Skip to main content
Nanomedicine logoLink to Nanomedicine
. 2024 Jul 16;19(18-20):1629–1641. doi: 10.1080/17435889.2024.2367958

Green-synthesized copper oxide nanoparticles induce apoptosis and up-regulate HOTAIR and HOTTIP in pancreatic cancer cells

Zahra Hosseini a, Amirhossein Ahmadi a,*, Ahmad Shadi a, Seyed Javad Hosseini a,b, Hossein Nikmanesh c
PMCID: PMC11389748  PMID: 39011923

Abstract

Aim: Cu2O nanoparticles were synthesized using an extract from S. latifolium algae (SLCu2O NPs). Their effect on PANC-1 cells and the expression of two drug resistance-related lncRNAs were evaluated in comparison with Arsenic trioxide.

Materials & methods: SLCu2O NPs were characterized using XRD, SEM, and TEM microscopies. The effects of SLCu2O NPs on cell cytotoxicity, cell cycle, and apoptosis, and expression of two drug resistance-related lncRNAs were examined using MTT assay, flow cytometry, and real-time PCR, respectively.

Results: SLCu2O NPs demonstrated anti-cancer properties against PANC-1 cells comparable to Arsenic trioxide, and the expression of lncRNAs increased upon treatment with them.

Conclusion: SLCu2O NPs demonstrate anti-cancer properties against PANC-1 cells; however, using gene silencing strategies along with SLCu2O NPs is suggested.

Keywords: : algae, arsenic trioxide, chemoresistance, copper, long non-coding RNA

Plain language summary

Article highlights.

  • Copper nanoparticles have been proposed as a novel anti-cancer agent in cancer research including Pancreatic cancer.

  • Copper oxide nanoparticles synthesis via physicochemical routs restricted their use in further cancer research or future clinical applications due to using highly toxic chemicals, high energy consumption, and high cost of synthesis.

  • In this study, Cu2O nanoparticles were synthesized using an extract from S. latifolium algae to address environmental concerns.

  • The XRD patterns of the S. latifolium-derived Cu2O nanoparticles showed the prominent peaks attributed to planes (110), (111), (200), (220), (311), and (222) that confirm the formation of the single-phase cubic structure with Pn-3m space group.

  • Scanning electron microscopy confirm the spherical-cubic shape of S. latifolium-derived Cu2O nanoparticles with the average size of 40–50 nm.

  • S. latifolium-derived Cu2O nanoparticles induced cell cytotoxicity on PANC-1 cells with IC50 of 72.75 and 65.4 μg/ml at 24 and 48 h, respectively.

  • S. latifolium-derived Cu2O nanoparticles induced S-phase cell cycle arrest and apoptosis in PANC-1 cells.

  • S. latifolium-derived Cu2O nanoparticles showed comparable effects to arsenic trioxide in terms of cell cytotoxicity, colony formation, cell cycle arrest, and apoptosis.

  • S. latifolium-derived Cu2O nanoparticles increased the expression of two drug-resistant related lncRNAs, HOTAIR and HOTTIP, which may suggest using gene silencing strategies along with these nanoparticles.

1. Background

Pancreatic cancer (PC) is one of the deadliest cancers with 5-year survival rates of 4% [1]. Chemoresistance poses a formidable hurdle in effective PC treatment, with chemotherapy and targeted therapies exhibiting limited efficacy. The mechanism of chemoresistance in PC is very complex, and there are many uncertain cases [2]. Recently, the role of long non-coding RNAs (lncRNAs), notably HOTAIR and HOTTIP, in the development of chemoresistance in PC has been considered [3–6]. HOTAIR, transcribed from the antisense strand of the homeobox C gene on chromosome 12, exerts transcriptional regulatory functions [7]. HOTTIP encoded from a genomic region at the end of the 5′ of the HOXA locus on chromosome 7 and acts as an oncogenic lncRNA in almost all cancers [8]. Studies have highlighted the upregulation of HOTAIR and HOTTIP in PC tissues and cell lines. Their knockdown enhances the chemosensitivity of human PC cells to agents like gemcitabine, a commonly used chemotherapeutic drug [9,10]. Furthermore, the expression of these lncRNAs can be induced by chemotherapeutic agents that may suggest their role in drug resistance mechanism in PC. These lncRNAs might foster chemoresistance by impeding cell apoptosis, promoting epithelial–mesenchymal transition, augmenting cell self-renewal, evading cell cycle checkpoints, modulating cell autophagy, influencing DNA repair, altering efflux pump function, and shaping the tumor microenvironment [11].

Several chemotherapeutic drugs, such as gemcitabine/nab-paclitaxel and 5-fluorouracil/leucovorin with irinotecan and oxaliplatin, can improve the prognosis of pancreatic cancer. However, the development of chemoresistance still leads to poor clinical outcomes [12]. To overcome chemoresistance, researchers are focusing on finding new drugs or repurposing existing ones in combination with chemotherapeutic agents for the treatment of PC [13]. For example, arsenic trioxide (ATO), an FDA-approved drug for acute promyelocytic leukemia, has been shown to be effective alone or in combination with other conventional chemotherapy drugs on PC cells [14]. Thus, the search for new therapeutic substances continues to combat mechanisms underlie drug resistance in PC treatment [15].

Metal nanoparticles (MNs), derived from metal-containing compounds, have garnered attention due to their potential in overcoming chemoresistance. MNs can reduce chemoresistance by disrupting the cellular redox balance, which is the equilibrium between the generation and elimination of reactive oxygen species (ROS) in cancer cells [16]. Additionally, MNs could enhance drug sensitivity by altering the tumor microenvironment through the reorganization of the extracellular matrix and by activating the immune response [17].

Copper oxide nanoparticles as one of the metal nanoparticles, show selective cytotoxicity to several cancer cells and suppress tumor cell proliferation [18]. These nanoparticles can regulate signaling pathways and affect cell movements through ROS production [19]. However, the synthesis of copper oxide nanoparticles via physicochemical routes restricts their use in further cancer researches or future clinical applications due to using highly toxic chemicals, high energy consumption, and high cost of synthesis [20]. Hence, the demand for synthesizing copper oxide nanoparticles using simple, cost-effective, and environmentally friendly approaches, such as green synthesis using natural extracts as reducing agents, has surged [21]. Various biological sources have been explored for their capabilities in the biosynthesis of copper oxide nanoparticles [22,23].

Sargassum latifolium is a brown alga belonging to the order Fucales and mainly exists in tropical and temperate waters. S. latifolium has a good habitat for other marine plants and animals and often comes ashore in masses with flowing water [24]. The S. latifolium extracts have been used to synthesize different kinds of metal nanoparticles [25,26]. Yet, the precise cytotoxic effects of S. latifolium-derived Cu2O nanoparticles compared with ATO on PC cells, alongside their impact on drug-resistant lncRNAs, remain largely unexplored.

The present study aimed to biosynthesize copper oxide nanoparticles using S. latifolium extract and evaluate their anti-cancer activity against PC cells in comparison to ATO, a known candidate for reducing chemoresistance in PC treatment research. Additionally, this study assessed the effects of biogenic copper (I) oxide (Cu2O) nanoparticles on the expression of HOTAIR and HOTTIP lncRNAs.

