Urchin-like piezoelectric ZnSnO3/Cu3P p-n heterojunction for enhanced cancer sonodynamic therapy

Qinyu Zhao Yunchao Zhao Songjing Zhong Zhaoyang Yue Zhuoheng Jiang Shaobo Wang Quanhong Hu Shuncheng Yao Kaikai Wen Linlin Li

Citation:  Qinyu Zhao, Yunchao Zhao, Songjing Zhong, Zhaoyang Yue, Zhuoheng Jiang, Shaobo Wang, Quanhong Hu, Shuncheng Yao, Kaikai Wen, Linlin Li. Urchin-like piezoelectric ZnSnO3/Cu3P p-n heterojunction for enhanced cancer sonodynamic therapy[J]. Chinese Chemical Letters, 2024, 35(12): 109644. doi: 10.1016/j.cclet.2024.109644 shu

Urchin-like piezoelectric ZnSnO3/Cu3P p-n heterojunction for enhanced cancer sonodynamic therapy

English

  • The fabrication of nanomaterials enables the production of cytotoxic reactive oxygen species (ROS), including singlet oxygen (1O2), superoxide anion (O2), and hydroxyl radicals (OH), providing new methods to overcome the limitations of conventional tumor treatments [16]. One notable example is sonodynamic therapy (SDT), which utilizes sonosensitizers combined with ultrasound (US) stimulation to generate ROS, offering minimal invasiveness and high tumor specificity [711]. As a mechanical wave, US has advantages of precise orientation and long-distance propagation [12,13], thereby achieving deeper tissue penetration and higher spatial accuracy to reduce damage to normal tissues [14,15]. However, the development of SDT still encounter challenges related to low-efficiency ROS generation and the complicated tumor microenvironment (TME) [16,17]. Generally, overproduction of glutathione (GSH) in TME can consume the generated ROS, thereby discounting the efficacy of SDT [1,18,19]. Therefore, it is necessary to develop a more efficient sonosensitizers with TME responiveness to increase ROS generation and SDT therapeutic efficiency.

    Recently, piezoelectric semiconductor nanomaterials, such as titanium oxide (TiO2) [20], zinc oxide (ZnO) [21], barium titanate (BaTiO3) [22,23], bismuth tungstate (Bi2WO6) [24,25], and bismuth oxychloride (BiOCl) [26], have gained recognition as a novel class of sonosensitizers for enhancing sonodynamic therapy (SDT). These piezoelectric materials can generate an inherently dynamic electric field under periodic mechanical forces, such as ultrasonic vibrations [2729], which facilitates the separation of electrons (e) and holes (h+), promoting the catalysis of ROS production [30]. Compared to nonpiezoelectric sonosensitizers, this inherently dynamic electric field piezoelectric semiconductor nanomaterials can efficiently reduce the recombination of carriers. And the band bending under US irradiation also promote the carrier separation and ROS production. Zinc stannate (ZnSnO3) has attracted much attention in various piezoelectric materials because of its piezoelectric and ferroelectric properties, with an electrical and structural ordering temperature up to 700 ℃ [3133]. ZnSnO3 boasts an impressive polarization of approximately 59 µC/cm2 along the c-axis, surpassing other non-lead-based piezoelectric materials such as and barium titanate (BaTiO3, 6 µC/cm2), zinc oxide (ZnO, 5 µC/cm2), potassium niobate (KNbO3, 23 µC/cm2) [34,35]. Similar to other piezoelectric semiconductors, the challenge for ZnSnO3 lies in the efficient separation of e and h+ as well as preventing their recombination [36]. Development of ZnSnO3 based piezoelectric sonosensitizers with high efficiency of e and h+ seperation and ROS generation is highly desired.

