funding Authority (STDF) in cooperation with The Egyptian Knowledge Bank (EKB). This paper is based upon work supported by Science, Technology & Innovation Funding Authority (STDF) under Grant 48864.

Author information

Authors and Affiliations

Authors

Contributions

K.B. carried out the laboratory work, participated in data analysis and shared in writing the original draft of the manuscript. M.A.R., M.H.S., A.M.A. and I.A.A. suggested the research plan, shared in writing the original draft of the manuscript, and followed up the experimental work. All authors gave final approval for publication.

Corresponding authors

Correspondence to Mohamed A. Ragheb or Ismail A. Abdelhamid.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Barakat, K., Ragheb, M.A., Soliman, M.H. et al. Novel thiazole-based cyanoacrylamide derivatives: DNA cleavage, DNA/BSA binding properties and their anticancer behaviour against colon and breast cancer cells. BMC Chemistry 18, 183 (2024). https://doi.org/10.1186/s13065-024-01284-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13065-024-01284-2

Keywords