Graphical abstract
Keywords: Antibacterial activity, Nanozyme, Peroxidase-like activity, Bacteria-binding, Photothermal effect
Abstract
Antibacterial agents with enzyme-like properties and bacteria-binding ability have provided an alternative method to efficiently disinfect drug-resistance microorganism. Herein, a Fe3O4@MoS2-Ag nanozyme with defect-rich rough surface was constructed by a simple hydrothermal method and in-situ photodeposition of Ag nanoparticles. The nanozyme exhibited good antibacterial performance against E. coli (~69.4%) by the generated ROS and released Ag+, while the nanozyme could further achieve an excellent synergistic disinfection (~100%) by combining with the near-infrared photothermal property of Fe3O4@MoS2-Ag. The antibacterial mechanism study showed that the antibacterial process was determined by the collaborative work of peroxidase-like activity, photothermal effect and leakage of Ag+. The defect-rich rough surface of MoS2 layers facilitated the capture of bacteria, which enhanced the accurate and rapid attack of •OH and Ag+ to the membrane of E. coli with the assistance of local hyperthermia. This method showed broad-spectrum antibacterial performance against Gram-negative bacteria, Gram-positive bacteria, drug-resistant bacteria and fungal bacteria. Meanwhile, the magnetism of Fe3O4 was used to recycle the nanozyme. This work showed great potential of engineered nanozymes for efficient disinfection treatment.
1. Introduction
The increase of drug-resistant microorganism, caused by over usage of antibiotics, has become a serious public threat to human [1], [2]. Therefore, enormous attention has been paid to urgent developing various antibacterial agents with broad-spectrum antimicrobial property and minor side effect [3]. As with the development of nanotechnology, noble metals, especially Ag nanoparticles, have been widely applied due to their potent antibacterial properties [4], [5]. Nevertheless, the effective application has been limited by the easy aggregation of small nanoparticles, which requires suitable matrix for accurate design of the nanoparticle loading. Moreover, the excessive leakage of metal ions can inevitably result in the toxicity to the organism [6]. To avoid the above issues, considerable efforts have been dedicated in alternative strategies without detrimental effects [7]. Recently, the property of natural enzymes to produce reactive oxygen species (ROS) has been used in disinfecting microorganism [8]. Unfortunately, the natural enzyme always suffers from high cost and environment-dependence [9], [10]. The further attraction has been focused on the construction of artificial nanozymes which endow stable materials with promising ROS production features [11]. Some inorganic materials, such as ferroferric oxide [12], graphene quantum dots [13] and cerium dioxide [14], have shown strong intrinsic peroxidase-like properties, which can mimic enzymes and effectively catalyze a low concentration of H2O2 into highly active •OH [15]. These nanozymes exhibited excellent antibacterial activity in disinfection treatment using •OH to destroy the membranes of bacteria via oxidization [16]. However, the effective bacteriotoxic application of nanozymes still needs extensive improvement in many aspects [17]. For instance, the low bacteria-binding ability of most artificial enzymes as well as the short lifetime and poor diffusivity of ROS greatly hindered their interaction with bacteria and limited the disinfection performance. Hence, it is still a great challenge to develop creative strategies to address the aforementioned issues and limitations.
During the disinfection process, the interaction between bacteria and antibacterial agents is well recognized to be one of the necessary steps to determine the disinfection performance [18], [19]. In this regard, the adhesive ability of the antibacterial agents plays a significant role in effective capture and exact attack process. Inspired by the natural trapping system, the rough material surface with protuberance or pili, such as pollen [20], spike on the coronaviruses and flagella of the bacteria, exhibits better adhesion towards the substrate compared to the flat surface. Studies have also shown that the promoted adhesive ability of nanomaterials mainly originates from regulating the topological structures on the surface to increase the roughness [21], [22]. Lately, remarkable works have been done by fabricating defect-rich 2D-layers MoS2 on the surface of Cu nanowires to develop a multifunctional artificial nanozyme with good capture ability to integrate the advantages of single component [23]. As a typical material of transition metal dichalcogenides (TMDs), MoS2 possesses 2D-layer structure with huge specific surface area, facile surface modification and good biocompatibility, making it a proper candidate as well as a supportive matrix for other material attachment [24] in the fields of catalysis [25], energy storage [26] and hydrogen evolution [27]. The excellent absorption in near-infrared range and the friendly elements of Mo and S to human body enable it accessible in photothermal therapy (PTT) [28], which can also be applied in disinfection treatment as a noninvasive method [29], [30], [31]. Meanwhile, the peroxidase (POD)-like property of MoS2 was explored in the colorimetric detection of H2O2 and glucose [32], which is rarely reported in antibacterial treatment. Therefore, the smart engineered nanozymes by combination of PTT and peroxidase-like properties can provide an intriguing alternative for disinfection treatment.
The recent reported literatures have combined different bactericidal modalities (such as metal ions, ROS, hyperthermia) to achieve the synergistic effect with decreased dose of antibacterial agents and increased efficiency. However, the long distance between antibacterial agents and bacteria still limited their interaction and negatively influenced the disinfection efficiency. To solve the aforementioned issues, herein, a Fe3O4@MoS2-Ag with rough surface was constructed by a simple hydrothermal method and in-situ photodeposition of Ag nanoparticles as shown in Scheme 1 . It was found that the surface topologies of Fe3O4 nanoparticles were modified by defect-rich MoS2 layers vertically growing on the surface, which showed excellent bacteria-binding ability. Notably, the combination of Fe3O4 and MoS2 further enhanced the intrinsic peroxidase-like properties. Moreover, the hyperthermia by the photothermal property of Fe3O4@MoS2-Ag assisted to prohibit the bacterial growth. The magnetism of Fe3O4 enable the nanozyme to be easily recycled. The advantages were: i) the intrinsic POD-like property of Fe3O4@MoS2-Ag could catalyze a low concentration of H2O2 into toxic •OH, which showed a great potential in inflammation treatment; ii) the released Ag+ played an auxiliary role to attack the bacterial membranes; iii) the photothermal effect of Fe3O4@MoS2-Ag not only generated hyperthermia but also improved the POD-like property; iv) the topological structure and S-vacancy of MoS2 nanosheets endowed Fe3O4@MoS2-Ag with potent adhesion ability to bacteria by forming chemical bonds, which shortened the diffusion distance of short-life •OH radicals and further enhanced the antibacterial effect. Therefore, this work provided a promising strategy for rapid and effective disinfection treatment and showed great potentials in practical inflammation treatment.
Scheme 1.
The schematic preparation of Fe3O4@MoS2-Ag with antibacterial function.
2. Material and methods
2.1. Materials
Iron (III) chloride hexahydrate (FeCl3·6H2O), sodium acetate (NaAc) and ethylene glycol (EG) were purchased from Tianjin Fuyu Fine Chemical Reagent Corporation. Ammonium molybdate tetrahydrate ((NH4)6Mo7O24·4H2O), thiourea (CH4N2S), polyethylene glycol (PEG 1000), silver nitrate (AgNO3), glutaraldehyde (2.5%), 3, 3, 5, 5-Tetramethylbenzidine (TMB) and terephthalic acid (TPA) were obtained from Aladdin Reagent Co., Ltd (Shanghai, China). Methanol, hydrogen peroxide (H2O2) and dimethyl sulfoxide (DMSO) were provided by Xilong chemicals (China). All chemicals were used without any treatment. E. coli (ATCC25922) was bought from Shanghai Luwei Microbial SCI. &TECH. Co. Ltd.
2.2. Preparation of Fe3O4@MoS2-Ag
Firstly, Fe3O4 particles were prepared via a solvothermal method in the reported literature with a little modifications [33]. In brief, 1.35 g FeCl3·6H2O was dissolved into 40 mL EG with stirring to form a yellow transparent solution, after which 3.6 g NaAc and 1.0 g PEG were added into the solution successively. The mixed solution was stirred vigorously, followed by transference to a 100 mL sealed Teflon-autoclave. The mixture was kept at 200 °C for 8 h. The obtained black sample was thoroughly washed three times with water and ethanol, respectively, and collected by magnetic separation. The final black product was dried in a vacuum oven at 60 °C overnight.