    In this work, an urchin-like piezoelectric p-n heterojunction was designed and synthesized by integrating p-type ZnSnO3 nanocubes (NCs) and n-type Cu3P nanoneedles (NNs) to form p-n heterojunction for enhancing cancer SDT (Scheme 1). The fabricated ZnSnO3/Cu3P NCs effectively narrow the bandgap of ZnSnO3 NCs, resulting in a higher redox activity to produce ROS. The urchin-like structure can further increase the US response to improve piezoelectric property. Additionally, Cu3P NNs can respond to the TME, thereby consuming intracellular glutathione (GSH) and producing ROS through Fenton-like reaction to achieve CDT. Combining together, the piezoelectric p-n heterojunction realizes a 70% tumor growth inhibition rate on 4T1 tumor mice model by combining CDT and SDT. This piezoelectric p-n sonosensitizer opens the way for the construction and application of heterostructured nanomaterials in improving sonodynamic therapy. ZnSnO3 NCs and Cu3P NNs were prepared separately, and then ZnSnO3/Cu3P heterojunction was formed by hydrothermal method (Fig. 1A) [37]. Scanning electron microscope (SEM) and transmission electron microscope (TEM) analysis revealed that ZnSnO3 exhibited a uniform cubic shape, with average size of 71.9 ± 13.3 nm (Fig. 1B and Fig. S1 in Supporting information). Upon formation heterojunction with Cu3P, the obtained ZnSnO3/Cu3P NCs displayed an urchin-like morphology, with numerous nanoneedles radiating outward from the center (Fig. 1C). High-resolution transmission electron microscopy (HRTEM) image displayed the lattice widths of 0.25 nm corresponding to (102) crystal plane of Cu3P and 0.33 nm corresponding to (012) crystal plane of ZnSnO3 (Figs. 1D and E). Elemental mapping analysis confirmed that Cu and P elements were uniformly distributed on the surface of ZnSnO3 (Fig. 1F and Fig. S2 in Supporting information). The phase structures of ZnSnO3 and ZnSnO3/Cu3P were characterized using X-ray diffraction (XRD), which revealed that only peaks corresponding to ZnSnO3 (JCPDS No. 11-0274, Fig. S4 in Supporting information) was observed in the pattern of ZnSnO3/Cu3P, due to the low content of Cu3P in the heterojunction (Fig. 1G). The chemical composition and valence state of the nanocubes were further analyzed by X-ray photoelectron spectroscopy (XPS). The XPS spectrum of ZnSnO3/Cu3P displayed characteristic peaks of Zn 2p, Sn 3d, Cu 2p and P 2p at 1021.6, 486.2, 933.8 and 129.6 eV, respectively (Fig. 1H and Table S1 in Supporting information). In the high-resolution Cu 2p spectrum (Fig. 1I), the peaks at 932.8 and 934.8 eV corresponded to Cu(I) and Cu(II), respectively. The calculation results of peak area showed that the ratio of Cu(I) to Cu(II) content was approximately 85:15 (Table S2 in Supporting information). The valence states of Zn 2p, Sn 3d and O 1s in ZnSnO3/Cu3P had no difference relative to ZnSnO3 (Fig. 1J and Fig. S5 in Supporting information). Fourier transform infrared spectroscopy (FTIR) spectrum of ZnSnO3/Cu3P showed that the characteristic peak of P−O bond appeared at 1100 cm−1. This result indicates that Cu3P have been successfully synthesized on the surface of ZnSnO3 (Fig. S6 in Supporting information).

    Scheme 1

    Scheme 1.  Schematic depiction of the synthesis process and mechanism in the antitumor process of ZnSnO3/Cu3P NCs.

    Figure 1

    Figure 1.  Synthesis and characterizations of ZnSnO3/Cu3P. (A) Schematic diagram of the fabrication process of ZnSnO3/Cu3P. (B) SEM image of ZnSnO3 NCs. (C) SEM image of ZnSnO3/Cu3P. (D) TEM image, (E) HRTEM image and (F) element mapping of ZnSnO3/Cu3P. (G) XRD, (H) XPS and high-resolution (I) Cu 2p and (J) O 1s XPS spectra of ZnSnO3/Cu3P. (K) hydrodynamic diameters and (L) zeta potential of ZnSnO3 and ZnSnO3/Cu3P. The data represent means ± standard deviation (S.D.) (n = 5).

    Dynamic light scattering (DLS) results illustrated that the mean hydrodynamic diameter of ZnSnO3 was 215.7 nm, while that of ZnSnO3/Cu3P slightly increased to 327.1 nm (Fig. 1K). Zeta potential of ZnSnO3 changed from positive to negative upon the formation of ZnSnO3/Cu3P heterojunction (Fig. 1L). Notably, the positive potential of ZnSnO3 gradually decreased over time, while the zeta potential of ZnSnO3/Cu3P remained stable, indicating better aqueous stability of the heterojunction. After dispersion of ZnSnO3/Cu3P in phosphate buffer saline (PBS) solution at pH 6.5 for 24 h, a partial of nanoneedle on ZnSnO3/Cu3P were fallen off from the nanocubes (Fig. S3 in Supporting information).

    The ROS production of ZnSnO3/Cu3P through sonodynamic, chemodynamic, and their combination processes was investigated (Fig. 2A). To evaluate the sonodynamic performance of ZnSnO3 and ZnSnO3/Cu3P, methylene blue (MB) was chosen as an indicator that can be degraded by OH and superoxide anion (O2). After 10 min of US stimulation (1.0 MHz, 1.0 W/cm2, 50% duty cycle), the presence of ZnSnO3 or ZnSnO3/Cu3P in the MB solutions resulted in significant degradation of MB (Fig. 2B and Fig. S6 in Supporting information), indicating that both ZnSnO3 and ZnSnO3/Cu3P NCs were capable of generating ROS under US stimulation. As shown in Fig. 2C, the first-order kinetic rate constant (k) of MB degradation by ZnSnO3/Cu3P + US was 0.036 (Table 1), higher than that by US (0.011) and ZnSnO3 + US (0.031).

    Figure 2

    Figure 2.  Sonodynamic and chemodynamic performance of ZnSnO3/Cu3P. (A) Schematic representation illustrating the sonodynamic and chemodynamic processes of ZnSnO3/Cu3P. (B) Degradation curves of MB and (C) corresponding kinetic curves in SDT process. Degradation curves of NBT for O2 detection in the SDT process. (E) Generation of OH in the SDT process with TA as the indicator. (F) Production of OH in the CDT process measured with TMB as the indicator. (G) Degradation curves of MB and (H) corresponding kinetic curves in the combination process of SDT and CDT. ESR spectra of (I) OH and (J) O2 trapped by DMPO under different conditions. (K) GSH oxidation by varying concentrations of ZnSnO3/Cu3P within 3 h. (L) Oxidation of GSH achieved by combining SDT and CDT.