2. Materials & methods

2.1. Materials

All chemicals, including copper (II) sulfate (CuSO4), sodium citrate (Na3C6H5O7), and sodium carbonate (Na2CO3), were purchased from Merck, Germany. Arsenic (III) oxide was purchased from Sigma, USA. All culture vessels were obtained from SPL Life Science, South Korea.

2.2. Collecting S. latifolium & preparing algae extract

The brown macro-alga S. latifolium was collected from the Persian Gulf on the coast of Bandar Bushehr, Iran (28° 58′ N 50° 50′ E). The S. latifolium algae were washed in running tap water to remove filth and dust and rinsed with distilled water. Algae were then left to dry at room temperature for about 72 h and then were incubated in an oven at 40°C for 48 h. Using a mechanical grinder, the dried algae were ground into powder. Subsequently, 5 g of algae powder was added in to a 250 ml of Erlenmeyer flask containing 200 ml double-distilled water. Then, the Erlenmeyer flask was placed on a heater stirrer for 20 minutes at 100°C with gentle stirring using a magnet. At the end of the boiling time, the algae extract was passed through filter paper and then centrifuged at 6000 rpm for 10 minutes to obtain a clear solution [27,28].

2.3. Biosynthesis of Cu2O NPs using Sargassum algae extract

The procedure for synthesizing Cu2O NPs using S. latifolium algae extract is presented in Figure 1. Nanoparticles were synthesized using Benedict's solution and S. latifolium algae extract, as reported previously [29]. Benedict's solution was prepared as described before with slight modifications [30]. Briefly, 4 g of copper (II) sulfate salt was completely dissolved in 20 ml double-distilled water. Subsequently, 25 g of sodium citrate was added to the solution and completely dissolved, which turned the solution dark blue. Then, 10 g of sodium carbonate was added to the solution, and the volume of the final solution was increased to 40 ml with double-distilled water.

Figure 1.

Figure 1.

The step-by-step procedure of the Cu2O NPs biosynthesis using the S. latifolium algae extract.

To synthesize nanoparticles using S. latifolium algae extract, a beaker containing 40 ml of the Benedict's solution was placed on a heater stirrer at 100° C and the solution was mixed until it reached boiling point. Subsequently, 3 ml of S. latifolium algae extract was added dropwise using a micropipette in several steps. By adding the extract, the color of the solution gradually changed from blue to light green, and the reddish-orange particles appeared. Stabilizing the color of the solution indicates the end of the reaction. Then, the solution was centrifuged for 10 minutes at 6000 g to precipitate Cu2O nanoparticles. The supernatant was discarded, and the nanoparticles were washed with double-distilled water and centrifuged at 6000 g. This step was repeated three-times. Finally, the precipitated nanoparticles were completely removed and placed in an oven at 60°C for 6 h to dry. For simplicity, the S. latifolium-derived Cu2O nanoparticles were named SLCu2O NPs in the rest of the manuscript.

2.4. Characterizing SLCu2O NPs

The crystalline structure of SLCu2O NPs was analyzed by advanced power x-ray diffraction (XRD), model D8 (Ultima IV Rigaku, Germany), with a Cu·Kα radiation source (1.54 Å) operating at 40 kV applied voltage and 30 mA current. XRD was performed at a scanning rate of 5° min-1. The diffractogram was recorded between 2θ = 10° and 80°. Transmission electron microscopy (TEM) and scanning electron microscope (SEM) were applied to determine the shape and surface morphology of the SLCu2O NPs.

2.5. Cell culture

Pancreatic cancer cell line (PANC-1) was purchased from the National Center for Genetic and Biological Resources (Tehran, Iran) and cultured in high glucose DMEM medium (BioIdea, Iran) supplemented with 10% FBS (Biosera, France). The cells were incubated at 37°C in a humidified incubator with 5% CO2 and 95% humidity. To detach cells, the culture media was removed and cells were washed with PBS. Then, 1 ml trypsin (0.025%)/EDTA (0.01%) solution was added, and cells were incubated for 5 minutes. The effect of trypsin was blocked by adding 2 ml DMEM containing 10% FBS.

2.6. MTT assay

To assess the cytotoxicity of the SLCu2O NPs, the MTT assay was used following the manufacturer's protocol (BioIdea, Iran). The optimal cell number for this assay was determined according to the manufacturer's instructions. To evaluate the cytotoxic effect of SLCu2O NPs on PANC-1 cells, 8 × 103 cells were cultured in each well of 96-well plate. After 24 h, the cells were exposed to different concentrations of SLCu2O NPs and arsenic trioxide at 24 and 48 h. Subsequently, the culture media was replaced with 100 μl serum- and phenol red-free DMEM media and 10 μl of MTT solution (5 mg/ml) was added. Then, the cells were incubated in the incubator for 4 h. Afterward, 50 μl of dimethyl sulfoxide was used to lyse the cells and dissolve the purple crystals of formazan. Using a 96-well spectrophotometer, the optical density (OD) of the plate wells was read at 570 nm [31]. The cytotoxicity of samples on cells was expressed relative to the control cells. The MTT assay was performed three-times with six replicates per treatment. To calculate the half maximal inhibitory concentration (IC50) of SLCu2O NPs, six concentrations of SLCu2O NPs (5, 10, 15, 30, 45 and 90 μg/ml) were used. The IC50 value was determined using the linear regression equation Y = Mx + C, where Y was set to 50 and x was the IC50 value.

2.7. Colony formation assay

The colony formation assay was performed as a golden standard method to evaluate the cell cytotoxicity. Briefly, 2.7 × 104 PANC-1 cells were seeded into a 6-well plate and incubated for 24 h at 37°C. Cells were then exposed to arsenic trioxide and/or nanoparticles for 24 h at concentrations of 20 μM and 65 μg/ml or left untreated as control cells. Then the culture medium was removed, and cells were washed with PBS and trypsinized. Subsequently, 350 cells from each well in the first plate were transferred to the wells of a new 6-well plate. After 13 days, the cells were stained with violet crystal dye and colonies with over 50 cells were counted under a magnifying glass. We calculated the colonization efficiency (PE) and survival fraction (SF) using these equations. PE = (Number of cultured cells/ number of colonies formed) × 100, and SF = (number of cultured cells/ number of colonies formed after treatment) × PE.

