Next Article in Journal
The Effect of Capping Agents on Gold Nanostar Stability, Functionalization, and Colorimetric Biosensing Capability
Next Article in Special Issue
Electrospun Biodegradable Poly(L-lactic acid) Nanofiber Membranes as Highly Porous Oil Sorbent Nanomaterials
Previous Article in Journal
Regulation of Kinetic Properties of Chemical Hydrogen Absorption and Desorption by Cubic K2MoO4 on Magnesium Hydride
Previous Article in Special Issue
Carbon Nanotube Coated Fibrous Tubes for Highly Stretchable Strain Sensors Having High Linearity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photothermal Regulated Nanozyme of CuFeS2 Nanoparticles for Efficiently Promoting Wound Healing Infected by Multidrug Resistant Bacteria

Institute of Biomedical Engineering, College of Life Sciences, Qingdao University, 308 Ningxia Road, Qingdao 266071, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2022, 12(14), 2469; https://0-doi-org.brum.beds.ac.uk/10.3390/nano12142469
Submission received: 10 June 2022 / Revised: 8 July 2022 / Accepted: 8 July 2022 / Published: 19 July 2022

Abstract

:
Peroxidase-mediated chemokinetic therapy (CDT) can effectively resist bacteria; however, factors such as the high dosage of drugs seriously limit the antibacterial effect. Herein, CuFeS2 nanoparticles (NPs) nanozyme antibacterial system with the photothermal effect and peroxidase-like catalytic activity are proposed as a combined antibacterial agent with biosafety, high-efficiency, and broad-spectrum antibacterial ability. In addition, the as-obtained CuFeS2 NPs with a low doses of Cu+ and Fe3+ can change the permeability of bacterial cell membranes and break the antioxidant balance by consuming intracellular glutathione (GSH), which results in more conducive ROS production. Meanwhile, the photothermal heating can regulate the CuFeS2 NPs close to their optimal reaction temperature (60 °C) to release more hydroxyl radical in low concentrations of H2O2 (100 µM). The proposed CuFeS2 NPs-based antibacterial system achieve more than 99% inactivation efficiency of methicillin-resistant Staphylococcus aureus (106 CFU mL−1 MRSA), hyperspectral bacteria β-Escherichia coli (106 CFU mL−1 ESBL) and Pseudomonas aeruginosa (106 CFU mL−1 PA), even at low concentration (2 μg mL−1), which is superior to those of the conventional CuO NPs at 4 mg mL−1 reported in the literature. In vivo experiments further confirm that CuFeS2 NPs can effectively treat wounds infected by MRSA and promote the wound healing. This study demonstrates that excellent antibacterial ability and good biocompatibility make CuFeS2 NPs a potential anti-infection nanozyme with broad application prospects.

1. Introduction

With the continuous development of drug-resistant strains, bacterial infectious diseases are a major threat to global health [1,2,3]. In traditional treatment strategy, antibiotics are the most effective and conventional method for the treatment of bacterial infection. However, the heavy use of antibiotics has led to the occurrence of multiple drug resistance bacteria, which causes a sharp decline in the efficacy of antibiotics [4]. Consequently, effective antimicrobial agents are severely needed, as a substitute to antibiotics. The development of new antibacterial platforms based on nanomaterials, Ag- or Cu-based, have been well-developed and regarded as efficiently antibacterial, but high-dose applications still suffered from potentially non-negligible biological toxicity, due to their intrinsic properties of heavy metals. Owing to the advantages of high catalytic activity, substrate selectivity, higher stability, and adaptability to special environments [5], nanozymes have been diffusely applied in the areas of antibacterial treatment, in order to treat the infected wounds in animals [6,7,8]. For example, peroxidase mimics can specifically facilitate the conversion of hydrogen peroxide (H2O2) to the more oxidizing hydroxyl radical (•OH), in order to achieve a better antibacterial effect, thus effectively avoiding the side effects of high concentration of H2O2 to animal healthy tissues [9,10]. Some inorganic nanomaterials with peroxidase-like activity have been reported, such as V2O5 QDs, Au NPs, and MoS2 nanomaterials, etc. [3,11]. Although these nanomaterials have shown great potential for antimicrobial use in peroxidase-mediated chemokinetic therapy (CDT), their antibacterial effects are limited, due to the insufficient production of hydroxyl radicals (•OH) and short effective action distance of hydroxyl radicals [12]. Photothermal therapy (PTT), which can use local near-infrared (NIR) light to cause high temperatures or improve the catalytic activity of nanozymes, which has been diffusely used in mice tumor treatment and antibacterial because of their advantages, i.e., less invasiveness, fewer side effects, and better controllability, compared with other antibacterial drugs. However, long time exposure to high power density NIR may induce skin destruction and bacterial heat resistance [13]. Since the antibacterial therapy alone is not efficient enough to reach satisfying efficiency, a combined therapy strategy is expected to compensate for each other, in order to overcome the shortcomings of single-mode therapy, thus greatly improving the antibacterial effect [14]. While enhancing the antibacterial efficiency of CDT, by means of multi-mode combination and ROS release, the antioxidant system of bacterial cells should not be ignored [15]. Glutathione (GSH) is an intracellular reducing substance that consumes hydroxyl radicals to inhibit the therapeutic effect of CDT. Therefore, the consumption of GSH helps to further improve the therapeutic effect of CDT [16,17]. In order to treat bacterial-infected wounds and promote wound healing, it is necessary to develop efficient and safe nano antibacterial materials. In addition, the efficient nanozyme against pathogenic bacteria with the synergistic effect of CDT, PTT, and GSH consumption capacity has rarely been reported.
Herein, we prepared and characterized the CuFeS2 nanoparticles (NPs) and evaluated the antibacterial efficacy against both Gram-positive and -negative bacteria. Under the irradiation of a NIR laser, the activity of CuFeS2 NPs peroxidase is effectively enhanced, under the condition that CuFeS2 does not produce temperatures that are too high, and more hydroxyl radicals are released, so as to kill pathogenic bacteria more effectively. In addition, under simulated physiological conditions, CuFeS2 NPs showed the ability to consume GSH. Based on the above advantages, the as-obtained CuFeS2 NPs can consume the GSH in bacteria and effectively fight pathogenic bacteria via the combination of PTT- and ROS-mediated CDT (Scheme 1). High-resolution transmission electron microscope (HRTEM) and transmission electron microscope (TEM) were used to characterize the morphology. The UV–VIS absorption spectrum and fluorescence methods were used to detect the hydroxyl radical and characterize the enzyme activity of CuFeS2 NPs.

2. Materials and Methods

2.1. Synthesis of CuFeS2 Nanoparticles

With deionized water as the solvent, 50 mL of deionized water was taken into a conical flask. The water bath was heated to 90 °C for more than 5 min, in order to effectively eliminate dissolved oxygen. Next, FeSO4·7H2O (5.6 mg) and CuCl2·2H2O (17.0 mg) were dispersed into the deionized water (1.2 mL), adjusting the pH of the mixed solution to about 12.0. Then, Na2S·9H2O (97.0 mg) was quickly added into the above mixed solution directly, and the solution turned black. Finally, the samples were converged by centrifugation, washed three times with deionized water, and freeze-dried for use.