    Table 1

    Table 1.  First-order reaction kinetics of MB degradation under various treatments.
    DownLoad: CSV

    To identify the specific types of ROS produced during the SDT process, terephthalic acid (TA) and nitro tetrazolium blue chloride (NBT) were used as specific probes of OH and O2, respectively. After 10 min of US stimulation, the OH generation by ZnSnO3 and ZnSnO3/Cu3P NCs was approximately 1.82-fold and 2.34-fold higher than that in the mere US group (Fig. 2E and Fig. S8 in Supporting information), which correlated well with the results of MB degradation. Similarly, ZnSnO3/Cu3P NCs produced a substantial amount of O2 under US stimulation (Fig. 2D and Fig. S7 in Supporting information). In contrast, the production of O2 ZnSnO3 was substantially lower, similar with that by mere US stimulation. These results demonstrated that ZnSnO3/Cu3P NCs can be excited by US irradiation to efficiently produce ROS, possessing superb SDT performance over ZnSnO3. Also, the partial degradation of Cu3P in a simulated tumor microenvironment (pH 6.5) did not affect the SDT performance of ZnSnO3/Cu3P (Fig. S10 in Supporting information). The specific reactions involved in the sonodynamic process are shown below. Firstly, upon US irradiation, ZnSnO3/Cu3P NCs, undergo polarization, leading to the segregation of electrons (e) and holes (h+) (Eq. 1). Subsequently, nearby oxygen and water molecules react with these e and h+, resulting in the formation of O2 and OH (Eqs. 2 and 3).

    (1)

    (2)

    (3)

    As previous reported, Cu(I) can initiate Fenton-like reaction and efficiently catalyze H2O2 to produce OH, enabling chemodynamic therapy [3843]. Given the presence of Cu(I) in ZnSnO3/Cu3P NCs, we investigated the ROS generation in the presence of H2O2 under weakly acidic conditions. To assess this, 3, 3′, 5, 5′-tetramethylbenzidine (TMB) was selected as a probe, which can react with OH to produce blue oxTMB. When mere H2O2 was added to the weakly acidic medium, the production of OH was negligible. Upon the addition of ZnSnO3/Cu3P NCs, OH generation significantly increased. Furthermore, as the pH values decreased towards a more acidic conditions, OH production gradually increased (Fig. 2F). Hence, ZnSnO3/Cu3P can serve as a potent chemodynamic agent for generating OH, as depicted in Eq. 4.

    (4)

    We further explored the synergistic effect of combining the sonodynamic and chemodynamic processes. We conducted a comparative analysis of MB degradation using ZnSnO3 and ZnSnO3/Cu3P NCs with both H2O2 and US irradiation. In MB solution containing H2O2 (pH 6.5), ZnSnO3/Cu3P NCs induced simultaneous sonodynamic and chemodynamic processes under US irradiation. Remarkably, the integration of SDT and CDT exhibited a notable enhancement in ROS generation using ZnSnO3/Cu3P NCs compared to the single sonodynamic or chemodynamic process (Fig. 2G and Fig. S8 in Supporting information). Consequently, the degradation of MB exhibited a faster first-order kinetic rate constant of 0.098, which was 2.7 times higher than of single SDT (Fig. 2H and Table 2). These finding demonstrated the enhanced efficiency of ROS generation through the combination effect, surpassing either SDT or CDT alone.

    Table 2

    Table 2.  First-order reaction kinetics of MB degradation under various treatments.
    DownLoad: CSV

    Additionally, electron spin resonance (ESR) spectroscopy was utilized to confirm the specific oxygen radicals produced in the combination process. For this purpose, a spin-trapping agent, 5, 5-dimethyl-1-pyrroline N-oxide (DMPO), was utilized in either water or dimethyl sulfoxide to detect OH and O2, respectively. The DMPO/OH adduct exhibits a characteristic quadruple signal with a 1:2:2:1 amplitude ratio, while the DMPO/O2 adduct displays a characteristic multiple signals with a 1:1:1:1:1:1 amplitude ratio. In the presence of H2O2, ZnSnO3/Cu3P displayed the highest signal intensity for both OH and O2 under US irradiation (Figs. 2I and J), verifying its superb combination effect of SDT and CDT.

    In TME, GSH is overexpressed and acts as an antioxidant, which may hinder the therapeutic effect by ROS-induced oxidative damage [44]. Therefore, depletion of GSH in tumor tissue can enhance the oxidative stress and the efficacy of dynamic therapies. Considering the valence-variable nature of Cu, we assessed the capacity of ZnSnO3/Cu3P to consume GSH. Upon incubating the GSH solution with ZnSnO3/Cu3P NCs for 3 h, a significant decrease in GSH concentration was observed (Fig. 2K). Notably, GSH was depleted by 81% at a ZnSnO3/Cu3P concentration of 250 µg/mL, indicating oxidative activity of Cu(II) in ZnSnO3/Cu3P towards GSH. Furthermore, the addition of US stimulation further enhanced the degradation of GSH. When the concentration of ZnSnO3/Cu3P was 250 µg/mL, 10 min of US stimulation (1.0 MHz, 1.0 W/cm2, 50% duty cycle) before 1 h of resting leading to a 1.82-fold improvement in GSH consumption (Fig. 2L). The SDT and CDT properties of ZnSnO3/Cu3P after 3 h of reaction with GSH were measured using MB and TMB as specific indicators. The reaction of ZnSnO3/Cu3P with GSH had negligible impact on the SDT performance of ZnSnO3/Cu3P (Fig. S12 in Supporting information), but improved its CDT performance (Fig. S13 in Supporting information). This may be due to the reduction of Cu(II) to Cu(I) in ZnSnO3/Cu3P NCs, which improved the Fenton-like activity. Therefore, the synergistic effect of SDT and CDT was increased after treated with GSH (Fig. S14 in Supporting information). Taking together, ZnSnO3/Cu3P NCs efficiently utilize the locally produced H2O2 in TME to catalyze the production of OH through CDT. Second, the application of external US stimulation enhances the generation of both OH and O2 through SDT. Third, the GSH consumption capacity of ZnSnO3/Cu3P NCs can reduce the antioxidant activity of tumor cells. This combination holds great potential for augmenting cancer therapy.