2.8. Evaluation of apoptotic & necrotic cells by Annexin V-FITC/ PI staining

Labeling of cells with fluorescein isothiocyanate (FITC)-conjugated Annexin V (Annexin V-FITC) together with propidium iodide (PI) is a common method to measure apoptosis and necrosis using flow cytometry. Annexin V-FITC can recognize phosphatidylserine on apoptotic cells, whereas PI as a DNA binding dye can only enter cells when their membranes are ruptured. The apoptosis detection kit (Biolegend, USA) was used to measure apoptotic and necrotic cells following the manufacturer's protocol. Briefly, 2 × 105 cells were cultured in a 6-well plate until reaching 70% confluence. Cells were then treated with 20 μM Arsenic trioxide or 65 μg/ml Cu2O nanoparticles or left untreated as control cells for 48 h. Then, the cells were washed twice with phosphate buffer saline, and 4 × 105 cells resuspended in 200 μl Annexin V binding buffer. Subsequently, 100 μl cell suspension was transferred to new microtube and 10 μl of PI and 5 μl of Annexin V-FITC solution was added and incubated for 15 min at room temperature. Finally, 400 μl binding buffer was added. The stained cells were analyzed using flow cytometry (BD Biosciences, USA) and FlowJo software to measure healthy cells (PI-,Annexin-V-), early apoptotic cells (PI-,Annexin-V+), necrotic cells (PI+,Annexin-V-), and necrotic or late apoptotic cells (PI+,Annexin-V+). The samples were analyzed by flow cytometry. All tests were performed in triplicates.

2.9. Flow cytometry-based cell cycle assay

Using flow cytometry, the cell cycle can be examined based on the DNA content of the cell using PI staining. To perform this test, the dye solution was prepared by mixing 20 μl of PI (1 mg/ml), 1 μl of Triton-X100 and 20 μl of RNaseA (5 mg/ml) and 950 μl PBS. 2 × 105 cells were seeded in a 96 well-plate. The cells were harvested 48 h using trypsin after treatment with 20 μM arsenic trioxide or 65 μg/ml nanoparticles. After centrifugation at 200 g, the cells were washed with cold PBS and fixed with 70% ethanol. After removal of ethanol, 300 μl of PI dye solution was added to the cell precipitate and cells were analyzed using flow cytometry. All tests were performed in triplicates. Using FlowJo software, the percentage of the cell population at different stages of the cell cycle was determined.

2.10. Real-time PCR

To analyze the expression of HOTAIR and HOTTIP lncRNAs real time PCR was performed. To do this, 2 × 105 cells were seeded in a 6-well plate. The cells were treated for 24 h with 20 μM arsenic trioxide or 65 μg/ml nanoparticles, then the supernatant was removed and washed with PBS. RNA extraction was performed using RNX plus solution (Sinaclone, Iran) according to the manufacturer's instructions. The extracted RNAs were treated with DNase I before cDNA synthesis. In order to synthesize cDNA, the Sinaclon First Strand cDNA Synthesis Kit-50T was used.

Real-time PCR was performed using the Rotor Gene Q system (Qiagen, USA), 2X RealQ Plus 2x Master Mix Green Without ROX (Ampliqon, Denmark), and 10 ng of cDNA template. The following primers were used in the PCR reactions: HOTTIP: Forward primer: 5′-CTGGTGAGGGGAGCTGAGTACG-3′, Reverse primer: 5′-GTCAGCGGCCAACCTTGATGC-3′; HOTAIR: Forward Primer: 5′-AGCACCCACCCAGGAATCCAC-3′, Reverse primer: 5′-AGGGTCCCACTGCATAATCACTCC-3′; ACTB: Forward primer: 5′-AGCCTTCCTTCCTGGGCATGG-3′; Reverse primer: 5′-AGCACTGTGTTGGCGTACAGGTC-3′. The thermal reaction conditions were as follows: initial denaturation at 95°C for 15 min, followed by 35 cycles of denaturation at 95°C for 15 s, annealing at 60°C for 30 s, and extension at 72°C for 20 s. Real-time PCR was carried out in the final reaction volume of 20 μl. To control DNA contamination, a no reverse transcribed RNA sample was used, and a negative control sample containing no cDNA template was used to check PCR false amplification for each gene amplification in each run of real-time PCR.

2.11. Statistical analysis

Statistical analyses were conducted using GraphPad Prism software (version 9). All data were presented as means ± standard deviation (SD) of three independent biological experiments. Comparisons between groups were performed using analysis of variance (ANOVA) followed by post-hoc tests or independent t-tests, as appropriate. Gene expression was analyzed by the 2-ΔΔCt method. Differences were considered significant at p <0.05.

3. Results

3.1. XRD analysis confirmed the formation of SLCu2O NPs

The XRD patterns of the SLCu2O nanoparticles are shown in Figure 2. The prominent peaks attributed to planes (110), (111), (200), (220), (311), and (222) show the formation of the single-phase cubic structure with the Pn-3m space group according to the standard Cu2O powder diffraction data (JCPDS card no. 01-074-1230). The XRD analysis with the Rietveld refinement through the Material Analysis Using Diffraction (MAUD) program confirmed a single-phase formation without evidence of a sondary phase of cubic structure with the experimental lattice parameter (a = 4.25Å), calculated density (6.18 g/cm3), measured density (5.92 g/cm3) and cell volume (76.87 106 pm3).

Figure 2.

Figure 2.

Characterization of SLCu2O NPs. The XRD analysis with the Rietveld refinement through MAUD program confirmed a single-phase crystal structure.

Also, the average crystallite size (D) of sample was calculated using the Williamson–Hall equation [32,33]:

βcos(θ)=kλD+4εsin(θ)

where β is the full-width at half maximum (FWHM) of the XRD peaks, θ is the Bragg angle, λ is the x-ray radiation wavelength, and k is the shape factor changing in the range from 0.9 to 1.15 (≈0.9 for spherical-shaped crystallites). The average crystallite size of SLCu2O nanoparticles was found to be 22.2 nm.

3.2. SEM confirmed the spherical shape of SLCu2O NPs

The morphological analyses of the specific sample SLCu2O nanoparticles were carried out using scanning electron microscopy (SEM). The SEM micrograph is provided in Figure 3A. The average size of the nearly spherical-cubic nanoparticles was found to be in the range of 40–50 nm, although the nanoparticles are agglomerated.

Figure 3.

Figure 3.

Electronic microscopic images of SLCu2O NPs. (A) SEM images of Cu2O NPs synthesized in the presence of sargassum algae extract at the 100 Kx magnification. (B) TEM image of SLCu2O NPs at the 50 nm scale.

3.3. TEM confirmed that SLCu2O NPs mean particle size is less than 50 nm

Figure 3B shows the transmission electron microscopy (TEM) image of the SLCu2O nanoparticles. It is evident from this figure that the majority of the particles are rounded with the diameters less than 100 nanometers (mean particle size is less than 50 nm). Considering the matter that each particle contains several crystallites, the particle size obtained by SEM and TEM images confirms the validity of the crystallite size calculated by XRD.

3.4. SLCu2O NPs induced cell cytotoxicity on PANC-1 cells

MTT assay was performed to study the cytotoxic effect of the SLCu2O NPs on PANC-1 cells. The results showed that Cu2O NPs reduced the viability of PANC-1 cells in a concentration-dependent manner at 24 and 48 h (Figure 4A). A significant increase in cell cytotoxicity was observed in concentrations over 45 μg/ml. The IC50 of SLCu2O NPs on PANC-1 cells was 72.75 and 65.4 μg/ml at 24 and 48 h, respectively. The cytotoxic effect of SLCu2O NPs was compared with ATO on PANC-1 cells. The ATO IC50 at 48 h was found to be 20 μM. As shown in Figure 4B, similar cytotoxic effect was observed at the IC50 for SLCu2O NPs and ATO at 48 h. Also, it was observed that the combination of SLCu2O NPs with ATO (65 μg/ml NP and 20 μM ATO) exhibited higher toxicity compared with the individual treatments of ATO and SLCu2O NPs at 48 h (Figure 4B).