Secondly, Fe3O4@MoS2 was prepared with a typical hydrothermal method according to the literature [34]. Briefly, 20 mg Fe3O4 nanoparticles were uniformly dispersed in 30 mL distilled water by strong sonication for 30 min. After that, 1.05 g (NH4)6Mo7O24·4H2O and 2.28 g thiourea were added into the solution with ultrasonication for another 30 min. The mixture was transferred into a 50 mL sealed Teflon-autoclave and heated at 180 °C for 10 h. The autoclave was cooled down naturally. The as-prepared sample was washed with water and ethanol in succession for 3 times and collected by magnetic separation to remove MoS2. The black product was dried in the vacuum oven at 60 °C overnight. The MoS2 was prepared with the same procedure without the addition of Fe3O4.
The in-situ loading of Ag was prepared by photo-deposition following our previous method [35]. Specifically, 100 mg Fe3O4@MoS2 powder was fully dispersed in 100 mL methanol by sonication for 20 min. The designed volume (46.7, 93.5, 280.4 and 467.3 μL corresponding to 0.5, 1, 3 and 5% compared to Fe3O4@MoS2 powder mass) of AgNO3 aqueous solution (0.1 mol/L) was slowly dropped into the suspension with the mechanical stirring for 15 min to electrostatically attract Ag+ onto MoS2 sheets, followed by bubbling N2 (99.999%) into the mixture for 30 min to remove O2. The photo-deposition process was conducted using a 300 W mercury lamp for 3 h. After washing with distilled water for 3 times and collecting by magnetism, the final product was dried in the vacuum oven at 60 °C overnight. The samples with different mass percentage of loading Ag compared to Fe3O4@MoS2 powder were termed as Fe3O4@MoS2-0.5%Ag, Fe3O4@MoS2-1%Ag, Fe3O4@MoS2-3%Ag and Fe3O4@MoS2-5%Ag, respectively.
2.3. Characterizations
X-ray diffraction (XRD) pattern from 10° to 90° was conducted to characterize the crystal structure of the as-prepared samples using Rigaku D/max-2500 diffractometer with Cu Kα radiation (λ = 0.1542 nm, 40 kV, 100 mA). Fourier-transform infrared (FT-IR) spectra were used to determine the chemical structure of the samples with Avatar 360 (Nicolet Instrument Corporation, US) in the range of 400–4000 cm−1. The morphology and microstructure of different samples were examined by scanning electron microscopy (SEM) on Carl Zeiss Microscopy GmbH 73,447 (Oberkochen, Germany) and transmission electron microscopy (TEM) on TecnaiG2 F30 (FEI Company, US). X-ray photoelectron spectroscopy (XPS) was carried out to investigate the chemical state of different elements in different as-prepared samples with ESCALAB 250 Xi (Thermo Fisher Scientific, US). The loading amounts of Ag in different samples and the Ag+ leakage were confirmed via Inductive Coupled Plasma Mass Spectrometer (ICP-MS) on an optical emission spectrometer of Optima 8300 (PerkinElmer, US). The magnetic properties were studied through Lakeshore 7404 (US). Raman spectra were recorded on XploRA PLUS (Japan) using a 532 nm laser. Ultraviolet–visible-near infrared (UV–Vis-NIR) diffuse reflectance spectra were employed to investigate the optical properties of different samples on a Hitachi U-4100 spectrophotometer (Japan). Photoluminescence (PL) spectra were performed on a fluorescence spectrometer (LS55, PerkinElmer, USA) and UV–Vis absorption spectra were obtained using Agilent Cary 60 (USA). The surface zeta potential of the as-prepared products were investigated using Zetasizer Nano ZSP (ZEN5600, Malvern Instruments Limited, U.K.). The Brunauer-Emmett-Teller (BET) analysis was conducted to investigate the surface structural change by the modification on ASAP 2460 Surface Area and Porosity Analyzer from Micromeritics Instrument Corporation (US). Before the adsorption-desorption process, the samples were degassed in vacuum at 100 °C for 6 h.
2.4. Peroxidase-like activity characterization
The peroxidase-like activity of different samples was evaluated via the oxidation properties of TMB with and without H2O2. Briefly, the as-prepared samples (final concentration of 50 μg/mL) were prepared in the acetic acid - sodium acetate buffer (pH of 4.0), followed by the addition of H2O2 (final concentration of 1 mmol/L) and TMB solution in DMSO (final concentration of 1 mmol/L) at room temperature. All catalytic activities were monitored by the absorbance change of 652 nm from the oxidation product of TMB. The temperature-dependent and pH-dependent peroxidase-like activities were explored by adjusting the temperature and pH with incubation for 5 min before measurement. The peroxidase-like activity under 808 nm laser irradiation was further investigated on the same mixture composition with UV–Vis absorbance spectra recording every 30 s during 10 min.
The final catalyst concentration for kinetics study was 50 μg/mL. Before catalysis, the aqueous solutions of different samples, H2O2 solution and TMB in DMSO were prepared at the concentrations of 500 μg/mL, 10 mmol/L and 10 mmol/L, respectively. For kinetics study, 100 μL of samples were fully mixed with 700 μL of acetic acid - sodium acetate buffer (pH of 4.0) and 100 μL of TMB. The additional 100 μL of H2O2 solution was added to above mixture to start the catalysis process. The kinetics were investigated using below Michaelis-Menton equation.
where Vo represents the initial velocity, Vmax refers to the maximum reaction velocity, and [S] is the concentration of substrate.
2.5. Detection of •OH
The production of •OH was assessed by the reaction between TPA and •OH to produce 2- hydroxyl terephthalic acid, which was monitored by PL with the excitement wavelength at 315 nm and emission wavelength at 435 nm. All mixtures were incubated in PBS (pH of 7.4) at room temperature for 6 h with TPA (0.5 mmol/L), H2O2 (1 mmol/L) and as-prepared samples (100 μg/mL). The mixtures without as-prepared samples were taken as control.
2.6. Photothermal effect of different samples
The PBS solution of different samples (100 μg/mL) was irradiated under an 808 nm NIR laser (MLL-III from Changchun, China) at a density of 1.0 W/cm2 for 15 min with the temperature recording every 30 s by a thermal probe (FLIR E6).
2.7. Bacteria-binding ability investigation
Gram-negative E. coli was taken as the target bacteria to investigate the bacteria-binding ability. The bacteria were cultured on the LB solid plate, and single colony was extracted and slowly dropped into the fluid nutrient medium. The fresh E. coli suspension can be obtained after incubation at 37 °C for 12 h in the water shaking bath. The E. coli was collected by centrifuging (5000 rpm, 3 min) and the E. coli suspension (~107 cfu/mL) was formed by diluting with PBS (pH of 7.4). All the glassware in the experiments was kept at 121 °C for 20 min to guarantee sterility.
The different as-prepared samples (final concentration of 100 μg/mL) were mixed with 1 mL bacterial suspension containing ~107 cfu/mL of E. coli. The mixture was incubated at room temperature in shaking bath for 30 min. The bacteria stuck onto the as-prepared samples were removed by magnetism. The bacteria mixed with MoS2 were collected by centrifuging at 1000 rpm for 1 min. The bacteria left in the suspension were withdrawn and diluted to proper concentrations, followed by being spread onto the sterilized LB agar plates. The viable colonies were enumerated and calculated.
2.8. Antibacterial experiments
The bacterial suspension (~107 cfu/mL) was incubated with different concentrations (50 μmol/L, 100 μmol/L, 500 μmol/L, 1 mmol/L, 5 mmol/L and 10 mmol/L) of H2O2. The antibacterial effect of different concentrations was examined by an oxford-cup method. The toxicity of different as-prepared samples was also investigated by an oxford-cup method.
The standard plate counting method was employed to investigate the antibacterial properties. The E. coli suspension (~107 cfu/mL) was mixed with different samples: Fe3O4, MoS2, Fe3O4@MoS2, Fe3O4@MoS2-Ag, H2O2, Fe3O4 + H2O2, MoS2 + H2O2, Fe3O4@MoS2 + H2O2 and Fe3O4@MoS2-Ag + H2O2 without and with 808 nm NIR laser irradiation for 15 min in the 24-well cell culture plates. The final concentrations of H2O2 and catalysts were 100 μmol/L and 100 μg/mL, respectively. After incubation for another 1 h at room temperature, 100 μL bacterial suspension was withdrawn and spread onto the LB solid plate for enumeration. The plates were cultured at 37 °C for 12 h, and the viable colonies can be observed. The E. coli with and without NIR irradiation was taken as the control. Each experiment was conducted at least three times.