    In order to understand the mechanism of the SDT performance of ZnSnO3/Cu3P NCs, the effect of piezotronic effect on carrier migration in ZnSnO3/Cu3P NCs is analyzed in detail (Fig. 3A). First, the band structures of ZnSnO3 and ZnSnO3/Cu3P NCs were studied by ultraviolet−visible diffuse reflection spectra (UV−vis DRS). The bandgaps (Eg) of ZnSnO3 and ZnSnO3/Cu3P NCs were determined by analyzing the intercepts of the tangent and the abscissa in the Kubelka-Munk conversion. The calculated bandgaps were found to be 3.7 eV for ZnSnO3 and 3.5 eV for ZnSnO3/Cu3P (Figs. 3B and C). By examining the XPS valence band spectra, the valence band (VB) positions of ZnSnO3 and ZnSnO3/Cu3P were estimated to be 3.0 eV and 2.2 eV, respectively (Figs. 3D and E). Consequently, the conduction band (CB) positions of ZnSnO3 and ZnSnO3/Cu3P were determined as −0.7 eV and −1.3 eV, respectively.

    Figure 3

    Figure 3.  Mechanism underlying the improved sonodynamic performance of ZnSnO3/Cu3P NCs. (A) Schematic diagram illustrating the separation and migration of electron-hole pairs in both ZnSnO3 NCs and ZnSnO3/Cu3P NCs. UV–vis diffuse reflection spectra and results of Kubelka–Munk conversion of (B) ZnSnO3 NCs and (C) ZnSnO3/Cu3P NCs. XPS valence band spectra of (D) ZnSnO3 NCs and (E) ZnSnO3/Cu3P NCs. PFM phase hysteresis loop, and amplitude butterfly loop of (F) ZnSnO3 NCs and (G) ZnSnO3/Cu3P NCs. (H) COMSOL simulation of influence of surface nanoneedles (n = 0, 1, 3, and 5) on the electric potential under ultrasonic field and (I) statistics of the maximum surface potential.

    The VB edge of pure ZnSnO3 NCs (3.0 eV) was more positive than the H2O/OH redox potential (2.01 eV vs. relative hydrogen electrode (RHE)), verifying that US-triggered h+ on VB have the ability to react with H2O to produce OH. Similarly, the CB level (−0.7 eV) of ZnSnO3 NCs is more negative compared with the redox potential of O2/O2 (−0.28 V vs. RHE), indicating that the e generated by US can react with O2 to produce O2. On the other hand, electron-hole pairs can rapidly recombine in pure ZnSnO3 NCs, which leads to low efficiency of ROS production and SDT.

    For ZnSnO3/Cu3P, the creation of p-n heterojunction effectively inhibited the recombination of e and h+. More importantly, US can induce a built-in electric field that drives e and h+ to opposite surfaces, leading to a band bending, which further promotes ROS generation. The piezoelectric properties of the ZnSnO3/Cu3P heterojunction were assessed by piezoresponse force microscopy (PFM, Fig. S15 in Supporting information). The phase images acquired at an applied voltage of 20 V clearly demonstrate the presence of piezoelectricity of both ZnSnO3 and ZnSnO3/Cu3P (Figs. 3F and G). When a DC voltage ranging from −20 V to +20 V is applied, the PFM phase hysteresis loop exhibits a 180° phase change, indicating a ferroelectric polarization switching process. To further quantify the piezoelectric behavior, the piezoelectric coefficients (d33) of ZnSnO3 and ZnSnO3/Cu3P NCs were measured using a typical amplitude butterfly loop method. The d33 values of ZnSnO3 and ZnSnO3/Cu3P NCs were about 200.0 and 219.5 pm/V, respectively. The results demonstrated that the piezoelectric response of ZnSnO3/Cu3P was slightly enhanced compared with that of ZnSnO3.

    To investigate whether the needle-like structure on the surface of ZnSnO3/Cu3P NCs enhanced the piezoelectric potential under an ultrasonic field, a COMSOL simulation was carried out (Figs. 3H and I). The simulation results showed that the more needle-like structures on the surface of ZnSnO3/Cu3P, the higher the US-induced piezoelectric potential, which may be due to the enhancement of mechanical effect on ZnSnO3 NCs surface caused by needle-like structures. Therefore, we can summarize that ZnSnO3/Cu3P p-n heterojunction can enhance the separation and migration of e and h+ under US irradiation, improving the piezoelectric properties and thus enhancing the sonodynamic properties.