2.2. POD-like Activity and Kinetic Assay

In order to explore the peroxidase-mimic activity of CuFeS2 NPs, TMB was selected as the chromogenic substrate, and the UV absorption changes (652 nm) of the reaction system were observed and recorded in the presence of H2O2 in the acetate buffer (0.1 M, pH 4.0). Accordingly, 0.1 mL CuFeS2 NPs suspension (40 μg mL−1), 0.025 mL H2O2 (0.5 mM), and 0.025 mL TMB (0.5 mM) were added to 0.85 mL acetate buffer. A total of four groups of reaction systems were set: TMB + H2O2, CuFeS2 + TMB, CuFeS2 + H2O2, and TMB + CuFeS2 + H2O2 at room temperature for 5 min. Then, the color pictures of each reaction system were taken, and the UV spectra of different systems were scanned by ultraviolet spectrophotometer. The temperature (22~90 °C) and pH (2.0~11.0) tolerance of the peroxidase of CuFeS2 NPs were studied with an acetate buffer (pH 4.0, 0.1 M) as the background, and the maximum absorbance of each group was defined as 100%. The peroxidase kinetics of CuFeS2 NPs was studied by absorbance monitoring at 652 nm under different conditions. In the pH 4.0 acetate buffer, the concentration of CuFeS2 NPs in the reaction system was 40 μg mL−1, and the concentration of H2O2 or TMB of one substrate was fixed, respectively, while the concentration of the other substrate was changed. The total system was 1 mL, and the reaction time was 10 min. After 10 min of reaction, the absorbance of each system was measured. The kinetic parameters of CuFeS2 NPs peroxidase were calculated according to the Michaelis equation:
1 V = K m V max × 1 [ S ] + 1 V max
where Km delegates the Mie constant, [S] delegates the substrate concentration, and Vmax stands for the maximum reaction rate. The reaction rate and substrate concentration were fitted by the Michaelis–Menten equation, Vmax and the kinetic parameters Km were obtained by Lineweave–Burk plot.

2.3. Detection of Hydroxyl Radical

The hydroxyl radical was detected by fluorescence method. In order to detect •OH in the reaction catalyzed by CuFeS2 NPs peroxidase, TA was selected as a fluorescent probe. TA itself had no fluorescence. When TA captured •OH, it formed 2-hydroxyl terephthalic acid (TAOH) with a strong fluorescence emission peak at 435 nm. In phosphate buffer solution (pH 7.4, 0.1 M), the concentration of each component in the reaction system was CuFeS2 NPs (40 μg mL−1), Ta (0.5 mM), and H2O2 (1.0 mM). Four groups of reaction systems were set: Ta, Ta + H2O2, Ta + CuFeS2, and Ta + H2O2 + CuFeS2. The samples were placed at room temperature for 12 h, and the fluorescence spectra of each sample were measured by a fluorometer (Edinburgh FS5).
Electron spin resonance spectroscopy (ESR) measurements were made at room temperature on a standard X-band Bruker E-500 EPR spectrometer (Germany). DMPO (5, 5-dimethyl-1-pyrrolin-N-oxide) was used as a trap to capture free •OH to form DMPO/•OH spin adducts. The concentrations of each component in the system were H2O2 (10 mM), CuFeS2 (40 μg mL−1), and DMPO (40 mM). Four groups of samples were set to be tested: DMPO, DMPO + H2O2, DMPO + CuFeS2, and DMPO + H2O2 + CuFeS2. The prepared sample solution was transferred into a quartz capillary tube and placed for 30 min to monitor the ESR signal. All of the above operations were performed in a NAAC buffer (0.1 M, pH 4.0). All ESR spectra were obtained with 1 G field modulation, 100 G scanning range, and 20 mW microwave power.

2.4. Bacterial Culture and In Vitro Bacteriostatic Experiment

Bacterial culture and in vitro bacteriostatic experiment. Pseudomonas aeruginosa (PA), hyperspectral β-lactamase-producing Escherichia coli (ESBL), and methicillin-resistant Staphylococcus aureus (MRSA) were inoculated in LB liquid medium (37 °C, 180 rpm) for 12 h, respectively. The obtained bacterial solution was diluted to 106 CFU mL−1, and the following treatments were performed: PBS, H2O2, CuFeS2, CuFeS2 + H2O2, PBS + NIR, H2O2 + NIR, CuFeS2 + NIR, and CuFeS2 + H2O2 + NIR, which were treated with NIR (1 W cm−2, 808 nm) for 5 min. The total system of each experimental group was 300 μL, including 106 CFU mL−1 bacterial suspension, 2 μg mL−1 CuFeS2 NPs, and 0.1 mM H2O2. After incubation at 37 °C for 30 min, 100 μL of the treated bacterial solution was taken and evenly coated on the plate. The bacteria were incubated overnight at 37 °C, observed, and photographed.

2.5. Fluorescence Detection of Living/Dead Bacteria

In order to visually identify the live/dead state of the bacteria, the bacterial samples treated with different methods were fluorescently stained. The concentration of bacterial liquid was 108 CFU mL−1, and 8 groups of experiments were set for the following treatments: PBS, H2O2, CuFeS2, CuFeS2 + H2O2, PBS + NIR, H2O2 + NIR, CuFeS2 + NIR, CuFeS2 + H2O2 + NIR, NIR refers to under the irradiation of 808 nm NIR light (808 nm, 1 W cm−2) for5 min and incubated with bacteria at 37 °C for 30 min. Then, the unbounded dye was washed using PBS after being incubated at 37 °C for 15 min using SYTO9 and PI fluorescent dye. Drop-coated glass slides were used for sample preparation. Bacteria were observed under a positive fluorescence microscope, and fluorescence staining images were taken. Bacteria in normal state showed green fluorescence, while bacteria in dead state showed red fluorescence.

2.6. Bacterial ROS Fluorescence Detection

DCFH-DA (10 μm) was added into 20 mL 0.9% NaCl solution and co-incubated with 108 CFU/mL bacteria for 30 min. The combined DCFH-DA was removed by centrifugation at 8000 RPM for 5 min. The bacterial cells were treated in different ways, as follows: PBS, H2O2, CuFeS2, CuFeS2 + H2O2, PBS + NIR, H2O2 + NIR, CuFeS2 + NIR, and CuFeS2 + H2O2 + NIR were treated with Nir (808 nm, 1 W cm−2) for 5 min and incubated at 37 °C for 4 h. Fluorescence intensity was measured with a fluorometer and observed and photographed under a fluorescence microscope.