Figure 4.

Figure 4.

Cytotoxic effect of SLCu2O NPs and ATO. (A) The OD (mean ± SD) absorbances of dissolved formazan in cells treated with different concentration of Cu2O NPs (n = 5). (B) The OD absorbances (mean ± SD) of dissolved formazan in cells treated with Cu2O NPs or/and ATO (n = 5). SLCu2O NPs showed synergistic cytotoxic effect in combination with ATO. (C) SLCu2O NPs decrease the number and size of colonies relative to untreated cells. (D) The mean ± SD of survival fraction among Cu2O NPs or ATO treated-cells versus untreated cells. SLCu2O NPs and ATO shows comparable effect on the survival fraction of PANC-1 cells.

**p <0.01; ns: Not significant; SD: Standard deviation.

3.5. SLCu2O NPs showed comparable effect to ATO on PANC-1 cell colony formation ability

In order to investigate the cytotoxic effect of SLCu2O NPs on PANC-1 cells, colony formation was performed as a gold standard assay (Figure 4C). As shown in Figure 4D, treatment of cells with 65 μg/ml SLCu2O NPs decreased the survival fraction by 45%. This effect was similar to the effect of 20 μM ATO which decreased the survival fraction by 40%.

3.6. SLCu2O NPs induced the S phase arrest in PANC-1 cells

Evaluating the effects of SLCu2O nanoparticles on cell cycle progression is crucial due to the propensity of cancer cells to proliferate uncontrollably. To study the impact of SLCu2O NPs on the cell cycle of PANC-1 cells, the cells were treated with SLCu2O NPs at a concentration of 65 μg/ml. After 48 h, the cells were stained with PI and analyzed using flow cytometry. Figure 5A & B shows SLCu2O NPs induced the S phase arrest in PANC-1 cells. Figure 5A shows the histograms of cell distribution in different phases of cell cycle and Figure 5B represents the percentage of cell distribution in different phases of cell cycle. As shown in Figure 5B, the treatment with SLCu2O NPs significantly reduced the population of PANC-1 cells in G1 phase by 17.88%. Indeed, the mean percentage of G1 phase-cell population was 49.65% among untreated cells, whereas the mean percentage of G1 phase-cell population was 31.77 among treated cells. A similar effect was observed when the ATO-treated cells were analyzed by flowcytometry. Also, SLCu2O NPs increased the population of S-phase cell population by 24.96%. This effect was similar to that of ATO which also increased the population of S-phase PANC-1 cells by 21.41% (23.78% among control cells vs 52.46% among treated cells).

Figure 5.

Figure 5.

The effect of SLCu2O NPs on PANC-1 cell cycle. (A) the histogram diagrams of cells. (B) The mean ± SD percentage of cell populations among SLCu2O NPs or ATO treated-cells versus untreated cells (n = 3). SLCu2O NPs induces S-phase cell cycle arrest in PANC-1 cells similar to ATO effect.

**p <0.01; n: Number of replicate; SD: Standard deviation.

3.7. SLCu2O NPs induced apoptosis in PANC-1 cells

To study the impact of SLCu2O NPs on PANC-1 cell death, flow cytometry analysis was conducted 48 h after cell treatment using Annexin V-PI staining. Figure 6A & B shows that SLCu2O NPs induced apoptosis in PANC-1 cells. Figure 6A represents the dot plot which shows that SLCu2O NPs induced early and late apoptosis in comparison with the untreated cells. As shown in Figure 6B, the mean percentage of apoptotic cells (early and late apoptosis) in control cells was 4.55%, while in cells treated with SLCu2O NPs was 39.6%. Also, the mean percentage of apoptotic cells was 27.2% among ATO-treated cells. Furthermore, as depicted in Figure 6A, the most dead cells among ATO-treated cells at 48 h were early apoptotic cells, whereas most dead cells among SLCu2O NPs-treated cells were necrotic/late apoptotic cells. This data shows that the cytotoxic effect of SLCu2O NPs in PANC-1 cells results in cell apoptosis.

Figure 6.

Figure 6.

The effect of SLCu2O NPs on PANC-1 cell apoptosis and drug resistant-related lncRNAs expression. (A) Flowcytometry analysis of Cu2O NPs and ATO-treated cells compared with untreated cells at 48 h. (B) The mean ± SD percentage of apoptotic cells among Cu2O NPs or ATO-treated cells and untreated cells (n = 3). (C) Alterations in the mean fold changes ± SD of HOTAIR and HOTTIP lncRNAs in Cu2O NPs and ATO-treated cells (n = 3). (SD = standard deviation, n = number of replicates, *p <0.05; **p <0.01; ***p <0.001).

3.8. SLCu2O NPs increased the expression of HOTTIP & HOTAIR LncRNAs

In order to investigate the effect of SLCu2O NPs on the expression of two drug-resistant lncRNAs, real-time PCR were performed. As shown in the Figure 6C, the expression of HOTAIR and HOTTIP was increased by about 2.4 and 2.8-fold, respectively in SLCu2O NPs-treated cells. This effect was comparable to the expression alterations of these two lncRNAs in cells treated with ATO.

4. Discussion

New anti-cancer drugs for PC treatment are being actively pursued through various initiatives. Copper oxide NPs have been found to exhibit anti-cancer effects against various types of cancer, including PC. Advancing research in this field requires affordable and environmentally friendly methods to synthesize copper oxide NPs. The abundance of algae and their metal ion reduction ability make them a promising candidate for green metal NP synthesis [34]. This study is the first to evaluate the anti-cancer activity of S. latifolium-derived Cu2O NPs and their mechanisms of action against PANC-1 PC cells.

In this study, our data demonstrated that S. latifolium algal extract can be utilized as a reducing and capping agent to synthesize Cu2O NPs. This is consistent with previous studies that have shown the potential of algal extracts in the biosynthesis of metal NPs [29,34,35]. Although the components of S. latifolium algal extract were not characterized in the present study, the successful synthesis of SLCu2O NPs suggests that S. latifolium algal extract could contain active compounds such as carbohydrates, vitamins, fats, polyunsaturated fatty acids, antioxidants, pigments, and phycobilins, which have been theoretically described as agents for reducing metal ions and stabilizing nanoparticle synthesis [36].

In our study, the XRD pattern of the synthesized NPs displayed peaks consistent with pure Cu2O NPs without any peaks corresponding to Cu or CuO NPs [37]. This data contrasts with some previous studies that used a green route to synthesize copper NPs, resulting in a mixture of Cu/CuO/Cu2O NPs [38,39]. The variance in our results compared with prior research may stem from differences in the biological sources of reducing and stabilizing agents employed in our study versus those in previous studies [40].

The SEM and TEM microscopic images of SLCu2O NPs revealed their spherical morphology, with most particles measuring less than 50 nm in diameter. These findings are in line with Soureshjani et al.'s study [29], which also reported spherical Cu2O NPs derived from S. latifolium. However, it's worth noting that the average size of Cu2O NPs in their study was smaller than those derived in our study. This discrepancy could potentially be attributed to slight differences in pH, precursor concentration, reaction time, exposure time, and temperature between our study and Soureshjani et al.'s investigation [29].