    We validated the killing effect of ZnSnO3/Cu3P NCs on tumor cells through the combination of SDT, CDT, and GSH consumption (Fig. 4A). First, L929 mouse fibroblasts were utilized to assess the cytotoxicity of ZnSnO3/Cu3P NCs to normal cells. L929 cells remained an over 82% survival after 24 h incubation with ZnSnO3/Cu3P NCs when the concentration reached 200 µg/mL (Fig. 4B). The cell viability of 4T1 mouse-derived breast cancer cells decreased to 30% after 24 h incubation with 200 µg/mL ZnSnO3/Cu3P NC, which can be attributed to the CDT effect produced by higher H2O2 concentration in tumor cells. When US irradiation (1.0 MHz, 1.0 W/cm2, 50% duty cycle) was imposed for 3 min, the cell killing rate remarkably increased to 80% (Fig. 4C). As shown in the biological TEM image (Fig. 4G), after endocytosis into the cells, ZnSnO3/Cu3P NCs nanoparticles located in both bilayer phospholipid membrane vesicles and cytoplasm of the cells.

    Figure 4

    Figure 4.  In vitro cell experiments demonstrating the combined SDT and CDT performance of ZnSnO3/Cu3P NCs for cancer cell killing. (A) Schematic representation depicting the mechanism of tumor cell killing by ZnSnO3/Cu3P NCs through the integration of SDT and CDT. (B) Viabilities of L929 cells after treatments with ZnSnO3/Cu3P NCs for 24 h. (C) Viability of 4T1 cells at 24 h post-treatment with different approaches. (D) Confocal images and green/red fluorescence (FL) ratios of Calcein AM/PI-stained 4T1 cells following various treatments. (E) Confocal images and corresponding quantitative analysis of intracellular ROS levels in 4T1 cells stained with DCFH-DA. (F) Confocal images and green/red FL ratios of JC-1-stained 4T1 cells. (G) Bio-TEM image of 4T1 cells after treatment with ZnSnO3/Cu3P NCs for 10 h. The data represent means ± S.D. (n = 5). ***P < 0.001.

    Calcein acetoxymethyl ester (calcein-AM) and propidium iodide (PI) were used for live/dead cell staining to further verify the cancer cell killing effect (Fig. 4D and Fig. S16 in Supporting information). The cell killing effect of mere US irradiation was negligible, while 31% of cells in the ZnSnO3/Cu3P group died due to the CDT effect. In ZnSnO3/Cu3P + US group, the percentage of dead cells was found to be 90%, which correlated with the results obtained from the MTT cell viability assay. 2′, 7′-Dichlorodihydrofluorescein diacetate (DCFH-DA) was used as a fluorescence probe to evaluate the intracellular levels of reactive oxygen species (ROS) to gain further insights into the mechanism of cell killing. DCFH-DA can react with ROS to produce 2′, 7′-dichlorofluorescein (DCF) which has green fluorescence. Compared to the US group, the intracellular fluorescence intensity in the ZnSnO3/Cu3P and ZnSnO3/Cu3P + US groups increased by 2.22-fold and 3.25-fold, respectively, demonstrating the capacity of ZnSnO3/Cu3P NCs to generate abundant ROS via CDT and SDT especially under US activation (Fig. 4E).

    Furthermore, to assess mitochondrial damage, the fluorescent probe 5, 5′, 6, 6′-tetrachloro-1, 1′, 3, 3′-tetraethyl-imidacarbocyanine iodide (JC-1) was employed to measure changes in mitochondrial membrane potential. Under normal conditions, JC-1 shows red fluorescence within intact mitochondrial membranes in the form of aggregates, while shows green fluorescence in damaged or depolarized mitochondria as monomers (Fig. S17 in Supporting information). As depicted in Fig. 4F, the ZnSnO3/Cu3P group exhibited both green and red fluorescence in the mitochondria of 4T1 cells, whereas the ZnSnO3/Cu3P + US group showed the strongest green fluorescence and red fluorescence, indicating significant mitochondrial oxidative damage.

    Prior to investigating the in vivo antitumor activity of ZnSnO3/Cu3P NCs, we evaluated their systemic toxicity. Acute toxicity experiments were performed by intravenously injecting female ICR mice with different doses of ZnSnO3/Cu3P NCs (50, 100, 200 mg/kg). Throughout the 14-day observation period, all mice displayed similar clinical manifestations. The trend of weight change in the treatment groups was similar to that of the control group (Fig. S18 in Supporting information).

    Blood biochemistry assay results showed that there were no abnormal changes in hepatic and renal function markers on the 14th day post-injection, including alanine aminotransferase (ALT), aspartate aminotransferase (AST), alkaline phosphatase (ALP), creatinine (CREA), and UREA, in mice injected with various doses of ZnSnO3/Cu3P NCs compared with those of mice in the control group. Blood routine test indicated normal levels of red blood cells (RBCs), hemoglobin (HGB), hematocrit (HCT), mean corpuscular volume (MCV), mean corpuscular hemoglobin (MCH), and platelets (PLT) in mice administered different doses of ZnSnO3/Cu3P NCs (Fig. S19 in Supporting information). Hematoxylin and eosin (H&E) staining revealed no histological abnormalities in major organs including heart, liver, spleen, lung and kidney in all mice, demonstrating the excellent biocompatibility of ZnSnO3/Cu3P NCs (Fig. S20 in Supporting information).