2.7. SEM Characterization of Bacteria

Methicillin-resistant Staphylococcus aureus (MRSA), hyperspectral β-lactamase-producing Escherichia coli (ESBL), and Pseudomonas aeruginosa (PA) were treated as follows: after incubation with CuFeS2 + H2O2 at 37 °C for 4 h, each reaction system was centrifuged at 3000 RPM for 5 min, and then washed with PBS for 3 times. Then, they were fixed with 2.5% glutaraldehyde, overnight at 4 °C, followed by dehydration with ethanol solution of different concentrations, followed by treatment with 50% and 100% isoamyl acetate, the final zero-point drying, smear sample preparation.

2.8. Biocompatibility Evaluation of CuFeS2 Nanomaterials

The biological toxicity of CuFeS2 nanomaterials was evaluated by hemolysis test. The specific procedures were as follows. Fresh blood of rats was taken, centrifuged (15 min, 1500 rpm) to concentrate red blood cells, and washed with normal saline for 3 times. The red blood cell suspension was diluted to 5%, treated with normal saline, water, and different concentrations of CuFeS2 NPs (25, 50, and 100 μg mL−1), incubated for 3 h at 37 °C, and centrifuged (5 min, 11,000 rpm). The absorbance value of the supernatant at 540 nm was determined. The hemolysis rate of CuFeS2 NPs was computed with water and normal saline treatment groups as controls.
MTT assay was carried out to assess the cytotoxicity of CuFeS2 NPs using L929 fibroblasts. The procedure was as follows. Approximately 5000 cells were inoculated in each well in a 96-well plate medium containing 10% FBS and 1% penicillin-streptomycin. The CO2 cell incubator was cultured until the cells were filled, and different concentrations of CuFeS2 NPs samples were added to continue the culture. After incubation for 24 h, 20 μL 5 mg mL−1 MTT were added to each well and continued to incubate for 4 h. Then, the obtained purple formazan was dissolved by DMSO (150 µL per well) with gentle shaking for 10 min. The absorbance at 490 nm was determined, and the cell viability was calculated.
In order to evaluate the compatibility of CuFeS2 in vivo, samples with concentrations of 40 and 80 μg mL−1 were prepared and injected into rats by intraperitoneal injection. The control group was given with the same volume of PBS buffer. After one-week treatment, the treated rats were killed, and the liver, heart, spleen, kidney, lung, and other organs were removed and fixed with formaldehyde; paraffin sections were made for histological analysis. At the same time, serum samples were collected to determine physiological indexes.

2.9. The Rats Woud Model

The Sprague–Dawley (SD) rats were aged 6 weeks, with a mean weight of 180 g. All the rats were randomly distributed into eight groups (each group of three) and slashed with about 0.6 cm wound area on their backs. After that, the wound area of the rats was injected with 1 × 106 of S. aureus, which was then treated with different antibacterial groups of PBS, H2O2, CuFeS2, CuFeS2 + H2O2, PBS + NIR, H2O2 + NIR, CuFeS2 + NIR, and CuFeS2 + H2O2 + NIR. To harvest the wound tissue, the mice were sacrificed on the seventh day of the experiment, and the skin was excised, including the entire wound with adjacent normal skin. For histological examination, the wound tissues were gathered after the seventh day of the experiment and fixed in 4% paraformaldehyde solution. Then, the tissue samples were embedded in paraffin and stained with hematoxylin and eosin (H&E).

3. Results and Discussion

3.1. Preparation and Characterization of CuFeS2 NPs

The antibacterial nanomaterials were synthesized using CuCl2·2H2O, FeSO4·7H2O, and Na2S·9H2O as the precursors [18]. The morphology of the as-prepared CuFeS2 NPs was characterized by TEM (Figure 1a) and HRTEM (Figure 1b). It can be seen that the as-obtained nanoparticles have uniform size, and the particle size is calculated to be 10.38 ± 1.39 nm (Figures S1 and S10), as computed to counting the CuFeS2 particles of the TEM images. Meanwhile, the HRTEM image shows that CuFeS2 NPs have a parallel and ordered crystal lattice with 0.305 nm lattice spacing, corresponding to the (111) crystal of CuFeS2 (JCPDS 41-1404), which fits in with the literature reports [18]. In order to demonstrate the crystal characteristics of the product, XRD was performed, and the diffraction peaks were basically consistent with the standard card of the CuFeS2 crystal and in good agreement with the HRTEM data results (Figure 1c). The above results preliminarily prove that the CuFeS2 NPs was successfully synthesized. In order to study the element composition and valence of CuFeS2 NPs, XPS analysis was performed. According to the XPS spectra (Figure 1d), the samples were composed of Cu, Fe, and S elements, and their corresponding signal peaks were located at 931.81 eV, 710.70 eV, and 161.90 eV, respectively [19]. The high-resolution XPS spectra of Fe, S, and Cu for CuFeS2 NPs were analysed, respectively. Figure 1e shows the high-resolution spectrogram of Cu 2p; the peaks of 2p1/2 and Cu2p3/2 at 952.70 eV and 932.90 eV were obtained by peak separation fitting, which are attributed to Cu(I) in CuFeS2 [20]. Anyway, there was no signal peak at 942 eV, reflecting that there was almost no Cu(II) in the prepared CuFeS2, which is consistent with the previously reported results that only monovalent copper atoms exist in chalcopyrite structures [21]. Figure 1f is the high-resolution XPS spectrum of Fe 2p, the peaks at 724.83 eV and 711.15 eV are attributed to Fe3+2p1/2 and Fe3+2p3/2, while the double peaks at 713.49 eV and 719.49 eV are attributed to Fe(II) [22,23,24]. Figure 1g shows the high-resolution spectrum of S 2p, and the peak at 162.18 eV is attributed to the presence of S2-. The peak at 163.97 eV belongs to the polysulfide (Sn2-) formed by sulphur vacancy on the surface of CuFeS2. The peaks at 168.66 eV and 169.74 eV are attributed to the presence of sulfate, possibly due to the oxidation of on the CuFeS2 surfaces [25,26]. The XPS characterization further confirmed that CuFeS2 was successfully synthesized, which is consistent with the XRD and HRTEM results. The photothermal conversion ability of the CuFeS2 NPs was studied by irradiating CuFeS2 NPs aqueous solution with an 808 nm NIR laser at 1.0 W cm−2. Figure 1h shows that the CuFeS2 NPs have effective photo thermal heating, and the corresponding solution temperature increased with the concentration of CuFeS2 NPs. In the presence of CuFeS2 NPs (1000 μg mL−1) and under 60 s NIR laser irradiation, the dispersion temperature increased to 50 °C (close to their optimum enzymatic temperature), demonstrating that CuFeS2 NPs have excellent photothermal conversion ability. In Figure S2, the CuFeS2 NPs still maintain good photothermal conversion ability after five heating/cooling processes under NIR laser irradiation, which proves the excellent photothermal stability of CuFeS2 NPs. According to Figure S2A,B, the photothermal conversion efficiency [27] is further computed to be 29.8%, which is higher than that of traditional Au NPs (22.1%) [28]. The near infrared absorption of material with different concentrations also illustrated in Figure S8. These results proved that the prepared CuFeS2 NPs could be used as a potential PTT material.

