Abstract
Electrochemical oxygen reduction reaction (ORR) is an attractive and alternative route for the on‐site production of hydrogen peroxide (H2O2). The electrochemical synthesis of H2O2 in neutral electrolyte is in early studying stage and promising in ocean‐energy application. Herein, N‐doped carbon materials (N‐Cx) with different N types are prepared through the pyrolysis of zeolitic imidazolate frameworks. The N‐Cx catalysts, especially N‐C800, exhibit an attracting 2e− ORR catalytic activity, corresponding to a high H2O2 selectivity (≈95%) and preferable stability in 0.5 m NaCl solution. Additionally, the N‐C800 possesses an attractive H2O2 production amount up to 631.2 mmol g−1 h−1 and high Faraday efficiency (79.8%) in H‐type cell. The remarkable 2e− ORR electrocatalytic performance of N‐Cx catalysts is associated with the N species and N content in the materials. Density functional theory calculations suggest carbon atoms adjacent to graphitic N are the main catalytic sites and exhibit a smaller activation energy, which are more responsible than those in pyridinic N and pyrrolic N doped carbon materials. Furthermore, the N‐C800 catalyst demonstrates an effective antibacterial performance for marine bacteria in simulated seawater. This work provides a new insight for electro‐generation of H2O2 in neutral electrolyte and triggers a great promise in ocean‐energy application.
Keywords: antibacterial, hydrogen peroxide, N‐doped carbon, oxygen reduction reaction, simulated seawater
The synthesized graphite nitrogen doped carbon material delivers an attracting 2e− ORR catalytic activity in 0.5 m NaCl simulated seawater solution, corresponding to a high H2O2 selectivity (≈95%), preferable stability and H2O2 production amount up to 631.2 mmol g−1 h−1. Additionally, the prepared catalyst demonstrates an effective antibacterial performance for marine bacteria in simulated seawater.

1. Introduction
H2O2 as a versatile clean oxidative chemical, is widely used for the degradation of organic dyes[ 1 ] and organic drugs,[ 2 ] water treatment,[ 3 ] bacteria killing/ disinfection,[ 4 ] and energy storage,[ 5 ] due to the only byproduct of water without hazardous residues. At present, H2O2 is commonly synthesized through an energy‐intensive anthraquinone oxidation‐reduction in industrial scale.[ 6 ] However, this method requires complex infrastructure and produces a substantial volume of organic byproduct wastes. Therefore, it is greatly important to develop highly efficient techniques for H2O2 synthesis. The direct synthesis of H2O2 through H2 and O2 is a straightforward and atom‐economic method.[ 7 ] Nevertheless, this method causes the potential explosion hazard of H2/O2 mixtures.[ 8 ] Electrocatalytic ORR is an attractive and alternative route for on‐site production of H2O2, which processes the advantages of low‐cost and convenience for in‐situ H2O2 generation at various application situations.
Substantial efforts in recent years have shown that electrocatalytic ORR is desirable and applicable for selective generation of H2O2.[ 9 ] However, the ORR always occurs according to four‐electron process to generate H2O, which is a strong completive reaction to generate H2O2. Developing highly efficient electro‐catalyst is crucial to refining the H2O2 generation via 2e− ORR. The previous reports indicate that noble metal‐based catalysts show high selectivity toward H2O2 under strongly acidic conditions, and transition metal‐based catalysts exhibit high 2e− catalytic selectivity under strongly acidic or basic solutions.[ 10 ] However, those noble/ transition metal catalysts always present low catalytic selectivity in neutral electrolytes.[ 11 , 12 ] Non‐metallic carbon‐based materials deliver good H2O2 selectivity in neutral electrolytes. Meanwhile, they process the advantages of high conductivity/stability, high mass transfer porosity, and low cost.[ 13 ] Therefore, it is a promising 2e− ORR catalyst, especially for large‐scale application. Heteroatom‐doping, especially N‐doping, is a useful strategy to improve the 2e− ORR catalytic performance of non‐metallic carbon‐based catalysts by refining the catalytic active sites.[ 14 , 15 ] Some previous reports have shown that nitrogen‐doped carbons are in favor of 2e− ORR catalytic activity in neutral phosphate buffer solutions.[ 16 , 17 ] However, the 2e− ORR catalytic active sites of nitrogen‐doped carbons, such as pyridinic N, pyrrolic N and graphite N, are still controversial and remain a matter of active debates.[ 18 , 19 ] Accurately refining the N‐species in N‐doped carbon materials and clarifying the catalytic mechanism are crucial to guide and optimize the synthesis of efficient 2e− ORR catalysts.
Noteworthily, most researchers focus on the electrochemical generation of H2O2 in strongly acidic or alkaline electrolyte, while the studies in neutral electrolyte are paid little attention.[ 20 ] H2O2 synthesis in neutral solution is actually very useful and flexible for practical application of H2O2, as it can avoid the influence of pH.[ 21 ] Seawater as an earth‐abundant resource stores pretty rich energy, and is investigated as an attractive electrolyte in energy‐field recently.[ 22 ] It is also a promising neutral electrolyte for in site electrocatalytic production of H2O2. Meanwhile, the as‐synthesized H2O2 displays great application potential in marine biofouling field for sterilization. However, it is still in the early stages that the research on the use of seawater as electrolyte in energy system. Generally, 0.5 m NaCl is used as simulated seawater in the lab‐scale research.[ 23 ] Whereas, the 2e− ORR catalytic performance and catalytic mechanism of catalysts in simulated seawater electrolyte are still unsatisfactory and ambiguous. Thus, it is urgent to explore suitable catalysts and recognize the 2e− ORR catalytic mechanism in neutral simulated seawater electrolyte.
Herein, N‐doped carbon materials are prepared using the facile pyrolysis method through calcinating Zeolitic imidazolate frameworks (ZIF‐8). The N species in the nitrogen‐doped carbon materials can be well refined by ranging the calcination temperature from 600 to 1000 °C. The structure, the species and content of N doping of resulting product are characterized by different analysis method. N‐doped carbon materials catalysts are investigated the effect of the different N doping on 2e− ORR activity, selectivity, and stability in the neutral electrolyte. The result shows that N‐doped carbon exhibit remarkable electrochemical performance, which is attributed to the species and content of nitrogen‐doped carbon structure. In addition, density functional theory (DFT) was calculated, and the results showed that the species of nitrogen‐doped carbon structure are active sites, and conducive to 2e− ORR. The amount of H2O2 generation were measured, and used to study the effect on the marine typical bacteria (Pseudomonas aeruginosa) in the simulated seawater. This study will provide an efficient alternative for water disinfection and other important application in the future.