    The in vivo anticancer efficacy of ZnSnO3/Cu3P NCs under US stimulation was assessed using a subcutaneous 4T1 tumor-bearing BALB/c mice model. All the procedures during the experiment firmly stuck to the "Beijing Administration Rule of Laboratory Animals" and the national standards "Laboratory Animal Requirements of Environment and Housing Facilities (GB 14925-2001)". The animal experiments were approved by the Committee on Ethics of Beijing Institute of Nanoenergy and Nanosystems (Approval Number: 2022A044). Mice which had been inoculated with tumor were randomly divided into four groups (n = 5): (1) PBS group; (2) PBS + US group, which intertumoral (i.t.) injection of PBS was followed by exposure to US irradiation; (3) ZnSnO3/Cu3P group, which received i.t. injection of 10 mg/kg ZnSnO3/Cu3P NCs; (4) ZnSnO3/Cu3P + US group, which i.t. injection of 10 mg/kg ZnSnO3/Cu3P NCs was followed by exposure to US irradiation. During the therapeutic period of 16 days, ZnSnO3/Cu3P NCs were administered twice on day 0 and day 2. Additionally, US irradiation (1.0 MHz, 2.5 W/cm2, 3 min, 50% duty cycle) was applied four times during the initial 4 days (Fig. 5A).

    Figure 5

    Figure 5.  In vivo cancer therapy on 4T1 tumor-bearing mice. (A) Schematic representation and timeline of the cancer treatment. (B–E) Individual tumor growth profiles and (F) average tumor growth profiles of mice following various treatments. (G) Average tumor weight and (H) representative tumor photographs from different groups on day 16. (I) H&E and Ki-67 staining of tumor sections in different groups. (J) Average body weight profiles. The data represent means ± S.D. (n = 5). ***P < 0.001.

    Based on the tumor volume growth curves recorded over the 16-day treatment period (Figs. 5B–F), these were no significant differences in tumor growth rate between the PBS + US group and the PBS group. By comparison, moderate inhibition of tumor growth was observed in the ZnSnO3/Cu3P group, mainly attributed to CDT effect from ZnSnO3/Cu3P NCs. In contrast, the ZnSnO3/Cu3P + US group exhibited the highest level of tumor suppression. At the end of the treatment, the ZnSnO3/Cu3P + US group exhibited a tumor growth inhibition rate of 70% based on tumor weights (Figs. 5G and H), which was the highest of all the groups.

    Histopathological H&E staining revealed extensive areas of tumor necrosis within the residual tissues of the tumors in the ZnSnO3/Cu3P + US group. Furthermore, Ki-67 immunohistochemical staining results showed a similar trend (Fig. 5I). Weight changes among all mice during the treatment exhibited no significant differences, indicating the compatibility and biosafety of the treatment modality (Fig. 5J). In addition, major organs such as the heart, liver, spleen, lung, and kidney exhibited no observable abnormalities or inflammation according to the results of H&E staining (Fig. S21 in Supporting information). These findings confirmed that ZnSnO3/Cu3P NCs, functioning as a sonosensitizer and chemodynamic agent, can achieve effective tumor therapy by combing SDT and CDT with a minimal systemic toxicity.

    In summary, an urchin-like piezoelectric ZnSnO3/Cu3P p-n heterojunction has been successfully synthesized as a sonosensitizer for enhancing cancer SDT. The p-n heterojunction optimizes the energy band structure of ZnSnO3/Cu3P NCs and promotes the US sensitization performance. Additionally, Cu3P exhibits chemodynamic activity and ability of GSH depletion, enabling the combination treatment of SDT and CDT. Ultimately, the piezoelectric sonosensitizer effectively generates ROS, reduces GSH, disrupts the redox balance within tumor cells, and promotes cell death, achieving satisfactory antitumor properties on 4T1 tumor-bearing BALB/c mice model. This strategy paves the way for constructing heterogeneous nanomaterials for cancer SDT therapy.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    The work was supported by the National Nature Science Foundation (No. 82072065), the Fundamental Research Funds for the Central Universities (No. E2EG6802X2), and the National Youth Talent Support Program.

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2024.109644.


    1. [1]

      S. Bai, N.L. Yang, X.W. Wang, et al., ACS Nano 14 (2020) 15119–15130. doi: 10.1021/acsnano.0c05235

    2. [2]

      Y.M. Ding, Z. Wang, Z.Y. Zhang, et al., Nano Res. 15 (2022) 7304–7312. doi: 10.1007/s12274-022-4336-0

    3. [3]

      Y. Zhao, T. Huang, X. Zhang, et al., BMEMat 1 (2023) e12006. doi: 10.1002/bmm2.12006

    4. [4]

      D.D. Liu, X.L. Dai, W. Zhang, et al., Biomaterials 292 (2023) 121917. doi: 10.1016/j.biomaterials.2022.121917

    5. [5]

      Y. Zhu, Y.J. Wang, G.R. Williams, et al., Adv. Sci. 7 (2020) 2000272. doi: 10.1002/advs.202000272

    6. [6]

      X. Xue, H. Qu, Y. Li, Exploration 2 (2022) 20210134. doi: 10.1002/EXP.20210134

    7. [7]

      Y. Zhang, X.Q. Zhang, H.C. Yang, et al., Chem. Soc. Rev. 50 (2021) 11227–11248. doi: 10.1039/d1cs00403d

    8. [8]

      L.A. Sun, X.W. Wang, F. Gong, et al., Theranostics 11 (2021) 9234. doi: 10.7150/thno.58755

    9. [9]

      Y.Q. Yang, X.W. Wang, H.S. Qian, L. Cheng, Appl. Mater. Today 25 (2021) 101215. doi: 10.1016/j.apmt.2021.101215