3.2. CuFeS2 NPs Peroxidase Activity and Hydroxyl Radical Production

To study the peroxidase-like catalytic properties of CuFeS2 NPs, the ultraviolet spectra were used to measure the absorbance of TMB, ABTS, and OPD as color substrates in the existence of H2O2. The three substrates all produced specific color changes, and their corresponding characteristic absorption peaks were also observed in Figure 2a, which preliminarily proved that CuFeS2 NPs had intrinsic peroxidase activity. The prepared material can facilitate the reaction of H2O2 to produce hydroxyl radical, leading to the oxidation of TMB to ox-TMB, which changes the dispersion from colorless to blue. Compared with the control group, with only TMB or H2O2, the reaction solution could change from colorless to dark blue, producing a strong UV absorption peak at 652 nm upon the addition of both TMB and H2O2 (Figure 2b). In addition, the absorbance of the reaction system at 652 nm increased monotonically with the concentration of CuFeS2 NPs (Figure 2f), and the UV–VIS absorption spectra of reaction systems containing different concentrations of CuFeS2 (0, 6.25, 12.5, 25.0, 50.0, and 100 μg mL−1), indicating that the conversion of H2O2 to hydroxyl radical (•OH) was accelerated, which led to the acceleration of TMB oxidation. To further verify the peroxidase catalytic activity of CuFeS2 NPs, the hydroxyl radical generated in the catalytic process was monitored via the terephthalic acid (TA)-based fluorescent assays. As shown in Figure 2c and Figure S3, a fluorescence peak appears at 435 nm in the presence of CuFeS2 NPs, which reflected the generation of hydroxyl radicals. After the addition of H2O2, the fluorescence intensity improved much more apparently, reflecting that CuFeS2 NPs could convert H2O2 to a hydroxyl radical. The hydroxyl radical produced in the catalysis of CuFeS2 peroxidase was further determined by ESR characterization. As can be seen from Figure 2d, four typical signal peaks, with an intensity ratio of 1:2:2:1, can be obtained when CuFeS2 and H2O2 are present in the DMPO system, significantly different from CuFeS2 or H2O2 alone, which proves the production of the hydroxyl radical in the system. All the results confirmed the POD-like activity of the as-obtained CuFeS2 NPs. Additionally, we verified the influence of NIR on the peroxidase catalytic activity of CuFeS2 NPs. As shown in Figure 2e and Figure S4, the absorbance of the reaction system CuFeS2 + H2O2 + TMB was significantly increased under NIR illumination, and a higher fluorescence intensity was detected by the TA method, which indicates that the peroxidase-like activity of CuFeS2 NPs can be further enhanced upon the NIR light irradiation.

3.3. Kinetic Studies of the Enzyme-Mimic Activities of the As-Prepared CuFeS2

The impacts of pH and temperature on the peroxidase-like activity of CuFeS2 NPs were explored. As shown in Figure 3a,b, the catalytic activity was good in the pH range (2.0–11.0) and a wide temperature range (22–90 °C), and the optimum pH and temperatures for CuFeS2 were 3.0 and 60 °C, respectively. It can be seen that the CuFeS2 NPs have strong temperature tolerance. Then, the steady-state kinetics analysis was carried out using TMB and H2O2 as substrates. A typical Michaelis–Menten curve was obtained by varying the concentration of one substrate, TMB or H2O2, as the concentration of the other substrate was kept constant (Figure 3c,e). The linear equation (Figure 3d,f) was obtained by the double-reciprocal plotting method, Lineweaver–Burk, and the apparent kinetic parameters Vmax and Km were finally calculated and summarized in Table S1. The results show that the behavior of CuFeS2 NPs on the substrates TMB and H2O2 followed the typical Michaelis–Menten equation. In addition, the CuFeS2 NPs have a lower Km value and a higher Vmax than natural horseradish peroxidase (HRP) and classical peroxidase analogs (Table S1) [29,30], indicating that CuFeS2 NPs have a stronger affinity for substrates TMB and H2O2, as well as a higher catalytic efficiency. This may be due to the large specific surface area and small size of the nanoparticles, which can expose more surface-active sites.

3.4. Evaluation of Antibacterial Performance of CuFeS2 NPs

Because of the remarkable photothermal conversion ability and peroxidase activity of CuFeS2 NPs, MRSA, ESBL, and PA were selected as the study objects to evaluate the antibacterial performance of CuFeS2 NPs. In the absence of NIR irradiation, after treatment with CuFeS2 NPs +H2O2, the amount of bacterial colonies was expressively less than other groups (Figure 4). This was because of the conversion of low concentration H2O2 into highly toxic hydroxyl radical catalyzed by CuFeS2 NPs peroxidase, which increased the killing ability of bacterial cells. However, the antimicrobial ability of this strategy is limited, and many bacteria remained. In order to improve the antibacterial efficiency, NIR illumination can achieve PTT treatment and improve the catalytic efficiency of peroxidase. It can be seen from the antibacterial experiment results that, under the NIR irradiation, CuFeS2 NPs + H2O2 group significantly improved the antibacterial effect. It was calculated by the plate counting CuFeS2 NPs (2 μg mL−1) on 1 × 106 CFU mL−1 MRSA, ESBL, and PA, with an antibacterial rate of more than 99%, which is superior to those of the other control groups and the conventional CuO NPs (4 mg mL−1) reported in the literature [31]. Thus, the antimicrobial strategy, based on the combination of PPT and CDT, can effectively inhibit bacteria, while evading the use of high concentration of H2O2 and CuFeS2 NPs and effectively avoiding the toxic and side effects. In order to evaluate the PPT/CDT synergistic antibacterial performance of CuFeS2 NPs and integrity of bacterial structure, live/dead staining assays were carried out (Figure S5). It was found that there was only a small amount of red fluorescence signals of the MRSA, ESBL, and PA cells after being treated with PBS, H2O2, CuFeS2, PBS + NIR, H2O2 + NIR, or CuFeS2 + NIR, indicating that the bacterial activity was normal. In contrast, the red fluorescence signals of the three bacterial cells were significantly enhanced after CuFeS2 + H2O2 treatment, indicating that the bacterial cell wall was seriously damaged, and the dye was bound to the nucleus. When NIR light was introduced, the red fluorescence signals of CuFeS2 + H2O2 + NIR-treated bacteria were enhanced, which proved that NIR light enhanced the antibacterial effects. These results were in line with the in vitro antibacterial experiments and SEM characterization, which all certified the enhanced antibacterial ability of CuFeS2 NPs under NIR light.
In order to reveal the antibacterial mechanism, DCFH-DA was used as an ROS fluorescent probe, which emitted green fluorescence after capturing ROS, for characterizing the difference of bacterial ROS in the experiment. The three strains treated by CuFeS2 + H2O2 showed a strong fluorescence signal, indicating that H2O2 decomposed into a hydroxyl radical with stronger oxidation under the catalysis of CuFeS2 peroxidase (Figure 5a,b). The ROS fluorescence signals in the three strains all increased significantly, indicating that the production of hydroxyl radical was enhanced with the assistance of NIR light, which led to an increase the ROS level in the bacteria, thus achieving better bacteriostatic effect. These results were in accordance with the in vitro bacteriostatic experiments. To further explore the mechanism of bacterial death, the cell morphology of MRSA, ESBL-E. coli, and PA were characterized by SEM. As can be seen from Figure 5c–e, the surface of bacterial cells treated with CuFeS2 + H2O2 was rough and wrinkled, or even ruptured. The cell structure of ESBL and PA was seriously damaged, and large areas of cavities could be clearly observed, while MRSA was characterized by severe bacterial shrinkage and a rough cell surface. In Figures S9 and S11, we added the SEM image of bacteria after the CuFeS2 + H2O2 + NIR and CuFeS2 + H2O2. The above results proved that both the Cu2+ release and ROS storm caused by CuFeS2 NPs peroxidase could cause severe damage to the cell structure, or even rupture, resulting in the death of bacteria. In addition, many particles were found to be attached to the surface of bacteria, which may be caused by the aggregation of the CuFeS2 NPs on the surface of cells. Meanwhile, the interaction between the CuFeS2 NPs and bacteria is conducive to the maximum killing of bacteria. In addition, the CuFeS2 NPs and GSH were co-incubated at 37 °C in phosphate buffer (pH 7.4, 0.1 M), and DTNB was used at the characteristic peak of GSH at 412 nm. The results showed that, with the increase of co-incubation time (0, 1, 3, 6 h), the UV absorption peak at 412 nm gradually weakened (Figure S6); when incubated for 6 h, the peak value was close to that of the PBS control group, which proved that CuFeS2 NPs could consume GSH under physiological conditions, and this GSH consumption ability might be attributed to the presence of Fe(III) in CuFeS2 NPs [32]. In the bacteriostatic process, the decrease in GSH level directly leads to the decrease of the bacteria’s antioxidant level, which is conducive to the production of ROS catalyzed by peroxidase, thus producing a stronger antibacterial ability.