Here, in this study, our data indicated that SLCu2O NPs have cytotoxic effects on PANC-1 cells with an IC50 of 65.4 μg/ml at 48 h. Additionally, the colony formation assay further confirmed the cytotoxicity of SLCu2O NPs on PANC-1 cells. While there are no previous reports specifically addressing the cytotoxic effect of Cu2O NPs on PANC-1 cells, Kodasi et al. demonstrated cytotoxic effects of Cu2O nanoparticles synthesized using kiwi fruit juice on breast cancer MCF-7 cells, with an IC50 of 6.25 μg/ml at 48 h. Furthermore, the S. latifolium- derived Cu2O NPs showed cytotoxic effects on K562 chronic myeloid leukemia cells, with an IC50 of 30 μg/ml at 48 h [29]. The difference in IC50 might be attributed to variation in drug sensitivity of different cell lines that were used in mentioned studies. Nonetheless, collectively, these reports support our findings regarding the cytotoxicity of SLCu2O NPs.

In another study, Benguigui et al. [15] reported that chemically synthesized CuO NPs showed a cytotoxic effect on PANC-1 cells with an IC50 of 20 μg/ml at 24 h. Consequently, it appears that chemically synthesized CuO NPs are more cytotoxic to PANC-1 cells compared with SLCu2O NPs. The higher cytotoxicity of CuO NPs compared with SLCu2O NPs on PANC-1 cells could be attributed to differences in their stability and mechanism of action [41,42], which could result in distinct patterns of cellular damage.

Our study revealed that SLCu2O NPs possess anti-cancer activity by inducing cell cycle arrest and apoptosis. This finding is supported by recent researches demonstrating the anti-cancer effects of copper oxide nanoparticles against various types of cancer cells. For instance, Karakus et al. reported that oolong tea extract-mediated CuO/Cu2O nanoparticles induced cell death and apoptosis in colon cancer (HCT116) and breast cancer cells (MCF-7) [39]. Bai et al. demonstrated that copper nanoparticles fabricated using Acroptilon repens leaf extract exhibit anti-lung cancer activity through cell cycle arrest, apoptosis, and modulation of mTOR signaling [43]. Additionally, Kodasi et al. showed the anti-cancer activity of Cu2O nanoparticles against MCF-7 breast cancer cells, primarily through apoptosis and inhibition of cell cycle progression [44]. Shinde et al. fabricated heterogeneous Cu/CuO/Cu2O NPs using groundnut shell extract and reported their anti-cancer activity on MCF-7 breast cancer cells by inducing oxidative DNA damage [38]. Mahmood et al. demonstrated that copper oxide nanoparticles synthesized by Annona muricata extract exert anti-cancer activity by inducing apoptosis through triggering caspase 3 and 9 expression in breast cancer cells [45].

Our results indicated that the SLCu2O NPs caused S-phase cell cycle arrest and apoptosis. Agents inducing cell cycle arrest can trigger apoptosis and offer potential therapeutic avenues for cancer treatment, either alone or in combination with other therapies [46]. Apoptosis is important in cancer therapies as helps prevent uncontrolled tumor growth while minimizing damage to healthy tissues [47]. Benguigui et al. [15] reported that CuO NPs induced G2/M cell cycle arrest and apoptosis in PANC-1-derived tumor-initiating cells. Wang et al. [48] reported that CuO NPs induced G2/M arrest and apoptosis in PC-3 cells by modulating the Wnt signaling pathway. Thit et al. [49] demonstrated that CuO NPs induced G2/M arrest in epithelial cells from Xenopus laevis. In contrast, Chen et al. [50] reported that carbon quantum dots/Cu2O composite induced S-phase arrest and apoptosis in SKOV3 ovarian cancer cells. The different effects of copper NPs on cell cycle arrest may be due to the varying cytotoxicity mechanisms of Cu (I) and Cu (II) [51].

In the current study, we did not examine how SLCu2O NPs cause cell cycle arrest and apoptosis. However, previous studies have proposed several mechanisms. Cu2O NPs can increase intracellular ROS generation and oxidative stress, which lead to DNA damages and activates signaling pathways that cause cell cycle arrest, ultimately leading to cell death through apoptosis [52]. S-phase arrest could be caused by the activation of DNA damage response and replication stress through ATR kinase activity, leading to induction of p21 [53,54]. ROS induced by copper NPs are known to stimulate DNA damage response and replication stress by causing oxidative damage [55]. Additionally, it has been reported that copper oxide NPs can target lipoacylated proteins involved in the tricarboxylic acid cycle and induce another type of cell death named cuproptosis, although this was not assessed in our study [56].

This study found that the cytotoxic effect of SLCu2O NPs is is comparable to that of ATO on PANC-1 cells, particularly in terms of cell apoptosis and cell cycle arrest. Additionally, SLCu2O NPs were observed to enhance the toxicity of ATO. Given that both ATO and Cu2O NPs have been reported to induce ROS [57], the increased cytotoxic effect of SLCu2O NPs in combination with ATO may be attributed to the excessive generation of ROS.

In our study, we observed that SLCu2O NPs increased the expression of two drug-resistant lncRNAs, HOTAIR and HOTTIP, similar to the effects of ATO. This finding aligns with the study by Tan et al.'s research [58], which demonstrated the upregulation of HOTAIR under ATO treatment in vivo and vitro settings. The up-regulation of drug-resistant-related lncRNAs in response to chemotherapeutic agents may be a part of the drug resistance mechanism. For example, Hu et al. reported that cisplatin treatment induces HOTAIR expression in bladder cancer cells through EGFR and NF-κB activation, leading to increased cisplatin resistance [59]. Upregulation of HOTAIR contributes to cisplatin resistance by inhibiting cell apoptosis, dysregulating cell cycle, enhancing epithelial-mesenchymal transition, autophagy, and self-renewal of cancer stem cells, interfering with DNA repair, and altering drug efflux pump mechanisms [11]. Similarly, Wang et al. found that Gemcitabine treatment causes resistance of pancreatic cancer stem-like cells via induction of HOTAIR, leading to promotion of proliferation and migration, reduced apoptosis rate, and increased chemoresistance [10]. Consistent with these prior studies, the increase in HOTAIR and HOTTIP expression in response to SLCu2O NPs may be a mechanism employed by PANC-1 cells to enhance their drug resistance. To overcome drug resistance, it is suggested to explore the use of Cu2O NPs in combination with gene silencing technologies for further research [9,60–62], although SLCu2O NPs may still induce chemosensitivity by disturbing redox balance [16].