    10. [10]

      M.F. Wang, P.A. Ma, J. Lin, Chin. Chem. Lett. 34 (2023) 108300. doi: 10.1016/j.cclet.2023.108300

    11. [11]

      T.T. Hu, W.C. Shen, F.Q. Meng, et al., Adv. Mater. 35 (2023) 2209692. doi: 10.1002/adma.202209692

    12. [12]

      A.B. Wu, L.D. Jiang, C. Xia, et al., Adv. Sci. 10 (2023) 2303016. doi: 10.1002/advs.202303016

    13. [13]

      A.Q. Ma, H.Q. Chen, Y.H. Cui, et al., Small 15 (2019) 1804028. doi: 10.1002/smll.201804028

    14. [14]

      S. Liang, X.R. Deng, P.A. Ma, Z.Y. Cheng, J. Lin, Adv. Mater. 32 (2020) 2003214. doi: 10.1002/adma.202003214

    15. [15]

      X.Q. Qian, Y.Y. Zheng, Y. Chen, Adv. Mater. 28 (2016) 8097–8129. doi: 10.1002/adma.201602012

    16. [16]

      Z.R. Gong, Z.F. Dai, Adv. Sci. 8 (2021) 2002178. doi: 10.1002/advs.202002178

    17. [17]

      X.H. Lin, J.B. Song, X.Y. Chen, H.H. Yang, Angew. Chem. Int. Ed. 59 (2020) 14212–14233. doi: 10.1002/anie.201906823

    18. [18]

      Y.F. Zhou, S.Y. Fan, L.L. Feng, X.L. Huang, X.Y. Chen, Adv. Mater. 33 (2021) 2104223. doi: 10.1002/adma.202104223

    19. [19]

      P.R. Zhao, H.Y. Li, W.B. Bu, Angew. Chem. Int. Ed. 62 (2023) e2022104.

    20. [20]

      S.P. Ning, X.L. Dai, W.W. Tang, et al., Acta Biomater. 152 (2022) 562–574. doi: 10.1016/j.actbio.2022.08.067

    21. [21]

      Q.T. Hoang, V. Ravichandran, T.G.N. Cao, et al., Chem. Eng. J. 435 (2022) 135039. doi: 10.1016/j.cej.2022.135039

    22. [22]

      P. Zhu, Y. Chen, J.L. Shi, Adv. Mater. 32 (2020) 2001976. doi: 10.1002/adma.202001976

    23. [23]

      Y.C. Zhao, S.B. Wang, Y.M. Ding, et al., ACS Nano 16 (2022) 9304–9316. doi: 10.1021/acsnano.2c01968

    24. [24]

      X.X. Deng, P. Chen, R.R. Cui, et al., Appl. Catal. B 339 (2023) 123148. doi: 10.1016/j.apcatb.2023.123148

    25. [25]

      H.Q. Ma, W.Y. Yang, S. Gao, et al., Chem. Eng. J. 455 (2023) 140471. doi: 10.1016/j.cej.2022.140471

    26. [26]

      J. Jang, K. Kim, J. Yoon, C.B. Park, Biomaterials 255 (2020) 120165. doi: 10.1016/j.biomaterials.2020.120165

    27. [27]

      Z.L. Wang, Nano Today 5 (2010) 540–552. doi: 10.1016/j.nantod.2010.10.008

    28. [28]

      Z.L. Wang, Adv. Mater. 24 (2012) 4632–4646. doi: 10.1002/adma.201104365

    29. [29]

      Z.L. Wang, W.Z. Wu, Natl. Sci. Rev. 1 (2014) 62–90. doi: 10.1093/nsr/nwt002

    30. [30]

      S. Chen, P. Zhu, L.J. Mao, et al., Adv. Mater. 35 (2023) 2208256. doi: 10.1002/adma.202208256

    31. [31]

      Z.J. Wang, J. Liu, F.J. Wang, et al., J. Phys. Chem. C 114 (2010) 13577–13582. doi: 10.1021/jp104733e

    32. [32]

      J. Zhang, K.L. Yao, Z.L. Liu, et al., Phys. Chem. Chem. Phys. 12 (2010) 9197–9204. doi: 10.1039/b920065g

    33. [33]

      Y. Inaguma, M. Yoshida, T. Katsumata, J. Am. Chem. Soc. 130 (2008) 6704. doi: 10.1021/ja801843v

    34. [34]

      K.Y. Lee, D. Kim, J.H. Lee, et al., Adv. Funct. Mater. 24 (2014) 37–43. doi: 10.1002/adfm.201301379

    35. [35]

      S. Shawon, Z.D. Carballo, V.S. Vega, et al., Nano Energy 92 (2022) 106653. doi: 10.1016/j.nanoen.2021.106653

    36. [36]

      X.S. Cao, M.X. Li, Q.Y. Liu, et al., Small 19 (2023) 2303195. doi: 10.1002/smll.202303195

    37. [37]

      G.R. Dillip, P.C. Nagajyothi, R. Ramaraghavulu, et al., Mater. Chem. Phys. 248 (2020) 122946. doi: 10.1016/j.matchemphys.2020.122946

    38. [38]

      B.J. Ma, S. Wang, F. Liu, et al., J. Am. Chem. Soc. 141 (2019) 849–857. doi: 10.1021/jacs.8b08714

    39. [39]