2. Results and Discussion
2.1. Characterization
ZIF‐8 as a class of metal‐organic framework contains 34% of N‐containing 2‐methylimidazole linker,[ 24 ] which is a good template and reactive precursor for nitrogen‐doped carbon materials. As shown in Figure 1a, the N‐doped carbon materials were synthesized by a two‐step strategy, including the fabrication of high N content ZIF‐8 precursor and a further thermal carbonization and acid treatment activation process. The N‐C materials with different nitrogen‐doped types were obtained by regulating the annealing temperature of ZIF‐8 precursor. As shown in Figure 1b, the synthesized ZIF‐8 displays a regular and rhomb dodecahedral morphology. It is also observed in the corresponding TEM image (Figure 1b1 ). The N‐Cx materials exhibit the different degrees of shrinkage with the increasing pyrolysis temperature in Figure 1c–g. Generally, the zinc atoms in the ZIF‐8 frameworks present the different degrees of volatilization at various pyrolysis temperatures,[ 25 ] which can cause the collapse of organic molecular skeleton. However, they still remain their basic polyhedron structure below the pyrolysis temperature of 900 °C, which can be observed by the corresponding TEM images (Figure 1c1–g1 ). Interestingly, the new structure was formed by the single rhomb dodecahedral accumulation, overlay and bond at the high temperature (1000 °C). It indicates that the framework structure is deformed and reorganized after the full volatilization of zinc atoms at this temperature. The above results demonstrate that the morphology/ micro‐structure of the N‐Cx materials is closely associated with the polymerization temperature and the volatilization of zinc atoms and is not affected by acid treatment.
Figure 1.

a) Synthesis schematic of nitrogen‐doped carbon materials, SEM and TEM images: b) and b1) ZIF‐8, c–g) and c1–g1) N doped carbon materials from annealing ZIF‐8 under 600, 700, 800, 900, and 1000 °C.
The high‐resolution XPS spectra of N‐doped carbon materials were further conducted to identify the binding mode of nitrogen and carbon atoms in the materials. The XPS survey spectra deliver typical C 1s, N 1s, O 1s, and Zn 2p signals (Figure S1, Supporting Information). As shown in Figure 2a, the N 1s XPS spectrum can be deconvoluted into three peaks at about 398.4, 399.6, and 400.9 eV, which are assigned to pyridinic N, pyrrolic N and graphitic N, respectively.[ 26 ] The accurate amounts of different N‐species are summarized in Figure 2b. The relative amount of pyrrolic N decreases whereas the relative amount of graphitic N increases with the increasing of annealing temperature, which is attributed to the stability of graphitic N being superior to pyrrolic N at the high annealing temperature.[ 26 ] With the difference, the relative amount of pyridinic N increases first and then decreases with the increasing annealing temperature. It suggests that pyridinic N is easier to form than graphitic N at the low annealing temperature (<800 °C), while graphitic N is more stable at the higher annealing temperature (>800 °C).[ 27 ] The high‐resolution C 1s XPS spectra show the presence of C═C (284.6 eV), C─C (285.4 eV), C─N (286.4 eV), and O─C═O (289.1 eV) functional groups (Figure 2c).[ 28 ] The O1s spectra of the samples can be assigned to C─O (532 eV) and C═O (533 eV) bond, respectively (Figure S2, Supporting Information). With the increasing of pyrolysis temperature, the ratio of C─N and C─O bond content decrease and that of C─C bond raise, due to the destruction of the ZIF‐8 framework at high temperature. Figure 2d presents the change of elemental contents (N, C, and H) in different N‐doped carbon materials with the increasing annealing temperature. The accurate contents of N, C, and H are shown in Table S1 (Supporting Information). It is obvious that the H content remained almost constant (near 2 wt.%) in all different N‐doped carbon materials. However, the C content significantly increases due to the evaporating of the zinc with the temperature increasing. The N content just slightly decreases from 16.27% of N‐C600 to 14.81% of N‐C800. As the annealing temperature continues to increase, the N content presents an obvious decrease. Because zinc is stable in the formation of “Zn‐N”, the volatilization of Zn atoms at high temperature also takes away part of N.[ 29 ] The above result is consistent with the results of XPS spectra N 1s (Figure 2a) and C1s (Figure 2c). Additionally, the Zn contents in all five samples are near 0.5–1.3 at.% (Table S2, Supporting Information) according to the quantitative XPS analysis results. The ICP‐OES results indicate that Zn content in all the five samples deliver comparable of near 2.1–2.8 wt.% (Table S3, Supporting Information). These results show that Zn content in all the five samples is low and comparable.
Figure 2.

The structure and content characterization of the nitrogen‐doped carbon materials. a) High‐resolution XPS spectra N 1s. b) The relative amount of the different nitrogen species from the N 1s. c) High‐resolution XPS spectra C 1s. d) The content of elements. e) Normalized and baseline corrected Raman spectra. f) The NMR of the N‐Cx materials.
To further explore the structure of nitrogen‐doped carbon catalysts, the Raman spectra were measured using a 532 nm solid‐state laser as an excitation source. The D‐band is associated with the defects and G bands represent to C═C stretching vibrations of carbon layers related to the sp2 hybridizations.[ 30 ] As shown in Figure 2e, the D peaks of the samples consist of two defect peaks at 1325 cm−1 (D1) and 1425 cm−1 (D3), which are attributed to defective edge carbon and amorphous sp2 carbon defects.[ 31 ] The 2D at around 2880 cm−1 is also defined, which attributes to the presence of structural defects. With the calcined temperature increasing, there is a slight change between D1 and D3 signals due to the conversion of defective edge carbon and amorphous sp2 carbon defects.[ 32 ] Furthermore, the ID1+D3/IG ratio (Table S4, Supporting Information) of the N‐Cx catalysts display a decreasing tendency, which is associated with the increased graphitic degree of the N‐Cx catalysts at high annealing temperature.