To confirm the broad-spectrum antibacterial performance of this method, Staphylococcus aureus (S. aureus, ATCC6538), Bacillus subtilis (B. subtilis, ATCC6633), Methicillin-resistant Staphylococcus aureus (MRSA, ATCC43300) and Candida albicans (C. albicans, ATCC10231) were selected as the representatives of Gram-positive bacteria, drug-resistant bacteria and fungal bacteria for the antibacterial application. All experiments underwent following the aforementioned procedures. For the disinfection process, the optimal nanozyme at concentration of 100 μg/mL and H2O2 at concentration of 100 μmol/L were employed. Four groups of each kind of bacteria: I) bacteria control, II) bacteria + H2O2, III) bacteria + nanozyme and IV) bacteria + nanozyme + H2O2 were treated without and with 808 nm NIR laser irradiation for 15 min. The viable colonies on LB solid plates cultured at 37 °C after 12 h were enumerated via standard plate counting method. Each experiment was conducted at least three times.
2.9. Morphology observation of live/dead bacteria
SEM images were obtained to observe the morphology changes of bacteria through different treatments. The bacteria with different samples were collected via centrifuging or magnetism. The mixture was treated by glutaraldehyde (2.5%) at 4 °C overnight. After fixation, the mixture was washed with distilled water. And then, gradual dehydration of bacteria was done by 10%, 30%, 50%, 70%, 90% and 100% ethanol treatment in sequence for separate 15 min. The final bacteria solution was dropped onto the silica substrate with gold coating for SEM observation. The fluorescence images of live/dead bacteria and magnetic recycle of the catalyst were described in supporting information.
3. Results and discussion
3.1. Synthesis and characterizations of Fe3O4@MoS2-Ag
The nanozyme was constructed by engineering MoS2 sheets on the surface of Fe3O4 nanoparticles with a two-step hydrothermal method. SEM image (Fig. 1 a) showed the magnetic nanospheres of Fe3O4 with average diameter of ~335.6 nm in a narrow distribution. The TEM image (Fig. 1b) exhibited the typical pomegranate-like microstructure of Fe3O4 consisting of many small magnetic particles [36]. The inset image of Fig. 1b confirmed the polycrystal of Fe3O4 by electron diffraction in selected area (SAED). The Fe3O4@MoS2-Ag composites were observed with MoS2 covering Fe3O4 (diameter of ~428.9 nm) by SEM and TEM images (Fig. 1c and d). The Fe3O4 nanoparticle was employed as the support with irregular and curvy MoS2 nanosheets vertically and densely growing on the surface to increase the exposed edges of MoS2 layers and form the rough surface, followed by loading Ag nanoparticles on MoS2 sheet surface (Fig. 1e). The morphology of MoS2 was shown in Fig. S1a and b with sheet structure. The high resolution TEM image (Fig. 1f) clearly depicted the crystal lattice of 0.618 nm and 0.238 nm corresponding to the (0 0 2) plane of MoS2 and the (1 1 1) plane of Ag nanoparticle, which revealed the formation of Ag nanoparticles (size of ~17 nm). Meanwhile, the discontinuous crystal lattice and atom loss in basal surface (Fig S1c and d) demonstrated the atom-vacancy defect in MoS2 sheets, which is favorable for the enhancing adhesive ability [23]. The energy dispersive spectroscopic (EDS) mapping images (Fig. 1g, h, i and j) of Fig. 1c showed the even distribution of Fe, Mo, S and Ag elements in Fe3O4@MoS2-Ag composites, which illustrated the coverage of MoS2 on Fe3O4 and the uniform distribution of Ag inside the composites. The surface EDS elemental atomic ratio of Mo and S (Fig. S2) was determined to be 1: 1.7, revealing the S-defect of MoS2. Also the zeta potential of as-prepared Fe3O4 at pH of ~7 was −26.7 mV, while the zeta potential of MoS2 at the same pH was −22.0 mV. After the integration of MoS2 on Fe3O4, the zeta potential at pH of ~7 was −25.0 mV, which confirmed the wrapping of MoS2. The negative charge facilitated the adsorption of Ag+, which was beneficial for further in-situ photo-reduction. According to the ICP-MS results, the final loading amounts of Ag were 3.875, 8.500, 25.251 and 44.125 μg of Ag in 1 mg Fe3O4@MoS2-0.5%Ag, Fe3O4@MoS2-1%Ag, Fe3O4@MoS2-3%Ag and Fe3O4@MoS2-5%Ag, respectively. The mass ratio of Fe and Mo in prepared Fe3O4@MoS2-Ag was determined to be ~1.71:1 by ICP-MS. After the deposition of Ag, the zeta potential of different composites at the same pH slightly increased from −22.3 mV to −21.4, −21.1, −20.9 mV with increasing loading amount of Ag (Fig. S3), respectively.
Fig. 1.
(a) SEM image of Fe3O4 nanoparticles with inset of size distribution. (b) High-resolution TEM image of single Fe3O4 with the inset image of its SAED pattern. (c) SEM image with inset of size distribution and (d) TEM image of Fe3O4@MoS2-Ag composite nanoparticles. (e) TEM image of single Fe3O4@MoS2-Ag nanoparticle with the inset image of Ag nanoparticles on the surface of MoS2 sheets. (f) High-resolution TEM image of MoS2 sheets and Ag nanoparticle. (g)-(j) The corresponding EDS mapping of Fe, Mo, S and Ag element profiles of (c).
The XRD patterns of different samples were presented in Fig. 2 a. The diffraction peaks at 18.2°, 30.1°, 35.4°, 37°, 43.1°, 53.4°, 56.9° and 62.5° were indexed to (1 1 1), (2 2 0), (3 1 1), (2 2 2), (4 0 0), (4 2 2), (5 1 1) and (4 4 0) planes of face-centered Fe3O4 nanoparticles [37]. The typical peaks appearing at 14.8°, 32.5° and 57.5° responded to (0 0 2), (1 0 0) and (1 1 0) planes of MoS2 (JCPDS NO. 75-1539) [38], which was in agreement with the TEM results. The XRD patterns of Fe3O4@MoS2-Ag exhibited the characteristic peak of Ag nanoparticles at 38° as well as the typical peaks of MoS2 and Fe3O4, which confirmed the construction of Fe3O4@MoS2-Ag composites. The weak peak of Ag was due to the relatively low loading amount of Ag nanoparticles. FT-IR spectra were used to testify the inorganic groups (Fig. S4). The spectrum of MoS2 confirmed the removal of thiourea. The peak around 586 cm−1 belonged to Fe-O vibration. Meanwhile, the incorporation of Ag did not change the structures of the composites. Raman spectroscopy was also employed to explore the surface modification with a laser of 532 nm (Fig. 2b). The double peaks at ~378.8 and ~404.8 cm−1 were attributed to the E1 2g and A1g vibrational modes of MoS2 [39], which showed slightly shift due to the coupling with Fe3O4 and Ag compared with those of pure MoS2. The chemical states of different elements in the composites were further investigated by XPS measurement. As shown in Fig. 2c, the broad survey spectrum of Fe3O4@MoS2-Ag composite contained the elements of Fe, O, Mo, S and Ag, demonstrating the components in the composite. Fig. 2d depicted the high-resolution spectra of Fe2p, where the characteristic binding energy peaks located at 724 and 711 eV assigned to Fe2p1/2 and Fe2p3/2, confirming the existence Fe3O4. In the Fe3O4@MoS2-Ag composite, the shifted peaks may be explained by the interaction of Fe with S by coupling with MoS2. Fig. 2e showed the specific information of Mo3d with two main peaks and two shoulder peaks. The two main distinct peaks can be fitted into two sets of 232.2 and 229.2 eV as well as 231.6 and 228.4 eV, which were attributed to 2H-MoS2 and 1T-MoS2, respectively. The peaks at 232 and 229 eV belonged to Mo3d3/2 and Mo3d5/2 of Mo (IV). The peak located at 235.3 eV was indexed to MoO3 due to the insufficient synthesis during hydrothermal process [40]. The peak centered at 225.6 eV was the result of S2s, accounting for the formation of 2H-MoS2 [41]. The characteristic peaks at 162.5 and 161.3 eV in Fig. 2f originated from S2p1/2 and S2p3/2 of S (II), respectively. For the Fe3O4@MoS2-Ag composite, the binding energy of S2p tinily shifted, which confirmed the interaction between S and Fe during the coupling of Fe3O4 and MoS2. The high-resolution spectrum of Ag in Fig. 2g exhibited two distinct peaks at 374.3 and 368.3 eV with a splitting energy of 6 eV corresponding to Ag3d3/2 and Ag3d5/2 of Ag (0) [42], which indicated the formation of metal Ag on the surface of MoS2. According to the specific surface XPS data of the Fe3O4@MoS2-Ag composite, the atom ratio of Mo and S was calculated to be 1:1.65, which verified the TEM and EDS results.