3.5. CuFeS2 NPs for the Treatment of Wound Infection in Rats

Through a series of characterization, it has been proven that CuFeS2 NPs have excellent antibacterial ability. In order to evaluate the therapeutic ability of CuFeS2 NPs in the process of promoting wound healing, we established the bacteria-infected wound healing model with an MRSA-infected rat back wound. A total of eight groups (three rats per group) were set up. Under the condition with or without the irradiation of NIR, the wounds of rats were treated with PBS, H2O2, CuFeS2, and CuFeS2 + H2O2, respectively. Due to the PPT/CDT synergistic antibacterial performance and GSH consumption ability under the physiological conditions, CuFeS2 could produce more ROS to inhibit bacteria and promote wound healing. In Figure 6a, the wound healing of rats at different time points (1, 3, 5, and 7 days) was compared. As expected, the CuFeS2 + H2O2 + NIR (NIR 5 min, 1.0 W cm−2, 808 nm)-treated rats had the best wound healing in all the control and experimental groups. In addition, the wound area memory of rats after 7 days of treatment was statistically analyzed, and the treatment effect was quantified based on the wound size on day 1 (Figure 6b). After CuFeS2 + H2O2 + NIR treatment, the wound healed faster, and the wound area was the smallest. This proves that the PPT/CDT synergistic antibacterial system based on CuFeS2 has great application potential in practical antibacterial applications. At the same time, on the 7th day, the wound tissues of the rats with different treatments were taken, and the wound healing was analyzed via the H&E staining of tissue sections. In Figure 6c, no significant inflammation was found in the CuFeS2 + H2O2 + NIR-treated wound tissue, indicating the good antibacterial and anti-wound infection ability of the system. In addition, compared to the control group, the new epithelial tissues in the CuFeS2 + H2O2 + NIR group were thicker, and the inflammatory cells were significantly reduced, proving that the PPT/CDT synergistic antibacterial system of CuFeS2 effectively inhibited wound infection and promoted the growth of new tissues.

3.6. Biocompatibility of CuFeS2 NPs

In vitro and in vivo antibacterial experiments have certified that the CuFeS2 NPs-based PTT/CDT antibacterial system has excellent antibacterial properties and can make infected wounds heal more easily. Therefore, it is indispensable to value the biosafety of CuFeS2 NPs. We used rat red blood cells as experimental materials to study the hemolysis characteristics of CuFeS2 NPs; no significant hemolysis was discovered, even in the presence of high concentration of CuFeS2 NPs (100 μg mL−1), and the hemolysis rate of CuFeS2 NPs was calculated to be about 1.4%, indicating that CuFeS2 NPS had good blood compatibility (Figure S7a). Figure S7b results showed that the L929 cells survival rate kept at about 100%, even while the CuFeS2 NPS concentration reached 32 μg mL−1, which proved that CuFeS2 NPs had almost no toxicity to normal cells and good biocompatibility.
In addition, blood biochemical markers were measured by intravenous injection of CuFeS2 NPs into healthy rats (Figure 7a), and the results showed that blood indexes were normal, compared to the control group treated with PBS. As shown in Figure 7b, a histological analysis of the internal organs (kidney, spleen, liver, lung, and heart) of the rats showed no significant damage or inflammation. In conclusion, the prepared CuFeS2 NPs have good biocompatibility and are a potential antibacterial material.

4. Conclusions

In this work, CuFeS2 NPs with small and uniform particle sizes were prepared at room temperature. The CuFeS2 NPs have high peroxidase catalytic activity, excellent photothermal conversion ability, and good biocompatibility. Based on these excellent characteristics, the PTT/CDT synergistic antibacterial system of CuFeS2 NPs was constructed, and the experimental results showed that the antibacterial system has high-efficiency and broad-spectrum antibacterial ability. Under NIR illumination, it can decompose a low concentration of H2O2 into hydroxyl radicals under the catalysis of the internal peroxidase-like activity, thus achieving the purpose of a multimodal antibacterial, through the combination of near-distance ROS-mediated CDT and PTT. In addition, CuFeS2 NPs can consume the GSH in bacteria and destroy the antioxidant balance of cells, which is more conducive to the generation of oxidative damage. Finally, the bacteriostatic rate of CuFeS2 NPs (2.0 μg mL−1) and H2O2 (100 μM) was more than 99%, which assured the role of copper in the antibacterial effect, while reducing the agent amount significantly. In addition, in the treatment of MRSA wound infection in rats, the combined antibacterial system effectively inhibited the wound infection and promoted wound healing. More importantly, no obvious damage was observed to the internal organs of rats, which fully proved the safety of the antibacterial system. Therefore, this work proposes a safe and efficient antibacterial system, which brings new possibilities for fighting pathogenic bacteria and bacterial resistance and will greatly promote the further application of nanozymes in the treatment of bacterial infections and other diseases.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/nano12142469/s1. Figure S1: Diameter distribution of CuFeS2 NPs. Figure S2: Temperature cycling test of CuFeS2 NPs. Figure S3: ΔPL intensity showed the catalytic activity of CuFeS2. Figure S4: Fluorescence intensity was measured by TA method. Figure S5: Fluorescence images of live/dead bacteria under different processing conditions. Bar scale: 20 μm. Figure S6: Evaluation of GSH consumption capacity of CuFeS2 NPs at different times. Figure S7: Cytotoxicity experiment of the CuFeS2. Figure S8: Visible−NIR absorption of CufeS2. Figure S9: The SEM images of PA, ESBL, MRSA treated with CuFeS2+NIR+H2O2. Figure S10: Different-sized TEM image, HRTEM image. Figure S11: The SEM images of PA, ESBL and MRSA treated with CuFeS2 +H2O2. Table S1: The Michaelis–Menten constants (Km) and maximum initial reaction rates (Vmax).