Our study also has some limitations that need to be addressed in future research. First, we did not evaluate the effect of SLCu2O NPs on ROS production in the PANC-1 cell, even though past studies have indicated that copper NPs can cause cell toxicity and death by generating ROS and leading to DNA damage [19,63]. It is possible that SLCu2O NPs may have a similar mechanism of action to ATO, generating ROS and affecting PANC-1 cells, but further studies are needed to confirm this. Additionally, our study was conducted solely in vitro, and we did not explore the therapeutic window for SLCu2O NPs. However, a previous study reported that the minimum effective dose of CuO NPs against PANC-1 cells in mice was 1 mg/kg Cu, while a dose of 12.5 mg/kg Cu resulted in death of all mice [15]. Therefore, it is imperative to delve into the therapeutic window of Cu2O NPs in future studies.

5. Conclusion

In conclusion, our study demonstrated the anti-cancer effects of SLCu2O NPs against PANC-1 cells, which were comparable to the effects observed with ATO. The resemblance in anti-cancer efficacy highlights the potential of SLCu2O NPs as a promising candidate for pancreatic cancer treatment. To enhance the effectiveness of pancreatic cancer therapy, future endeavors should focus on combining SLCu2O NPs with strategies aimed at mitigating drug-resistant lncRNAs. This integrated approach holds promise for advancing pancreatic cancer treatment modalities.

Acknowledgments

The authors are grateful to the research management of the Persian Gulf University for providing this research opportunity.

Author contributions

Z Hosseini participated in data curation, formal analysis, and writing- original draft preparation. A Shadi, SJ Hosseini and H Nikmanesh contributed to conceptualization, review and editing. A Ahmadi contributed to the supervision of the study, data validation, writing-review and editing of the manuscript.

Financial disclosure

The authors have no financial involvement with any organization or entity with a financial interest in or financial conflict with the subject matter or materials discussed in the manuscript. This includes employment, consultancies, honoraria, stock ownership or options, expert testimony, grants or patents received or pending, or royalties.

Competing interests disclosure

The authors have no competing interests or relevant affiliations with any organization or entity with the subject matter or materials discussed in the manuscript. This includes employment, consultancies, honoraria, stock ownership or options, expert testimony, grants or patents received or pending, or royalties.

Writing disclosure

No writing assistance was utilized in the production of this manuscript.

References

Papers of special note have been highlighted as: • of interest; •• of considerable interest