The variation of N‐Cx catalysts in Raman spectra are highly consistent with the results in elemental contents (Figure 2d), XPS (Figure 2c) and TEM (Figure 1) measurements. In addition, the electrical conductivity of N‐Cx significantly increases with the improved thermal‐pyrolysis temperature (Table S5, Supporting Information). The enhanced electrical conductivity is also associated with their improved graphitization degree, which is well consistent with the XPS (Figure 2a) and Raman results (Figure 2e).
Solid‐state cross‐polarization/ magic angle spinning nuclear magnetic resonance (CP/MAS NMR) 13C spectra were then recorded to determine the chemical structure of the prepared samples. As displayed in Figure 2f, all the samples calcined at different temperature deliver a broad peak with the chemical shift of 100–150 ppm, which can be assigned to the sp2‐hybridized carbon.[ 33 ] Especially, the spectra featured strong peaks at ≈ 130 and 155 ppm, which correspond to the aromatic carbon and graphite C═N groups.[ 34 ] For the samples calcined below 800 °C, two distinct C signals are observed at around 125 and 170 ppm in the 13C NMR spectra, which are attributed to C≡N and carbonyl groups (C═O/COOH). Meanwhile, the intensity of the peaks at 125 and 170 ppm decreases as the carbonation temperature increases, implying a higher carbonization temperature can cause further aromatization and N content reduction. The above result is well consistent with the XPS result (Figure 2a) and element content analysis (Figure 2c). The BET measurements were conducted to investigate the specific surface area and pore size distribution of the materials (Figure S3, Supporting Information). The N‐Cx samples mainly display mesoporous structure and the specific surface area of the N‐C600, N‐C700, N‐C800, N‐C900 and N‐C1000 is 43, 56, 434, 1088 and 438.2 m2 g−1, respectively.
2.2. Electrochemical Property
To reveal the electrochemical behavior of N‐Cx catalysts, CV curves were measured in O2 or N2‐saturated 0.5 m NaCl solution. All the potentials in this work were referenced to the reversible hydrogen electrode (RHE). The CV curves indicate an obvious oxygen reduction peak in O2‐saturated electrolyte compared with in N2‐saturated electrolyte (Figure S4, Supporting Information). The ORR catalytic activity of N‐Cx catalysts were measured using the RDE in an O2‐saturated 0.5 m NaCl solution (pH = 6.88). Figure 3a shows the ORR polarization curves of the N‐doped carbon materials with the annealing temperature from 600 to 1000 °C. Similar with the results of CV, the LSV curves exhibit an obvious oxygen reduction peak in the O2‐saturated 0.5 m NaCl solution. Apparently, the current densities of N‐C600 and N‐C700 are significantly lower than those of other catalysts, which maybe is related to their low content of graphitic N and the poor conductivity. Some reports indicate that graphitic N is responsible for the ORR activity of N‐doped carbons.[ 35 ] The inferior LSV curves and catalytic performance of the N‐C600, N‐C700 and N‐C1000 are associated with their poor catalytic sites, electrical conductivity, and specific surface area.[ 36 , 37 ] Figure 3b displays the values of the onset potential for various catalysts, which are selected at a current density of 0.1 mA cm−2 as the previous reports.[ 38 ] Notably, the N‐C800 and N‐C900 exhibit an outstanding and almost identical onset potential (E onset) of about 0.61 V (vs RHE). Their values are much more positive than N‐C600 (0.381 V) and N‐C700 (0.412 V), even much more positive than N‐C1000 (0.506 V). The ORR catalytic kinetics of the catalysts are further estimated by the Tafel plots (Figure 3c). The N‐C800 displays a smaller Tafel slope (79 mV decade−1) than N‐C900 (89 mV decade−1), N‐C1000 (95 mV decade−1), N‐C700 (114 mV decade−1) and N‐C600 (116 mV decade−1), indicating a superior catalytic kinetic of the N‐C800, which is consistent with the results of onset potentials. Furthermore, to establish the intrinsic activity of the N‐Cx catalysts, electrochemical double‐layer capacitance (Cdl) was investigated by CV at different sweep rates (Figure S5, Supporting Information). The Cdl of N‐Cx catalysts were calculated and shown in Figure 3d. The results demonstrate N‐C800 possesses a significantly larger ECSA than other N‐Cx catalysts, which greatly contributes to its ORR catalytic performance. The above results indicate that N‐C800 delivers the best 2e− ORR catalytic activity among various N‐Cx catalysts.
Figure 3.

ORR performance of N‐doped carbon materials. a) ORR polarization curves in O2‐saturated 0.5 m NaCl (rotation rate: 1600 rpm, sweep rate: 10 mV s−1), b) Bar plots of E onset, c) Tafel plots derived from panel (a), and d) Current density as a function of scan rate. e) linear sweep voltammetry performed by a RRDE technique where the ring current is collected on the Pt ring at a constant potential of 1.5 VRHE and f) calculated n and H2O2 selectivity (%), as a function of electrode potential.
The catalytic selectivity of N‐Cx catalysts is the key electrocatalytic performance for 2e− ORR. It is further investigated using RRDE in an O2‐saturated 0.5 m NaCl solution. Figure 3e shows the RRDE curves of the N‐Cx catalysts, in which the ORR current and the simultaneous H2O2 detection current are obtained on the disk electrode and Pt ring electrode, respectively. The N‐C800 exhibits the highest ring current among all the N‐Cx catalysts. It is attributed to its abundant N content and graphite‐N type in N‐C800 catalyst. Meanwhile, the N‐C800 achieves almost the highest ORR disk current (comparable with that of N‐C900) among those N‐Cx catalysts. Therefore, the N‐C800 presents a superior performance, not only in catalytic activity but also in catalytic selectivity for H2O2 production. Figure 3f shows the catalytic selectivity for H2O2 production with the number of transferred electrons (n) and H2O2 selectivity (H2O2%). For all the N‐Cx materials, both H2O2 selectivity and n variation tendency depend on the applied potential.[ 10 ] The corresponding n value is given near two, which is consistent with the H2O2 selectivity, suggesting N‐Cx catalysts mainly follow 2e− oxygen reduction pathway. The N‐C800 exhibits a high H2O2 selectivity at the whole potential range. Especially, the H2O2 selectivity of N‐C800 is above 90% at a high potential range (>0.446 V), which is better than other catalysts. The N‐C800 also exhibits a comparable H2O2 selectivity at the low potential (<0.446 V). Additionally, the N‐C800 sample was immersed into aqua regia solution (HNO3: HCl = 1:3) for 12 h to fully remove the Zn species, which was labeled as N‐C800 (0 wt.% Zn). Compared to N‐C800 (0 wt.% Zn), the N‐C800 delivers an almost same onset potential, a slightly higher disk current density and lower 2e− ORR selectivity (Figures S6 and S7, Supporting Information). Both the N‐C800 and N‐C800 (0 wt.% Zn) deliver an attracting and similar 2e− ORR catalytic performance, implying the catalytic effect of Zn concentration in this N‐Cx catalyst is weak. These results confirm that the N types of the N‐Cx are crucial to the 2e− ORR selectivity.