Author Contributions

Conceptualization, Z.L. (Zezhong Liu) and Z.L. (Zengxu Liu); methodology, Z.L. (Zengxu Liu); validation, Z.Z. and D.L.; formal analysis, Y.Z.; investigation, X.L.; resources, P.Z.; data curation, Y.X.; writing—original draft preparation, Z.L. (Zezhong Liu); writing—review and editing, X.D.; funding acquisition, Y.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Nos. 21874079 and 21575071), Natural Science Foundation for Outstanding Young Scientists of Shandong Province (ZR2018JL011), Key R&D Project of Shandong Province (GG201809230180), Qingdao Science and Technology Planning Project (17-6-3-15-gx), Qingdao Postdoctoral Applied Research Project (grant No. RZ2000003874), and Science and Technology Fund Planning Project of Shandong Colleges and Universities (J16LA13 and J18KA112).

Institutional Review Board Statement

In this experiment, the mouse fibroblast cell line L929 was obtained from the Institute of Biochemistry and Cell Biology of the Chinese Academy of Sciences.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, Q.; Jiang, J.; Gao, L. Catalytic antimicrobial therapy using nanozymes. WIREs Nanomed. Nanobiotechnol. 2021, 14, e1769. [Google Scholar] [CrossRef] [PubMed]
  2. Li, T.; Sun, M.; Wu, S. State-of-the-art review of electrospun gelatin-based nanofiber dressings for wound healing applications. Nanomaterials 2022, 12, 784. [Google Scholar] [CrossRef] [PubMed]
  3. Jones, K.E.; Patel, N.G.; Levy, M.A.; Storeygard, A.; Balk, D.; Gittleman, J.L.; Daszak, P. Global trends in emerging infectious diseases. Nature 2008, 451, 990–993. [Google Scholar] [CrossRef]
  4. Yin, W.; Yu, J.; Lv, F.; Yan, L.; Zheng, L.R.; Gu, Z.; Zhao, Y. Functionalized Nano-MoS2 with peroxidase catalytic and near-infrared photothermal activities for safe and synergetic wound antibacterial applications. ACS Nano 2016, 10, 11000–11011. [Google Scholar] [CrossRef] [PubMed]
  5. Cao, F.; Zhang, L.; Wang, H.; You, Y.; Wang, Y.; Gao, N.; Ren, J.; Qu, X. Defect-rich adhesive nanozymes as efficient antibiotics for enhanced bacterial inhibition. Angew. Chem. 2019, 58, 16236–16242. [Google Scholar] [CrossRef]
  6. Fang, Y.F.; Ding, X.T.; Xu, G.F.; Gong, S.D.; Niu, Y.S.; Yao, Z.Y.; Jin, Z.Y.; Wang, Y.; Xu, Y.H. Direct synthesis of bienzyme-like carbide-derived carbons via mild electrochemical oxidation of Ti3AlC2 MAX. Biomed. Env. Sci. 2022, 35, 215–224. [Google Scholar]
  7. Zhang, Y.F.; Li, D.X.; Xu, Y.H.; Niu, Y.S. Application of a cascaded nanozyme in infected wound recovery of diabetic mice. ACS Biomater. Sci. Eng. 2022, 8, 1522–1531. [Google Scholar] [CrossRef]
  8. Raj, V.; Kim, Y.; Kim, Y.G.; Lee, J.H.; Lee, J. Chitosan-gum arabic embedded alizarin nanocarriers inhibit biofilm formation of multispecies microorganisms. Carbohyd. Polym. 2022, 284, 118959. [Google Scholar] [CrossRef]
  9. Liu, B.; Wang, Y.; Chen, Y.; Guo, L.; Wei, G. Biomimetic two-dimensional nanozymes: Synthesis, hybridization, functional tailoring, and biosensor applications. J. Mater. Chem. B 2020, 8, 10065–10086. [Google Scholar] [CrossRef]
  10. Wang, L.W.; Gao, F.N.; Wang, A.Z.; Chen, X.Y.; Li, H.; Zhang, X.; Zheng, H.; Ji, R.; Li, B.; Yu, X.; et al. Defect-rich adhesive molybdenum disulfide/rGO vertical heterostructures with enhanced nanozyme activity for smart bacterial killing application. Adv. Mater. 2020, 32, 2005423. [Google Scholar] [CrossRef]
  11. Lin, Y.; Liu, X.; Liu, Z.; Xu, Y. Visible-light-driven photocatalysis-enhanced nanozyme of TiO2 Nanotubes@MoS2 nanoflowers for efficient wound healing infected with multidrug-resistant bacteria. Small 2021, 17, 210327. [Google Scholar] [CrossRef] [PubMed]
  12. Lin, J.J.; Lin, W.C.; Li, S.D.; Lin, C.Y.; Hsu, S.H. Evaluation of the antibacterial activity and biocompatibility for silver nanoparticles immobilized on nano silicate platelets. ACS Appl. Mater. Inter. 2013, 5, 433–443. [Google Scholar] [CrossRef] [PubMed]
  13. Yu, Z.; Sun, Q.; Pan, W.; Li, N.; Tang, B. A near-Infrared triggered nanophotosensitizer inducing domino effect on mitochondrial reactive oxygen species burst for cancer therapy. ACS Nano 2015, 9, 11064–11074. [Google Scholar] [CrossRef] [PubMed]
  14. Gao, Q.; Zhang, X.; Yin, W.; Ma, D.; Xie, C.; Zheng, L.; Dong, X.; Mei, L.; Yu, J.; Wang, C.; et al. Functionalized MoS2 nanovehicle with near-infrared laser-mediated nitric oxide release and photothermal activities for advanced bacteria-infected wound therapy. Small 2018, 14, 1802290. [Google Scholar] [CrossRef]
  15. Aksoy, İ.; Küçükkeçeci, H.; Sevgi, F.; Metin, Ö.; Hatay Patir, I. Photothermal antibacterial and antibiofilm activity of black phosphorus/gold nanocomposites against pathogenic bacteria. ACS Appl. Mater. Inter. 2020, 12, 26822–26831. [Google Scholar] [CrossRef]
  16. He, H.; Yang, Q.; Li, H.; Meng, S.; Xu, Z.; Chen, X.; Sun, Z.; Jiang, B.; Li, C. Hollow mesoporous MnO2-carbon nanodot-based nanoplatform for GSH depletion enhanced chemodynamic therapy, chemotherapy, and normal/cancer cell differentiation. Mikrochim. Acta 2021, 188, 141. [Google Scholar] [CrossRef]
  17. Meng, X.; Zhang, X.; Lei, Y.; Cao, D.; Wang, Z. Biodegradable copper-metformin nanoscale coordination polymers for enhanced chemo/chemodynamic synergistic therapy by reducing oxygen consumption to promote H2O2 accumulation. J. Mater. Chem. B 2021, 9, 1988–2000. [Google Scholar] [CrossRef]
