Introduction

The application of metallic nanoparticles is now a popular solution in various fields, including catalysis, semiconductors, as well as for biomedical uses1,2,3,4. The unique physiochemical properties of the high surface-volume ratio, localized surface plasmon resonance effect, and internalization, owing to the small size of nanoparticles, provide new perspectives for solving a variety of biomedical problems5,6,7,8,9,10.

Among the various applications of nanoparticles in the biomedical field, cancer treatment is a suitable subject. Due to the short outgrowth of tumors, several differences exist between tumor tissue and normal tissue, including low pH value, hypoxia microenvironment, and defective vessels11,12. The abnormal structure and function of tumor vessels with leaks and defects provide sites for nanoparticles to approach the tumor tissues and passively accumulate. This phenomenon is called the enhanced permeability and retention effect13.

Photodynamic therapy (PDT), an innovative and promising therapeutic modality for treating cancers, involves light, photosensitizers, and oxygen around the tissue14,15. Among them, photosensitizers essentially function as catalysts that can be activated by light of a specific wavelength and transfer the energy to surrounding oxygen molecules, resulting in the generation of reactive oxygen species (ROS) via Type I and Type II reactions16,17,18. With light absorption, these specific photosensitizers can undergo an intersystem crossing and further react in two ways: with biomolecules to generate ROS indirectly (Type I reaction) or directly with oxygen, which creates an energy transfer and results in singlet oxygen (1O2) generation (Type II reaction)18. Between them, 1O2 is extremely electrophilic and can directly oxidize electron-rich double bonds in biological molecules and macromolecules. It is believed to be the prime cytotoxic agent related to PDT19.

Gold nanoparticles have been extensively studied within biomedicine because of their photochemical, photophysical, and optical properties20,21,22. In particular, these have been widely used to deliver photosensitizer agents for PDT16,23,24,25. Gold nanoparticles are also found to enhance the singlet oxygen generation rate26,27. In addition, the biocompatibility and low toxicity of gold nanoparticles make it ideal for biomedical applications20,28. Methylene blue is one of the appropriate photosensitizers commonly used in photodynamic therapy due to its outstanding photochemical properties19,29. It has been reported that the monomer species of methylene blue favors the type II pathway30.

Chemodynamic therapy is another way to trigger cancer-cell apoptosis. Ferrous ion (Fe2+) and cuprous ion (Cu+) can catalyze H2O2 to generate highly cytotoxic hydroxyl radicals (·OH) via Fenton and Fenton-like reactions, respectively when tailored to the specific tumor microenvironment, which involves acidity and the overproduction of hydrogen peroxide (H2O2)31,32,33,

Table 1 Metal concentration of Au/Cu ferrite oxide-polymer core–shell NPs. (n = 6).

Figure 2 shows transmission electron microscopy (TEM) images of Au-doped Cu/Fe oxide-polymer nanoreactors prepared with different synthesis parameters (Fig. 2a-d). The structures were confirmed with TEM images. The dark region, representing metal cores, is encapsulated in the light region, representing PSMA. In Fig. 2a-b, for the Au(f)/CuFe(4:1) NPs, the angular AuNPs were formed. For the Au(s)/CuFe(4:1) NPs, the single and round-shaped cores of AuNPs were produced. In Fig. 2c-d TEM images and Fig. 2e,f HRTEM images, the averagely-doped AuNPs were observed, evenly distributed with Cu and ferrite NPs to form multiple-core structures.

Figure 2
figure 2

TEM images of the Au-doped Cu/Fe oxide-polymer nanoreactor synthesized with the different reaction parameters (a) Au(f)/CuFe(4:1), (b) Au(s)/CuFe(4:1), (c) Au(f)/CuFe(1:4), (d) Au(s)/CuFe(1:4), ), (e) HRTEM images of Au(f)/CuFe(1:4), (f) HRTEM images of Au(s)/CuFe(1:4) (g) UV–vis spectra, (h) XRD pattern.

Figure 2g shows the UV–vis spectra of the Au-doped Cu/Fe oxide-polymer nanoreactor. The primary absorption band at ~ 540–580 nm was attributed to the typical SPR property of Au nanostructures in the Au-doped Cu/Fe oxide-polymer nanoreactor. The Au(f)/CuFe NP shows a broad absorption bandwidth after 500 nm, while the Au(s)/CuFe NP shows a relatively narrow one. Additionally, after HCl corrosion, the remaining NPs in the cores were supposed to be AuNPs. In Fig. 2e, Au(f)/CuFe(4:1) NPs transformed into multiple round-shaped and relatively small particles, inferred to be AuNPs. In Fig. S1, a portion of the metal core was removed by HCl, deduced to be Cu and ferrite NPs. In Fig. S1a–d, the cores were still occupied by widespread AuNPs after HCl corrosion, confirming that Au was successfully doped in the core of the nanoreactors. We can see that the absorption peaks contributed by the AuNPs still existed after corrosion by 0.05 M HCl (Fig. S2), which can be inferred that the AuNPs were not etched by the acid. As shown in Fig. 2h, the Au-doped Cu/Fe oxide-polymer nanoreactor in the product was not detected by X-ray diffraction (XRD). The fcc-structured Au material in the resulting crystal indicated that Fe was immobilized on the PSMA nanoparticles instead of releasing Fe ions to react with Au.