  • 1.Ryu JK, Hong S-M, Karikari CA, et al. Aberrant MicroRNA-155 expression is an early event in the multistep progression of pancreatic adenocarcinoma. Pancreatology. 2010;10(1):66–73. doi: 10.1159/000231984 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Coughlin SS, Calle EE, Patel AV, et al. Predictors of pancreatic cancer mortality among a large cohort of United States adults. Cancer Causes Control. 2000;11(10):915–923. doi: 10.1023/A:1026580131793 [DOI] [PubMed] [Google Scholar]
  • 3.Li D, Qian X, Xu P, et al. Identification of lncRNAs and their functional network associated with chemoresistance in SW1990/GZ pancreatic cancer cells by RNA sequencing. DNA Cell Biol. 2018;37(10):839–849. doi: 10.1089/dna.2018.4312 [DOI] [PubMed] [Google Scholar]
  • 4.Lang M, Wang X, Wang H, et al. Arsenic trioxide plus PX-478 achieves effective treatment in pancreatic ductal adenocarcinoma. Cancer Lett. 2016;378(2):87–96. doi: 10.1016/j.canlet.2016.05.016 [DOI] [PubMed] [Google Scholar]
  • 5.Gao J, Wang G, Wu J, et al. Skp2 expression is inhibited by arsenic trioxide through the upregulation of miRNA-330-5p in pancreatic cancer cells. Mol Ther Oncolytics. 2019;12:214–223. doi: 10.1016/j.omto.2019.01.006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Xu C, Wang X, Zhou Y, et al. Synergy between arsenic trioxide and JQ1 on autophagy in pancreatic cancer. Oncogene. 2019;38(47):7249–7265. doi: 10.1038/s41388-019-0930-3 [DOI] [PubMed] [Google Scholar]; • Highlights the importance of ATO alone or in combination of other agents in pancreatic ductal adenocarcinoma treatment researches.
  • 7.Cai B, Song X, Cai J, et al. HOTAIR: a cancer-related long non-coding RNA. Neoplasma. 2014;61(4):379–391. doi: 10.4149/neo_2014_075 [DOI] [PubMed] [Google Scholar]
  • 8.Wang KC, Yang YW, Liu B, et al. A long noncoding RNA maintains active chromatin to coordinate homeotic gene expression. Nature. 2011;472(7341):120–124. doi: 10.1038/nature09819 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Li Z, Zhao X, Zhou Y, et al. The long non-coding RNA HOTTIP promotes progression and gemcitabine resistance by regulating HOXA13 in pancreatic cancer. J Transl Med. 2015;13(1):1–16. doi: 10.1186/s12967-015-0442-z [DOI] [PMC free article] [PubMed] [Google Scholar]; • Focuses on the lncRNA HOXA transcript at the distal tip (HOTTIP) and its role in pancreatic ductal adenocarcinoma progression and chemoresistance.
  • 10.Wang L, Dong P, Wang W, et al. Gemcitabine treatment causes resistance and malignancy of pancreatic cancer stem-like cells via induction of lncRNA HOTAIR. Exp Ther Med. 2017;14(5):4773–4780. doi: 10.3892/etm.2017.5151 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Zhu C, Wang X, Wang Y, et al. Functions and underlying mechanisms of lncRNA HOTAIR in cancer chemotherapy resistance. Cell Death Discov. 2022;8(1):383. doi: 10.1038/s41420-022-01174-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Zeng S, Pöttler M, Lan B, et al. Chemoresistance in pancreatic cancer. Int J Mol Sci. 2019;20(18):4504. doi: 10.3390/ijms20184504 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Yauch RL, Gould SE, Scales SJ, et al. A paracrine requirement for hedgehog signalling in cancer. Nature. 2008;455(7211):406–410. doi: 10.1038/nature07275 [DOI] [PubMed] [Google Scholar]
  • 14.Lew YS, Brown SL, Griffin RJ, et al. Arsenic trioxide causes selective necrosis in solid murine tumors by vascular shutdown. Cancer Res. 1999;59(24):6033–6037. [PubMed] [Google Scholar]
  • 15.Benguigui M, Weitz IS, Timaner M, et al. Copper oxide nanoparticles inhibit pancreatic tumor growth primarily by targeting tumor initiating cells. Sci Rep. 2019;9(1):1–10. doi: 10.1038/s41598-019-48959-8 [DOI] [PMC free article] [PubMed] [Google Scholar]; •• Represents one of the foremost research efforts in examining the anti-cancer properties of CuO nanoparticles on pancreatic cancer both in vitro and in vivo.
  • 16.Liu J, Yuan Y, Cheng Y, et al. Copper-based metal-organic framework overcomes cancer chemoresistance through systemically disrupting dynamically balanced cellular redox homeostasis. J Am Chem Soc. 2022;144(11):4799–4809. doi: 10.1021/jacs.1c11856 [DOI] [PubMed] [Google Scholar]; •• Discusses how copper nanoparticles can be applied to overcome drug resistance.
  • 17.Yang R, Chen L, Wang Y, et al. Tumor microenvironment responsive metal nanoparticles in cancer immunotherapy. Front immunol. 2023;14:1237361. doi: 10.3389/fimmu.2023.1237361 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Aishajiang R, Liu Z, Wang T, et al. Recent advances in cancer therapeutic copper-based nanomaterials for antitumor therapy. Molecules. 2023;28(5):2303. doi: 10.3390/molecules28052303 [DOI] [PMC free article] [PubMed] [Google Scholar]; • Comprehensively introduces different types of copper nanoparticles, discusses their mechanisms of action as anti-cancer agents and their therapeutic potential.
  • 19.Angelé-Martínez C, Nguyen KV, Ameer FS, et al. Reactive oxygen species generation by copper(II) oxide nanoparticles determined by DNA damage assays and EPR spectroscopy. Nanotoxicology. 2017;11(2):278–288. doi: 10.1080/17435390.2017.1293750 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Murugan B, Rahman MZ, Fatimah I, et al. Green synthesis of CuO nanoparticles for biological applications. Inorg Chem Commun. 2023;155:111088. doi: 10.1016/j.inoche.2023.111088 [DOI] [Google Scholar]
  • 21.Siddiqui MA, Alhadlaq HA, Ahmad J, et al. Copper oxide nanoparticles induced mitochondria mediated apoptosis in human hepatocarcinoma cells. PLOS ONE. 2013;8(8):e69534. doi: 10.1371/journal.pone.0069534 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Fuku X, Modibedi M, Mathe M. Green synthesis of Cu/Cu2O/CuO nanostructures and the analysis of their electrochemical properties. SN Appl Sci. 2020;2(5):902. doi: 10.1007/s42452-020-2704-5 [DOI] [Google Scholar]
  • 23.Momeni S, Sedaghati F. CuO/Cu2O nanoparticles: a simple and green synthesis, characterization and their electrocatalytic performance toward formaldehyde oxidation. Microchem J. 2018;143:64–71. doi: 10.1016/j.microc.2018.07.035 [DOI] [Google Scholar]
  • 24.Britton-Simmons KH. Direct and indirect effects of the introduced alga Sargassum muticum on benthic, subtidal communities of Washington State, USA. Mar Ecol Prog Ser. 2004;277:61–78. doi: 10.3354/meps277061 [DOI] [Google Scholar]
  • 25.Abo El-Maaty HA, El-Khateeb AY, Al-Khalaifah H, et al. Effects of ecofriendly synthesized calcium nanoparticles with biocompatible Sargassum latifolium algae extract supplementation on egg quality and scanning electron microscopy images of the eggshell of aged laying hens. Poult Sci. 2021;100(2):675–684. doi: 10.1016/j.psj.2020.10.043 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.El-Sheekh MM, Deyab MA, Hassan NI, et al. Green biosynthesis of silver nanoparticles using sodium alginate extracted from Sargassum latifolium and their antibacterial activity. Rend Lincei Sci Fis Nat. 2022;33(4):867–878. doi: 10.1007/s12210-022-01102-8 [DOI] [Google Scholar]
  • 27.Prasad MH, Ramesh C, Jayakumar N, et al. Biosynthesis of bimetallic Ag/Cu2O nanocomposites using Tridax procumbens leaf extract. Adv Sci Eng Med. 2012;4(1):85–88. doi: 10.1166/asem.2012.1124 [DOI] [Google Scholar]
  • 28.Azizi S, Namvar F, Mahdavi M, et al. Biosynthesis of silver nanoparticles using brown marine macroalga, Sargassum muticum aqueous extract. Materials. 2013;6(12):5942–5950. doi: 10.3390/ma6125942 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Taherzadeh Soureshjani P, Shadi A, Mohammadsaleh F. Algae-mediated route to biogenic cuprous oxide nanoparticles and spindle-like CaCO(3): a comparative study, facile synthesis, and biological properties. RSC Adv. 2021;11(18):10599–10609. doi: 10.1039/D1RA00187F [DOI] [PMC free article] [PubMed] [Google Scholar]; •• Many methods in our study referred to this literature, including synthesis of Cu2O nanoparticles using Sargassum latifolium extracts.
  • 30.Ackerman F, Redlich O. On the thermodynamics of solutions. IX. Critical properties of mixtures and the equation of Benedict, Webb, and Rubin. J Chem Phys. 1963;38(11):2740–2742. doi: 10.1063/1.1733582 [DOI] [Google Scholar]
  • 31.Sylvester PW. Optimization of the tetrazolium dye (MTT) colorimetric assay for cellular growth and viability. Methods Mol Biol. 2011; 716:157–168. doi: 10.1007/978-1-61779-012-6_9 [DOI] [PubMed] [Google Scholar]