  18. Chen, Q.; Luo, Y.; Du, W.; Liu, Z.; Zhang, S.; Yang, J.; Yao, H.; Liu, T.; Ma, M.; Chen, H. Clearable theranostic platform with a pH-independent chemodynamic therapy enhancement strategy for synergetic photothermal tumor therapy. ACS Appl. Mater. Inter. 2019, 11, 18133–18144. [Google Scholar] [CrossRef]
  19. Velásquez, P.; Leinen, D.; Pascual, J.; Ramos-Barrado, J.R.; Grez, P.; Gómez, H.; Schrebler, R.; del Río, R.; Córdova, R. A chemical, morphological, and electrochemical (XPS, SEM/EDX, CV, and EIS) analysis of electrochemically modified electrode surfaces of natural chalcopyrite (CuFeS2) and pyrite (FeS2) in alkaline solutions. J. Phys. Chem. B 2005, 109, 4977–4988. [Google Scholar] [CrossRef]
  20. Yang, A.; Huangfu, X.; Liu, L.; Luo, W.; Zhao, W.; Yin, J. Electrochemiluminescence immunosensor based on signal probe CuFeS2 quantum dots for ultrasensitive detection of cyclin D1. Electroanal. Chem. 2020, 871, 114269. [Google Scholar] [CrossRef]
  21. Ghahremaninezhad, A.; Dixon, D.G.; Asselin, E. Electrochemical and XPS analysis of chalcopyrite (CuFeS2) dissolution in sulfuric acid solution. Electrochim. Acta 2013, 87, 97–112. [Google Scholar] [CrossRef]
  22. Sahoo, S.; Pazhamalai, P.; Mariappan, V.K.; Veerasubramani, G.K.; Kim, N.-J.; Kim, S.-J. Hydrothermally synthesized chalcopyrite platelets as an electrode material for symmetric supercapacitors. Inorg. Chem. Front. 2020, 7, 1492–1502. [Google Scholar] [CrossRef]
  23. Xu, X.; Tang, D.; Cai, J.; Xi, B.; Zhang, Y.; Pi, L.; Mao, X. Heterogeneous activation of peroxymonocarbonate by chalcopyrite (CuFeS2) for efficient degradation of 2,4-dichlorophenol in simulated groundwater. Appl. Catal. B Environ. 2019, 251, 273–282. [Google Scholar] [CrossRef]
  24. Zhou, J.; Jiang, F.; Li, S.; Xu, Z.; Sun, W.; Ji, X.; Yang, Y. CuFeS2 as an anode material with an enhanced electrochemical performance for lithium-ion batteries fabricated from natural ore chalcopyrite. J. Solid State Electr. 2019, 23, 1991–2000. [Google Scholar] [CrossRef]
  25. Deen, K.M.; Asselin, E. On the use of a naturally-sourced CuFeS2 mineral concentrate for energy storage. Electrochim. Acta 2019, 297, 1079–1093. [Google Scholar] [CrossRef]
  26. Da Silveira Salla, J.; Dotto, G.L.; Hotza, D.; Landers, R.; da Boit Martinello, K.; Foletto, E.L. Enhanced catalytic performance of CuFeS2 chalcogenide prepared by microwave-assisted route for photo-fenton oxidation of emerging pollutant in water. J. Environ. Chem. Eng. 2020, 8, 104077. [Google Scholar] [CrossRef]
  27. Zhang, Y.; Li, D.; Tan, J.; Chang, Z.; Liu, X.; Ma, W.; Xu, Y. Near-infrared regulated nanozymatic/photothermal/photodynamic triple-therapy for combating multidrug-resistant bacterial infections via oxygen-vacancy molybdenum trioxide nanodots. Small 2021, 17, 2005739. [Google Scholar] [CrossRef]
  28. Zeng, J.; Goldfeld, D.; Xia, Y. A plasmon-assisted optofluidic (PAOF) system for measuring the photothermal conversion efficiencies of gold nanostructures and controlling an electrical switch. Angew. Chem. 2013, 52, 4169–4173. [Google Scholar] [CrossRef] [Green Version]
  29. Gao, L.; Zhuang, J.; Nie, L.; Zhang, J.; Zhang, Y.; Gu, N.; Wang, T.; Feng, J.; Yang, D.; Perrett, S.; et al. Intrinsic peroxidase-like activity of ferromagnetic nanoparticles. Nat. Nanotechnol. 2007, 2, 577–583. [Google Scholar] [CrossRef]
  30. Song, C.; Ding, W.; Zhao, W.; Liu, H.; Wang, J.; Yao, Y.; Yao, C. High peroxidase-like activity realized by facile synthesis of FeS2 nanoparticles for sensitive colorimetric detection of H2O2 and glutathione. Biosens. Bioelectron. 2020, 151, 111983. [Google Scholar] [CrossRef]
  31. Manyasree, D.; Kiran Mayi, P.; Ravikumar, R. CuO nanoparticles: Synthesis, characterization and their bactericidal efficacy. Int. J. App. Pharm. 2017, 9, 71–74. [Google Scholar]
  32. Li, J.; Hu, Z.E.; We, Y.J.; Liu, Y.H.; Wang, N.; Yu, X.Q. Multifunctional carbon quantum dots as a theranostic nanomedicine for fluorescence imaging-guided glutathione depletion to improve chemodynamic therapy. J. Colloid Interf. Sci. 2022, 606, 1219–1228. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Scheme of the treatment of bacterial wound infection in mice.
Scheme 1. Scheme of the treatment of bacterial wound infection in mice.
Nanomaterials 12 02469 sch001
Figure 1. (a) TEM image, (b) HRTEM image, (c) XRD pattern, and (d) XPS survey of the prepared CuFeS2 NPs; high-resolution XPS spectra of (e) Cu 2p, (f) Fe 2p, and (g) S 2p. (h) Heating curves of CuFeS2 NPs with different concentrations (0, 62.5, 125, 250, 500, and 1000 μg mL−1) under near-infrared light (808 nm, 1.0 W cm−2) irradiation.
Figure 1. (a) TEM image, (b) HRTEM image, (c) XRD pattern, and (d) XPS survey of the prepared CuFeS2 NPs; high-resolution XPS spectra of (e) Cu 2p, (f) Fe 2p, and (g) S 2p. (h) Heating curves of CuFeS2 NPs with different concentrations (0, 62.5, 125, 250, 500, and 1000 μg mL−1) under near-infrared light (808 nm, 1.0 W cm−2) irradiation.
Nanomaterials 12 02469 g001