The stabilities of N‐Cx materials were further studied in O2‐saturated 0.5 m NaCl under their optimal applied potential according to H2O2 selectivity. As shown in Figure 4a, the current of the disk/ ring electrode and H2O2 selectivity of the N‐C800 remain pretty stable for 10 h without an obvious decay. It indicates that the N‐C800 catalyst keeps a high structural and 2e− ORR catalytic stability during the electrochemical process. Meanwhile, the other catalysts also deliver comparable stabilities at their optimal applied potentials (Figure S8, Supporting Information). The results indicate that N‐C800 catalyst presents good catalytic activity, selectivity and stability at the same time. In addition, the electrocatalytic activity and H2O2 selectivity of N‐Cx were tested in 0.1 m H2SO4 and 0.1 m KOH electrolyte solution with different pH values. Figure 4b shows the 2e− ORR catalytic performance of N‐C800 in 0.1 m H2SO4, 0.5 m NaCl and 0.1 m KOH electrolyte solution, separately. The results show that N‐C800 presents high disk currents in above three electrolyte solutions, illustrating a good catalytic activity. Meanwhile, the ring current tested in 0.5 m NaCl is the highest, indicating the best H2O2 selectivity (Figure S9, Supporting Information). The other types of N‐Cx catalyst also present the best H2O2 selectivity in 0.5 m NaCl solution (Figures S10–S13, Supporting Information). In a neutral electrolyte, the 2e− ORR catalytic performance of prepared catalyst in this work is compared with those in previous reports (Figure 4c; Table S6, Supporting Information).[ 9 , 13 , 16 , 22 , 23 , 26 , 39 ] It is clear that the N‐C800 catalyst exhibits a better/ attracting onset potential and H2O2 selectivity, which are even higher than those of some metal‐based catalysts (Co─N─C).
Figure 4.

a) Stability tests of N‐C800 at 0.51 V versus RHE, b) LSV of N‐C800 recorded in 0.1 m H2SO4, 0.5 m NaCl and 0.1 m KOH at 1600 rpm, with the ring electrode at a constant potential of 1.5 VRHE, c) Comparison of the onset potential and H2O2 selectivity in the neutral solutions, d) H‐Cell electrolyzer for H2O2 production in 0.5 m NaCl solutions, e) H2O2 production amount normalized to catalyst loading amount over the reactive time at applied potentials and f) The Faraday efficiency of H2O2 for N‐Cx catalysts under the flow of O2.
The H2O2 yield is a crucial parameter to evaluate the catalytic performance of the materials especially in practical applications. Herein, an H‐Cell electrolyzer is used for producing H2O2 according to ORR process in 0.5 m NaCl solutions (Figure 4d). The generation amount of H2O2 can be detected through a photometric method.[ 13 ] Based on the standard curve of H2O2 concentration measurement by cerium method (Figure S14, Supporting Information), Figure 4e displays the accumulated amounts of H2O2 for various catalysts in O2‐saturated 0.5 m NaCl, which are normalized by catalyst loading amount over the reaction time. Remarkably, for the H2O2 production of ≈410–630 mmol g−1 h−1, the catalytic rate of various catalysts delivers a list of N‐C800 > N‐C1000 > N‐C900 > N‐C700 > N‐C600, which is better than that of the reported nitrogen‐doped carbon catalyst.[ 39 ] The results show that N‐C800 possesses the highest H2O2 production amount up to 631.2 mmol g−1 h−1. The H2O2 production amount of N‐C900 is lower than that of N‐C1000, which is attributed to the consecutive decomposition reaction of H2O2. The peroxide reduction reaction (PRR) result displays the N‐C900 catalyst a high cathodic current, indicating a H2O2 reduction tendency (Figure S15, Supporting Information). In addition, the Faraday efficiency of H2O2 is another important parameter for evaluating the performance of the catalyst. It was calculated according to the real amount of produced H2O2 and H2O2 selectivity of the catalysts in their investigated applied potential (Figure 4f). The result indicates that N‐Cx catalysts exhibit Faraday efficiency ranging from 31.2% to 79.8% H2O2. Especially, the N‐C800 catalyst stands the top level Faraday efficiency (79.8%) for H2O2 production.
2.3. Theoretical Calculations
To further confirm the catalytic active sites of various N species in N‐Cx catalysts, the DFT calculations were carried out to evaluate the free energies of the adsorbed intermediate of the catalytic reaction. First, the geometrical structure of N‐graphene (pyridinic, pyrrolic or graphitic N) were optimized using vibration analysis in Figure 5a. The optimized geometry structure is nearly perfect, indicating that these optimized structures are stable. The 2‐electron ORR mechanism is generally considered to proceed through the following steps (1 and 2):[ 40 ]
| (1) |
| (2) |
where asterisks (*) indicate unoccupied active sites. *OOH denotes the important and single intermediate for the reaction. Among the adsorption energies between catalysts and *OOH intermediate determines the reaction product, which is recognized as the rate‐determining step.[ 41 ] The DFT calculations work well in describing adsorption energies of intermediates on the different N‐graphene surface. As shown in Figure 5b, the geometrical structure of carbon adjacent to graphitic N obviously distorted after adsorbing O2 reactant and *OOH intermediate. Thus, the carbon adjacent to graphitic N is the main catalytic active site to promote H2O2 generation. The schematic illustration of pyridinic N and pyrrolic N for 2e‐ ORR were also given in Figures S16 and S17 (Supporting Information). In addition, the computational hydrogen electrode model (CHE) was used in the catalytic process.[ 42 ] Figure 5c shows the DFT calculated free diagram of the various N‐graphene (pyridinic‐N, pyrrolic‐N and graphitic‐N) for the 2e− ORR process using the CHE model. The free energy of various N‐graphene to catalyze the H2O2 generation delivers the order of pyridinic‐N > pyrrolic‐N > graphitic‐N. According to the DFT calculation, the graphitic‐N doped graphene displays the best 2e− ORR catalytic performance, which is well consistent with the experimental results.
Figure 5.

a) Schematic illustration of the three nitrogen doping configurations, b) The schematic of 2e− ORR catalytic active sites from carbon adjacent to graphitic N and c) Free energy diagrams of 2e− ORR on the different typical nitrogen doped carbon materials.
2.4. Antibacterial Properties
Utilizing seawater as electrolyte can reduce the cost and broaden the on‐site production of H2O2 along the coast in energy‐related field. Meanwhile, H2O2 as an oxidizer has been used for bacteria killing.[ 43 ] Exploring a kind of low‐cost catalyst for efficient H2O2 generation is very promising in the practical applications of the marine field, especially for antibacterial and antifouling of the offshore engineering facilities. In the current lab‐scale research, we conduct the disinfection properties of the N‐C material in 0.5 m NaCl solution to mimic the practical bacterial killing applications. The schematic of the electro‐chemical 2e− ORR progress for bacteria killing in simulated seawater is shown in Figure 6a. The sterilization effect of the prepared electro‐catalyst is further evaluated using a H‐cell electrolytic cell and Pseudomonas aeruginosa (P. Aeruginosa) marine bacteria. The obtained P. Aeruginosa concentration is ≈108 c.f.u. mL−1 by plate counting and is operated with a suitable dilution magnification. The electrolyte with different concentrations of H2O2 is picked up at a series of time (0, 30, 60, 120, 180, and 300 min) during i–t measurement (0.51 V vs RHE). The picked electrolyte is further diluted before sterilization, and the images of agar plates with cultured bacteria colony after sterilization are displayed in Figure 6c. The number of colonies gradually decreases with the increasing of electro‐catalytic time (as more H2O2 is generated), and the number of colonies decreases to be almost negligible after 300 min. It is obvious that N‐C800 demonstrates a promising disinfection efficiency for P. Aeruginosa according to the calculated killing rate plotted (Figure 6b). These above results indicate that the synthesized N‐C electrocatalyst presents a good sterilization effect in the simulated seawater. This could provide an experimental basis for practical marine antibacterial and antifouling applications in the future.
Figure 6.

a) schematic of electrochemical synthesis of H2O2 for antibacterial. b) The disinfection efficiency as a function of treatment time. c) Photos of cultured plates with spread droplets taken from different time slots during the electrolysis. Dilution factor is labeled of each image.
3. Conclusion
In summary, the N‐Cx materials with different types of doped‐nitrogen (graphitic N, pyridinic N, and pyrrolic N) are synthesized by varying the pyrolysis temperatures of ZIF‐8. The N‐Cx catalysts exhibit a distinguishing 2e− ORR performance in 0.5 m NaCl solution due to their different N‐doping content and types. Especially, N‐C800 exhibits the best catalytic activity with an onset potential of ≈0.6 V (vs RHE), catalytic selectivity with ≈95%, and high catalytic stability in 0.5 m NaCl solution. This correlates with the high ratio and content of graphitic N in N‐C800 catalysts. N‐C800 catalyst possesses a high H2O2 production amount up to 631.2 mmol g−1 h−1, and stands the top level Faraday efficiency (79.8%). The DFT calculation shows that the graphitic N displays higher 2e− ORR catalytic activity than pyridinic N and pyrrolic N, and the carbon atoms adjacent to graphitic N act as the active sites to interact with reaction intermediates. The results strongly suggest that graphitic N is more favorable to the 2e− ORR performance. In addition, the H2O2 generated by N‐C800 electro‐catalyst in simulated seawater displays a favorable sterilization effect, which is meaningful for in‐site bacteria killing in marine field. This work may provide a promising possibility for converting the marine resource to energy storage and microbial fouling protection.
4. Experimental Section
Synthesis ZIF‐8
The ZIF‐8 framework was prepared by reacting 2‐Methylimidazole with Zn(NO3)2 at room temperature according to the previous report.[ 44 ] Typically, 1.752 g of 2‐methylimidazole was dissolved in 200 mL methanol to form solution A. 7.93 g of Zn(NO3)2∙6H2O was dissolved in 200 mL of methanol to form solution B. Solution B was poured in solution A all at once. The mixture was kept under stirring for 24 h. The precipitate was separated by filtration and washed with methanol for three times. After drying under vacuum at 60 °C, the white powder precursors were obtained.
Synthesis of N‐C Materials
The synthesized precursors were pyrolyzed at a series of temperatures (600, 700, 800, 900, and 1000 °C) for 2 h with the ramping rate of 2 °C min−1 under an argon atmosphere. The calcined samples were washed with moderate hydrochloric acid solution and labeled as N‐Cx, where x was defined as the calcinated temperature.
Characterization
The morphologies of ZIF‐8 and N‐C were characterized by scanning electron microscopy (SEM) and transmission electron microscopy (TEM) on HITCH microscope (Regulus and HT7700, Japan). The electronic structure and chemical bonding of materials were determined using the X‐ray photoelectron spectroscopy (XPS, K‐alpha 250Xi, England). The elemental contents of C, H, N, and S in the materials were conducted on an elemental analysis instrument (Elementar Unicube, EA 3000, Germany) and the Raman spectra were collected on Renishaw MZ20‐FC Raman microscope. Metal content were evaluted by the inductively coupled plasma‐optical emission spectrometer (ICP‐OES, iCAP RQ). Solid‐state nuclear magnetic resonance (NMR) of 13C was performed on the Bruker AVANCE III HD NMR 400MWB spectrometer, equipped with a 4.0 mm MASDVT BL4.0 X/Y/F‐H resonance probe head. Cross‐polarization/magic angle spinning (CP/MAS) sequence was performed to enhance the 13C NMR signal response. The synthesized powder samples were packed inside Zirconia MAS rotor with a diameter of 4.0 mm and a vespel cap. The 13C Hahn‐echo MAS spectra were acquired with a 2 s recycle delay, a 13C excitation (90o) pulse length of 2 s and 20 kHz MAS. The electrical conductivity of the materials was determined using two‐probe measurement method (ROOKO FT‐300L) at series of testing pressure (12, 15, or 20 MPa). The specific surface area of materials was measured with Brunauer–Emmett–Teller analyzer (BET, ASAP2460). The absorbance of the solutions was measured by UV–vis spectroscopy (HITCH 3900, Japan).
Electrochemical Measurements
Electrochemical tests were performed using a CHI electrochemical workstation (CHI 760E) coupled with a rotating‐ring disc electrode (RRDE, Pine) in a three‐electrode cell. A graphite rod and a saturated calomel electrode (SCE) were used as the counter electrode and reference electrode, respectively. The reference electrode was calibrated to a reversible hydrogen electrode (RHE) before each measurement. All potentials measured against SCE were converted to the RHE scale using E (vs RHE) = E (vs SCE) + Eɵ (SCE)+ 0.059*pH, where Eɵ (SCE) value was the calibrated value, pH value of electrolytes was determined by the pH meter (lightning magnetism, PHSJ‐3F). A RRDE assembly (258051, Pine Instruments) consisting of a glass carbon rotation disk electrode (0.196 cm2) and a Pt ring (0.2475 cm2) was used, with the collection efficiency of 37%. The electrocatalysts inks with the mass concentration of 5 mg mL−1 catalysts were prepared by dispersing a certain amount of catalyst in an isopropanol solution with Nafion (5%). 6 µL of each catalyst ink was pipetted on a precleaned glass carbon disk electrode and dried at room temperature to yield a uniform thin‐film electrode.
The ORR catalytic activity was measured by linear sweep voltammetry (LSV) with a rotation rate of 1600 rpm in an O2‐saturated electrolyte solution (the solution resistance was compensated). Cyclic voltammetry (CV) with N2‐saturated 0.5 m NaCl solution were measured at the different scan rates to estimate the electrochemical active surface area (ECSA).[ 45 ] ECSA was estimated according to electrochemical double‐layer capacitance (Cdl), based on the positive proportional relationship between ECSA and Cdl (ECSA = Cdl /Cs),[ 46 ] where Cs is the specific capacitance of carbon materials, and its real value is unknown. Cdl was determined by conducting CV with a 3.5% NaCl electrolyte solution with a potential range of non‐faraday reaction at increasing scan rates of 5, 10, 20, 40, 50, 60, 80, 100, 150, 200, 250, 300 mV s−1. The value of Cdl was obtained from the slope of a derived plot of average current density versus the scan rate. The average current density is equal to (Ia+Ic)/2, where Ia and Ic are the anodic current and the cathodic current, respectively, which can be read from the CV curves.[ 47 ]
The catalytic selectivity was measured by an RRDE in an O2‐saturated electrolyte solution at a rotation rate of 1600 rpm. The Pt‐ring electrode was polarized at 1.2 VRHE to further reduce the as‐formed H2O2 from the disk electrode. The electron transfer number and selectivity for the H2O2 yield were calculated from the disk current and ring current according to the corresponding formula.[ 44 ] Among that, the current collection efficiency of the Pt‐ring was 0.37.
The stability of catalysts was determined by chronoamperometry with a rotation rate of 1600 rpm in an O2‐saturated 0.5 m NaCl solution at room temperature. Meanwhile, a potential of 1.2 VRHE was used to the Pt‐ring electrode during the entire testing process. It is noted that the pH value remained almost unchanged before and after the continuous electrolysis.
H2O2 production in 0.5 m NaCl was carried out in the H‐cell electrolyzer. The cathode anode chambers were separated with Nafion 117 membrane. 6 µL of each catalyst ink was loaded on working electrode. A chronoamperometry measurement was performed to H2O2 production. The H2O2 concentration was quantified by cerium sulfate titration method.[ 13 ] A yellow solution of Ce4+ would be reduced by H2O2 to colorless Ce3+ (2Ce4+ + H2O2 — 2Ce3+ + 3H+ + O2). Based on this mechanism, the concentration of Ce4+ before and after the reaction can be measured by UV–vis spectroscopy. The characteristic absorption peaks appear at wavelength of 316 nm. The 1 mm Ce(SO4)2 solution was prepared by previous reported.[ 13 ] The calibration curve of H2O2 were determined by measuring the mixture of known concentration H2O2 and Ce(SO4)2 solution. Based on the linear relationship between the signal absorption value and know concentration H2O2, the H2O2 concentrations of the samples could be obtained.
Computational Details
The density functional theory (DFT) was conducted using the Guassian09 program package to simulate the mechanistic process of O2 reduction reaction on various N‐graphene substrates. No geometric constraints were assumed in geometry optimization. The nonlocal correlation functional of Lee, Yang, and Parr3 (B3LYP) with the 6–31++G** basis set was used for C, N, H and O atom. The pristine graphene, with 7 hexagonal rings with delocalized π electron, was used as the baseline model. With the same configuration as pristine graphene, N‐graphene with three different N‐graphene models were constructed, including pyridinic N, pyrrolic N, and graphitic N. The hydrogen atoms terminate the carbon atoms at the edge of the graphene. The relative energies of the molecules presented in this study were zero‐point‐energy (ZPE) obtained from frequency calculations at the same level of optimization.
O2 reduction reaction was simulated, beginning with the adsorption of an O2 molecule onto an N‐graphene surface. In this step, we placed O‐O near the N‐graphene plane, then the structures of O─O adsorption on N‐graphene was optimized structures by the structural optimization calculations. The adsorption energy was derived from frequency analysis after the initial structure optimization. At the same process, the structure of OOH and HOOH adsorption on N‐graphene were optimized, respectively. Finally, the relative energy between the reactants and products was calculated. The other types of N‐graphene model were optimized and calculated at the same method.
Antibacterial Tests
A typical marine bacterial Pseudomonas aeruginosa (P.aeruginosa) strain was provided by our laboratory. P. Aeruginosa was cultured in 2216E liquid medium at 37 °C for 24 h with 130 rpm shaking. 1 mL of the obtained green culture liquid was centrifugated at 4000 rpm to separate the bacterial biomass from the medium. Then, the bacterial cells were diluted and suspended in 0.1 m phosphate buffer saline (PBS) for plate counting. Others were added into 0.5 m NaCl. The electrochemical antibacterial measurements were run at room temperature in H‐type glass cell separated by Nafion 117 membranes. 50 mL of the prepared P. Aeruginosa (≈108 c.f.u. mL−1) in 0.5 m NaCl were added into the cathodic chamber, and 50 mL of bacteria‐free 0.5 m NaCl were added into the anode chamber under sterile conditions. Glass carbon electrode (Ф = 5.0 mm) with 3 µg N‐C800 catalysts was served as the working electrode. A chronoamperometry curve measurement at the same voltage value (0.51 VRHE) was carried out to verify the antibacterial performance. The bacteria killing rates were measured using standard spreading plating techniques. 200 µL electrolyte with bacteria were taken at different time interval during the electrolysis process. Samples were then serially diluted and plated on 2216E agar plates in triplicate. Plates were incubated at 37 °C for 24 h or more to make the larger colony and easier to observe and count. The photos of cultured plates were taken with the fully automatic colony counter (icount 30, Hangzhou Xunshu Technology Co., Ltd).
Conflict of Interest
The authors declare no conflict of interest.
Supporting information
Supporting Information
Acknowledgements
The present work was supported by the Special Research Assistant Program of the Chinese Academy of Sciences, Shandong Provincial Natural Science Youth Fund Project (ZR2022QD001, ZR2023QB193), 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.
Wang N., Ma S., Zhang R., Wang L., Wang Y., Yang L., Li J., Guan F., Duan J., Hou B., Regulating N Species in N‐Doped Carbon Electro‐Catalysts for High‐Efficiency Synthesis of Hydrogen Peroxide in Simulated Seawater. Adv. Sci. 2023, 10, 2302446. 10.1002/advs.202302446
Contributor Information
Ruiyong Zhang, Email: ruiyong.zhang@qdio.ac.cn.
Jizhou Duan, Email: duanjz@qdio.ac.cn.
Baorong Hou, Email: brhou@qdio.ac.cn.
Data Availability Statement
The data that support the findings of this study are available from the corresponding author upon reasonable request.
References
- 1. Zhou W., Gao J., Ding Y., Zhao H., Meng X., Wang Y., Kou K., Xu Y., Wu S., Qin Y., Chem. Eng. J. 2018, 338, 709. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2. Ko Y.‐J., Kim H.‐G., Seid M. G., Cho K., Choi J.‐W., Lee W.‐S., Hong S. W., ACS Sustainable Chem. Eng. 2018, 6, 14857. [Google Scholar]
- 3. Yang S., Verdaguer‐Casadevall A., Arnarson L., Silvioli L., Čolić V., Frydendal R., Rossmeisl J., Chorkendorff I., Stephens I. E. L., ACS Catal. 2018, 8, 4064. [Google Scholar]
- 4. Alvarez P. J. J., Chan C. K., Elimelech M., Halas N. J., Villagrán D., Nat. Nanotechnol. 2018, 13, 634. [DOI] [PubMed] [Google Scholar]
- 5. Wang W., Hu Y., Liu Y., Zheng Z., Chen S., ACS Appl. Mater. Interfaces 2018, 10, 31855. [DOI] [PubMed] [Google Scholar]
- 6. Campos‐Martin J. M., Blanco‐Brieva G., Fierro J. L., Angew Chem Int Ed Engl 2006, 45, 6962. [DOI] [PubMed] [Google Scholar]
- 7. Xia C., Xia Y., Zhu P., Fan L., Wang H., Science 2019, 366, 226. [DOI] [PubMed] [Google Scholar]
- 8.a) Yamanaka I., Murayama T., Angew Chem Int Ed Engl 2008, 47, 1900; [DOI] [PubMed] [Google Scholar]; b) Jung E., Shin H., Hooch Antink W., Sung Y.‐E., Hyeon T., ACS Energy Lett. 2020, 5, 1881. [Google Scholar]
- 9. Sun Y., Silvioli L., Sahraie N. R., Ju W., Li J., Zitolo A., Li S., Bagger A., Arnarson L., Wang X., Moeller T., Bernsmeier D., Rossmeisl J., Jaouen F., Strasser P., J. Am. Chem. Soc. 2019, 141, 12372. [DOI] [PubMed] [Google Scholar]
- 10. Wang N., Ma S., Zuo P., Duan J., Hou B., Adv. Sci. 2021, 8, 2100076. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11. Hu C., Paul R., Dai Q., Dai L., Chem. Soc. Rev. 2021, 50, 11785. [DOI] [PubMed] [Google Scholar]
- 12. Wang Y., Shi R., Shang L., Peng L., Chu D., Han Z., Waterhouse G. I. N., Zhang R., Zhang T., Nano Energy 2022, 96, 107046. [Google Scholar]
- 13. Lu Z., Chen G., Siahrostami S., Chen Z., Liu K., Xie J., Liao L., Wu T., Lin D., Liu Y., Jaramillo T. F., Nørskov J. K., Cui Y., Nat. Catal. 2018, 1, 156. [Google Scholar]
- 14. Bu Y., Wang Y., Han G. F., Zhao Y., Ge X., Li F., Zhang Z., Zhong Q., Baek J. B., Adv. Mater. 2021, 33, 2103266. [DOI] [PubMed] [Google Scholar]
- 15. Qiao M., Tang C., He G., Qiu K., Binions R., Parkin I. P., Zhang Q., Guo Z., Titirici M. M., J. Mater. Chem. A 2016, 4, 12658. [Google Scholar]
- 16. Iglesias D., Giuliani A., Melchionna M., Marchesan S., Criado A., Nasi L., Bevilacqua M., Tavagnacco C., Vizza F., Prato M., Fornasiero P., Chem 2018, 4, 106. [Google Scholar]
- 17. Liang H.‐W., Zhuang X., Brüller S., Feng X., Müllen K., Nat. Commun. 2014, 5, 4973. [DOI] [PubMed] [Google Scholar]
- 18. Ding W., Wei Z., Chen S., Qi X., Yang T., Hu J., Wang D., Wan L. J., Alvi S. F., Li L., Angew Chem Int Ed Engl 2013, 52, 11755. [DOI] [PubMed] [Google Scholar]
- 19. Guo D., Shibuya R., Akiba C., Saji S., Kondo T., Nakamura J., Science 2016, 351, 361. [DOI] [PubMed] [Google Scholar]
- 20. Geng S., Ji Y., Yang S., Su J., Hu Z., Chan T.‐S., Yu H., Li Y., Chin Y.‐Y., Huang X., Shao Q., Adv. Funct. Mater. 2023, 33, 2300636. [Google Scholar]
- 21. Hwang S. M., Park J. S., Kim Y., Go W., Han J., Kim Y., Kim Y., Adv. Mater. 2019, 31, 1804936. [Google Scholar]
- 22.a) Yu L., Zhu Q., Song S., McElhenny B., Wang D., Wu C., Qin Z., Bao J., Yu Y., Chen S., Ren Z., Nat. Commun. 2019, 10, 5106; [DOI] [PMC free article] [PubMed] [Google Scholar]; b) Yu J., Li B.‐Q., Zhao C.‐X., Zhang Q., Energy Environ. Sci. 2020, 13, 3253. [Google Scholar]
- 23.a) Ito K., Takayama Y., Makabe N., Mitsui R., Hirokawa T., J. Chromatogr. A 2005, 1083, 63; [DOI] [PubMed] [Google Scholar]; b) Zhao Q., Wang Y., Lai W.‐H., Xiao F., Lyu Y., Liao C., Shao M., Energy Environ. Sci. 2021, 14, 5444. [Google Scholar]
- 24. Jiang H.‐L., Liu B., Lan Y.‐Q., Kuratani K., Akita T., Shioyama H., Zong F., Xu Q., J. Am. Chem. Soc. 2011, 133, 11854. [DOI] [PubMed] [Google Scholar]
- 25. Zheng F., Yang Y., Chen Q., Nat. Commun. 2014, 5, 5261. [DOI] [PubMed] [Google Scholar]
- 26. Sun Y., Sinev I., Ju W., Bergmann A., Dresp S., Kühl S., Spöri C., Schmies H., Wang H., Bernsmeier D., Paul B., Schmack R., Kraehnert R., Roldan Cuenya B., Strasser P., ACS Catal. 2018, 8, 2844. [Google Scholar]
- 27. Lai L., Potts J. R., Zhan D., Wang L., Poh C. K., Tang C., Gong H., Shen Z., Lin J., Ruoff R. S., Energy Environ. Sci. 2012, 5, 7936. [Google Scholar]
- 28. Peng L., Hung C.‐T., Wang S., Zhang X., Zhu X., Zhao Z., Wang C., Tang Y., Li W., Zhao D., J. Am. Chem. Soc. 2019, 141, 7073. [DOI] [PubMed] [Google Scholar]
- 29. Li J., Chen M., Cullen D. A., Hwang S., Wang M., Li B., Liu K., Karakalos S., Lucero M., Zhang H., Lei C., Xu H., Sterbinsky G. E., Feng Z., Su D., More K. L., Wang G., Wang Z., Wu G., Nat. Catal. 2018, 1, 935. [Google Scholar]
- 30. Niu W.‐J., Wang Y.‐P., He J.‐Z., Liu W.‐W., Liu M.‐C., Shan D., Lee L., Chueh Y.‐L., Nano Energy 2019, 63, 103788. [Google Scholar]
- 31. Pfau S. A., La Rocca A., Haffner‐Staton E., Rance G. A., Fay M. W., Brough R. J., Malizia S., Carbon 2018, 139, 342. [Google Scholar]
- 32. Gohda S., Saito M., Yamada Y., Kanazawa S., Ono H., Sato S., J. Mater. Sci. 2021, 56, 2944. [Google Scholar]
- 33. Wang Y., Xu L., Wang M., Pang W., Ge X., Macromolecules 2014, 47, 3901. [Google Scholar]
- 34. Ma S., Zhang Z., Wang Y., Yu Z., Cui C., He M., Huo H., Yin G., Zuo P., Chem. Eng. J. 2021, 418, 129410. [Google Scholar]
- 35. Wang N., Lu B., Li L., Niu W., Tang Z., Kang X., Chen S., ACS Catal. 2018, 8, 6827. [Google Scholar]
- 36. Liu D., Dai L., Lin X., Chen J.‐F., Zhang J., Feng X., Müllen K., Zhu X., Dai S., Adv. Mater. 2019, 31, 1804863. [DOI] [PubMed] [Google Scholar]
- 37. Xu Y., Sumboja A., Zong Y., Darr J. A., Catal. Sci. Technol. 2020, 10, 2173. [Google Scholar]
- 38. Wang N., Ma S., Duan J., Zhai X., Guan F., Wang X., Hou B., Electrochim. Acta 2020, 340, 135880. [Google Scholar]
- 39. Sun Y., Li S., Jovanov Z. P., Bernsmeier D., Wang H., Paul B., Wang X., Kühl S., Strasser P., ChemSusChem 2018, 11, 3388. [DOI] [PubMed] [Google Scholar]
- 40. Siahrostami S., Verdaguer‐Casadevall A., Karamad M., Deiana D., Malacrida P., Wickman B., Escudero‐Escribano M., Paoli E. A., Frydendal R., Hansen T. W., Chorkendorff I., Stephens I. E. L., Rossmeisl J., Nat. Mater. 2013, 12, 1137. [DOI] [PubMed] [Google Scholar]
- 41. Gao J., Yang H. b., Huang X., Hung S.‐F., Cai W., Jia C., Miao S., Chen H. M., Yang X., Huang Y., Zhang T., Liu B., Chem 2020, 6, 658. [Google Scholar]
- 42. Kulkarni A., Siahrostami S., Patel A., Nørskov J. K., Chem. Rev. 2018, 118, 2302. [DOI] [PubMed] [Google Scholar]
- 43. Jiang K., Back S., Akey A. J., Xia C., Hu Y., Liang W., Schaak D., Stavitski E., Nørskov J. K., Siahrostami S., Wang H., Nat. Commun. 2019, 10, 3997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44. Wang N., Zhao X., Zhang R., Yu S., Levell Z. H., Wang C., Ma S., Zou P., Han L., Qin J., Ma L., Liu Y., Xin H. L., ACS Catal. 2022, 12, 4156. [Google Scholar]
- 45. Han L., Sun Y., Li S., Cheng C., Halbig C. E., Feicht P., Hübner J. L., Strasser P., Eigler S., ACS Catal. 2019, 9, 1283. [Google Scholar]
- 46. San Roman D., Krishnamurthy D., Garg R., Hafiz H., Lamparski M., Nuhfer N. T., Meunier V., Viswanathan V., Cohen‐Karni T., ACS Catal. 2020, 10, 1993. [Google Scholar]
- 47. Jia N., Yang T., Shi S., Chen X., An Z., Chen Y., Yin S., Chen P., ACS Sustainable Chem. Eng. 2020, 8, 2883. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Supporting Information
Data Availability Statement
The data that support the findings of this study are available from the corresponding author upon reasonable request.