  • 32.Jafarpour M, Rostami M, Khalkhali SMH, et al. The effect of lanthanum substitution on the structural, magnetic, and dielectric properties of nanocrystalline Mn-Ni spinel ferrite for radio frequency (RF) applications. Phys Lett A. 2022;446:128285. doi: 10.1016/j.physleta.2022.128285 [DOI] [Google Scholar]
  • 33.Nikmanesh H, Jaberolansar E, Kameli P, et al. Structural and magnetic properties of CoFe2O4 ferrite nanoparticles doped by gadolinium. Nanotechnology. 2021;33(4):045704. doi: 10.1088/1361-6528/ac31e8 [DOI] [PubMed] [Google Scholar]
  • 34.Chugh D, Viswamalya VS, Das B. Green synthesis of silver nanoparticles with algae and the importance of capping agents in the process. J Genet Eng Biotechnol. 2021;19(1):126. doi: 10.1186/s43141-021-00228-w [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Acharya D, Satapathy S, Somu P, et al. Apoptotic effect and anticancer activity of biosynthesized silver nanoparticles from marine algae Chaetomorpha linum extract against human colon cancer cell HCT-116. Biol Trace Elem Res. 2021;199(5):1812–1822. doi: 10.1007/s12011-020-02304-7 [DOI] [PubMed] [Google Scholar]
  • 36.Sampath S, Madhavan Y, Muralidharan M, et al. A review on algal mediated synthesis of metal and metal oxide nanoparticles and their emerging biomedical potential. J Biotechnol. 2022;360:92–109. doi: 10.1016/j.jbiotec.2022.10.009 [DOI] [PubMed] [Google Scholar]
  • 37.Chen K, Xue D. pH-assisted crystallization of Cu2O: chemical reactions control the evolution from nanowires to polyhedra. CrystEngComm. 2012;14(23):8068–8075. doi: 10.1039/c2ce26084k [DOI] [Google Scholar]
  • 38.Shinde S, Parjane S, Turakane H, et al. Bio-inspired synthesis and characterizations of groundnut shells-mediated Cu/CuO/Cu2O nanoparticles for anticancer, antioxidant, and DNA damage activities. J Sol-Gel Sci Technol. 2023;106(3):737–747. doi: 10.1007/s10971-023-06109-7 [DOI] [Google Scholar]
  • 39.Karakuş EE, Sert E, Erol A, et al. Enhancement of cytotoxic and apoptotic activity through oolong tea extract-mediated CuO/Cu2O nanoparticles. J Drug Deliv Sci Technol. 2024;94:105512. doi: 10.1016/j.jddst.2024.105512 [DOI] [Google Scholar]
  • 40.Mukherjee A, Sarkar D, Sasmal S. A review of green synthesis of metal nanoparticles using algae. Front Microbiol. 2021;12:693899. doi: 10.3389/fmicb.2021.693899 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Karlsson HL, Cronholm P, Gustafsson J, et al. Copper oxide nanoparticles are highly toxic: a comparison between metal oxide nanoparticles and carbon nanotubes. Chem Res Toxicol. 2008;21(9):1726–1732. doi: 10.1021/tx800064j [DOI] [PubMed] [Google Scholar]
  • 42.Seo Y, Cho YS, Huh YD, et al. Copper ion from Cu2O crystal induces AMPK-mediated autophagy via superoxide in endothelial cells. Mol Cells. 2016;39(3):195–203. doi: 10.14348/molcells.2016.2198 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Bai X, Tang J, Safaralizadeh R, et al. Green fabrication of bioactive copper nanoparticles using Acroptilon repens extract: an enhanced anti-lung cancer activity. Inorg Chem Commun. 2024;164:112393. doi: 10.1016/j.inoche.2024.112393 [DOI] [Google Scholar]
  • 44.Kodasi B, Kamble RR, Manjanna J, et al. Adept green synthesis of Cu2O nanoparticles using Kiwi fruit (Actinidia deliciosa) juice and Studies on their cytotoxic activity and antimicrobial evaluation. JTEMIN. 2023;3:100044. doi: 10.1016/j.jtemin.2022.100044 [DOI] [Google Scholar]
  • 45.Mahmood RI, Kadhim AA, Ibraheem S, et al. Biosynthesis of copper oxide nanoparticles mediated Annona muricata as cytotoxic and apoptosis inducer factor in breast cancer cell lines. Sci Rep. 2022;12(1):16165. doi: 10.1038/s41598-022-20360-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Almalki SG. The pathophysiology of the cell cycle in cancer and treatment strategies using various cell cycle checkpoint inhibitors. Pathol Res Pract. 2023;251:154854. doi: 10.1016/j.prp.2023.154854 [DOI] [PubMed] [Google Scholar]
  • 47.Carneiro BA, El-Deiry WS. Targeting apoptosis in cancer therapy. Nat Rev Clin Oncol. 2020;17(7):395–417. doi: 10.1038/s41571-020-0341-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Wang Y, Yang QW, Yang Q, et al. Cuprous oxide nanoparticles inhibit prostate cancer by attenuating the stemness of cancer cells via inhibition of the Wnt signaling pathway. Int J Nanomed. 2017;12:2569–2579. doi: 10.2147/IJN.S130537 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Thit A, Selck H, Bjerregaard HF. Toxicity of CuO nanoparticles and Cu ions to tight epithelial cells from Xenopus laevis (A6): effects on proliferation, cell cycle progression and cell death. Toxicol In Vitro. 2013;27(5):1596–1601. doi: 10.1016/j.tiv.2012.12.013 [DOI] [PubMed] [Google Scholar]
  • 50.Chen D, Li B, Lei T, et al. Selective mediation of ovarian cancer SKOV3 cells death by pristine carbon quantum dots/Cu(2)O composite through targeting matrix metalloproteinases, angiogenic cytokines and cytoskeleton. J Nanobiotechnol. 2021;19(1):68. doi: 10.1186/s12951-021-00813-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Wang X, Wang WX. Cu(I)/Cu(II) released by Cu nanoparticles revealed differential cellular toxicity related to mitochondrial dysfunction. Environ Sci Technol. 2023;57(26):9548–9558. doi: 10.1021/acs.est.3c00864 [DOI] [PubMed] [Google Scholar]
  • 52.Sajjad H, Sajjad A, Haya RT, et al. Copper oxide nanoparticles: in vitro and in vivo toxicity, mechanisms of action and factors influencing their toxicology. Comp Biochem Physiol Toxicol Pharmacol. 2023;271:109682. doi: 10.1016/j.cbpc.2023.109682 [DOI] [PubMed] [Google Scholar]
  • 53.Gu L, Hickey RJ, Malkas LH. Therapeutic targeting of DNA replication stress in cancer. 2023;14(7):1346. doi: 10.3390/genes14071346 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Wang G-F, Dong Q, Bai Y, et al. Oxidative stress induces mitotic arrest by inhibiting Aurora A-involved mitotic spindle formation. Free Radic Biol Med. 2017;103:177–187. doi: 10.1016/j.freeradbiomed.2016.12.031 [DOI] [PubMed] [Google Scholar]
  • 55.Srinivas US, Tan BWQ, Vellayappan BA, et al. ROS and the DNA damage response in cancer. Redox Biol. 2019;25:101084. doi: 10.1016/j.redox.2018.101084 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Wu S, Wang Q, Li Y, et al. Recent advances in the biomedical applications of copper nanomaterial-mediated cuproptosis. Adv Nanobiomed Res. 2024;4:2400018. doi: 10.1002/anbr.202400018 [DOI] [Google Scholar]
  • 57.Hu Y, Li J, Lou B, et al. The role of reactive oxygen species in arsenic toxicity. Biomolecules. 2020;10(2):240–270. doi: 10.3390/biom10020240 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Tan J, Sun M, Luo Q, et al. Arsenic exposure increased expression of HOTAIR and LincRNA-p21 in vivo and in vitro. Environ Sci Pollut Res Int. 2021;28(1):587–596. doi: 10.1007/s11356-020-10487-8 [DOI] [PubMed] [Google Scholar]
  • 59.Hu C-Y, Su B-H, Lee Y-C, et al. Interruption of the long non-coding RNA HOTAIR signaling axis ameliorates chemotherapy-induced cachexia in bladder cancer. J Biomed Sci. 2022;29(1):104. doi: 10.1186/s12929-022-00887-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Fang S, Gao H, Tong Y, et al. Long noncoding RNA-HOTAIR affects chemoresistance by regulating HOXA1 methylation in small cell lung cancer cells. Lab Inves. 2016;96(1):60–68. doi: 10.1038/labinvest.2015.123 [DOI] [PubMed] [Google Scholar]
  • 61.Yin F, Zhang Q, Dong Z, et al. Lncrna hottip participates in cisplatin resistance of tumor cells by regulating mir-137 expression in pancreatic cancer. Onco Targets Ther. 2020;13:2689. doi: 10.2147/OTT.S234924 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Liu MY, Li XQ, Gao TH, et al. Elevated HOTAIR expression associated with cisplatin resistance in non-small cell lung cancer patients. J Thorac Dis. 2016;8(11):3314–3322. doi: 10.21037/jtd.2016.11.75 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Henson TE, Navratilova J, Tennant AH, et al. In vitro intestinal toxicity of copper oxide nanoparticles in rat and human cell models. Nanotoxicology. 2019;13(6):795–811. doi: 10.1080/17435390.2019.1578428 [DOI] [PMC free article] [PubMed] [Google Scholar]; • Comprehensively reviewed the toxicology of copper oxide nanoparticles in rat and human models.

Articles from Nanomedicine are provided here courtesy of Taylor & Francis

RESOURCES