Fig. 2.
(a) XRD patterns of Fe3O4, MoS2 and Fe3O4@MoS2-Ag. (b) Raman spectra of MoS2 and Fe3O4@MoS2-Ag. (c) XPS survey spectrum of Fe3O4@MoS2-Ag. (d), (e), (f) and (g) The Fe2p, Mo3d, S2p and Ag3d spectra of the Fe3O4@MoS2-Ag. (h) Magnetic properties of Fe3O4, Fe3O4@MoS2 and Fe3O4@MoS2-Ag with inset digital photos of Fe3O4@MoS2-Ag change under external magnetic field. (i) Temperature change of Fe3O4, MoS2, Fe3O4@MoS2 and Fe3O4@MoS2-Ag (100 μg/mL) irradiated by 808 nm NIR laser (1.0 W/cm2).
The BET analyses of different samples were carried out to investigate the specific surface area. The N2 adsorption-desorption isotherms of MoS2, Fe3O4, Fe3O4@MoS2 and Fe3O4@MoS2-1%Ag and corresponding pore width distributions were shown in Fig. S5. The curves of all samples exhibited obvious hysteresis at the relative high pressure (P/P0) between 0.6 and 1.0, which was in accordance with type IV isotherms. The BET surface area was calculated to be 7.746, 13.464, 16.607 and 17.446 m2/g for MoS2, Fe3O4, Fe3O4@MoS2 and Fe3O4@MoS2-1%Ag, respectively, showing the superiority of MoS2 sheets vertically growing on Fe3O4. Meanwhile, the Ag nanoparticles on MoS2 sheets slightly enhanced the specific surface area, which was favorable for the catalysis. Fig. S5b presented the wide pore width distribution of different samples centering between 2 and 25 nm, indicating the mesoporous nanostructure. The magnetic properties of different samples were investigated at 300 K (Fig. 2h). All samples exhibited good superparamagnetic properties with little hysteresis and coercivity, indicating the excellent magnetic response under magnetic fields. The magnetization saturation values decreased from 76.9 emu/g for Fe3O4 to 39.8 emu/g for Fe3O4@MoS2 and 33.3 emu/g for Fe3O4@MoS2-Ag, which could be explained by the less mass ratio of Fe3O4 in unit composites [43]. The inset photos showed the rapid response (20 s) of Fe3O4@MoS2-Ag to external magnetic field and the suspension quickly became even after removing the magnetism. The optical properties of different samples were investigated by UV–Vis-NIR spectrophotometer. Fig. S6a presented the UV–Vis-NIR diffuse reflectance spectra of different sample powders, which showed the good absorbance of MoS2 from 300 to 1000 nm. An obvious absorbance region around 800 nm of Fe3O4 nanoparticles appeared mainly attributed to the tiny particles of a single nanoparticle. The surface modification of MoS2 on Fe3O4 nanoparticles promoted its absorbance around 800 nm. The increased absorbance at ~420 nm is due to the surface plasmon resonance (SPR) effect of Ag nanoparticles on the surface of nanozymes, meanwhile the enhanced absorbance around 800 nm is possibly the result of the quick electron transfer by Ag nanoparticles. The UV–Vis-NIR absorbance spectra of different samples in aqueous solution displayed the similar trend as the UV–Vis-NIR diffuse reflectance spectra of corresponding powders (Fig. S6b), demonstrating excellent absorbance at ~800 nm of the as-prepared nanozymes in aqueous solution. The photothermal effect of different samples (100 μg/mL) at PBS (pH of 7.4) was conducted under an 808 nm NIR irradiation (1.0 W/cm2). The temperatures were recorded every 30 s as shown in Fig. 2i. As the control, the PBS almost stayed a slight increase with ΔT of 2.7 °C in 15 min. The temperatures of all samples increased with the irradiation time, among which the temperature of MoS2 increased to 43.3 °C and the temperature of Fe3O4 suspension increased to 45.4 °C after 15 min irradiation. For the coupling of these two components, the temperature reached a significantly higher temperature of 48.3 °C, while the incorporation of Ag improved the photothermal property to 50.7 °C, which displayed excellent photothermal effect of the composites. The photothermal effect of composites (100 μg/mL) with different Ag loading was investigated. Fig. S7a showed the temperature change of different samples. The temperature of Fe3O4@MoS2-0.5%Ag irradiated after 15 min rose to 49.1 °C, while the final temperatures of Fe3O4@MoS2-1%Ag, Fe3O4@MoS2-3%Ag and Fe3O4@MoS2-5%Ag were around 50.5 °C, exhibiting little difference. The higher temperatures were obtained with increasing concentrations of Fe3O4@MoS2-Ag (Fig. S7b).
3.2. Investigation of peroxidase-like property
The peroxidase-mimicking properties of different as-prepared samples were investigated by fluorescent experiments and UV–Vis response in different conditions. TPA was employed as the fluorescent probe to check the existence of •OH, where higher intensity meant more •OH. It can be found that TPA alone and H2O2 alone showed negligible fluorescent intensity at 435 nm after incubation for 6 h, while single TPA or H2O2 with catalyst did not show much difference in Fig. 3 a. The intensity of TPA and H2O2 in the presence of catalyst exhibited greatly improvement, which suggested that the catalyst could convert H2O2 into •OH. The intensity gradually increased by chemical coupling Fe3O4 with MoS2, and the addition of Ag nanoparticles also considerably increased the fluorescent intensity, which was due to the assisting adsorption of TPA by interaction between Ag nanoparticles and –COO– group in TPA. However, with more loading amount of Ag, the intensity of nanozymes increased. Fe3O4@MoS2-1%Ag showed the strongest intensity. The peroxidase-like ability was subsequently verified with the catalytic oxidation of TMB monitored by UV–Vis spectrometer (Fig. 3b). The control experiment only contained H2O2 and TMB. After incubation for 5 min, a typical absorbance at 652 nm was observed with different as-prepared samples, whereas the control showed no absorbance, indicating no peroxidase reaction happened. The higher absorbance intensity referred to the strong peroxidase ability. The intensity of Fe3O4@MoS2 was much higher than that of Fe3O4 alone and MoS2 alone. The incorporation of Ag into Fe3O4@MoS2 promoted the intensity by the enhanced •OH production property. Fe3O4@MoS2-1%Ag showed the highest intensity. The peroxidase-mimicking ability can also be observed by the color change seen from the inset of Fig. 3b. The higher peroxidase-like property was, the darker blue the solution became, which confirmed the excellent peroxidase-like property of Fe3O4@MoS2-1%Ag. Hence, the Fe3O4@MoS2-1%Ag was chosen as the optimal sample for the further kinetic study and antibacterial applications.
Fig. 3.
(a) PL spectra of different samples incubated with TPA for 6 h with excitement wavelength of 315 nm and emission wavelength of 435 nm. (b) UV–Vis absorption spectra of different samples incubated with TMB for 5 min. Inset digital photos: color change of corresponding samples. (c)-(f) Steady-state kinetic assay and catalytic ability of Fe3O4@MoS2-1%Ag towards [H2O2] and [TMB].
The kinetic study of the catalytic process in our work was analyzed according to Michaelis-Menton equation. The Michaelis Menten constant (Km) was calculated through the Lineweaver Burk plot:
Vo was obtained through the absorption intensity at 652 nm and the molar absorption coefficient of oxidation product of TMB (39000/(cm·mol/L)) [44]. Based on the above parameters, the Michaelis-Menton curves of Fe3O4@MoS2-1%Ag were measured by separately manipulating the concentrations of TMB (Fig. 3c and d) and H2O2 (Fig. 3e and f). As a contrast, the Michaelis-Menton curves of Fe3O4@MoS2 were shown in Fig. S8. Km and Vmax of different samples were obtained through the formulation (Table S1). Normally, the lower Km represents the stronger attraction between the catalyst and the substrate, while the higher Vmax refers to the better catalytic ability. The whole catalytic process should be assessed by combining Km and Vmax. Compared to the data of HRP in previous reports [45], both Fe3O4@MoS2 and Fe3O4@MoS2-1%Ag showed smaller values of Km, indicating better affinity to TMB and H2O2 to promote catalytic performance, which makes it possible to act as peroxidase. Moreover, the value of Vmax of Fe3O4@MoS2-1%Ag was a little higher than that of Fe3O4@MoS2, which may be the result of peroxidase-like and electron transfer accelerating property of Ag [46].
Similar to the natural enzyme, the catalytic response of Fe3O4@MoS2-1%Ag to pH and temperature was also considered. As shown in Fig. S9, Fe3O4@MoS2-1%Ag composites exhibited best catalytic activity at pH of 4. With the increasing pH, the relative activity of Fe3O4@MoS2-1%Ag decreased, but it still showed catalytic capability at pH of 7.4, which was favorable for the in vitro disinfection treatment. Unlike natural enzyme, the artificial nanozyme showed increasing catalytic activity with the higher temperature in the range from 25 to 55 °C. Besides, four groups of (i) TMB only (1 mmol/L), (ii) TMB and H2O2 (1 mmol/L), (iii) TMB and Fe3O4@MoS2-1%Ag (50 μg/mL), (iv) TMB, Fe3O4@MoS2-1%Ag and H2O2 were irradiated by 808 nm NIR laser to examine the influence of hyperthermia on POD-like activity. Fig. S10a clearly showed the UV–Vis absorbance spectra with strong absorbance at 652 nm of TMB, Fe3O4@MoS2-1%Ag and H2O2 mixture without NIR irradiation, which indicated the increased catalytic reaction with time at room temperature. Fig. S10b exhibited the UV–Vis absorbance spectra of the same mixture composition with that in Fig. S10a under 808 nm laser irradiation. The intensity at 652 nm of the mixtures without and with irradiation was recorded in Fig. S10c. The catalytic reaction in mixture under irradiation showed growingly difference with that of mixture without irradiation for the same reaction time, which was due to the increased temperature with time caused by Fe3O4@MoS2-1%Ag under 808 nm irradiation. Hence, POD-like activity of Fe3O4@MoS2-1%Ag can be evidently promoted by hyperthermia from Fe3O4@MoS2-1%Ag under 808 nm irradiation.
3.3. The adhesive properties against E. coli
The adhesive properties of different samples were explored against the fresh E. coli suspension with concentration of 107 cfu/mL. The E. coli suspension was incubated with different samples: (I) control; (II) Fe3O4; (III) MoS2; (IV) Fe3O4@MoS2; (V) Fe3O4@MoS2-1%Ag for 30 min. And then MoS2 with E. coli was collected by centrifuging at 1500 rpm, while all other samples with attached E. coli were removed by external magnetic field. The E. coli suspensions left were spread on the agar plate and cultured for 24 h. The viable colonies were shown in Fig. S11a with decreasing numbers from group (I) to group (IV). Specifically, Fig. 4 c showed that Fe3O4 only took away 4.2% E. coli, while MoS2 adhered 5.7% E. coli. Regarding Fe3O4@MoS2 and Fe3O4@MoS2-1%Ag, the removal rate against E. coli of Fe3O4@MoS2-1%Ag (~22.5%) was almost same as that of Fe3O4@MoS2 (~21.9%). The promising adhesive ability was promoted by surface engineering of sharp MoS2 sheet, which was beneficial for the disinfection treatment. The morphology of removed as-prepared samples was observed by SEM images in Fig. S11b. It can be clearly seen that the viable E. coli was integral rhabditiform with no deformation and surface damage. Several E. coli without deformation was observed inside the Fe3O4 nanoparticles, which were extracted by the crowded particles. There were a few bacteria inside the MoS2 sheets with the nanosheets around bacteria with slight deformation, showing the little adhesive ability. Unsurprisingly, more E. coli was found attaching to Fe3O4@MoS2 and Fe3O4@MoS2-1%Ag, which confirmed the enhanced interaction between bacteria and the catalyst. The deformation with little destruction on the bacterial membrane of the E. coli attached to the surface of Fe3O4@MoS2 and Fe3O4@MoS2-1%Ag confirmed the notable adhesive ability of the catalysts due to the rough surface formed by sharp MoS2 sheets engineered on Fe3O4 nanoparticles, which would be an essential part for the highly efficient attacking during disinfection process. Besides, the interaction between different samples and bacteria was further investigated. The cell wall of E. coli was mainly composed of peptidoglycan with amino acid residues with negative charge (Fig. S12), which acted as the functional sites to tightly combine with S-vacancy of S-defect MoS2 to form chemical bonds [47], [48]. It is also confirmed by phase transfer of samples from water to oil phase with oleylamine (–NH2) and oleic acid (–COOH) as the hydrophobic ligands. Fig. S13 showed great affinity of Fe3O4@MoS2 to –NH2 and Fe3O4@MoS2-1%Ag to both –NH2 and –COOH. Hence, the pilus on bacteria and the sharp MoS2 sheets first pierced the energy barrier between the two negatively charged objects. The interaction between Fe3O4@MoS2-Ag and –NH2 and –COOH facilitated the further affinity. The chemical bonds provided the potent adhesion of Fe3O4@MoS2-Ag to bacteria.
Fig. 4.
(a) Digital photos of viable colonies on standard agar plates cultured with different samples without and with 808 nm NIR laser irradiation. (b) SEM images of corresponding samples and E. coli. (c) Bacteria percentage left in the suspension after removing the catalysts. (d) Survival of bacteria after different samples treatment determined by standard plate counting method.
3.4. In vitro antibacterial experiments
Considering its excellent peroxidase-like property and promising adhesive ability, the antibacterial performance was further evaluated against E. coli by a standard plate counting method. Because H2O2 at high concentration could result in damage to organism, the toxicity of H2O2 in different concentrations was assessed by oxford-cup method prior to antibacterial experiments. Fig. S14a demonstrated that H2O2 at a low concentration of 100 μmol/L barely prohibited the growth of E. coli, while H2O2 at concentration of 500 μmol/L and 1 mmol/L could inhibit the growth inside the oxford cup. H2O2 at concentration of 5 and 10 mmol/L displayed extended inhibition zones. Thus, H2O2 at a moderate concentration of 100 μmol/L was the optimal concentration for future use. The disinfection performance was tested in different groups: (I) PBS control, (II) H2O2, (III) Fe3O4, (IV) MoS2, (V) Fe3O4@MoS2, (VI) Fe3O4@MoS2-1%Ag, (VII) Fe3O4 + H2O2, (VIII) MoS2 + H2O2, (IX) Fe3O4@MoS2 + H2O2, (X) Fe3O4@MoS2-1%Ag + H2O2 without and with 808 nm NIR irradiation at a density of 1.0 W/cm2. As presented in Fig. 4a and b, negligible inhibition against E. coli could be observed when cultured with Fe3O4, MoS2, Fe3O4@MoS2 and Fe3O4@MoS2-1%Ag (100 μg/mL) without NIR irradiation, implying their insignificant damage on E. coli, which could be confirmed by the oxford-cup method with almost no inhibition zone around the samples (Fig. S12b). The SEM images showed no obvious destruction on the surface structure of the bacteria, indicating that the Fe3O4@MoS2-1%Ag could not cause fatal damage on bacteria. By combining the catalysts with H2O2, distinctly reduced viable colonies were shown with decreasing survival percentage of 92.1%, 90.3%, 80.0% and 69.4% for groups of Fe3O4, MoS2, Fe3O4@MoS2 and Fe3O4@MoS2-1%Ag, respectively, illustrating the significant role of produced •OH. SEM images confirmed the partial deformation and membrane damage of the E. coli. Subsequently, photothermal effect was used to assist the antibacterial treatment. The result of PBS control group as well as the H2O2 alone group demonstrated little influence of NIR irradiation against viable E. coli. The visual colony numbers of different samples under irradiation without H2O2 evidently decreased, which illustrated that the hyperthermia could effectively cause the bacteria destruction (Fig. 4b). Obviously, the survival percentage of Fe3O4@MoS2-1%Ag (46.6%) was much lower than that of Fe3O4@MoS2 (60.3%), indicating that the higher temperature showed better disinfection performance. After combining with the moderate concentration of H2O2 under NIR irradiation, the survival percentages of different groups continued reducing, implying the enhanced antibacterial ability due to the increasing property to convert H2O2 into highly toxic •OH and the heat. Fe3O4@MoS2-1%Ag showed prominent antibacterial performance that nearly disinfected all bacteria, which was better than Fe3O4@MoS2 with survival percentage of 13.5%. The survival rates of all samples at different conditions were exhibited in Fig. 4d. SEM images verified the morphology of bacteria together with catalysts, where the bacteria were apparently adhesive to Fe3O4@MoS2-1%Ag with distorted and destroyed membrane. It can be deduced that the adhesive ability of Fe3O4@MoS2-1%Ag paved the way for interaction between bacteria and catalyst, and at the same time the collaborative work of •OH and hyperthermia dominated the accurate disinfection process.
A concentration-dependent disinfection study was explored with ten groups: E. coli, E. coli + H2O2, E. coli + Fe3O4@MoS2-0.5%Ag, E. coli + H2O2 + Fe3O4@ MoS2-0.5%Ag, E. coli + Fe3O4@MoS2-1%Ag, E. coli + H2O2 + Fe3O4@MoS2-1%Ag, E. coli + Fe3O4@MoS2-3%Ag, E. coli + H2O2 + Fe3O4@MoS2-3%Ag, E. coli + Fe3O4@ MoS2-5%Ag, E. coli + H2O2 + Fe3O4@MoS2-5%Ag without and with 808 nm NIR irradiation for 15 min. The viable colonies of E. coli in each group were enumerated and disinfection results were calculated by standard agar plate method. The viable colonies and survival rates of different samples without NIR irradiation (Figs. S15 and S16) manifested that all composites alone cannot affect the growth of bacteria. However, the numbers of E. coli incubated with different samples and irradiated by NIR exhibited clear decrease, indicating the roles of hyperthermia against bacteria. The E. coli incubated with different samples and H2O2 without NIR irradiation also deceased in agreement with the POD-like properties, showing the •OH attack on bacteria. As with the synergistic effect, ~6% of E. coli with Fe3O4@MoS2-0.5%Ag and H2O2 under NIR irradiation were still alive, whereas the E. coli with other samples were all dead, implying the prominent disinfection performance of all samples. Moreover, ICP-MS was employed to detect the released Ag+ during the 15 min treatment in four groups of only nanozyme (100 μg/mL), nanozyme (100 μg/mL) and 808 nm NIR irradiation, nanozyme (100 μg/mL) and H2O2 (100 μmol/L), and nanozyme (100 μg/mL), H2O2 (100 μmol/L) and 808 nm NIR irradiation. Fig. S17a displayed the increased Ag+ leakage with the increased Ag load in samples. The hyperthermia caused by NIR irradiation and oxidizing of H2O2 could accelerate the Ag+ leakage. The released Ag+ could play an auxiliary role in disinfection process and enhanced the antibacterial performance. The highest amount of released Ag+ in Fe3O4@MoS2-1%Ag suspension was ~62.6 μg/L. Fig. S17b showed the releasing trend of Ag+ in Fe3O4@MoS2-1%Ag incubated with H2O2 under NIR irradiation, indicating the gradual release of Ag+ with time.
The rupture of E. coli incubated with different catalysts under 808 nm NIR irradiation was examined by blue fluorescence (DAPI) staining all E. coli and red fluorescence (PI) only staining membrane-damaged E. coli. As shown in Fig. 5 a, the fluorescence images demonstrated that the NIR irradiation of PBS and H2O2 cannot destroy the membrane integrity and cause the red fluorescence stains. When the E. coli suspension with different catalysts was irradiated by 808 nm for 15 min, the fluorescence images showed increasing red fluorescence stains with the higher temperature by photothermal effect, which illustrated that the local hyperthermia can cause the membrane disorganization of E. coli. The number of red fluorescence stains rose in the presence of H2O2, implying the attack of •OH towards the bacterial membrane. Compared with Fe3O4@MoS2, Fe3O4@MoS2-1%Ag showed the highest numbers of dead bacteria, which was in agreement with the results of viable colonies on agar plates. The micro structure of single E. coli was further observed by SEM images. It can be seen that the fresh live E. coli was smooth and rod-like with intact membrane, while the bacteria treated by H2O2 showed no obvious difference on the structure. All the dead bacteria exhibited the clear deformation and disruption of membrane, finally resulting in the leakage of cytoplasm. Based on above description, the possible synergetic antibacterial mechanism was that the adhesion of Fe3O4@MoS2-1%Ag to bacteria facilitated the precise and rapid attacking of •OH caused by peroxidase-like property and Ag+ leaking from the catalyst surface assisted by the local hyperthermia under 808 nm NIR irradiation, thus leading to the deformation and disruption of bacterial membrane with leakage of cytoplasm. Fe3O4@MoS2-1%Ag was recycled by external magnetic field and reuse to disinfect E. coli, which showed excellent inactivation rate (~95%) and stability after 5 times reuse (Fig. S18).
Fig. 5.
(a) The fluorescence images of live/dead bacteria after 808 nm NIR treatment in different groups with DAPI staining all bacteria and PI staining membrane-damaged bacteria. (b) The SEM images showing the microstructure of single E. coli in corresponding groups.
The broad-spectrum antibacterial performance of Fe3O4@MoS2-1%Ag was confirmed by their disinfection effect on S. aureus, B. subtilis, MRSA and C. albicans. Fig. S19a and b showed disinfection results of S. aureus. The viable colonies on standard agar plates exhibited no difference in S. aureus, S. aureus + NIR, S. aureus + H2O2, S. aureus + H2O2 + NIR, indicating the negligible effect of only NIR and H2O2 without and with NIR against S. aureus. However, the numbers of S. aureus colonies cultured with Fe3O4@MoS2-1%Ag greatly differed from those cultured with Fe3O4@MoS2-1%Ag and NIR, illustrating the hyperthermia killing ~57% of S. aureus. The viable colonies treated by Fe3O4@MoS2-1%Ag and H2O2 decreased by ~27% as the presence of ·OH. Whereas, almost 99% of S. aureus was disinfected by the treatment of Fe3O4@MoS2-1%Ag, H2O2 and NIR due to the synergistic effect of ·OH, hyperthermia and Ag+. Fig. S19c and d displayed the morphology change of S. aureus of fresh cells and treated cells. The excellent disinfection performance was universal towards B. subtilis (Fig. S20a and b), MRSA (Fig. S21a and b) and C. albicans (Fig. S22a and b). B. subtilis was more sensitive to hyperthermia than ·OH, while MRSA showed the contrary result, which was possibly due to the component difference of membranes. The viable colonies of both bacteria dramatically decreased under treatment of •OH, hyperthermia and Ag+. For C. albicans, •OH or hyperthermia alone killed small portion of C. albicans, while the synergy of them damaged 80% of C. albicans. The SEM images of all morphology comparisons between fresh cells and dead cells distinctly elucidated the damage on cell membranes of dead cells (Fig. S20c and d, Fig. S21c and d, Fig. S22c and d). The abovementioned results confirmed the broad-spectrum antibacterial property of Fe3O4@MoS2-1%Ag against Gram-negative bacteria, Gram-positive bacteria, drug-resistant bacteria and fungal bacteria.
4. Conclusion
In summary, a facile artificial nanozyme of Fe3O4@MoS2-1%Ag was constructed by a two-step hydrothermal method with in-situ growth of Ag nanoparticles. The composite showed attractive rough surface by engineering sharp MoS2 sheet around Fe3O4 surface, which exhibited potent adhesive ability towards bacteria. The peroxidase-mimicking properties were further explored by TMB and TPA, which confirmed the effective converting a low concentration of H2O2 into toxic •OH. When exposed to 808 nm NIR irradiation, the local hyperthermia and enhanced peroxidase-mimicking property can prominently improve the disinfection activity. The antibacterial experiments against E. coli revealed that Fe3O4@MoS2-1%Ag can efficiently inactivate bacteria by capturing E. coli and releasing toxic •OH and Ag+ assisted by local hyperthermia to attack the membranes. The magnetic property made it feasible to reuse it. The broad-spectrum antibacterial performance against Gram-negative bacteria, Gram-positive bacteria, drug-resistant bacteria and fungal bacteria was demonstrated. The fabrication of this nanozyme with promising adhesive ability could effectively shorten the diffusive distance of •OH, which provided a potential option for rapid and effective antibacterial treatment.
Declaration of Competing Interest
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.
Acknowledgements
This work was supported by the National Key R & D Program of China (2017YFA0207203), the National Natural Science Foundation of China (Grant No. 21773050), the Harbin Distinguished Young Scholars Fund (No. 2017RAYXJ024), and the State Key Laboratory of Urban Water Resource and Environment (Harbin Institute of Technology) (No. 2020DX03).
Footnotes
Supplementary data to this article can be found online at https://doi.org/10.1016/j.cej.2020.127240.
Appendix A. Supplementary data
The following are the Supplementary data to this article:
References
- 1.Miao H., Teng Z., Wang C., Chong H., Wang G. Recent progress in two-dimensional antimicrobial nanomaterials. Chem. Eur. J. 2019;25:929–944. doi: 10.1002/chem.201801983. [DOI] [PubMed] [Google Scholar]
- 2.Zhao X., Su Y., Qi X., Han X. A facile method to prepare novel Ag2O/Ag2CO3 three-dimensional hollow hierarchical structures and their water purification function. ACS Sustain. Chem. Eng. 2017;5:6148–6158. [Google Scholar]
- 3.Zhu W., Liu X., Tan L., Cui Z., Yang X., Liang Y., Li Z., Zhu S., Yeung K.W.K., Wu S. AgBr nanoparticles in situ growth on 2D MoS2 nanosheets for rapid bacteria-killing and photodisinfection. ACS Appl. Mater. Inter. 2019;11:34364–34375. doi: 10.1021/acsami.9b12629. [DOI] [PubMed] [Google Scholar]
- 4.Zheng K., Setyawati M.I., Leong D.T., Xie J. Antimicrobial silver nanomaterials. Coordin. Chem. Rev. 2018;357:1–17. [Google Scholar]
- 5.Liang Y., Wang M., Zhang Z., Ren G., Liu Y., Wu S., Shen J. Facile synthesis of ZnO QDs@GO-CS hydrogel for synergetic antibacterial applications and enhanced wound healing. Chem. Eng. J. 2019;378 [Google Scholar]
- 6.Cao F., Ju E., Zhang Y., Wang Z., Liu C., Li W., Huang Y., Dong K., Ren J., Qu X. An efficient and benign antimicrobial depot based on silver-infused MoS2. ACS Nano. 2017;11:4651–4659. doi: 10.1021/acsnano.7b00343. [DOI] [PubMed] [Google Scholar]
- 7.Wang Z., Dong K., Liu Z., Zhang Y., Chen Z., Sun H., Ren J., Qu X. Activation of biologically relevant levels of reactive oxygen species by Au/g-C3N4 hybrid nanozyme for bacteria killing and wound disinfection. Biomaterials. 2017;113:145–157. doi: 10.1016/j.biomaterials.2016.10.041. [DOI] [PubMed] [Google Scholar]
- 8.Huang Y., Ren J., Qu X. Nanozymes: classification, catalytic mechanisms, activity regulation, and applications. Chem. Rev. 2019;119:4357–4412. doi: 10.1021/acs.chemrev.8b00672. [DOI] [PubMed] [Google Scholar]
- 9.Tao Y., Ju E., Ren J., Qu X. Bifunctionalized mesoporous silica-supported gold nanoparticles: intrinsic oxidase and peroxidase catalytic activities for antibacterial applications. Adv. Mater. 2015;27:1097–1104. doi: 10.1002/adma.201405105. [DOI] [PubMed] [Google Scholar]
- 10.Wei H., Wang E. Nanomaterials with enzyme-like characteristics (nanozymes): next-generation artificial enzymes. Chem. Soc. Rev. 2013;42:6060–6093. doi: 10.1039/c3cs35486e. [DOI] [PubMed] [Google Scholar]
- 11.Sang Y., Li W., Liu H., Zhang L., Wang H., Liu Z., Ren J., Qu X. Construction of nanozyme-hydrogel for enhanced capture and elimination of bacteria. Adv. Funct. Mater. 2019;29:1900518. [Google Scholar]
- 12.Wei H., Wang E. Fe3O4 magnetic nanoparticles as peroxidase mimetics and their applications in H2O2 and glucose detection. Anal. Chem. 2008;80:2250–2254. doi: 10.1021/ac702203f. [DOI] [PubMed] [Google Scholar]
- 13.Sun H., Zhao A., Gao N., Li K., Ren J., Qu X. Deciphering a nanocarbon-based artificial peroxidase: chemical identification of the catalytically active and substrate-binding sites on graphene quantum dots. Angew. Chem. Int. Edit. 2015;54:7176–7180. doi: 10.1002/anie.201500626. [DOI] [PubMed] [Google Scholar]
- 14.Zhao H., Dong Y., Jiang P., Wang G., Zhang J. Highly dispersed CeO2 on TiO2 nanotube: a synergistic nanocomposite with superior peroxidase-like activity. ACS Appl. Mater. Inter. 2015;7:6451–6461. doi: 10.1021/acsami.5b00023. [DOI] [PubMed] [Google Scholar]
- 15.Wu J., Wang X., Wang Q., Lou Z., Li S., Zhu Y., Qin L., Wei H. Nanomaterials with enzyme-like characteristics (nanozymes): next-generation artificial enzymes (II) Chem. Soc. Rev. 2019;48:1004–1076. doi: 10.1039/c8cs00457a. [DOI] [PubMed] [Google Scholar]
- 16.Wang H., Wan K., Shi X. Recent advances in nanozyme research. Adv. Mater. 2019;31:1805368. doi: 10.1002/adma.201805368. [DOI] [PubMed] [Google Scholar]
- 17.Liang M., Yan X. Nanozymes: from new concepts, mechanisms, and standards to applications. Acc. Chem. Res. 2019;52:2190–2200. doi: 10.1021/acs.accounts.9b00140. [DOI] [PubMed] [Google Scholar]
- 18.Shan J., Li X., Yang K., Xiu W., Wen Q., Zhang Y., Yuwen L., Weng L., Teng Z., Wang L. Efficient bacteria killing by Cu2WS4 nanocrystals with enzyme-like properties and bacteria-binding ability. ACS Nano. 2019;13:13797–13808. doi: 10.1021/acsnano.9b03868. [DOI] [PubMed] [Google Scholar]
- 19.Rizzello L., Sorce B., Sabella S., Vecchio G., Galeone A., Brunetti V., Cingolani R., Pompa P.P. Impact of nanoscale topography on genomics and proteomics of adherent bacteria. Acs Nano. 2011;5:1865–1876. doi: 10.1021/nn102692m. [DOI] [PubMed] [Google Scholar]
- 20.Song H., Ahmad Nor Y., Yu M., Yang Y., Zhang J., Zhang H., Xu C., Mitter N., Yu C. Silica nanopollens enhance adhesion for long-term bacterial inhibition. J. Am. Chem. Soc. 2016;138:6455–6462. doi: 10.1021/jacs.6b00243. [DOI] [PubMed] [Google Scholar]
- 21.Li Y.Q., Zhu B., Li Y., Leow W.R., Goh R., Ma B., Fong E., Tang M., Chen X. A synergistic capture strategy for enhanced detection and elimination of bacteria. Angew. Chem. Int. Edit. 2014;53:5837–5841. doi: 10.1002/anie.201310135. [DOI] [PubMed] [Google Scholar]
- 22.Liu L., Chen S., Xue Z., Zhang Z., Qiao X., Nie Z., Han D., Wang J., Wang T. Bacterial capture efficiency in fluid bloodstream improved by bendable nanowires. Nat. Commun. 2018;9:1–9. doi: 10.1038/s41467-018-02879-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.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. Int. Edit. 2019;58:16236–16242. doi: 10.1002/anie.201908289. [DOI] [PubMed] [Google Scholar]
- 24.Zhang X., Zhang W., Liu L., Yang M., Huang L., Chen K., Wang R., Yang B., Zhang D., Wang J. Antibiotic-loaded MoS2 nanosheets to combat bacterial resistance via biofilm inhibition. Nanotechnology. 2017;28 doi: 10.1088/1361-6528/aa6c9b. [DOI] [PubMed] [Google Scholar]
- 25.Guo N., Zeng Y., Li H., Xu X., Yu H. MoS2 nanosheets encapsulating TiO2 hollow spheres with enhanced photocatalytic activity for nitrophenol reduction. Mater. Lett. 2017;209:417–420. [Google Scholar]
- 26.Hu Z., Wang L., Zhang K., Wang J., Cheng F., Tao Z., Chen J. MoS2 nanoflowers with expanded interlayers as high-performance anodes for sodium-ion batteries. Angew. Chem. Int. Edit. 2014;53:12794–12798. doi: 10.1002/anie.201407898. [DOI] [PubMed] [Google Scholar]
- 27.Wang C., Su Y., Zhao X., Tong S., Han X. MoS2@HKUST-1 flower-like nanohybrids for efficient hydrogen evolution reactions. Chem. Eur. J. 2018;24:1080–1087. doi: 10.1002/chem.201704080. [DOI] [PubMed] [Google Scholar]
- 28.Murugan C., Murugan N., Sundramoorthy A.K., Sundaramurthy A. Nanoceria decorated flower-like molybdenum sulphide nanoflakes: an efficient nanozyme for tumour selective ROS generation and photo thermal therapy. Chem. Commun. 2019;55:8017–8020. doi: 10.1039/c9cc03763b. [DOI] [PubMed] [Google Scholar]
- 29.Yang X., Li J., Liang T., Ma C., Zhang Y., Chen H., Hanagata N., Su H., Xu M. Antibacterial activity of two-dimensional MoS2 sheets. Nanoscale. 2014;6:10126–10133. doi: 10.1039/c4nr01965b. [DOI] [PubMed] [Google Scholar]
- 30.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. doi: 10.1021/acsnano.6b05810. [DOI] [PubMed] [Google Scholar]
- 31.Wang W.-N., Zhang C.-Y., Zhang M.-F., Pei P., Zhou W., Zha Z.-B., Shao M., Qian H.-S. Precisely photothermal controlled releasing of antibacterial agent from Bi2S3 hollow microspheres triggered by NIR light for water sterilization. Chem. Eng. J. 2020;381 [Google Scholar]
- 32.Zhao K., Gu W., Zheng S., Zhang C., Xian Y. SDS–MoS2 nanoparticles as highly-efficient peroxidase mimetics for colorimetric detection of H2O2 and glucose. Talanta. 2015;141:47–52. doi: 10.1016/j.talanta.2015.03.055. [DOI] [PubMed] [Google Scholar]
- 33.Deng H., Li X., Peng Q., Wang X., Chen J., Li Y. Monodisperse magnetic single-crystal ferrite microspheres. Angew. Chem. 2005;117:2842–2845. doi: 10.1002/anie.200462551. [DOI] [PubMed] [Google Scholar]
- 34.Guo N., Li H., Xu X., Yu H. Hierarchical Fe3O4@MoS2/Ag3PO4 magnetic nanocomposites: enhanced and stable photocatalytic performance for water purification under visible light irradiation. Appl. Surf. Sci. 2016;389:227–239. [Google Scholar]
- 35.Wei F., Li J., Dong C., Bi Y., Han X. Plasmonic Ag decorated graphitic carbon nitride sheets with enhanced visible-light response for photocatalytic water disinfection and organic pollutant removal. Chemosphere. 2020;242 doi: 10.1016/j.chemosphere.2019.125201. [DOI] [PubMed] [Google Scholar]
- 36.Xu X., Deng C., Gao M., Yu W., Yang P., Zhang X. Synthesis of magnetic microspheres with immobilized metal ions for enrichment and direct determination of phosphopeptides by matrix-assisted laser desorption ionization mass spectrometry. Adv. Mater. 2006;18:3289–3293. [Google Scholar]
- 37.Liang X., Wang X., Zhuang J., Chen Y., Wang D., Li Y. Synthesis of nearly monodisperse iron oxide and oxyhydroxide nanocrystals. Adv. Funct. Mater. 2006;16:1805–1813. [Google Scholar]
- 38.Ramakrishna Matte H., Gomathi A., Manna A.K., Late D.J., Datta R., Pati S.K., Rao C. MoS2 and WS2 analogues of graphene. Angew. Chem. Int. Edit. 2010;49:4059–4062. doi: 10.1002/anie.201000009. [DOI] [PubMed] [Google Scholar]
- 39.Gao Q., Zhang X., Yin W., Ma D., Xie C., Zheng L., Dong X., Mei L., Yu J., Wang C. Functionalized MoS2 nanovehicle with near-infrared laser-mediated nitric oxide release and photothermal activities for advanced bacteria-infected wound therapy. Small. 2018;14:1802290. doi: 10.1002/smll.201802290. [DOI] [PubMed] [Google Scholar]
- 40.Lin T., Wang J., Guo L., Fu F. Fe3O4@ MoS2 core–shell composites: preparation, characterization, and catalytic application. J. Phys. Chem. C. 2015;119:13658–13664. [Google Scholar]
- 41.Zhu M., Liu X., Tan L., Cui Z., Liang Y., Li Z., Yeung K.W.K., Wu S. Photo-responsive chitosan/Ag/MoS2 for rapid bacteria-killing. J. Hazard. Mater. 2020;383 doi: 10.1016/j.jhazmat.2019.121122. [DOI] [PubMed] [Google Scholar]
- 42.Qiao X.-Q., Zhang Z.-W., Tian F.-Y., Hou D.-F., Tian Z.-F., Li D.-S., Zhang Q. Enhanced catalytic reduction of p-nitrophenol on ultrathin MoS2 nanosheets decorated with noble metal nanoparticles. Cryst. Growth Des. 2017;17:3538–3547. [Google Scholar]
- 43.Xu Y., Geng J., Peng Y., Liu Z., Yu J., Hu X. Lubricating mechanism of Fe3O4@ MoS2 core-shell nanocomposites as oil additives for steel/steel contact. Tribol. Int. 2018;121:241–251. [Google Scholar]
- 44.Metelitza D., Naumchik I., Karasyova E., Polozov G., Shadyro O. Inhibition of peroxidase-catalyzed oxidation of aromatic amines by substituted phenols. Appl. Biochem. Micro+ 2003;39:352–362. [Google Scholar]
- 45.Liu J., Du J., Su Y., Zhao H. A facile solvothermal synthesis of 3D magnetic MoS2/Fe3O4 nanocomposites with enhanced peroxidase-mimicking activity and colorimetric detection of perfluorooctane sulfonate. Microchem. J. 2019;149 [Google Scholar]
- 46.Chen T., Wu X., Wang J., Yang G. WSe2 few layers with enzyme mimic activity for high-sensitive and high-selective visual detection of glucose. Nanoscale. 2017;9:11806–11813. doi: 10.1039/c7nr03179c. [DOI] [PubMed] [Google Scholar]
- 47.Zhang P., Wang Z., Liu L., Klausen L.H., Wang Y., Mi J., Dong M. Modulation the electronic property of 2D monolayer MoS2 by amino acid. Appl. Mater. Today. 2019;14:151–158. [Google Scholar]
- 48.Satheeshkumar E., Bandyopadhyay A., Sreedhara M., Pati S.K., Rao C., Yoshimura M. One-step simultaneous exfoliation and covalent functionalization of MoS2 by amino acid induced solution processes. ChemNanoMat. 2017;3:172–177. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.