To characterize the behavior of Au/CuFe NPs in the aqueous phase, dynamic light scattering was applied to determine the hydrodynamic diameter and zeta potential. In Fig. S3, after the Au do** reaction, the surface charge changes from negative to positive due to the presence of CTAB in the shell of Au/CuFe NPs. To understand the stability of NPs, different Au/CuFe NPs were dispersed in PBS solvent with three different pH values, including 4.0 (acidic), 7.4 (neutral), and 10.0 (basic). The solution was centrifuged at different time intervals, and optical properties were measured (Fig. S4). In the PBS-NPs system, the optical change can be ascribed to salt-induced aggregation. For each group of Au/CuFe NPs, the gradual decrease of the absorption peak in intensity could be observed in pH 4.0 and 7.4 PBS; however, in pH 10.0 PBS, the absorption peaks were quickly weakened after 4 h quiescence.

Based on the characterization test results and the NP-cell interaction performance, CuFe(4:1) NPs, Au(f)/CuFe(4:1) NPs, and Au(s)/CuFe(4:1) NPs were selected and introduced to present the ROS-mediated effect of Au/CuFe NPs. To undergo photodynamic ablation of cancer cells, methylene blue (MB) was applied as a photosensitizer and embedded between the layers of PSMA with the help of a pi-pi stacking interaction, which formed MB-immobilized CuFe NPs and Au/CuFe NPs (MB-CuFe NPs and MB- Au/CuFe NPs). The optical property was changed after MB loading and showed an absorption peak at around 660 nm for both Au(f)/CuFe(4:1) NPs and Au(s)/CuFe(4:1) NPs (Fig. S2), which confirmed the successful loading of MB. In Fig. 3a-b, the hydrodynamic size and surface charge are also changed. A larger structure was formed, and the nanocrystal was more negatively-charged on the surface, which was more obvious for the Au/CuFe NPs that were with the positively-charged surface before MB-immobilization. In Fig. 3c, two groups of Au/CuFe(1:4) NPs showed a similar EE% at around 25%, while CuFe(1:4) NPs showed about 39%. In Fig. 3d, two groups of Au/CuFe(1:4) NPs also showed a similar LD%, slightly lower than that of CuFe(1:4) NPs. The results were advantageous for the upcoming tests due to the relative total mass amount of Fenton catalysts, copper, and iron for the three groups as the MB concentration was fixed. MB-CuFe(4:1) NPs, MB-Au(f)/CuFe(4:1) NPs, and MB-Au(s)/CuFe(4:1) NPs were later diluted with DI water to 10 μM of MB concentration for further use, and the composition of three nanocrystals as shown in Table S1.

Figure 3
figure 3

Characterization of MB-NPs. (a) Hydrodynamic diameter. (n = 3) (b) Zeta potential. (n = 3) (c) Encapsulation efficiency. (n = 3) (d) Loading capacity. (n = 3). (e) The ability of hydroxyl radical generation.

After RNO is added and mixed with the samples, the absorption peak at the wavelength of 440 nm was noticed (Fig. S7a). After 10 min of light irradiation, the reduction of the characteristic peak in intensity could also be observed and measured, implying the generation of singlet oxygen (1O2). In Fig. S7b–c, DI water showed almost no 1O2 generation, while MB showed a certain amount of 1O2 generated, set as 100%. In Fig. S7d–f, the three groups with the same concentration of MB showed the equally matched ability of 1O2 generation, which were all around 75% compared to MB only (Table S2); however, this was expected since MB was fixed in the assay.

To understand which component of the Au/CuFe NPs contributed to the ability of hydroxyl radical (·OH) generation, all the groups with and without immobilized MB underwent a TA test. The concentration of Cu, Fe, Au, and MB was fixed at 1 ppm, 20 ppm, 60 ppm, and 10 uM, respectively. The results were presented by TAOH-contributed fluorescent intensity folds compared to DI water only. (Fig. 3e) As the Cu and Fe concentrations were fixed, CuFe(1:4) NP had the highest ability of ·OH generation among non-MB-immobilized groups due to the highest Fenton catalyst concentration either when Cu or Fe was fixed. As Au concentration was fixed, Au(f)/CuFe(1:4) NPs and Au(s)/CuFe(1:4) NPs showed a similar ability of ·OH generation. However, the Au(f)/CuFe(1:4) NP showed the highest production of ·OH after MB immobilization, whether Fe or MB concentration was fixed.

After 4 h of incubation with MB-CuFe NPs or MB-Au/CuFe NPs, the uptake of MB was quantified (Fig. 4a). Furthermore, a magnetic field (MF) was applied to understand whether or not the behavior could be manipulated by magnetic force. Without a MF, three groups showed a similar uptake of MB at around 18%, and the values were lifted to an average of 24% with MF, thanks to the ferromagnetic property of ferrite. MB was qualified to be fixed for each group in the later cell experiments based on the result. The corresponding cell activity was also estimated (Fig. 4b). With a MF, the cell activity of all groups decreased, which verified the elevated uptake of MB with a MF. Among them, Au(f)/CuFe(1:4) NP showed the highest toxicity, with CuFe(1:4) NP next, and Au(s)/CuFe(1:4) NP at the lowest, which corroborates the result in the TA test, and the toxicity contributed by the chemodynamic effect was also confirmed.

Figure 4
figure 4

(a) Cellular uptake and (b) cell activity after 4 h incubation with and without a magnetic field. (n = 4).

After 24 h of incubation with MB-CuFe NPs or MB-Au/CuFe NPs containing different MB concentrations, the dark cytotoxic activity was estimated. In Fig. 5a, at a low concentration of MB, MB-CuFe(1:4) NP shows the highest toxicity, while at a high concentration of MB, MB-Au/CuFe(1:4) NP becomes the most toxic to cells. Compared to the incubation time of 4 h, the cell activity also decreases due to the long-term action of the chemodynamic effect as the incubation time stretches to 24 h. (Fig. 5b).

Figure 5
figure 5

Dark toxicity of MB-NPs. (a) Cell activity after 24 h of incubation. (b) After 4 h and 24 h of incubation, cell activity with 10 μM MB of MB-NPs. (n = 4).

The ROS generation at different MB concentrations and different incubation times was detected via DCFH-DA assay. The DCF fluorescence intensity was proportional to the concentration of ROS. In Fig. 6a, there is a weak detected signal even at the MB concentration of 10 uM for all groups before light irradiation. However, after light irradiation and another 24 h of incubation, the concentration of ROS is elevated with the increase of MB concentration. Among three groups, CuFe (1:4) NPs and Au(f)/CuFe NPs show the relatively more vigorous DCF fluorescence intensity, compared to Au(s)/CuFe NPs (Fig. 6b). With another 48 h of incubation after laser treatment, the DCF fluorescence could still be detected, which implies the chemodynamic effect continued working (Fig. 6c).

Figure 6
figure 6

The DCFH-DA performance. (a) Without light irradiation. (b) With light irradiation after 24 h of incubation. (c) With light irradiation after 48 h of incubation.

Figure 7
figure 7

(a) Cell activity of NPs with and without light irradiation for 10 min after 24 h and 48 h incubation. (n = 6). (b) The Live/Dead assay of the groups without light irradiation. (c) The Live/Dead assay of the groups with light irradiation after 24 h. (d) The Live/Dead assay of the groups with light irradiation after 48 h.

Photoinduced toxicity was estimated by MTT assay and Live/Dead assay to understand the numerical and morphological presentation of cell activity before and after light irradiation. The photoinduced groups were tracked for a further 24 and 48 h (Fig. 7a). Without the help of a laser, each group showed weak toxicity resulting from a chemodynamic effect. In contrast, the toxicity increased after light irradiation, which was averagely a drop of 30% in cell activity contributed by the photodynamic effect. After the irradiation, the ROS-mediated effect still occurred, resulting in second-stage cell apoptosis led by a chemodynamic effect, especially for the groups with 10 uM of MB. With the light irradiation after 24 h, each group showed matchable toxicity. However, it could still be observed that the Au(f)/CuFe(1:4) NP had the highest toxicity, with the CuFe(1:4) NP the next, and the Au(s)/CuFe(1:4) NP the lowest as the previous result. However, with the elevation in doses and the passing of time, the toxicity of the Au(f)/CuFe(1:4) NP was gradually revealed, which showed the lowest cell activity at around 39% after 48 h of incubation. Compared to the CuFe(1:4) NP and Au(s)/CuFe(1:4) NP, it was determined to be the best material for performing the ROS-mediated effect.

The cell activity of HeLa cells could also be morphologically observed by Live/Dead assay. Before light treatment, there were more live cells than dead ones in each group. (Fig. 7b) After light treatment and another 24 h of incubation, the number of live cells decreased, and the dead cells increased with the elevation of doses of MB concentration (Fig. 7c). When the incubation time was stretched to 48 h, the amount of the live cells increased for the groups with low doses of NPs because of cellular differentiation. Nevertheless, for the groups with high doses of NPs, the amount of the dead cells continued increasing with the help of the second-stage chemodynamic effect (Fig. 7d).