Figure 2. (a) The UV–VIS absorption spectra and color photos of TMB + H2O2, CuFeS2 + TMB, CuFeS2 + H2O2, and CuFeS2 + TMB + H2O2 reaction systems in NaAc HAC buffer (pH 4.0, 25 °C) after 10 min incubation. (b) The UV–VIS absorption spectra and corresponding color photos of TMB, ABTS, and OPD color systems in NaAc HAC buffer (pH 4.0, 25 °C) after incubation for 10 min. (c) The fluorescence spectra of different reaction systems: TA, TA + H2O2, TA + CuFeS2, and TA + H2O2 + CuFeS2. (d) ESR. (e) The UV–VIS absorption spectra of CuFeS2 peroxidase catalysis with or without NIR. (f) The UV–VIS absorption spectra of reaction systems containing different concentrations of CuFeS2 (0, 6.25, 12.5, 25, 50, and 100 μg mL−1).
Figure 2. (a) The UV–VIS absorption spectra and color photos of TMB + H2O2, CuFeS2 + TMB, CuFeS2 + H2O2, and CuFeS2 + TMB + H2O2 reaction systems in NaAc HAC buffer (pH 4.0, 25 °C) after 10 min incubation. (b) The UV–VIS absorption spectra and corresponding color photos of TMB, ABTS, and OPD color systems in NaAc HAC buffer (pH 4.0, 25 °C) after incubation for 10 min. (c) The fluorescence spectra of different reaction systems: TA, TA + H2O2, TA + CuFeS2, and TA + H2O2 + CuFeS2. (d) ESR. (e) The UV–VIS absorption spectra of CuFeS2 peroxidase catalysis with or without NIR. (f) The UV–VIS absorption spectra of reaction systems containing different concentrations of CuFeS2 (0, 6.25, 12.5, 25, 50, and 100 μg mL−1).
Nanomaterials 12 02469 g002
Figure 3. (a) The pH of CuFeS2 peroxidase activity. (b) The temperature curves of CuFeS2 peroxidase activity. (ce) and (f) steady-state kinetic assay of CuFeS2.
Figure 3. (a) The pH of CuFeS2 peroxidase activity. (b) The temperature curves of CuFeS2 peroxidase activity. (ce) and (f) steady-state kinetic assay of CuFeS2.
Nanomaterials 12 02469 g003
Figure 4. Bacterial colonies of MRSA, ESBL, and PA after different treatments of PBS, H2O2, CuFeS2, CuFeS2 + H2O2, PBS + NIR, H2O2 + NIR, CuFeS2 + NIR, and CuFeS2 + H2O2 + NIR.
Figure 4. Bacterial colonies of MRSA, ESBL, and PA after different treatments of PBS, H2O2, CuFeS2, CuFeS2 + H2O2, PBS + NIR, H2O2 + NIR, CuFeS2 + NIR, and CuFeS2 + H2O2 + NIR.
Nanomaterials 12 02469 g004
Figure 5. (a) After different treatments of PBS, H2O2, CuFeS2, CuFeS2 + H2O2, and CuFeS2 + H2O2 + NIR, the ROS in MRSA, ESBL, and PA were labeled with fluorescent probes; (b) the corresponding fluorescence spectrum. Bar scale: 200 μm. The SEM images of (c) PA, (d) ESBL, and (e) MRSA treated with CuFeS2 + H2O2.
Figure 5. (a) After different treatments of PBS, H2O2, CuFeS2, CuFeS2 + H2O2, and CuFeS2 + H2O2 + NIR, the ROS in MRSA, ESBL, and PA were labeled with fluorescent probes; (b) the corresponding fluorescence spectrum. Bar scale: 200 μm. The SEM images of (c) PA, (d) ESBL, and (e) MRSA treated with CuFeS2 + H2O2.
Nanomaterials 12 02469 g005
Figure 6. (a) Wound healing of rats at different times (day 1, 3, 5, and 7) after different treatments; (b) the corresponding wound area statistics (n = 3). (c) Histological analysis of wound tissue after different treatments. Bar scale: 200 μm (* p < 0.05, *** p < 0.001, **** p < 0.0001).
Figure 6. (a) Wound healing of rats at different times (day 1, 3, 5, and 7) after different treatments; (b) the corresponding wound area statistics (n = 3). (c) Histological analysis of wound tissue after different treatments. Bar scale: 200 μm (* p < 0.05, *** p < 0.001, **** p < 0.0001).
Nanomaterials 12 02469 g006
Figure 7. (a) Blood biochemical markers of rats after different treatments (PBS, 80 μg mL−1 CuFeS2 NPs). The terms are following: alkaline phosphatase (ALP), alanine transaminase (ALT), albumin (ALB), globulin (GLOB), aspartate transaminase (AST), total protein (TP), urea nitrogen (UREA), and the ratio of albumin and globulin (A/G). Error bars represented the standard deviation (n = 3). (b) After different concentrations (40, 80 μg mL−1) of CuFeS2 NPs and PBS treatment, histological sections of internal organs of rats.
Figure 7. (a) Blood biochemical markers of rats after different treatments (PBS, 80 μg mL−1 CuFeS2 NPs). The terms are following: alkaline phosphatase (ALP), alanine transaminase (ALT), albumin (ALB), globulin (GLOB), aspartate transaminase (AST), total protein (TP), urea nitrogen (UREA), and the ratio of albumin and globulin (A/G). Error bars represented the standard deviation (n = 3). (b) After different concentrations (40, 80 μg mL−1) of CuFeS2 NPs and PBS treatment, histological sections of internal organs of rats.
Nanomaterials 12 02469 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, Z.; Liu, Z.; Zhao, Z.; Li, D.; Zhang, P.; Zhang, Y.; Liu, X.; Ding, X.; Xu, Y. Photothermal Regulated Nanozyme of CuFeS2 Nanoparticles for Efficiently Promoting Wound Healing Infected by Multidrug Resistant Bacteria. Nanomaterials 2022, 12, 2469. https://0-doi-org.brum.beds.ac.uk/10.3390/nano12142469

AMA Style

Liu Z, Liu Z, Zhao Z, Li D, Zhang P, Zhang Y, Liu X, Ding X, Xu Y. Photothermal Regulated Nanozyme of CuFeS2 Nanoparticles for Efficiently Promoting Wound Healing Infected by Multidrug Resistant Bacteria. Nanomaterials. 2022; 12(14):2469. https://0-doi-org.brum.beds.ac.uk/10.3390/nano12142469

Chicago/Turabian Style

Liu, Zezhong, Zengxu Liu, Zhen Zhao, Danxia Li, Pengfei Zhang, Yanfang Zhang, Xiangyong Liu, Xiaoteng Ding, and Yuanhong Xu. 2022. "Photothermal Regulated Nanozyme of CuFeS2 Nanoparticles for Efficiently Promoting Wound Healing Infected by Multidrug Resistant Bacteria" Nanomaterials 12, no. 14: 2469. https://0-doi-org.brum.beds.ac.uk/10.3390/nano12142469

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop