Next Article in Journal
CRISPR/Cas9-Targeted Disruption of Two Highly Homologous Arabidopsis thaliana DSS1 Genes with Roles in Development and the Oxidative Stress Response
Next Article in Special Issue
Silver-Based Surface Plasmon Sensors: Fabrication and Applications
Previous Article in Journal
MoaE Is Involved in Response to Oxidative Stress in Deinococcus radiodurans
Previous Article in Special Issue
A Recent Progress in the Leachate Pretreatment Methods Coupled with Anaerobic Digestion for Enhanced Biogas Production: Feasibility, Trends, and Techno-Economic Evaluation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

N-Doped Carbon/CeO2 Composite as a Biomimetic Catalyst for Antibacterial Application

1
Key Laboratory of Marine Environmental Corrosion and Bio-Fouling, Institute of Oceanology, Chinese Academy of Sciences, 7 Nanhai Road, Qingdao 266071, China
2
Open Studio for Marine Corrosion and Protection, Pilot National Laboratory for Marine Science and Technology (Qingdao), 1 Wenhai Road, Qingdao 266237, China
3
Center for Ocean Mega-Science, Chinese Academy of Sciences, 7 Nanhai Road, Qingdao 266071, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(3), 2445; https://doi.org/10.3390/ijms24032445
Submission received: 3 January 2023 / Revised: 22 January 2023 / Accepted: 23 January 2023 / Published: 26 January 2023

Abstract

:
Exploring new and high efficiency mimic enzymes is a vital and novel strategy for antibacterial application. Haloperoxidase-like enzymes have attracted wide attention thanks to their amazing catalytic property for hypohalous acid generation from hydrogen peroxide and halides. However, few materials have displayed halogenating catalytic performance until now. Herein, we synthesized N-doped C/CeO2 (N-C/CeO2) composite materials by a combination of the liquid and solid-state method. N-C/CeO2 can possess haloperoxidase-like catalytic activity by catalyzing the bromination of organic signaling compounds (phenol red) with H2O2 at a wide range of temperatures (20 °C to 55 °C), with a solution color changing from yellow to blue. Meanwhile, it exhibits high catalytic stability/recyclability in the catalytic reaction. The synthesized N-C/CeO2 composite can effectively catalyze the oxidation of Br with H2O2 to produce HBrO without the presence of phenol red. The produced HBrO can resist typical marine bacteria like Pseudomonas aeruginosa. This study provides an efficient biomimetic haloperoxidase and a novel sustainable method for antibacterial application.

1. Introduction

Microbiological contamination constitutes one of the fatal worldwide issues facing both environmental sustainability and public healthcare [1,2]. Various antibacterial methods have been developed to limit microbial growth, such as the addition of excess chlorine dioxide [3] or antibiotics [4,5], even new types of antibacterial materials, such as nano silver [6,7]. However, the toxicity of disinfection byproducts and the bacterial resistance lead to a quest for novel and effective methods [8,9]. Therefore, it is necessary to exploit novel and effective environment-friendly and nontoxic antibacterial materials and antibacterial technology. Different conventional antibacterial agents and biomimetic catalyst emulate nature enzymes to produce intermediates such as halogenated metabolite, which target specific bacterial signaling and regulatory systems for preventing bacterial colonization or biofilm development [10].
In nature, some marine algae can effectively prevent the attachment of microorganisms by self-secreting a haloperoxidase [11,12]. This kind of haloperoxidase can catalyze the oxidation of halides (Cl, Br, I) with H2O2 to the corresponding hypohalous acid [13]. Inspired by this phenomenon, natural haloperoxidase and functional recombinant haloperoxidase, especially vanadium haloperoxidases (V-HPOs), are applied to paint as an additive, which can effectively prevent the growth and attachment of microorganisms [14,15]. However, these natural and functional recombinant enzymes were restricted to large-scale application because of their high production costs, short-term stability, and specific reaction conditions (pH and temperature) [16]. Exploring high-performance artificial V-HPOs mimic enzymes is a useful strategy to replicate natural enzymes.
Attempts to mimic haloperoxidases reactions with synthetic enzymes have been successful in catalytic activity [17]. Research on functional enzyme mimics has seen an upsurge in recent years [10]. Several vanadium complexes [18,19] or V2O5 nanoparticles [16] have been reported as mimicking V-HPOs, which display catalytic efficiency and selectivity in oxidative halogenation reactions [20,21]. However, the vanadium compounds are mutagenic, carcinogenic, and teratogenic [17]. It is an urgent need to develop efficient and non-toxic materials to replace vanadium-based complexes. Inspired by catalyzing oxidation/halogenation reaction of cerium oxide in organic synthesis, cerium-based materials were reported to have haloperoxidase-like activity [22,23,24,25,26]. For example, cerium oxide nanorods as haloperoxidase mimic have been used in antimicrobial membranes [23,24,25]. Cerium oxides present good catalytic performance, which is attributed to the self-structural properties and environmental compatibility [25]. However, the extreme low abundance of rare-earth metallic cerium on Earth limits its large-scale application. Doping is an efficient strategy to reduce the usage amount of cerium and increase the utilization of cerium. In our previous report, compared with cerium oxide, the same amount carbon-doped cerium oxide exhibited better haloperoxidase mimicry for antimicrobial [26]. Therefore, specific doping and complex can effectively reduce resource utilization and facilitate the widespread application of cerium-based materials.
Herein, N-C/CeO2 composite was prepared and studied as haloperoxidase mimicry for antibacterial, as shown in Scheme 1. N-C/CeO2 composite as a biomimetic catalyst possesses haloperoxidase-like catalytic activity by catalyzing the bromination of phenol red in the presence of H2O2 with a solution color changing from yellow to blue. Meanwhile, it can possess antibacterial application by catalyzing the oxidation of Br with H2O2 (without phenol red) to produce HBrO. The haloperoxidase activity of prepared material and the factors affecting the mimicry activity, such as temperature and concentration of catalysts, were studied. The kinetics of the catalytic reaction were investigated by varying the concentration of one reactant while keeping the concentration of others constant. Consequently, the stability and recyclability of N-C/CeO2 composite were proved through the reutilization test. It can effectively catalyze the oxidative bromination of Br and H2O2 to produce HBrO. The produced HBrO with a strong antibacterial activity was used to resist microorganisms such as Escherichia coli (E. coli), Pseudomonas aeruginosa (P. aeruginosa), and Staphylococcus aureus (S. aureus). The proposed high efficiency N-C/CeO2 artificial enzyme mimic may represent a novel strategy to emulate a natural defense system for restraining biofilm growth and bacterial colonization.

2. Results and Discussion

2.1. Characterization

The morphology and crystal structure of the synthesized composites were investigated by SEM, TEM, and XRD. Figure 1A is the SEM image of the prepared N-C/CeO2 composite. N-C/CeO2 composite presents a kind of sheet morphology, and the sheets branch off from each other. The high-resolution TEM (HRTEM) image (Figure 1B) of N-C/CeO2 shows clear lattice fringes with an interlayer distance of 0.32 nm, which can be indexed to the (002) plane of CeO2. As shown in Figure 1C, the XRD pattern of N-C/CeO2 composite shows typical peaks at around 28°, 33°, 47°, and 56°, corresponding to the (111), (200), (220), and (311) planes of the cubic CeO2 (PDF#34-0394), respectively. The inset of Figure 1C displays the crystal structure illustration of CeO2 with cubic space group (Fm3m). The diffraction peaks are strong and sharp, implying that the N-C/CeO2 sample maintains good crystallinity. In addition, XRD patterns of the other N-C/CeO2 composites are shown in Figure S1. These results suggest that all of the XRD patterns of N-C/CeO2 composites are nearly identical to the pure CeO2. Moreover, the loading of N-doped carbon did not change the crystalline phase of the composites. As displayed in Figure 1D, the STEM elemental mapping images reveal a uniform distribution of N, C, O, and Ce. This indicates the high homogeneity of the synthesized N-C/CeO2 composite. To summarize, the doping modification in the current study has not changed the morphology and crystal structure of the N-C/CeO2 composites.
XPS was further employed to elucidate the electronic structure and chemical state of N-C/CeO2. As shown in Figure 1E, the refined Ce 3d XPS spectrum is composed of multiple couples of peaks, corresponding to a mixture of Ce3+ and Ce4+ oxidation states. Generally, cerium switches reversibly between its Ce(Ⅲ) and Ce(Ⅳ) states owing to the non-stoichiometric nature and multiple d-splitting of Ce element [27]. The Ce 3d XPS peaks located at 885/904 eV can be assigned to Ce3+ states, while the peaks at 883/901 eV, 889/908 eV, and 898/916 eV are attributed to Ce4+ states [28]. According to the integrated area, the N-C/CeO2 composite consisted of a major amount of Ce(Ⅳ) oxide (85.93%) and small amount of Ce(Ⅲ) oxide (14.08%), which is related to the lattice strain induced by Ce3+ and the presence of oxygen vacancies [29]. The O 1s XPS spectrum is given in Figure S2. The O1s XPS spectrum displays typical peaks at 529.5 and 531.6 eV, which can be assigned to the Ce4+-O and Ce3+-O bond, respectively. After high-temperature calcination, the final N-C/CeO2 product also contains abundant nitrogen-doped carbon components. The N 1s XPS spectrum (Figure 1F) of the N-C/CeO2 composite can be de-convoluted into three species, corresponding to pyridinic N, pyrrolic N, and oxidized N, respectively [30]. Additionally, the C 1s XPS spectrum (Figure 1G) exhibits significant signals at 284.5, 285, and 286 eV, which can be identified as C-C/C=O, C-N, and O-C=O/N-C=O functional groups, respectively [31]. In Raman spectra (Figure S3), the pure CeO2 delivers typical peaks at 460 cm−1 and 600/1170 cm−1, which can be assigned to the F2g vibration model of the CeO2 and oxygen defects, respectively [32]. However, the N-C/CeO2 displays additional Raman spectrum peaks at around 1680 and 2900 cm−1, corresponding to the C-N band in the composite [33]. Meanwhile, N-C does not change the crystalline phase of the composites, which is consistent with the results of XRD. The above results indicate that the N-C/CeO2 composite contains ceric oxide and abundant N (O and C-)-functional groups. These are crucial to the high homogeneity of the prepared composites.

2.2. Haloperoxidase Mimicry Activity

To study the haloperoxidase mimicry activity of N-C/CeO2 composites, phenol red (PR) was used as the color substrate. N-C/CeO2 composite, as the haloperoxidase mimicry, can catalyze the bromination and PR in the presence of H2O2 with a solution color change from yellow to blue. N-C/CeO2 composite (6:1) exhibits the best haloperoxidase mimicry activity out of all of the N-C/CeO2 composites in Figure S4. Thus, this doping composite was used throughout the study. As shown in Figure 2A, the absorbance spectra of the solution were obtained in different reaction systems. No obvious absorption signals are detected in the c blank system with the components of NH4Br + H2O + H2O2 + c. However, there is an obvious peak at ~430 nm in systems a, b, and d, which belongs to the absorption of PR. System e reveals a distinct absorbance at 590 nm, which is attributed to the product bromophenol blue. The N-C/CeO2 composite can catalyze Br and PR in the presence of H2O2 to produce bromophenol blue. The corresponding color changes in different systems are shown in Figure 2A (insert). The solution is colorless in system c and yellow in systems a, b, and d. This yellow color comes from the color of dilute PR dye. The solution is blue only in system e. It indicates that the N-C/CeO2 composite can catalyze Br and PR in the presence of H2O2 to produce a blue-color reaction and exhibit good haloperoxidase-like catalytic activity. In addition, to further test the haloperoxidase-like activity of N-C/CeO2 composites, the catalytic activities of N-C composites and pure CeO2 were investigated as controls in Figure 2B. There is a faint peak at 590 nm of the N-C composite, which shows that the N-C composite has a certain catalytic activity. N-C/CeO2 delivers a significant absorbance peak at 590 nm, and its adsorption peak is higher absorbance than that of pure CeO2, indicating the N dopants carbon as an electron donor atom can facilitate the catalytic activity of CeO2 to produce HBrO [10]. These above results indicate that the N-C/CeO2 composite possesses excellent haloperoxidase-like activity higher than that of N-C composites and pure CeO2. Therefore, the use of nitrogen-doped carbon as a substrate is beneficial to improve the haloperoxidase-like catalytic activity of the N-C/CeO2 composite.
In order to study the haloperoxidase-like properties of N-C/CeO2 composite, the UV–Vis spectra of aqueous reaction were measured every 2 min within a total testing time of 40 min. As shown in Figure 2C, the absorbance at 590 nm increases quickly at the early time, while the increase of absorbance at 590 nm slows down after 30 min and tends to be stable at 40 min. The formation rate of bromination product (Br4PR) was evaluated by the accurate absorbance intensity of 590 nm, as displayed in Figure S5. These results show that the N-C/CeO2 composite has the same haloperoxidase-like catalytic activity as natural enzymes [34]. In general, the catalytic activity of artificial mimicry is associated with the working temperature. The optimal temperature of the N-C/CeO2 composite was investigated from 20 °C to 55 °C. The result shown in Figure 2D indicates that the catalytic activities of the N-C/CeO2 composite were high at various temperatures. These are only slightly affected by the temperature. Therefore, room temperature was chosen in the following experiments.

2.3. Kinetics Constant Determination

The reaction kinetics of the N-C/CeO2 composite were further studied. The Michaelis–Menten constant of substrates (H2O2, NH4Br, and PR) was measured by changing the concentration of one substrate, while keeping all other concentrations constant. Then, kinetics graphs (Figure 3) were obtained by the calculated initial velocity rates using kinetic data according to Equations (2) and (3). For the N-C/CeO2 composite, Figure 3A (left) shows the kinetic function as the concentration of the N-C/CeO2 composite, when the concentrations of other substrates are fixed. The kinetic values are fitted nonlinearly (blue line) according to the Michaelis–Menten equation, and the partial kinetic values are fitted linearly (dark dashed line). The results show that the kinetic data fit perfectly with the nonlinear fitting line based on the Michaelis–Menten equation. Thus, the N-C/CeO2 composite complies with the Michaelis–Menten kinetic of natural enzymes. The middle graph and the right graph in Figure 3A show the corresponding Lineweaver–Burk linearizations and logarithmic correlations, respectively.
The Michaelis–Menten constants (Km) and the maximal reaction rates (vmax) values of all substrates (H2O2, NH4Br, and PR) calculated are shown in Table 1. The kinetic function, corresponding Lineweaver–Burk linearizations, and logarithmic correlations of other substrates, including H2O2, NH4Br, and PR, are treated with the similar method to that of the N-C/CeO2 substrate (Figure 3B–D). The above results show that the N-C/CeO2 composite as haloperoxidase mimicry matches with the catalytic reaction kinetics of natural enzymes.
The Michaelis–Menten constant of the N-C/CeO2 composite and vanadium bromoperoxidase (V-BPO) in the previous reports are summarized in Table 2. Compared with the V-BPO biological sample, the prepared N-C/CeO2 composite delivers an obviously smaller Km of H2O2 and bigger Km of bromide (NH4Br) [16,34]. Generally, Km indicates the affinity between the substrate and catalyst. The lower Km value of H2O2 suggests that H2O2 has a higher affinity for the surface of the N-C/CeO2 composite than V-BPO. The higher Km value of Br indicates that Br has a lower affinity for the surface of the N-C/CeO2 composite.
To evaluate the effect of bromide source on the catalytic reaction, KBr and NaBr are used as control samples. Figure 4A shows the absorbance of the solution at 590 nm in the presence of NH4Br, KBr, and NaBr, respectively. The solution absorbances are almost identical for different bromide sources, meaning that the reaction is independent of the bromide source. As the stability of the catalyst is essential for real applications, reutilization tests of N-C/CeO2 composite are performed with the same concentration of PR, NH4Br, and H2O2 at room temperature. After each reaction cycle, the N-C/CeO2 composite is separated by centrifugation at 3020× g and washed with ultrapure water. Then, the obtained N-C/CeO2 composite is treated again with PR, NH4Br, and H2O2 under identical experimental conditions. As shown in Figure 4B, the absorbance at 590 nm stays almost constant through ten cycles. This clearly illustrates that the activity of the N-C/CeO2 composite has not decreased. The above results suggest that the catalytic activity of the N-C/CeO2 composite is independent of the bromide source, and it also exhibits high catalytic stability.

2.4. Antibacterial Test

N-C/CeO2 composites have good haloperoxidase mimicry activity by catalyzing the bromination of organic signaling compounds. Therefore, N-C/CeO2 composites can catalyze the reaction of H2O2 and Br to produce HBrO. In order to investigate the antibacterial property of N-C/CeO2 composites as haloperoxidase mimicry, N-C/CeO2 composites are applied onto the titanium plates’ surfaces and antibacterial tests are conducted. As shown in Figure 5A, the bare titanium plate/N-C/CeO2 composites modified titanium plates are exposed to P. aeruginosa suspensions at 37 °C for 4 h. Bacterial cell density and adhesion is further evaluated by fluorescence microscopy. As a control, a dense P. aeruginosa population is observed on the bare titanium plates surfaces in the absence of N-C/CeO2 composites in the medium without H2O2 and Br (Figure 5A, left column “Blank”). The same experimental set up is conducted without adding the substrates H2O2 and Br in P. aeruginosa suspensions. In this case, high P. aeruginosa adhesion/density is also observed on the N-C/CeO2 modified titanium plates (Figure 5A, middle column). In contrast, the absence of P. aeruginosa adhesion is detected on the N-C/CeO2 composites modified titanium plates in the presence of substrates H2O2 and Br (Figure 5A right column “N-C/CeO2 + Br + H2O2”). The above results indicate that the system of “N-C/CeO2 + Br + H2O2” exhibits the best antibacterial adhesion property. As shown in Figure S6, the proposed catalytic reaction with the prepared N-C/CeO2 catalyst can also work to suppress the microbial adhesion of E. coli and S. aureus. As shown in Figure 5B, the blank sample without adding N-C/CeO2 catalyst and Br reveals abundant P. aeruginosa colonies on the entire plate. Figure 5C displays the plate treated with only N-C/CeO2 catalyst (without Br and H2O2), and Figure 5D reveals the plate photo treated with all components of the catalytic condition (with N-C/CeO2, Br, and H2O2). The sample with the addition of only the N-C/CeO2 catalyst exhibits a reduced number of P. aeruginosa colonies, implying that the N-C/CeO2 composite itself has weak antibacterial activity (Figure 5C). However, almost no P. aeruginosa colonies are detected on the plate of Figure 5D because of the generation of HBrO. These above results demonstrate that the N-C/CeO2 composite itself has weak antibacterial activity, while the N-C/CeO2 composites can catalyze the reaction of Br and H2O2 to produce HBrO, which plays a major role in antibacterial properties. In addition, the antibacterial properties of N-C/CeO2 as haloperoxidase mimicry were compared with the previously reported CeO2-based materials, as shown in Table S1. N-C/CeO2 presents lower bacterial attachment on the titanium plates than other CeO2-based materials for E. coli, which indicates that they have good antibacterial adhesion properties. This provides a novel way to prevent biofouling and attachment to marine facilities. Therefore, N-C/CeO2 composites as haloperoxidase mimics have excellent bromination activity, and the produced hypobromous acid exhibits superb antibacterial activity.

3. Methods

3.1. Reagents and Apparatus

Cerium(III) nitrate hexahydrate (Ce(NO3)3•6H2O) was purchased from Aladdin Chemical Reagent Co., Ltd. (Shanghai, China). Phenol red (PR), melamine, ammonium bromide, NaCl, KCl, Na2HPO4, KH2PO4, acetic acid, and hydrogen peroxide solution (30%) were purchased from Sinopharm Chemical Reagent Co. Ltd. (Shanghai, China). Cell staining kit (K2081) was purchased from APExBIO (Houston, TX, USA). All the reagents and chemicals were used without further purification. All aqueous solutions were prepared with ultra-pure water (18.2 MΩ·cm) throughout this experiment. Phosphate buffered saline (PBS, 0.1 mmol·L−1) was prepared with 8.0 g·L−1 NaCl, 0.2 g·L−1 KCl, 1.44 g·L−1 Na2HPO4, and 0.44 g·L−1 KH2PO4 in ultra-pure water. Then, the pH of the solution is regulated to 7.0 by NaOH solution. PBS (0.1 mmol·L−1, 7.0) was used in the whole experiment. The aqueous standard solutions of H2O2 were stored in the dark because of their photosensitivity.
The morphology and structure investigation of the synthesized N-C/CeO2 composites were carried out by scanning electron microscopy (SEM, Reguas, Japan). The phase structures of these electrodeposits were determined using X-ray diffraction (XRD, Rigaku D/max-Ultima IV, Tokyo, Japan). The fine structures of these samples were further investigated by transmission electron microscopy TEM (JEM 2100F, Tokyo, Japan). The heteroatoms and functional groups were determined by X-ray photoelectron spectroscopy XPS (Escalab K-alpha 250Xi). The Raman spectra were collected on Renishaw MZ20-FC Raman microscope. The absorption spectra of UV–Vis and absorbance-time were measured with an UV–Vis spectrophotometer (UV–Vis, U-3900 HITACHI, Tokyo, Japan). Observation of bacteria was performed using a fluorescence microscopy (BX-51 with image software of Cellsens, Olympus, Japan) after staining with K2081 kit, as previously described [35].

3.2. Synthesis of N-C/CeO2 Composites

Herein, 3.0 g melamine and different amounts of Ce(NO3)3•6H2O were dissolved into 40 mL acetic acid and 40 mL ultra-pure water. The obtained solutions were mixed by ultrasonication for 30 min and transferred into a stainless-steel vessel. The hydrothermal reaction was carried out at 120 °C for 12 h. Thereafter, the solvent was removed from the product by the freezing drying process using vacuum equipment. The resultant materials were annealed at 520 °C for 4 h at a ramp rate of 5 °C·min−1 in the air. Different N-C/CeO2 composites were obtained by varying the mass ratio of melamine and Ce(NO3)3•6H2O (6:1, 3:1, 2:1, 1:1, 1:2).

3.3. Haloperoxidase-like Activity of N-C/CeO2 Composites

The haloperoxidase-like activity of the synthesized N-C/CeO2 composites was analyzed using an optical absorption spectroscopy. The reaction scheme was as follows: N-C/CeO2 composite catalyzes the oxidative bromination H2O2 and Br in the presence of PR, resulting in the color change from yellow to blue. The 950 μL mixed solutions (containing 28 μmol·L−1 PR, 69.4 mmol·L−1 NH4Br, 830 μmol·L−1 H2O2, and 50 μg·mL−1 N-C/CeO2 composites) reacted at room temperature for 40 min. Afterwards, the absorption was measured by UV–Vis spectroscopy. As a control, the absorption spectra of mixtures were measured when one of the substrates was absent in all mixtures. The amount of the added reagent was quantified. In addition, the optimal reaction conditions such as temperature, H2O2 concentration, and N-C/CeO2 composite concentration were tested by changing one reaction condition while leaving other conditions unchanged. Three replicate experiments were performed.

3.4. Determination of Kinetic Constant

The kinetic constants were carried out in time course mode of UV–Vis by fixing the wavelength at 590 nm [20]. The absorbance of mixed solutions was measured by changing the concentration of one reactant while keeping others constant in kinetic tests. In order to obtain the optimal concentration of all reactants, each measurement was carried out at 590 nm for 40 min. In addition, kinetic parameters were calculated based on the slopes (dA590nm/dt), which were kept constant over 5 min. The kinetic constants (Michaelis–Menten constant Km and the maximum reaction velocity vmax) were obtained using the Linewaver–Burk linearization (Equation (1)) [36,37].
1/v = Km/vmax[C] + 1/vmax
where v is the initial velocity and C is the concentration of substrate. In order to evaluate the Km and vmax, v was calculated. In our experiments, the product Br4PR was used as a measure of the reaction rate to obtain the initial reaction rate (Equation (2)).
v = d[Br4PR]/dt
The Br4PR concentration was obtained according to the Lambert–Beer law [25] (Equation (3)).
[Br4PR] = A590/dεBr4PR
where εBr4PR is the extinction coefficient of Br4PR and its value is 72,200 L·mol−1·cm−1.
In addition, in order to test the dependence on the Br source, some Br-salts such as KBr, NaBr, and NH4Br served as the Br source in the mixture reaction solutions. The reutilization test of N-C/CeO2 composite was carried out in ten recycles with N-C/CeO2, PR, NH4Br, and H2O2.

3.5. Bacterial Adhesion Tests

For bacterial adhesion tests, P. aeruginosa (Gram-negative, typical marine bacterium, risk group 2 organism) as a model bacteria was grown in Luria Broth (LB) medium with shaking at 160 rpm and 37 °C for 12 h. The cell concentration of P. aeruginosa in the medium was calculated using the plate colony counting method [38]. Here, 107 colony-forming units (cfu) mL−1  P. aeruginosa were separately obtained by centrifugation. These cells were resuspended in 0.1 mmol·L−1 PBS to obtain a cell concentration of 107 cfu·mL−1. Multiple sets of 10 mL of this PBS cell suspension solutions were placed into 50 mL inoculation tubes and used for later bacterial adhesion tests.
Titanium plates (Ti, 0.1 × 1 × 1 cm3), with/without N-C/CeO2 composite, were placed into the above bacterial solution and cultivated with agitation at 37 °C for 4 h in different systems: (1) Ti without Br and H2O2 (blank), (2) modified Ti without Br and H2O2 (N-C/CeO2), and (3) modified Ti with Br and H2O2 (N-C/CeO2 + Br + H2O2). Three replicate experiments were performed per system. Afterwards, Ti was stained using a staining kit (K2081) for 15 min in the dark. Excess stain was gently removed by sterile PBS. The stained samples were examined by fluorescence microscopy. Bacterial solutions containing the same concentration of N-C/CeO2 composite, Br, and H2O2 were cultured on an agar plate at 37 °C for 24 h. Three parallel agar plates were painted for each bacterial solution. As a control, PBS bacteria solutions containing the N-C/CeO2 composite were cultured under the same conditions. Three parallel experiments were performed. E. coli (Gram-negative) and S. aureus (Gram-positive) were also studied using the same experimental method. These above results show that the N-C/CeO2 composite as haloperoxidase mimicry presents good antibacterial activity.

4. Conclusions

In summary, the N-C/CeO2 composite was successfully synthesized using melamine as a carbon and nitrogen source. The N-C/CeO2 composite can effectively catalyze the oxidation of H2O2 with bromination of organic signaling compounds to produce a blue-color reaction, and presents excellent intrinsic haloperoxidase mimicry activity. The catalytic activity of the N-C/CeO2 composite is influenced by the substrate concentration and almost not influenced by temperature. The N-C/CeO2 composite as haloperoxidase mimicry can catalyze the reaction process of H2O2, Br, and PR, which complied with the typical Michaelis–Menton kinetics process. The N-C/CeO2 composite shows good catalytic stability and recyclability in multiple reaction cycles. In the absence of phenol red, the produced HBrO catalyzed by N-C/CeO2 composites presents good antibacterial activity against the model bacteria, especially P. aeruginosa. The N-doped carbon/CeO2 composite as a biomimetic catalyst for antibacterial application is a novel and efficient “green” strategy to emulate and utilize a natural defense system for preventing bacterial colonization and biofilm growth. However, the catalytic activity of the N-doped carbon/CeO2 composite is mainly attributed to the action of CeO2, and the formation mechanism of the halogenated reactive oxygen species needs to be further improved. This work introduces a stable, green, and environment-friendly biomimetic material for antibacterial applications.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/ijms24032445/s1, Additional information including XRD pattern, XPS curves, UV/Vis absorption spectra, time-dependent kinetics spectra, and live/dead staining images of E. coli and S. aureus.

Author Contributions

Conceptualization, N.W.; methodology, N.W.; software, F.G.; validation, N.W. and R.Z.; formal analysis, X.Z.; investigation, N.W.; resources, R.Z.; data curation, N.W.; writing—original draft preparation, N.W.; writing—review and editing, R.Z.; visualization, N.W.; supervision, J.D.; project administration, N.W., J.D. and B.H.; funding acquisition, N.W. All authors have read and agreed to the published version of the manuscript.

Funding

The present work was supported by China Postdoctoral Science Foundation (2021M703246), Shandong Provincial Natural Science Youth Fund Project (ZR2022QD001), the Key Research Program of Frontier Sciences, Chinese Academy of Sciences (ZDBS-LY-DQC025), National Natural Science Foundation of China for Exploring Key Scientific Instrument (No. 41827805), Postdoctoral Innovation Project of Shandong Province, and Applied Basic Research Programs of Qingdao.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the plots within this paper are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mazurkiewicz, K.; Jeż-Walkowiak, J.; Michałkiewicz, M. Physicochemical and microbiological quality of rainwater harvested in underground retention tanks. Sci. Total Environ. 2022, 814, 152701. [Google Scholar] [CrossRef]
  2. Wang, J.; Wang, Y.; Zhang, D.; Xu, C.; Xing, R. Dual response mimetic enzyme of novel Co4S3/Co3O4 composite nanotube for antibacterial application. J. Hazard. Mater. 2020, 392, 122278. [Google Scholar] [CrossRef]
  3. Han, J.; Zhang, X.; Li, W.; Jiang, J. Low chlorine impurity might be beneficial in chlorine dioxide disinfection. Water Res. 2021, 188, 116520. [Google Scholar] [CrossRef]
  4. Luther, A.; Urfer, M.; Zahn, M.; Müller, M.; Wang, S.-Y.; Mondal, M.; Vitale, A.; Hartmann, J.-B.; Sharpe, T.; Monte, F.L.; et al. Chimeric peptidomimetic antibiotics against Gram-negative bacteria. Nature 2019, 576, 452–458. [Google Scholar] [CrossRef]
  5. Patteson, J.B.; Putz, A.T.; Tao, L.; Simke, W.C.; Bryant, L.H.; Britt, R.D.; Li, B. Biosynthesis of fluopsin C, a copper-containing antibiotic from Pseudomonas aeruginosa. Science 2021, 374, 1005–1009. [Google Scholar] [CrossRef]
  6. Jin, Y.; Yang, Y.; Duan, W.; Qu, X.; Wu, J. Synergistic and on-demand Release of Ag-AMPs loaded on porous silicon nanocarriers for antibacteria and wound healing. ACS Appl. Mater. Interfaces 2021, 13, 16127–16141. [Google Scholar] [CrossRef]
  7. Abdulsada, Z.; Kibbee, R.; Schwertfeger, D.; Princz, J.; DeRosa, M.; Örmeci, B. Fate and removal of silver nanoparticles during sludge conditioning and their impact on soil health after simulated land application. Water Res. 2021, 206, 117757. [Google Scholar] [CrossRef]
  8. Lv, L.; Yu, X.; Xu, Q.; Ye, C. Induction of bacterial antibiotic resistance by mutagenic halogenated nitrogenous disinfection byproducts. Environ. Pollut. 2015, 205, 291–298. [Google Scholar] [CrossRef]
  9. Stracy, M.; Snitser, O.; Yelin, I.; Amer, Y.; Parizade, M.; Katz, R.; Rimler, G.; Wolf, T.; Herzel, E.; Koren, G.; et al. Minimizing treatment-induced emergence of antibiotic resistance in bacterial infections. Science 2022, 375, 889–894. [Google Scholar] [CrossRef]
  10. Herget, K.; Frerichs, H.; Pfitzner, F.; Tahir, M.N.; Tremel, W. Functional enzyme mimics for oxidative halogenation reactions that combat biofilm formation. Adv. Mater. 2018, 30, e1707073. [Google Scholar] [CrossRef]
  11. de Boer, E.; Plat, H.; Tromp, M.G.M.; Wever, R.; Franssen, M.C.R.; van der Plas, H.C.; Meijer, E.M.; Schoemaker, H.E. Vanadium containing bromoperoxidase: An example of an oxidoreductase with high operational stability in aqueous and organic media. Biotechnol. Bioeng. 1987, 30, 607–610. [Google Scholar] [CrossRef] [PubMed]
  12. Sandy, M.; Carter-Franklin, J.N.; Martin, J.D.; Butler, A. Vanadium bromoperoxidase from Delisea pulchra: Enzyme-catalyzed formation of bromofuranone and attendant disruption of quorum sensing. Chem. Commun. 2011, 47, 12086–12088. [Google Scholar] [CrossRef] [PubMed]
  13. Butler, A.; Sandy, M. Mechanistic considerations of halogenating enzymes. Nature 2009, 460, 848–854. [Google Scholar] [CrossRef]
  14. Kristensen, J.B.; Meyer, R.L.; Laursen, B.S.; Shipovskov, S.; Besenbacher, F.; Poulsen, C.H. Antifouling enzymes and the biochemistry of marine settlement. Biotechnol. Adv. 2008, 26, 471–481. [Google Scholar] [CrossRef]
  15. Hasan, Z.; Renirie, R.; Kerkman, R.; Ruijssenaars, H.J.; Hartog, A.F.; Wever, R. Laboratory-evolved vanadium chloroperoxidase exhibits 100-fold higher halogenating activity at alkaline pH: Catalytic effects from first and second coordination sphere mutations. J. Biol. Chem. 2006, 281, 9738–9744. [Google Scholar] [CrossRef] [Green Version]
  16. Natalio, F.; Andre, R.; Hartog, A.F.; Stoll, B.; Jochum, K.P.; Wever, R.; Tremel, W. Vanadium pentoxide nanoparticles mimic vanadium haloperoxidases and thwart biofilm formation. Nat. Nanotechnol. 2012, 7, 530–535. [Google Scholar] [CrossRef]
  17. Assem, F.L.; Levy, L.S. A review of current toxicological concerns on vanadium pentoxide and other vanadium compounds: Gaps in knowledge and directions for future research. J. Toxicol. Env. Health B Crit. Rev. 2009, 12, 289–306. [Google Scholar] [CrossRef] [PubMed]
  18. Mimoun, H.; Saussine, L.; Daire, E.; Postel, M.; Fischer, J.; Weiss, R. Vanadium(V) peroxy complexes. New versatile biomimetic reagents for epoxidation of olefins and hydroxylation of alkanes and aromatic hydrocarbons. J. Am. Chem. Soc. 1983, 105, 3101–3110. [Google Scholar] [CrossRef]
  19. Yalcin, O.; Molinari Erwin, J.E.; Gerceker, D.; Onal, I.; Wachs, I.E. Role of local structure on catalytic reactivity: Comparison of methanol oxidation by aqueous bioinorganic enzyme mimic (vanadium haloperoxidase) and vanadia-based heterogeneous catalyst (Supported VO4/SiO2). ACS Catal. 2019, 10, 1566–1574. [Google Scholar] [CrossRef]
  20. Colpas, G.J.; Hamstra, B.J.; Kampf, J.W.; Pecoraro, V.L. Functional models for vanadium haloperoxidase:  Reactivity and mechanism of halide oxidation. J. Am. Chem. Soc. 1996, 118, 3469–3478. [Google Scholar] [CrossRef]
  21. Conte, V.; Coletti, A.; Floris, B.; Licini, G.; Zonta, C. Mechanistic aspects of vanadium catalysed oxidations with peroxides. Coord. Chem. Rev. 2011, 255, 2165–2177. [Google Scholar] [CrossRef]
  22. Cheng, Y.; Liang, L.; Ye, F.; Zhao, S. Ce-MOF with intrinsic haloperoxidase-like activity for ratiometric colorimetric detection of hydrogen peroxide. Biosensors 2021, 11, 204. [Google Scholar] [CrossRef] [PubMed]
  23. Hu, M.; Korschelt, K.; Viel, M.; Wiesmann, N.; Kappl, M.; Brieger, J.; Landfester, K.; Therien-Aubin, H.; Tremel, W. Nanozymes in nanofibrous mats with haloperoxidase-like activity to combat biofouling. ACS Appl. Mater. Interfaces 2018, 10, 44722–44730. [Google Scholar] [CrossRef] [PubMed]
  24. He, X.; Tian, F.; Chang, J.; Bai, X.; Yuan, C.; Wang, C.; Neville, A. Haloperoxidase mimicry by CeO2–x nanorods of different aspect ratios for antibacterial performance. ACS Sustain. Chem. Eng. 2020, 8, 6744–6752. [Google Scholar] [CrossRef]
  25. Herget, K.; Hubach, P.; Pusch, S.; Deglmann, P.; Gotz, H.; Gorelik, T.E.; Gural’skiy, I.A.; Pfitzner, F.; Link, T.; Schenk, S.; et al. Haloperoxidase Mimicry by CeO2-x Nanorods Combats Biofouling. Adv. Mater. 2017, 29, e1603823. [Google Scholar] [CrossRef]
  26. Wang, N.; Li, W.; Ren, Y.; Duan, J.; Zhai, X.; Guan, F.; Wang, L.; Hou, B. Investigating the properties of nano core-shell CeO2@C as haloperoxidase mimicry catalyst for antifouling applications. Colloids Surf. A: Physicochem. Eng. Asp. 2021, 608, 125592. [Google Scholar] [CrossRef]
  27. Barth, C.; Laffon, C.; Olbrich, R.; Ranguis, A.; Parent, P.; Reichling, M. A perfectly stoichiometric and flat CeO2(111) surface on a bulk-like ceria film. Sci. Rep. 2016, 6, 21165. [Google Scholar] [CrossRef] [Green Version]
  28. Baldim, V.; Yadav, N.; Bia, N.; Graillot, A.; Loubat, C.; Singh, S.; Karakoti, A.S.; Berret, J.-F. Polymer-coated cerium oxide nanoparticles as oxidoreductase-like catalysts. ACS Appl. Mater. Interfaces 2020, 12, 42056–42066. [Google Scholar] [CrossRef] [PubMed]
  29. Baldim, V.; Bedioui, F.; Mignet, N.; Margaill, I.; Berret, J.F. The enzyme-like catalytic activity of cerium oxide nanoparticles and its dependency on Ce3+ surface area concentration. Nanoscale 2018, 10, 6971–6980. [Google Scholar] [CrossRef] [Green Version]
  30. Dong, Y.; Deng, Y.; Zeng, J.; Song, H.; Liao, S. A high-performance composite ORR catalyst based on the synergy between binary transition metal nitride and nitrogen-doped reduced graphene oxide. J. Mater. Chem. A 2017, 5, 5829–5837. [Google Scholar] [CrossRef]
  31. Wang, H.; Maiyalagan, T.; Wang, X. Review on recent progress in nitrogen-doped graphene: Synthesis, characterization, and its potential applications. ACS Catal. 2012, 2, 781–794. [Google Scholar] [CrossRef]
  32. Du, X.; Zhang, D.; Shi, L.; Gao, R.; Zhang, J. Morphology Dependence of Catalytic Properties of Ni/CeO2 Nanostructures for Carbon Dioxide Reforming of Methane. J. Phys. Chem. C 2012, 116, 10009–10016. [Google Scholar] [CrossRef]
  33. Maiti, S.; Dhawa, T.; Mallik, A.K.; Mahanty, S. CeO2@C derived from benzene carboxylate bridged metal-organic frameworks: Ligand induced morphology evolution and influence on the electrochemical properties as a lithium-ion battery anode. Sustain. Energy Fules 2017, 1, 288–298. [Google Scholar] [CrossRef]
  34. de Boer, E.; Wever, R. The reaction mechanism of the novel vanadium-bromoperoxidase. A steady-state kinetic analysis. J. Biol. Chem. 1988, 263, 12326–12332. [Google Scholar] [CrossRef]
  35. Zhai, X.; Ju, P.; Guan, F.; Duan, J.; Wang, N.; Zhang, Y.; Li, K.; Hou, B. Biofilm inhibition mechanism of BiVO4 inserted zinc matrix in marine isolated bacteria. J. Mater. Sci. Technol. 2021, 75, 86–95. [Google Scholar] [CrossRef]
  36. Liu, M.; Zhao, H.; Chen, S.; Yu, H.; Quan, X. Interface engineering catalytic graphene for smart colorimetric biosensing. ACS Nano 2012, 6, 3142–3151. [Google Scholar] [CrossRef]
  37. Wang, N.; Li, B.; Qiao, F.; Sun, J.; Fan, H.; Ai, S. Humic acid-assisted synthesis of stable copper nanoparticles as a peroxidase mimetic and their application in glucose detection. J. Mater. Chem. B 2015, 3, 7718–7723. [Google Scholar] [CrossRef]
  38. Kim, D.J.; Chung, S.G.; Lee, S.H.; Choi, J.W. Relation of microbial biomass to counting units for Pseudomonas aeruginosa. Afr. J. Microbiol. Res. 2012, 6, 4620–4622. [Google Scholar]
Scheme 1. Bactericidal mechanism of N-C/CeO2 as haloperoxidase-like.
Scheme 1. Bactericidal mechanism of N-C/CeO2 as haloperoxidase-like.
Ijms 24 02445 sch001
Figure 1. (A) SEM image; (B) HRTEM image; (C) XRD pattern of N-C/CeO2 composite and the standard PDF cards of CeO2, inset: crystal structure of CeO2; (D) STEM-mapping and XPS pattern of N-C/CeO2 composite; (E) Ce 3d; (F) N 1s; (G) C 1s.
Figure 1. (A) SEM image; (B) HRTEM image; (C) XRD pattern of N-C/CeO2 composite and the standard PDF cards of CeO2, inset: crystal structure of CeO2; (D) STEM-mapping and XPS pattern of N-C/CeO2 composite; (E) Ce 3d; (F) N 1s; (G) C 1s.
Ijms 24 02445 g001
Figure 2. (A) Absorbance spectra in different reaction systems: (a) H2O + PR + H2O2 + C, (b) NH4Br + PR + H2O + C, (c) NH4Br + H2O + H2O2 + C, (d) NH4Br + PR + H2O2 + H2O, and (e) NH4Br + PR + H2O2 + C. C: N-C/CeO2. (B) Absorbance spectra of solutions catalyzed by N-C composites, CeO2, and N-C/CeO2 composites. (C) Time-dependent UV–Vis spectra showing the kinetics of the oxidative bromination of PR catalyzed by the N-C/CeO2 composite (40 min, 23–25 °C) and (D) point plot of solutions catalyzed by the N-C/CeO2 composite at different temperatures.
Figure 2. (A) Absorbance spectra in different reaction systems: (a) H2O + PR + H2O2 + C, (b) NH4Br + PR + H2O + C, (c) NH4Br + H2O + H2O2 + C, (d) NH4Br + PR + H2O2 + H2O, and (e) NH4Br + PR + H2O2 + C. C: N-C/CeO2. (B) Absorbance spectra of solutions catalyzed by N-C composites, CeO2, and N-C/CeO2 composites. (C) Time-dependent UV–Vis spectra showing the kinetics of the oxidative bromination of PR catalyzed by the N-C/CeO2 composite (40 min, 23–25 °C) and (D) point plot of solutions catalyzed by the N-C/CeO2 composite at different temperatures.
Ijms 24 02445 g002
Figure 3. Kinetics of the fourfold oxidative bromination of PR as a function of the substrate concentrations. The kinetics as a function of the substrate (A) C; (B) H2O2; (C) NH4Br and (D) PR. Note: C represents the N-C/CeO2 composite, blue line: nonlinearly fitted line of kinetic values, dark dashed line: the linearly fitted line of partial kinetic values, pink arrow: the fitting region.
Figure 3. Kinetics of the fourfold oxidative bromination of PR as a function of the substrate concentrations. The kinetics as a function of the substrate (A) C; (B) H2O2; (C) NH4Br and (D) PR. Note: C represents the N-C/CeO2 composite, blue line: nonlinearly fitted line of kinetic values, dark dashed line: the linearly fitted line of partial kinetic values, pink arrow: the fitting region.
Ijms 24 02445 g003
Figure 4. (A) Dependence on the bromine source and (B) reutilization test of the N-C/CeO2 composite.
Figure 4. (A) Dependence on the bromine source and (B) reutilization test of the N-C/CeO2 composite.
Ijms 24 02445 g004
Figure 5. (A) Cell staining images of P. aeruginosa treated with titanium plate in different systems: blank, N-C/CeO2, and N-C/CeO2 + Br + H2O2, P. aeruginosa agar photo; (B) blank; (C) after adding N-C/CeO2; and (D) after adding N-C/CeO2 + Br + H2O2. Note: the concentration of P. aeruginosa is 107 cfu·mL−1.
Figure 5. (A) Cell staining images of P. aeruginosa treated with titanium plate in different systems: blank, N-C/CeO2, and N-C/CeO2 + Br + H2O2, P. aeruginosa agar photo; (B) blank; (C) after adding N-C/CeO2; and (D) after adding N-C/CeO2 + Br + H2O2. Note: the concentration of P. aeruginosa is 107 cfu·mL−1.
Ijms 24 02445 g005
Table 1. The Michaelis–Menten constant (Km) and maximal reaction rate (vmax) of N-C/CeO2 composites.
Table 1. The Michaelis–Menten constant (Km) and maximal reaction rate (vmax) of N-C/CeO2 composites.
SubstratesKm (μ·mol·L−1)vmax (μmol·L−1·min−1)
H2O2 (mmol·L−1)0.2460.669
NH4Br (mmol·L−1)1731.11
PR (mmol·L−1)0.01302.48
Table 2. Comparison of the apparent Michaelis–Menten constant and maximal reaction rate between the N-C/CeO2 composite and V-BPO.
Table 2. Comparison of the apparent Michaelis–Menten constant and maximal reaction rate between the N-C/CeO2 composite and V-BPO.
MaterialsSubstratesKm (μmol·L−1)
N-C/CeO2 compositesH2O20.246
Br173
vanadium bromoperoxidase (V-BPO)H2O222.0
Br18.1
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, N.; Zhai, X.; Guan, F.; Zhang, R.; Hou, B.; Duan, J. N-Doped Carbon/CeO2 Composite as a Biomimetic Catalyst for Antibacterial Application. Int. J. Mol. Sci. 2023, 24, 2445. https://doi.org/10.3390/ijms24032445

AMA Style

Wang N, Zhai X, Guan F, Zhang R, Hou B, Duan J. N-Doped Carbon/CeO2 Composite as a Biomimetic Catalyst for Antibacterial Application. International Journal of Molecular Sciences. 2023; 24(3):2445. https://doi.org/10.3390/ijms24032445

Chicago/Turabian Style

Wang, Nan, Xiaofan Zhai, Fang Guan, Ruiyong Zhang, Baorong Hou, and Jizhou Duan. 2023. "N-Doped Carbon/CeO2 Composite as a Biomimetic Catalyst for Antibacterial Application" International Journal of Molecular Sciences 24, no. 3: 2445. https://doi.org/10.3390/ijms24032445

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