The X-ray diffraction (XRD) pattern for the [email protected]/TSC nanoparticles was displayed in Figure 1. The XRD pattern for [email protected]/TSC showed the typical reflection peaks at 2θ values of 32.3°, 38.6°, 48.7°, 58.1.0° and 68.0° corresponding to (110), (111), (202), (202) and (220) planes of CuO nanoparticles. The reflections could be indexed to face-centered cubic phase of CuO, in accordance with the JCPDS card no. 80-1268 [13]. The sharpness and the intensity of the peaks imply the high crystallinity of the [email protected]/TSC NPs.
The FT-IR spectra of [email protected] and [email protected]/TSC NPs in the range of 400 cm–1 to 4000 cm–1 were presented in Figure 2. There are two sharp absorption peaks at 427 cm–1 and 535 cm–1 that are corresponding to the stretching vibrations of metal–O bonds supporting the presence of the monoclinic phase of copper oxide NPs [14]. The band at 1607 cm–1 corresponds to the C=O vibration of −COOH groups for glutamic acid on the surface of CuO NPs.
The band at 3392 cm–1 corresponding to the O-H stretching vibrations show the vibrational modes of water molecules adsorbed on the surface of the NPs. Also, the peak at around 1955 cm–1 in FT–IR spectrum of [email protected]/TSC is associated with the O-H bending. The comparison of FT‒IR spectra of synthesized [email protected] and [email protected]/TSC confirms the successful conjugation of thiosemicarbazide compound on the surface of [email protected] NPs. The strong band observed at 1339 cm−1 in the spectrum of [email protected]/TSC can be devoted to the ν (C‒N) stretching of thiosemicarbazone in the synthesized nanoparticles. The comparison of FT‒IR spectra of the [email protected] NPs and synthesized [email protected]/TSC confirms the successful conjugation of this organic compound on the surface of [email protected] NPs. For instance, the band around 800 cm−1 in the IR spectrum is corresponding to ν(C=S) stretching [11, 14].
The morphology and size of the [email protected]/TSC NPs were evaluated by scanning and transmission electron microscopy. The particles were approximately spherical with the particle size in a range of 20-39 nm. Figures 3 and 4 display the SEM and TEM images of the [email protected]/TSC NPs.
Energy dispersive X-Ray (EDX) was employed to analyze the Elemental composition of [email protected]/TSC NPs. The EDX analysis showed strong signals for the presence of Cu element, and also signals for the C, O, N, and S atoms. The presence of no other EDX signal clearly indicates the purity of the synthesized NPs (Figure 5, Table 1). Moreover, the hydrodynamic size and zeta potential of [email protected]/TSC NPs were measured 710 nm and 27.5 mv, respectively. The proper zeta potential of the [email protected]/TSC NPs could provide sufficient repulsive force between particles to avoid agglomeration of the NPs and provides satisfactory colloidal stability [15] (Figure 6).
Table 1
EDS analysis of [email protected]/TSC NPs
Elements
|
K
|
Kr
|
Weight (%)
|
Atomic (%)
|
ZAF
|
C
|
0.0707
|
0.0357
|
17.9
|
31.72
|
0.1996
|
N
|
0.0720
|
0.0364
|
17.2
|
26.14
|
0.2114
|
O
|
0.0828
|
0.0418
|
17.1
|
22.78
|
0.2442
|
S
|
0.1689
|
0.0853
|
10.2
|
6.78
|
0.8357
|
Cu
|
0.6055
|
0.3058
|
37.6
|
12.58
|
0.8140
|
Total
|
1.0000
|
0.5050
|
100.00
|
100.00
|
|
The anti-proliferative potential of [email protected]/TSC NPs against AGS and HEK293 cell lines was evaluated. A considerable anti-proliferative effect of [email protected]/TSC NPs against AGS cells was observed with IC50 value of 203µg/mL. Also, the prepared NPs showed a significantly less anti-proliferative effect on the HEK293, as normal human cells, with an IC50 value of 435µg/mL. However, the anticancer potential of the NPs against AGS cells was significantly less than cisplatin, which exhibited the IC50 value of 65µg/mL. The relative viability of the cells exposed to different concentrations of [email protected]/TSC NPs and cisplatin was presented in Figure 7.
As a trace element with specific functions inside the eukaryotic cells, this element has lower toxicity to normal human cells [3]. Also, due to the higher physiological state of cancer cells, they show a higher demand for this element, which could provide selective toxicity of copper-containing drugs against cancer cells [3, 6]. Combination of copper with a variety of therapeutic ligands, including imidazoles and phosphines, as well as TSC derivatives showed a promising anticancer potential [3, 16].
Several anticancer mechanisms have been proposed to be mediated by copper. The anticancer potential of copper has been mainly associated with the oxidative stress caused by the production of reactive oxygen species (ROS) and also, Glutathione (GSH) oxidation [3, 6]. The oxidative imbalance damages cellular molecules, including nucleic acids, lipids, and proteins, which triggers cell apoptosis [6, 17-18].
According to the literature, copper complexes have shown proteasome inhibitory potential in different cell lines and copper element had a crucial role in this issue. The proteasome, a large and multi-catalytic protease, is responsible for the degradation of damaged proteins and also regulating the concentrations of cell cycle regulatory proteins, including cyclins and p53 (144 and 145 from 4). Due to the higher load of copper inside cancer cells than normal cells, a higher proteasome inhibitory potential and thus, apoptosis induction by copper complexes was reported, which efficiently inhibited the proliferation of cancer cells [19-20].
Similar to other metal complexes, copper complexes can undergo redox reaction and ligand substitution which enable them to interact with a variety of biological substrates. Thus, conjugation of therapeutic ligands with copper could provide efficient cytotoxicity against cancer cells by the action of both partners [3, 6].
TSCs are considered as efficient cytotoxic agents, which supress cancer cells proliferation by inhibition of cellular ribonucleotide phosphate reductase, the enzyme necessary for DNA replication [6]. Also, it was reported that the cytotoxicity of copper complexes containing coordinating ligands, including TSCs could be associated to the binding of the ligands to cell DNA which inhibits genome replication and transcription, and finally results in cell apoptosis [6].
Based on the results and above considerations we found that the conjugation of TSC to copper oxide NPs could inhibit the proliferation of cancer cells and results in higher toxicity to AGS cells compared with normal human cells. In addition, the functionalization of [email protected]/TSC NPs using glutamic acid seems to improve their stability and bioavailability. In agreement with our finding, several studies aimed to increase the hydrophilicity of Cu/TSC complexes by modification of TSCs framework to improve their stability and reported promising anticancer characteristics [8, 21].
Moreover, investigating the antitumor effect of copper- 2-keto-3- ethoxybutyraldehyde-bis(thiosemicarbazone) (KTS) revealed that the KTS ligand could stabilize the oxidation state of +2copper. Following uptake of the complex by cancer cells, the complex is dissociated by interaction with cellular thiols. This dissociation generates Cu(I), which may act as heavy metal poison [6, 22]. The similar synergistic potential could be also hypothesized to other CuO-TSC complexes.
Apoptosis induction of [email protected]/TSC NPs in AGS cells
Figure 8 displays the apoptosis of [email protected]/TSC treated and control AGS cells. The results were presented in the diagrams with four regions where Q1 region displays cell necrosis, Q2 and Q3 represent late and early apoptosis, and Q4 for live cells. Our results showed that early apoptosis occurred among the majority of [email protected]/TSC treated cells (69 %) and 11.5 % of cells experienced late apoptosis, which was significantly higher than control cells. In addition, 96.7 % and 19.3 % of control and NPs treated cells remained alive. Also, only 0.27 % of cell necrosis was found among [email protected]/TSC treated cells. The apoptosis induction by copper NPs conjugated to the TSC derivatives against Burkitt’s lymphoma cells has been reported previously, which is in agreement with our finding [22]. As described above, the apoptotic effect of [email protected]/TSC NPs could be associated with the DNA damages caused by oxidative stress, inhibition of DNA replication, and cellular proteasome inhibition [3, 6].
Hoechst 33258 staining
Hoechst staining was performed on [email protected]/TSC treated and control cells to evaluate the presence of compact chromatin of apoptotic nuclei. The apoptosis morphological characteristics, including condensed chromatin and chromatin fragmentation, and apoptotic bodies were found among the AGS cells treated with [email protected]/TSC NPs. In contrast, the control cells did not show considerable changes (Figure 9). As stated above, the increased ROS generation by [email protected]/TSC could results in DNA strand breakage and apoptosis induction [6]. The apoptotic morphological characteristics of [email protected]/TSC treated cells comply well with the results from the MTT and flow cytometry assays.
Caspase 3 activity
The activity of caspase-3 among [email protected]/TSC treated AGS cells and control cells was investigated. The results showed that treating AGS cells with the NPs resulted in the increased activation of caspase-3. In other words, among NPs treated cells the caspase-3 activity significantly increased by 1.4 folds in comparison with the control cells (Figure 10). Caspases play a critical role in triggering of apoptosis [23]. In this study a significantly increased activity of caspase-3 was noticed, confirming the apoptosis induction in AGS cells following exposure to [email protected]/TSC NPs. In agreement with our work, previous studies showed increased transcription and activation of caspase-3 in a variety of cancer cells following treatment with copper NPs or copper complexes [4, 6, 11, 23]. Generation of ROS and subsequently damages to cellular DNA is considered a major signaling for triggering apoptosis pathway [23]. In our opinion, after exposure of AGS cells to [email protected]/TSC NPs, generation of ROS by copper and also inhibition of cell proteasome, which regulates the concentration of p53 proteins, are major causes of apoptosis induction following upregulation of caspase proteins. In agreement with this, previous works reported that the activation of p53 among cancer cells treated with copper NPs indirectly promotes caspase-3 activity [24-25].