      S.J. Zhong, C. Xiong, Y.C. Zhao, et al., Adv. Funct. Mater. 33 (2023) 2305625. doi: 10.1002/adfm.202305625

    40. [40]

      Y.T. Zhao, L. Tao, Chin. Chem. Lett. 35 (2024) 108571. doi: 10.1016/j.cclet.2023.108571

    41. [41]

      S. Wang, S.Q. Yang, Z.L. Cui, et al., Chem. Eng. J. 457 (2023) 141186. doi: 10.1016/j.cej.2022.141186

    42. [42]

      H. Lei, Z. Pei, C. Jiang, L. Cheng, Exploration 3 (2023) 20220001. doi: 10.1002/EXP.20220001

    43. [43]

      D. Jana, Y. Zhao, Exploration 2 (2022) 20210238. doi: 10.1002/EXP.20210238

    44. [44]

      J.A. Malla, R.M. Umesh, S. Yousf, et al., Angew. Chem., Int. Ed. 59 (2020) 7944–7952. doi: 10.1002/anie.202000961

  • Scheme 1  Schematic depiction of the synthesis process and mechanism in the antitumor process of ZnSnO3/Cu3P NCs.

    Figure 1  Synthesis and characterizations of ZnSnO3/Cu3P. (A) Schematic diagram of the fabrication process of ZnSnO3/Cu3P. (B) SEM image of ZnSnO3 NCs. (C) SEM image of ZnSnO3/Cu3P. (D) TEM image, (E) HRTEM image and (F) element mapping of ZnSnO3/Cu3P. (G) XRD, (H) XPS and high-resolution (I) Cu 2p and (J) O 1s XPS spectra of ZnSnO3/Cu3P. (K) hydrodynamic diameters and (L) zeta potential of ZnSnO3 and ZnSnO3/Cu3P. The data represent means ± standard deviation (S.D.) (n = 5).

    Figure 2  Sonodynamic and chemodynamic performance of ZnSnO3/Cu3P. (A) Schematic representation illustrating the sonodynamic and chemodynamic processes of ZnSnO3/Cu3P. (B) Degradation curves of MB and (C) corresponding kinetic curves in SDT process. Degradation curves of NBT for O2 detection in the SDT process. (E) Generation of OH in the SDT process with TA as the indicator. (F) Production of OH in the CDT process measured with TMB as the indicator. (G) Degradation curves of MB and (H) corresponding kinetic curves in the combination process of SDT and CDT. ESR spectra of (I) OH and (J) O2 trapped by DMPO under different conditions. (K) GSH oxidation by varying concentrations of ZnSnO3/Cu3P within 3 h. (L) Oxidation of GSH achieved by combining SDT and CDT.

    Figure 3  Mechanism underlying the improved sonodynamic performance of ZnSnO3/Cu3P NCs. (A) Schematic diagram illustrating the separation and migration of electron-hole pairs in both ZnSnO3 NCs and ZnSnO3/Cu3P NCs. UV–vis diffuse reflection spectra and results of Kubelka–Munk conversion of (B) ZnSnO3 NCs and (C) ZnSnO3/Cu3P NCs. XPS valence band spectra of (D) ZnSnO3 NCs and (E) ZnSnO3/Cu3P NCs. PFM phase hysteresis loop, and amplitude butterfly loop of (F) ZnSnO3 NCs and (G) ZnSnO3/Cu3P NCs. (H) COMSOL simulation of influence of surface nanoneedles (n = 0, 1, 3, and 5) on the electric potential under ultrasonic field and (I) statistics of the maximum surface potential.

    Figure 4  In vitro cell experiments demonstrating the combined SDT and CDT performance of ZnSnO3/Cu3P NCs for cancer cell killing. (A) Schematic representation depicting the mechanism of tumor cell killing by ZnSnO3/Cu3P NCs through the integration of SDT and CDT. (B) Viabilities of L929 cells after treatments with ZnSnO3/Cu3P NCs for 24 h. (C) Viability of 4T1 cells at 24 h post-treatment with different approaches. (D) Confocal images and green/red fluorescence (FL) ratios of Calcein AM/PI-stained 4T1 cells following various treatments. (E) Confocal images and corresponding quantitative analysis of intracellular ROS levels in 4T1 cells stained with DCFH-DA. (F) Confocal images and green/red FL ratios of JC-1-stained 4T1 cells. (G) Bio-TEM image of 4T1 cells after treatment with ZnSnO3/Cu3P NCs for 10 h. The data represent means ± S.D. (n = 5). ***P < 0.001.

    Figure 5  In vivo cancer therapy on 4T1 tumor-bearing mice. (A) Schematic representation and timeline of the cancer treatment. (B–E) Individual tumor growth profiles and (F) average tumor growth profiles of mice following various treatments. (G) Average tumor weight and (H) representative tumor photographs from different groups on day 16. (I) H&E and Ki-67 staining of tumor sections in different groups. (J) Average body weight profiles. The data represent means ± S.D. (n = 5). ***P < 0.001.

    Table 1.  First-order reaction kinetics of MB degradation under various treatments.

    下载: 导出CSV

    Table 2.  First-order reaction kinetics of MB degradation under various treatments.

    下载: 导出CSV
  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  41
  • HTML全文浏览量:  1
文章相关
  • 发布日期:  2024-12-15
  • 收稿日期:  2023-12-12
  • 接受日期:  2024-02-07
  • 修回日期:  2024-02-06
  • 网络出版日期:  2024-02-10
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章